Sunteți pe pagina 1din 566

Gas Research Institute

Underbalanced
Drilling Manual

Underbalanced Drilling Manual

Published by
Gas Research Institute
Chicago, Illinois

GRI Reference No. GRI-97/0236

Copyright 1997 by Gas Research Institute


All Rights Reserved

This work is the property of Gas Research Institute. No part of this work may be used or
reproduced without prior written permission from Gas Research Institute, and no part of this
work may be transmitted to any other party in any form or by any means, electronic or
mechanical, including without limitation, photocopy, recording or input into any information
storage or retrieval system without prior written permission of Gas Research Institute.
Requests for permission to reproduce any part of the work should be mailed to:
Contract and License Management Group
Gas Research Institute
8600 West Bryn Mawr Avenue
Chicago, Illinois 60631

LEGAL NOTICE
This publication was prepared as an account of work sponsored by Gas Research Institute (GRI)
and other organizations. Neither GRI, members of GRI, nor any person or organization acting on
behalf of either:
A. MAKES ANY WARRANTY OR REPRESENTATION, EXPRESS OR IMPLIED
WITH RESPECT TO THE ACCURACY, COMPLETENESS, OR USEFULNESS OF
THE INFORMATION CONTAINED IN THIS PUBLICATION, THAT THE USE OF
ANY INFORMATION, APPARATUS, METHOD, OR PROCESS DISCLOSED IN
THIS PUBLICATION MAY NOT INFRINGE PRIVATELY OWNED RIGHTS, OR
B. ASSUMES ANY LIABILITY WITH RESPECT TO THE USE, OR FOR DAMAGES
RESULTING FROM THE USE OF, ANY INFORMATION, APPARATUS, METHOD,
OR PROCESS DISCLOSED IN THIS PUBLICATION.

ii

About this Manual


One of GRIs primary exploration and production goals has been to lower the cost of finding and
developing natural gas reserves. Appropriate application of underbalanced drilling techniques
has the potential to significantly impact drilling costs. In the consummate, original manual on air
drilling, Lyons, 1984, estimated that up to 30% of the wells drilled in the United States were
candidates for air drilling. Today, with improved technology, the number of wells that could be
safely and cost-effectively drilled with air or other underbalanced technologies is probably
significantly higher than this. However, some estimates are that only about ten percent of the
current wells in the United States are being drilled underbalanced. Initial evaluations by GRI
attributed this low percentage to lack of public-domain knowledge and experience. This manual
is an attempt to consolidate some of the publicly available knowledge, protocol and experience in
one reference.
Industry experience has been drawn on to assemble a thorough presentation of underbalanced
drilling. The goal has been to characterize various techniques and methodologies air, N2,
natural gas, mist, foam, mudcap, flowdrilling, coiled tubing drilling, snub drilling ). Various
methods are described and the concepts and operational considerations are indicated to assist in
designing, planning and running underbalanced drilling operations. Underlying concepts are
emphasized so that engineers can evaluate new developments and techniques that are occurring
rapidly.

iii

About the Authors


John McLennan is Vice President of TerraTek, Inc., in Salt Lake City, Utah. John has a Ph.D. in
Civil Engineering (Rock Mechanics); University of Toronto, 1980. Prior to joining TerraTek, he
worked for Dowell Schlumberger, Inc. He has varied engineering experience in rock mechanics,
analytical/numerical modeling, hydraulic fracturing, horizontal wellbore stability, coalbed methane
exploitation and core analysis. He has more than fifty technical publications.
Richard S. Carden is Manager of Special Projects for Grace, Shursen, Moore and Associates, Inc.
(GSM), in Amarillo, Texas. He earned a B.S. degree in Petroleum Engineering from Montana
College of Mineral Science and Technology in 1977. Rich has more than eighteen years of drilling
engineering and operational experience, including geothermal wells; deep, high pressure gas wells;
air drilled wells; as well as directional and horizontal wells. He has worked on the Grand Canyon
Waterline Project (an air and mist drilled, directional well), on DOE horizontal air drilling projects
and on numerous frontier technology drilling and well control operations. Rich teaches classes
worldwide on many aspects of drilling, including air and gas drilling, horizontal operations and
advanced well control. He has authored numerous technical papers on directional and air drilling.
David Curry has fourteen years of experience in drilling research and developing drilling
technology. He is currently Technical Support Manager for Hughes Christensens OASIS (Drilling
Performance Optimization) group. Previously, he has been the Drilling Manager for TerraTek,
Inc.s Drilling and Completions Laboratory, Managing Director of the International Drilling and
Downhole Technology Center, and Drilling Mechanics Program Leader at Schlumberger
Cambridge Research. David has an M.A. in Natural Sciences and a Ph.D. in Fracture Mechanics,
both from the University of Cambridge.
C. Rick Stone is founder, President and owner of Signa Engineering Corporation, a contract
petroleum engineering and project management company, based in Houston, Texas. He currently
designs and supervises the drilling and completion of wells, many of them horizontal, throughout
the world. This includes many underbalanced Austin chalk wells. Rick has over sixteen years of
experience in drilling and production engineering, mostly with Sun Exploration Production and
then Oryx Energy Company. During his tenure with Oryx, Rick devised the technique for
flowdrilling horizontal wells in the Austin chalk and helped to design the Rotating Blowout
Preventer (RBOP). He was an SPE Distinguished Lecturer, on underbalanced drilling, in 19931994. Rick received his B.S. degree in Mechanical Engineering from Texas A&M University in
1979.
Richard E. Wyman recently retired as Vice President of Canadian Hunter Exploration, Ltd. He
received his B.S. degree in Mechanical Engineering from the University of California. Prior to
joining Canadian Hunter, Dick was with Shell Oil Company for twenty-three years. During this
time, he was responsible for the Pacific Coast area logging and wellbore evaluation, and for
petrophysical research in Houston; he also held positions of Chief Engineer for Shell and Manager
of Engineering Research for Shell Development. Twenty years ago, Dick joined Canadian Hunter
as Director of Research. He has authored a number of technical papers and holds several patents.

iv

Acknowledgments
Contributors and Reviewers
Gas Research Institute and the authors wish to thank the people and organizations listed below
for contributing to and reviewing this manual. Their comments and recommendations were a
great asset.
Santos Ltd provided unpublished case studies of successful air drilling operations in Australia.
Brian Tarr, Mobil Technology Company, provided detailed reviews, suggestions and guidance
for all chapters of this manual.
Mike Akins, Chevron Petroleum Technology Corporation, reviewed all chapters of the manual.
Reuben Graham, Reuben L. Graham, Inc., provided important practical observations on the first
two chapters and provided an unsolicited protocol for unloading holes.
Les Shale, Baker Hughes Inteq, reviewed and critiqued the first two chapters of the manual.
Mike Tweedy, Chevron Petroleum Technology Company, reviewed all chapters of the manual.
Jim Williams, Parker Drilling Company, provided valuable review and critique of the first two
chapters of the manual.
Mike Weiss, Gas Research Institute, was the driving force for production of this manual, and
provided in-depth review and editorial revisions for the entire manual.

Manuscript Preparation
Sherri Heroux, TerraTek, Inc., prepared and laid out the entire manuscript, text and illustrations.

Gas Research Institute and the authors wish to express their appreciation to all other
individuals in the natural gas industry, especially members of the GRI Natural Gas
Project Advisory Group, who offered support and useful suggestions for the
development of a complete and balanced Underbalanced Drilling Manual.

The cover artwork was conceived by Rick Stone and prepared by Ben Siegel, both of Signa
Engineering Corporation, in Houston Texas.

Table of Contents
About this Manual.................................................................................................................. iii
About the Authors..................................................................................................................iv
Acknowledgments ..................................................................................................................v

CHAPTER 1

INTRODUCTION

1.1

What Is Underbalanced Drilling? ............................................................................... 1-1

1.2

Why Drill Underbalanced?......................................................................................... 1-2

1.3

Underbalanced Drilling Techniques ........................................................................... 1-4

1.4

Limitations To Underbalanced Drilling ..................................................................... 1-12

1.5

Summary................................................................................................................. 1-15

CHAPTER 2

UNDERBALANCED DRILLING TECHNIQUES

2.1

Dry Air Drilling ........................................................................................................... 2-1

2.2

Nitrogen Drilling....................................................................................................... 2-55

2.3

Natural Gas Drilling ................................................................................................. 2-62

2.4

Mist Drilling.............................................................................................................. 2-70

2.5

Stable Foam Drilling .............................................................................................. ..2-88

2.6

Stiff Foam Drilling .................................................................................................. 2-138

2.7

Gasified Liquids..................................................................................................... 2-152

2.8

Flowdrilling ............................................................................................................ 2-205

2.9

Mudcap Drilling...................................................................................................... 2-215

2.10 Snub Drilling .......................................................................................................... 2-220


2.11 Closed Systems .................................................................................................... 2-226

CHAPTER 3

BENEFITS OF DRILLING UNDERBALANCED

3.1

Penetration Rate ....................................................................................................... 3-1

3.2

Bit Life ..................................................................................................................... 3-17

3.3

Differential Sticking.................................................................................................. 3-22

3.4

Lost Circulation........................................................................................................ 3-24

3.5

Formation Evaluation............................................................................................... 3-25

3.6

Formation Damage.................................................................................................. 3-28

vi

CHAPTER 4

SELECTING AN APPROPRIATE TECHNIQUE

4.1

Introduction ............................................................................................................... 4-1

4.2

Potential Applications ................................................................................................ 4-1

4.3

Technical Feasibility .................................................................................................. 4-7

4.4

Economic Analysis .................................................................................................. 4-36

CHAPTER 5

WELL ENGINEERING

5.1

Circulation Programs ................................................................................................. 5-1

5.2

Circulation Calculations (Air, Gas, Mist) .................................................................... 5-6

5.3

Circulation Calculations (Gasified Liquids)............................................................... 5-10

5.4

Wellhead Design ..................................................................................................... 5-22

5.5

Casing Design ......................................................................................................... 5-31

5.6

Completion Design .................................................................................................. 5-33

5.7

Bit Selection ............................................................................................................ 5-36

5.8

Underbalanced Perforating...................................................................................... 5-43

5.9

Drillstring Design ..................................................................................................... 5-45

CHAPTER 6

SPECIAL CONSIDERATIONS

6.1

Safety in Underbalanced Drilling ............................................................................... 6-1

6.2

Regulatory Requirements.......................................................................................... 6-4

6.3

Environmental Issues ................................................................................................ 6-8

6.4

Directional Drilling.................................................................................................... 6-10

6.5

Percussion Drilling................................................................................................... 6-34

6.6

High Pressure Drilling.............................................................................................. 6-47

6.7

Cementing............................................................................................................... 6-53

6.8

Formation Evaluation............................................................................................... 6-61

CHAPTER 7

CASE STUDIES

7.1

Introduction ............................................................................................................... 7-1

7.2

Case Study 1 - Controlling Bottomhole Pressure....................................................... 7-3

7.3

Case Study 2 - Barrolka 3 ....................................................................................... 7-12

7.4

Case Study 3 - Swan Lake-1 ST ............................................................................. 7-16

7.5

Case Study 4 - Karwin-1 ST .................................................................................... 7-19

7.6

Case Study 5 - Unloading the Hole From the Bottom .............................................. 7-22

vii

7.7

Case Study 6 - Gasified Liquid (Concentric String Injection) ................................... 7-32

7.8

Case Study 7 (Underbalanced Re-Entry) ................................................................ 7-35

7.9

Case Study 8 - Controlled Tripping.......................................................................... 7-38

7.10 Case Study 9 - Flowdrilling...................................................................................... 7-40


7.11 Case Study 10 - Coiled Tubing Drilling .................................................................... 7-42
7.12 Case Study 11 - Cementing .................................................................................... 7-47
7.13 Case Study 12 - The Friction Dominated Regime.................................................... 7-49
APPENDIX A......................................................................................................................A-1
APPENDIX B......................................................................................................................B-1
APPENDIX C......................................................................................................................C-1
APPENDIX D......................................................................................................................D-1

viii

INTRODUCTION

his chapter introduces underbalanced drilling, summarizes the different techniques available
for drilling underbalanced, and indicates the various benefits and restrictions to their use.

1.1

What is Underbalanced Drilling?

Most oil and natural gas wells are drilled using rotary techniques, in which a drill bit
disaggregates rock at the base of the well. A drilling fluid is pumped to the bottom of the hole
and then back up to the surface. The fluid is pumped down the well inside the drillstring and it
returns to the surface, in the annulus between the drillstring and the borehole wall. As it flows
over the hole bottom, the drilling fluid entrains the rock cuttings and removes them to the
surface.
In conventional drilling operations, the drilling fluid serves several other functions. These
include stabilizing the borehole, cooling the bit, and most importantly, controlling the formation
fluids. The well is said to be at balance if the borehole and formation fluid pressures are equal.
In this situation, there is no net fluid flow into or out of the borehole.
The composition and properties of the drilling fluid are often chosen to ensure that the fluid
pressure in the borehole exceeds the pore fluid pressure in the formations penetrated by the
wellbore, at all depths where the formation is open to the borehole. In this overbalanced
situation, the drilling fluid pressure prevents formation fluids from flowing into the well during
drilling. There is some fluid flow from the borehole into the rock around the borehole. Materials
are added to the drilling fluid to restrict this flow, by depositing low permeability filter cake on
the borehole wall and in the pores and fractures adjacent to the borehole.
In underbalanced drilling operations, the pressure of the drilling fluid in the borehole is
intentionally maintained below the formation pore fluid pressure, in the openhole section of the
well. As a result, formation fluids flow into the well when a permeable formation is penetrated
during underbalanced drilling. For this reason, underbalanced drilling is sometimes referred to as
flowdrilling.

1-1

Chapter 1

Introduction

Special equipment and procedures are required to control formation fluid inflow during
underbalanced drilling. Nevertheless, drilling underbalanced offers several significant benefits
over conventional drilling techniques. These include:








Increased penetration rate and bit life,


Reduced probability of sticking the drillstring downhole,
Minimized lost circulation while drilling,
Improved formation evaluation,
Increased well productivity, and,
The requirement for primary stimulation treatments can be reduced or eliminated.

The extent to which it is possible to achieve any of these benefits is generally controlled by the
properties of the target reservoir and overlying formations; and, in some instances, even by the
specific characteristics of the well being drilled.

1.2

Why Drill Underbalanced?

The simple answer to the question Why drill underbalanced? is that it can improve the
financial returns on drilling the well. This improvement can come from a variety of different
factors that reduce the cost of drilling the well or increase its productivity once drilled.
Increased Penetration Rate
Drilling underbalanced can lead to increased penetration rate. Most references, describing
drilling operations with air or lightened drilling fluids, report penetration rates which are greater
than these for wells drilled overbalanced with conventional liquid drilling fluids. A systematic
review of air drilling operations, conducted by Carden, 1993,1 cited that air drilling penetration
rates could be as much as ten times greater than those for mud drilling in equivalent formations.
Increased Bit Life
It is often claimed that bit life is increased when lightened fluids are used instead of conventional
drilling muds. Drilling underbalanced removes the confinement imposed on the rock by the
overbalance pressure. This should decrease the apparent strength of the rock and reduce the
work that must be done to drill away a given volume of rock. It is reasonable that this increased
drilling efficiency should increase the amount of hole that can be drilled before the bit reaches a
critical wear state.
Minimized Lost Circulation
Lost circulation occurs when drilling fluid enters an open formation downhole, rather than
returning to the surface. It is possible for drilling fluid to be lost by flow into a very permeable
zone. More frequently, lost circulation involves flow into natural fractures that intersect the
wellbore or into fractures induced by excessive drilling fluid pressure. Lost circulation can be
very costly during conventional drilling. The lost fluid has to be replaced, and the losses have to
1-2

be mitigated, usually by adding lost circulation material to the mud (to plug off the path by which
the fluid is entering the formation), before drilling can safely be resumed. Since there is no
physical force driving drilling fluid into the formation if the well is drilled underbalanced,
underbalanced drilling effectively prevents lost circulation problems.
This is not to say that lost circulation cannot occur when drilling with lightened fluids. It is
possible to lose circulation whenever the wellbore pressure exceeds the formation pore pressure.
Using a lightened fluid does not, by itself, guarantee underbalanced conditions. This is clearly
illustrated by a well, drilled with mist, in the Grand Canyon National Park.2 The pore pressure
gradient was almost zero and air circulation could only be achieved for less than one-half of the
drilling time.
It is possible for chemical driving forces, caused by activity differences between the aqueous
phase of a drilling fluid and the formation, to cause fluid to enter the formation, even though
there is a pressure gradient driving flow in the opposite direction. These low rates are usually so
low that they are undetectable while drilling although they can affect a wells subsequent
productivity.
Minimized Differential Sticking
In a well drilled conventionally, a filter cake forms on the borehole wall from solids deposited
when liquid flows from the drilling mud into permeable zones, due to an overbalance pressure.
If the drillstring becomes embedded in the filter cake, the pressure differential between the
wellbore and the fluid in the filter cake can act over such a large area that the axial force required
to move the string can exceed its tensile capacity. The drillstring is then differentially stuck.
There will be no filter cake and no pressure acting to clamp the drillstring if the well is
underbalanced. Other mechanisms can cause sticking; underbalanced drilling does not eliminate
the possibility of a stuck drillstring (refer to Section 1.4).
Reduced Formation Damage
Anticipated well productivity is often reduced by regions of impaired permeability, formation
damage, adjacent to the wellbore. Formation damage can occur when liquid(s), solid(s) or both
enter the formation, during drilling.3,4 If the drilling fluid pressure in the wellbore is less than the
pore pressure, the physical driving force causing penetration of material from the drilling fluid is
removed. That is not to say that the possibility of formation damage from the drilling fluid is
completely removed. In some circumstances, chemical potential differences between drilling and
pore fluids could cause filtrate to enter the formation against the pressure gradient.3 Also, there
are instances in which a well, that is drilled nominally underbalanced, experiences transient
overbalanced conditions, due to less than perfect control of circulating pressures or possibly due
to fluid inflow while the well is not being circulated.5
In any case, there are many examples of wells drilled underbalanced with higher productivity
than adjacent wells drilled conventionally.

1-3

Chapter 1

Introduction

Earlier Production
When a well is drilled underbalanced, hydrocarbon production can begin as soon as a productive
zone is penetrated. With suitable surface equipment, it is possible to collect oil while drilling.
Some underbalanced wells have paid for themselves entirely from production before drilling
operations were completed.
Reduced Stimulation Requirements
Following conventional drilling operations, wells are often stimulated to increase their
productivity. Stimulation can include acidizing or surfactant treatments, to remove formation
damage; or hydraulic fracturing can be used to guarantee adequate production in low
permeability reservoirs or to bypass damage in higher permeability formations. Reduced
formation damage means lower stimulation costs.
Improved Formation Evaluation
Drilling underbalanced can improve the detection of productive hydrocarbon zones, even
identifying zones that might otherwise have been bypassed if the well had been drilled
conventionally.
When a well is drilled underbalanced, formation fluids flow into the wellbore from any
permeable formation in the openhole section. Penetrating any hydrocarbon-bearing formation
with adequate drive and permeability will result in an increased hydrocarbon cut in the drilling
fluid returning to the surface. With adequate mud logging and drilling records, underbalanced
drilling can indicate potentially productive zones, as the well is drilled. Conversely, during
conventional drilling, the overbalance pressure prevents formation inflows; hydrocarbon-bearing
zones have to be identified from cuttings, core analysis, logging or DSTs.
It should be possible to use the volumes of produced hydrocarbons, from a well drilled
underbalanced, to give an indication of the productivity of any pay zones that have been
penetrated. Shutting down circulation will allow flow from the well to be measured; for example
with a pitot tube or flow prover.1 This is a straightforward practice when a gas well is drilled
with a dry gas or mist. The length of the flare at the flare pit can give a qualitative indication of
productivity. Since this is also influenced by the rate of circulation, the size of the blooie line
and the wind direction, it is difficult to quantify gas production rates solely from the flare length.
If a drilling fluid with non-negligible liquid content is being used, the well will have to be
allowed to unload before stable flow is established. The flow measured is the sum of the
production from all open zones; different zones are not being selectively tested.
The reduction or elimination of formation damage that results from drilling underbalanced will
also improve the interpretation of openhole logs. For example, there should be no modification
of formation fluid composition adjacent to the wellbore, that might otherwise mask the presence
of hydrocarbons.

1-4

Environmental Benefits
There can be environmental benefits associated with properly managed, underbalanced drilling
operations. These depend on the exact drilling technique adopted. With dry, gaseous drilling
fluids there is no potentially damaging liquid drilling mud to dispose of after drilling is
completed. The chemicals used in mist and foam drilling are often benign and biodegradable
surfactants that do not pose significant environmental concerns.
On the other hand, formation fluids are produced while drilling underbalanced. Particularly with
open surface systems, these have to be handled carefully, to avoid environmental contamination.
However, with closed surface systems, there is no reserve pit and both cuttings and produced
fluids are contained in a way that minimizes the potential for environmental contamination.

1.3

Underbalanced Drilling Techniques

Many different techniques are available for intentionally achieving underbalanced conditions.
These mostly involve circulating a drilling fluid with a density that gives a hydrostatic pressure
gradient in the wellbore that is less than the pore pressure gradient. The drilling fluid may be a
single gas or liquid phase, or a two-phase gas-liquid mixture. When there is any significant
volume fraction of gas (injected or produced) present, the drilling fluid will be compressible.
Underbalanced drilling does not require the use of a compressible drilling fluid - a conventional,
liquid drilling fluid can give underbalanced conditions in normally or over-pressured formations,
if the circulating pressure (sum of the hydrostatic head and the frictional pressure drop to the
open end of the fluid return line) is less than the pore pressure. Using a drilling fluid with a
density less than the reservoir pressure gradient does not guarantee underbalanced conditions.
Particularly with foamed fluids, the frictional pressure drop can be substantial. This can result in
a circulating wellbore pressure that exceeds the pore pressure even when the hydrostatic head of
the drilling fluid does not.
Gaseous Drilling Fluids
Probably the simplest and oldest of all underbalanced drilling techniques is to use dry air as the
drilling fluid.6 Compressors pump air to the swivel attached to the top of the drillstring, down
the drillstring, through the drill bit and back up the annulus. Figure 1-1 is a schematic of a
simplified air drilling flow system. A rotating head provides a low pressure seal around the
drillstring between the wellhead and the drill rigs rotary table, which diverts the return flow
away from the rig floor. For this reason, it is sometimes referred to as a diverter. A blooie line
takes the returning, cuttings-laden air flow from immediately beneath the rotating heads seal to a
safe distance away from the rig. The cuttings are allowed to collect in a pit. It is common to use
some form of water spray close to the exit point on the blooie line, to prevent dust clouds. A
lighted flame is maintained at the exit from the blooie line to ignite any hydrocarbon gas in the
return flow.

Detailed rig-site layouts are shown in Chapter 2.

1-5

Chapter 1

Introduction

It is possible to use an inert gas, instead of air, as the circulating fluid; nitrogen is almost
invariably chosen. Other inert gases are usually too expensive. Nitrogen may be brought to
location as a liquid in cryogen tanks. In this case, heaters are used to boil the liquid nitrogen
before it is compressed and pumped downhole. It is also possible to generate nitrogen, using
membrane-type filters that remove oxygen from the air flow delivered by the compressors before
it is pumped downhole.7
Another option is to use natural gas as the drilling fluid. This can be less costly than using
nitrogen, when drilling in a producing gas field or close to a natural gas pipeline. The pipeline
pressure may be large enough that compressors are not needed. Pressure boosters will normally
be required. API and NFPA (National Fire Protection Agency) guidelines must be followed.
Whatever gas is used, the compression system has to be have sufficient flow rate and delivery
pressure capacities to be able to lift drilled cuttings and fluid influxes from the wellbore.
Circulating pressure and cuttings transport are not independent of one another. The weight of
cuttings being transported up the annulus increases the wellbore pressure. As the penetration rate
(the rate of cuttings generation) increases, so does the wellbore pressure. If the circulation rate is
too low, cuttings will accumulate in the wellbore, the wellbore pressure will increase and flow
will be choked.

C o m p r esso r

C o m p r esso r

B o o ster

M i st/F o a m er
Pump

Blooie Line

W ellbore

Figure 1-1.

Simplified air drilling flow system.


8

1-6

Charts prescribing minimum air flow rates for adequate hole cleaning were developed by Angel,
1957.9 These charts are still widely used in the design of air drilling operations. The circulation
rates they suggest correspond to an annular velocity that would be equivalent to 3,000 ft/min at
atmospheric pressure. Angel argued that this was the minimum velocity for effective cuttings
transport. Predicting circulating pressures and appropriate flow rates for gaseous drilling fluids
is not, however, trivial. It will be discussed in much greater detail in Section 2.1. For the time
being, it is sufficient to note that several of the assumptions made in the development of Angels
charts are not conservative10 and that higher gas flow rates are often required than would be
predicted by the charts.11,12 Today, most drilling service companies have predictive simulators.
Two-Phase Drilling Fluids
Mixing of gaseous and liquid phases is a way to achieve any desired drilling fluid density, from
pure gas to pure liquid. These mixtures of gas and liquid are sometimes collectively referred to
as lightened drilling fluids. Lightened drilling fluids can be classified as mist, foam or aerated
liquid, according to the structure and relative volumes of the gaseous and liquid phases. Their
structure and properties depend critically on the relative volume fractions of gas and liquid at the
prevailing conditions of pressure and temperature. Different lightened drilling fluid densities
are shown in Table 1-1.
Table 1-1.

Densities (ppg) of Lightened Drilling Fluids

Description

Density
(ppg)

Gas

0.01 - 0.1

Mist

0.1 - 0.3

Foam

0.3 - 3.54
3.5 - 6.95 with backpressure

Gasified Liquid

4.0 - 6.95

Liquid

6.95 - 19.0

If the volume fraction of an unviscosified liquid is less than about 2.5 percent, the liquid will be
suspended as discrete droplets in a continuous gaseous phase. Drilling with these low liquid
volume fractions is usually referred to as misting or mist drilling. A small triplex pump is
used for low-rate, liquid injection, into the circulating gas at the surface before the fluid enters
the drillstring. The injected liquid is usually water, a surfactant and a corrosion inhibitor.
Sometimes, polymers or salts are added to inhibit interaction with water-sensitive shales. Since
the liquid is present as discrete droplets, it has little direct impact on the rheology of the
circulating gas. The liquid droplets do, however, affect wellbore pressure; in a manner that is
analogous to the influence of drilled cuttings in dry air.

1-7

Chapter 1

Introduction

If the relative liquid volume is higher, a stable foam results. A stable foam is usually generated
when liquids, similar to those used in mist drilling, are injected into the gas flow, at rates giving
downhole liquid volume fractions in the range 2.5 to 25 percent. The liquid forms a
continuous cellular structure, entrapping the gaseous phase. The gas and liquid move together
with nominally the same velocity. Foams are often described in terms of their quality and their
texture. Foam quality is the gas volume fraction, usually expressed as a percentage, at the
prevailing pressure and temperature. For example, a 90 quality foam is 90 percent gas and 10
percent liquid, by volume. The texture describes the bubble structure of the foam - a fine foam
has small gas bubbles and a coarse foam has large bubbles.
Foams have high viscosities, enabling very good cuttings transport. Foam rheology is largely
controlled by the liquid volume fraction, at the prevailing pressure.13 Foam viscosities have been
measured ranging from 115 cP at a liquid volume fraction of 2.5 percent, to 35 cP, at a liquid
volume fraction of 25 percent.13 In some instances, viscosifiers are added to the injected liquid.
These are termed stiff foams.
If cuttings are to be efficiently removed from a well, the foam needs to be sufficiently stable to
retain its structure until it is discharged from the blooie line. Untreated drilling foam can have a
lifetime of several hours after it returns to the surface. Since the circulating rate is usually
hundreds of cfm, very large volumes of foam might have to be contained at the surface. This
foam containment problem can be overcome by using an appropriate defoaming system.
Once the liquid volume fraction exceeds about 25 percent, the foam structure breaks down. The
gas forms isolated bubbles that are independent of the liquid phase to the extent that the two
phases can move with different velocities. When gas-liquid mixtures with this structure are used
as drilling fluids, they are usually described as aerated muds. They can, be formed with
combinations of gas and liquid other than air and drilling mud. It is not unusual for the liquid to
be fresh water or brine, with or without viscosifiers, diesel oil or even crude oil.
The drilling rigs mud pumps are used to pump the liquid phase to the standpipe and from there
into the drillstring. Compressors, suitable for the gas in use, are normally arranged to inject the
gas into the flowing mud at or close to the standpipe, aerating the drilling fluid before it is
pumped down the drillstring. The aerated liquid returning from the well is passed through a gasliquid separation system. The gas is then directed to a flare pit, while the liquid flows through a
conventional solids removal system. Additional gas separation measures may be necessary before
the liquid can be pumped downhole again (if at all). Some form of oil-mud separation system
may be required if liquid hydrocarbons are produced. Figure 1-2 schematically illustrates the
main elements of an aerated mud drilling system.
It is possible to create an aerated drilling fluid downhole rather than at the surface. This is most
often done by using a parasite string. This is a small-bore tubular that penetrates the wellhead
and leads into the wellbore just above the last casing shoe. By circulating a gas, which may be
air, nitrogen or natural gas, down the parasite string and into the wellbore, the density of the fluid
returning up the annulus is reduced (above the injection point) and the wellbore pressure
1-8

decreases correspondingly. The same effect can be achieved during re-entry drilling of
previously gas-lifted wells if the drilling assembly can be run downhole without pulling the gas
lift tubing and production casing string. This requires a small diameter, drilled hole. This is
probably only practical when drilling with coiled tubing. It is also possible to aerate the returning
drilling fluid by injecting gas outside an uncemented casing string, or by using a dual wall drill
pipe, where the drilling liquid is pumped down the central conduit of the pipe and the gas down
the outer conduit. These techniques will be explained in more detail in Section 2.7.
Mist drilling is often used after a significant water influx is encountered while dry air drilling.
Normally air is used in mist drilling. Other than cost, there is, no fundamental reason why other
gases, such as nitrogen, carbon dioxide or natural gas, should not be used in lightened drilling
fluids. If the liquid phase is flammable, i.e., crude oil or diesel, nitrogen may be preferred over
air, for reasons of safety. Also, using air in lightened drilling fluids can create conditions
downhole that are ideal for corrosion of the drillstring and any exposed casing. Finally, gas can
dissolve in liquid; some more than others; for example carbon dioxide in aqueous liquids, or
natural gas in crude oil.
Because gases are much more compressible than liquids, the liquid volume fraction in a lightened
drilling fluid will vary as the drilling fluid is circulated around the well. As pressure increases,
the liquid volume fraction and drilling fluid density will also increase. This situation is further
complicated by formation fluids flowing into the well. For example, when a water inflow occurs,
the drilling fluid may be a mist on its way down the drillstring but change to a foam when it
enters the annulus and picks up the additional water from the inflow. Calculating circulating
pressures is critical.

1-9

Chapter 1

Introduction

Compressor

Compressor

Booster

Mud Pump

Mud Pits

Figure 1-2.

O il/Mud
Separator

Shale
Shaker

Oil

Solids

Gas/Liquid
Separator

Flare Pit

Simplified aerated liquid drilling flow system.

Concepts for the design of lightened drilling fluid operations are described in Chapter 2;
commercial simulators are available.
Liquid Drilling Fluids
The formation pore fluid pressure often exceeds the hydrostatic pressure of fresh or saline water
at the same depth. In this environment, it is possible to drill under-balanced using a liquid. It is
not uncommon for conventional drilling oper-ations to become underbalanced (un-intentionally)
if the wellbore penetrates a region of higher than anticipated pore pressure.
In certain circumstances it is possible to achieve underbalanced conditions even though the
drilling fluid has a density exceeding the pore pressure gradient. For example, loss of drilling
fluid into a low pressure zone can reduce the wellbore pressure, allowing formation fluids to flow
into the well from higher up the hole. The inflowing fluids then reduce the drilling fluid density
until circulation is regained and a mixture of drilling and formation fluids flows to the surface.
This is the case in the Pearsall Field in Texas, which has seen one of the most extensive and
successful recent applications of underbalanced drilling in the United States.14
Surface Systems
Probably the key distinction between underbalanced and conventional drilling operations is that
additional surface equipment is required if a well is to be drilled underbalanced. This equipment
essentially diverts all return flow away from the rig floor and separates produced hydrocarbons
1-10

from the drilling fluid in a way that allows them to be contained. In this way, underbalanced
drilling can continue safely once a permeable formation is penetrated.
The complexity of the surface system is influenced by the choice of drilling fluid and the nature
and quantity of formation fluids produced while drilling. In the case of dry air drilling, with
natural gas as the only potential inflow and no potential for hydrogen sulfide, it is often sufficient
to have the blooie line discharge flared over an open, earthen pit in which the cuttings collect. At
the other extreme, a closed, multi-phase separator, used with a nitrified water drilling fluid, has
to handle cuttings, produced oil, produced gas, circulating water, and nitrogen. Such systems
allow oil to be collected for storage, gas to be flared, and water to be re-cycled to the rig pumps.
Broadly, it is possible to characterize the separation systems as open or closed, depending on
whether or not the separation vessels themselves are open to the atmosphere or sealed. Closed
separators are not normally used with drilling fluids containing air, in order to minimize any
explosion hazard when hydrocarbons are encountered. Conversely, a closed system should be
used if hydrogen sulfide may be present in the produced fluids. Specific requirements for
various drilling fluids will be discussed in more detail in the relevant sections of Chapter 2.
In many instances, surface equipment incorporates an adjustable choke in the drilling fluid return
line, between the diverter and the separation system. Back pressure on the well provides some
degree of control over the wellbore pressure, independently from the drilling fluid density and
rheology. If this is to be done, a rotating seal element in the stack is normally required, to
provide sufficient pressure bearing capacity to seal the back pressure generated by the choke.
This technique provides the flexibility in controlling wellbore pressure that can be particularly
important when drilling through poorly consolidated or very productive formations, where it may
be necessary to restrict the underbalance pressure (differential) to a few hundred psi. In air or
mist drilling, if back pressure is increased, annular velocities are reduced and hole cleaning may
be jeopardized. Applying a back pressure can also help to control changes in the liquid volume
fraction with depth. This may be required if a foam is to be maintained throughout the annulus. 15

1-11

Chapter 1

Introduction

1.4 Limitations to Underbalanced Drilling


Along with their benefits, there are technical and economic limitations of underbalanced drilling.
Carden, 1993,1 reported that, in the United States, wellbore instability and water inflow were the
two main reasons for terminating air drilling operations. Other technical factors restricting
underbalanced drilling include downhole fires, directional drilling difficulties, and excessive
hydrocarbon production. Various limitations on underbalanced drilling are outlined below.
Wellbore Instability
As in conventional drilling, wellbore instability may arise from mechanical or chemical
mechanisms. These may be accentuated by drilling underbalanced. Whatever the underlying
mechanism, wellbore instability can result in the drillstring becoming stuck downhole. Rock
fragments, too large for the drilling fluid to lift from the hole, may fall, accumulate and stick the
drillstring or the formation may swell or creep, reducing clearance to the point where the string
sticks.
In conventional, overbalanced drilling operations, the excess of wellbore pressure over the
formation pore pressure provides some degree of support at the borehole wall. In underbalanced
drilling this support is missing; as the degree of underbalance is increased, so too does the
tendency for wellbore instability. This can put a lower limit on the wellbore pressure; below
which it is effectively impossible to drill. This limiting underbalance pressure is principally
influenced by the prevailing in-situ-stresses, the formations strengths, the actual reservoir
pressures and the wellbore geometry. In general, it is normally only older, harder and more
competent formations that have sufficient strengths to allow drilling with dry air without
wellbore instability problems. In some instances, mechanically-induced wellbore instability may
be controlled by adopting a drilling technique that restricts the degree of underbalance to less
than the critical level. In other cases, particularly in tectonically active areas, the wellbore may be
inherently unstable under any conditions.
Chemically-induced wellbore instability can occur when drilling formations with significant
amounts of water-sensitive clays. These may be dehydrated when drilling with a gaseous fluid.
Conversely, these water-sensitive clays may absorb water from an aqueous phase present in the
well when mist, foam or aerated liquid drilling fluids are used. In either case, the change in shale
water content induces additional stresses in the near-wellbore region. These can destabilize the
wellbore. In principle, it is possible to adjust the activity of the aqueous phase; for example, by
the addition of suitable electrolytes, to match the exposed shale and prevent chemically induced
stress changes.16 This can be a challenging task and complicating factors, such as variations in
the water inflow rate salinity, may render it impossible.
Water Inflows
Water inflow can impede underbalanced drilling for several reasons. When drilling with gas,
formation water can moisten the cuttings downhole, causing them to stick together and
accumulate on the drillstring and on the borehole wall. This is most likely to occur at the top of
the drill collars, where the decrease in drillstring diameter leads to a sudden decrease in annular
1-12

velocity. This cuttings accumulation is sometimes called a mud ring. If unchecked, a mud
ring can grow to the point where the string is trapped. Paradoxically, adding water to the
circulating fluid can control the formation of mud rings, by saturating the cuttings and preventing
them from adhering to each other.1 For this reason, it is normal to change from dry gas to mist
drilling when water inflow first occurs.
Metering in foaming agent can allow relatively large water inflows to be lifted from the well.
However, foam stability can be compromised by saline water inflows, or water inflows can be
encountered that are too rapid to be removed from the well, even when circulating foam. In these
circum-stances, the buildup of water in the well can increase the circulating pressure to the point
that the surface equipments pressure capacity is exceeded, requiring a change in drilling fluid.
However, the major concern when large water inflows occur is economic; i.e., disposal costs.
Water inflows may be further controlled by using an aerated drilling fluid. Since these fluids
tend to have lower underbalance pressures (differential) than occur when drilling with gas, mist
or foam, the rate of water inflow will be reduced. Produced water mixes with the liquid phase of
the drilling fluid and is circulated from the well. Inflow may, however, lead to difficulties in
determining the air injection rate required to maintain target wellbore pressures.
There are some circumstances in which it is possible to seal off water-bearing zones, by pumping
chemicals that penetrate the formation and react with one another or with the formation water to
form flow barriers.
Downhole Fires
Downhole fires should perhaps more properly be termed downhole explosions.17 They are
infrequent but their consequences are spectacular18 - the drill collars and bit can be melted or
burnt away. For a fire to occur, the downhole composition of the hydrocarbon and air mixture
has to be in a flammable range. There also has to be an ignition mechanism, such as a mud ring,
downhole sparking, or a small hole or washout in the drillstring.12 A mud ring can lead to
ignition when it seals the annulus and continued circulation causes the pressure of the
hydrocarbon-air mixture to increase to the point at which combustion ignition occurs, much as it
does in a diesel engine. Sparks can result downhole from contact between the drillstring and hard
minerals in the borehole wall. Circulating air, leaking through a small washout in the drillstring,
can cause a local hot spot with the potential to ignite a combustible hydrocarbon/air mixture.
Downhole fires can be avoided by using a non-flammable circulating fluid. It may not be
practical to change from air to a non-flammable gas, such as nitrogen, part way through drilling a
well. In that case, changing from dry air to mist drilling can help by reducing the probability of
forming mud rings. Since the structure of a stable foam made with air isolates the air in separate
bubbles, the air in foam is not normally available for combustion. This is one reason why air
foams are widely used to extinguish hydrocarbon fires; using them in drilling fluid may well
provide a means to avoid downhole fires.

1-13

Chapter 1

Introduction

Directional Drilling Equipment


Difficulties with directional drilling equipment have caused some operators to abandon
underbalanced drilling prior to undertaking directional work. The issue here is the requirement
for the hole to be surveyed frequently, particularly in the case of horizontal wells. Conventional,
mud pulse telemetry Measurement While Drilling (MWD) tools cannot operate with the
compressible fluids often used in underbalanced drilling - the pressure pulses they generate to
convey their signals do not propagate back to the surface with sufficient amplitude to be
detectable. Electromagnetic MWD systems do exist, but there are concerns about their
reliability, although tool development is improving rapidly.1,19 Wireline steering tools can be
used. These have a hard-wired connection from the surveying instruments downhole back to the
surface. They cannot be left downhole if the drillstring is to be rotated, as is normally the case.
The additional time taken to trip the steering tool in and out of the hole can make underbalanced
drilling unattractive. There are wet connect and cartridge wireline systems that allow the
wireline tool to remain in the hole while rotating the drillstring, but they also result in some
slowing of the drilling operation.20,21
Conventional downhole motors were designed to operate on incompressible fluids. Their
performance deteriorates when they are run with compressible fluids. They tend to give high
circulating pressures that can require additional compression equipment in the circulating system.
High levels of energy stored in the drillstring can lead to disastrous overspeeding of the motor, if
the bit is pulled off bottom without first venting the drillstring pressure. Downhole motors have
recently been (and are being) developed specifically for operation with compressible fluids,22
capable of generating penetration rates that match those in rotary drilling.21
Excessive Hydrocarbon Production
Well control concerns are fundamentally not a limitation on underbalanced drilling. Because the
formation fluids are not prevented from flowing into the wellbore by the drilling fluid, as they are
in conventional drilling, different well control practices and procedures are required. Under most
circumstances, suitable surface equipment can contain and control produced fluids while drilling
underbalanced. High hydrocarbon production rates and high pressures are desirable from the
point-of-view of the long term profitability of the well. They can, however, prevent, or at least
complicate, some underbalanced drilling operations. The surface equipment should be able to
safely handle the maximum rate of production. It should also be able to contain the maximum
probable surface pressure, which could be substantial. If excessive production rates are
encountered, there may be little alternative but to kill the well and switch to overbalanced
drilling.
Economic Factors
While it may be technically possible to drill a well underbalanced, it may not always be
economical. Factors that can prevent under-balanced drilling from being cost effective include
large water inflows, good penetration rates or productivity with conventional drilling techniques,
and local logistics.

1-14

In many locations, environmental restrictions make produced water disposal expensive. With
large water inflows, disposal costs can negate any reductions in well cost associated with
underbalanced drilling.
The increased penetration rate due to drilling underbalanced may not always reduce the drilling
cost. If the penetration rate with mud drilling is already quite high, (for example, 50 ft/hr or
more) or if only a short interval is to be drilled underbalanced, the total drilling time for that
interval may not be decreased sufficiently to pay for the additional mobilization and daily costs
associated with the underbalanced drilling equipment.
Similarly, if the well productivity is high when drilled conventionally, there may be little to be
gained by drilling underbalanced. At the other extreme, there are many reservoirs which have
such low undamaged permeability that they would have to be stimulated by hydraulic fracturing
even if drilled underbalanced (presuming high ROP when drilling conventionally).
Finally, in some areas, it can be uneconomic to drill underbalanced if the required equipment and
materials, such as com-pressors, boosters, foaming agents, etc., are not available locally and the
cost of their mobilization or transport exceeds the benefits of drilling underbalanced.

1.5

Summary

This chapter has introduced the different techniques that can be used to generate underbalanced
conditions while drilling. The benefits of drilling underbalanced can be considerable, in terms of
reduced drilling cost and increased productivity. However, underbalanced drilling is not suitable
for all wells. It requires special equipment and procedures, not used in conventional drilling
operations. A number of technical and economic factors limit the application of underbalanced
drilling. Sometimes, the limitations seem to be very daunting. However, it will often turn out
that the benefits outweigh the disadvantages.
Chapter 2 provides, more detail on the different techniques of underbalanced drilling, describing
how the underbalance pressure is generated, summarizes the concepts for predicting circulating
pressures, cuttings and liquid lift capacities, and highlights the regimes of application and
specific limitations for the various specific techniques.

Return to SC Manual

1-15

References
1.

Carden, R.S: Technology Assessment of Vertical and Horizontal Air Drilling Potential in
the United States, U.S. Department of Energy Report No. DOE/MC/28252-3514
(DE94000044), (August 1993).

2.

Lattimore, G.M., Carden, R.S. and Fisher, T.: Grand Canyon Directional Drilling and
Water Line Project, paper SPE 16169 presented at the 1987 SPE/IADC Drilling
Conference, New Orleans, LA.

3.

Bennion, D.B. and Thomas, F.B.: Underbalanced Drilling of Horizontal Wells: Does it
Really Eliminate Formation Damage? paper SPE 27352 presented at the 1994 SPE
International Symposium on Formation Damage Control, Lafayette, LA.

4.

Bennion, D.B., Thomas, F.B., Bennion, D.W. and Bietz, R.F.: Formation Damage Control
and Research in Horizontal Wells, presented at 1993 International Conference on Horizontal
Well Technology, Houston, TX.

5.

Graham, R.L., Foster, J.M., Amick, P.C. and Shaw, J.S.: Reverse Circulation Air Drilling
Can Reduce Wellbore Damage, Oil and Gas J. (March 22, 1993) 85-94.

6.

Brantley, J.E.: History of Oil Well Drilling, Gulf Publishing Company, Houston, TX (1971).

7.

Allan, P.D.: Nitrogen Drilling System for Gas Drilling Applications, paper SPE 28320
presented at the 1994 SPE Annual Technical Conference and Exhibition, New Orleans, LA
September 25-28.

8.

Tian, S. and Adewumi, M.A.: Multiphase Hydrodynamic Model Predicts Important


Phenomena in Air-Drilling Hydraulics, SPE Drill. Eng. (June 1991) 145-152.

9.

Angel, R.R.: Volume Requirements for Air and Gas Drilling, Pet Trans., AIME (1957)
210, 325-220; also Volume Requirements for Air and Gas Drilling, Gulf Publishing Co.,
Houston, TX (1958).

10.

Johnson, P.W.: Design Techniques in Air and Gas Drilling: Cleaning Criteria and
Minimum Flowing Pressure Gradients, J. Cdn. Pet. Tech. (May 1995) 34, No. 5, 18-26.

11.

Gray, K.E.: The Cutting Carrying Capacity of Air at Pressures above Atmospheric, SPE
paper 874G, Pet. Trans. AIME (1958) 23.

12.

Hook, R.A., Cooper, L.W. and Payne, B.R.: Air, Gas and Foam Drilling: A Look at Latest
Techniques, World Oil (April 1977) 95-103.

1-16

Chapter 1

1-17

Introduction

13.

Beyer, A.H., Millhone, R.S. and Foote, R.W.: Flow Behaviour of Foam as a Well
Circulating Fluid, paper SPE 3986 presented at the 1972 SPE Annual Fall Meeting, San
Antonio, TX.

14.

Stone, C.R.: Horizontal Underbalanced Drilling, SPE Distinguished Lecture Series, 199394.

15.

Okpobiri, G.A. and Ikoku, C.U.: Volumetric Requirements for Foam and Mist Drilling
Operations, SPE Drill. Eng. (February 1986) 71-88.

16.

Hale, A.H., Mody, F.K. and Salisbury, D.P.: The Influence of Chemical Potential on
Wellbore Stability, SPEDC (September 1993) 207.

17.

Grace, R.D. and Pippin, M.: Downhole Fires During Air Drilling, World Oil (May 1989)
42-44.

18.

Shale, L.: Underbalanced Drilling Equipment and Techniques, presented at 1995 ASME
Energy Technology Conference, Houston, TX (January 30 - February 1).

19.

Carden, R.S.: Air Drilling has some Pluses for Horizontal Wells, Oil and Gas J. (April 8,
1991) 76-78.

20.

Shale, L. and Moberley, G.T.: Development of a Cartridge Data Transmission System for
Use with Air Drilling Motor, IADC/SPE paper 23907 presented at the 1992 IADC/SPE
Drilling Conference, New Orleans, LA.

21.

Shale, L. and Curry, D.A.: Drilling a Horizontal Well Using Air/Foam Techniques, paper
OTC 7355 presented at 1993 Annual Offshore Technology Conference, Houston, TX.

22.

Shale, L.: Development of Air Drilling Motor Holds Promise for Specialized Directional
Drilling Applications, paper SPE 22564 presented at the 1991 SPE Annual Technical
Conference and Exhibition, Dallas, TX.

UNDERBALANCED
DRILLING TECHNIQUES

his chapter provides detailed descriptions of the different techniques of underbalanced


drilling. The major function of the circulating drilling fluid in underbalanced drilling is to lift
cuttings from the hole. This aspect of each technique is considered in some detail. Methods for
analyzing hole cleaning and circulating pressures are reviewed. In each case, the required
equipment is described. Any special operating procedures that may have to be adopted are
described, as are any limitations.

2.1
2.1.1

Dry Air Drilling


Hole Cleaning

It will rapidly become impossible for the drill bit to deepen a hole, if the cuttings it generates are
not removed from the wellbore. In air drilling, the main function of the circulating air is to lift
cuttings from where they are generated at the hole bottom to the surface and out of the wellbore.
If the air flow is not adequate to do this, there is a real danger that the drillstring will become
stuck by cuttings that settle back downhole and pack around the bottomhole assembly (BHA)
when circulation is stopped (for example, to make a connection). A key concern, in any air
drilling job, is to determine the air flow rate required for adequate hole cleaning. As a result, the
processes of cuttings transport and hole cleaning have been studied quite extensively. These
processes are reviewed below.
How does the air lift cuttings from the hole? The flowing air exerts a drag force on each
individual cutting, opposing gravity. If the drag force is larger than the gravitational force, the
cutting will move up hole. Conversely, if the drag force is too small, the cutting will fall back
down hole. Intuitively, it is clear that the drag force will increase as the air flow rate past the
cutting increases. The gravitational force on the cutting will not be influenced by the air flow
rate. There should, therefore, be some threshold air flow rate at which cuttings begin to move up
hole. As the flow rate increases, the rate at which the cuttings move upwards should also
increase.
It is also intuitively clear that the air velocity required to lift a cutting will increase as the cutting
size increases. Day-to-day experience tells us that dust is frequently picked up and carried along
by even modest winds, whereas it is an unusually strong wind that can move sand, grit or loose
2-1

Chapter 2

Underbalanced Drilling Techniques

gravel. Cuttings, recovered at the surface in air drilling operations, are often almost exclusively
very fine. In fact, dry air drilling is sometimes termed dusting. These are probably not
representative of the cuttings generated by the bit. Laboratory air drilling experiments have
shown that cuttings with dimensions of one-half inch or more are formed, although even under
these conditions the majority of cuttings are fine enough to be classified as dust. Johnson, 1995,1
reported collecting cuttings that were close to one-half inch in maximum diameter, from two
different shallow air drilled wells. Other authors2,3 have collected air drilled cuttings as large as
one inch in dimension, in junk baskets located at the bit and at the top of the collars. It seems
reasonable to conclude that these large cuttings normally stay downhole, possibly accumulating
at the top of the collars, until they are broken down to a sufficiently small size that the air flow
can lift them from the well.
The process of cuttings transport is complicated by the compressibility of air. As air flows up the
annulus, the frictional pressure drop increases the air pressure downhole. So too does the mass
of cuttings and air in the annulus, and the air density increases in direct proportion to its pressure.
In consequence, the air velocity decreases with increasing depth, provided that the annular
geometry remains unchanged. The drag force on the cuttings increases with increasing air
pressure and decreases with decreasing air velocity. The net effect is that a higher air flow rate,
when expressed in standard cubic feet per minute, is required to lift cuttings as a well becomes
deeper. Air temperature also influences density, and the temperature will also change as the air
flows around the well.
Strictly, there is a buoyant force on the cutting due to the displacement of its volume of air. This
will be influenced by the air flow rate since the flow rate influences the local air pressure and
hence the density. However, this buoyant force is, small for cuttings in air even when the air
pressure is quite high, since the air density remains much less than that of the cuttings.
At high air flow rates, the cuttings move at more or less the same speed as the air and are lifted
efficiently out of the well. Under these circumstances, the bottomhole air pressure is largely
controlled by the frictional pressure loss up the annulus. If the flow rate is decreased, the
frictional pressure loss will fall and initially the bottomhole air pressure will also decrease. As
the air flow rate is further decreased, the efficiency of cuttings removal decreases. This causes
the mass of cuttings in the annulus to increase. At some point, the increase in air pressure due to
cuttings accumulation, with falling air flow rate, exceeds the decrease in frictional pressure drop,
and the air pressure downhole will actually increase with decreasing air flow rate. The pressure
then increases rapidly as the flow rate is reduced until the air flow is no longer capable of
supporting the cuttings. This phenomenon is sometimes referred to as choking and is well
recognized in the pneumatic transport literature.4

2-2

In the context of air drilling, choking flow has been studied experimentally by Supon and
Adewumi, 1991,5 using a 3-inch internal diameter, 26-foot long model wellbore. Figure 2-1
summarizes their findings. The choking velocity is defined as the velocity below which the
cuttings are not supported by the air flow. Supon and Adewumi suggested that the air velocity
leading to the minimum bottomhole air pressure represents the optimum flow rate for cuttings
transport. They found that this optimum velocity increased, with increasing cuttings diameter
and penetration rate.

Frictional
resistance
predominates
Minimum
Pressure Drop

Choking
Velocity

Optimal
Air Velocity

Annulus Pressure Drop

Annulus Pressure Drop

Static head
predominates

Annulus Air Velocity

Figure 2-1.

The influence of air flow rate on annular pressure drop (after Supon and Adewumi,
19915).

An alternative way of expressing this is that the minimum annular pressure drop increases with
increasing cuttings size and penetration rate. Interestingly, it was the diameter of the largest few
percent of cuttings that seemed to control the so-called optimum air velocity, as opposed to an
average cuttings diameter.
Deviated Holes
In a deviated hole, the situation is complicated by friction between the cuttings and the borehole
wall. The drag force must exceed the combined effect of gravity and friction, for a cutting to
move up hole. Conversely, the gravitational force must exceed the combined effect of the drag
force and friction if the cutting is to move downhole. There will be a regime in which a more or
less stationary bed of cuttings will build on the lower side of an inclined hole. Hagar et al., 1995,6
extended earlier cuttings transport experiments, to examine modestly inclined wellbores. The
optimum air velocity increased with increasing wellbore inclination, at least up to an inclination
of 12, the highest angle studied. This increase in air flow requirement was ascribed to friction
between the cuttings and the borehole wall. They observed that the bed of cuttings would be
removed when the air flow rate was increased to a level corresponding to the minimum pressure
drop. Shale, 1995,7 noted that higher air flow rates are required to efficiently clean a deviated
hole drilled using a downhole motor, if the drillstring is not rotated. Drillpipe rotation aids hole

2-3

Chapter 2

Underbalanced Drilling Techniques

cleaning by grinding up the cuttings. This reduces the air velocity required to transport them.
Rotation also agitates the cuttings, restricting the formation of a cuttings bed.
The Annular Geometry
The annular geometry of a well being drilled influences cuttings transport. The lowest portion of
the drillstring, i.e. the bottomhole assembly (BHA), almost invariably has larger diameter
components (drill collars) than the drillpipe above it. Thus, the air velocity drops significantly as
it flows past the top of the drill collars. This is the most demanding region for cuttings transport.
It is here that cuttings will accumulate first, if the flow rate is not sufficient to remove all cuttings
from the well. Because the air velocity is proportional to the square of the wellbore diameter,
even modest increases in wellbore diameter (washouts ... ) can reduce the air velocity sufficiently
to degrade cuttings transport efficiency.
Gas Inflows
Significant gas inflows can also degrade cuttings transport efficiency, if they occur above the
BHA. Such inflow increases the flow rate from the point of influx to the surface, causing the
annular pressure drop and the air pressure below the influx to increase. The increased air
pressure reduces the air velocity lower down the well and cuttings are not lifted as efficiently in
the region below the influx.
The foregoing discussion should illustrate the main factors involved in the transport of cuttings
out of an air drilled wellbore, and the ways in which they interact. It should be clear that cuttings
transport and circulating air pressure are closely linked. Both must be considered in any air
circulation model.
Drag Force and Terminal Velocity
In principle, it is possible to model the cuttings removal process by computing the drag force on
the cuttings. This is probably best done by considering the terminal or free settling velocity, Vt,
of the cuttings. This is the maximum velocity that is attained by a particle falling freely in an
infinite quantity of the fluid in question. In a vertical hole, if the upward air velocity is Vf, the
cuttings velocity, Vc, will be:
Vc = Vf Vt

(2.1)

The higher the cuttings terminal velocity, the higher the air velocity that will be required to
transport the cuttings. Several authors have predicted minimum volumetric air flow rates on the
basis that the air velocity downhole should be at least equal to the cuttings terminal velocity.8,9
A knowledge of cuttings terminal velocity is therefore important for fundamentally-based models
of cuttings transport during air drilling. The fluid dynamics literature contains a number of
different correlations for determining terminal velocities. The following discussion is based on
an experimental study conducted by Gray, 1958.8 For a spherical particle of diameter, dc, and
density, c, falling through a fluid of density, f, the terminal velocity is given by:

2-4

Vt = 4gd c

c f
3C d f

(2.2)

where:
g......... gravitational acceleration
(32.17 ft/s2),
dc ........ characteristic particle diameter
(feet),
Cd ....... drag coefficient,
c ....... density of cuttings (lbm/ft3), and,
f ........ density of fluid (lbm/ft3).
The drag coefficient is, in general, a function of the Reynolds number, which is in turn
determined by the velocity of the cutting relative to the fluid through which it is moving.
However, Gray found that the drag coefficient of a cutting can safely be regarded as a constant,
independent of velocity, if the air flow around the cutting is turbulent. This will normally be the
case in air drilling. On the other hand, the drag coefficient was found to be influenced by the
particle shape. Flat particles had drag coefficients of about 1.4, while particles classified as
angular to sub-rounded had drag coefficients of about 0.8. Combining these findings with the
gas law, Gray derived the following approximate expressions for terminal velocity:
Vt 3.369

d c Tc
P

for flat cuttings, and


d c T c
Vt 4.164
P

(2.3)

(2.4)

for sub-rounded cuttings.


where T and P are the bottomhole temperature (R) and pressure (psia), respectively. The
terminal velocity of cuttings in air drilling is determined mainly by the cuttings diameter, shape
and density, and by the bottomhole temperature and pressure. The above expressions do not
account for interactions between adjacent cuttings or between cuttings and the borehole wall.
Since these interactions should tend to reduce the terminal velocity, it is probably conservative to
discount their effects when modeling cuttings removal.

2-5

Chapter 2

Underbalanced Drilling Techniques

I t can be seen from Equations (2.3) and (2.4) that the terminal velocity of cuttings will increase
as their diameter increases. Higher air velocities (i.e., higher volumetric flow rates) are required
to lift larger cuttings from a well. Rounded cuttings will require more air flow to remove them
than flat cuttings.
The air pressure and temperature downhole vary around the well, and differ from those at the
surface. As the air pressure goes up, the cuttings terminal velocity goes down, indicating that the
local air velocity required to lift cuttings will also decrease. However, air density increases more
or less in direct proportion to the pressure. The local air velocity will be inversely proportional to
the air pressure, if the mass flow rate of gas remains constant (as it will be if the flow rate at
standard temperature and pressure is constant). The terminal velocity only decreases as the
square root of the pressure. As a result, the air flow rate required to lift a given size of cutting
will increase as the bottomhole air pressure increases.
Friction Pressure
The pressure at the hole bottom is increased by the frictional pressure drop as the cuttings laden
air flows up the annulus and out of the surface equipment, and by the mass of air and cuttings in
the annulus. A further pressure drop occurs if the cuttings are being accelerated. It is usually
sufficient to assume that the cuttings have reached a steady velocity and that the acceleration
term can be neglected. The pressure gradient, dP/dL (psf/ft), in the annulus can then expressed
by:
f m m v 2m
dP

= m +
dL
2g ( D h D p )

(2.5)

where:
fm ....... friction factor of the mixture of air
and cuttings,
m....... mixtures density (lbm/ft3),
vm ....... mixtures velocity (ft/s),
Dh....... hole diameter (feet), and,
Dp....... drillstring outer diameter (feet).
The friction factor of the air/cuttings mixture is commonly regarded as the sum of the friction
factors due to the air alone, fa, and due to the presence of the cuttings, fc:
f m = fa + fc

(2.6)

There are a number of expressions that predict the frictional pressure drop for gas flow in pipes.
The best known of these expressions is probably the Weymouth equation.10 Whichever frictional
analysis is used, it has to be modified for the annular geometry of a well being drilled. This is

2-6

usually done through the concept of the hydraulic radius, which is defined as the cross-sectional
area of the conduit divided by its perimeter. The hydraulic radius of an annulus is 0.25(Dh - Dp),
where Dh is the diameter of the hole and Dp is the outer diameter of the body (in this case the
drillpipe) inside that hole. The hydraulic radius of a pipe of internal diameter D is simply 0.25D.
When modified in this way, Angel, 1957,11 showed that the friction factor determined from the
Weymouth equation is:
fa =

0.14
( D h D p ) 0.333

(2.7)

This does not strictly apply to the case of flow in rough-walled pipes, and as argued by Guo et al.,
1994,12 a borehole wall will appear rough to the circulating air flow. As air flow during drilling
is nearly always turbulent, Guo et al., 1994, argued that the friction factor correlation developed
by Nikuradse, 1933,13 for turbulent flow in rough pipes, is more applicable to air drilling. For
annular flow, this correlation is:
2
1

= 114
. 0.86 ln
fa
Dh Dp

(2.8)

where:
......... the absolute roughness of the pipe
(feet).
The component of friction due to the presence of cuttings in the air flow was studied by Machado
and Ikoku, 1981,9 who experimentally determined friction factors for the cuttings, fc.
The density of the mixture of air and cuttings in the annulus is determined by the mass of the
cuttings and the density of the air. The mass of cuttings is itself controlled by the penetration rate
(the rate at which cuttings are generated) and by the efficiency with which the cuttings are
removed from the well. As the penetration rate increases, the mass of cuttings injected into the
airflow downhole increases. If these cuttings are not removed efficiently from the wellbore, the
mass of cuttings in the annulus will increase, as will the density of the air and cuttings mixture.
The density of the air is controlled by the local pressure; increasing pressure increases the air
density.
The additional circulating pressure drops down the drillstring and across the bit are relevant to
cuttings transport because they can influence the air temperature. Temperature changes result
from heat exchange with the cuttings, the formation and (through the drillstring) with air flowing
in the opposite direction, from frictional heating, and from adiabatic effects as the pressure
changes around the well.

2-7

Chapter 2

Underbalanced Drilling Techniques

Required Air Injection Rates


It should be clear from this discussion that relating the air velocity downhole to the surface
injection rate is not trivial. Prediction of surface air injection rates required for efficient cuttings
removal, by comparing air velocity and cuttings terminal velocity, is not readily accomplished.
Nevertheless, several methods for so doing have been presented in the petroleum
literature.9,14,15,16,17
All of these analyses have the same major problem when they come to be applied to field
operations. They require knowledge of the cuttings shape and size. This is rarely available.
They also require knowledge of hole geometry. The annular velocity will fall markedly, and with
it the cuttings transport efficiency, opposite any washouts (sections of enlarged hole diameter).
An alternative analysis of cuttings transport forms the basis of what is probably the most widely
used method to predict air injection rate requirements. This was presented by Angel, 1957.11
Angel assumed that, for efficient cuttings transport downhole, the kinetic energy of the air (or
gas) striking each cutting should be the same as that of air giving efficient cuttings transport at
standard pressure and temperature, i.e.:
1
1
min v 2min = stp v 2 stp
2
2

(2.9)

where:
min .... density of air (or gas) at the
minimum required downhole
injection rate (lbm/ft3),
vmin ..... air (or gas) velocity downhole
(ft/min),
stp...... density of air at standard pressure
and temperature (lbm/ft3), and,
vstp ...... minimum air velocity for efficient
cuttings transport, at standard
pressure and temperature (ft/min).
Alternatively, this can be expressed as:
v min = v stp

2-8

stp
min

(2.10)

It was seen earlier that a cuttings terminal velocity is more or less inversely proportional to the
square root of the density of the gas through which it is falling. Angels analysis indicates the
same dependence of minimum air velocity on downhole pressure as do the analyses based on
terminal velocity.
Experience from shallow blast holes, drilled in limestone quarrying operations, indicated that
cuttings were transported efficiently if the air velocity equaled or exceeded 3,000 feet per minute.
In choosing this cuttings transport criterion, Angel implicitly assumed that all cuttings lifted from
a well would have the same size and shape as these limestone cuttings that were transported
efficiently at 3,000 ft/min air velocity. Johnson, 1995,1 has shown that this corresponds to Grays
(1958)8 value for terminal velocity; for flat cuttings with a major diameter of 0.46 inches, and for
sub-rounded particles with a diameter of 0.26 inches.
Angel computed the air pressure downhole using Equation (2.5), assuming that the cuttings
traveled uphole at the same speed as the air (no slippage of the cuttings relative to the airflow).
The density of the mixture of air and cuttings is given by:
w
m = a 1 + c
wa

(2.11)

where:
wc ....... mass of cuttings generated in a
given time; the mass flow rate of
cuttings (lbm/min), and,
wa ....... mass of air flowing past any point
in the well in a given time; the
mass flow rate of air (lbm/min).
If the cuttings do slip, the mixture density will be higher than predicted by Equation (2.11), by an
amount that will increase as the rate of slippage increases. This means that Angels analysis will
predict lower downhole pressures than would be observed in practice, and that the discrepancy
will become greater as the air flow rate is decreased and choking starts to occur. As shown by
Supon and Adewumi, 1991,5 Angels analysis fails to predict a minimum bottomhole pressure
(refer to Figure 2-1); instead it gives a pressure that decreases monotonically with decreasing air
flow rate. If air velocity downhole is sufficient to give reasonable cuttings transport (that is
limited cuttings slippage), this does not lead to major practical concern.
Angel used Weymouths friction factor to compute bottomhole pressure. As noted above, this
strictly applies for flow in smooth pipes. Borehole walls are normally far from smooth. Annular
pressure drop tends to be higher than Angel computed. The tacit assumption that the cuttings do
not contribute to the frictional pressure drop will also tend to make the predicted pressure drop
lower than actually seen. Other assumptions made were that the annular geometry is

2-9

Chapter 2

Underbalanced Drilling Techniques

appropriately represented by a pipe with the same hydraulic diameter as the annulus (as in
Equation (2.8)), that interactions between cuttings can be neglected, that the surface air
temperature is 80F, that the downhole air temperature increases by 1F per 100 feet of depth,
and that the cuttings have a specific gravity of 2.70.
The expression Angel obtained for the annulus air pressure at the hole bottom, Pb (lbf/ft2), is:
where:
Ps........ surface air pressure
(lbf/ft2, absolute),
Ts ....... surface temperature (F), and,
G ........ annular temperature gradient,
(assumed to be 1F per 100 feet).
T, the downhole temperature, is given by:
T = Ts + Gh

(2.13)

where:
h......... hole depth (feet).
a is given by:
a=

SQ + 28.8 ROP D 2h
53.3 Q

(2.14)

where:
S......... gas specific gravity (air = 1),
Q ........ gas flow rate (scfm), and,
ROP ... rate of penetration (ft/hr).
b is given by:
b=

1.625 10 6 Q 2
(Dh Dp )

1.333

( D 2h D 2p ) 2

where:
Dh....... hole diameter (feet), and,
Dp....... drillpipe diameter (feet).
2-10

(2.15)

This was combined with the cuttings transport criterion defined in Equation (2.10) to deduce the
minimum air (gas) flow rate, as a function of hole depth, annular geometry and penetration rate.
Angel simplified the analysis by using the average downhole temperature, Tav, when calculating
the downhole pressure. The resultant expression is:
6.61S (Ts + Gh) Q 2
=
2
( D 2h D 2p ) 2 v stp
(2.16)
( P + bT ) e
2
s

2
av

2 ah Tav

bT

2
av

This was solved numerically for the gas injection rate required to give an annular velocity
equivalent in cuttings lifting power to air with a velocity of 3,000 ft/min. A series of charts was
generated for different combinations of hole size, drillpipe diameter and penetration rate. These
charts are commonly used in estimating air flow rate requirements.
It is possible to approximately represent the recommended minimum injection rate, Qmin, by the
following expression:
Q min = Q o + NH

(2.17)

where:
Qo....... injection rate (scfm) at zero depth
that corresponds to an annular
velocity of 3,000 ft/min,
N ........ factor dependent on the
penetration rate, (refer to
Appendix C), and,
H ........ hole depth (thousands of feet).
Differences between the minimum injection rate estimated with Equation (2.16) and the exact
solution of Angels analysis are considerably less than 10 percent for all except the smallest hole
sizes at high penetration rates.
Angel also calculated bottomhole pressures for air and gas, as functions of depth and penetration
rate, at a gas injection rate corresponding to a surface annular velocity of 3,000 ft/min. The
pressure depends on the annular geometry, through the hole and drillpipe diameters. The
predicted pressures for some typical cases are shown in Figure 2-2. At 12,000 feet hole depth,
2a

Pb =

2 abT T G abT 2
Ps
+

G a Ts
Ga
2
s

(2.12)
2-11

Chapter 2

Underbalanced Drilling Techniques

they range from 40 psi, for an 11-inch hole drilled with 5-inch pipe at a zero penetration rate, to
over 200 psi, for a 4-inch hole drilled with 2 7/8-inch pipe and a penetration rate of 90 feet per
hour.
Grace18 measured bottomhole annular pressure in an air drilled hole and compared his
measurements with values predicted by Angels analysis. A 7 7/8-inch hole was being drilled
below 3800 feet, with 3-inch drillpipe and 6-inch drill collars. A caliper log had been run to
determine actual hole size prior to starting the experiment. Initially, pressure losses were
measured at various flow rates, while circulating with the bit off bottom, without cuttings in the
air stream. The results of these measurements are presented in Table 2-1, together with the
predicted pressures. At all flow rates studied, the predictions were about one psi less than the
actual measured values.
Bottomhole pressures were then measured while drilling, to determine how cuttings affected the
circulating air pressure in the annulus. The penetration rate was held constant at approximately
30 feet per hour, while air flow rates were changed from 600 to 1,200 scfm. The pressures
measured at a depths of 3,812 feet and 4,488 feet are plotted, as functions of the air injection rate,
in Figure 2-3. Figure 2-3 also shows predictions made using Angels method ology. Not
surprisingly, the measured pressures were all somewhat higher than those predicted, the
differences ranging from five to fourteen psi. All other factors being equal, it would be expected
that the pressure should increase with increasing depth. It is interesting to note that the pressures
measured at higher air flow rates were lower at 4,488 feet than they were at 3,812 feet. The
cause of these anomalous observations is not certain. It is plausible, however, that there was a
change in cuttings size or shape, perhaps resulting from a change in formation type. This would
lead to a change in the concentration of cuttings in the annulus. If the cuttings were smaller, they
would be lifted more efficiently from the hole and this would reduce the bottomhole pressure. It
is also possible that there was a washout close to the hole bottom. This would reduce the local
annular velocity, reducing the frictional pressure losses in the washed out zone. At the same time
the cuttings concentration in the air flow would increase at the washout, causing an increase in
bottomhole pressure. Without details of the hole geometry and the cuttings size and shape, it is
not possible to determine the relative impact of these different factors on the air pressure.

2-12

250

Bottomhole Pressure (psia)

4 3/4-Inch Hole and 2 7/8-Inch Pipe


Air (0 ft/hr)

200

Gas (S=0.6), (0 ft/hr)


Air (90 ft/hr)
Gas (S=0.6), (90 ft/hr)

150

100

50

0
0

2000

4000

6000

8000

10000

12000

Depth (feet)

80
11-Inch Hole and 5 1/2-Inch Pipe

Bottomhole Pressure (psia)

70
Air (0 ft/hr)
Gas (S=0.6), (0 ft/hr)

60

Air (90 ft/hr)


Gas (S=0.6), (90 ft/hr)

50
40
30
20
10
0
0

2000

4000

6000

8000

10000

12000

Depth (feet)

Figure 2-2.

Bottomhole pressures predicted for injection rates equivalent to 3,000 feet per minute
standard air velocity (after Angel, 1957 11).

2-13

Chapter 2

Underbalanced Drilling Techniques

Table 2-1.

Annular Bottomhole Pressures in an Air Drilled Hole - Comparison of


Predictions and Measurements Made While Circulating Off-Bottom.

Flow Rate,
(scfm)

Measured Pressure,
(psia)

Calculated Pressure,
(psia)

589.79

19.33

18.40

699.52

20.34

19.27

771.92

21.05

19.90

856.96

21.85

20.69

1007.34

23.26

22.10

1238.52

25.43

24.76

These measurements show that some of the assumptions made in Angels analysis are nonconservative. They lead to under-prediction of downhole air pressures, whether or not cuttings
are present in the annulus. Guo et al., 1994,12 overcame what is probably the most significant of
these simplifications by incorporating Nikuradses friction factor into Angels analysis. This is
more appropriate for flow through a rough-walled borehole than Weymouths factor. They also
introduced varying hole inclination into the analysis.
Summary
It should be apparent from this discussion that there are a number of different methods of
analyzing cuttings transport and circulating pressures in air drilling. These analyses tend to give
somewhat different recommendations for air flow requirements. Figure 2-4 compares the
recommendations from several available analyses with injection rates found from experience to
give adequate hole cleaning.12 There is no uniquely correct flow rate for any one well. Neither
is there a universally correct way to specify air flow rate. There will be a minimum flow rate
below which the well will choke, and increasing the flow rate will lift progressively larger
cuttings from the well. Those cuttings that are not lifted from the well are ground up until they
become fine enough for the air flow to lift them to the surface. As a result, the drilling process
may be slowed, but the increased cost of a lower penetration rate will be offset by the lower
compressor rental and fuel requirements; i.e., operational and economic factors will influence the
choice of air flow rate, in addition to cuttings transport considerations.

2-14

45

Bottomhole Pressure (psia)

Penetration Rate: 29 ft/hr


Depth: 3812 feet

40

35

30

25
Angel's Predicted Pressure
Measured Pressure

20
500

600

700

800

900

1000

1100

1200

1300

Flow Rate (scfm)

34

Bottomhole Pressure (psia)

32

Penetration Rate: 28 ft/hr


Depth: 4488 feet

30
28
26
24
Angel's Predicted Pressure
Measured Pressure

22
20
500

600

700

800

900

1000

1100

1200

1300

Flow Rate (scfm)

Figure 2-3.

Measured versus calculated bottomhole pressures (after Grace 18).

2-15

Chapter 2

Underbalanced Drilling Techniques

3.5
Angel
Schoeppel and Spare
Wolcott and Sharma
Macado and Ikoku
Guo, Miska and Lee
Amoco's Experience

Required Rate of Air (scfm)

2.5

1.5

0.5

0
0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

Depth (feet)

Figure 2-4.

2.1.2

Comparison of air flow rates recommended by several different cuttings transport


analyses (after Guo et al, 1994 12).

Circulating Pressures

In mud drilling, the standpipe pressure is a major source of information on downhole conditions.
The relationship between the standpipe and downhole pressures is easily established if the
drilling fluid is incompressible. If the drilling fluid has significant compressibility, as it does
when drilling with air, the relationship between these pressures is more complex. The standpipe
pressure is a result of the pressure beneath the bit, the pressure drop across the bit, and the
pressure change down the drillstring between the standpipe and the bit. The various analyses of
hole cleaning that were described in the preceding section all required computation of the annular
pressure downhole. This was done by evaluating the pressure loss as the air and cuttings flowed
uphole and discharged to atmospheric pressure at the blooie line exit.
Bit Pressure Drop
Air flow is restricted at the bit. In air drilling, it is common practice not to fit nozzles into the
bit. Even so, the empty jets, which are typically less than one inch in diameter, constrain the air
to flow through a passage with a much smaller cross-sectional area than the drillstring. As air
flows through the jet, it expands in response to the decreased pressure and its velocity increases.
Once the pressure drop exceeds a certain level, the air velocity reaches the prevailing speed of
sound. At this point, the air cannot expand any faster and the upstream pressure becomes
independent of the pressure into which the jet is discharging. This implies that, under sonic
discharge conditions, standpipe pressure is independent of the annular pressure. Flow through

2-16

the jet is then said to be critical or sonic. The condition for the onset of sonic flow in ideal gases
is:19
k

Pa 2 k 1
=

Pb k + 1

(2.18)

where:
Pb ....... downstream pressure (psia),
Pa ....... upstream pressure at the onset of
sonic flow (psia), and,
k......... ratio of the specific heat at
constant pressure to that at
constant volume (dimensionless).
For air, k = 1.40, and:
Pa = 189
. Pb

(2.19)

If the upstream pressure is more than 1.89 times the annulus pressure beneath the bit, flow
through the bit will be sonic. In this case, the pressure upstream of the bit, Pa (psia), will be
given by:20
G
Pa =
An

k +1

Ta R 2 1 k

Sgk k + 1

0.5

(2.20)

where:
G ........ mass flow rate of air in lbm/s,
An....... total area of the bit nozzles (in2),
Ta ....... air temperature above the bit (R),
R ........ the universal gas constant
53.3 ftlbf/lbmR for air),
S......... gas gravity (1 for air), and,
g......... gravitational constant (32.17 ft/s2).
Noting that the density of air under standard conditions is 0.0764 lbm/ft3, for air Equation (2.20)
reduces to:
GT 0 .5
QT 0.5
Pa = 1.88 a = 0.00239 a
An
An

2-17

Chapter 2

Underbalanced Drilling Techniques

(2.21)
where:
Q ........ air flow rate (scfm).
If the air flow velocity through the jets remains sub-sonic, the pressure above the bit is related to
the mass flow rate and the annulus pressure beneath the bit, Pb (psia), by:21
k

R( k 1)G 2 Tb k 1
Pa = Pb 1 +

2gkSA 2n Pb2

(2.22)

For air, this becomes:


0.236G 2 Tb
Pa = Pb 1 +

A 2n Pb2

G=

Qg

3.5

(2.23)

60

where:
Tb ....... temperature beneath the bit (R).
g ....... gas density at STP (lbm/ft3), and
air ...... 0.0764 at STP (lbm/ft3).
The pressure beneath the bit, Pb, can be computed following Angels analysis, using Equation
(2.12).
The circulating air cools as it expands through the flow restriction at the bit. Assuming ideal gas
behavior, the temperature decrease can be estimated from the relationship:
P
Tb = Ta b
Pa

k 1
k

(2.24)

This indicates that the absolute air temperature beneath the bit will be approximately 17 percent
lower than that above the bit, if flow through the jets is sonic.

2-18

Standpipe Pressure
Any one of several friction factors can be used to establish the relationship between pressure
above the bit and pressure at the standpipe. If the flowing air pressure above the bit has been
determined, this can be used to calculate the standpipe pressure. The change in pressure over a
small element of depth down the drillstring will be the difference between the pressure loss due
to friction and the pressure increase due to the change in hydrostatic head.
In this instance, the air is flowing down a more or less smooth pipe and there are no cuttings in
the air flow. It is reasonably accurate to take Weymouths equation to represent the friction
between the flowing air and the inner bore of the drillstring. If this is done, the standpipe
pressure, Ps (psfa), is related to the pressure above the bit by:22
Pa2 + Tav2 (e 2 h / Tav 1)
e 2 h / Tav

Ps =

(2.25)

where and are given by:


=

S
53.3

(2.26)

1625
.
106 Q 2
=
D5i .333
(2.27)
where:
Di ....... internal diameter of the drillstring
(feet).
The standpipe pressure can be determined by first assessing whether or not flow through the bit is
sonic. If it is sonic, the pressure above the bit can be determined from Equation (2.21). This
value can be used in Equation (2.25) to predict the standpipe pressure. If the flow through the bit
is sub-sonic, the annulus pressure beneath the bit must first be predicted. This can be done,
following Angels analysis, by using Equation (2.12). The pressure above the bit can then be
predicted using Equation (2.23) and this value can be used in Equation (2.25) to arrive at the
standpipe pressure.

2-19

Chapter 2

Underbalanced Drilling Techniques

A very important point to note when air drilling is that large changes in annulus pressure may
result in smaller changes in standpipe pressure, or in the case of sonic flow through the bit, no
change in standpipe pressure at all. Consequently, hole problems that lead to an increase in
annulus pressure may be indicated by little or no change in standpipe pressure. For this
reason, it is important to monitor standpipe pressures closely and to react promptly to any
unanticipated changes. It is also important to know whether or not flow downhole is sonic. If
it is sonic, the standpipe pressure will not change in response to hole problems.
This is illustrated with an example. An 8-inch hole is being drilled at 6,000 feet with 4-inch
drillpipe. The air rate is 1,400 scfm. For simplicity, it is assumed that there are no collars in the
string. Standpipe pressures have been predicted following the method described above; for a
range of penetration rates up to 300 feet per hour, and for a bit equipped with no nozzles and
with three 14/32-inch diameter nozzles. Table 2-2 and Figure 2-5 show the resulting predictions.
As shown in Figure 2-5, for sub-sonic flow through a bit without nozzles, as the penetration rate
increases from zero (i.e. off-bottom) to 300 feet per hour, the bottomhole annular pressure
increases from 33 psia to 88 psia. At the same time, the standpipe pressure increases from 94
psia to 114 psia. A 55 psi (164 percent) increase in annular pressure has caused an increase of
only 21 psi (22 percent) in the standpipe pressure. For a bit with nozzles, flow through the bit is
sonic and the standpipe pressure does not change at all in response to the increasing annular
pressure.
Inadequate hole cleaning can increase annular pressures, by increasing the quantity of cuttings in
the annulus, exactly as would happen when the penetration rate increases. There may be little or
no increase in the standpipe pressure to indicate the onset of hole cleaning problems. The fact
that sonic flow effectively isolates the standpipe pressure from the annulus should be considered
very carefully before deciding to fit nozzles to the bit for dry air drilling.
2.1.3
Equipment
Air Compression System
Figure 2-6 shows a typical layout of an air compression system for dry air drilling. The major
components are described below.
Compressors
As their name indicates, these are the primary means of compressing air to the pressure required
to circulate it round the well. Several different types of compressor units are available - rotary
vane, straight-lobe, reciprocating, and rotary screw. Of these, the reciprocating and rotary screw
type are the most widely used for drilling applications. Local availability of compressors, with
the necessary delivery rate and pressure, tends to control compressor selection. For example,
both reciprocating and rotary screw types are commonly used in the Arkoma Basin and in West
Texas, whereas rotary screw compressors predominate in the Appalachian Basin.

2-20

An internal diesel engine provides power to drive the air compressing unit itself. Specially
silenced compressors exist. These units are capable of meeting EPA noise restrictions for
industrial compressors (76 dBA at 22 feet) and have sufficiently high delivery rate and operating
pressure capacities for oil and gas drilling applications. Consideration should be given to the
use of these units wherever rig-site noise levels may be a concern.
Table 2-2.

Predicted Bottomhole Annular and Standpipe Pressures at Various Penetration Rates


in a 6,000 Foot Dry, Air Drilled Hole.

Penetration Rate
(ft/hr)

Annular Pressure
(psia)

Standpipe Pressure Without


Nozzles
(psia)

Standpipe Pressure With


Nozzles
(psia)

33

93

167

30

38

94

167

60

43

95

167

120

53

98

167

180

64

103

167

240

75

108

167

300

88

114

167

2-21

Chapter 2

Underbalanced Drilling Techniques

180
160
Bottomhole Annular Pressure
Standpipe Pressure (Without Nozzles)

140

Pressure (psia)

Standpipe Pressure (With Nozzles)

120
100
80
60
40

Depth: 6000 feet


Air Drilled

20
0

50

100

150

200

250

300

Penetration Rate (ft/hr)

Figure 2-5.

Predicted standpipe pressures as functions of penetration rate, for a bit with and
without nozzles, in a 6000 foot deep, air drilled hole.

TO
KELLY
HOSE

Figure 2-6.

2-22

A typical layout of an air compression system, (modified from Cooper, et al., 1977 24).

Compressor output is usually expressed in terms of the free volume that the output air would
occupy under the prevailing input conditions. Delivery capacities of 750 to 1,000 cubic feet per
minute (cfm) are common in oilfield applications. Compressor output is sometimes expressed in
standard cubic feet per minute (scfm). This is the volume that the air delivered by the compressor
in one minute would occupy under standard conditions of temperature and pressure (STP, 60 F
and 14.7 psia). When expressed in scfm, the output decreases with increasing altitude and
temperature because of the accompanying reduction in density of the free air drawn into the
compressor. Assuming that air behaves as an ideal gas, the volume, V1, occupied by a given
quantity of air at pressure, P1 (psia), and temperature, T1 (F), is related to the volume, V0, at
standard pressure (14.7 psia) and temperature (60F or 520R) by:
14.7 (T1 + 460)
V1 = V0
(2.28)
520 P1
The air delivery rate, Q0, expressed in scfm, can be found from the free air delivery rate, Q,
(cfm), the ambient pressure, P (psia), and the temperature T(F), using:
Q0 = Q

520 P1
14.7(T1 + 460)

(2.29)

The influence of reduced ambient pressure on air delivery rate, due to operating at high
elevations, can be significant in some parts of the United States. Appendix A includes a table
showing normal atmospheric pressure as a function of elevation. As a rule-of-thumb,
atmospheric pressure decreases by 0.5 psi for each 1,000 feet of elevation increase. In the Rocky
Mountains, it is not unusual to have wells located at 6,000 feet above sea level, where the
ambient pressure is around 11.8 psia. At this elevation, a compressor rated at 1,000 scfm free air
delivery will deliver only 803 scfm, if the ambient temperature is 60F. The influence of
temperature on delivery rate is smaller, but not necessarily negligible. Considering the same well
location, if the ambient air temperature is 100F, the delivery rate will drop further to 745 scfm.
The wellsite elevation and ambient temperature should therefore be considered when
determining compressor requirements.
Wellsite elevation can have a further impact on compressor output because of its effect on the
power generated by the compressors diesel engine. Lyons, 1984,20 indicate that the power
output of internal combustion engines decreases linearly with increasing altitude. A normally
aspirated diesel engine will lose 22 percent of its sea-level power rating when operated at an
altitude of 6,000 feet; a turbo-charged engine will lose approximately 15 percent of its power
rating. This will be significant if the compressors are to be operated close to both their
volumetric delivery and pressure ratings.

2-23

Chapter 2

Underbalanced Drilling Techniques

Measurements of compressor delivery rates, made with an orifice meter during field operations,
have indicated that the delivery rate actually achieved by different compressors can vary between
50 and 95 percent of rated capacity. A common average is 70 to 75 percent of the rated inlet
capacity. The efficiency of the compressor is primarily a function of how well it has been
maintained. As a result, it is not possible to determine the discharge volume simply by
measuring the compressor rpm. An orifice meter is the only practical way to determine actual
volumes delivered to the standpipe.
Very often, two or more compressors are used to provide the required flow rate. Depending on
the daily rental rate for compressors, in comparison to the total daily drilling cost, there can be
advantages in having an extra compressor on site, in addition to those necessary to give the
desired flow rate. In this way, one compressor can be pulled out of service for maintenance
without impeding drilling operations.
Single stage compressors typically have a maximum discharge pressure of about 135 psi. Most
compressors that are used for air drilling are multi-stage (usually two-stage in the case of rotary
screw compressors). These have maximum discharge pressures that range from 250 to 350 psi.
In many instances, this pressure capacity is sufficient for dry air drilling. However, particularly
when drilling deeper wells, when using a downhole motor or a percussion hammer, or when
significant water inflows are expected, the required air delivery pressure can exceed the pressure
rating of conventional compressors. The booster is then used.
Boosters
Boosters are essentially positive displacement pumps (compressors) that take the output of the
compressor(s) and boost the pressure to as high as 1,500 psi. Boosters for oilfield use are either
single- or two-stage. Single-stage boosters are usually adequate only for lower pressure
operations. Almost all of the boosters in the Appalachian area are single-stage. In combination
with the low pressure compressors used in this region, the total system can typically operate at a
maximum pressure of about 500 psi for extended periods. They are capable of higher pressures
but not for long periods of time. The high compression ratio causes excessive heat that can
damage the booster. The majority of the boosters available from air drilling service companies
are two-stage and are capable of sustaining much higher delivery pressures. A typical two-stage
booster system can operate for extended periods, at 1500 psi, without overheating.
Normally one booster unit will have sufficient volume capacity for most air drilling operations. If
necessary, two or more boosters can be run in parallel to handle higher flow rates. The booster is
always required to unload water from casing after cementing.

2-24

Air Header and Valves


A large diameter, ideally 4 inches, hose or pipe, with a pressure rating matching or exceeding the
boosters should be used to take the air from the compression system to the rigs standpipe
manifold. This is sometimes termed the air header. The compressors are often connected to this
line with ball valves, to avoid problems if it is necessary to take one out of service. Similarly, ball
valves are usually used to isolate the lower pressure portion of the compression system from
possible flowback through the booster when that is shut down. Check valves in these latter
locations can provide an additional degree of security should there be concerns about possible
operator error.
There should be a valve for venting the compressor end of the delivery line. It is also advisable
to install a pressure relief valve, set to open just above the delivery pressure rating, in order to
protect the compressors. If a booster is installed in the circulating system, provision has to be
made for venting high air pressures between the booster and the standpipe. Since these pressures
can potentially be 1,500 psi or more, attention must be paid to how the high pressure vent line is
sited and restrained.
Mist and Soap Pumps
Figure 2-6 shows mist and foamers (soap) pumps. These are not strictly required for dry air
drilling. However, mist pumps are always required when unloading casing after cementing, even
if the well is later dried out and dusted to TD. All air packages have a mist pump. Also, the
possibility always exists that it will be necessary to switch from dry air to mist or foam as the
circulating fluid, in which case these pumps will be needed. Further details are given in Sections
2.4 and 2.5 on mist and foam drilling, respectively.
It should be possible, when necessary, to divert air from the air header, directly to the blooie line,
through a bypass line, so that the compressors can be kept running during connections. The
bypass is sometimes also referred to as a blow-down line. There should be a second, bleedoff
line between the standpipe and the blooie line, downstream of the bypass line, to allow air
pressure in the standpipe and drillstring to be bled down before making a connection. Two-inch
diameter pipes are normally adequate for both bypass and bleedoff lines. It may be possible to
arrange the standpipe manifold to provide the necessary control of air flow. If not, suitable
valves that can be operated from the rig floor, must be installed.
If there is any potential of having to circulate air down the annulus, a connection with appropriate
valving should be made from the air delivery line to the wellhead.
It is advisable to have mud pumps rigged up so that liquids can be circulated into the well
without delay if downhole conditions so dictate. The mud pump delivery line can be connected
to the standpipe manifold to allow selection of the circulating fluid. Even so, it is advisable to
install check valves in the air and mud lines to prevent flow from one line entering the other.
Once the circulating air reaches the standpipe, conventional rig equipment is used to connect to
the top of the kelly (or drillstring if the rig is fitted with a top drive).
2-25

Chapter 2

Underbalanced Drilling Techniques

Drillstring
There will normally be one or two non-return or float valves placed in the drillstring; one near
the surface and one just above the bit. These may be flapper or dart (piston) type valves. Both
are illustrated in Figure 2-7.

Flapper Style

Figure 2-7.

Dart (or Piston) Style

Drillstring float valves for air drilling.

The drillstring should never be run without at least the lower float valve. This is often a dart-type
valve. It prevents cuttings from flowing back through the bit when they could plug the bit, or any
downhole motor or hammer that might be in use. In the absence of a string float, the lower float
prevents any flow from the annulus back up the string. It also prevents flow of gas up the
drillstring while tripping.
The upper float is commonly termed the string float. This is usually a flapper-type valve. It
prevents most of the compressed air in the drillstring from being lost at the rig floor when a
connection is made. It reduces the time required to bleedoff the air pressure in the string, and
thereby reduces the time taken to make a connection. It is not normally necessary when
standpipe pressures are low. Typically, a string float will not be added to the string until the well
depth reaches 3,000 to 4,000 feet. If a downhole motor or percussion hammer is in use, the
standpipe pressure will be higher than for conventional rotary drilling operations. In these
instances, a string float will often be added to the drillstring at shallower depths.
It is common practice to remove the spring that closes the flapper valve in the string float before
it is put in the string. When the standpipe manifold is opened to bleedoff air pressure before
making a connection, the air flow up the string will close the valve. Then, after the air pressure
below the valve has bled down through the annulus, the valve will fall open. Thus, if the flapper
valve spring has been removed, it is possible to run wireline tools through the string float. It is
possible to run inclination survey tools through the string float in this way, although great care
has to be exercised when pulling the survey tool back up through the float valve. Alternatively,
the string can be tripped, the float valve removed and the string run back to bottom before taking

2-26

a survey. The string float will then be installed at the surface before resuming drilling operations.
Custom-modified float valves can be included in the drillstring to prevent the flow of air down
the drillstring in the event of a downhole fire (refer to Figure
Figure 2-8).
2-8 A fire stop will normally be
run near the top of the collars. This is basically an inverted flapper valve, in which the flap is
held open by a zinc ring. If a downhole fire occurs, the zinc ring melts and allows the flap valve
to close. As with the conventional flapper valve run near the top of this string, the fire stop does
not obstruct the inner bore of the drillstring; wireline tools can safely be run through it when it is
open. A fire float is a modified dart-type float valve, run instead of the conventional lower float.
Under normal circumstances, it functions like a conventional dart-type valve. It opens in
response to air pressure in the string above the valve and closes when circulation is shut down. If
there is a fire downhole, the zinc ring melts, allowing a sleeve to close the flow ports in the valve
body, preventing flow in either direction through the valve.

Figure 2-8.

Drillstring floats for air drilling; fire stop and fire float valves.

2-27

Chapter 2

Underbalanced Drilling Techniques

Return System
A conventional open bell nipple will not direct the returning air flow away from the rig
substructure. Additional equipment, a diverter, is required above the Blowout Preventer (BOP)
stack to do this. Although it is possible to use various types of equipment as a diverter, it is now
normal to use either a rotating head or a rotating BOP. These are shown in Figure 2-9. Both of
these use elastomeric elements to seal around the kelly and direct the return flow laterally
through the outlet and into the blooie line. The principal difference between these two types of
diverters is that the sealing element in a rotating head is actuated by the air pressure that it seals,
whereas the element in a rotating BOP is actuated hydraulically. Typically, rotating heads have a
pressure limitation of 500 psi, and rotating BOPs can seal higher pressures, up to 1,500 psi while
drilling (2000 psi static).23 It is important that both are operated according to their
manufacturers recommendations. Unacceptably rapid wear of the seal element and mechanism
will occur if the axis of the diverter is not aligned directly with the center of the rotary table or if
lubrication is inadequate.
The diverter system does not remove the need for a conventional BOP stack. This should
comply with local regulatory requirements. At a minimum, it should contain pipe and blind rams
so that the well can be shut in with the string in or out of the well. For gas wells at least, the pipe
and blind rams should be able to support the highest anticipated formation fluid pressure. Where
it can be accommodated beneath the rig floor, it is desirable to have a full stack consisting, from
the wellhead up, of pipe rams, blind rams, pipe rams and annular.
This provides operational flexibility. For example, it allows stripping pipe back into the well
under pressure if high pressure is encountered that cannot be contained within the lower pressure
capacity of the diverter.
The returning air flow is taken from the diverter to a flare pit through the blooie line (Figure 2-10).
Ideally, the blooie line should be oriented at 45 away from the direction of the prevailing
wind, and must be firmly restrained along its full length. It should be sufficiently long that any
flared or unignited gases and cuttings are kept well away from the rig. Regulations often set a
minimum length of 100 or 150 feet for the blooie line, but lengths of up to 300 feet have been
recommended.7
If the diameter of the blooie line is too large, cuttings may not be efficiently carried. Conversely,
if it is too narrow, the pressure drop down the line can become excessive. As a rule-of-thumb, the
blooie line diameter should have a cross-sectional area that is equal to that in the annulus over
the longest section of hole to be drilled with air. Most blooie lines are 7-inch or 8 5/8-inch
casing. Seven-inch casing would have the same cross-sectional area as a 7 7/8-inch by 4-inch
annulus. An 8 5/8-inch blooie line would have a cross-sectional area equivalent to a 9 1/8-inch
by 4-inch annulus. It is unusual for a blooie line to have an internal diameter greater than nine
inches. Larger diameter blooie lines are very heavy and difficult to rig up. Cranes may be
required to set very large diameter lines in place. It is normal for the blooie line to have a crosssectional area that is less than the annulus when drilling surface hole. Figure 2-11 shows

2-28

Higher rated devices are now available.

frictional pressures losses, calculated for typical, 150 foot long blooie lines, made from 7-inch
and 8 5/8-inch casing.

Rotating Head

Quick-change packer
assembly

Kelly driver
assembly

Hydraulic fluid Inlet


Inner packer
Outer packer
Bearings

Mechanical seal

Outlet flange

Hydraulic fluid return

Casing flange

Rotating Blowout Preventer

Figure 2-9.

Diverters; rotating head and rotating blowout preventer (after Cooper et al., 1977;24
and courtesy of Signa Engineering Corporation).

2-29

Chapter 2

Underbalanced Drilling Techniques

Propane
Secondary
Jet

Degasser
Pump

Primary
Jet
Gas
Sampler

Rotating Head

Pilot
Light
Annular BOP
Dual Ram BOP
Sample
Catcher

Figure 2-10.

Blooie Line.

4
7" Blooie Line

3.5

8 5/8" Blooie Line

Pressure Loss (psi)

3
2.5
2
1.5
1
0.5
0
2000

2200

2400

2600

2800

3000

3200

3400

3600

3800

4000

Flow Rate (scfm)

Figure 2-11.

Frictional pressure losses down two different 150 foot long blooie lines.

Even at 4000 scfm, the friction losses for the 8 5/8-inch line are only 1 psi and would not be
detrimental to drilling operations. The 7-inch blooie line has a friction loss of 3 psi at 4000 scfm.
This will have a slightly greater effect on drilling operations.

2-30

The bypass line should be connected to the primary jet, which is located inside the blooie line.
When air is pumped through the bypass line during a trip, the primary jet reduces pressure in the
blooie line. Provided that the pressure decrease is sufficient, this will pull any produced gas
flowing from the well into the blooie line and away from the rig floor. The rotating head is
always kept in place when making a connection so that any gas flow is automatically diverted
into the blooie line without having to rely on the primary jet.
Figure 2-12 shows the most effective way to install a primary jet. It should be located at a
distance of about four times the blooie lines internal diameter from the exhaust end.24 This
location maximizes the pressure drop in the blooie line for any given air flow rate through the jet.
If it is much closer to the exhaust end, the air stream from the jet will not have expanded to the
full internal diameter of the blooie line. On the other hand, if the primary jet is placed farther up
the blooie line, the pressure decrease is reduced by the friction losses downstream of the jet.
Cuttings will eventually erode the tube extending into the blooie line and the jet will have to be
replaced. This jet should be constructed of 3/4- to 1-inch diameter tubing. To extend the useful
life of the jet, it can be covered with rubber hose. Another method for protecting the jet is to tack
weld used tong dies on the upstream side of the tube. In either case, the jet should be designed so
that the tube can be easily replaced.
The bleedoff line should be connected to a secondary jet, which is normally located at the other
end of the blooie line, close to the diverter. A simpler design, as shown in Figure 2-12 (b), is
normally adequate for the secondary jet.
Cooper et al., 1977,24 measured the pressure reduction that occurred while flowing air at 1200
scfm, sequentially through identical primary and secondary jets, into a 7-inch diameter blooie
line. With the primary jet located approximately 28 inches from the blooie line exhaust end, a 6
psi pressure drop was generated at the diverter end of the blooie line. The same flow, through
the secondary jet, located much closer to the diverter, resulted in a pressure drop of only 2 psi.
If too small a diameter blooie line is used, the amount of gas that can be jetted out of the wellbore
can be restricted. When the combined pressure drop due to the full production rate of gas
flowing from the well down the blooie line exceeds the pressure drop (including momentum
changes due to the jet configuration) created by the jet in use, some of the produced gas will be
vented to the rig floor when the rotating head is pulled on a trip. The pressure drop down the
blooie line will be higher for smaller diameter blooie lines. Less gas can be jetted away from the
rig floor with a smaller blooie line than with a larger line.
In normal drilling operations, the noise generated by the air flow exhausting from the end of the
blooie line is low. Noise levels can become significant when the compressor delivery is
bypassed through the primary jet or when bleeding pressure off the drillstring. In these
situations, air flow through the smaller diameter lines and the primary and secondary jets creates
the noise, rather than the exhaust from blooie line itself.

2-31

Chapter 2

Underbalanced Drilling Techniques

2" Union

4" x 2" Swedge


1/4" Plate
4" Collar cut in half

4D

3/4" to 1" Tubing


covered with
rubber hose

2" Union
20 o

2"

Air flow

Blooie Line

(a) Primary Jet


Figure 2-12.

Air from
standpipe
manifold

Blooie Line

(b) Secondary Jet

Recommended jet construction; primary jet and secondary jet (after Cooper, et al.,
1977 24).

A gas detector, capable of discriminating hydrogen sulfide and hydrocarbon gases, should be
installed in the blooie line, close to the diverter if at all possible. Gas detectors located at the
exhaust end of the blooie line can be damaged by the flare if the wind is blowing towards the rig.
It is good practice to have a similar gas detector on the rig floor or in the substructure.
Some arrangement should be made to collect cuttings samples. This can be done with a valved
stub line. This would normally be approximately two inches in diameter, leading from the
bottom of the blooie line at some convenient point along its length. A design that has proven to
be effective is shown in Figure 2-13. A used tong die, welded into a 3-inch to 1-inch pipe
swedge, deflects cuttings from the return flow into the swedge, where they accumulate until the
ball valve is opened. Alternatively, if there is no flare burning, it is possible to use a plate
downstream of the blooie line exhaust to deflect cuttings into a suitable container.
Cuttings coming from the blooie line are often very fine, dry dust. These pose both an
environmental and a health hazard, if they can blow freely about the rig site. Some form of dust
control is required. This normally involves dampening the cuttings with a water spray either at
the end of the blooie line or preferably shortly before the cuttings leave the blooie line. Cooper et
al., 1977,24 described one system that used a one-inch diameter water line discharging into the
blooie line through two jets, separated by about eighteen inches. Alternatively, it is possible to
use a ring arrangement of jets, as is sometimes used for defoaming when drilling with mist or
foam.25 This provides more uniform coverage of the returning cuttings flow.

2-32

Figure 2-13.

Cuttings sample catcher.

The exit end of the blooie line should be above a flare pit. A pilot light should be kept ignited in
the exit air flow adjacent to the end of the blooie line at all times when there is any prospect of
producing hydrocarbon gas.
The requirements for flare and reserve pits will differ from well to well. For example, Cooper et
al., 1977,24 recommended a flare pit at least 30 feet square in plan, with a separate reserve pit.
When the well is to be drilled from surface to total depth with dry air or mist, the reserve pit and
flare pit can be combined into one relatively small pit, as illustrated in Figure 2-14. A separate
reserve pit is not required if this combined pit can accommodate all produced water and cement
displacement water. If drilling mud is to be used in addition to air or mist, a more extensive
reserve pit will be required. Sometimes, as shown in Figure 2-15, this is constructed separate
from the flare pit. Then, if the flare pit is not big enough to hold all of the liquids produced when
air drilling, a spillway can be constructed to allow the surplus, produced liquids to flow into the
reserve pits. Care is required when the well produces a large amount of oil. All of the oil will
not burn at the end of the blooie line. Some of the oil will burn on top of the reserve pit. A large
burn pit is best when oil is expected.
The reserve pit wall, at the back of the reserve or flare pit, should be far enough from the end of
the blooie line that water unloaded from the hole will not wash it out. When unloading cement
displacement water or produced water, the water will come out of the end of the blooie line at a
high velocity. If the blooie line is too close to the reserve pit wall, the water can erode a hole and
breach the reserve pit. The wall at the back of the reserve pit should be higher than the other
walls. This high wall prevents cuttings, dust and produced fluids from drifting over the wall with
the air.
Instrumentation
Some supplementary instrumentation, in addition to that found on a conventional drilling rig, is
advisable for air drilling. A low pressure gauge, with a range somewhat greater than the
compressor pressure rating, should be installed in the air delivery line, close to the compressors.
An equivalent high pressure gauge, with a range exceeding the delivery pressure rating of the
booster, should be installed between the booster and the standpipe.
2-33

Chapter 2

Underbalanced Drilling Techniques

An orifice meter should be installed in the air line between the booster and the mist pump to
measure the air injection rate. An air pressure gauge and a thermometer should be located
upstream of, but close to, the orifice meter. The flow rate, Q (scfm), is calculated from the
pressure differential measured across the orifice, hw (inches of water), the flowing air pressure, Pf
(psia), and the temperature, Tf (F), by:
Q = Fb Fg

520h w Pf
Tf

(2.30)

where:
Fb ....... orifice flow factor, (Appendix B),
Fg ....... (1/S)0.5, and,
S......... gas gravity (1 for air).
The flowing air pressure can be converted from gage to absolute pressure by adding the normal
atmospheric pressure for the rig-site altitude (Appendix A).
High W all At Back of the Pit

Reserve and Flare Pit

Mud Tank

Mud Tank

Blooie Line

Mud
Tank

Rig Substructure

Figure 2-14.

2-34

Combined flare and reserve pit.

Figure 2-15.

Separate flare and reserve pits (Cooper et al., 1977 24).

A good indication of the standpipe pressure is required. This pressure will typically be much
lower than during mud drilling operations. A standpipe pressure gauge with a lower range than
normal, for example, 0 to 1,000 psi, should be installed. It should be possible to isolate this
value from the standpipe if higher pressures are required. Small changes in standpipe pressure
when air drilling can be an indication of larger changes in bottomhole pressure. Therefore, it is
strongly recommended that the standpipe pressure gauge should be capable of reliably indicating
any pressure change of 10 psi or more.
2.1.4

Operating Procedures

This section gives general guidelines on operating procedures, when drilling with dry air. These
are guidelines only and will need to be changed to fit specific well conditions.
Standpipe Pressure
It is important to monitor standpipe pressure carefully when circulating or drilling ahead. Large
pressure changes downhole due to major hole problems may only cause small changes in
standpipe pressure. For example, Figure 2-16 is an orifice meter chart from an air drilled hole.
The well was drilled in the Arkoma Basin, using air as the circulating medium. The chart shows
the pressure in the air header, which is effectively the same as the standpipe pressure. The scale
is from 0 to 500 psi. The differential record is the differential pressure across the orifice, on a
scale of 1 to 100 inches of water. As can be seen on the chart, the drillstring became stuck at
3:45 a.m., while drilling at 10,854 feet. During the day, the circulating pressure was between 140
and 150 psi. When the drillstring became stuck, the pressure had increased to only 160 to 170
psi. A 20 psi pressure increase was the difference between drilling and being stuck.
2-35

Chapter 2

Underbalanced Drilling Techniques

Any detectable change in standpipe pressure should be treated as a warning sign of a potential
problem downhole, its cause determined and, if necessary, appropriate corrective action taken.
Making Connections
Connections, when drilling with dry air, are a little more complicated than when drilling with
mud. Air is a compressible fluid and the pressure must be bled from the drillstring before
breaking off the kelly. If this is not done, the energy stored in the compressed air inside the
drillstring, kelly hose, and standpipe will be released violently when the kelly is broken off. This
poses a considerable hazard to the rig crew. The standpipe manifold is used to bypass the flow
from the compressors and bleed the pressure off the drillstring. Figure 2-17 is a schematic of a
standpipe manifold. In this figure, Valves 1 and 2 would be open during air drilling. This would
allow air to pass into the standpipe. Valves 3, 4 and 5 would be closed.
After drilling the kelly down, it is time to make a connection. Unlike mud pumps, compressors
are not normally turned off while making a connection. Instead the air is bypassed to the blooie
line through the primary jet. Valve 3 (Figure 2-17) is opened to allow the air to bypass the
drillstring. Valve 2 is closed to isolate the compressors from the drillstring. Valve 4 is opened to
bleed the pressure off of the drillstring. Valve 4 allows the compressed air to be vented to the
blooie line through the secondary jet. Once the pressure has bled off the drillstring, the kelly can
be disconnected and a joint added.

2-36

Drillstring became
stuck at 3:34 a.m.

Increased Rate

Figure 2-16.

Meter chart from a well in the Arkoma Basin. The drillstring became stuck at 3:45
a.m. while drilling at 10,854 feet.

After the connection is made, Valve 2 is opened first. Valves 4 and 3 are then closed. Valve 3 is
not closed before opening Valve 2 because the compressors will be shut in and the pressure will
increase rapidly. Unlike mud drilling, it takes time for the drillstring to re-pressurize and for
returns to be observed at the blooie line.
After adding another joint, it is very common in air drilling operations to leave the pipe on the
slips until circulation is established; i.e. returns are noted at the blooie line. Since differential
pressure sticking is not a problem in air drilled holes, pipe movement is not a requirement. With
air in the annulus, the cuttings will drop to the bottom while a connection is made. If the volume
of cuttings in the annulus is greater than the hole volume below the bit or if the cuttings bridge

2-37

Chapter 2

Underbalanced Drilling Techniques

off at the bit, the cuttings will fill the annulus above the bit. As air is circulated into the annulus,
the cuttings will again be suspended in the air column and drilling can resume. Increased
standpipe pressure when breaking circulation is an indication that cuttings are above the bit.
When the pressure is high enough, the cuttings will start to move and the standpipe pressure will
decrease.
S ta n d p ip e
T o K e lly

F ro m M u d
Pumps

B le e d -O ff L in e
T o S e co n d a r y
Je t

Figure 2-17.

B yp a s s L in e
T o P r im a ry
Je t

F rom
C o m p r e ss o rs

Schematic of a standpipe manifold.

Shortly after this, returns should be observed at the blooie line. If the drillstring is picked up
before the cuttings are moved, the cuttings can be tightly packed in the annulus, making it
more difficult or impossible to break circulation. The drillstring will be stuck. It is good
practice to observe circulation at the blooie line before moving the drillstring.
Tripping
Tripping from an air drilled hole is not substantially different from tripping with mud in the hole,
except that it is a much cleaner operation. Generally, it is a good idea to circulate bottoms up
prior to tripping out. In an air hole, this will only take a few minutes, since annular velocities are
on the order of 1000 to 6000 feet per minute, depending on the depth of the well. The trip can
begin when the volume of cuttings observed at the blooie line decreases.
Do not drop the survey instrument before tripping out of the hole. Survey instruments fall much
too rapidly in air. They will be destroyed when they hit bottom.
Any surveys should be run on wireline.

2-38

Unlike mud drilling, the rotating head rubber is left in the bowl while pulling the drillpipe. Any
gas flowing from the well will be diverted down the blooie line, keeping it off the rig floor.
Stripping through the rotating head rubber is usually cheaper than the cost of fuel to keep the
compressors running through the primary jet. If the well is not making gas, the rotating head
rubber can be pulled to extend its life.
Prior to pulling the bottomhole assembly, the rotating head rubber is pulled. However, collars
with a diameter equal to the drillpipe couplings can be stripped through the rubber if desired. The
primary jet is used to keep gas off the rig floor after pulling the rotating head rubber. The
amount of gas that can be pulled down the blooie line is extremely variable. It depends on the
gas volume, the gas velocity in the annulus, the blooie line diameter and how the primary jet was
installed (refer to the previous section). It is not uncommon to jet two to five million cubic feet
of gas per day down the blooie line.
With the gas being jetted down the blooie line, the bottomhole assembly can be pulled and
changed. Once the bottomhole assembly is run back into the wellbore, the rotating head rubber
can be installed and the trip continued to bottom. After reaching bottom, it is good practice to
start the air and observe returns before starting to wash to bottom. This should indicate whether
or not water is present in the well. If it is, the water must be unloaded from the hole and the hole
dried before drilling can resume. Unloading the hole is described below. With small quantities of
water in the well, it may be possible to unload with the bit at bottom. Otherwise, it may be
necessary to stage the string into the hole. This is also described in the following section. Water
in the hole must be unloaded after running casing and when formation water is encountered.
Unloading the Hole
When the casing is cemented, the cement is displaced with water. The displacement stage is left
in the casing until the cement sets up. In order to drill the section of the hole below the casing
with air, this water must be unloaded from the hole. There are two ways to unload the casing.
One method is to trip in the hole to the float collar:
 Start circulating water with the mud pump(s), at a low enough rate that the standpipe pressure
remains low. Cooper et al.,1977,24 suggested a flow rate as low as 1.5 to 2 BPH (1 to 1.5
gpm).
 Start up one compressor and booster. Deliver air to the standpipe so that it begins to aerate
the water being delivered to the standpipe by the mud pumps. If necessary, reduce the air
delivery rate so that the standpipe pressure remains below the air delivery pressure capacity.
 Pump approximately 10 BPH of mist fluid into the air flow with the mist pump. The mist
fluid should contain 0.1 to 0.25 percent, by volume, foaming agent, or soap.
 After air returns to surface, reduce the mud pump volume and increase the air volume. As the
standpipe pressure decreases, the mud pumps can be turned off.

2-39

Chapter 2

Underbalanced Drilling Techniques

Another method, staging into the hole, unloads the hole without using mud pumps:
 The string should be tripped part-way into the hole. With a float valve near the bit, water will
not enter the string as it is tripped downhole. Instead, it will be displaced into the annulus,
increasing the hydrostatic head and the bottomhole pressure.
 Attach the kelly and start air circulation. The air in the drillstring will be compressed
progressively until the air pressure at the float valve exceeds the water pressure below the
valve. Water will then be lifted up the annulus and discharged from the blooie line.
 Once water flow has stopped, remove the kelly and start tripping.
 The cycle of tripping and unloading is repeated; unloading the hole in stages, until the float
collar is reached and the casing is empty.
The depth to which the string should be tripped before circulating air is partly determined by the
pressure capacity of the air compression system. Low pressure systems will not be able to unload
water from as deep a stage as high pressure systems. If the delivery pressure capacity of the air
system is Pmax (psi) and the formation water pressure gradient is 0.433 psi/ft, the maximum stage
depth, Dmax (feet), from which water can be unloaded is:
D max =

Pmax
0.433

(2.31)

For an air pressure capacity of 350 psi, the maximum stage depth is little more than 800 feet.
With a pressure capacity of 1,500 psi, it would be possible to unload over 3,500 feet of water per
stage. This is not advisable. Once air passes the bit and enters the annulus, the hydrostatic
pressure resisting air circulation will reduce. The compressed air will expand and cause the rate
at which the water is displaced from the well to accelerate. Water discharge from the blooie line
can become very violent if the standpipe pressure necessary to initiate water displacement is high
and significant energy is stored in the air compressed inside the drillstring. The high annular
velocities near the surface can damage the well and surface equipment. For this reason, it is
unusual to unload more than 2,000 feet in any one stage.
Even with restricted stage lengths, this method of unloading the hole causes greater water surges
at the blooie line than when the mud pumps are used in conjunction with the air delivery system.
The blooie line must be securely anchored. Water surges are also more likely to wash out the
bank of the reserve pit opposite the blooie line. To some extent, surging can be limited by using
only one compressor and booster, until air returns to the surface.
After unloading, the float equipment and shoe are drilled out, using mist at about 6 to 10 BPH
(refer to Section 2.4 Mist Drilling). Once new formation is encountered, the hole must be
dried out before the well can be drilled with dry air. This should be done carefully. Before drying
out the well, it should be circulated until the mist comes back clean; with very few cuttings in it.
The mist pump should be turned off.

2-40

The well is sometimes then slugged with extra foaming agent, in an attempt to remove as much
water as possible. To do this, the air compressors are turned off and the kelly is removed. Ten to
twenty gallons of foaming agent and two to four barrels of water are poured down the drillpipe.
The kelly is re-attached, the bit is run to bottom and air is circulated at the rate that will be used
when drilling. The slug will be seen as a stiff foam at the blooie line; like shaving cream in
consistency.
Whether or not the well is slugged with foam, the hole should be blown with dry air (usually for
one-half hour; one hour at the most). The longer circulating times are required for deeper wells.
If the well continues to make water throughout this period, it is likely that formation water is
being produced and that the well cannot be dried out.
When the well quits making water at the blooie line, it is ready to be dried out. The moisture
remaining on the pipe and borehole wall must be removed. Simply circulating air will not
normally dry out a well in an economically acceptable timeframe. If the surface humidity is high,
air circulation alone may never entirely dry the hole. The most economical drying agent is
cuttings. As cuttings travel up the wellbore, they absorb moisture and carry it out of the hole.
Drill ahead in five to ten foot intervals, circulating off bottom for a few minutes between each
interval. While circulating off bottom after each interval, pass a tool joint to make sure that the
drillstring can be moved freely. This will indicate that no mud ring is forming (the formation of
mud rings, their detection and treatment are described below).
The well should start dusting (returning dry cuttings to the surface), before more than 30 feet
have been drilled. If the well is only making a little water, two to three joints may need to be
drilled down before it starts dusting. If the well does not start dusting after drilling 90 feet, the
well is making too much water. In this case, the prospect of further dry air drilling should be
abandoned and mist drilling should be adopted. The total time required to dry out the hole
depends on depth. It should not exceed two to six hours.
Economics of drilling can be improved if unloading and tripping is done more rapidly. Williams,
1997, (personal communication), provided an example of how a top drive rig can reduce required
times, since connections can be broken every ninety feet rather than every thirty feet. A comparison
is shown in Figure 2-18.
An example of a slightly different unloading protocol is included in Chapter 7, Section 7-6
(Graham, 1997, personal communication).
Water Inflows While Drilling
Water will flow into the wellbore when a permeable, water-bearing formation, or a water-bearing
fracture system, is penetrated. As water enters the wellbore, it will be broken up into droplets that
are lifted out of the hole; much like cuttings are. The dry cuttings will tend to absorb water as
they travel up hole. They can effectively dry the well if the water inflow is slight.24

2-41

Chapter 2

Underbalanced Drilling Techniques

There is a dangerous regime when a modest water inflow occurs. This moistens the cuttings and
they tend to build up into a mud ring. A mud ring can stick the drillstring and may lead to a
downhole fire. Higher inflow rates saturate the cuttings and reduce their tendency to agglomerate
as a sticky mass that forms a mud ring. At this stage, water droplets are likely to be seen in the
return air flow at the end of the blooie line. If the inflow rate increases, the air flow may
ultimately be incapable of separating the water into droplets downhole. At this point, flow tends
to slug; the water is lifted out of the well as discrete volumes or slugs. These slugs of water can
aggravate wellbore instability problems as they pass uphole. If the flare/cuttings pit has earth
walls, these may be damaged or cut out by water slugs when they reach the surface.7
Water inflows cause the standpipe pressure to increase because of the increase in density of the
air/cuttings/water mixture flowing up the annulus. A pressure increase of five to ten psi can
accompany an inflow, with the potential to form mud rings.19 This can occur before any water is
detected at the blooie line. If slug flow occurs, the standpipe pressure will show large surges (50
psi or more24) due to the additional hydrostatic head of the water in the hole.

200
180

Effective ROP (ft/hr)

160
140

Without Top Drive


With Top Drive

120
100
80
60
40
20
0
0

10

12

14

16

18

Connection Time (minutes)

Figure 2-18.

2-42

An example comparison of effective ROP when a top drive rig is used.

20

The consequences of water inflow and the required remedial action depend on the amount of
water flowing into the well. Small water inflows are not readily detected. Water may not be seen
at the surface because it can be absorbed downhole by the cuttings. The first indication of a
small water inflow is usually that the well will stop dusting; i.e. the discharge of cuttings at the
blooie line will stop. This is not always easy to detect. In most air drilling operations, the dust is
suppressed with water at the blooie line, and no change will be visible from the drilling floor.
Careful and frequent checking of the sample catcher will reveal if cuttings return has stopped,
since the catcher should be upstream of the deduster.
A small water inflow has the potential to stick the drillstring, unless it is caught in time. The
water mixes with the cuttings downhole and creates a mud-like slurry of moistened cuttings.
Since cuttings transport efficiency is lowest at the top of the drill collars, the slurry of cuttings
will tend to accumulate here, on the drillstring and the walls of the hole. The clearance in the
annulus progressively decreases until the annulus is blocked and the drillstring is stuck. This
restriction is known as a mud ring. This is illustrated in Figure 2-19.
The sooner the formation of a mud ring is detected, the less likely it is that the drillstring will
become stuck. The formation of a mud ring can be detected by an increase in standpipe pressure.
Since the mud ring forms a restriction in the annulus, it increases the frictional pressure losses
and in turn increases the standpipe pressure.

Figure 2-19.

Cuttings mixed with a small amount of water will form a mud ring at the top of the
drill collars where hole cleaning is critical.

2-43

Chapter 2

Underbalanced Drilling Techniques

While an increase in standpipe pressure is not always associated with the formation of a mud
ring, it should be considered as a warning of potential problems. If the standpipe pressure
increases, stop drilling and check to see if the well is still dusting properly. If the well has quit
dusting, there is a good chance that a mud ring has formed. Cuttings are one of the primary
ingredients necessary to form a mud ring. By stopping drilling, additional cuttings will not be
generated. Any mud ring will be inhibited from growing further and becoming more serious.
String Washouts
Although string washouts are not especially common when air drilling, they can cause problems.
They occur when a hole develops in the drillstring, allowing air to escape into the annulus
without passing through the bit. The washout may be due to a fully-penetrating fatigue crack, in
the pipe itself or in the tool joint, or due to a poor seal at the threaded connection.
In dry air drilling, the cuttings-laden air flowing at high velocity up the annulus tends to erode the
drillstring. This erosion is concentrated at the lower portions of the tool joints. The tool joints
also tend to wear due to rotating contact with the hole wall. Since there is no lubricating liquid,
abrasive wear occurs more rapidly than if mud was present. Dry air drilling is also associated
with greater downhole vibration than if mud was in the hole. These vibrations are often
attributed to dramatically lower viscous damping of string motions in air.7,26 As a consequence,
fatigue cracks can initiate and propagate more rapidly in air drilling than in conventional mud
drilling operations.
Washouts associated with poor seals may deteriorate more slowly and may be less troublesome
when drilling with air than with mud. The air flow down the pipe and through the washout
should not contain any solids. As a result, metal around the leakage path is not eroded away as
rapidly as it would be by mud. Air drilling operations can continue for extended periods with
minor string washouts at poor connection seals. In a mud drilled hole, washouts can be found in
a drillstring that is used after the string has been used for air drilling.
When a washout occurs in a mud drilled well, standpipe pressure reduces. When this is detected,
the string is tripped wet. When the washout comes through the rig floor, mud will flow out of it.
The rig crew removes the section(s) of pipe with the washout and returns to bottom.
In air drilling operations, a washout is also indicated by a reduction in the standpipe pressure,
although the pressure change is smaller than it would be with mud. Finding the washout is
difficult because the drillstring cannot be pulled wet. Attempts should be made to locate
suspected washouts. Trip out of the hole and install small jets in the bit. Attach the bottomhole
collar to the kelly and record the standpipe pressure, while circulating with the compressors and
booster. Trip the drillstring in the hole and periodically (every five to six stands) pick up the
kelly and observe the standpipe pressure while circulating. A substantial reduction in the
standpipe pressure indicates that the washout was in the last section of pipe run. Pipe is pulled
from the hole, testing each stand. When the standpipe pressure increases, the washout is in the
last stand pulled. Now it is a matter of finding which joint or connection contains the washout so
that it can be laid down. Overall, this is a time consuming process.
2-44

There is a temptation to drill ahead until the string parts and then fish for the parted string. This
can be a dangerous strategy if the string parts when being tripped. The fish will fall more rapidly
through air than through mud. It is more likely to corkscrew inside the hole when it hits bottom.
Cooper et al., 1977,24 noted that these corkscrewed fish can be very difficult to retrieve. If it is
believed that the washout is so severe that the string could part when tripped, they recommended
that no attempt be made to trip the string. Instead, they suggested the following procedure.
 Set the bit on bottom.
 Locate the washout by reverse circulating air and running a wireline spinner tool inside the
string.
 Adjust the neutral point and back off the string below the washout. This should leave a
readily-retrieved fish.

2.1.5

Limitations of Dry Air Drilling

The three main limitations of dry air drilling are water inflows, downhole fires, and wellbore
instability.
Water Inflows
The flow of water into a well being drilled with dry air can cause problems that are significant
enough to preclude dry air drilling.
Methods are available to shut off water inflow. All involve attempting to inject material into the
formation that is producing water, where it sets to form a barrier to water flow. The oldest,
established technique is to squeeze cement into the water producing zone. This is only likely to
be successful if the water is flowing from natural fractures; cement is unlikely to enter matrix
porosity. Hower et al., 1958,27 described field trials of two different water shutoff systems. One
of these was based on a resin and catalyst system, which could be used either neat or in a slurry
containing inert solids for bulking and for fluid loss control. Water further catalyzed the resin
setting, which tended to restrict preferential channeling of the resin into the most permeable
zones. The set resin had a compressive strength of 2,000 psi, giving it the ability to shut off
substantial water flows once set. The neat slurry could be squeezed into the matrix of reasonably
permeable formations, although it seemed to be more suitable for fractured zones. The other
water shutoff system they described was a water-based polymer solution that set as a stiff gel.
This system had a lower initial viscosity, allowing it to enter low permeability zones. Core tests
showed that the gel was capable of sealing differential water pressures of up to 2,000 psi.

2-45

Chapter 2

Underbalanced Drilling Techniques

Particularly in the case of cement or resin-based treatments, high pressures have to be applied to
squeeze the treatment into the formation. These pressures can jeopardize wellbore integrity.
Treatment is less likely to be successful if the formation is fractured during the squeeze. The
open fracture(s) will accept the shut-off material at the expense of the water-producing
permeability. Water shut-off methods that rely on liquid material setting all involve a waiting
period of several hours after placement to set, before it will resist water flow and the wellbore
pressure can be reduced.
There are several water shut-off methods that use gas. One requires pumping an aluminum
sulfate solution into the formation, followed by gaseous ammonia.28 These react to form a solid
aluminum hydroxide precipitate, in the formation. This obstructs porosity and shuts off the water
flow. The gas and liquid tend to effectively mix inside the formations porosity. Ammonia has a
low vapor pressure and liquefies at quite low pressures. If the interval to be treated is deeper
than one or two hundred feet, it may be necessary to use a mixture of ammonia and an inert gas,
such as nitrogen or natural gas, to avoid liquefaction due to the injection pressure. In this case, it
has been suggested that a second liquid treatment should follow the gas injection. Otherwise, the
quantity of precipitate may be insufficient to shut-off the water flow.
This shutoff method has several advantages. It does not require any waiting time for setting and
there is no significant risk of it setting prematurely. The materials involved are also inexpensive
(but, gaseous ammonia is toxic). However, it does not seem to be particularly successful in
fractured formations and care needs to be taken to avoid fracturing while treating. It has been
suggested28 that injection pressures should not exceed 0.6 psi/ft.
Another water shutoff method29 uses only gaseous silicon tetrafluoride. This reacts with
formation water to generate a blocking precipitate of hydrated silicon dioxide. In formations
containing calcium, magnesium or sodium salts, additional solids will precipitate and further
contribute to blocking. Laboratory tests have shown that Berea sandstone cores plugged by this
method retained only one percent of their untreated permeability, even when subjected to a
pressure differential of 4,000 psi. Field treatments involve pumping a spearhead of liquid
hydrocarbon or a non-reactive gas, such as air or nitrogen, followed by the silicon tetrafluoride,
and finally tailing-in with additional inert gas or a rubber plug and water.
With all of these water shutoff systems, a single packer has to be set above the zone to be treated
if it is close to the hole bottom, or a straddle packer has to be set across the zone. In either case,
the water producing zone must be located accurately. This may require wireline logging.
Treatments take time, particularly if they involve cement or resin placement. A shutoff treatment
is probably only worth considering if there is a substantial interval still to be drilled with dry air
and the operator can be confident that further water producing zones will not be encountered
between the treated zone and the projected end of the air drilled interval.

2-46

The usual response to a water influx is to switch from dry air to mist or foam. This can lift the
inflowing water, without slugging or mud ring formation. The produced water still has to be
handled at the surface and disposed of in an acceptable manner. The cost of this can outweigh
the reduction in drilling costs for underbalanced drilling. Conventional mud drilling may
become more cost effective.26
Downhole Fires
The possibility for downhole fires is a potential limitation on the use of dry air drilling. Fires
occur when a mixture of oil or natural gas and air, with the hydrocarbon concentration in a
combustible regime, is exposed to an ignition source. A concentration of five to fifteen percent
natural gas is combustible at atmospheric pressure. The upper limit is extended with
increasing pressure, reaching thirty percent when the pressure is 300 psi.
Figure 2-20 shows the influence of pressure on the combustible regime, for a typical natural gas.
These limits vary somewhat with gas composition. Consider a typical air drilling flow rate of
2,000 scfm. Combustion limits suggest that gas inflows of 100 to 600 scfm (144 to 864 Mscf/D)
would be required for a combustible mixture to occur during normal air drilling. However, as
reported by Cooper et al., 1977,24 and by Grace and Pippin, 1989,30 most downhole fires are
preceded by formation of a mud ring that obstructs air flow. Carden, 1993,26 has noted that
downhole fires have not occurred when dry gas is encountered while drilling with dry air. Some
liquid, either water or oil, had to be present. The role of the liquid in causing a fire is presumably
to moisten the cuttings, permitting the formation of a mud ring.
Once flow is obstructed, the air pressure will rapidly rise to the delivery pressure limit of the
compressor system. This increases the temperature of the gas below the mud ring. Since air flow
has been obstructed, even low hydrocarbon inflow rates can rapidly lead to combustible
mixtures. Once the gas mixture is in the combustible regime, the compression itself can often
ignite the mixture. Sparking, from the drillstring rubbing against the borehole wall, particularly
where the rock penetrated is strong and has a high quartz content, can ignite the gas mixture.
Frictional heating of the string by air flow through pinholes has also been cited as a potential
sources of ignition.7
Downhole fires can be very difficult to detect since the combustion rarely reaches the surface. It
may be necessary to run a temperature survey through the drillstring to confirm that a fire has
occurred. Downhole fires usually melt the drillstring at the combustion site. It has been
reported24 that slag from a downhole fire could be blown several hundred feet uphole. The
damaged string can be very difficult to fish and downhole fires usually require sidetracking.

2-47

Chapter 2

Underbalanced Drilling Techniques

400

350

le

Ar
ea

250

Inf
lam
ma
b

Pressure (psia)

300

200

150

100

50

Atmospheric Pressure
0
0

10

20

30

40

Natural Gas in Mixture (% by Volume)

Figure 2-20.

Effect of pressure on combustible concentrations of natural gas in air.

Obvious ways to avoid a downhole fire are to prevent formation of a combustible mixture and to
remove the source of ignition. Changing the circulation medium, from air to natural gas or an
inert gas, will prevent the formation of a combustible mixture. This may not be economically or
logistically feasible. Preventing the formation of mud rings will effectively avoid downhole fires.
Probably the most common method used to avoid a fire when natural gas is encountered is to use
mist rather than dry air.24
Cooper et al., 1977,24 recommended steps to minimize the probability of a downhole fire. First,
attentive monitoring of the standpipe pressure is necessary. An increase of five to ten psi,
without any accompanying increase in compressor output, can indicate the onset of mud ring
formation or a gas inflow. Once a mud ring has developed fully, the flow of air and cuttings
from the blooie line will stop. By then, the string may be stuck or a fire may already have started.
If the standpipe pressure increases when approaching a possible gas bearing zone, or when there
is a gas show at the flare pit, the following steps should be taken:
2-48

Stop drilling.

Shut off the air and monitor the gas flare at the pit. If the flare continues to burn, any of the
following indicates that the gas is wet: wet cuttings, black smoke, yellow flame, sparking at
the blooie line exit (the last three conditions indicate condensate).

If the flare does not continue to burn with the air shut off, resume air circulation and
determine if the gas is wet, using the indicators in (2) above. DO NOT DRILL. New cuttings
will promote the formation of a mud ring. Reciprocate the string to avoid formation of a mud
ring.

If the gas is wet, switch to mist or natural gas drilling.

If the gas is dry, drill ahead in five to ten foot intervals, reciprocating the string between
intervals. Continue to do this until it is certain that there is no further possibility for wet gas
to be encountered.

Wellbore Instability
Dry air drilling usually leads to the lowest wellbore pressures of any drilling method. These low
wellbore pressures can cause mechanically-induced instability, especially in weak formations.
Alternatively, if there is a significant water inflow with water-sensitive shales exposed uphole, it
is possible for the produced water to cause wellbore instability as it is lifted out of the well.
Wellbore instability, during air drilling, is especially problematic when large fragments of rock
break away or slough from the borehole wall. In most cases, the sloughed rock fragments are
much larger than the drilled cuttings. Fragments, as large as 1.5 inches in diameter, may come
over the shale shakers in a mud drilled hole. Figure
Figure 2-21 shows how particle terminal velocity
(under standard conditions, computed using Equation (2.2)) increases with increasing particle
size. Terminal velocities of these large rock fragments can be much higher than the 3,000 feet
per minute annular velocity used in Angel's predictions of recommended air flow rates. The rock
sloughing fragments will not be lifted from the well by circulation rates normally used in air
drilling.
Those fragments that are too large to be lifted from the well will remain downhole until grinding
action of the string breaks them down into particles sufficiently small to be lifted by the
circulating air. If the rate of sloughing exceeds the rate at which the debris is broken up
downhole, fragments will accumulate and eventually stick the string. This process is accelerated
by the increased concentration of rock fragments in the air flowing up the annulus. The resulting
increase in air pressure reduces the annular velocity and progressively leads to poorer hole
cleaning so that even smaller rock particles cannot be removed from the hole. Sloughing from
the borehole wall also increases the hole diameter. This reduces annular velocity and further
degrades hole cleaning.

2-49

Chapter 2

Underbalanced Drilling Techniques

The time, over which the annulus packs off with enough rock particles for the string to become
stuck, depends on the rate at which the formation sloughs, the size of the sloughed rock
fragments, and the rate at which they can be ground up downhole.
If the formation sloughs too fast or in big pieces, air drilling must be terminated. It is unlikely
that the air circulation rate can be increased sufficiently to lift enough of the sloughed rock
fragments from the borehole to prevent the annulus from eventually packing off. Furthermore,
the increased annular velocity will tend to increase the rate at which rock fragments are dislodged
from the borehole wall. A drilling fluid with greater lifting capacity and higher wellbore
pressures must be used.
8000
C d = 0.85

Terminal Velocity (ft/min)

7000
6000
5000
4000
3000
2000
1000
0
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Particle Diameter (inches)

Figure 2-21.

Terminal velocity as a function of particle size at standard conditions (assuming a


drag coefficient of 0.85).

2.1.6 Other Limitations


Other factors that can restrict the use of air drilling are higher friction between the drillstring and
the wellbore, and the difficulties in operating conventional downhole motors and MWD systems
with compressible drilling fluids. These are normally only relevant to directional wells.

2-50

2.1.7

Reverse Circulation Air Drilling

Some of the problems described in this section on conventional air drilling may be overcome or
mitigated by reversing the circulation of the air. In this procedure, air is injected down the
annulus and returned with cuttings up the drillstem. This procedure, still considered
experimental, has several important advantages indicated from tests by Graham, 1986.31
 Reduced Damage to Permeable Formations
Tests strongly suggested less damage to the formation than with conventional air drilling.
Conventional air drilling, in turn, normally resulted in less damage than conventional mud
drilling.
 Quality and Size of Drill Cuttings Improved
Drill cuttings are larger and less contaminated. With the larger cuttings, it is possible to run
quantitative petrophysical analyses. This is virtually impossible with conventional dusting.
 Wellbore Integrity Improved
There is a potential for reduced wellbore damage because there is less hole erosion by
cuttings or water influx that could promote uphole sloughing of sensitive shale.
 Less Air Volume Required
Tests have indicated that reverse drilling used less air. This would be expected since the
velocity of air in the larger annular space would no longer be critical to cuttings removal.
 Limitations
There is a greater likelihood of cuttings plugging the bit. Surface equipment design needs
improvement. Large inflows above the bit may cause problems in circulating down the
annulus.

2-51

References
1.

Johnson, P.W.: Design Techniques in Air and Gas Drilling: Cleaning Criteria and
Minimum Flowing Pressure Gradients, J. Cdn. Pet. Tech. (May 1995) 34, No. 5, 18-26.

2.

Bruce, G.H., Simons, L.H. and Whitaker, W.W.: You Can Recover Large Cuttings When
Air Drilling, Oil and Gas J. (May 1962) 112-116.

3.

Pratt, C.A.: Modifications to and Experience with Air-Percussion Drilling, SPE Drill.
Eng. (December 1989) 315-320.

4.

Zenz, F.A. and Othmer, D.F.: Fluidization and Fluid-Particle Systems, Reinhold
Publishing Corp., New York (1960).

5.

Supon, S.B. and Adewumi, M.A.: An Experimental Study of the Annulus Pressure Drop
in a Simulated Air-Drilling Operation, SPE Drill. Eng. (March 1991) 74-80.

6.

Hagar, J.M., Tian, S., Adewumi, M.A. and Watson, R.W.: An Experimental Study of
Particle Transport in a Deviated Wellbore, J. Cdn. Pet. Tech. (February 1995) 34, No.2,
51-54.

7.

Shale, L.: Underbalanced Drilling Equipment and Techniques, presented at the 1995
ASME Energy Technology Conference, Houston, TX, January 30- February 1.

8.

Gray, K.E.: The Cutting Carrying Capacity of Air at Pressures Above Atmospheric,
paper SPE 874-G, Pet. Trans., AIME (1958) 213, 180-185.

9.

Machado, C.J. and Ikoku, C.U.: Experimental Determination of Solids Friction Factor and
Minimum Volumetric Requirements in Air and Gas Drilling, paper SPE 9938 presented at
the 1981 California Regional Meeting, Bakersfield, CA.

10.

Weymouth, T.R.: Problems in Natural Gas Engineering, Trans., ASME (1912) 34, 185.

11.

Angel, R.R.: Volume Requirements for Air and Gas Drilling, Pet. Trans., AIME, (1957)
210, 325-330; also Volume Requirements for Air and Gas Drilling, Gulf Publishing Co.,
Houston, TX (1958).

12.

Guo, B., Miska, S.Z. and Lee, R. L.: Volume Requirements for Directional Air Drilling,
IADC/SPE paper 27510 presented at the 1994 IADC/SPE Drilling Conference, Dallas, TX.

13.

Nikuradse, J., Forschungshelf, (1933) 301.

14.

Schoeppel, R.J. and Sapre, A.R.: Volume Requirements in Air Drilling, paper SPE 1700
(1967).

2-52

Chapter 2

2-53

Underbalanced Drilling Techniques

15.

Mitchell, R.F.: Simulation of Air and Mist Drilling for Geothermal Wells, J. Pet. Tech.
(November 1983).

16.

Puon, P.S. and Ameri, S.: Simplified Approach to Air Drilling Operations, paper SPE
13380 presented at the 1984 SPE Eastern Regional Meeting, Charleston, WV, October 31November 2.

17.

Wolcott, P.S. and Sharma, M.P.: Analysis of Air Drilling Circulating Systems with
Application to Air Volume Requirement Estimation, paper SPE 15950 presented at the
1986 SPE Eastern Regional Meeting, Columbus, OH.

18.

Grace, R.D.: A Field Study of Annular Pressure Losses In Rotary Drilling with Air-Solids
Mixtures, MSc. Thesis, University of Oklahoma.

19.

Streeter, V.L.: Fluid Mechanics, McGraw-Hill Inc., New York (1971) 337.

20.

Lyons, W.C.: Air and Gas Drilling Manual, Gulf Publishing Co., Houston, TX (1984) 46.

21.

Lyons, W.C.: Air and Gas Drilling Manual, Gulf Publishing Co., Houston, TX (1984) 84.

22.

Lyons, W.C.: Air and Gas Drilling Manual, Gulf Publishing Co., Houston, TX (1984)
Appendix A.

23.

Cress, L.A., Stone, R. and Tangedahl, M.: History and Development of a Rotating
Blowout Preventor, paper IADC/SPE 23931 presented at the 1992 IADC/SPE Drilling
Conference, New Orleans, LA.

24.

Cooper, L.W., Hook, R.A. and Payne, B.R.: Air Drilling Techniques, paper SPE 6435
presented at the 1977 Deep Drilling and Production Symposium, Amarillo, TX.

25.

Shale, L. and Curry, D.A.: Drilling a Horizontal Well Using Air/Foam Techniques, paper
OTC 7355 presented at the 1993 Annual Offshore Technology Conference, Houston, TX.

26.

Carden, R.S.: Technology Assessment of Vertical and Horizontal Air Drilling Potential in
the United States, Final Report, Contract No. DOE/MC/28252-3514 (DE94000044), U.S.
DOE (August 1993).

27.

Hower, W.F., McLaughlin, C., Ramos, J. and Land, J.: Water Can be Controlled in Air
and Gas Drilling, paper SPE 1099-G presented at the 1958 SPE Annual Fall Meeting,
Houston, TX.

28.

Goodwin, R.J. and Teplitz, A.J.: A Water Shut-off Method for Sand-Type Porosity in Air
Drilling, paper SPE 1098-G, Pet. Trans., AIME (1959) 216, 163-167.

29.

Becker, F.L. and Goodwin, R.J.: The Use of Silicon Tetrafluoride Gas as a Formation
Plugging Agent, paper SPE 1098-G, addendum, Pet. Trans., AIME (1959) 216, 168.

30.

Grace, R.D. and Pippin, M.: Downhole Fires During Air Drilling, World Oil (May 1989)
42-44.

31.

Graham, R.L: Exploration-Production Studies in Newly Drilled Devonian Shale Gas


Wells, Annual Report, Gas Research Institute, Chicago, IL (November 1986).

2-54

Chapter 2

2.2

Underbalanced Drilling Techniques

Nitrogen Drilling

In underbalanced drilling operations, nitrogen can be used as the drilling fluid, or as a component
of the drilling fluid. The major advantage over air is that mixtures of nitrogen and hydrocarbon
gases are not flammable. This removes the possibility of downhole fires.
The circulating gas does not have to be pure nitrogen to prevent downhole fires. Mixtures of air,
nitrogen and hydrocarbon are not capable of combustion, provided that the oxygen concentration
is kept below a critical level. Allan, 1994,1 demonstrated that the flammability of natural gas is
quite well represented by the flammability of methane, which has been studied extensively by the
U.S. Bureau of Mines.2,3 At atmospheric pressure, at least 12.8 percent oxygen is required before
it is possible to create a flammable mixture of oxygen, nitrogen and methane. The minimum
oxygen concentration required for a flammable mixture is influenced by the prevailing pressure
and can be represented by the following correlation: 3
O min = 13.98 168
. log ( P)

(2.32)

where:
Omin .... percent oxygen, and,
P......... absolute pressure, psia.
This correlation is shown in Figure 2-22. For a pressure of 3,000 psi, a mixture of oxygen,
nitrogen and methane has to contain slightly more than eight percent oxygen to be flammable.
As Allan indicated,1 this represents the level of oxygen in the drilling fluid, which effectively
eliminates any possibility for downhole fires.
2.2.1 Hole Cleaning
Circulating nitrogen will lift cuttings and liquid inflows, in the same way that air does. As with
air, flow in the annulus is usually turbulent. Consequently, density is the fluid property with the
most impact on cuttings transport. Since the density of nitrogen is only slightly (about three
percent) less than that of air at standard temperature and pressure, cuttings transport efficiency
will be effectively the same as it is for air at the same surface injection rate.
When using air or nitrogen, selecting an injection rate is usually a compromise between increased
cleaning efficiency, leading to increased overall penetration rate, and increased nitrogen costs,
compressor rental and fuel charges. The cost of increasing the nitrogen injection rate is far
greater than for air. The most cost efficient injection rate (not total costs) for nitrogen drilling
will likely be lower than for dry air drilling.

2-55

2.2.2

Equipment

The principal differences between nitrogen and air drilling equipment are due to the substitution
of nitrogen for air, as the circulating fluid. There are currently two main methods of supplying
nitrogen for drilling operations-cryogenic and membrane filter.
Cryogenic Nitrogen Supply
Nitrogen is widely used in well completion, stimulation and production operations. For these
applications, nitrogen is usually transported to the wellsite as a liquid. Since the boiling point of
liquid nitrogen, at atmospheric pressure, is -321F, cryogenic tanks are necessary for
transportation to location. A nitrogen pumping unit consists of a diesel-driven, positive
displacement pump and a heat exchanger. The pump takes liquid nitrogen from the cryogenic
tank and delivers it through the heat exchanger, where heat from the pump engines exhaust is
used to evaporate the liquid nitrogen. There is a large rental fleet of nitrogen units available in
most parts of the United States. Various sized units are available. They have maximum delivery
rate and pressure specifications, but generally power restrictions will prevent them from reaching
both limits simultaneously. For example, a small unit will typically be able to deliver 1,100 scfm
at pressures of up to 3,000 psi, but this rate will then fall if the delivery pressure is increased
towards its delivery pressure rating of 4,500 psi. Large units can often deliver 6,000 scfm at up
to 8,000 psi. The delivery pressure of a cryogenic nitrogen unit will not be a problem for dry
nitrogen drilling.

2-56

Oxygen Required for a Flammable Mixture (%)

12
11.5
11
10.5
10
9.5
9
8.5
8
0

500

1000

1500

2000

2500

3000

Pressure (psia)

Figure 2-21.
22

Influence of pressure on the minimum concentration of oxygen required for a


flammable mixture of oxygen, nitrogen and methane (after Allan, 1994,1 and
Zabetakis, 1964 3).

For drilling applications, the nitrogen pump unit replaces the bank of compressors and the
booster. Since the nitrogen is pumped as a liquid and the conversion from liquid volume to gas
volume, at standard conditions, is well characterized, it is straightforward to accurately measure
the nitrogen delivery rate. Control of delivery rate is also good.
If the pump is at ambient temperature when it is started up, some liquid nitrogen is effectively
lost in cooling the unit down to where the nitrogen remains as a liquid until it is delivered to the
heat exchanger. Before then, formation of gaseous nitrogen in the pump reduces its efficiency.
Controlling the delivery rate is uncertain during cooldown. However, this does not take more
than a few minutes, and should not affect operations substantially.
Each gallon of liquid nitrogen generates approximately 100 scf of nitrogen gas. For a typical
drilling operation, with a volumetric flow rate of up to 2,000 scfm, this means that up to 30 bbl,
or roughly 5 tons per hour, of liquid nitrogen will be required.

2-57

Chapter 2

Underbalanced Drilling Techniques

Membrane Filter
On-site generation of nitrogen by membrane filtration can provide a feasible alternative nitrogen
source. It does not involve transporting and storing large quantities of cryogenic liquid.1 It
requires that air coolers and a filter array are added to an otherwise conventional air drilling
compression system (refer to Figure 2-23).
As shown in Figure 2-23, conventional air compressors deliver the air, at a pressure of 150 psi.
The compressed air is cooled to approximately 80F and passed through a series of primary
filters. These remove contaminants, such as dust, compressor lubricant oil, and atmospheric
water. The air flow then passes through the membrane filter. This consists of an array of many,
very fine, hollow polymeric fibers. The lighter nitrogen molecules pass down the fibers, while
heavier oxygen molecules penetrate the fiber walls. The two gases are separated. Nitrogen is
delivered to the booster unit and then to the standpipe. Oxygen is vented to the atmosphere.
Nitrogen concentration in the gas flow delivered by these membrane filters can be readily
controlled, and can range from 92 to 99.5 percent. Nitrogen purity is controlled by varying the
air input rate and the back pressure on the filter unit.
Other Equipment
Firestops or firefloats in the drillstring are not needed when drilling with nitrogen. Otherwise,
the equipment used is effectively the same as that used for dry air drilling, as described earlier, in
Section 2.1.
2.2.3

Operational Procedures

No special operating procedures, different from those described for dry air drilling, are required
for nitrogen drilling. The risk of downhole fires is removed. Since formation of a mud ring can
cause stuck pipe, timely detection of the symptoms of mud ring formation is still necessary. The
release of large quantities of nitrogen and enriched oxygen into the atmosphere do not normally
pose any risks. Care should nevertheless be taken so that dispersion of the discharged oxygen is
not hindered. Even a modest increase in oxygen concentration can result in dramatic changes in
the combustibility of familiar materials.

2-58

Oxygen Out
Ambient Air In

Compressor (s)

Air Cooler

Water Filter

Hydrocarbon Filter

Oxygen
Filter
Membrane

Particulate Filter
Membrane Skid

Booster(s)

Mist Pump
Nitrogen
Into
Standpipe

Figure 2-23.

Schematic of a nitrogen drilling system using membrane filter generation (after Allan,
1994 1).

Limitations
An appropriate concentration of nitrogen, in the circulating medium, removes the risk of
downhole fires. This overcomes one of the major limitations of dry air drilling. However, the
other limitations of dry air drilling apply equally when nitrogen is used.
The formation of mud rings remains a concern. It is possible to use nitrogen as the gaseous
phase in mist or foam drilling, overcoming problems with excessive water production.
The principal limitation on the use of nitrogen for drilling is economic. The nitrogen supply is
costly, whether liquid nitrogen or membrane filters are used. The quantities of liquid nitrogen
required can easily cost from $10,000 to $35,000 per day of drilling. Allan, 1994,1 cited a daily
incremental cost of over $7,000, to use a membrane filter unit, capable of delivering up to 3,600
scfm of nitrogen. This cost includes rental of the compressors and boosters, and all mobilization.
As a result of its high cost, nitrogen is normally only used when drilling through a long
producing interval, as would be the case in a horizontal well. It is less likely that the additional
cost/ can be justified (particularly cryogenic liquid nitrogen), for drilling long vertical (or near
vertical) intervals, unless multiple zones are likely to be penetrated.

2-59

Chapter 2

Underbalanced Drilling Techniques

There is a possibility that nitrogen could be re-cycled if a closed surface system is used. This
could make the use of nitrogen more attractive, although the savings in nitrogen cost would be
partly offset by the additional daily cost of the surface equipment that might be involved. This
will be discussed again in Section 2.7, Gasified Liquid.

2-60

Chapter 2

Underbalanced Drilling Techniques

References

2-61

1.

Allan, P.D.: Nitrogen Drilling System for Gas Drilling Applications, paper SPE 28320
presented at the 1994 SPE Annual Technical Conference and Exhibition, New Orleans,
LA.

2.

Coward, H.F. and Jones, G.W.: Limits of Flammability of Gases and Vapors, U.S.
Bureau of Mines Bulletin No. 503, Washington, DC (1952).

3.

Zabetakis, M. G.: Flammability Characteristics of Combustible Gases and Vapors, U.S.


Bureau of Mines Bulletin No. 627, Washington, DC (1964).

2.3

Natural Gas Drilling

As was the case with nitrogen, natural gas can be used (instead of air) as the circulating fluid for
drilling underbalanced. Using natural gas will prevent the formation of a flammable gas mixture
downhole when a hydrocarbon producing zone is penetrated. Unlike nitrogen, however, natural gas
will almost invariably form a combustible mixture when it is released into the atmosphere. This
inherently higher potential for surface fires requires few changes in operating procedures from those
used in dry air drilling.

2.3.1

Hole Cleaning

The density of natural gas is generally different from the density of air, at the same temperature and
pressure. It is convenient to represent a gass density by its specific gravity, S. The specific gravity
is defined as the ratio of the weight of a unit volume of the gas at standard conditions to the weight
of the same volume of air. Gaseous hydrocarbons have specific gravitys that increase with
increasing molecular weight. For example, methane has a gravity of 0.55, ethane 1.05 and propane
1.55 at STP. The gravity of natural gas (gas density divided by the density of air, both densities at
STP) depends on its composition, but is often in the range of 0.6 to 0.7. Thus, the density of natural
gas circulating round a well differs sufficiently from air injected at the same volumetric rate that
cuttings transport efficiency is influenced.
Equation (2.33) shows that a cuttings terminal velocity is effectively inversely proportional to the
square root of the density of the fluid through which it is falling:
4g d c c
Vt
(2.33)
3Cd f
where:
Vt ....... terminal velocity (ft/s),
dc ........ diameter of the particle (ft),
c ....... density of the particle (lbm/ft3),
f ........ density of the fluid (lbm/ft3),
g......... gravitational acceleration
(32.17 ft/s2), and,
Cd ....... drag coefficient.
This equation requires that the density of the carrier fluid is much less than that of the cutting. This
is the case when drilling with a gaseous fluid. At atmospheric pressure, the terminal velocity in
natural gas, Vtg, is related to the terminal velocity in air, Vtair, by:
Vtg = Vtair

1
S

(2.34)

2-54
2-62

Chapter 2

Underbalanced Drilling Techniques

The minimum gas injection rate required for efficient cuttings transport at atmospheric pressure
increases in inverse proportion to the square root of the gass specific gravity.
Equation (2.5) shows that the variation of the pressure gradient with depth is proportional to the
density of the gas/cuttings mixture in the annulus. The pressure gradient will be lower for natural
gas than for air, and the downhole pressure will also be lower. However, the mixture density is not,
directly proportional to the specific gravity of the gas. It also depends on the mass of the cuttings
and moisture in the annulus (and hence on the penetration rate). As a result, there is no simple
relationship between the gas pressure downhole and its specific gravity. The full frictional pressure
drop and cuttings transport analysis has to be repeated using the specific gravity of the appropriate
gas. When this is done, the net effect is that the gas injection rate required for efficient cuttings
transport is higher for natural gas than it is for air, but the fractional difference decreases with
increasing depth. The bottomhole pressure is lower for natural gas than it is for air, at an injection
rate providing equivalent cuttings transport. As can be seen from Figure 2-2, Angels analysis
indicates that bottomhole pressures with natural gas are ten to fifteen percent lower than those for
air.

2-62

Chapter 2

Underbalanced Drilling Techniques

An additional complication is departure from ideal gas law behavior. Natural gas is characterized
by a phenomenon referred to as supercompressibility. This means that it compresses more
readily at some pressures than an ideal gas. The deviation from ideal compressibility can be
significant at pressures around 2,200 psi, where it is possible for natural gas to compress to twice
the density that would be estimated for an ideal gas. In most gas drilling situations, downhole
pressures will be well below this level. Given the overall uncertainty of the computations, it is still
probably reasonable to neglect departures from ideal gas compressibility when estimating required
injection rates. If it is necessary to have good control of bottomhole pressure - for example, to
maintain the underbalanced pressure within a specific range - then the real compressibility of
natural gas may have to be considered.
Natural gas is more expensive than compressed air. The incremental cost of increasing the
injection rate is higher when drilling with natural gas than it is when drilling with air. The most
cost effective injection rate is likely to be closer to the recommended minimum rate than would
be the case with air.
Cummings, 1987,1 presented a case history of natural gas drilling in the San Juan Basin, New
Mexico. Gas injection rates of between 900 and 1,200 scfm were found to be adequate for
drilling 6-inch hole to depths of between 6,900 and 7,500 feet, at penetration rates ranging from
40 to 75 feet per hour. The gas used had a specific gravity of 0.65. Angels analysis indicated
minimum air injection rates of between 815 and 950 scfm, for this range of depths and
penetration rates. For natural gas, a first order estimate of the minimum injection rates required
could be derived by taking Angels figures for air drilling at the appropriate depth and
penetration rate and dividing these by the square root of the gass specific gravity. When this is
done, minimum injection rates, ranging from 1,000 to 1,180 scfm, are inferred. These are very
close to the rates that proved acceptable in practice.
When drilling with natural gas, the size of the hole to be drilled should be carefully considered.
Even minor reductions in hole diameter could lead to significant reductions in gas injection rates
and savings. Cummings, 1987,1 noted that increasing the hole size from 6-inches to 6-inches
would have required a fifteen to twenty percent increase in the gas injection rate, while going to a 7
7/8-inch hole would have increased gas consumption (and the gas cost) by as much as 40 percent.

2-63

2.3.2
Equipment
Currently, natural gas drilling is only undertaken if the natural gas can be obtained directly from a
supply pipeline. Usually, the pipeline operator will provide a drill gas unit, which consists of a
scrubbing vessel and an orifice meter,1 connected to the pipeline. A three-inch diameter line is used
to take the gas supply to the rig location. Cummings, 1987,1 noted that this line could be as much
as one-half mile long. The pressure drop along the supply line needs to be considered, to ensure
that adequate delivery is available at the rig site. With a three-inch diameter supply line, the
pressure drop is unlikely to be much more than one psi per 100 feet of pipeline, at a gas flow rate of
1,200 scfm and an average line pressure of 150 psi. This will probably not have significant impact
on gas delivery rate.
Compressors may or may not be required, depending on the pipeline pressure and the anticipated
gas injection (standpipe) pressure. It may be advisable to have a booster available, in the event that
downhole problems require a standpipe pressure higher than pipeline delivery pressure.
It is important to remove, as much as possible, any water or condensate from the supply gas before
injecting it downhole. Otherwise, problems with mud ring formation and wellbore stability may
occur. Cummings, 1987,1 suggested installing a three-phase production vessel between the drill gas
supply unit and the booster. It should have working pressure and temperature ratings that at least
match the pipeline delivery pressure and temperature and it should be able to handle flow in excess
of the highest required gas injection rate. These requirements are not difficult to meet. For example,
the vessel used by Cummings was sixteen inches in outside diameter by six feet long, with a 1,000
psi and a 100F working pressure and temperature. It had a rated capacity of 3,125 scfm (4.5
MMscf/D). Collected water was discharged into a water tank and hydrocarbon vented into a stock
tank.
Downstream of the booster, the gas should flow through an adjustable choke, to control the flow
rate during drilling and tripping. Otherwise, the gas delivery rate can be controlled by a pressure
regulating valve. The gas flow is then directed to a three-inch valve manifold, ideally located on
the rig floor, adjacent to the drillers console. This should allow gas to be sent to the blooie line
primary jet or to the standpipe. It should also be possible to independently vent the gas delivery line
and the standpipe. Both of these vent lines, which should be two-inches in diameter, should run to
a flare pit. Cummings recommended that this flare pit should be separate from the main flare pit
into which the blooie line discharges (refer to Figures 2-24 and 2-25).
2-25).
Additional lines, from the choke manifold direct to the flare pit, and through a flow tester to the
flare pit, should be installed if it is intended to measure the gas production rate. These lines must be
rated for the maximum anticipated well flow rate and wellhead pressure.

If water sensitive shales are penetrated.

2-64

Chapter 2

Underbalanced Drilling Techniques

As natural gas is always present on the rig site, it is sensible to have hydrocarbon gas and
hydrogen sulfide detectors located on the rig floor and in the well cellar. A hydrogen sulfide
detector should be also be sited at the blooie line exit. As with air and nitrogen drilling operations,
care should be taken when locating this detector, to prevent it from being damaged by the flare
while still guaranteeing its exposure to the returning gas flow. There is little point in having a
hydrocarbon detector at the blooie line exit.

2-65

Chapter 2

Underbalanced Drilling Techniques

Booster
Drill Gas Unit
3 Phase Separator
with Meter Run (16" x 6' 1000 psi)

1500 psi
2500 cfm

Water Oil
Dump Dump

Standpipe
Driller's Manifold

3" Line

Supply
Gas Well

Adjustable
Choke

To Blooie
Line Jet

Figure 2-24.

Emergency
Vent Line

Standpipe
Relief Line

Surface equipment required for drilling with natural gas (after Cummings, 1987 1).

F lare P it
7" B looie Line 150 Feet Long

2" Line to B looie Line J et


C hoke M anifold

F lare P it

O rifice
W ell T ester

R otary T able

3" S upply G as Line

2" V ent Line


D og H ouse

Figure 2-25.

2-66

D riller's
M anifold

Flaring arrangements for drilling with natural gas (after Cummings, 1987 1).

Other required surface equipment is similar to that needed for air drilling. Cummings indicated that
a conventional rotating head could be used to divert gas flow into the blooie line in certain
situations (i.e. low permeability wells that need to be hydraulically fractured). Particularly if it is
anticipated that high formation gas pressures or production rates may be encountered, the additional
pressure capacity of a rotating blow out preventer is worth considering.
Large volumes of gas have to be flared when drilling with natural gas. The flare pit has to be
designed accordingly. It may be necessary to construct banks, or even possibly use a vertical flare
stack. Since the flare is larger and is burning whenever natural gas is circulated, there is a greater
risk of fire damage to the hydrogen sulfide detector located at the end of the blooie line, than when
dry air, nitrogen or mist are used.
Instead of using pipeline gas, in rare instances, it may be possible to transport natural gas to the rig
site in large pressure vessels. Typically, these can hold up to 160,000 scf of gas at pressures up to
2,400 psi. The gas would be discharged from the vessel into a compressor and re-compressed to
the required delivery pressure, rather than discharging directly to the standpipe. This allows much
more of the gas in the vessel to be used. Assuming ideal gas behavior, only approximately 10,800
sft3 could be delivered at 200 psi, while 116,000 sft3 of charge could be delivered at an absolute
pressure of 20 psi.
If natural gas were supplied in this way, surface equipment used would be substantially the same as
that used for dry air drilling, except that the compressors and booster would have to be rated for
natural gas service. Requirements for removal of water and hydrocarbon upstream of the standpipe
would depend on the purity of the gas supplied. As with nitrogen, when drilling with natural gas,
fire floats and fire stops are not required in the drillstring.
Using a top drive may provide significant savings in connection times (J. Williams, personal
communication). When using any gaseous fluids, modifications are required to provide adequate
sealing, beyond that used for mud systems. These modifications depend on the specific
manufacturer, i.e. some top drives have integrated swivels while others do not.
2.3.3 Operating Procedures
Operating procedures are generally similar to those used when drilling with dry air or nitrogen. The
gas delivery rate is controlled by adjusting the choke in the supply line. Alternatively, rather than
measuring the volumetric rate, it is possible to control the gas delivery by computing the standpipe
pressure for a desired rate and adjusting the pressure regulating valve to this target standpipe
pressure.
When tripping, the drillstring should be stripped through the rotating head as far as possible, before
pulling the rotating head rubber seal element. The gas flow should then be directed to the primary
jet to keep produced gas away from the rig floor. The cost of replacing rotating head rubbers worn
by the stripping operations is almost invariably lower than the cost of the gas used to jet the blooie
line.
2-67

Chapter 2

Underbalanced Drilling Techniques

While there is no risk of downhole fires when drilling with natural gas, formation of a mud ring can
still stick the pipe. It is possible to mist when using natural gas as the injected gas. If the well starts
to make limited quantities of water, mist drilling should be adopted.
Because the returning gas is flared at all times, water droplets in the return flow are not as readily
observed at the blooie line as when drilling with air. Similarly, low cuttings return may not be
easily detected. However, the standpipe pressure will increase with water inflow or the formation of
a mud ring. It is important that the standpipe pressure gauge have a resolution of 5 psi and that
it is monitored at all times. There should also be a change in the character of the flame at the flare
pit when significant water inflow occurs.
2.3.4 Limitations
In almost all instances, natural gas drilling currently requires that there is a supply of pipeline gas,
within about one-half mile of the rig site. This is probably the greatest limitation on the use of
natural gas as a drilling fluid.
Even when the rig site is close to a supply pipeline, the cost of using natural gas is high, in
comparison to dry air drilling. Allan, 1994,2 cited an example where the daily incremental cost of
drilling with natural gas (at an injection rate of 3,000 scfm) was over $8,500 in the San Juan Basin,
a region where natural gas is readily available. Although directly comparable figures were not
given, it is almost certain that the equivalent cost for dry air drilling would have been less than onehalf of this. The cost of drilling with nitrogen generated on-site was not much lower than for
natural gas.
Using natural gas prevents the formation of a combustible mixture downhole and eliminates the
possibility of downhole fires. The potential for a surface fire is present at all times and the volume
of gas to be handled is greater than it would be for dry air drilling. A typical gas injection rate of
2,000 scfm corresponds to 2.88 MMscf/D, the output of a respectably productive gas well. Careful
attention to the design of surface equipment and to operating procedures is necessary to handle this
volume of gas, although it is not a technical limitation on the applicability of natural gas drilling.
In some locations, such as close to habitation, there may be problems of environmental
acceptability, associated with flaring large volumes of gas. Similar problems would exist if natural
gas were encountered when drilling with dry air, in the same location. Natural gas is unlikely to
pose significantly greater environmental problems than dry air.
Finally, mud rings can still form when water inflows occur, the disposal of produced water remains
costly, and there is still a potential for wellbore instability.

2-68

Safety conformance on the drilling rig using natural gas (in the United States) must
meet the following guidelines:

American Petroleum Institute (API) RP 500B: Recommended Practice for


Classification of Areas for Electrical Installations at Drilling Rigs and Production
Facilities on Land and on Marine Fixed and Mobil Platforms - 1973

National Fire Protection Association (NFPA) 70: National Electric Code - 1990

NFPA 496: Purged and Pressurized Enclosures for Electrical Equipment in


Hazardous (Classified) Locations - 1988

The cost savings while drilling with natural gas should offset any additional expenses
associated with regulatory compliance.

References
1.

Cummings, S.G.: Natural Gas Drilling Methods and Practice: San Juan Basin, New
Mexico, SPE/IADC paper 16167 presented at the 1987 SPE/IADC Drilling Conference,
New Orleans, LA.

2.

Allan, P.D.: Nitrogen Drilling System for Gas Drilling Applications, paper SPE 28320
presented at the 1994 SPE Annual Technical Conference and Exhibition, New Orleans,
LA.

2-69

Chapter 2

2.4

Underbalanced Drilling Techniques

Mist Drilling

2.4.1 Mist versus Foam


It is normal practice to switch from dry air to mist drilling, if a modest water influx is encountered.
To do this, a small quantity of water, with a foaming agent, is injected into the compressed air flow
before it enters the drillstring. This liquid and any produced formation water are dispersed into a
mist of independent droplets of liquid, which move at approximately the same velocity as the gas.
Mist drilling is only one of several different drilling techniques in which the drilling fluid is a twophase mixture of gas and liquid. Other drilling fluids which contain gaseous and liquid phases
include foams and aerated (or gasified) muds. These are sometimes collectively termed lightened
drilling fluids.1 Since the droplets in a mist are not connected to one another, the liquid phase is
discontinuous. In a foam, a continuous liquid phase forms the walls of closed cellular structures
that entrap the discontinuous gaseous phase. Finally, in gasified muds, the gas exists as discrete
and independent bubbles.
As indicated in Chapter 1, the liquid volume fraction largely determines the structure of lightened
drilling fluids. Gases have significant compressibility (their density increases with increasing
pressure), whereas liquids are largely incompressible. As a result, the volume fractions of gas and
liquid in a lightened drilling fluid change in response to the pressure changes that occur as the fluid
circulates round the well. It is quite possible to have more than one fluid structure, such as mist and
foam, present at different points in the same well.
Since mist drilling entails circulating a fluid with essentially the same components as are used in
foam drilling, (water, a foaming agent and a gas), there is a possibility for confusion between these
two drilling techniques. A mist is formed if the liquid volume fraction is below approximately one
to two percent, at the prevailing pressure and temperature. In mist drilling, the volumes of liquid
and gas injected into the well guarantee that the drilling fluid is a mist as it flows down the
drillstring. If there is a significant water inflow, the liquid volume fraction downhole can increase
to where a foam is formed. As the drilling fluid moves up the annulus, the pressure will decrease
and any foam may or may not revert to a mist, before it returns to surface. Consider mist drilling to
refer to those operations where the drilling fluid is a mist for at least part of its circulation history.
During foam drilling, (Section 2.5) the volumes of liquid and gas injected into the well are carefully
controlled. This ensures that foam forms when the liquid enters the gas stream, at the surface. The
drilling fluid remains foamed throughout its circulation path down the drillstring, up the annulus
and out of the well. Foam has a dramatically higher viscosity than dry air or mist. Effective hole
cleaning during true foam drilling will occur at much lower circulation rates than are required for
mist drilling.

2-70

2.4.2 Effect of Pressure on Volume Fractions in Lightened Drilling Fluids


Before discussing mist drilling in more detail, the impact of changing pressure on the volume
fractions and density of a lightened drilling fluid is considered.
Generally, a lightened fluid consists of gas, liquid and solid phases. The solid phase consists of
cuttings that enter the circulating fluid at the bit. The volume fraction, FP, of each component at
any particular pressure, P, is defined as the volume, VP, which the individual component occupies
at that pressure, divided by the volume of the mixture, VmP, at the same pressure:
FgP =

VgP
VmP

VgP
VgP + VfP + VsP

FfP =

VfP
VfP
=
VmP VgP + VfP + VsP

FsP =

VsP
VsP
=
VmP VgP + VfP + VsP

(2.35)

The subscripts g, f, and s denote gas, liquid (fluid) and solids, respectively. The sum of the volume
fractions is, by definition, unity:
FgP + FfP + FsP = 1

(2.36)

Over the range of pressures encountered in most underbalanced drilling operations, it is usually
adequate to assume that the gaseous phase of a lightened drilling fluid acts as an ideal gas and that
the liquid and solid phases are incompressible. That is, if Vo is the volume each component
occupies at pressure, Po, then, under isothermal conditions:
P
VgP = Vgo o
P
VfP = Vfo = Vf

(2.37)

VsP = Vso = Vs
Considering the volume fraction of the gaseous phase first, and substituting from Equation (2.37)
into (2.35),
P
Vgo o
P
FgP =
(2.38)
Po
Vgo + Vf + Vs
P

Assumptions of an ideal gas can be easily changed if necessary, by considering Z, the real gas deviation factor.

2-71

Chapter 2

Underbalanced Drilling Techniques

From this, the gas fraction at pressure is:


Fgo

FgP =

P
Fgo + (Ffo + Fso )
Po
(2.39)
Fgo

P
Fgo + 1 Fgo
Po

Similarly, for the liquid and solid volume fractions:


FfP =

(1 F )

Ffo

go

FsP =

(2.40)

P
+ Fgo o
P
Fso

(1 F )
go

(2.41)

P
+ Fgo o
P

The finite compressibility of the gaseous phase in lightened drilling fluids means that their density
increases with increasing pressure. The presence of incompressible liquid and solid phases means
that the density does not change in direct proportion to the pressure, as would be the case for an
ideal gas. Denoting the mixture density at pressure, P, by mP and the volume occupied by a fixed
mass of the mixture at that pressure as Vmp, then:
V
mP = mo mo
(2.42)
VmP
Assuming that the gaseous phase obeys the ideal gas law and that the liquid and solid phases are
incompressible, it follows that:
mP =

2-72

mo Vgo + Vf + Vs

P
Vgo o + Vf + Vs
P

mo Vmo
P
Vmo Ffo + Vmo Fso + Vmo Fgo o
P

(2.43)

Finally:
mP =

mo
P
1 Fgo 1 o

(2.44)

These equations can be used to demonstrate the significance of pressure changes during mist
drilling. Consider a mist drilling operation where liquid is metered at 10.7 BPH (1 cfm) into 2,000
scfm of dry air, being delivered to the standpipe at 210 psig. The liquid volume fraction at
atmospheric pressure is:
1 cfm (liquid)
2000 scfm (dry air) + 1 cfm (liquid)

1
= 5 10 4 ( 100 0.05% )
2000

To simplify the ratios involved, atmospheric pressure will be assumed to be 15 psi. The ratio of the
absolute standpipe pressure to the atmospheric pressure is (210+15)/15, that is (P/Po) = 15. The air
at the standpipe will be compressed to one-fifteenth of its volume under standard conditions, and
the volumetric air flow rate at this point will be 133 cfm (one-fifteenth of the value under standard
conditions). Substituting these pressure and flow rate values into Equation (2.40), the liquid
volume fraction at the standpipe is 5 10-4/(5 10-4 + 0.9995/15) 7.5 10-3 ( 100 0.75%).
This value is still below the one to two percent threshold at which a foam may start to form. This
drilling fluid will probably remain a mist during all stages of its circulation.
Suppose that there is a water inflow of 96.3 BWPH (i.e. 9 cfm), and that the bottomhole pressure is
135 psig. Also suppose that the volume of cuttings added to the circulating fluid at the bit is
negligible. The total volume of liquid in the mixture flowing up the annulus is now 10 cfm (the
injected plus produced fluid). At the blooie line exit, the liquid volume fraction will be 10/2000 x
100, or 0.5 percent. Presuming water inflow only at the hole bottom, the bottomhole liquid volume
fraction will be 4.78 percent. A foam will probably form when the flowing air and foamer solution
mix with the water flowing into the wellbore. The liquid volume fraction will drop as the foam
moves up the annulus and the foam will have broken down by the time it reaches the surface.
In practice, non-uniformity in the drilling fluid can lead to the formation of slugs of foam and slug
flow. Particularly if the penetration rate is high, the solid volume fraction added to the circulating
fluid at the bit may not be negligible. This will reduce the volume fraction of compressible gas in
the circulating fluid and influence the effect of pressure on the mixture density. This will probably
not greatly influence the circulating fluids structure because it will not change the continuity of the
liquid and gaseous phases.

2-73

Chapter 2

Underbalanced Drilling Techniques

In addition to pressure, temperature also regulates the density and volume fractions of lightened
drilling fluids. A gas expands significantly on heating. The fractional volume increase, at constant
pressure, depends on the fractional increase in the absolute temperature. Assuming ideal gas
behavior, with temperature in R, Equation (2.37) can be re-written as:
P T
VgP ,T = Vgo o
P To

(2.45)

The various equations for density and volume fractions can be modified to incorporate the effects of
temperature changes. In practice, the impact is often small. The surface gas temperature is likely to
be around 60F or 520R. The geothermal gradient is typically around 1.6F per 100 feet. The
bottomhole temperature at 5,000 feet will then be approximately 140F or 600R. At 10,000 feet,
the temperature is about 220F or 680R. Comparing these temperatures, it is seen that the effect on
volume fractions is unlikely to exceed 30 percent and for many underbalanced drilling applications
it will be less than 15 percent. Taking the example given above, if the bottomhole pressure is 135
psig and air is injected at 2,000 scfm, the volumetric air flow rate at bottomhole pressure would be
200 cfm (2,000 15/(135 + 15)) at a bottomhole temperature of 60F. If the bottomhole
temperature were 140F or 220F, the volumetric air flow rate would be 231 or 262 cfm,
respectively. The corresponding liquid fractions would be 4.15 percent and 3.68 percent.
In contrast, pressure changes in a circulating, lightened drilling fluid can lead to more than ten-fold
changes in the volume fractions. In many instances, at least for liquid volume fractions, it may be a
sufficient to neglect temperature changes.
2.4.3 Hole Cleaning
The liquid droplets in mist can be considered to be somewhat analogous to cuttings. They have a
lower density than cuttings (less than one-half of the density of typical cuttings) and they tend to be
smaller than most cuttings. As a result, the droplets, in many cases, can be assumed to move with
the same velocity as the gas;2 this means that their slip velocity is zero. They do not markedly
change the flow properties of the gas in which they are dispersed. Consequently, mist is not
fundamentally more efficient than dry air for transporting cuttings from the wellbore. Theoretically,
high annular velocities are therefore required in mist drilling.
The circulating fluid density is, however, increased by the presence of the liquid droplets in mist.
Droplets may also add to the frictional pressure losses. Both of these factors increase the
bottomhole pressure over that seen with dry air circulating at the same volumetric rate. The higher
fluid density reduces the terminal velocity of the cuttings, in comparison to dry air. At the same
time, the increased bottomhole pressure leads to a lower annular velocity for mist (in comparison to
dry air at the same volumetric injection rate). As a result, higher air injection rates are necessary
when mist drilling, to achieve the same annular velocity as for dry air. The net result of these
various factors is that the air injection rate may need to be increased when changing from dry
air to mist drilling, if the level of hole cleaning efficiency is to be maintained.

2-74

It is possible to adapt Angels method of estimating air injection rates required for adequate hole
cleaning with dry air3 to mist drilling. In this method, cuttings are assumed to travel with the same
velocity as the air. This would be the case for the liquid droplets in mist drilling. Angels analysis
incorporates the influence of cuttings in the annulus (generated at a specified penetration rate) on
the bottomhole pressure and determines the air injection rate needed to give the circulating fluid
kinetic energy (downhole) equivalent to that for air flowing at 3,000 ft/min under standard
conditions.
The first step in modifying the analysis for mist drilling is to determine the penetration rate that
would generate the same mass of cuttings as the mass of liquid entering the well over a unit time
period (i.e., represent the liquid as cuttings). The liquid includes base liquid and foamer injected at
the surface, as well as any formation water inflow into the wellbore. If this total liquid rate is L
(BPH), the mass flow rate of liquid entering the well will be 350.5L lbm/hr, assuming that the
liquid is water (62.4 lbm/ft3). Angel assumed the bulk density of the rock being drilled to be 169
lbm/ft3 (2.70 g/cm3). For consistency, this density is used below. If the bit diameter is Db (inches),
the apparent penetration rate, ROPe (ft/hr), giving the same mass rate (lbm/hr) as the combined
liquid rate L (BPH), is:
ROPe =

350.5 L
380 L
2 =
D 2b
D
169 b
4 12

(2.46)

This apparent penetration rate, ROPe, computed for the expected liquid rate, is added to the actual
anticipated penetration rate ROP. Finally, the minimum air injection rate, required for good hole
cleaning during mist drilling, is determined; either from Angels charts or from the approximation
in Equation (2.17). This would be the air injection rate for dry air drilling the same hole at a
penetration rate of ROP + ROPe.
As an example, consider drilling a 7 7/8-inch diameter hole, at a depth of 5,000 feet, using 4-inch
diameter drillpipe. A penetration rate of 30 ft/hr is anticipated. In this case, using the nomenclature
of Equation (2.17), Qo = 670, N = 65, H = 5000/1000 = 5 (refer to Appendix C), and the minimum
air rate for dry air drilling, Qa (scfm), is:
Q a = Q o + NH

(2.47)
= 670 + 65 5 = 995 scfm
Suppose that, for this example well, liquid is injected into the air stream at the surface, at a rate of 6
BPH. Assume that water is flowing into the well at 3.8 BPH. The total volume of liquid entering
the well is 9.8 BPH (3435 lbm/hr). The penetration rate that would give this mass of cuttings per
hour (Equation 2.46) is 60 ft/hr. The minimum air rate for mist drilling at 30 ft/hr actual
penetration rate should be the air rate required for dry air drilling at a penetration rate of 90 ft/hr.
Using the value of N at 90 ft/hr, N = 98.3 (Appendix C), the minimum air rate predicted would be
1,162 scfm.

2-75

Chapter 2

Underbalanced Drilling Techniques

Mitchell, 1981,4 developed a model for dry air and mist circulation, with cuttings transport. It
accounted for the effect of the masses of cuttings and mist on the momentum of the flowing fluid,
and for the settling velocity of the cuttings in the mist. The model assumed that cuttings were 0.375
inches in diameter, and used drag coefficients from Gray, 1958.5 The liquid droplets were treated
as cuttings with the density of water, moving with the same velocity as the air. Mitchell modeled
mist droplet flow, using an expression for the settling velocity of the droplets. These simulations
confirmed that the droplets travel with the same velocity as the air.
This model also represented temperature changes that occur around a well. Mitchell presented a
sample temperature analysis for a dry, air drilled well. This analysis indicated that flowing air
temperatures, at all points in the annulus, are within 10F of the undisturbed geothermal gradient,
over most of the depth of the well. This supports Angels simplifying assumption that the gas
temperature is equal to the geothermal temperature.
Mitchells model predicts higher dry air volume rates (for efficient hole cleaning) than does Angels
analysis. For an 8-inch diameter hole, drilled using 5-inch drillpipe, at a penetration rate of 90
ft/hr, Mitchell computed that an air injection rate of 1,575 scfm would give a downhole air velocity
equivalent to 3,000 ft/min under standard conditions. Angels analysis indicated that an injection
rate of 1,387 scfm would give this air velocity. Mitchell did, however, find that the lower air rate
would be sufficient to lift the 0.375-inch diameter cuttings uphole. Slippage between cuttings and
air would increase the cuttings load in the annulus, leading to a higher bottomhole pressure and
lower air velocity than would be predicted by Angels method.
Mitchell compared dry air and mist drilling for an 8-inch diameter hole, drilled with 5-inch
drillpipe, at 90 ft/hr, to a depth of 10,000 feet. Mist was generated by injecting liquid at 2 BPH.
This is rather low, in comparison with the 6 to 12 BPH liquid rates frequently used in practice.6 He
used the air injection rates for dry air drilling indicated by Angels analysis as the minimum
required for effective hole cleaning. Mitchell computed the downhole air velocity, at these air rates,
using his model. The air injection rates were then determined that would give the same air velocity
downhole for mist drilling. The air injection rates are compared in Figure 2-26. The rates for mist
drilling were typically 30 to 40 percent higher than the dry air rates. Mist drilling standpipe
pressures and cuttings velocities were also 30 to 40 percent higher than for dry air drilling. The air
injection rates shown for mist drilling are higher than would be necessary to achieve equivalent
cuttings velocities and efficient hole cleaning for dry air.
As indicated earlier, it is possible for a foam to form when a significant water influx is encountered.
Foam has a much greater viscosity than mist or dry air and is much more efficient at transporting
cuttings. The various analyses described above have assumed that the circulating fluid remains a
mist everywhere in the annulus. From the point-of-view of hole cleaning efficiency, this is a
conservative assumption. Standpipe pressures will, however, increase if a foam forms and this
could influence equipment specification.

2-76

Field experience provides guidance on air injection rates and pressures for mist drilling. Cooper et
al., 1977,7 reported that mist drilling requires 30 to 40 percent higher air injection rates for
acceptable hole cleaning; in comparison to dry air drilling. These higher injection rates lead to
standpipe pressures during mist drilling that are typically 100 psi (30 to 50 percent) higher than
those in dry air drilling.
Equipment
The liquid phase in a drilling mist is normally water, a surfactant, and a corrosion inhibitor. The
surfactant, often referred to as soap, reduces the surface tension of the water and promotes the
formation of fine, stable liquid droplets in the flowing gas. These same surfactants are commonly
used, at higher concentrations, to create stable foams (Section 2.5).

2-77

Chapter 2

Underbalanced Drilling Techniques

Required Volumetric Flow Rate (scfm)

2500

Mist
5-inch Drill Pipe
8 3/4- inch Bit
3 3/4- inch Nozzles
600 ft of 8- inch Drill Collars
Drilling Rate: 90 ft/hr
Mist: 2 bbl/hr Water

2000

1500

Air

1000

500

0
0

1000

2000

3000

4000

5000

6000

7000

8000

9000

10000

Depth (feet)

Figure 2-26.

Comparison of air injection rates for dry air and for mist drilling giving the same air
velocity downhole (from Mitchell, 19814).

The corrosion inhibitor (refer to Section 2.5) protects the drillstring and any exposed casing strings
if, as is generally the case, air is used as the gaseous phase in mist drilling. The corrosion inhibitor
must be compatible with the foaming agent.
Mist is created by injecting the liquid into the gas flow, between the gas delivery system and the
rigs standpipe. There can be benefits in using separate water and foaming agent injection systems.7
This allows the water and foaming agent to be controlled independently. This can be useful, for
example, if slug flow occurs when a significant water influx is encountered downhole. On the other
hand, there are many successful mist drilling operations that do not use a separate foaming agent
injection pump; the foaming agent and other additives are mixed with the water in the mist pump
tanks.
Mist injection rates are often in the range of 6 to 20 BPH, depending on hole size.7 A 40 to 50
horsepower triplex pump, capable of delivering 25 to 35 gpm (36 to 50 BPH), will normally be
adequate. The pumps delivery pressure must be at least equal to the highest anticipated gas
delivery pressure. The size of the water reservoir should be large enough to avoid frequent refilling,
but small enough for accurate monitoring of the water injection rate. A 12 bbl reservoir has been
recommended for routine mist drilling operations.7

2-78

Skid-mounted mist pump units are generally available in those areas where air and mist drilling are
common. A typical mist pump will come with two compartmentalized tanks on the same skid. The
tanks each have a volume of 10 to 20 barrels and resemble the displacement tanks on a cementer.
The tanks are usually equipped with a simple gauge, consisting of a steel rod in the tank, with
marks for each barrel of tank volume. Mist injection rates, which are usually reported in barrels per
hour (BPH), can be measured adequately using these gauges.
Enough water must be available on location, to refill the water reservoir without interrupting
drilling. It is possible to recycle water from the reserve pit. Although, this will greatly reduce water
requirements, several factors must be considered before reusing this water. First, any injection
water must be relatively solids free; a high solids content could damage the water injection pump.
This means that water returning from the well has to stay in the pits long enough for all cuttings to
settle out. There has to be sufficient water above the cuttings and the pit bottom so that it can be
drawn off without entraining solids. Second, formation water lifted from the well must not
interfere with the foaming agent. Finally, it can be difficult to determine the concentration of
various additives, such as foaming agents and corrosion inhibitors, remaining in the recycled water.
If water is to be recycled from the reserve pit, a suitable air-driven or centrifugal pump should be
rigged up to transfer the water from the pit to the injection pump reservoir. The suction hose should
be fitted with a good strainer. It should be supported to prevent the suction end from falling too
close to any solids in the pit.
If a separate foaming agent injection pump is used, it will need to deliver from 0.25 to 5 GPH, or
occasionally more.7 An air-operated pump may be adequate. It should have a delivery pressure that
matches or exceeds the highest anticipated gas delivery pressure. It requires a much smaller
reservoir than is necessary for the water injection system. It is often possible to feed the foam pump
directly from the drums containing the foaming agent.
When separate water and foamer injection pumps are used, they are normally manifolded together
before entering the air header, downstream of the compressors and booster. Isolating and check
valves should be installed in the delivery lines from both pumps. This will allow them to be
disconnected, if required, and will prevent air from entering the pumps when they are operating.
Figure 2-6 shows a typical surface system arrangement, with separate mist water and foamer
injection pumps.
Since air injection rates for mist drilling tend to be 30 to 40 percent higher than those for dry air
drilling, an additional compressor or compressors may be required. Standpipe pressures during
mist drilling can be 100 psi higher than during dry air drilling. For deeper wells or when a
downhole motor is used, a booster may be needed when compressors alone might have been
adequate for dry air drilling.

2-79

Chapter 2

Underbalanced Drilling Techniques

Normally, the return flow of mist and cuttings is directed to a system of flare and reserve pits,
similar to that sometimes used in dry air drilling. When mist drilling, significant volumes of
formation water may have to be contained at the surface, possibly more than 2,000 BWPD. Surface
equipment must be capable of holding this liquid until it can be properly disposed of. In some
locations, it may be possible to arrange for continuous disposal of the returning liquid, if all
additives and any produced water are environmentally benign. If the well is close to others in a
producing field, it may be possible to dispose the waste water down a water injection well. In some
instances, it may be possible to re-inject the water into a permeable zone, cased-off above the
interval being drilled.
Otherwise, the waste water will have to be hauled off in tankers. The logistics for liquid collection
and disposal have to be confirmed before starting drilling operations. It is probably appropriate to
design surface equipment with the capacity to contain one or two days worth of liquids; any
disruption of liquid removal could force drilling to be stopped. The pits liquid holding capacity is
in addition to the volume required for cuttings.
It is useful to be able to measure the rate at which liquid is returning from the well. Comparing this
with the liquid injection rate indicates the rate of any influx. One way to do this is to route the
returns initially through a steel tank, comparable in size to a typical mud pit. Since cuttings will
tend to accumulate in this tank, there should be ready access, so that these drilled solids can be
periodically removed with a grab-crane or bucket digger. The liquid will have to be defoamed first.
During mist drilling, the liquid volume fraction in the circulating fluid is normally too low for foam
to exit the blooie line. Foam returns may occur if there is substantial water influx or if the water
injection rate is excessive. Foamer remaining in the returned liquid can lead to foaming in the
cuttings pit, if it is agitated for any reason. It is advisable to be prepared for defoaming, particularly
if substantial water inflow is expected. Adequate foam control should be possible by chemical
means alone. Defoaming agents will be described in Section 2.5, Stable Foam Drilling. A
suitable defoamer should be sprayed onto the pit if the liquid foams there, or injected into the blooie
line if foam returns are experienced. A ring-type spray, in which the defoamer enters the blooie line
through a number of small nozzles, can help in defoaming by mixing the defoamer uniformly with
the foam. This unit should be located three to six feet from the exit end of the blooie line.
Other surface equipment is the same as for dry air drilling. There is no fundamental reason why a
mist should not be formed with other circulating gases, such as nitrogen or natural gas. If this is
done, the mist and soap pumps should be incorporated into the gas delivery manifold, upstream of
the standpipe manifold. The increased gas flow rate and standpipe pressure required for mist
drilling would have to be considered when designing the gas delivery system. Particularly for
misting with natural gas, the increased surface pressure might require additional compres-sors or
boosters.

2-80

2.4.5

Operating Procedures

Most mist drilling operating procedures are similar to those used when dry air drilling. Procedures
that are specific to mist drilling are described below.
With a twin-tank mist pump unit, the mist liquid is pumped out of one tank, while the other is
isolated and re-filled with water. After filling, corrosion inhibitor and any other additives are mixed
in the tank. The tank can be physically stirred or rolled using a small amount of air from the
compressors. The foaming agent is added last to prevent excess foaming.
Liquid additives are much easier to mix in than solids. Solids often require long periods of stirring
before they go into solution. Potassium chloride (KCl) is a common additive used to help stabilize
shales. It is supplied in bags. Large quantities of undissolved KCl often remain on the bottom of
the mist tank when the job is complete. Powdered polymers, used for shale stabilization, are even
worse and can form small globules because of incomplete dispersion in the water. These globules
are commonly called fish eyes. If anything other than liquids is mixed with the mist liquid, this
should be done in the rigs premix or suction tank, using the mud hopper. The shearing action
through the mud hopper does a much better job of dispersing and dissolving powdered additives.
The premix can be transferred to the mist tanks with a small centrifugal pump. The foaming agent,
however, should always be added to the mist tank. Otherwise, it will foam excessively during
circulation through the hopper and pumping to the mist tank.
When separate water and foaming agent injection systems are used, corrosion inhibitor can be
added either to the water or to the foaming agent reservoir. If the foamer injection rate changes
during drilling, it may be preferable to add the corrosion inhibitor to the injection water reservoir.
The water injection rate is less likely to be varied. This will help to ensure that adequate volumes
of inhibitor are pumped.
Water and foaming agent injection rates are dictated by hole conditions. Cooper et al., 1977,7
suggested that a good rule-of-thumb was to pump one and one-half barrels of water per hour per
inch of hole diameter. They also suggested that the foaming agent should be injected at between
0.5 and 6 GPH. Suitable concentrations will vary, depending on the foaming agent. Typically,
mist drilling requires only around one-fifth of the concentration of foamer that would be needed
for foam drilling. For example, if 0.75 to 2 percent foamer is required to create a stable foam,
then the concentration of the same agent for mist drilling would probably be in the range 0.1 to
0.5 percent. If the injection water is re-cycled from the reserve pits, there will be some foamer in
this water and it may be possible to reduce the rate at which foamer is added to the injection
water. It may be possible to estimate the foamer concentration in the re-cycled water if the water
inflow rate can be determined. If this cannot be done, it is probably appropriate to maintain the
foamer injection rate, unless foam slugging occurs.
The type and volume of fluid flowing from the blooie line should be monitored carefully when mist
drilling. It is important that there are continuous returns. If the water injection rate is too low, a
mud ring may form and restrict circulation, with the attendant risk of stuck pipe or a downhole fire.
If returns stop, the water injection rate should be increased immediately.
2-81

Chapter 2

Underbalanced Drilling Techniques

Slugs of liquid will be seen at the blooie line if the liquid in the well is not dispersed as
droplets. This happens if the gas injection rate is too low or if the concentration of foaming agent is
too low. Slugging causes the standpipe pressure to fluctuate markedly.
Normally, a foaming agent concentration of between 0.1 and 0.25 percent should be sufficient for
misting.6 If the foaming agent concentration in the returning liquid is in this range or higher and
slugs of liquid are seen at the blooie line, the air injection rate should be increased.
The rate of water inflow will determine what corrective action should be taken if the foaming agent
concentration is too low. Usually, the foamer concentration should be increased until the return
flow becomes steady.
If there is significant water influx, it is tempting to reduce the water injection rate if the well begins
slugging. If this were done, it would not be necessary to increase the foamer injection rate.
However, this should be considered only when it is certain that the bit is still drilling the water
producing formation. Once the bit is below the inflow, it is possible for cuttings traveling up the
annulus to form a mud ring when they mix with the inflowing water at the bottom of the water
producing zone. If the liquid rate has been reduced, it should be re-established when drilling below
the water zone.
It is critical to avoid too high a foamer injection rate. This leads to unnecessarily high chemical
costs and can cause slugs of high viscosity foam. The standpipe pressure will fluctuate widely as
each foam slug leaves the well. The high viscosity of foam increases the pressure drop up the
annulus and causes the standpipe pressure to increase. The foaming agent concentration should be
reduced until steady flow is re-established. The liquid volume fraction downhole must be exceeding
the threshold for foam formation, at least intermittently. Reducing the water injection rate may also
help to stabilize flow.
Water is probably flowing into a mist drilled wellbore when the drillstring is tripped. When the
string is run back in the hole, water may have accumulated in the wellbore. The well will have to
be unloaded before drilling can resume. Procedures for doing this were given in Section 2.1, Dry
Air Drilling. If the amount of water is significant, it may not be advisable or possible to break
circulation with the bit on bottom. If this is the case, the hole will have to be unloaded in stages
while tripping in, as would be done to remove water from the casing after cementing. The length of
each unloading stage will have to be smaller than when emptying casing because formation water
will continue to enter the wellbore while tripping after each unloading stage. The hydrostatic
pressure that will have to be overcome is due to the length of each stage plus the influx volume.
After staging to bottom and before resuming drilling, the well should be circulated for some time to
reduce the amount of water in the annulus.

2-82

2.4.6

Limitations

The main reason for mist drilling is to avoid forming mud rings, if a water producing zone is
penetrated when dry air drilling. As discussed in Section 2.1, Dry Air Drilling, a mud ring is
often the precursor to stuck pipe or a downhole fire. Water injected during mist drilling saturates
the cuttings. This, in combination with the surfactant properties of the foaming agent, prevents
the cuttings from sticking together downhole. Also, liquid in the circulating fluid significantly
increases its thermal capacity. This reduces the temperature increase that occurs if the circulating
fluid is compressed by a flow obstruction, further decreasing the potential for ignition. Misting
greatly reduces the probability of a downhole fire, overcoming one of the major limitations of dry
air drilling.
If the annular velocity is insufficient to lift rock fragments, it is possible for the annulus to close,
even if there is no mud ring. This may happen in highly deviated or horizontal holes, where the
circulation rates required for efficient hole cleaning are much higher than in vertical and nearvertical holes. It may also happen when large fragments spall from an unstable formation into
the wellbore. Once the annulus has packed off, the gas below will be compressed if circulation is
continued. If the gas in the annulus is combustible, there is a potential for downhole ignition,
just as if the annulus were closed by a mud ring.
Hole drag and an increase in standpipe pressure indicate that an accumulation of cuttings and/or
rock fragments may be beginning to pack off the annulus. If this is suspected, pull off bottom to
stop generating cuttings that would otherwise add to the obstruction. Continue circulating and
work the pipe in an attempt to break up the obstruction. Do not attempt to pull the string without
air circulation. The standpipe pressure will continue to rise until the obstruction is cleared or
circulation is shut off. A downhole fire or stuck pipe both require fishing or sidetracking. The
decision on how high the standpipe pressure can be allowed to rise before shutting off
circulation, if the string has not been freed, will be influenced by hole conditions. It also depends
on the bottomhole assembly. If the risk of downhole fire is great, (i.e., if the well is producing
wet gas), and the bottomhole assembly includes expensive components (such as an MWD unit or
a downhole motor), circulation should be shut down earlier than if the only components at risk
are the bit and a few drill collars.
Cooper et al., 1977,7 reported successfully blowing free a stuck pipe, using high pressure
nitrogen from an oilfield liquid nitrogen unit. They noted that this procedure was only likely to
be successful with a pipe that had not been pulled excessively; where the cuttings had not been
packed too tightly.
Mist drilling helps to reduce the impact of water inflow on drilling operations. The foamer
disperses water into droplets downhole. It should, therefore, reduce the air flow rate at which the
transition from slug to mist flow occurs downhole. That is, mist drilling permits larger volumes
of water to be lifted from the well than would be possible in dry air drilling, while avoiding
unsteady slug flow. However, the ability to handle larger water inflows does require greater
compressor capacities.
2-83

Chapter 2

Underbalanced Drilling Techniques

The other limitations of dry air drilling remain, particularly those relating to wellbore instability
and hydrocarbon inflows. Mist drilling has its own characteristic limitations, including increased air compression and waste water disposal demands, additional potential for wellbore
instability as a result of the misting liquid injected into the well, and corrosion of downhole
equipment. These limitations are discussed below.
Air Compression
Mist drilling usually requires air injection
rates that are 30 to 40 percent higher than are required for dry air drilling at the same depth and
penetration rate. Similarly, standpipe pressures are higher than they would be for dry air drilling,
typically by around 100 psi. Not only will more compressor capacity be required, but it is also
more probable that the booster will need to be used. These increased air requirements lead to
higher daily fuel costs. They may not, however, lead to higher equipment rental charges. The
compressor capacity on site depends on the capacity of the individual units available in that area.
For example, consider a 7 7/8-inch hole, being drilled to 6,000 feet, using 4-inch drillpipe.
Angels charts indicate that a minimum dry air injection rate of 1,200 scfm is required. If 1,000
cfm rated compressors are the only units available, two would be rented. Even when dry air
drilling, these two units would probably be used at more than the minimum circulation rate,
perhaps at 1,500 scfm. Mist drilling would require an air rate about 30 percent greater than the
1,200 psi minimum, which would be little different from the 1,500 scfm in use. This suggests
that there would be no automatic requirement to have an additional compressor on site in order to
switch from dry air to mist drilling in this case. The procedure normally followed is to start
misting and then to determine whether or not more air volume is required, by observing hole
conditions. If hole cleaning problems are encountered, an additional compressor can be added.
If suitable compressors are not readily available on location and there is the prospect of having to
drill with mist, appropriate compressors should be selected at the outset.
Waste Water Disposal
The cost of waste water disposal can be an economic limitation. Typically, 200 to 500 bbls of
injected water will be circulated through a well each day. This water is not normally recirculated, and incurs disposal costs; often more than $1 per barrel.
When a large water inflow occurs, the produced water can rapidly exceed the surface storage
capacity. In some areas, large reserve pits are constructed to handle anticipated water production.
Once these are filled, the options are to abandon mist drilling and mud up, or to haul away some
of the produced water for off-site disposal. Cost per foot calculations indicate when it will be
more cost effective to mud up.
The various additives injected into the well, including the foaming agent, corrosion inhibitor and
any salts or polymers, add to the well cost. Finally, the supply cost of the injected water may be
significant in remote locations. These cost-related factors all serve to limit the application of
mist drilling.

2-84

Wellbore Instability
When drilling with dry gas, wellbore instability can occur because of the large difference
between the stresses in the rock adjacent to the wellbore and the pressure of the drilling fluid in
the wellbore. The wellbore pressure is normally higher when drilling with mist than with dry
gas, but the difference is not very large in comparison with the in-situ stresses. There is little
prospect for mist drilling to improve wellbore stability if mechanically-induced instability
problems have been encountered when dry air drilling.
During mist drilling, the gas flow rate tends to be higher and the density of the circulating fluid is
greater than it is for dry air drilling. These factors increase the potential for wellbore erosion, if
weak or poorly consolidated formations are penetrated. The much lower annular velocities
experienced during stable foam drilling may be more appropriate if wellbore erosion is
suspected.
The aqueous phase in the mist drilling fluid can cause chemically-induced wellbore instability if
water-sensitive shales are encountered. These shales tend to dehydrate and slough into the
wellbore during dry air drilling. On the other hand, they will tend to hydrate and swell, leading
to undergauge hole, on exposure to certain circulating liquids during mist drilling. Swelling may
be reduced by the addition of salts, such as potassium chloride (KCl), to the injected water.8
These reduce the activity of the water phase and this decreases the rate of shale hydration.9 It is
also possible to add polymers to inhibit shale hydration. These include polyanionic cellulose
(PAC) and partially hydrolyzed polyacrylamide (PHPA). Since these can add considerably to the
cost of the well, it may be more cost effective to switch to conventional mud drilling if shale
hydration is leading to wellbore instability.
Attempts to chemically stabilize shale may not, in any case, prove very successful. One example
of mist-induced wellbore instability is the Wolfcamp shale, in West Texas. There are very few
sloughing problems if these shales are drilled with dry air. Formations above the Wolfcamp
produce water. If these are left open, drilling cannot continue into the Wolfcamp without
misting. However, if mist is used, sloughing occurs. Often the well cannot reach total depth
without having to mud up. Some operators have elected to run intermediate casing to isolate the
water producing zones. The Wolfcamp interval can then be dry air drilled and sloughing is
normally not experienced. Even though analysis of the shale suggested that it was not
particularly sensitive to brine exposure, sloughing occur-red even when salts were added to the
mist water.

2-85

Chapter 2

Underbalanced Drilling Techniques

Corrosion
During mist drilling, there is a significant potential for rapid corrosion of downhole equipment.
When mist is formed with compressed air, there is a high oxygen concentration in the aqueous
phase. This promotes corrosion of exposed steel. The rotating drillstring impacts against the
hole wall and the casing. The resultant local deformation creates anodic regions that are more
prone to corrosion than undeformed steel. Any oxide film that forms on exposed steel downhole
tends to be removed by these impacts and by the erosive action of the cuttings being carried
uphole. This allows corrosion to proceed unhindered. Any ferrous steel equipment downhole
will corrode rapidly if protective measures are not taken.
The best protection against downhole corrosion is adding a suitable inhibitor to the injected water
or the foaming agent. The corrosion inhibitor must be compatible with the foamer and with any
other chemicals added to the injected water. Since many foaming agents used in mist and foam
drilling are anionic, anionic corrosion inhibitors will normally be required. Of the readily
available anionic inhibitors, complex organo-phosphate esters are probably the most successfully
and widely used in mist and foam drilling applications. Film-forming inhibitors are not usually
successful (they are the most commonly used inhibitors
in liquid environments). They are not normally compatible with foaming agents, any film they
form will be continuously disrupted downhole, and they are not particularly successful in
preventing corrosion in the presence of high dissolved oxygen concentrations.

References

2-86

1.

Rankin, M.D., Friesenhahn, T.J. and Price, W.R.: Lightened Fluid Hydraulics and
Inclined Boreholes, paper SPE/IADC 18670 presented at the 1989 SPE/IADC Drilling
Conference, New Orleans, LA, February 28-March 3.

2.

Govier, G.W. and Aziz, K.: The Flow of Complex Mixtures in Pipes, Robert E. Kreiger
Publishing Company, Malabar, Florida (1982) 419.

3.

Angel, R.R.: Volume Requirements for Air or Gas Drilling, Pet. Trans., AIME (1957)
210, 325-330.

4.

Mitchell, R.F.: The Simulation of Air and Mist Drilling for Geothermal Wells, paper
SPE 10234 presented at the 1981 SPE Annual Fall Technical Conference and Exhibition,
San Antonio, TX.

5.

Gray, K.E.: The Cutting Carrying Capacity of Air at Pressures Above Atmospheric, Pet.
Trans., AIME (1958) 213, 180-185.

6.

Shale, L.: Underbalanced Drilling Equipment and Techniques, ASME paper PD-Vol. 65,
Drilling Technology, presented at the 1995 ASME Energy and Environment Expo '95,
Houston, TX, January 29-February 1.

7.

Cooper, L.W., Hook, R.A. and Payne, B.R.: Air Drilling Techniques, paper SPE 6435
presented at the 1977 SPE Deep Drilling and Production Symposium, Amarillo, TX.

8.

Carden, R.S.: Technology Assessment for Vertical and Horizontal Air Drilling Potential
in the United States, Final Report, Contract No. DOE/MC/28252-3514 (DE94000044),
U.S. DOE (August 1993).

9.

Hale, A.H., Mody, F.K. and Salisbury, D.P.: The Influence of Chemical Potential on
Wellbore Stability, SPEDC (September 1993) 207.

2-87

2.5

Stable Foam Drilling

Foam can be used as the circulating fluid during drilling and for many well completion and
production operations. Foams consist of a continuous liquid phase, forming a cellular structure
that surrounds and entraps a gas. Foams can have extremely high viscosities; in all instances
their viscosities are greater than that of both the liquid and the gas that they contain. At the same
time, their densities are usually less than one-half that of water.
With this combination of high viscosity and low density, foamed drilling fluids can provide
several benefits to drilling operations.
 The high viscosity allows efficient cuttings transport, at annular velocities that are much
lower than those required for dry air or mist drilling. The gas injection rates for foam drilling
can be much lower than those for dry gas or mist drilling.
 The low density of foam allows underbalanced conditions to be established in almost all
circumstances. Bottomhole pressures with foam tend to be higher than those in dry gas or
mist drilling. This may reduce penetration rates below those for dry gas. However,
penetration rates with foam are often still considerably higher than can be achieved in mud
drilling.
 The higher annular pressures with foam can potentially reduce mechanical wellbore
instability experienced when drilling with a dry gas or mist. At the same time, the low
annular velocities, typical of foam drilling, greatly reduce the possibility of erosion of the
borehole wall or the drillstring.
Air is most commonly used as the gaseous phase in foam drilling. It is possible to make foam
with other gases. The liquid phase is aqueous. Because the liquid phase is continuous, a foam
formed with air will not normally permit combustion of produced hydrocarbons. Air foams are
frequently used in fire-fighting to extinguish burning hydrocarbons.
Probably the greatest benefit of foam as an underbalanced drilling fluid, and the main reason for
its use, is its ability to lift large quantities of produced liquids. When water inflows are too large
to be efficiently lifted with mist, it is often possible to continue drilling underbalanced by
switching to foam.
During foam drilling, a surfactant solution is mixed with a gas flow and the mixture is injected
into the drillstring. Foam generation may not be complete until the gas and liquid phases are
mixed thoroughly as they pass through the bit. In pre-formed stable foam drilling, the foam is
intentionally formed at the surface by mixing the surfactant and gas in a foam generator. The
term foam drilling will be used in this report for those operations in which the circulating fluid
remains a foam at all points in the well. Circulating conditions have to be chosen or controlled to
ensure this. Principally, the circulating pressures have to guarantee that the liquid volume
fraction remains within a reasonably well defined range. In this way, the beneficial properties of
foam, notably its viscosity, can be adequately exploited.

2-88

Chapter 2

Underbalanced Drilling Techniques

2.5.1 Foams
A foam consists of an aggregate of gas bubbles in a continuous liquid matrix.1 Water alone will
not form a foam - any bubbles that are created coalesce as soon as they touch one another. A
surfactant, or foaming agent, in the liquid phase, stabilizes the films that form the bubble walls
which allow the foam structure to persist.
Bubble Shape
Foams are often classified according to the shape of the bubbles which they contain.1,2 In a
freshly generated foam, or one containing very small bubbles, the bubbles will be spherical.
These are called sphere foams. Conversely, polyhedron foams contain polyhedral bubbles.
Generally, sphere foams have higher liquid volume fractions than polyhedron foams - undistorted
spheres do not pack as closely together as do polyhedra, and the liquid phase between spherical
bubbles must be thicker than that between polyhedral bubbles. In an ideal polyhedron foam,
consisting of equally-sized bubbles, the bubbles would be pentagonal dodecahedra, that is
twelve-sided, with each side consisting of a pentagon. It is possible to achieve perfect close
packing with this shape, minimizing the liquid volume required to form the films between
bubbles. In practice, a number of factors prevent foams from containing equally sized and shaped
bubbles.
Quality and Texture
Two other terms used to characterize foams are quality and texture.3 A foams quality is its gas
volume fraction, expressed in percent. A 65 quality foam contains 65 percent gas by volume,
while a 90 quality foam contains 90 percent gas by volume. A low quality foam (wet foam)
contains more liquid than a high quality foam (dry foam). Texture describes the size and
distribution of the bubbles. A fine foam has small bubbles and a coarse foam has large bubbles.
Combining these various terms, a sphere foam tends to be a low quality, fine foam and
conversely a polyhedron foam tends to be a high quality, coarse foam.
Once a foams quality exceeds some threshold level, the liquid phase becomes discontinuous and
the foam breaks down into a mist of dispersed droplets. The upper limit for a stable foam is not
clearly defined. It depends on the shear rate. Okpobiri and Ikoku, 1986,3 found that foam
collapsed to mist at a quality of 94 percent, for shear rates below 5,000 s-1, but would persist up
to 96 percent quality, for shear rates above this level. Beyer et al., 1972,4 reported foam
becoming unstable at qualities of 97 to 98 percent, and that foam flowed as slugs of foam and
gas, once the quality exceeded 98 percent.
The upper limit of foam stability also depends on the composition of the liquid phase. Russell,
1993,5 reported foam drilling operations in which a quality of at least 99.1 percent was required
to generate a stable foam using surfactant solution alone, but that the addition of polymeric
viscosifiers to the liquid phase gave stable foam at qualities as high as 99.65 percent. It is not
clear in this case, how much, if any, liquid was being added to the foam by formation water
inflows as the foam traveled uphole, but the effect of the viscosifiers was clear.

2-89

The lower limit of foam stability is more a question of definition. Bubbles dispersed in liquid
can remain stable down to vanishingly small gas volume fractions. The bubbles do not interact at
all, in flowing foam, until the quality reaches about 55 percent.3,6 If the quality exceeds about 75
percent, flow is dominated by deformation of adjacent bubbles, and the foam viscosity increases
rapidly with increasing quality. It is probably this observation that led Rankin et al., 1989,7 to
describe stable foams used for drilling as having qualities ranging from 75 to 97.5 percent. A
lightened fluid will be regarded as a stable foam if its liquid phase is continuous and its quality
exceeds 55 percent.
Breaking
All foams are ultimately unstable. However, fine, low quality sphere foams decay much more
slowly than do coarse polyhedron foams. Two processes disrupt the foam structure - thinning of
the bubble walls, and growth of large bubbles at the expense of smaller ones.
Bubble walls thin due to gravity. The bubbles tend to rise towards the top of the foam, while the
liquid drains through the bubble walls towards the base of the foam. At some point, the walls
become so thin that they rupture. This is energetically favored since it reduces the surface area of
the liquid and therefore reduces the surface free energy. Stirring a sphere foam to re-distribute
the bubbles can prevent excessive thinning. Any agitation of a high quality polyhedron foam
will, however, promote rupture of the thinned bubble walls.
Surface tension of the liquid in a bubble wall tends to collapse the bubble. This is balanced by
the gas pressure inside the bubble. The gas pressure within a bubble is inversely proportional to
the bubble size. When a large bubble contacts a smaller bubble, the higher gas pressure inside the
smaller cell causes the gas inside it to diffuse through the liquid separating the two bubbles, until
the smaller bubble is fully absorbed by the larger.
Foams can be stabilized by increasing the strength of the bubble walls and by slowing drainage
of liquid through those walls. Surfactants strengthen the bubble walls against excessive thinning.
So too can certain proteins which, if present in the liquid phase of an air foam, react with oxygen
at the air-liquid interface, to form a skin. Increasing the bulk viscosity of the liquid phase slows
drainage. Surfactant mixtures can increase the surface viscosity of the base fluid and this can
also slow drainage through bubble walls.
2.5.2 Foaming Agents
The primary agents used to generate foams are surfactants. All surfactants consist of molecules
having a hydrophilic group attached to a long hydrophobic tail, which is usually a fatty
hydrocarbon chain.8 They are classified according to the nature of the hydrophilic group; which
may be anionic, cationic, amphoteric, or non-ionic. Whenever possible, surfactant molecules
orient themselves so that the hydrophilic group is in an aqueous environment and the
hydrophobic tail is in a non-aqueous environment. They will therefore concentrate at the liquidgas interfaces in foams. Here, they may increase or decrease the surface tension of the liquid and
may strengthen or weaken the bubble walls. Not all surfactants will act as foaming agents - some
will destabilize the foam structure and can be used instead as defoamers.
2-90

Chapter 2

Underbalanced Drilling Techniques

At one time or another, all classes of surfactant have been considered for drilling applications.
Ammonium salts of alcohol ether sulfates are probably the most widely used. These anionic
surfactants are highly soluble in most liquids. They provide a foam with good thermal stability
and are suitable for use at very low surface temperatures. Their main drawback is comparatively
high cost. Other anionic surfactants that can be used as foaming agents are alpha olefin
sulphonates. These are inexpensive, perform well in fresh water and are very resistant to
hydrocarbon contamination. They do not, however, work well in brines and will not tolerate low
surface temperatures.2
Very often, commercial foaming agents contain both anionic primary foaming agents and foam
boosters. The latter may be ethylene or propylene glycol butyl ethers, or amphoteric betaines.
Their function is primarily to enhance foam stability.
Cationic surfactants are not commonly used as foaming agents in drilling operations. Quaternary
ammonium chlorides do not perform well in fresh water, give poor to mediocre foam stability
and must be used at high concentrations.2 Nevertheless, cationic surfactants may be worth
considering for drilling water-sensitive shales, because of their ability to stabilize clays. 7
In general, increasing the foaming agent concentration in the liquid phase will increase the
stability of the foam. A convenient way to assess foam stability is to measure its half-life. Halflife is the time required for the foam volume to decrease to one-half of its original value. Initially,
foam half-life will increase more or less in direct proportion to the concentration of foaming
agent. With common foaming agents, once the concentration exceeds 0.5 percent, the half-life
does not increase as rapidly. Rankin et al., 1989,7 presented data showing an increase in half-life
from about 200 to 300 seconds, as foaming agent concentration was increased from 0.5 percent
to 1.5 percent.
There is a critical foaming agent concentration above which the foam half-life will tend to
decrease with further increases in concentration.9 This concentration should not be encountered
in normal drilling operations.
Contamination of foam by brine or hydrocarbon can significantly reduce its stability. For
example, 6 percent oil and 12 percent sodium chloride each reduced the half-life of one drilling
foam by about 50 percent, and the simultaneous presence of both contaminants reduced it to less
than 25 percent of its original value.7 Some foaming agents are more sensitive to brine
contamination than others. In extreme cases, fluids will not foam at all in the presence of even
modest chloride concentrations.2
Temperature also influences foam stability. As temperature increases, the rate of foam decay
increases. It is necessary to increase the foaming agent concentration, as the downhole
temperature increases. In one example, foaming agent concentration had to be increased to ten
times that required at 70F, in order to generate foam with the same half-life at 400F.2

2-91

2.5.3 Defoaming
When correctly formulated, drilling foam can have a half-life of many minutes, if not hours.
Unmanageably large volumes of foam can rapidly accumulate at the surface when it is circulated
at typical rates. In many instances, it is necessary to accelerate the decay of the foam once it
returns to the surface.
Several different chemical means are available to destroy a foam. Very often, defoamers are
themselves surfactants. The effectiveness of a defoamer usually depends on the specific foaming
agent in use. For example, iso-octanal, an oil-soluble alcohol, is effective at destroying foam
made with the commonly used anionic foaming agents. So too are various silicone emulsions.
The addition of very fine solids, that are not wetted by the foams liquid phase, can enhance a
defoamers effectiveness.1
It may be necessary to stir the foam in order to achieve effective defoaming. The initial foam
collapse that follows application of a defoamer can create a film of liquid that effectively shields
the foam beneath from the defoamer. This is more likely to occur with oil-soluble defoamers,
such as iso-octanol.
If the foam is generated with an ionic surfactant, contact with a small quantity of foam generated
with an oppositely-charged surfactant can cause rapid collapse of both foams.1
It is also possible to destroy foams mechanically. In some instances, particularly with high
quality foams, it is sufficient to agitate the foam and rupture the bubble walls. This will not be
effective on lower quality sphere foams, and may actually increase their half-life by reversing any
gravity-induced phase segregation. Drainage of the liquid phase can be accelerated and the foam
destabilized by rotating or swirling the foam so that it is subjected to radial, inertial forces.
Various mechanical defoaming systems exist. These either flow the foam through a
hydrocyclone or a tightly spiraling tube to induce rotation, or pass it through a rotating chamber.
Further details on chemical and mechanical defoamers are provided later.
2.5.4 Hole Cleaning
A good drilling foam resembles shaving cream. It would be expected that drilling foam should
be able to lift cuttings from a borehole, at even very modest annular velocities. A number of
factors interact to make hole cleaning with foam a very difficult process to model. First, foam
rheology is complex and strongly depends on foam quality. Foam viscosity is sufficiently high
that the flowing pressure drop around the well is much greater than it would be with dry gas or
mist as the drilling fluid. As shown previously, the gaseous volume fraction in a lightened
drilling fluid, and by definition, foam quality, strongly depend on pressure. There is considerable
interaction between rheology and circulating pressure. The situation is further complicated by
inflow of formation fluid. A gas inflow will increase foam quality, possibly to the extent that the
foam breaks down into a mist and loses its viscosity. A liquid inflow will reduce foam quality.
This will lead to a loss of viscosity and will increase the foam density.

2-92

Chapter 2

Underbalanced Drilling Techniques

Rheology
There have been numerous studies of foam rheology, covering wide ranges in gas and liquid flow
rates, compositions and flow geometries. Although there are differences in the results of the
various studies, some general conclusions can be drawn.
 The two factors having the greatest impact on the flow behavior of foams are quality and
flow rate.4,10
 Foam viscosity is largely independent of the foaming agents concentration in the liquid
phase, at least when using oilfield foaming agents at concentrations typical of drilling
foams.4,6,10,11
 When viscosifying agents are added to the liquid phase, the foam viscosity increases with
increasing liquid phase viscosity.12
Foam rheology is not very sensitive to other flow variables. Beyer et al., 1972,4 studied foam
flow in several different internal diameter pipes, under a wide range of pressures, temperatures,
qualities, flow velocities, and foaming agent concentrations. They found that pressure and
temperature influenced foam rheology mainly by regulating foam quality. Increasing pressure
reduces the volume occupied by the gaseous phase, reducing the foam quality. When allowance
was made for this, pressure had only a minor impact on the pressure drop characterizing foams
flowing at the same rate and quality. Increasing temperature causes the gas to expand and
increases foam quality. There is a secondary effect of temperature on viscosity. As the
temperature increased from 70F to 180F, the pressure drop at constant rate and quality
decreased by about 25 percent.4 By comparison, the viscosity of water decreases by 65 percent
over the same temperature interval.
Mitchell, 1971,6 measured foam rheology, by flowing through small diameter tubes. He found
that foam was effectively Newtonian, at qualities up to about 55 percent. This means that the
viscosity is independent of the shear rate, up to this quality level at least. In this regime, the foam
viscosity was related to its quality and to the viscosity of the liquid phase by:
F = L (1 + 3.6 )

(2.48)

where:
F ..... foam viscosity (cP),
L ..... base liquid viscosity (cP), and,
....... foam quality (fractional).
This expression has the same form as theoretical models of two-phase viscosity proposed by
Einstein, 1906,13 and Hatschek, 1910,14 although the numerical values of the coefficients are
somewhat different.

2-93

Mitchell also found that foams with qualities greater than approximately 55 percent were nonNewtonian, at shear rates below about 20,000 s-1. He represented their rheology, in this regime,
using a Bingham plastic model, frequently used to describe the flow of drilling muds:
= y +

PV
, 2 104 s1
479

(2.49)

where:
....... shear stress (lbf/ft2),
y ..... yield point (lbf/ft2),
PV ..... plastic viscosity (cP), and,
....... shear rate (s-1).
Mitchell determined values for plastic viscosity and yield point, as functions of foam quality.
These are shown in Figure 2-27
2-27. Both plastic viscosity and yield point increased with increasing
quality, up to 96 percent; the highest quality studied. Plastic viscosity ranged from less than 4 cP
at 60 percent quality to more than 15 cP at ~ 96 percent quality. Yield point increased from less
than 0.1 lbf/ft2 to more than 2 lbf/ft2 over the same range. Bearing in mind that drilling mud
yield points are usually reported in units of lbf/100 ft2, it is clear that a high quality foam can
have a resistance to flow that is much greater than many liquid drilling muds.

2-94

20

2.5
Viscosity

18

Yield Stress
Mist Region

14
12

1.5

Foam Region

10

1
Bubble
Deformation

Dispersed Bubbles

4
2

Foam Yield Stress (psf)

Bubble
Interference

Foam Viscosity (cP)

16

0.5

0
0

0.2

0.4

0.6

0.8

Foam Quality (fractional)

Figure 2-27. Plastic viscosity and yield point of foam as functions of foam quality (after
Mitchell, 1971 6)

Beyer et al., 1972,4 also concluded that foam could be described as a Bingham plastic. They
noted that there was some slippage between the flowing foam and the walls of the pipes used in
their experiments. This complicated the determination of plastic viscosity and yield point from
the measured flowing pressures. Making allowance for this slippage, they found the yield point
to be independent of foam quality, in the range from 75 percent to 98 percent, with a value of 0.1
lb/ft2. They reported higher plastic viscosities than Mitchells, reflecting the lower yield point
value and slippage. Their plastic viscosity ranged from close to 40 cP at 75 percent quality to
over 100 cP at 97 percent quality.
Several authors15,16,17 have concluded that a power law model is a better representation of the
flow behavior of foams than a Bingham model. For a power law model:
= K n

(2.50)

where:
K ........ consistency index (lbfsn/ft2), and,
n......... flow behavior index.

2-95

Chapter 2

Underbalanced Drilling Techniques

Sanghani and Ikoku, 1983,17 reported values for K and n, determined from experiments
performed using an annular geometry. These data covered shear rates from 100 to 1,000 s-1.
They reported effective viscosities that ranged from 60 to 500 cP. Again, viscosity increased
with increasing quality, up to about 94 percent. Thereafter, viscosity decreased with increasing
quality, reflecting the breakdown of foam to mist. However, even at 97.5 percent quality,
viscosity was still comparable to an 80 quality foam. Above this, viscosity dropped very rapidly,
implying that cuttings transport efficiency would be poor if foam quality were allowed to rise to
more than 97.5 percent. These data were interpreted by Okpobiri and Ikoku, 1986,3 to indicate
that foam flow would be laminar for Reynolds numbers of up to at least 3,000. This would imply
that most foam drilling operations could be performed in laminar flow.
Effective viscosity (shear stress divided by shear rate) is one means of comparing the results of
these different foam rheology studies. Figure 2-28 shows the effective viscosities of foam, as
reported by Mitchell, 1971,6 Beyer et al., 1972,4 and Sanghani and Ikoku, 1983,17 at several
different shear rates and foam qualities. Despite differences in reported values, it is clear that
drilling foam has a high effective viscosity under all conditions, and that this viscosity will
increase as the foam quality increases, at least up to 90 percent quality.
Estimating Circulating Pressure
Flowing pressure drop will be strongly influenced by viscosity and quality. Both viscosity and
quality change, as the prevailing pressure changes. Changing the relative volumes of gas and
liquid injected into the well is not the only way to vary foam quality downhole. Increasing the
annular pressure at the surface, by choking back the return flow, will reduce the extent to which
foam quality increases between the hole bottom and the wellhead.
This is illustrated by a simple, and very approximate, example. Consider a foam circulating
round a moderately deep well. Foam quality at the surface is 95 percent (a gas volume fraction of
0.95) with the return line open to atmosphere (14.7 psia). If the bottomhole pressure is 1,470
psia, Equation (2.39) indicates that the gas volume fraction, FgP, at the hole bottom is:
FgP =

Fgo
P
Fgo + 1 Fgo
Po

0.95

0.95 + 0.05 (1470 14.7)

(2.51)
= 0.16

That is, the foam quality would be only 16 percent. The large increase in pressure between the
surface and the bottom of the hole compresses the gas to one-hundredth of its surface volume,
and the gas volume fraction in the foam decreases correspondingly.

2-96

Now, suppose that the surface pressure is increased to be 161.7 psia, by choking back the flow.
At the same time, the air and liquid injection rates are changed so that the foam quality at this
elevated surface pressure is still 95 percent. Ignoring the considerable changes in annular
pressure drop associated with this backpressure, assume that the bottomhole pressure would
increase by the amount of the imposed backpressure, so that it becomes (1,470 + 147) = 1,617
psia. In this instance, the gas volume fraction at the hole bottom would be:
0.95
= 0.66 (2.52)
0.95 + 0.05 (1617 161.7)

FgP =

80 Quality Foam
1000

Effective Viscosity (cP)

Mitchell, 1971
Sanghani and Ikoku, 1983
Beyer et al., 1972

100

10

1
1

10

100

1000

10000

-1

Shear Rate (s )

90 Quality Foam
10000

Effective Viscosity (cP)

Mitchell, 1971
Sanghani and Ikoku, 1983
Beyer et al., 1972

1000

100

10

1
1

10

100

1000
-1

Shear Rate (s )

2-97

10000

95 Quality Foam
10000

Effective Viscosity (cP)

Mitchell, 1971
Sanghani and Ikoku, 1983
Beyer et al., 1972

1000

100

10

1
1

10

100

1000

10000

-1

Shear Rate (s )

Figure 2-28.

Effective viscosity of foam as a function of shear rate and foam quality (after Mitchell,
1971,6 Beyer, et al., 1972,4 and Sanghani and Ikoku, 1983 3).

The foam quality at the hole bottom would be 66 percent. In reality, the new bottomhole
pressure will not simply be the sum of the original bottomhole pressure and the additional
backpressure. Changes in foam density and viscosity combine to make the computation of the
new bottomhole pressure complex, but the trend is the same. Applying a surface backpressure to
circulating foam will reduce the change in foam quality between the hole bottom and the surface.
As noted above, any change in quality will change the effective viscosity, which influences the
flowing pressure drop, which will further influence the quality. The situation is even more
complicated because the foam density changes in response to changing pressure, and the
hydrostatic head will depend on the coupling of viscosity and quality. Consequently, it is not
possible to relate circulating pressures explicitly to flow parameters. Instead, numerical methods
are required to predict foam circulating pressures.
Numerical Pressure Predictions
Numerical models are available to predict circulating pressures and rates, when foam is used.
Recent improvements in computing power allow representative calculations to be done on a PC
(for example, Liu and Medley, 199618). Some of the earlier numerical approaches are described
below. These are useful for understanding the concepts of the interrelationships between
pressure, rate and quality.
Krug and Mitchell, 1972,19 approximated foams with a Buckingham-Reiner model, for laminar
flow (without slippage) of a Bingham plastic fluid through a pipe. They divided the flow path
into small increments of equal pressure drop, and assumed that the foam properties, including
quality and rheology, were constant over each increment. Foam temperatures in the pipe and the
annulus were assumed to be equal to the undisturbed formation temperature at the depth in
question. The pressure drop, down each increment of drillpipe length, incorporated the flowing

pressure drop due to friction and due to loss of hydrostatic head, are additive. The flow
increment length was expressed as (in the annulus):
p i +1 p i

Li =
i g c

(48v PV ) +
+

(D

Dp

6 yi
Dh Dp

(2.53)

where:
Li ........ the length of the ith increment in
the circulation path (ft),
pi ........ pressure at i (lbf/ft2),
pi+1 .... pressure at i + 1 (lbf/ft2),
i ........ flowing density at i (lbm/ft3),
gc ....... 32.2 lbmft/lbfs2,
vi ........ average flow velocity (ft/s),
PVi ..... plastic viscosity (lbf-s/ft2),
Dh....... hole diameter (ft), and,
Dp....... pipe/collars diameter (ft).
Similar formulations were used for flow through the drillpipe. A finite difference scheme was
used, taking each interval as having a pressure drop of 5 psi and iterating until the sum of the
interval lengths was equal to the length of the circulating system. The foam density, quality and
flowing velocity, in each interval, were adjusted to allow for cuttings and it was assumed that
there was no slippage of the cuttings relative to the foam. Again, it was necessary to iterate until
the sum of the flow increment lengths equaled the circulation length.
Charts were developed that showed the air and liquid volumetric injection rates and injection
pressure predicted to give a bottomhole annular velocity of 90 ft/min, with an annular surface
pressure of 14.7 psia, and an annular surface foam quality of 96 percent. These conditions were
taken to represent practical limits for adequate hole cleaning, with a minimized circulating power
requirement.
Beyer et al., 1972,4 developed a similar finite difference model of foam circulation. This model
made explicit allowance for foam slippage and for possible eccentricity of the drillstring in the
annulus. It did not allow for cuttings in the annulus or their impact on foam quality. The
rheological data assumed in the model only covered foams with qualities from 75 to 98 percent.
For lower foam qualities, the frictional pressure loss was found by interpolating between the
pressure losses computed for 75 quality foam and for water at the same total velocity. For
qualities above 98 percent, when foam would be expected to break down to mist, the frictional
pressure loss was found by interpolation between the pressure losses computed for 98 quality
foam and dry air, at the same total velocity. This model was tested against data collected while

To convert from cP to lb-s/ft2, multiply by 4.788 x 104.

2-99

Chapter 2

Underbalanced Drilling Techniques

circulating pre-formed foam around two test wells. In one of these wells, foam was circulated
down a one-inch diameter tubing string and back up the annular spacing between this and twoinch diameter production tubing. In the second well, circulation was down 2 7/8-inch diameter
tubing and back up the annulus (2 7/8-inch tubing 9 5/8-inch casing). In both cases, the total
depth was approximately 3,000 feet. Foam pressure was measured at the surface and at several
different depths. The model was used to predict circulating pressures, for the different
geometries and flow conditions examined. The average difference between predicted and
measured pressures was less than 10 percent.
An approximate method for predicting foam pressure losses and injection pressures, that does not
require iterative solution, was developed by Blauer et al., 1974.20 This model assumed that the
density of the gaseous phase is negligible, throughout the well, irrespective of the prevailing
pressure and temperature. Lord, 1979,21 showed that this is unreasonable. Despite this, the
model did predict injection pressures that were reasonably close to those measured.
Lord, 1979,21 developed an equation of state for foam. This related foam density to pressure,
temperature, liquid density, and the gas mass fraction, presuming real gas behavior. A
mechanical energy balance for the circulating system related the pressure at the entrance and exit
of the circulating system in terms of the equation of state. Frictional pressure losses were
estimated by assuming a constant friction factor throughout the system, taken as the average of
the friction factors computed at the inlet and the outlet. This model still requires numerical
solution, if flowing pressures are to be predicted, and the friction factor has to be carefully
selected. The model accurately predicted downhole pressures when proppant-laden foam was
pumped down a well.
These foam circulation models did not account for flow restrictions at the bit (i.e. nozzles), or for
the influence of cuttings on annular friction pressure. Okpobiri and Ikoku, 1986,3 developed a
foam circulation model that incorporated these two factors. It used rheological properties
measured by Sanghani and Ikoku, 1983,17 together with friction factors for clean and solids-laden
foam (from Okpobiri and Ikoku, 1986).22 This model included settling of the cuttings through the
foam. This model again required iterative solution. Some sample predictions of circulating
pressures and gas and liquid injection rates required for efficient hole cleaning at various depths
were presented. The predicted bottomhole pressures were fairly close to those predicted, for the
same conditions, using the Beyer et al., 1972, model.4
In practice, the additional complexity of allowing for the bit pressure drop may not be necessary.
Okpobiri and Ikoku22 showed that the bit pressure drop is small, in comparison with other
circulating pressure losses, if the bit is not fitted with nozzles. Also, additional frictional
pressure due to cuttings may be negligible, unless the foam velocity is so low that cuttings
accumulate in the well. At the highest solids concentration studied, the friction factor was
approximately 30 percent higher than for clean, solids-free foam of the same quality.22

2-100

Recently, Guo et al., 1995,23 developed a simpler method of predicting bottomhole pressures,
during foam drilling. The required bottomhole foam velocity and the bottomhole and surface
foam qualities are specified as input. The bottomhole pressure and the average foam density and
quality are approximated (neglecting friction). Effective viscosity is estimated from Sanghani and
Ikoku, 1983.17 This, in turn, is used to determine a friction factor and the resulting frictional
pressure loss. This is added to the first estimate of bottomhole pressure; the computation is
repeated until the predicted bottomhole pressure converges. Predictions made by this method
compared well with those from Beyer et al., 1972,4 for a case of circulation without drilling. In
addition to its basic simplicity, this model uses a simple enough representation of wellbore
geometry that it can be used for deviated or horizontal wells. No allowance is made for the
increase in foam density from cuttings in the annulus. As a result, this model predicts bottomhole
pressures that are lower than those predicted by Okpobiri and Ikoku,1983.22
Most of these methods for predicting foam circulating pressures do not allow for formation fluid
inflows. One of the principal reasons for drilling with foam, rather than dry gas, is to handle
large inflows. For example, a water inflow of 60 BPH (42 gpm) can occur during foam drilling.
Liquid injection rates are often less than this. This indicates that liquid inflows can significantly
change the foam quality at all points above the inflow. They will also influence foam quality
before the inflow point, because the additional liquid changes the borehole pressures. The
situation is further complicated by the interaction between borehole pressure and inflow rate as
the borehole pressure increases, the difference between borehole and formation pressure (that
drives the inflow) will decrease.
Millhone et al., 1972,24 demonstrated the impact of water inflow on bottomhole pressure. Using
Beyer et al.s model, they computed bottomhole pressures for four gas and liquid injection rate
combinations (ranging from 0 to 20 BPH) with different water inflows in a 3,000 foot vertical
well (2 7/8-inch diameter tubing inside 7-inch casing). The predictions are shown in Figure 2-29.
They showed an increase in bottomhole pressure, with increasing liquid injection rates. For this
well at least, the bottomhole pressure would decrease as the gas injection rate increased. In other
words, the extra friction pressure drop, due to the increased quality, is more than offset by the
simultaneous reduction in foam density.
2-29 the predicted bottomhole pressures also increased as the water inflow rate
In Figure 2-29,
increased. At a liquid injection rate of 10 gpm (14.3 BPH), bottomhole pressures could be
doubled by an influx of about 30 BPH. Increasing the specified water inflow rate is analogous to
increasing the area of productive formation exposed to the wellbore or increasing the formation
pressure. If these parameters remain fixed, the inflow rate will be controlled by the bottomhole
pressure. The well productivity line shows how, with one particular producing interval exposed,
the inflow rate would decrease in response to increasing bottomhole pressure (the formation was
assumed to be normally pressured). The intersections between the productivity line and the
predicted bottomhole pressure lines show the relevant rates of inflow and bottomhole pressure.
Particularly when drilling for natural gas, the probability of a large gas inflow needs to be
recognized. The volume of gas flowing into the wellbore can easily exceed that injected at the

2-102

Chapter 2

Underbalanced Drilling Techniques

surface. The additional gas will tend to reduce the liquid volume fraction, increase the quality
and reduce the foam density. Increased foam quality, providing that the quality is below the
threshold at which foam breaks down, would increase the frictional pressure loss. So too would
the increased volumetric flow rate. Both of these effects oppose the reduced foam density. If the
liquid volume fraction falls so low that the foam breaks down into a mist, the viscosity will fall
dramatically. There is a strong coupling between borehole pressure and inflow rate; the overall
impact of the gas inflow on borehole pressure is not easy to predict.
Newer generation models (Liu and Medley, 199618) solve the mechanical energy balance and
also incorporate inflows of gas, liquid and oil. The pressure profile, foam quality, density and
cuttings transport are predicted. Inclined and horizontal wells are represented.
Cuttings Transport
Since foams effective viscosity is large, it is reasonable to expect that it should be an efficient
medium for transporting cuttings. Beyer et al., 19724 measured the lifting force exerted by foam,
flowing past a 0.1875-inch diameter sphere. Figure 2-30 shows these lifting forces as functions
of foam quality; expressed as a fraction of the highest force that they observed. Lifting force
increased with increasing foam quality (decreasing liquid volume fraction) to a peak at the
transition from foam to slug and mist flow. At this point, the force exerted by foam was over ten
times greater than by water alone (liquid volume fraction of one). Much of the increase occurred
as the quality increased above 80 percent; for a 60 quality foam, the lifting force was little more
than twice that for water.
500
Foam Gas/Liquid
Rates (scfm/gpm)

Bottomhole Pressure (psi)

100/40

400

300
400/40
100/10

200
400/10

100
Well Productivity

0
0

10

15

20

25

30

35

40

45

Formation Fluid Influx (BWPH)

Figure 2-29. Predicted influence of water inflow on bottomhole pressure (after Millhone et al., 1972 24).

2-103

Relative Lifting Force

0.9
0.8
0.7
0.6

Relative Velocity 2

0.5
0.4
0.3
Relative Velocity 1

0.2
0.1
0
0

0.2

0.4

0.6

0.8

Liquid Volume Fraction

Figure 2-29.
30

Lifting forces acting on a 0.1875-inch diameter sphere for different quality foams
(after Beyer et al., 1972 4).

Rankin et al., 1989,7 used a model horizontal well to compare the transport capacities of several
different lightened drilling fluids and water. Two different, unspecified foam qualities were both
more efficient at transporting sand (when flowing at an annular velocity of 30 ft/min) than water
flowing at 120 ft/min. An annular foam velocity of 100 ft/min was adequate to lift sand along
the horizontal section, around the build and out of the wellbore. In contrast, dry air needed an
annular velocity of 2,500 to 3,000 ft/min and aerated water needed 300 to 400 ft/min.
Cuttings transport can be analyzed by comparing the drilling fluids uphole velocity, with the
settling velocity of the cuttings. The lifting force does not directly correlate with the settling
velocity (the maximum velocity that a single cutting would reach if falling freely through the
fluid). It is the velocity at which the buoyant weight of the cutting is exactly opposed by the drag
force generated by the fluid flowing past the cutting.
Abbott, 1974, measured terminal velocities of different sized spherical particles falling through
foams.25 He found that the terminal velocities through foam were significantly less than through
water, and were equivalent to those in many drilling muds. For particles, approximately 0.1
inches in diameter, terminal velocities were 10 to 20 ft/min. Surprisingly, terminal velocities
increased as the foam quality increased. Abbott calculated that the increase in a foams annular
velocity, due to the reduction in borehole pressure as the foam moved uphole, would be more
than sufficient to offset the increase in terminal velocity measured for increasing foam quality.
Cuttings transport efficiency would increase as the cuttings move uphole.

2-104

Chapter 2

Underbalanced Drilling Techniques

In dry air drilling, flow past the cutting is almost always turbulent. This is not necessarily the
case for foam. Moore, 1974,26 forecasted the terminal velocities for laminar, transitional, and
turbulent flow past a cutting. His predictions were based on Equation (2.2) and drag coefficients,
determined experimentally for limestone and shale cuttings flowing through mixtures of water
and glycerin:
in laminar flow:
f

v t = 4980 d c2 c
e
in transitional flow:
( )2 / 3
c
f

v t = 175 d c
1/ 3
( f e )
in fully turbulent flow:
f
v t = 92.6 d c c

(2.54)

where:
vt ........ terminal velocity of a cutting
(ft/min),
dc ........ the cuttings diameter (inches),
c....... the cuttings density (ppg),
f ....... the drilling fluids density (ppg),
and,
e....... the fluids effective viscosity at the
rate flowing up the annulus (cP).
A cuttings Reynolds number, NRec, can be expressed as:
N Re c =

15.47 f v t d c
e

(2.55)

Theoretically, flow past the cutting will be laminar if NRec < 1, transitional if 1 < NRec < 2000,
and turbulent if NRec > 2000.

2-105

In summary, the following generalizations are possible:


 If the flow is laminar, an increase in foam viscosity with increasing quality will dominate the
reduction in foam density, and the terminal velocity will decrease with increasing foam
quality, until the foam breaks down into mist.
 If the flow is turbulent, the terminal velocity is independent of the foams viscosity. vt will
increase with increasing foam quality, due to the reduction in density.
 For typical foam drilling conditions, Guo et al., 1995,23 presented an example showing that
flow past a one-half inch diameter cutting, in a 60 quality foam, at nearly 10,000 feet, was
transitional. The terminal velocity computed was about 60 feet per minute.
 In transitional flow, the terminal velocity is sensitive to the density difference between the
cutting and the foam, as well as the effective viscosity of the foam. This dependence on fluid
density and foams greatly reduced density (compared to drilling muds), is probably why
foam does not show as much increase in cuttings transport capacity (over water) as might be
expected from its viscosity.
Recent foam circulation models incorporate cuttings transport and can be used to predict gas and
liquid injection rates required for efficient hole cleaning. Guo et al., 1995,23 calculated the
bottomhole pressure first and from this determined the circulating foam velocity. Moores
terminal velocity model24 was then used, together with Sanghani and Ikokus rheological data,17
to compute a terminal velocity for some user-selected cuttings diameters. This was added to an
arbitrary minimum, required cuttings transport velocity, to determine the necessary foam annular
velocity. For the conditions considered, Guo et al., 1995,23 found that an annular velocity of 90
ft/min would be sufficient for good transport of 0.5-inch diameter cuttings.
Okpobiri and Ikoku, 1986,3 coupled cuttings transport with their circulating model. They
developed charts, showing the injection rates and pressures necessary for good hole cleaning; as
functions of depth, hole geometry, penetration rate, cuttings diameter and annulus backpressure
(choking back the return flow at the surface).
These models have the same practical problem as many of dry air drilling cuttings transport
models - the cuttings size needs to be known before the model can be used precisely.
Design Charts
For dry air (and for mist) Angels charts, while not necessarily precise, provide convenient
guidelines. The best foam drilling analog to Angels dry gas drilling model is probably charts
2-31 They indicate air
from Krug and Mitchell, 1972.19 These charts are reproduced in Figure 2-31.
and water volumes requirements (straight hole) for an annular bottomhole velocity of 90 ft/min,
at a surface annular pressure of 14.7 psia (i.e. no backpressure) and a surface foam quality of 96
percent. Different charts interrelate depth, penetration rate, hole size, drillpipe diameter and
predicted injection pressure.

2-106

Chapter 2

Underbalanced Drilling Techniques

Krug and Mitchell also computed bottomhole pressures that would be seen if their recommended
gas and liquid schedules were followed. These are shown in Figure 2-32. The nonlinearity is
due to the increase in average foam density with depth. As an example, when drilling 12-inch
hole with 5-inch drillpipe, at a penetration rate of 90 ft/hr, the bottomhole pressure is predicted
to be ~850 psi at a depth of 3,000 feet and ~3,000 psi at a depth of 9,000 feet. These correspond
to average densities of 5.5 to 6.4 ppg. Too much reliance should not be put on design charts.
With currently available PC-based software, they are unnecessary except for approximate
comparisons of different drilling systems.
The different models of foam circulation used to assess cuttings transport show that:
 Both air and liquid injection rates should increase with increasing drilling depth.
 There should be little need to increase air or liquid injection rates if the penetration rate
increases. However, the injection pressure will increase with an increased penetration rate. If
annular velocity has to be increased to improve hole cleaning, both air and liquid injection
rates must be increased; the injection pressure will increase as a result.
 Higher quality foams generally give lower bottomhole pressures.
 In order to maintain foam quality within an acceptable range, as well depth increases, it may
be necessary to apply a backpressure by choking back the return flow.
Field Evidence
Typical air injection rates, reported for foam drilling, are much lower than those normally used in
dry air drilling. Hutchinson and Anderson, 1972, 27 described a 20-inch diameter surface hole
that could not be cleaned with 4,000 scfm of dry air, but which was effectively cleaned with
stable foam, pre-formed from 700 scfm air and 12 gpm surfactant solution. Russell, 1993,5
reported cases where 17-inch and 26-inch surface holes were drilled to about 1,000 feet, using
only 375 to 460 scfm air to form foam of about 99 percent quality. These particular holes were
drilled at low penetration rates, and hole cleaning was apparently less than perfect; but the holes
were nevertheless drilled satisfactorily. Polymer-stiffened foam and higher injection rates
improved hole cleaning (refer to Section 2.6).
Foam has also been used to drill at significantly greater depths (Dupont, 198428). A seven-inch
hole was drilled through a depleted Gulf Coast gas reservoir, at 15,000 feet, using 350 scfm air
and 15 gpm foaming agent solution. No annular backpressure was applied. Despite high (99.4
percent) surface foam quality, no hole cleaning problems were reported. The reservoir pressure
was so low that the pay (at 15,000 feet) was drilled overbalanced, even though foam was used
and the bottomhole pressure was predicted to be less than 2,000 psi. This does not mean that
there would have been no inflows that might have modified the foam quality. Dupont also
indicated that good hole cleaning had been achieved in other foam drilling jobs, at annular
velocities as low as 50 ft/min.

2-107

These recommendations assumed that there is no formation fluid inflow.

180
Hole Size 6.75 inches
Pipe Size 3.50 inches

350

Drilling Rate = 1.5 fps

300

170
160

250

150
1.0 fps

200

140

150

Air

0.5 fps

100

130

Injection Pressure (psi)

Air Volume Rate (scfm) and Water Rate (gpm)

400

120
0 fps

Water

50

110

0
0

2000

4000

6000

8000

10000

100
12000

Depth (feet)

Hole Size 9.87 inches


Pipe Size 5.50 inches

180

Drilling Rate = 1.5 fps

440

160

340

140
1.0 fps

240

120

Air

Injection Pressure (psi)

Air Volume Rate (scfm) and Water Rate (gpm)

540

0.5 fps

Water

100

140
0 fps

40
0

2000

4000

6000

8000

10000

80
12000

Depth (feet)

31a.
Figure 2-30a.

Recommended air and liquid injection rates and predicted injection pressures for
19
foam drilling (after Krug and Mitchell, 1972 ); no inflow continued..

2-108

Chapter 2

Underbalanced Drilling Techniques

450

180

400

170

Hole Size 7.87 inches


Pipe Size 4.50 inches

350

160
Air

300

150
1.0 fps

250

140

200

130
Water

150

120

Injection Pressure (psi)

Air Volume Rate (scfm) and Water Rate (gpm)

Drilling Rate = 1.5 fps

0.5 fps

100

110
0 fps

50

100

0
0

2000

4000

6000

8000

10000

90
12000

Depth (feet)

150
Hole Size 12.5 inches
Pipe Size 5.50 inches

1050

140

900

Drilling Rate = 1.5


fps

750

130
120

Air

600

1.0 fps

450

110
100

Water

Injection Pressure (psi)

Air Volume Rate (scfm) and Water Rate (gpm)

1200

0.5 fps

300

90
0 fps

150

80

0
0

2000

4000

6000

8000

10000

70
12000

Depth (feet)

31b.
Figure 2-30b.

2-109

Recommended air and liquid injection rates and predicted injection pressures for
19
foam drilling (after Krug and Mitchell, 1972 ); no inflow continued..

180
Hole Size 9.00 inches
Pipe Size 5.50 inches

400

170

Drilling Rate
= 1.5 fps

350

160

300

150
Air

1.0 fps

250

140

200

130

150

Water

0.5 fps

100

120

Injection Pressure (psi)

Air Volume Rate (scfm) and Water Rate (gpm)

450

110

50

100
0 fps

0
0

2000

4000

6000

8000

90
12000

10000

Depth (feet)

140
Hole Size 15.0 inches
Pipe Size 5.50 inches

1600

130

1400

120
Drilling Rate = 1.5 fps
Air

1200

110
1.0 fps

1000

100
0.5 fps

800
600

90

Injection Pressure (psi)

Air Volume Rate (scfm) and Water Rate (gpm)

1800

80
0 fps

400

70

Water

200
0

2000

4000

6000

8000

10000

60
12000

Depth (feet)

31c.
Figure 2-30c.

Recommended air and liquid injection rates and predicted injection pressures for
19
foam drilling (after Krug and Mitchell, 1972 ); no inflow.

2-110

Chapter 2

Underbalanced Drilling Techniques

4500
4000

Bottomhole Pressure (psi)

Drill Rate = 1.5 fps

Hole and Pipe Sizes


6.75 - 3.50 in.
7.87 - 4.50 in.
9.00 - 5.50 in.

3500
3000

Drill Rate = 0 fps

2500
2000
1500
1000
500
0
0

2000

4000

6000

8000

10000

12000

Depth (feet)

5000
4500

Hole and Pipe Sizes


9.87 - 5.50 in.
12.5 - 5.50 in.
15.0 - 5.50 in.

Bottomhole Pressure (psi)

4000

Drill Rate = 1.5 fps

3500
3000
2500
Drill Rate = 0 fps

2000
1500
1000
500
0
0

2000

4000

6000

8000

10000

12000

Depth (feet)

Figure 2-32 Predicted bottomhole pressures during foam drilling, no inflow (after Krug
and Mitchell, 1972 19)

2-111

Foam has been used to successfully drill through permafrost. Anderson, 1971,29 reported a 17inch diameter hole, drilled to 700 feet, using 300 scfm of ~99 percent quality foam. Fraser and
Moore, 1987,30 presented guidelines for permafrost drilling with foam, based on experience from
drilling ten wells in the Canadian arctic. Twelve and one-quarter-inch surface hole was drilled
with 4-inch drillpipe. Recommended air injection rates ranged from as low as 215 scfm near
surface to 465 scfm at 2,000 feet. Suggested surface foam qualities were 98 to 99 percent. Since
there were water inflows, the annular qualities could have been lower than these. For
comparison, Krug and Mitchells charts recommended about 650 scfm air for 12-inch hole at
2,000 feet, and Angels analysis indicates rates of approximately 2,500 scfm if dry air were to be
used.
Without giving specific injection rates, Shale, 1995,31 noted that annular velocities during foam
drilling are often as low as 100 feet per minute, and that surface foam qualities typically range
from 95 to 99.7 percent. Finally, in a thorough review of lightened drilling fluid use in the U.S.
mid-continent, Sheffield and Sitzman, 1985,32 recommended that surface foam quality should be
kept between 95 and 98 percent, to avoid hole cleaning problems.

2-112

Chapter 2

Underbalanced Drilling Techniques

Summary
 Foam is often selected as the drilling fluid when water inflows are anticipated. With inflows,
the foam quality in the annulus will often be lower than that anticipated from the injected air
and liquid volumes. This almost certainly contributes to the frequent reports of good hole
cleaning and stable foam returns with pre-formed foam injected at rates that the theoretical
work would suggest should have broken down to mist. Nevertheless, it does seem that
surface foam qualities may be over 95 percent, without catastrophic degradation of hole
cleaning efficiency.
 Overall, during foam drilling, it is concluded that efficient cuttings transport will normally
occur at relatively low annular velocities (~100 ft/min), provided that the foam quality is
maintained between 60 and 98 percent.
 Convincing evidence of the hole cleaning that can be achieved with foam are reports of
granite fragments, as large as 2-inches in diameter, being lifted from a well during a foam
cleanout operation.33
 Unless a large backpressure is applied to the annulus, foam injection pressures are not very
different from those occurring in dry air or mist drilling, although air injection rates are
usually much lower. Foam drilling generally requires significantly lower compressor power,
in comparison to dry or mist drilling techniques.
Fluid Removal
Circulating foam can effectively accommodate large inflows of formation water. Any water
entering the annulus is likely to be rapidly incorporated into the structure of the flowing foam.
This is accelerated by the stirring effect of the rotating drillstring. Bentsen and Veny, 1976,34
observed that small water inflows, 10 to 20 BPH into 4-inch diameter holes, were so effectively
absorbed by 150 to 350 scfm of 99 percent or higher quality foam that the inflow could go
undetected.
With large water inflows, the air injection rate may have to be increased in order to keep the
foam quality within acceptable limits. If the quality is allowed to fall too low, hole cleaning
efficiency will be impaired. It may also be necessary to increase the concentration of foaming
agent in the injected liquid, to prevent the foam from deteriorating into slugs of air and water.
Provided that the air rate is adjusted accordingly, inflows as large as 500 BPH may be lifted from
a well with foam.31
Water inflow will normally increase both the bottomhole and the injection pressures.24,30
Increasing the air injection rate will, however, tend to offset the increase in bottomhole pressure.
In some circumstances, the amount of water that can be lifted from the well will ultimately be
limited by the available air rate and pressure capacity. The cost of foam consumables and waste
water disposal will normally intervene to make it uneconomic to continue foam drilling, long
before an inflow becomes too large to be lifted from the well.

2-113

Problems may arise if the foaming agent in use is not compatible with one or more of the
components of the formation fluids. Chlorides, for example, can cause some foamers to loose
their efficiency. Similarly, hydrocarbon inflows can destabilize some foams by interacting with
the foamer. Any liquid hydrocarbon flowing into the well will not mix with the aqueous liquid
phase of the foam and will not, therefore change the foam quality.
Any gas flowing into the borehole will increase the foam quality between the site of the inflow
and the surface. For the foam to remain stable to surface, the pumps should be able to deliver a
sufficient volume of liquid to keep the quality below 98 or 99 percent. If the foam quality rises
above this, the foam will break down into slugs of foam and gas, or even into mist. This can
jeopardize cuttings transport and wellbore stability.
For example, consider a well being drilled with 300 scfm of air and 99 quality foam at the
surface. The liquid injection rate to achieve this quality would be 3 cfm (22 gpm). If a gas inflow
of 432 Mscf/D (300 scfm), is encountered, the total gas rate flowing up the annulus (natural gas
plus injected air) will increase to 600 scfm. If the liquid injection rate is not increased, the liquid
volume fraction will fall too low and the foam will collapse. At the same time, the annular
velocity may be too low for the flowing mist and slugs of foam to provide hole cleaning. The
drillstring can rapidly become stuck under these circumstances.32 Increasing the liquid rate will
restore foam circulation and efficient hole cleaning, if there is a modest gas inflow, as in this
example. If this is done and drilling continues below the inflow, the foam quality between the bit
and the site of the inflow may become so low that hole cleaning will suffer in this region.
Ideally, a foam circulation model should be used to determine the impact of the increased liquid
rate on foam quality and hole cleaning below the inflow.
Now, suppose that a large gas inflow is encountered; for example, 5 MMscf/D. This corresponds
to approximately 3,500 scfm; the total gas rate flowing from the well would be 3,800 scfm. This
would require 38 cfm, which is 285 gpm, of liquid for the foam to remain below 99 quality at
surface. Few foam drilling units can come close to that liquid rate. In any case, there would be
very high consumable costs and too low a foam quality below the inflow. It would be effectively
impossible to continue drilling with foam throughout the borehole if such a large inflow were
encountered. However, in most circumstances, the annular velocity of the mist would be
sufficient for good hole cleaning. Drilling could possibly continue safely even though the returns
would be mist rather than foam.
Circulating Program Design
To summarize, a good starting point for designing a foam circulation program is to target at least
65 quality foam downhole, with a minimum annular velocity of 100 feet per minute, and a foam
quality of not more than 95 to 98 as the foam reaches the surface. It may be necessary to have an
annular backpressure to achieve this range of foam qualities. Particularly if there is no local
experience, it is advisable to carefully design the circulation program. Formation fluid inflows
can influence circulating foam quality to the extent that they should be considered as part of this
design exercise.

2-114

Chapter 2

Underbalanced Drilling Techniques

2.5.5 Equipment
In most respects, the same equipment is used to drill with pre-formed foam as is used in dry air
and mist drilling. The following sections summarize the additional equipment required for foam
drilling.
Compressors
The gas is usually provided by air compressors. Normally, the air rates used in foam drilling are
lower than those for dry air or mist. It may be possible to use fewer or smaller compressors. For
example, Fraser and Moore, 1987,30 used two 300 scfm reciprocating compressors, with a 1,000
psi delivery pressure capacity. Surface injection pressures, during routine foam drilling, rarely
exceed 200 psi. Exceptions may occur if an annular backpressure is applied, if jets are run in the
bit (which is not normal practice), or if a downhole motor is used. Higher surface pressures will
be necessary if there is a large liquid inflow, or if liquid has to be unloaded from the borehole
following a trip. For this reason, if low delivery pressure compressors are used, it is advisable to
have a booster that can increase the air pressure to 1,000 psi or more, if necessary. Boosters are
commonly rated to 1,500 psi.
The Gas
Air is most frequently used in foam drilling. Other gases, such as nitrogen, natural gas, carbon
dioxide, or exhaust gas, can be used instead. It is unlikely that these specialty gases will be less
expensive than compressed air, although the comparatively low gas rates used in foam drilling
can reduce the additional cost associated with these alternatives. Whatever gas is used, adequate
volume and pressure capacity are essential (Sections 2.2 and 2.3).
The Base Fluid
Liquid mixing tanks and injection pumps, similar to those used in mist drilling, are required.
Two 10 bbl mixing tanks are normally used. The liquid injection pump (often referred to as the
mist pump) is fed from one tank, while mixing fresh liquid in the other. A higher capacity pump
may be required for foam drilling. Liquid rates typically reported for foam drilling are in the
range of 10 to 20 gpm. Rates close to 100 gpm have been recommended for good hole cleaning
in deep, large diameter wells.13,19
Because of the great impact of foam quality on hole cleaning, good metering of the gas and liquid
is essential. Although monitoring the fall in liquid level in the mix tanks is usually adequate in
mist drilling, a flow meter, in the mist pump suction line, may be necessary for foam drilling.
Foam Generator
The one addition to a conventional air/mist drilling compressor system, for pre-formed foam
drilling, is the foam generator. It ensures thorough mixing of the two phases. One type of foam
generator is located where the gas and liquid flows meet. It introduces the liquid into the gas
flow through a small bore tube centered in the gas flow path, and then directs the mixture
through a venturi-type flow constriction. Other foam generators are located downstream from
where the air and liquid flows meet. These foam generators may contain baffle plates, or even
sand beds, to promote mixing.

2-115

In practice, it is not clear that a foam generator is specifically required.35 The air and liquid
mixture will invariably flow through a number of valves and experience many changes in
direction, before it enters the kelly; these alone may produce a good foam. Even if foam does not
form at the surface, it will when the mixture passes through the bit. There is evidence that foam
generated at the surface is more tolerant of contaminants, such as formation water or
hydrocarbons, than foam formed in their presence.33 Unless there are specific reasons not to do
so, it makes sense to use a foam generator.
Portable Units
As an alternative to conventional air drilling equipment, there are a number of custom-built,
portable, air foam units in the continental U.S.36 Although these units have mostly been
assembled for completion and workover operations, some have sufficient output for many foam
drilling operations. They combine air compressors, sometimes with a booster as well, a divided
mixing tank, liquid pumps, foam generators and metering systems. Volumetric output from a
single unit can range from 650 to 1,500 scfm, with pressure capacities ranging from 800 to 1,500
psi, for low pressure units, and up to 3,000 psi for a high pressure unit. The mixing tank
normally has a 60 bbl capacity, divided into two sections to allow liquid mixing to take place in
one section while pumping from the other. The liquid pump is generally capable of pumping at
least 50 gpm.
Mud Pumps
As was the case for dry air drilling, it is advisable to have mud pumps hooked up, to allow
liquids to be pumped into the well rapidly, if necessary. If there is any risk of hydrocarbon
production, at rates greater than the surface equipment can handle, or if there is uncertainty about
the maximum inflow rate that might be encountered, an adequate volume of kill weight mud
should be on location.
Drillstring
Float valves are necessary in the drillstring, just above the bit and near surface. If long intervals
are to be drilled, it may be advisable to reposition the upper string float (or install another) to
minimize the time taken to bleed down the string pressure before making a connection. Fire
stops are probably not necessary, except possibly for wells with long horizontal sections that are
to be drilled with an air foam.
Return System
A choke should be installed in the blooie line, close to the rotating head or rotary blowout
preventer, to pressurize the annulus, if necessary. Dupont, 1984,28 recommended a bladder-type
choke. If the circulating program indicates that annular backpressure may be necessary, then this
additional pressure should be considered when specifying the pressure capacity of the rotating
head or rotating blowout preventer. The section of the blooie line between the choke and the
rotating head should also have a pressure rating sufficient to support the highest backpressure
likely to be imposed.

2-116

Chapter 2

Underbalanced Drilling Techniques

In very cold conditions, there is the possibility for foam returns to freeze and plug the blooie line.
In these areas, Fraser and Moore30 recommended using an additional foam discharge line, with
both the blooie and foam discharge lines leading to the flare pit. This redundancy is not
necessary for normal operations, where there will be a line from the choke manifold to the flare
pit that could be used to continue circulation, but not drilling, if the blooie line were to plug.
Small rigs may not have enough clearance below the floor for two diverters on top of a
conventional BOP stack.
There is no fundamental reason why the blooie line should be any longer for foam than for dry air
or mist drilling. The flare pit should be far enough from the well that heat from the flare cannot
ignite any gas inadvertently released onto the rig floor. At the same time, as discussed in Section
2.1, Dry Air Drilling, the backpressure created by the flowing pressure drop down the blooie
line should not be excessive, if the primary jet is to be able to draw produced gas away from the
rig floor during trips.
For most operations, a 150-foot long blooie line has proven to be satisfactory. Four hundred feet
of separation between the wells centerline and the flare pit was used for a recent foam drilling
project in Oman.37 In this case, the blooie line was only 6 inches in diameter, and the jets were
placed close to the rig and about half way along the blooie line. The wells were, however, killed
prior to pulling off bottom; there were no concerns about jetting gas away from the rig during
trips. Both natural gas and crude oil were flowing into the wells during drilling. To eliminate
problems with the flared gas igniting crude oil in the return pit, a gas separator was installed in
the blooie line, 330 feet from the rig. Salt water was injected into the separator, presumably to act
as a defoaming agent. Two lines ran from the separator. One line, between 100 and 130 feet
long, took gas to the flare pit. The flare pit was banked on three sides to reduce heat radiation at
ground level. The other line, which was 130 feet long, took the liquids (foaming solution,
produced water and crude oil) to a lined liquid pit, approximately 130 feet by 130 feet in plan and
6.5 feet deep.
In foam drilling operations, it is more normal to discharge the returns into a large, combined flare
and reserve pit, as described in Section 2.1. The volumes of liquid discharged into the pits are
likely to be greater than those in air and mist drilling, considering that foam is often used to
handle large water inflows. Adequate arrangements must be made to handle the anticipated
volumes of liquid.
A large, open return pit is not essential for foam drilling. Effective defoaming and waste water
disposal are the keys to decreasing the size of the return pit. It is even possible to use a closed,
surface system when drilling with stable foam. The equipment used to do this is described in the
Section 2.9, Closed Flow Systems.

2-118

2.5.6

Injected Liquid

The Foaming Agent


The injected liquid includes (at a minimum) water, a foaming agent, and a corrosion inhibitor.
The foaming agent should be selected to suit the anticipated downhole conditions.
API RP4638 describes a method for testing foaming agents to be used in mist drilling. It should
be equally applicable for foam drilling:
 The standard test liquids are fresh water, fresh water with 15 percent kerosene, 10 percent
brine, and 10 percent brine with 15 percent kerosene.
 Ten grams of silica flour are added to one liter of test liquid, to simulate the presence of
cuttings.
 Foam, generated with the specific agent, is used to lift each of the four test liquids up a 10foot long, 2.5-inch diameter model wellbore.
 The quantity of test liquid collected in 10 minutes, at the top of the wellbore is taken as an
indication of the foaming agents suitability for use in saline or hydrocarbon environments.
 If possible, samples of actual formation fluids and cuttings should be substituted for the
regular test liquids and solids.
Downhole conditions, and the interaction between the foaming agent and formation fluids, will
dictate the concentration of the foaming agent in the injected liquid. Most commercial foaming
agents are used in concentrations that range from 0.5 to 2 percent. A concentration of 1 percent is
often a good starting point. The concentration of foaming agents should be adjusted to achieve
a level of foam stability that balances good hole cleaning with easy defoaming.
Other Additives
Particularly if air foam is used at depths greater than about 4,000 feet, careful evaluation and
selection of the corrosion inhibitor are vital to prevent severe corrosion of downhole equipment.
As was the case with mist drilling, KCl, or other shale hydration inhibitors, may be added.
Viscosifiers may be added to the liquid phase to create a so-called stiff foam, (refer to Section
2.6).
Disposal
The surfactants in foaming agents are almost invariably bio-degradable. Some of the
components may pose minor health or environmental hazards. Any hazard should be identified
on the Material Safety Data Sheets (MSDSs) for the material in question. These should be
considered when selecting a foaming agent, particularly if there are constraints on disposal of
waste liquid for the well in question. Some commonly used foaming agents are so
environmentally benign that they can be disposed into urban waste water drainage systems.
Contamination of the waste liquid by formation fluids or by a chemical defoamer may create
disposal problems, even if the foaming agent itself is environmentally acceptable.

2-119

Chapter 2

Underbalanced Drilling Techniques

The liquid injected into a well is not normally recycled. If the foam decays so rapidly at the
surface that chemical defoaming is not required, it may be possible to recycle the spent liquid
collected in the return pits. To do this, Anderson, 1984,33 recommended using a large, U-shaped
pit, to hold the returned foam until it collapses, returning the spent liquid from the pit to the
mixing tank, and re-conditioning it to the original specifications (foaming agent concentration,
corrosion inhibitor concentration, etc.) before reuse. When this can be done, savings of 25 to 50
percent in consumable costs are possible. It may also be possible to re-cycle injected liquid
recovered from closed surface systems.39 Some form of solids control is advisable to minimize
the concentration of formation solids in recycled liquid. This is particularly important if
downhole tools are used. These include motors which might be susceptible to erosion damage or
plugging.
Finally, ensure that an adequate water supply is available at the rig site to maintain the planned
injection rate.
Chemical Defoamers
Very large volumes of foam can rapidly accumulate at the surface, particularly if the foam quality
is not very high (for example 95 percent or less), and if there are no formation fluid inflows to
destabilize the foam. It is normal practice to take measures to accelerate foam decay. This can
be done chemically and/or mechanically.
It is possible to destroy foam simply by spraying a suitable defoamer on top of the foam in the
return pit. More effective foam destruction results if the defoamer is mixed thoroughly with the
returning foam. This can be done by injecting the defoamer into the blooie line using a ring
spray, as described in Section 2.4 Mist Drilling or using a turbolizer tube.40 When possible,
the defoamer should be diluted, before use, with an appropriate liquid (water whenever possible),
to assist in mixing with the foam.
The effectiveness of a particular defoaming agent often varies markedly from one foaming agent
to another. It is important, to match the defoamer to the foaming agent being used. Simple pilot
tests can be performed:
 Make a small batch (for example, one-half liter) of foam, by stirring a solution of the foaming
agent in a high shear mixer.
 Spray a small quantity, (perhaps 5 ml), of the candidate defoamer onto the newly formed
foam, using a hypodermic syringe.
 An effective defoamer will cause a significant portion of the foam to collapse in seconds.
If there are environmental restrictions on the disposal of the waste liquid, a silicone antifoam
should be considered. These are effective on many commercial foaming agents and are
environmentally benign.

2-120

Rates for adding defoamers will be controlled by the concentration and rate of foaming agent
injection, the quality of the foam at the surface, the extent of any contamination by formation
fluid inflows, the efficiency of the defoamer and the rate at which foam has to be destroyed.
Effective foam destruction usually requires smaller volumes of defoamer than foamer. For
silicone antifoams, which are available in different concentrations of active silicone, significantly
smaller volumes of defoamer will probably be required. Even with only 10 percent active
silicone, the volume of silicone antifoamer, required for very rapid defoaming, will likely be less
than one-tenth of the volume of foaming agent used.
Mechanical Defoamers
There are several different types of mechanical defoamers. All of these work by accelerating
gravity-induced separation of the foams liquid and gaseous phases. Hall and Roberts, 1984,40
described using an hydrocylone, in conjunction with a chemical defoamer. This combination was
sufficiently successful in destroying foam that stable foam could be used in offshore drilling
operations (because of disposal and storage restrictions). Defoamer was injected into the blooie
line and returning foam was passed through a hydrocyclone. The solid and liquid discharge
stream was passed over a shale shaker and into a surge tank, from where it was disposed of down
an adjacent water injection well. This is shown in Figure 2-33
2-33. In this instance, fines in the
waste liquid plugged the injection well and, ultimately, the waste liquid had to be disposed of at
an onshore treatment facility. If this system is used when natural gas might be encountered, the
gas emerging from the hydrocyclone should be piped to a flare.
There are two other generic types of mechanical defoaming. One, is conceptually similar to a
hydrocyclone. It uses a corkscrew-like internal flow path to subject the foam to radial inertial
forces and accelerate drainage. The other has a spinning, perforated drum inside an outer vessel.
Foam flows into the spinning drum. There, it breaks down under the centrifugal force. Liquid
passes through the holes, into the drum, and is discharged through the base of vessel, while gas is
discharged through the top. It is not clear how well this system will handle large quantities of
solids, such as cuttings entrained in the foam.
Combined Chemical and Mechanical Defoaming
It may be necessary to combine chemical and mechanical defoaming, if the discharge is into a
small tank or a closed system. MacDonald and Crombie, 1994,41 conducted yard trials with a
polyurethane hydrocyclone. This system decreased the volume of foam passing through it by
approximately 97 percent. As they noted, the remaining 3 percent would be sufficient to fill a
conventional well test separator within an unacceptably short time. Injecting defoamer into the
blooie line, before passing the foam through the hydrocyclone, increased the efficiency of the
foam destruction process to the extent that a closed surface system could be used.
When drilling with an open return system, it is worthwhile to check the quality and level of foam
in the pits. Particularly if defoamer efficiency is low or if defoamer is not being used, it is
possible for foam to be blown around the rig location or even off the site altogether. Although
this should not have permanent impact on the environment, it can cause much annoyance in
urban or semi-urban areas.
2-121

Chapter 2

Underbalanced Drilling Techniques

W ater
(80 gpm)
Foamer
(0.08 gpm)
Foam
G enerator

W ell

ABC
Valve

Blooie Line

Onshore
W ater
Treatment
Facility

Foam Killer
Injector

W ater
(3 gpm)

Foam Killer
(0.03 gpm)

Figure 2-33.

AntiFoam
Cyclone

Shale Shaker
Disposal
W ell

SurgeTank

A foam disposal system used offshore (after Hall and Roberts, 1984 40).

Recycling
Using most chemical defoamers effectively precludes any possibility of recycling the killed foam
liquid. A technique has recently been developed,42 in which foam is destroyed by changing its
pH. Foam is generated and circulated through the well in the normal manner, except that a pH
sensitive foaming agent is used. Acid, injected into the blooie line, reduces the pH and
destabilizes the foam. The treated foam passes through a gas/liquid separator. The gas is
directed to the flare pit. The cuttings-laden liquid discharge is passed over the shale shaker and
into a conventional arrangement of mud pits. Some of the liquid is drawn off the shaker tank to
re-inject into the blooie line; extra acid is added to decrease the pH, as necessary. If necessary,
the liquid can be treated with flocculant to remove clays or centrifuged to remove fines. Lime is
added to the liquid in the pits to raise its pH and condition it for foaming. The liquid can then be
re-injected into the foam generator. A 90 percent reduction in the consumables required for a
given foam rate was claimed for this system. The necessary surface equipment is shown in
Figure 2-34. There is little information currently available to gauge the effectiveness of this
system. The compatibility of the foaming agent with the make-up water, formation water and
hydrocarbons would have to be established, just as with conventional foaming operations. The
anticipated savings in consumables would then have to be weighed against the additional cost of
the system.

2-122

Blooie Line Return


To Air Pit or
Nitrogen Recycle

Separator
(Gas Buster)
pH

Cuttings

Shaker

Acid Feed

Acid
Storage

Recirculation
Line

pH

Supplemental Defoamer
Addition (If Necessary)

Alkali
Storage

Alkali Feed
Excess Foam
Overflow

Slop Pit

Figure 2-33.
34

Free Water
Return

Mud Pits

Polymer
Flocculant
Addition
(if Necessary)

Recovery Solution
Ready for Reuse
To Centrifuge
for Solids Control

A recyclable foam system (after Clearwater Inc., 1996 42).

Operating Procedures
In many instances, drilling with stable foam is similar to dry air or mist drilling. The procedural
differences are highlighted below.
Standpipe Pressure
As with dry air or mist drilling, the standpipe pressure should be monitored carefully. Mud rings
should not form when foam drilling. An increase in standpipe pressure usually indicates a
formation fluid inflow. Changes in the foam quality at the blooie line can often confirm the
influx.
Foam Quality
It is important to monitor the returning foam quality carefully. A steady discharge of a stiff, high
quality foam that collapses rapidly in the flare pit is preferred. With a properly designed
circulation program, this will occur, unless formation fluids are flowing into the well. Without
careful pre-planning, it may be necessary to adjust injection rates and foaming agent
concentration, to achieve the desired foam.

2-123

Chapter 2

Underbalanced Drilling Techniques

If the foam is wet (quality substantially less than 95 percent) at the surface, its quality downhole
may be too low for effective hole cleaning. There is no point in injecting more liquid than
necessary, incurring unnecessary expenditures. Furthermore, a wet foam takes longer to decay
and requires more defoamer than a dry (high quality) foam. If the foam seems wet at the surface
and there is no evidence of a water inflow, reduce the rate at which liquid is injected into the
well.
A water inflow can be detected from increased standpipe pressure and, if the surface system
permits, by comparing the liquid injection and production rates. If a water inflow is suspected, it
may be necessary to increase the concentration of the foaming agent.
If the gas volume fraction (the quality) has been allowed to become too high, the foam may
destabilize as the pressure falls on approaching the surface, and slugs of foam and gas will be
seen at the blooie line. In the extreme, the foam may collapse to a mist before reaching surface.
These conditions will normally arise as a result of a gas influx, or too low a liquid injection rate.
Gas Inflow
A gas inflow will increase the standpipe pressure and burn at the flare. If the rate of inflow is
much larger than the air injection rate, it can displace the foam above it. Initially, this correlates
with an increase in the rate at which foam is discharged from the blooie line. Once the foam has
been blown from the well, the returns are likely to be either slugs of foam and gas or mist, and
there will be a large flare. When a gas inflow is suspected because of slugs of foam and gas in
the return flow, stop drilling and increase the liquid injection rate until stable foam returns
are re-established. If this is not done, cuttings may not be lifted efficiently from the well and
the string can become stuck.
Liquid hydrocarbons can also be produced. These can reduce the effectiveness of some foaming
agents. If a foaming agent is not resistant to hydrocarbon contamination, it may be necessary to
increase the foamer concentration and injection rate, in order to re-establish stable foam returns.
As noted earlier, a very high rate gas inflow may give an annular velocity sufficient to effectively
lift cuttings from the well. To determine if this is the case:
 Estimate the inflow rate.
 Compare the total gas rate (inflow plus injected air) with the injection rate required for
efficient mist drilling for the relevant depth and hole size.
 If the total gas rate is sufficient for the mist to clean the hole, there is little point in increasing
the liquid rate. If the total gas rate is not sufficient, and it is also not possible to increase the
liquid injection rate to re-establish foam returns, it may still be possible to increase the air
rate to where mist drilling become feasible.
 Otherwise, it may be necessary to switch to mud.
 If drilling does continue with mist, with a gas inflow, the potential for a downhole fire should
be considered and the appropriate precautions taken.

2-124

Liquid Inflow
Slugs of liquid and gas may be seen at the blooie line. This will happen when liquid inflow is
not compatible with the foamer. The foamer concentration should be increased. It may also be
necessary to increase the foamer concentration simply because the well is getting deeper and the
foam temperature is increasing corres-pondingly.
If it is necessary to reduce the liquid injection rate, increase the concentration of corrosion
inhibitor in the injected liquid to compensate for the reduction in injection rate.
Lost Circulation
It is possible to lose circulation when drilling with foam. A temporary loss of returns, followed
by a violent surge of foam, may be observed following a break in circulation if a water inflow has
occurred and the hydrostatic head of water has to be overcome before circulation can be
established. Under some circumstances, however, the circulating pressures can be sufficient for
foam to be lost into downhole cavities or fractures. Particularly if this happens near surface,
careful consideration must be given to where the foam may be going. This becomes even more
important if hydrocarbons are flowing into the well when returns are lost. There are anecdotes of
foam finding its way through aquifers, into nearby streams and rivers, or emerging from a hillside
below the rig site. Any apparently uncontrolled release of foam could have damaging
consequences in terms of local opinion, even though its environmental impact should be small if
no hydrocarbons are released.
Hole Cleaning
The well should be monitored for characteristic signs of poor hole cleaning, such as excessive
drag or fill. Poor cuttings transport may simply indicate too low an annular velocity, if steady
returns of dry foam are occurring. If the well is deep, or if a wet, low quality foam is being
discharged at the blooie line, the foam quality downhole may be too low for good cuttings
transport. This may be improved by reducing the liquid injection rate or, if the surface foam
is dry, by increasing the air rate and applying backpressure to the annulus.
Backpressure
In some foam drilling jobs, particularly in deep wells, it may be necessary to apply a
backpressure, to maintain foam quality everywhere in the annulus. The circulating program
should specify a backpressure schedule, likely increasing with increasing depth. If backpressure
is required, the choke should be adjusted, as necessary, to follow the predetermined pressure
schedule. Whenever possible, the impact of any change in backpressure on foam quality, annular
velocity and hole cleaning should be computed and evaluated before adjusting the choke.
There is a large volume of compressible fluid in the well. The circulating pressures only respond
slowly to changes in the choke setting. Care should be taken, to avoid over-controlling the
choke. It is better practice to make small adjustments to the choke setting and allow the
pressure to stabilize before making any further adjustment.

2-125

Chapter 2

Underbalanced Drilling Techniques

If possible, do not use backpressure to reduce foam quality. It increases circulating pressures and
compressor power require-ments. If annular velocities are high enough, increasing the liquid
injection rate or even reducing the air injection rate, is more efficient than imposing a
backpressure.
Making Connections
Connections are made in the same way as they would be in mist drilling. Stop injecting liquid,
before bypassing the air flow to the primary jet. Continue jetting the blooie line with air, while
making the connection. This will draw any natural gas that might be flowing from the well away
from the rig floor.
Provided that the foam quality is not too high, most of the foam in the annulus will not collapse
while making a connection. With foams high viscosity and low settling velocities, cuttings will
tend to be held in suspension by the foam in the annulus. Although it is still good practice to
circulate the annulus clear of cuttings before making the connection, this is not as important with
foam as it is with dry air or mist drilling. However, with a poor, unstable foam, (for example, if
the quality is too high or if there is a de-stabilizing liquid inflow), the foam half-life can be less
than one minute. In this case, a significant proportion of the foam in the annulus will collapse
during the connection, and any cuttings will settle rapidly. If this is likely, the hole should be
circulated clean before making the connection.
As with dry air drilling, it is advisable to wait until returns are established, before resuming
drilling after a connection. Cuttings fill around the bottomhole assembly is less likely when
drilling with foam. Very often, foam is used because there is a significant liquid influx, and there
will normally be liquid in the hole after making a connection. This will not trap the string as a
cuttings accumulation could. Liquid accumulation could, however, prevent re-establishing
circulation, until the foam pressure below the bit overcomes the hydrostatic pressure. The
hydrostatic head of the accumulated liquid can be noticeable, even with a rapid connection.
Consider an 8-inch diameter hole, being drilled with 6-inch collars, and making 60 BWPH. If
circulation is shut off for 5 minutes, 5 bbl of water would flow into the well. This volume,
flowing into a dry hole, would fill 143 feet in the annulus, corresponding to a bottomhole
pressure increment of over 60 psi. The standpipe pressure will progres-sively increase after foam
injection is resumed, until the pressure of the liquid that accumulated in the well is overcome and
the liquid is lifted from the well. It can take several minutes, after resuming foam injection,
before foam actually begins to flow up the annulus. Until that happens, any cuttings that are
generated will not be lifted from the hole. Do not resume drilling, at least until the standpipe
pressure has begun to fall. Depending on hole conditions, it may be advisable to wait until
foam returns are established before resuming drilling.

2-126

Tripping
Over the time taken for a round trip, foam will separate into its liquid and gas constituents in all
except the shallowest of holes. This will allow all cuttings to settle to the hole bottom. It is
important to circulate the hole clear of cuttings before tripping.
Liquid from a collapsed, wet (low quality) foam can provide a significant hydrostatic pressure at
the hole bottom. For example, if the average foam quality around the well is 95 percent, the
liquid phase is filling about 5 percent of the hole volume. In a 5,000 foot deep hole, the foams
base liquid alone could fill at least 250 feet, giving a bottomhole pressure increase of over 100
psi. If the well is not making any liquid (at the current depth), displacing the hole with air before
tripping will avoid problems in re-establishing circulation.
If liquids are flowing into the well at a significant rate, a much larger volume of formation
liquids can accumulate during a trip. As an example, consider a 5,000 foot deep, 8-inch
diameter hole, making 60 BWPH. A round trip will take approximately 5 hours. During this
time, approximately 300 bbl, or 1,685 ft3, will flow into the hole. This would fill over 4,000
feet of the hole. The volume of the foams liquid phase is negligible in comparison. In this case,
there is no point in displacing the well to air before tripping.
The most efficient tripping procedure will depend on the volume of natural gas being produced.
If the well is producing gas, or even suspected of producing gas, steps have to be taken to prevent
gas from reaching the rig floor.
 After circulating the hole clean and, if appropriate, displacing the well to air, shut off the air
injection and strip the string back through the rotating head or rotating BOP, as far as
possible before pulling the rubber.
 If the well is not making more than about 2 MMscf/D, turn on the compressors and jet the
blooie line, as described in the section on tripping when drilling with dry air. This volume of
gas should be jetted away from the rig floor.
 Continue tripping with the rotating head rubber removed.
 If the well is making much more than 2 MMscf/D, close the annular BOP before pulling the
rotating head rubber and strip the BHA through the annular as far as possible.
 It may be necessary to jet the blooie line, to draw gas leaking past the BHA away from the rig
floor.
 Close the blind rams and flow the produced gas through an open choke to the flare pit while
the string is out of the hole.
 Once the blind rams are closed, the compressors can be shut down.
 Conventional air drilling surface equipment should be able to handle over 5 MMscf/D of gas

Neglecting any inflow.


Neglecting reduction in the inflow rate due to increasing hydrostatic pressure.

2-127

Chapter 2

Underbalanced Drilling Techniques

production, when tripping. In extreme cases of very high production rates, it may be
necessary to kill the well before tripping. Since this has the potential to damage a productive
formation, it should only be considered when it is felt to be unsafe to allow the well to flow
while tripping.
 Resume jetting, open the blind rams and close the choke before re-running the string.
 With significant gas production, it may be advisable to strip as much of the BHA as possible
back into the well through the annular.
 When possible, re-install the rotating head rubber and stop jetting.
 If the well is not producing liquids, continue the trip back to bottom and re- establish foam
circulation, before resuming drilling, as described above.
 If the well is making water, it will be
necessary to unload this from the hole before resuming drilling. The method adopted to do
this will influence how the string is tripped back into the hole. When drilling with foam, the
procedures for unloading water from the hole are similar to those used when drilling with dry
air.36 There are options for staging into the hole or for unloading in one stage.
Staging into the Hole
As was the case for dry air drilling, the length of stage that can be unloaded is determined by the
maximum available injection pressure. A general rule-of-thumb, reported by Scott et al., 1995,36
is to limit the stage length (feet) to no more than twice the injection pressure capacity (psi). For
example, up to 1,600 feet could be unloaded with 800 psi air injection capacity. This differs little
from the stage length estimation given in Section 2.1, for unloading with dry air.
 Run the string to the selected stage depth.
 Install the string float.
 Connect the kelly.
 Begin injecting a low (50 to 60 percent) quality foam.
 Once circulation is established, the standpipe pressure will fall and the well will begin to
unload.
 At this point, increase the foam quality to the level specified for drilling, by increasing the air
rate, and reducing the liquid injection rate, as appropriate. Circulating pressures will stabilize
once the formation liquid is removed from the stage.
 Shut down foam injection, trip the string down to the next stage depth, and continue
unloading in stages, until bottom is reached.
 Once the bit is on bottom, with the hole unloaded and steady foam circulation is reestablished, there is no need to dry the hole further before resuming drilling.

2-128

Unloading in One Stage


Alternatively, it is possible to unload the hole in one stage. This method uses the hydrostatic
head of a low quality foam in the drillstring, to reduce the surface pressure needed to displace
liquid from the annulus.
 Run to bottom, and begin injecting liquid at up to 30 gpm.
 Turn on the compressors and slowly inject air, to create a low quality foam. If the air rate is
too high, the injection pressure will rise quite rapidly. In this case, shut down the air and slug
the well with liquid, before resuming air injection at a lower rate.
 The standpipe pressure should begin to fall when liquid reaches the surface, or when foam
enters the annulus, if the borehole was liquid-filled.
 When foam returns are seen, the air and liquid injection rates should be adjusted to give the
foam quality and rate required for drilling to resume.
 Once the bit is on bottom, with the hole unloaded and steady foam circulation is reestablished, there is no need to dry the hole further before resuming drilling.
This procedure should only be used if it is clear that the wellbore can support the bottomhole
pressure and allow circulation to be re-established, without losing injected fluids into the
formation.
Mudding Up
It can be difficult to mud up a well that is producing gas at a high rate. The gas can blow the
mud out of the well as fast as it is pumped downhole. If this happens, the well can be shut-in and
flow diverted through the choke line to restrict the gas production rate. Since this will increase
the annular pressures, care must be taken not to subject the openhole section and casing shoe to a
pressure higher than they can support. If there is any concern about the openholes ability to
withstand shutting-in the well, it should be mudded up as soon as the gas production rate
approaches the level at which filling the hole with mud becomes difficult. For example,
Sheffield and Sitzman, 1985,32 advised mudding up, in the Arkoma Basin, if a gas flow rate of 3
MMscf/D is encountered.

2-129

Chapter 2

2.5.8

Underbalanced Drilling Techniques

Limitations Of Foam Drilling

Why is Foam Drilling Done?


 The principal reason for using stable foam as an underbalanced drilling fluid is its ability to
lift large quantities of formation water from the borehole, without requiring excessive air
(gas) volumes or injection pressures.
 Foam gives good hole cleaning, at lower annular velocities than other compressible drilling
fluids, reducing the potential for erosion of the borehole wall.
 It is possible to conduct foam drilling operations, using only biodegradable consumables.
This means that cuttings disposal can be much easier than it would be for a mud drilled hole,
if no liquid hydrocarbons are encountered. This is also the case for dry air (or gas) and mist
drilling.
 The control of foam returns is sometimes thought to restrict the use of foam drilling. With
proper attention to foam control, as described previously, there is no reason why this should
pose any technical limitation to foam drilling.
What are the Limitations?
There are a number of factors that do limit the applicability of foam drilling. These include
corrosion of downhole equipment, wellbore instability, downhole fires in horizontal wells, waste
water disposal, and consumable costs. Some of these limitations are also common to dry air and
mist drilling.
Corrosion
If air is used as the gas phase in a drilling foam, the corrosion rates of ferrous downhole
equipment can be high. As in mist drilling, oxygen and water in the drilling fluid allow corrosion
downhole. A combination of flowing conditions and mechanical contact between the drillstring
and borehole can remove potentially protective corrosion products and prevent significant local
depletion of oxygen by the corrosion reactions. Both effects promote further corrosion. Foam
has a higher liquid content than mist and the liquid phase is continuous, encouraging corrosion.
Any saline formation fluid flowing into the borehole will increase the electrical conductivity of
the foams liquid phase; this will accelerate corrosion. So too will salts, such as potassium
chloride, added to the injected liquid to control shale hydration.
Sour gas will further increase corrosion problems during foam drilling, (and other types of
underbalanced drilling). Corrosion will reduce the wall thickness of downhole equipment, as
steel is converted into oxides that have little strength. If hydrogen sulfide is present, there is real
potential for stress corrosion cracking, in addition to loss of bulk material. Sulfide scavengers in
the injected liquid will prevent small quantities of hydrogen sulfide from attacking downhole
equipment. It will not be economic to treat large quantities of hydrogen sulfide in this way.
Instead, downhole equipment should be specified for sour service. Furthermore, for safety
reasons, a closed system should be specified to handle the wellbore returns if hydrogen sulfide is anticipated.

2-130

Corrosion is not an insurmountable limitation to foam drilling. Adding an effective corrosion


inhibitor to the injected liquid will, in many instances, slow corrosion of downhole equipment to
an acceptably low level. The inhibitor must be compatible with the foaming and defoaming
agents in use. Its function should not be impaired by any formation fluid inflows that might
occur. As with mist drilling, it is advisable to conduct corrosion rate studies, to confirm the
effectiveness of the selected corrosion inhibitor.

Corrosion problems with foam increase with increasing depth, principally because of the
associated increases in temperature. For example, Scott et al., 1995,36 described a series of air
foam workover jobs, in which the workstring showed a distinct line where corrosion began, at a
depth of 4,000 feet. The corrosion inhibitor was ineffective at greater depths. Corrosion inhibitor
evaluations should be performed under conditions that replicate downhole temperatures and
pressures, if air foam is to be used in protracted drilling operations, at great depths.
Using an inert gas to make the foam will prevent corrosion. Except in special circumstances, this
will be uneconomical, even when allowance is made for the cost of corrosion and its inhibition.
Wellbore Instability
In naturally fractured formations, drilled overbalanced,43 wellbore erosion has been lessened by
reducing wellbore boundary shear stress. This happens because foams are efficient at cuttings
transport, at low annular velocities. Annular velocities with foam tend to be much lower than
with dry air or mist. As has been discussed, high quality foams typically used in drilling, have
high viscosities at low shear rates. It is not unreasonable that foam should have a lower tendency
to erode the borehole wall than dry air or mist.
 Mechanical Instability
When drilling with foam, borehole pressures tend to be higher than those encountered when
drilling with dry gas or mist. The difference may or may not be sufficient enough to have a
beneficial effect on wellbore stability. Consider, for example, a 5,000 foot deep, normally
pressured well (refer to the example in Section 2.4). Assume that the maximum effective
circumferential stress concentrated at the borehole wall is approximately 4,875 psi. The
bottomhole pressure would be around 25 psig with dry air and around 50 psig with mist.
Krug and Mitchells charts19 indicate that the circulating bottomhole pressure could be as
high as 1,500 psi, if this well were drilled with stable foam. Using foam could increase the
borehole pressure by as much as 60 times. This would decrease the difference between
circumferential stress and borehole pressure by 30 percent, from 4,875 psi to 3,400 psi, but
still may not provide adequate support. This will help reduce mechanically-induced
wellbore in-stability in weak rock. In many instances, mudding up is the only option when
major wellbore instability problems are encountered while drilling with foam. Stability
calculations are discussed in Chapter 4.
2-131

Chapter 2

Underbalanced Drilling Techniques

 Chemical Instability
As with mist drilling, it is possible to include shale inhibiting salts and polymers in the
liquid phase of a foam drilling fluid. If the additives are carefully selected, they can certainly
reduce chemically-induced instability problems, when water-sensitive shales are penetrated.
Foam is often used when significant formation water inflows are encountered. These
inflows will alter the composition of the foams liquid phase and may promote interaction
with water-sensitive shales, exposed further uphole. Foam drilling uses larger quantities of
liquid than mist drilling; the costs of treating the foam will be correspondingly higher. The
cost of shale stabilizing additives can become prohibitive when large water inflows occur.
Workovers
Many shallow wells have been successfully worked over by drilling a liner into place in an
unconsolidated and underpressured reservoir, using stable foam as the drilling fluid.44 Essary
and Rogers, 1976,44 reported that conventional drilling operations with foam were difficult
because, although stable when circulating, the formations would slough into the wellbore as soon
as foam circulation was stopped. Under some circumstances, the circulating pressure of foam
can provide sufficient support to prevent wellbore instability. Loss of borehole pressure when
circulation was shut down in these wells caused rapid wellbore collapse. Wellbore instability
that accompanied stopping circulation was overcome by attaching the bit to the liner and drillingin the liner. Even though circulation would be shut down for up to 30 minutes per connection,
few difficulties with liner sticking were reported.
Downhole Fires
Downhole fires, when drilling with dry air, were discussed in Section 2.1. The continuous liquid
phase in a foam provides a barrier to combustion, even if the gaseous phase is air. There are
many reported instances of wells drilled and worked over using air foam, without any fire safety
problems.36,45 Regardless, proceed intelligently.
The reported instances of downhole fires, when drilling with foam, are in horizontal wells.37
With the low annular velocities typically used in foam drilling, gravity-induced separation may
occur in a long, horizontal section. It is suspected that the reported fires occurred when the foam
separated, forming an air-continuous phase, which could support combustion on the top side of
the hole. The potential for downhole fires when drilling with an air foam should be
considered when planning a well, particularly a horizontal well.
Foam made with inert gases will provide more complete protection against downhole fires,
irrespective of the borehole trajectory. In foam drilling, the cost of using any gas other than air is
often prohibitive.
Waste Water Disposal
Stable foam can lift enormous quantities of water from a well, and foam drilling can safely
continue, even in the presence of very large inflows. The economic benefit of continuing to drill
underbalanced with foam must be balanced against the cost of handling and disposing of the

2-132

waste water. Because larger quantities of water will probably be produced during foam drilling
than during dry air or mist drilling, the cost of waste water disposal becomes more of a
limitation. The options for water disposal should be carefully considered, taking into account
formation water composition and liquid additives.
In some locations, it may not be possible to build a reserve pit large enough to hold all of the
waste water that may be produced when drilling a particular interval. If it is not possible to
dispose of the water while drilling, it may be necessary to mud up, once the pit level reaches a
critical limit.

Consumable Cost
A final limiting factor in foam drilling is the cost of the consumables. If saline or liquid
hydrocarbon inflows are experienced, it may be necessary to use large quantities of expensive
foaming agents, to maintain foam stability in the annulus. In combination with the additional
costs of the foam drilling equipment (compressors, liquid pumps, rotating head, etc.), the cost of
consumables may become too high for foam drilling to be economically more effective than mud
drilling. In general, this is not the case and stable foam drilling can often provide a flexible and
economically beneficial method of drilling underbalanced.

2-133

Chapter 2

Underbalanced Drilling Techniques

References

2-134

1.

Sebba, F.: Foams and Biliquid Foams - Aphrons, J. Wiley and Sons, Chichester (1987)
Chapter 4.

2.

Reduced Pressure Drilling Systems, Bachman Drilling and Production Specialties, Inc.,
Oklahoma City, OK.

3.

Okpobiri, G.A. and Ikoku, C.U.: Volumetric Requirements for Foam and Mist Drilling
Operations, SPEDE (February 1986) 71-88.

4.

Beyer, A.H., Millhone, R.S. and Foote, R.W.: Flow Behavior of Foam as a Well
Circulating Fluid, paper SPE 3986 presented at the 1972 SPE Annual Fall Meeting, San
Antonio, TX (October 1972).

5.

Russell, B.A.: How Surface Hole Drilling Performance was Improved 65 Percent, paper
SPE/IADC 25766 presented at the 1993 SPE/IADC Drilling Conference, Amsterdam, The
Netherlands, February 23-25.

6.

Mitchell, B.J.: Test Data Fill Theory Gap on Using Foam as a Drilling Fluid, Oil and
Gas J. (September 1971) 96-100.

7.

Rankin, M.D., Friesenhahn, T.J. and Price, W.R.: Lightened Fluid Hydraulics and
Inclined Boreholes, paper SPE/IADC 18670 presented at the 1989 SPE/IADC Drilling
Conference, New Orleans, LA, February 28-March 3.

8.

Kouloheris, A.P.: Surfactants: Important Tools in Chemical Processing, Chemical


Engineering (October 1989) 130-136.

9.

Burcik, E.J.: J. Coll. Sci., (1950) 5, 421.

10.

Clark, N.O.: Special Report No. 6, Department of Scientific and Industrial Research, Her
Majestys Stationary Office, London (1947).

11.

Fried, A.N.: The Foam Drive Process for Increasing the Recovery of Oil, U.S. Bureau of
Mines, Report of Investigation No. 5866 (1961).

12.

Reidenbach, V.G, Harris, P.C., Lee Y.N. and Lord, D.L.: Rheological Study of Foam
Fracturing Fluids using Nitrogen and Carbon Dioxide, SPE Production Engineering
(1986) 39-41.

13.

Einstein, A.: Eine Neue Bestimmung der Molekuldimensionen, Annalen der Physik, Vol.
19, Ser 5 (1906) 289.

14.

Hatschek, E.: Die Viskositat der Dispersoide II. Die Emulsionen und Emulsoide,
Kolloide, Vol. 8 (1910) 34.

15.

Raza, S.H. and Marsden, S.S.: The Streaming Potential and the Rheology of Foam, SPEJ
(1967) 7, 4.

16.

Wendorff, C.L. and Ainley, B.R.: Massive Hydraulic Fracturing of High-Temperature


Wells with Stable Frac Foam, paper SPE 10257 presented at the 1981 SPE Annual Fall
Meeting, San Antonio, TX.

17.

Sanghani, V. and Ikoku, C.U.: Rheology of Foam and its Implications in Drilling and
Cleanout Operations, ASME paper AO-203 presented at the 1983 ASME Energy Sources
Technology Conference and Exhibition, Houston, TX, January 30-February 3.

18.

Lui, G. and Medley G.H. Jr.: Foam Computer Model Helps in Analysis of Underbalanced
Drilling, Oil and Gas J. (July 1996) 114-119.

19.

Krug, J.A. and Mitchell, B.J.: Charts Help Find Volume, Pressure Needed for Foam
Drilling, Oil and Gas J. (February 7, 1972) 61-64.

20.

Blauer, R.E., Mitchell, B.J., and Kohlhaas, C.A.: Determination of Laminar, Turbulent
and Transitional Foam Flow Friction Losses in Pipes, paper SPE 4885 presented at the
1974 Annual SPE California Regional Meeting, San Francisco, CA.

21.

Lord, D.L.: Mathematical Analysis of Dynamic and Static Foam Behavior, paper SPE
7927 presented at the 1979 SPE Symposium on Low-Permeability Gas Reservoirs, Denver,
CO.

22.

Okpobiri, G.A. and Ikoku, C.U.: Experimental Determination of Friction Factors for Mist
and Foam Drilling and Well Cleanout Operations, J. Energy Resources Tech. (December
1983) 542-553; Trans., ASME, 105.

23.

Guo, B., Miska, S. and Hareland, G: A Simple Approach to Determination of Bottomhole


Pressure in Directional Foam Drilling, presented at the 1995 ASME Energy and
Environmental Expo 95, Houston, TX, January 29 -February 1.

24.

Millhone, R.S., Haskin, C.A. and Beyer, A.H.: Factors Affecting Foam Circulation in Oil
Wells, paper SPE 4001 presented at the 1972 SPE Annual Fall Meeting, San Antonio,
TX.

25.

Abbott, W.K.: An Analysis of Slip Velocities of Spherical Particles in Foam Drilling


Fluid, MSc. thesis, Colorado School of Mines, Golden, CO (1974).

2-135

Chapter 2

2-136

Underbalanced Drilling Techniques

26.

Moore, P.L.: Drilling Practices Manual, The Petroleum Publishing Co., Tulsa, OK (1974),
Chapter 8.

27.

Hutchinson, S.O. and Anderson, G.W.: Preformed Stable Foam Aids Workover,
Drilling, Oil and Gas J. (May 15, 1972) 74-79.

28.

Dupont, J.: Foam Used to Drill, Gravel-Pack Deep Gas Well, Oil and Gas J. (May 7,
1984) 192-194.

29.

Anderson, G.W.: Near-Gauge Holes Through Permafrost, Oil and Gas J. (September 20,
1971) 132-142.

30.

Fraser, I.M. and Moore, R.H.: Guidelines for Stable Foam Drilling through Permafrost,
paper SPE/IADC 16055 presented at the 1987 SPE/IADC Drilling Conference, New
Orleans, LA.

31.

Shale, L.: Underbalanced Drilling Equipment and Techniques, presented at the 1995
ASME Energy Technology Conference, Houston, TX, January 30 - February 1.

32.

Sheffield, J.S. and Sitzman, J.J.: Air Drilling Practices in the Mid-Continent and Rocky
Mountain Areas, paper SPE/IADC 13490 presented at the 1985 SPE/IADC Drilling
Conference, New Orleans, LA.

33.

Anderson, G.W.: Use of Preformed Stable Foam in Low Pressure Reservoir Wells, paper
SPE 12425 presented at the 1984 Offshore South East Asia Conference, Singapore.

34.

Bentsen, N.W. and Veny, J.N.: Preformed Stable Foam Performance in Drilling and
Evaluating Shallow Gas Wells in Alberta, paper SPE 5712, J. Pet. Tech. (October 1976)
1237-1240.

35.

Cobbett, J.S.: Application of an Air-Drilling Package in Oman, paper SPE 9600


presented at the 1981 SPE Middle East Oil Technical Conference, Manama, Bahrain.

36.

Scott, S.L., Wu, Y. and Bridges, T.J.: Air Foam Improves Efficiency of Completion and
Workover Operations in Low-Pressure Gas Wells, SPEDC (December 1995) 219-225.

37.

Kitsios, E., Kamphuis, H., Quaresma, V., Rovig, J.W. and. Reynolds, E.: Underbalanced
Drilling Through Oil Production Zones with Stable Foam in Oman, paper IADC/SPE
27525 presented at the 1994 IADC/SPE Drilling Conference, Dallas, TX.

38.

API Recommended Practice for Testing Foam Agents for Mist Drilling, API RP-46,
American Petroleum Institute, Washington, D.C. (1966).

39.

Rovig, J.: The Evolution of Stable Foam as an Underbalanced Drilling Medium, paper
presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

40.

Hall, D.L. and Roberts, R.D.: Offshore Drilling with Preformed Stable Foam, paper SPE
12794 presented at the 1984 SPE California Regional Meeting, Long Beach, CA.

41.

MacDonald, R.R. and Crombie, D.L.: Balanced Drilling with Coiled Tubing, paper
IADC/SPE 27435 presented at the 1994 IADC/SPE Drilling Conference, Dallas, TX.

42.

Clearwater, Inc., Underbalanced Drilling Fluids, (Air, Mist, Foam and Mud), Product
Information, Pittsburgh, PA (1996).

43.

Santarelli, F.J., Dardeau, C. and Zurdo, C.: Drilling Through Highly Fractured
Formations: A Problem, A Model and A Cure, paper SPE 24592 presented at the 1992
SPE Annual Technical Conference and Exhibition, Washington, D.C.

44.

Essary, R.L. and Rogers, E.E.: Techniques and Results of Foam Redrilling Operations San Joaquin Valley, California, paper SPE 5715 presented at the 1976 SPE Symposium
on Formation Damage Control, Houston, TX.

45.

Hutchinson, S.O.: Stable Foam Lowers Production, Drilling and Remedial Costs, 17th
Annual Southwestern Petroleum Short Course (April 1970).

2-137

2.6

Stiff Foam Drilling

The previous section described drilling with pre-formed or stable foam. The liquid phase in a
stable foam contains surfactants, and possibly salts and corrosion inhibitors, none of which has a
significant impact on the viscosity of the liquid. It is also possible to make what is
conventionally termed a stiff foam, by using viscosified water. This results in a more viscous
and stable foam than can be produced from surfactants alone.
The use of stiff foam as a drilling fluid was pioneered by the US Atomic Energy Commission,
who used it to drill a series of 64-inch diameter test holes.1 In this application, the principal
benefit claimed over conventional stable foam was improved wellbore stability in unconsolidated
formations. Subsequent oilfield applications have established that stiff foam can be used at
lower annular velocities and higher qualities (lower liquid volume fractions) than stable foam. In
some circumstances, the associated reductions in compressor power requirements and water
consumption can make stiff foam economically attractive, in comparison to other lightened
drilling fluids.2,3 Bottomhole pressures may also be lower than for a stable foam. This may be
beneficial; for example, during percussion drilling.3
In practice, stiff foam drilling involves similar equipment and procedures to those used when
drilling with stable foam. The properties of stiff foams do, however, differ from those of
unviscosified stable foam.
2.6.1

Stiff Foams

What is a stiff foam?


Consider a stiff foam to be one with an intentionally viscosified liquid phase. In the field, stiff
foams are usually made by adding a foaming agent to a fairly thin, unweighted drilling mud and
using this mixture in place of the foamer solution used in stable foam drilling.
Essentially the same structures are seen in stiff and non-viscosified (stable) foams. In stiff
foams, the increased viscosity of the liquid phase seems to provide a more stable structure,
perhaps by slowing gravity drainage and rupture of the cell walls. This allows a foam to be
created with a higher quality (lower liquid volume fraction) than would be possible without the
viscosifiers. As might be expected, the composition of the liquid phase influences the extent to
which the liquid volume fraction can be reduced before the foam structure collapses. For
example, Russell, 1993,3 described drilling operations that used both stable and stiffened foams.
A liquid volume fraction of around 1 percent (i.e., 99 percent quality), in the injected fluid, was
required to give steady foam returns and avoid air slugging, if the liquid phase was not
viscosified. By adding 0.125 ppb of each of PAC (polyanionic cellulose) and XC (xanthan gum)
polymers to the injected liquid, steady returns were achieved, at liquid volume fractions as low as
0.35 percent. The minimum liquid volume fraction for adequate foam stability was further
reduced to 0.15 percent when 0.5 ppb CMC (carboxymethyl cellulose) was substituted for the
other polymers. These wells were drilled in the Yemeni desert, where water is scarce and
expensive. Lowering the liquid volume fraction necessary for adequate foam stability resulted in
2-138

Chapter 2

Underbalanced Drilling Techniques

worthwhile cost reductions at this location.


2.6.2 Rheology
The rheological properties of stiff foams depend on the composition of the liquid phase and the
foam quality. This is different from stable foam, where viscosity is more or less independent of
the foaming agent concentration and is controlled by quality. This is not surprising since the
concentration of clays or viscosifying polymers has a considerable impact on the viscosity of the
base of liquids, whereas low concentrations of surfactants have rather little impact on base
viscosity.
Hatschek, 1910,4 developed a theoretical model of foam viscosity, for qualities between 74 and
96 percent. The foams viscosity, F, was related to the base liquid viscosity, L , and the foam
quality (expressed as a fraction), using:
F =

L
1 0.33

(2.56)

Mitchell, 1971,5 found a better fit to his experimental data (for stable foams), if the exponent
acting on the quality was increased to 0.49:
F =

L
1 0. 49

(2.57)

An analogous expression for the viscosity of lower quality foams was presented as Equation
(2.48) in Section 2.5. In the context of stiff foams, these expressions indicate that foam
viscosity increases as the viscosity of the liquid phase increases.
Few systematic studies of the rheology of stiff foam drilling fluids have been reported. There
have, however, been studies of foam fracturing fluids that indicate, qualitatively at least, the sort
of rheological behavior to be expected in stiff foam drilling operations. Reidenbach et al., 1986,6
investigated the flow properties of nitrogen- and carbon dioxide-based fracturing foams, at
different qualities, with aqueous surfactant solutions, containing various concentrations of
hydroxypropyl guar (HPG) as the viscosifying agent. They used their experimental data to
develop a model describing the rheology in terms of its quality and the rheology of the liquid
phase. Effective viscosities computed using this model are shown in Figures 2-35a and b.
These figures show that increasing foam quality and increasing polymer concentration both
increase a foams viscosity. The viscosity of a foam made with 1.68 ppb HPG (40 lbm/1000 gal)
was typically three to four times that of the unstiffened foam made with the same foaming agent
concentration. Addition of the HPG had the greatest relative impact on the viscosity at low
qualities. This is reasonable, since bubble deformation plays a large part in controlling foam
rheology of high quality foams.

2-139

It does indicate that the assumption of direct proportionality between foam and liquid phase
viscosity implicit in Equations (2.56) and (2.57) may not be correct.
2.6.3

Hole Cleaning

Benefits of Stiff Foam


Because the effective viscosity of stiff foam is higher than that of stable (unstiffened) foam, it is
possible to use even lower annular velocities, while maintaining acceptable hole cleaning
efficiency. In foam drilling, the annular velocity and the efficiency of cuttings transport are
generally at their lowest close to the top of the BHA.

Apparent Pipe Viscosity (cP)

1000
Water
10 lbm HPG/1000 gal
20 lbm HPG/1000 gal
40 lbm HPG/1000 gal

100

10

60 Quality Stiff Foam

1
10

100

1000

10000

-1

Shear Rate (s )

2-140

Chapter 2

Underbalanced Drilling Techniques

Apparent Pipe Viscosity (cP)

1000
Water
10 lbm HPG/1000 gal
20 lbm HPG/1000 gal
40 lbm HPG/1000 gal

100

10

70 Quality Stiff Foam

1
10

100

1000

10000

-1

Shear Rate (s )

Figure 2-35a.

Effective viscosity of stiffened nitrogen-based fracturing foams, 60 and 70 quality


(after Reidenbach et al., 1986 6).

Apparent Pipe Viscosity (cP)

1000
Water
10 lbm HPG/1000 gal
20 lbm HPG/1000 gal
40 lbm HPG/1000 gal

100

10

80 Quality Stiff Foam

1
10

100

1000
-1

Shear Rate (s )

2-141

10000

Apparent Pipe Viscosity (cP)

10000
Water
10 lbm HPG/1000 gal
20 lbm HPG/1000 gal
40 lbm HPG/1000 gal

1000

100

90 Quality Stiff Foam

10
10

100

1000

10000

-1

Shear Rate (s )

Figure 2-35b.

Effective viscosity of stiffened nitrogen-based fracturing foam, 80 and 90


(after Reidenbach et al., 1986 6).

quality

Foam quality here is lower than at the surface, because of the elevated borehole pressure. Since
viscosifying the base liquid has the most dramatic effects at low foam quality, the improved
viscosity is most pronounced at the point in the well where it is needed the most.
Another benefit of the increased liquid phase viscosity is that foam structure will persist at higher
qualities (lower liquid volume fractions). Since foam viscosity drops rapidly as the foam
structure breaks down, stiff foams maintain hole cleaning efficiency at low liquid volume
fractions.
Predictive Models
There are models for calculating cuttings transport in stiff foam. The restrictions of these models
lie in adequately characterizing friction, slippage and viscosity. Rheological models, such as the
one developed by Reidenbach et al., 1986,6 have been incorporated into foam circulation models.
Despite accurate numerical protocol, field calibration of these models (as with any oilfield
models) is generally necessary.

2-142

Chapter 2

Underbalanced Drilling Techniques

There are some back-of-the-envelope methods. For example, Garavini et al., 1971,7 presented a
chart that indicated suitable air and liquid (which they termed mud) injection rates for stiff foam
drilling, in various hole sizes. This is shown in Figure 2-36. For a 17-inch hole it suggests an
air rate of 300 to 450 scfm and a liquid rate of 8 to 35 gpm. These limits correspond to surface
foam qualities of 99 to 99.6 percent.
One rule-of-thumb for designing stiff foam circulating programs is to aim for a surface foam
quality from 99 to 99.7 percent and an annular velocity from 100 to 200 feet per minute, making
no correction for the compressibility of the foam.8 Using this approach, foam quality
recommendations are similar to those from Garavini et al., but the suggested injection rates are
somewhat lower. For the same example used above (17-inch diameter hole, drilled with 5-inch
drillpipe), the annular area is 1.53 ft2. According to the rule-of-thumb, a foam rate of 150 to 300
scfm should be sufficient for adequate hole cleaning (compared to 300 to 450 scfm, predicted
using Figure 2-36).
Field Cases
Cobbett, 1981,2 described 17-inch diameter holes, drilled successfully to approximately 2,500
feet, using stiff foam. Typical air rates were 150 to 175 scfm and liquid rates were 14 to 18 gpm.
These correspond to surface qualities of slightly less than 99 percent. In this instance, the liquid
phase of the foam contained 2 ppb of hydroxyethyl cellulose (HEC) and 1.25 to 1.5 percent
foaming agent. Cobbett also reported that an annular velocity of 15 to 20 ft/min was sufficient to
keep the hole clean, during coring operations with stiff foam. In these operations, annular
backpressure was used to keep the bottomhole pressure high enough to prevent the well from
flowing; foam qualities downhole were relatively low.
Russell, 1993,3 found that instantaneous penetration rates and hole cleaning efficiency were
improved by increasing the foam injection rate to levels substantially higher than those suggested
above. When drilling a series of 17-inch diameter surface holes to depths of around 1,000 feet,
using a very low liquid fraction polymer stiffened foam, increasing the air rate from 750 to 1,600
scfm increased the penetration rate from 65 to 85 ft/hr. Some cuttings fill, as well as tight hole,
were still noted at the higher injection rate. Increasing the air rate still further, to 2,100 scfm,
eliminated any indications of hole cleaning problems and minimized water requirements in an
arid location.

2-143

Mud Injection Rate (gpm)


35

30

25

20

15

10

18

Hole Diameter (inches)

17
16

ted
es
g
g
Su

15
14

e
im
eg
R
te
Ra

13
12
11
10
9
8
50

75

100 125 150 175 200 225 250 275 300 325 350 375 400 425 450

Air Injection Rate (cfm)

Figure 2-36.

Suggested air and liquid (mud) injection rates for stiff foam drilling (after Garavini et
al., 1971 7).

Higher injection rates may be necessary to drill to great depth with stiff foam. Sheffield and
Sitzman, 1985,9 reported using 1,000 scfm air to drill a 12-inch diameter hole, to a depth of
11,725 feet, using a foam containing a polymeric viscosifier. In contrast, at 7,300 feet, 3,400
scfm of air had been required for mist drilling. Foam chemistry was not specified precisely and
the degree of viscosification is uncertain. In any case, it is clear that using a foam drilling fluid
permitted underbalanced drilling to continue to considerable depth, with only modest air
compressor power.
Field and laboratory data indicate that there can be considerable variation in the injection rates
required for adequate hole cleaning when drilling with stiff foam. This partly reflects the
variations in foam viscosity that can be achieved by varying the liquid phase composition. In
addition, different operations can tolerate different levels of cuttings buildup in the annulus.
There is an economic balance between reduced air and liquid injection rates and increased
consumables costs, as viscosifying agent concentration is increased to increase foam viscosity.
The optimum is likely to vary from well to well.

2-144

Chapter 2

Underbalanced Drilling Techniques

2.6.4 Circulating Pressures


Typically, injection pressures of around 200 psi have been reported when drilling to depths of
2,500 feet2. These are not very different from those typical of unstiffened (stable) foam. The
higher viscosity of the stiff foam is offset by the lower injection rates. Without anything to rely
on, it is probably reasonable to anticipate that circulating pressures will be similar to those
computed for stable foam drilling.
There is no fundamental reason why annular backpressure should not be used to control the
downhole quality of stiff foams. In practice, it is less likely to be necessary than with unstiffened
foams. Any increase in frictional pressure drop for stiff foams is often largely offset by the lower
annular velocities that are required. It is possible to maintain a foam structure at lower liquid
volume fractions, and, the density of stiff foams can be lower, than for unstiffened foams. This
reduces the rate of hydrostatic pressure increase with depth; this means that these foams do not
compress as rapidly as do unstiffened foams. As a result, stiff foam quality may not decrease as
rapidly with increasing depth (compared to stable foams). This reduces the need to choke back
the return flow in order to maintain foam throughout the well.
2.6.5 Equipment
Much of the same equipment is used for stiff foam as for stable foam drilling. This section will
highlight those features that are distinctive to stiffened foam.
Surface Equipment
Because of lower air injection rates, it may be possible to reduce the air compression capacity on
location. Unless a very large hole will be drilled, it may not be necessary to have more than 450
scfm injection capacity. Although many operations will not require more than 250 to 300 psi
delivery pressure, it is advisable to have a booster on site, for unloading water.
With stiff foam, liquid injection rates tend to be lower than in stable foam drilling. Even when
drilling large holes (for example, 17-inch and larger), it is unlikely that more than 35 gpm will
be required. The soap pump on an air/mist drilling rig will have adequate capacity for many stiff
foam drilling jobs.
One difference from stable foam drilling is that a separate tank, in addition to the injection pump
tanks, is used to prepare the liquid phase. A hopper system, which allows solids to be added to a
flowing liquid stream, is usually required for effective mixing of polymeric viscosifiers in water.
On a conventional drilling rig, the mud tank and mud hopper will normally be used for mixing
the injection liquid. A suitable pump is necessary to transfer the hydrated polymer to the
injection pumps suction tanks. If a custom-built air drilling rig is being considered for stiff, as
opposed to stable, foam drilling, then allowance must be made for the additional cost of these
liquid mixing facilities.

2-145

Stiff foam tends to have a longer half-life than stable foam. Even though the foam flow rate may
be lower, more attention has to be paid to foam control at the surface. If the foam is to be
allowed to collapse naturally, it may be necessary to increase the size of the reserve pits from
those shown in Figure 2-15. Pilot tests with the intended foam composition can indicate the halflife of the foam and the effectiveness of any proposed defoaming measures. If these tests are
conducted before the location is prepared, they can be used as a guide to sizing of the pits.
Foam Consumables - Viscosifiers
Originally, stiff foams were commonly made by adding bentonite to water, at a rate of 10 to 12
ppb, to create the viscosified base liquid. Often, up to 0.5 ppb of polymer, such as guar gum or
CMC, were added to augment the bentonite. Recently, the trend has been to use polymers
exclusively. Formulations that have been reported include:
 2.0 lbm/bbl HEC2, or,
 0.5 lbm/bbl PAC and 1 lbm/bbl Cypan (sodium polyacrylonitrile)8, or,
 0.125 lbm/bbl PAC and 0.125 ppb XC3, or,
 0.5 ppb CMC3.
The CMC has shown better resistance to calcium and chloride contamination than PAC and XC.
Similarly, the PAC/Cypan mix has been reported to be more resistant to minor saline and oil
contamination than bentonite-based mixes.
Where there is significant calcium in the make-up water, or if calcium contamination of the foam
by formation water is possible, it may be necessary to add soda ash to the liquid phase to prevent
the base fluid from breaking.
Increasing the concentration of viscosifiers, whether bentonite or polymers, will increase the
foam viscosity and tend to improve cuttings transport efficiency. In general, higher foam
injection rates are required for adequate hole cleaning with the low polymer content formulations
(listed above) than when higher concentrations were used. Increasing the polymer concentration
will increase the consumable cost, although this will be partly offset if the higher viscosity allows
a lower foam injection rate to be used. Further, an excessively stiff foam will be very stable and
difficult to break at the surface, potentially leading to foam control problems. The base liquid
should not be thicker than necessary to achieve the desired hole cleaning capability. Cobbett,
1981,2 recommended a Marsh funnel viscosity in the range 50 to 80 seconds; this is probably a
good starting point.

2-146

Chapter 2

Underbalanced Drilling Techniques

Foam Consumables - Foamers


There is some possibility for confusion between the influence of the foaming agent and the
viscosifiers on foam stability. Constien, 1989,10 noted that the polymer concentration in
fracturing foams contributed less to foam stability than did the type and concentration of the
foaming agent. The same foaming agents are used for stiff foam as for stable foam drilling.
Typical foaming agent concentrations will be around 1 percent. Because the foaming agent plays
the principal role in foam generation, there has to be an adequate concentration of foaming agent
for a stable foam to form. Contamination by saline formation water or liquid hydrocarbons, as
well as elevated temperatures, will reduce a foaming agents effectiveness and de-stabilize a stiff
foam, in exactly the same way as they do an unstiffened foam. Higher foaming agent
concentrations will be required under these circumstances.
Foam Consumables - Corrosion Inhibitors
Corrosion of downhole equipment is a concern when drilling with stiff foam. Suitable corrosion
inhibitors, which must be compatible with the other components of the foam and with any
potential contaminants, should be added to the liquid phase. In some instances, caustic soda is
added to increase the pH to between 9 and 10; this will tend to retard corrosion.2
The optimum foam composition will probably vary from well to well, and may have to be
determined by on-site experimentation. If at all possible, pilot tests should be conducted to
confirm the foams stability in the presence of anticipated contaminants, regardless of the foam
composition.
Foam Control
Because of its stability, stiff foam can be resistant to defoaming measures. The high viscosity of
the liquid phase makes mechanical defoaming less effective than it is with unstiffened foams.
Chemical defoamers are effective, but the high viscosity of stiff foam can slow their dispersion
through the foam. For this reason, thorough mixing of the defoamer with the foam is required.
Where location constraints demand that defoaming measures be taken, the consumption of
defoamer is likely to be higher than for an unstiffened foam.

2-147

2.6.6 Operating Procedures


Operating procedures are virtually identical to those for stable foam drilling. The only significant
differences relate to mixing the injected liquid and controlling foam composition and quality.
Mixing the Injected Liquid
Base liquid is prepared in the mixing tank, rather than in the injection pumps suction tanks.
Many polymeric viscosifiers can only be added slowly, if they are to disperse, mix and hydrate
satisfactorily in the liquid. The polymer should be trickled into water flowing through the rigs
mud mixing hopper. The liquid should be circulated through the hopper, to allow viscosity to
build. If the liquid phase contains bentonite, this should be allowed to hydrate, preferably
overnight, before being used.
Large quantities of liquid may have to be mixed. A liquid rate of 25 gpm corresponds to over
850 BPD of continuous foam injection. Because the mixing operation can take some hours, it is
necessary to start mixing a fresh batch of base liquid well before the previous batch is exhausted.
It may be advisable to transfer each batch of base liquid, once mixed, to another tank so that the
next batch can be prepared.
Base liquid is pumped to the injection pumps suction tank that is not in use. The foaming agent
is added to the suction tank to avoid any possibility of foam formation in the main mud tanks.
Inflows
In addition to a stiff foams quality, the concentration of viscosifying agent also influences its
viscosity and stability. This provides another parameter that can be manipulated if there are
problems with hole cleaning or foam stability.
 Stiff foam returns should have a consistency similar to shaving cream. Provided that the
foamer concentration is sufficient and that the surface liquid volume fraction is not more than
99 percent, steady foam returns should be observed. If slugs of liquid and air are seen,
either the foamer concentration is insufficient or it is incompatible with a formation fluid
inflow. In either case the response should be to increase the foamer concentration in the
injected liquid.
 If hole cleaning problems are experienced with steady foam returns, the options are to
increase foam injection rate, to increase the concentration of viscosifiers in the injected
liquid, or some combination of both. The economics of each specific operation will dictate
which of these options is likely to be most cost effective.
 Water inflow can reduce the hole cleaning efficiency of stiff foams. It will reduce both the
foam quality and the concentration of viscosifier in the liquid phase, possibly causing the
viscosity downhole to fall considerably. If a water inflow occurs and hole cleaning
problems are suspected, the air rate should be increased. It may also be necessary to
increase the concentration of foaming agent and viscosifiers in the injected liquid. The
liquid injection rate should not be reduced. Reducing the liquid injection rate would lower
the concentrations of foamer and viscosifiers in the foam above the inflow, which would
further reduce cuttings transport efficiency.

2-148

Chapter 2

Underbalanced Drilling Techniques

 Because stiff foams tend to be used at lower annular velocities and higher qualities than
stable foams, they are more vulnerable to gas inflows. A modest gas inflow can increase the
quality to the point that the foam breaks down and cuttings transport efficiency is lost. If this
happens sufficiently rapidly, the string could become stuck. Slugs of foam and gas will be
seen at the surface and the gas will burn at the flare. If a gas inflow is suspected, stop
drilling and increase the liquid rate to re-establish foam returns if at all possible.
Increasing the foamer concentration may help if liquid hydrocarbons are entering the well
along with gas. If the inflow is too rapid, it may not be possible to achieve steady foam
returns. Then, if possible, increase the air rate to a level sufficient to permit mist drilling. If
there is insufficient compressor power to do this, the well must be mudded up.
2.6.7 Limitations
Stiff foam can effectively lift cuttings, at very low annular velocities. This makes it particularly
suitable for drilling large diameter holes, where the gas injection rates required for other
lightened drilling fluids may not be economically feasible. Otherwise, most of the factors that
limit stable foam drilling also limit drilling with stiffened foams.
Wellbore Instability
As was noted earlier in this section, stiff foams can give better wellbore stability through poorly
consolidated formations than other lightened drilling fluids, including stable foams.8 It has been
claimed that bentonite in a foam can create a stabilizing cake on the borehole wall.1 It is not
clear how any cake could deposit if the well were truly underbalanced. The Atomic Energy
Commissions large diameter holes were located in desert areas, and it is plausible that they were
above the water table. In this case, they would have been overbalanced and there would have
been a differential pressure tending to push foam into the formation. The very low annular
velocities that can be used when drilling with stiff foam reduce the potential for erosion of the
borehole wall below that of unstiffened foam and other lightened fluids. This probably
contributes to wellbore stability in unconsolidated formations. Reduced fluid loss and less
change in capillary pressure may also afford stability.
Gas Inflows
Gas inflows can pose a problem for stiffened foams. Since stiff foams tend to be used at higher
qualities, there is more chance for the foam structure to collapse downhole. At the same time, air
rates tend to be lower than for unstiffened foam, and there is a greater probability that the annular
velocity of the collapsed foam will be too low for efficient cuttings transport.
This does not mean that stiff foam should not be used to drill gas wells. In many circumstances,
it should be possible to increase the liquid injection rate and maintain a foam until the annular
velocity is sufficient to lift cuttings, even after collapse. Difficulties may arise if drilling is to
continue beneath a gas inflow. Then, the foam quality may be too low between the bit and the
inflow site, if the liquid injection rate is to be sufficient to keep a foam stable to the surface, with
the additional gas flowing up the annulus.

2-149

Downhole Fires
Stiff foams tend to be more stable than unstiffened foams. For this reason, they should be more
resistant to the gravity separation that can lead to downhole combustion in long, horizontal
sections. The lower gas phase rates required with a stiff foam may also make it less costly to
make the foam with an inert gas, such as nitrogen, than would be the case if the foam were
unstiffened. This would completely remove the risk of downhole fires and would also greatly
retard downhole corrosion.
Corrosion
Stiffened foams are characterized by the same combination of an ample oxygen supply and a
continuous aqueous phase as are unstiffened foams. This offers the same potential for rapid
corrosion of downhole equipment. While this should be controllable in most applications, by
correct selection and use of corrosion inhibitor, it does add to the cost of all foam drilling
operations.
Consumable Costs
Viscosifying agents in the injected liquid add considerably to the cost of the operation. In those
locations where water is expensive, the low water volumes required for stiff foam drilling can
sometimes offset the high product costs. So too can any reduction in compressor power
requirements. Nevertheless, there are applications where high consumable costs make stiff foam
economically unattractive, compared to other lightened drilling fluids.
Waste Water Disposal
Disposal of waste water may also limit the use of stiff foam. Some viscosifying agents have
restrictions placed on their disposal. The increased viscosity of the killed foam could also limit
the possibility of re-injecting it in disposal wells.
Formation Damage
Finally, the potential for formation damage should be considered. Provided that the well is kept
underbalanced through the producing zone, liquid from the foam will not mechanically enter the
formation. Should overbalanced conditions be inadvertently created, the liquid phase of the foam
can enter the pore structure of exposed productive zones. The clay or polymers in the base liquid
would reduce fluid loss, but could impair permeability much more than a simple surfactant
solution.

2-150

Chapter 2

Underbalanced Drilling Techniques

References

2-151

1.

Shale, L.T.: Underbalanced Drilling: Formation Damage Control During High-Angle or


Horizontal Drilling, paper SPE 27351 presented at the 1994 SPE International
Symposium on Formation Damage Control, Lafayette, LA.

2.

Cobbett, J.S.: Application of an Air-Drilling Package in Oman, paper SPE 9600


presented at the 1981 SPE Middle East Oil Technical Conference, Manama, Bahrain.

3.

Russell, B.A.: How Surface Hole Drilling Performance was Improved 65 Percent, paper
SPE/IADC 25766 presented at the 1993 SPE/IADC Drilling Conference, Amsterdam, The
Netherlands, February 23-25.

4.

Hatschek, E.: Die Viskositat der Dispersoide II. Die Emulsionen und Emulsoide,
Kolloide, Vol. 8 (1910) 34.

5.

Mitchell, B.J.: Test Data Fill Theory Gap on Using Foam as a Drilling Fluid, Oil and
Gas J. (September 6, 1971) 96-100.

6.

Reidenbach, V.G, Harris, P.C., Lee, Y.N. and Lord, D.L.: Rheological Study of Foam
Fracturing Fluids Using Nitrogen and Carbon Dioxide, SPEPE (1986) 1, 39-41.

7.

Garavini, O., Radenti, G. and Sala, A.: How Foam Aids Drilling Operations on Zagros
Mountain in Iran, presented at the 1971 Weatherford International Round-Up, London.

8.

Anon: Stiff Foam, Magcobar Technical Memorandum.

9.

Sheffield, J.S. and Sitzman, J.J.: Air Drilling Practices in the Mid-Continent and Rocky
Mountain Areas, paper SPE/IADC 13490 presented at the 1985 SPE/IADC Drilling
Conference, New Orleans, LA.

10.

Constien, V.G.: Fracturing Fluid and Proppant Characterization, Reservoir Stimulation,


M.J. Economides and K.G. Nolte (eds.), Prentice Hall, Englewood Cliffs, NJ (1989).

2.7

Gasified Liquids

Foams are not the only lightened drilling fluids in which the liquid is the continuous phase. A
flowing liquids density will be reduced when a gas is mixed into it, whatever the structure of the
resultant mixture. Gasified liquids are sometimes used as underbalanced drilling fluids.
Generally, these do not contain surfactants and often have high liquid volume fractions under
downhole conditions. A bubbly flow regime, rather than a stable foam, is created in the
wellbore. Bottomhole pressures are generally higher for gasified liquids than for other lightened
drilling fluids. Effective densities of gasified liquids are usually in the range 4 to 7 ppg. Unlike
other lightened drilling fluids, the liquid phase of gasified liquids is normally cleaned and
pumped back downhole, after it returns to surface.
Although they had been previously used by the oil and gas industry, the first engineered approach
to drilling with gasified liquids was made in Emery County, Utah in the early 1950s.1
Compressed air was injected into drilling mud at the standpipe, to reduce the borehole pressure
and avoid the lost circulation that occurred with mud alone.
Originally, aerated muds constituted the majority of gasified liquid drilling fluids. Their primary
application was to avoid lost circulation, rather than specifically for drilling underbalanced.
Severe lost circulation can be an expensive and time consuming drilling problem. There are
areas where lost circulation material, cement and proprietary additives are, at best, partially
successful in regaining circulation. In these areas, gasifying the drilling fluid, to reduce its
density until the circulating pressure is equal to or less than the formation pressure, can be a costeffective way to prevent lost circulation.
Recently, gasified liquids have become the predominant underbalanced drilling fluids used in
Canada and some other regions of the world. The base liquids are normally unviscosified water,
or crude oil, gasified with nitrogen or, less frequently, with natural gas. Many of these
applications involve horizontal wells. These are particularly prone to formation damage, as the
producing interval tends to be exposed to the drilling fluid for longer periods of time than it
would be when drilling vertically. Stimulation of long horizontal intervals, to overcome
formation damage, is difficult and expensive. Drilling underbalanced often avoids formation
damage and the need for any stimulation. This can reduce the total well cost, even if the drilling
costs are increased by the additional equipment required.
Circulating pressures can be controlled by adjusting the gas and liquid injection rates. Often, the
intended underbalanced pressure differential between the wellbore and the formation fluid is
fairly small, in the range of 250 to 500 psi. These low underbalance pressures, which are typical
of drilling with a gasified liquid, give fewer wellbore instability problems and lower formation
fluid inflow rates than would be seen with other lightened drilling fluids. As a result, gasified
liquids allow less consolidated and more prolific formations to be drilled underbalanced.

Occasionally gel mud or condensate.

2-152

Chapter 2

Underbalanced Drilling Techniques

Gasified liquids have other benefits over dry gas or mist drilling. They provide a degree of
hydraulic damping of drillstring vibrations. Directional drilling, with dry gases or mist, often
involves frequent failures of steering tools and MWD systems due to high vibration levels.
Vibrational failure of directional tools is not a major problem when drilling with a gasified
liquid. Drillstring components tend to show less erosive wear with gasified liquids than they do
when drilling with dry gases or mist, reflecting the lower annular velocities used when drilling
with gasified liquids. Finally, in vertical holes at least, the potential for downhole fires is
eliminated.
2.7.1 Gasification Techniques
There are two basic techniques for gasifying the drilling fluid; injecting the gas into the liquid at
the surface before it enters the drillpipe, and injecting it downhole into the liquid in the annulus.
It is also possible to use both techniques in combination. Downhole gas injection requires a connection from the surface to the annulus. This may be provided by a parasite tubing string run
outside the last string of casing, by the annulus between a temporary casing string hung off inside
the last cemented casing string, or if deepening a previously gas-lifted producing well, by the gas
lift system. Options for gasification are illustrated schematically in Figure 2-37.

Drillstring Injection

Annulus Injection: Parasite Casing

Figure 2-37.

2-153

Annulus Injection: Parasite String

Annulus Injection: Through Completion

Gas injection methods for drilling with a gasified liquid (after Bieseman and Emeh,
1995 2).

Drillpipe Injection
Drillpipe injection has certain advantages.
 It does not require any additional downhole equipment, and therefore has a lower capital cost
than gasifying with either a parasite tubing string or with a temporary casing.
 Because gasified fluid fills the entire annulus, lower bottomhole pressures can be achieved
than if the gas is injected part-way up the annulus.
 For the same reason, the gas injection rate needed for any particular bottomhole pressure is
lower than with annular injection.
 There are better hydraulics at the bit.
There are also drawbacks to drillpipe gas injection.
 It is impossible to continue gasification when circulation is shut down to make connections or
for tripping. As a result, it becomes difficult to maintain a specific underbalance pressure.
Procedures exist that allow underbalanced conditions to be preserved when circulation is shut
down, but these do not give good bottomhole pressure control.
 Since gas is trapped under pressure in the drillpipe by the various string floats, the drillpipe
pressure takes a finite time to bleed down when making connections and tripping. These
operations are slower than they would be with only liquid in the drillpipe.
 Flow down the drillpipe will be two-phase. This tends to give higher frictional pressure
losses than would be seen with single-phase, liquid flow. Standpipe pressures will, therefore,
be higher than they would be with annular injection. The compressible phase (the gas
bubbles), in both the annulus and the drillpipe, causes rapid attenuation of any Measurement
While Drilling (MWD) pressure pulse signals. Conventional mud pulse telemetry MWD
cannot be used. With drillpipe injection, both gas and liquid flow through any downhole
motor. This can reduce the motors efficiency. It can also cause downhole vibrations that
may shorten the operating life of mud motors and other downhole equipment.2
 Another potential problem with drillpipe gas injection can arise when gasification is resumed
with the bit opposite or below a permeable formation. If the wellbore pressure exceeds the
formation fluid pressure and fluid loss control is not good, it is possible for the injected gas to
flow into the exposed permeable formation rather than up the annulus.3
 Drillpipe gas injection exposes the entire drillstring to higher corrosion rates.
Annular Gas Injection
With annular gas injection:
 It is possible to continue gasification of the fluid in the annulus when making connections or
tripping. However, liquid circulation will be suspended and control of the bottomhole
pressure is not perfect.

2-154

Chapter 2

Underbalanced Drilling Techniques

 Flow down the drillstring is single-phase (liquid); mud pulse MWD systems can be used,
downhole motors operate efficiently, and downhole vibration levels may be lower.
 The gas injection point will normally be inside casing, and the potential for injected gas
flowing into exposed permeable formations is more or less eliminated.
 Although the gas pressure required to initiate injection may be higher than for drillstring
injection, it will normally fall to less than the standpipe pressure, once the target bottomhole
pressure is established. As a result, the gas compression power required for annular injection
may be less than for drillstring injection.
 Finally, if the injected gas is compressed air, not all of the drillstring will be exposed to the
potentially very corrosive aerated liquid.
General
The different annular gas injection techniques all require higher gas rates, and higher gas costs, to
achieve a given underbalance pressure. There are further limitations specific to each technique.
There is additional capital cost associated with injection down either parasite tubing or a
temporary casing string. The possibility for using either method in a re-entry drilling operation is
severely limited, unless existing casing or large-bore production tubing can be pulled.
There has to be sufficient clearance between the casing to which a parasite tubing string is
attached and both the casing and the openhole section inside which it is run, to accommodate the
tubing and its injection sub. Parasite tubing strings slow down running the casing to which they
are attached. Mechanical damage is a risk while they are run. This may be a particular concern
in deviated wells. The injection sub through which they port into the casing can be a weak point
in the casing string.2
If a parasite, or micro-annulus casing string is used, the hole size that can be drilled is restricted
to the drift diameter of the temporary casing. The annular volume outside of the temporary
casing will be large in comparison to the volume of a parasite tubing string. This slows down the
rate at which gas injection and bottomhole pressures respond to changes in gas injection rate, to
the extent that bottomhole pressure control can be difficult. The casing will normally be
retrieved once the target section has been drilled; the time needed for this will add to the well
cost. One benefit of the temporary casing string is that it can be safely run into deviated or even
horizontal wells, facilitating a deeper gas injection point than could be achieved with a parasite
tubing string.
It may be possible to deepen a well by drilling inside the existing completion. If there is a gaslift system in place, this can be used to lighten the drilling fluid in the annulus without installing
any additional equipment downhole. The hole size that can be drilled will in general be severely
restricted. Since the completion is also vulnerable to damage by a rotating drillstring, this
gasification technique may only be feasible when drilling with coiled tubing.4

2-155

Finally, it is possible to use more than one gasification technique at the same time. This can be an
attractive option in some circumstances, notably for maintaining a very low borehole pressure
when hole cleaning demands high annular velocities. Using both micro-annulus and drillpipe
injection can allow higher gas injection rates, at an acceptable surface pressure, than could be
achieved with either annular or drillpipe injection alone.5
2.7.2 The Liquid Phase
Originally, unweighted drilling muds were used as the liquid phase in gasified liquid drilling
fluids. The recent trend is to use unviscosified fluids, such as water, brine, diesel, crude oil or
condensate.
If possible, drilling mud should not be used as the liquid phase for gasified liquid drilling.
Formation fluid inflows are likely during underbalanced drilling. These will contaminate and
dilute the mud. Re-conditioning this mud can be expensive (removing comtamination,
countering dilution).
The base liquid should be non-damaging to any producing formation to be drilled and compatible
with formation fluids that may be encountered. Interactions between the drilling fluid liquid
phase and formation fluids are possible in the formation, in the annulus and in the surface
separation system. As long as the well is underbalanced, formation fluids will flow into the
borehole and mix with the drilling fluid. Because pressure control is not easy when drilling with
gasified liquids, it is probable that transient overbalanced conditions will occur and that drilling
fluid will periodically flow into the formation. It is also possible, particularly when drilling a
dry, gas-saturated formation with an aqueous fluid, for fluid to flow from the borehole into the
formation, even when drilling underbalanced.6
Emulsions
The formation of emulsions can be troublesome.7 Emulsions can occur when formation water
flows into a hydrocarbon-based, gasified liquid or when oil flows from the formation into a
water-based, gasified liquid. Emulsions can cause high and irregular annular pressure losses.
These not only increase the gas injection rate required to maintain the desired underbalance
pressure, but also make maintaining a steady downhole pressure very difficult. Commonly used
viscosifying and fluid loss control polymers can promote the formation of oil and water
emulsions. If emulsions are likely to form downhole, adding a de-emulsifier to the injected
liquid should be considered.7
Foaming
Foaming at the surface can also be a problem when liquid formation fluid is produced and mixes
with the injected liquid.2 Minimizing the use of viscosifying agents will restrict foaming.
Adding small quantities of a silicone anti-foam or other defoaming agent to the mud pits of an
open system may be required to control foaming. If a closed surface system is used, the defoamer
may have to be added to the injected liquid.

2-156

Chapter 2

Underbalanced Drilling Techniques

Additives
Water is often the most cost-effective liquid for gasification. It is relatively resistant to
contamination by produced fluids. Drilling with water can lead to excessive shale sloughing in
some formations. Adding salts, such as potassium chloride, and suitable polymers, can inhibit
shale hydration and restrict sloughing problems, but these add to the cost of the drilling fluid.
Separators
Water, and other non-viscosified liquids, may not provide sufficient hole cleaning capacity, in
large diameter holes or when sloughing leads to hole enlargement. It may be necessary to use a
drilling mud for the liquid phase. If this has to be done, the liquid phase viscosity and gel
strength should be as low as possible to meet the requirements of good hole cleaning. High
viscosities cause high annular pressure losses and can require increased gas injection rates to
achieve the target pressure downhole. High viscosity and gel strength make separation of gas
and liquid more difficult. If natural gas is flowing into the borehole, gas breakout from the
drilling liquid in any open tanks has to be prevented. This requires efficient de-gassing, upstream
of any open portion of the surface system. To avoid malfunctioning of the mud pumps, good gas
removal is also necessary before the liquid is returned to them.
Gases readily break out of water when it passes through a simple mud/gas separator. An
atmospheric de-gasser will probably be sufficient for gas removal from low viscosity muds. A
vacuum de-gasser may be required for higher viscosity muds.
Field Examples
It has been reported that some increase in liquid phase viscosity, above the viscosity of water,
was beneficial in preventing air slugging in the annulus and flow surging, when drilling
geothermal wells.8 Slugging of the return flow during normal circulation has not been a major
problem in the many wells recently drilled with gasified water and diesel. The higher liquid rates
typically used in these wells may effectively suppress slug flow. However, slugs of liquid and
gas can form with these unviscosified liquids when circulation is resumed, after it has been
suspended for some time; for example after a connection or a trip.
Claytor et al., 1991,9 described drilling a horizontal well in Michigan, using an aerated mud.
This provided an illustration of the problems that can result from inadequate de-gassing. They
used a mud with a yield point in the range of 6 to 12 lbm/100 ft 2. Initially, the surface system
routed the returning mud through parallel gas separators and through two of the rigs mud pits,
before it went back to the pumps. Sufficient air remained in the mud when it was pumped
downhole again that the pumps cavitated. The volume fraction of air in the mud became so high,
due to the air entrapped in the liquid being pumped downhole, that a large air bubble developed
in the annulus and unloaded the well. No additional problems, related to inefficient air removal
from the mud, were experienced after two additional pits and a de-gasser were added to the
surface system.
The liquid phase is normally re-used after it returns to surface. Removal of fine drilled solids can
be important for controlling viscosity and limiting erosion. A high liquid yield point can restrict
solids removal, and should be avoided, if possible.10

2-157

2.7.3 The Gas Phase


Air, nitrogen and natural gas have all been used to gasify drilling fluids. The primary advantage
of air is that it is less expensive than nitrogen. Compressors and a booster are required to supply
the air at the necessary injection pressures and rates. In order to maintain the downhole pressure
within a tolerable range, more precise control of the air rate is needed than for dry air or mist
drilling.
Aerated liquids can be very corrosive. Careful attention should be paid to corrosion inhibition.
Some drilling contractors refuse to allow their drillstrings to be used in aerated liquids because of
high corrosion rates. With careful selection of water, good pH control and appropriate selection
and use of corrosion inhibitors, corrosion rates can be reduced to acceptable levels. The
inhibitors involved are, however, expensive.
Combustion
The risk of combustion rules out using air to lighten hydrocarbon liquids. There is no significant
risk of a downhole fire with an aerated aqueous drilling fluid, provided that the liquid phase
remains continuous. However, downhole fires are possible in horizontal wells drilled with
aerated liquids. As was the case with air foams, the horizontal hole acts as a separator, with air
migrating towards the top or high side of the hole and liquid to the bottom or low side. If the
annulus bridges and the circulating system pressure increases, ignition can occur. To avoid this,
nitrogen or natural gas can be used.
Aerated liquids should not be used with closed surface systems where combustible mixtures of
air and hydrocarbon vapor could form inside the separators (if hydrocarbons are produced).
Nitrogen may be supplied either cryo-genically or by membrane filters operating on location, as
described in Section 2.2 Nitrogen Drilling. Regardless of the nitrogen source, it eliminates the
possibility of downhole fires. Pure, cryogenic nitrogen also prevents downhole corrosion.
Membrane-generated nitrogen contains some oxygen, and downhole corrosion remains a major
concern.
It is possible to inject both nitrogen and compressed air simultaneously.5 Particularly for wells
requiring high gas injection rates, this could be a cost effective procedure, if the reduced nitrogen
cost outweighs the additional cost of the air compression equipment. As was the case with
membrane-generated nitrogen, corrosion of downhole equipment will be a problem. Correct
specification of the proportions of air and nitrogen will, however, prevent the composition of the
drilling fluid from entering the combustible region. Different hydrocarbons will have different
combustible limits. If samples of produced hydrocarbons are available, experiments can be
performed to determine the minimum proportion of nitrogen required to prevent combustion. As
noted in Section 2.2 Nitrogen Drilling, a mixture of air, nitrogen and natural gas will not
normally be combustible if the oxygen concentration is less than 8 percent. Noting that the
concentration of oxygen in air is around 21 percent, using 38 percent air and 62 percent nitrogen
in the injected gas will keep the oxygen concentration at or below 8 percent at all times.
2-158

Chapter 2

Underbalanced Drilling Techniques

Teichrob, 1994,5 found that injecting a 60 percent air - 40 percent nitrogen blend, in water, was
sufficient to prevent combustion in a heavy oil well.
It has been reported11 that hydrogen sulfide can reduce the concentration of oxygen at which a
mixture of natural gas, nitrogen and oxygen can ignite. Until more information is available, it is
advisable not to use mixed nitrogen and air or membrane-generated nitrogen (which can contain
up to 5 percent oxygen) in any application where sour gas production is possible.
Natural gas may be taken from a gas supply line close to the well site, using the equipment
described in Section 2.3 Natural Gas Drilling. Using natural gas demands very efficient gas
removal from the returning liquid, under conditions that allow all of the gas to be collected and
either flared or treated and transferred to the local gas production system.
Provided that the risk of combustion can be controlled, the selection of gas for a gasified drilling
operation ultimately depends on the comparative economics. Gas injection rates tend to be lower
when drilling with an aerated liquid than during conventional air or mist drilling. The gas is
being injected to reduce the drilling fluid density, not to clean the hole. The lower injection
rates can make the high unit costs of liquid nitrogen and natural gas less restrictive than they are
in dry gas drilling.
Recycling
Finally, there is no method to recycle the gas returns. Superficially, recycling looks attractive. If
gas could be collected from the separator, compressed and re-injected, there could be real savings
in the cost of injected nitrogen or natural gas. However, several factors combine to make this
process impractical.11 First, the quality and rate of return gas are not steady. A secondary gas
supply would be necessary to guarantee that a consistent injection rate is maintained. The return
gas would have to be dried, to prevent liquid condensation in the compressors. Since most
compressors are not rated for sour operation, the recycled gas would have to be scrubbed to
remove any hydrogen sulfide that might enter the returns from the formation.
2.7.4

Rheology

Elevated viscosity and yield point should be avoided where possible.


In practice, the flow of a gasified liquid drilling fluid will almost always be highly turbulent. The
Reynolds number, Re, for flow of a liquid up an annulus is given by:
Re =

2-159

15.47 D h W v an

(2.58)

where:
Dh....... hydraulic diameter of the annulus
(the difference between the hole
and pipe diameters) (inches),
W ...... liquid density (ppg),
van ...... average annular velocity (ft/min),
and,
......... liquid viscosity (cP).
Turbulent flow (at least transitional) is often designated as being established once the Reynolds
number exceeds 4,000. Consider water flowing up the annulus of an 8-inch diameter hole
being drilled with 5-inch drillpipe. An annular velocity of 7 ft/min would be sufficient for
turbulent flow. Typically, liquid rates giving annular velocities in excess of 100 ft/min are used
in gasified liquid drilling. Also, the gas bubbles are more likely to promote turbulence than to
suppress it. Turbulent flow should be expected.
The practical consequence of this is that the gasified liquids rheology will have little impact on
circulating pressure losses or hole cleaning, at least when the liquid phase is unviscosified and
annular velocities are reasonably high.
There may be circumstances, particularly when drilling large hole sizes, when it is not possible to
achieve an annular velocity sufficient for good hole cleaning with an unviscosified liquid. In
these instances, drilling mud is used as the base liquid. The combination of low annular velocity
and high viscosity may then prevent the development of turbulent flow. If required, the viscosity
of the gasified fluid can be estimated from the gas volume fraction and the viscosity of the base
fluid using relationships for low quality foams (Equation (2.48), Section 2.5, Stable Foams).
2.7.5

Circulating Pressures

Multi-Phase Flow
Prediction of circulating pressures, when drilling with a gasified fluid, accounts for multi-phase
flow. There are at least three phases in the annulus - liquid, gas, and solid (cuttings). Depending
on the injected liquid and any formation fluid inflow, there may be both aqueous and
hydrocarbon liquid phases. With drillpipe gas injection, gas and liquid phases flow down the
drillstring. With annular gas injection, the flow down the drillstring is purely liquid.
Generally, four different flow regimes have been identified in gas-liquid flow; bubbly, slug,
churn and annular.12,13 Guo et al., 1993,12 noted that the flow regime in gasified liquid drilling is
normally bubbly and it is reasonable to treat the drilling fluid as a homogenous mixture of gas,
liquid(s) and solids.
Because of the compressibility of its gaseous phase, the density of a gasified liquid changes with
pressure. This means that the density of the gasified liquid decreases as it flows up the annulus.

2-160

Chapter 2

Underbalanced Drilling Techniques

The rate at which the density changes can be influenced by the frictional pressure loss. As flow
rates are increased, the frictional pressure loss up the annulus increases, as does the bottomhole
pressure. This compresses the gaseous phase and increases the downhole density of the
circulating fluid. Consequently, the bottomhole pressure of a circulating gasified liquid is
determined principally by the interaction between the hydrostatic head and the frictional pressure
change up the annulus. Strictly, there is also a component of downhole pressure due to
acceleration of the fluid up the annulus. This is small and is generally neglected in any analysis
of circulating pressures.
Formation fluid inflows can be expected when drilling underbalanced with gasified liquids.
These add to the liquid and gas flowing up the annulus. Their effects on the circulating pressure
can be considerable.
Pressure Prediction
Drilling with a gasified liquid can be considered analogous to conditions in a gas-lifted well.7
Although a full analysis of circulating pressures requires computer simulation, in some instances
it is possible to use a static analysis to roughly approximate the bottomhole pressure, without
recourse to more complex simulation. The various factors that influence gasified liquid
circulating pressures are considered below.
Hydrostatic Pressure in the Annulus
The bottomhole pressure, due to a static mud column, can be analyzed in terms of the gas and
liquid volume fractions. Poettmann and Bergman, 1995,14 developed charts showing the volume
of gas required (per barrel of mud) to achieve any particular reduction in effective density of the
drilling fluid. For a static column of fluid:
144

P1
P2

VdP + h = 0

(2.59)

where:
V ........ specific volume of the fluid
(ft3/lbm),
P......... pressure (psia),
P1 ....... pressure at the top of the column
(psia),
P2 ....... pressure at the bottom of the
column (psia), and,
h......... height (feet).
In oilfield units, assuming ideal gas behavior:

14.7 S Tavg + 460


Vm = 5.61 +
520P

2-161

(2.60)

M = 42 MW + 0.0764 GS

(2.61)

where:
Vm ...... total volume (ft3) of gas/bbl liquid
at pressure,
P......... pressure (psia),
Tavg..... average temperature (F),
M ....... mass of mixture (lbm/bbl) of
liquid,
S......... volume of gas (scf/bbl) of mud,
G ........ gas gravity, and,
MW ... mud weight (ppg).
For a static column of mixed gas and liquid in a well of depth, h, Equation (2.59) can be rewritten as:

h=

1 P1
1 P1 2117S Tavg + 460 dP
+
dP
808
M P2
M P2
520P

(2.62)
This can be integrated and re-arranged to find S, the volume of gas (scf/bbl of liquid):
S=

808 ( Pb Ps ) 42 hMW
0.0764 h 4.071 (Tavg + 460) ln (Pb Ps )
(2.63)

where:
Ps........ surface pressure (psia), and,
Pb ....... desired bottomhole pressure (psia).
When the fluid column is flowing up the annulus, work is done against friction between the fluid
and the annular walls (the hole wall, the casings inside surface, and the drillstring). Neglecting
acceleration, the pressures at the top and bottom of a vertical flowing column of fluid are related
to the fluids specific volume, the height of the column and the energy lost to friction, W f, by:
144 VdP + h + Wf = 0
P1

P2

(2.64)

2-162

Chapter 2

Underbalanced Drilling Techniques

Poettmann and Bergman related Wf to a Fanning friction factor, f:


Wf =

2.85 109 fQ 2 V 2 mavg

(D

h + D s ) (D h D s )
2

(2.65)

where:
Q ........ flow rate of liquid (gpm),
Vmavg .. integrated average of Vm between
the surface and the bottomhole
pressures (ft3/bbl),
Dh....... hole diameter (inches), and,
Ds ....... drillstring diameter (inches).
This relationship implies that an average friction factor is taken to represent frictional effects up
the full length of the annulus. Substituting Equation (2.65) into Equation (2.64) and integrating,
the following relationship between well depth, surface and bottomhole pressures is obtained:
808 ( Pb Ps ) + 4.071 (Tavg + 460) ln ( Pb Ps )

h=

(42 MW + 0.0764GS) 1 +

2.85 10 9 f Q 2 Vmavg

2
3
(D h + Ds ) (Dh Ds )

(2.66)
Poettmann and Carpenter, 1952,15 determined the friction factor, f, using a correlation with a
reduced Reynolds number for flow of gas and liquid mixtures up gas wells. The reduced
Reynolds number, RePC, was defined (in oilfield units) as:
R ePC

516
. 10 6 MQ
=
Dh + Ds

(2.67)

where:
Q ........ liquid flow rate, stock tank (gpm).
The correlation between the friction factor and this reduced Reynolds number is shown in Figure
2-38. This correlation was used by Poettmann and Bergman in conjunction with Equation (2.66),

2-163

to compute the quantity of gas required to give a bottomhole pressure of 2,497 psig, when
drilling a 6,000-foot deep 8-inch hole with 4-inch drillpipe and circulating an 8.6 ppg mud at
350 gpm. This example situation would require 14.9 scfm of air per barrel of mud, or 124 scfm
(added to the 350 gpm mud rate) to give the desired bottomhole pressure. Equation (2.63) was
used to compute the volume of air required to give this bottomhole pressure if the mud column
were static. This was 13.4 scf/bbl of mud (112 scfm).
Poettmann and Bergman, 1955, 14 concluded that the difference between static and flowing states
was sufficiently small to allow the much simpler static analysis to be used to estimate the air rate
that would be required to achieve a target reduction in the effective fluid weight. They generated
charts using their static analysis, for different average fluid temperatures. These relate the air
volume required, in scf/bbl of liquid, to the desired reduction in mud density, the initial mud
density, and the bottomhole depth. Figure
Figure 2-39
2-39 provides charts for average fluid temperatures of
100, 150, and 200F. The fluid temperature is often not well known and it does not have a large
impact on the air volume required. These charts are probably sufficient to estimate air
requirements, in applications where the target bottomhole pressure is poorly defined and when
large formation fluid inflows are not anticipated.
When drilling, the cuttings in the annulus will add some component to the bottomhole pressure.
In most instances, this is small and can safely be ignored when predicted circulating pressures.
In order to predict standpipe pressures, flow down the drillstring and through the bit nozzles has
to be considered. If the gas is injected into the annulus, flow down the string and through the
nozzles will be single (liquid) phase.
For liquid only, the prediction of hydrostatic pressure and frictional pressure losses is relatively
straightforward. The flowing fluid compressibility is negligible and, its density does not change
measurably until it reaches the gas injection point. Any conventional drilling hydraulics model
can be used to predict the pressure change through the bit nozzles and up the annulus. The
standpipe pressure will be the sum of the bottomhole pressure, the bit pressure drop, and the
drillstring frictional pressure drop, minus the hydrostatic head in the drillstring. Because the
fluid flowing through the nozzles is incompressible, there is no prospect for sonic flow to isolate
the standpipe pressure from the annulus, unlike dry gas drilling.

2-164

Chapter 2

Underbalanced Drilling Techniques

100

Friction Factor, f

10

0.1

0.01

0.001
1

10

100

Reduced Reynolds Number, RePC

2-165

Figure 2-38.

Correlation between friction factor, f, and the reduced Reynolds number, RePC, (after Poettmann and
Carpenter, 1952 15).

Figure 2-39 (a).

Air volumes required to achieve desired mud weight reductions; average fluid temperature 100F
(Poettmann and Bergman, 1955 14).

Figure 2-39 (b).

Air volumes required to achieve desired mud weight reductions; average


temperature 150F (Poettmann and Bergman, 1955 14).

Figure 2-39 (c).

Air volumes required to achieve desired mud weight reductions; average


temperature 200F (Poettmann and Bergman, 1955 14).

2-166

Chapter 2

Underbalanced Drilling Techniques

If gas is injected down the drillstring, flow will be two-phase down the string and through the
nozzles. The finite compressibility of the gasified liquid means that the pressure drop through
the bit nozzles, if there are any, cannot be computed from the conventional, liquid pressure drop
formula. Guo et al.,1993,12 gave an expression for computing the pressure drop of a gasified
liquid flowing through a nozzle:
G2 1
1
Pa = Pb +

2
g c A n b a

(2.68)

where:
G ........ mass flow rate of drilling fluid
(lbm/s),
An....... total flow area of the nozzles
(feet),
gc ........ gravitational conversion factor
(32.17 ftlbm/lbfs2),
a....... fluids density above the bit
(lbm/ft3),
b ...... fluids density below the bit
(lbm/ft3),
Pa ....... pressure above the bit (psfa), and,
Pb ....... bottomhole pressure (below the
bit) (psfa).
This relationship neglects any energy loss through the nozzles due to frictional effects and any
change in potential energy of the fluid due to the vanishingly small height difference across the
nozzles. Substituting from Equation (2.44), for the density of a lightened drilling fluid, this
becomes:
G 2 Fgo Po Po

Pa = Pb +
2
g c A n o Pb Pa

(2.69)

where:
Fgo ...... volume fraction of gas in the liquid
under standard conditions, and,
o ...... density of the fluid under standard
conditions (pressure, Po) (lbm/ft3).
This expression can be used to compute the pressure above the bit from the bottomhole pressure.

2-167

The effect of temperature changes on fluid density is neglected in this expression. As the liquid
volume fraction is quite high, the thermal capacity of the drilling fluid will be large in
comparison to that of a gas. The temperature change across the nozzles will be much smaller
than for a pure gas.
Hydrostatic and frictional pressures mutually interact during flow down the string. The increase
in flowing pressure, due to friction, causes changes in the hydrostatic pressure. The annular
geometry term (Dh + Ds)2 (Dh - Ds)3 in Equations (2.65) and (2.66) simplifies to D5 for flow down
the string. With this substitution and with the appropriate sign changes, Equation (2.66) can now
be used to predict the standpipe pressure from the pressure above the bit.
Circulation Simulations
Several computer simulators, representing flow of circulating gasified liquids, have been
developed; for example, by Guo et al., 1993,12 Saponja,1995,7 and, by Wang et al., 1995.16
Guo et al.s model was validated by comparing its standpipe pressure predictions with
measurements made when drilling at depths from 3,000 to 7,000 feet, in three different wells,
with liquid rates ranging from 180 to 300 gpm and gas rates ranging from 0 to 680 scfm.
Predicted and measured standpipe pressures ranged from less than 300 psi to 700 psi. In most
instances, the differences between predictions and measurements were 10 percent or less.
The frictional pressure drop, up the annulus, increases rapidly with increasing gas injection rate,
to the extent that it can control the bottomhole pressure at high gas rates. Figure 2-40 illustrates
the effect of gas and liquid injection rates on bottomhole pressures predicted, using Gou et al.s
methodology. This example is for a vertical 8-inch diameter hole, 6,000 feet deep, being
drilled with 4-inch drillpipe and 6-inch collars. Bottomhole pressure is plotted as a function
of the gas injection rate, for liquid rates of 200 and 350 gpm. It is assumed that the annulus is
open to atmospheric pressure at the surface.
Initially, the bottomhole pressure decreases as the gas injection rate is increased. The reduction
in fluid density in the annulus is much more than the increase in frictional pressure drop due to
the increased gas rate. As the gas rate increases, the frictional pressure drop becomes more
significant. There is a critical gas rate for which the bottomhole pressure is a minimum. As the
gas rate is increased beyond this critical value, the frictional pressure drop dominates any
reduction in fluid density and the bottomhole pressure actually increases slightly. Guo et al.,
1993,12 pointed out that the critical rate depends on the annular geometry, the well depth and the
liquid density.
Saponja, 1995,7 termed the region where increasing gas injection rate reduces bottomhole
pressure as hydrostatic-dominated, while the region where increasing the gas rate increases
bottomhole pressure was labeled friction-dominated.

2-168

Chapter 2

Underbalanced Drilling Techniques

In the hydrostatic-dominated regime, the bottomhole pressure is sensitive to small changes in the
gas injection rate. Figure 2-40 shows that an increase in the gas rate from 100 to 200 scfm,
injected into a 350 gpm liquid flow, would reduce the bottomhole pressure from 1,150 psi to 825
psi. The bottomhole pressure is also sensitive to gas inflow. If gas flows from the formation into
the annulus, the bottomhole pressure will be reduced if the flow is in the hydrostatic-dominated
regime. This will allow more gas to flow into the annulus, further reducing the pressure. The
bottomhole pressure will only stabilize when the gas rate becomes sufficiently high to move flow
into the friction-dominated regime.
On the other hand, bottomhole pressures do not substantially change in response to small changes
in gas injection rate, if the flow is friction-dominated. Again, using the example in Figure 2-40,
an increase in gas rate from 500 to 1,000 scfm would cause an increase of only 75 psi in the
bottomhole pressure. This contrasts with the 325 psi bottomhole pressure reduction when gas
flow was increased from 100 to 200 scfm. In the friction-dominated regime, gas inflows tend to
be self-stabilizing. Bottomhole pressure increases in response to the inflow and this reduces the
gas inflow rate. Saponja recommended designing the circulating program so that flow would be
friction-dominated, whenever possible. However, there is no point in using flow rates that are
higher than necessary for adequate hole cleaning. As can be seen from Figure 2-40, the critical
gas rate is roughly proportional to the liquid rate. It is possible to achieve friction-dominated
flow at a lower gas rate, by reducing the liquid rate. Particularly when using nitrogen, the cost of
an unnecessarily high gas rate can be considerable and should be avoided, if at all possible.
Optimization of flow rates will be discussed in more detail later.
2000

Predicted Bottomhole Pressure (psia)

Specifications
Depth: 6000 feet
Hole Diameter: 8 1/2-inches
Drill Pipe: 4 1/2-inches
Drill Collars: 6 1/4-inches
Standpipe Injection

200 gpm
350 gpm

1800
1600
1400
1200
1000
800
600
400
200
0
0

100

200

300

400

500

600

700

800

900

1000

Air Rate (scf/bbl)

Figure 2-40.

2-169

Influence of gas and liquid injection rates on predicted bottomhole pressures.

The impact of formation liquid inflows on bottomhole pressure can be estimated by adding the
formation fluid rate to the injected rate of the appropriate phase. Consider the example shown in
Figure 2-40, with a liquid injection rate of 200 gpm. A liquid inflow of 214 BPH (150 gpm)
would increase the liquid rate in the annulus to 350 gpm. This would increase the bottomhole
pressure at all except very low gas injection rates. At low gas rates, the change in pressure due to
the liquid inflow becomes larger with increasing gas rate. Approaching the onset of frictiondominated flow, the bottomhole pressure becomes less sensitive to the inflow and changes in
bottomhole pressure are smaller with increasing gas rate.
Gas inflows can cause bottomhole pressures to decrease or to increase. The rate of gas inflow
can be added to the gas injection rate to estimate the change in bottomhole pressure. If flow is
hydrostatic dominated, the bottomhole pressure is reduced by the inflow, whereas it is increased
if the flow is friction-dominated.
Whether the inflow is gas or liquid, the inflow rate will be influenced by the bottomhole
pressure. For gas inflow into a hydrostatic-dominated flow regime, reduction in bottomhole
pressure will increase the inflow rate, tending to amplify the pressure change. In all other
instances, formation fluid inflows increase the bottomhole pressure and therefore tend to be selfregulating.
In general, it will be necessary to adjust the gas and/or liquid injection rates to re-establish the
desired bottomhole pressure if an inflow occurs. For a liquid inflow, a corresponding reduction
in liquid injection rate will restore the bottomhole pressure. If the inflow occurs well above the
bit, this could lead to hole cleaning problems. The correct response to a gas inflow will
depend on the flow regime prevailing downhole. It can be possible to exploit formation fluid
inflows to reduce gas injection rates along with the cost of the injected gas. This will be
discussed further in the section on operating procedures.
To adequately balance an injection gas rate with a gas or formation fluid inflow requires a great
deal of knowledge about the flowing formation. This becomes more difficult if the inflowing
formation is damaged during tripping. Response needs to be monitored and changes made on the
fly.
Drillstring injection reduces gas requirements and bottomhole pressures. Figure
Figure 2-41
2-41 compares
bottomhole pressures, predicted for the example 6,000 foot well (Figure 2-40) with drillstring
injection and with injection through a parasite string ported into the borehole at 2,000 feet. The
minimum bottomhole pressure for drillstring injection with a liquid rate of 350 gpm is around
650 psi, whereas injection down the parasite string cannot reduce the bottomhole pressure below
1,200 psi at this liquid rate. A bottomhole pressure of 1,250 psi would require less than 100 scfm
of gas injected into the drillstring, but would require 200 scfm if it were injected down the
parasite string. The same effects would be seen with other methods of annular gas injection.

2-170

Chapter 2

Underbalanced Drilling Techniques

Due to the compressibility of a gasified liquid, its equivalent circulating density, ECD, changes
with depth. Figure 2-42 shows the equivalent circulating densities, predicted for the 6,000 foot
hole used above, with air being injected into the liquid at a rate of 100 scf per barrel. With both
drillstring and parasitic string gas injection, the ECD is 2 ppg at about 300 feet from surface.
The ECD increases to 5 ppg at 6,000 feet, if the gas is injected into the drillstring, while it
exceeds 6 ppg at this depth with the same injection rate down the parasite string. This
demonstrates that it is possible to be at balance or even overbalanced at the hole bottom while at
the same time being substantially underbalanced near surface. It is possible to lose circulation
into a formation near the hole bottom while at the same time formation fluids flow into the
borehole from a permeable zone up the hole.
It is possible to choke back the return flow, to pressurize the annulus, at least within the pressure
limitations of the rotating blowout preventer. When drilling with foam, this may be necessary to
maintain foam throughout the annulus. Good hole cleaning, with a gasified liquid, does not
require that the liquid volume fraction has to be maintained within any specific range. Choking
back the return flow will increase the bottomhole pressure. It will also increase the standpipe
pressure and the gas compression requirements. Any choke system should be monitored for
erosion by cuttings in the return flow. Bottomhole pressure can normally be increased by
reducing the gas injection rate or by increasing the liquid rate, without significant change in gas
consumption.

Predicted Bottomhole Pressure (psia)

2500
Standpipe
Parasite String

Specifications
Liquid Rate: 350 gpm
Depth: 6000 feet
Drilled Diameter: 8 1/2-inches
Drillpipe: 4 1/2-inches
Drill Collars: 6 1/4-inches
Parasite string at 2000 feet

2000

1500

1000

500

0
0

100

200

300

400

500

600

700

800

900

1000

Air Rate (scf/bbl)

Figure 2-41.

2-171

Comparison of bottomhole pressure predicted for drillstring (standpipe) and annular


(parasite string) gas injection.

10

1
Standpipe (Equivalent Circulating Density
Parasite (Equivalent Circutating Density)
Standpipe (Volume Fraction Air)

0.9
0.8

Parasite (Volume Fraction Air)

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0

1000

2000

3000

4000

5000

Gas Fraction

Equivilent Circulating Density (ppg)

0
6000

Measured Depth (feet)

Figure 2-42.

Predicted equivalent circulating densities and gas volume fractions as functions of


depth.

In general, a backpressure should not be imposed on the annulus when drilling with a
gasified liquid.
Finally, as is the case when drilling with other compressible fluids, the standpipe pressure does
not necessarily respond to changes in annular pressure. Saponja, 1995,7 compared pressures
recorded downhole and at the surface, during a gasified liquid drilling operation (refer to Figure
2-43
2-43). The annular pressure was in the range of 325 to 450 psi when drilling normally, but
increased to over 870 psi when the annulus apparently packed off due to inadequate hole
cleaning. The standpipe pressure did not give any indication of a downhole problem.
2.7.6 Hole Cleaning
The methods for ensuring good hole cleaning with a gasified liquid are influenced by the
properties of the base liquid, by the hole geometry, and to some extent by the formation fluid
inflow. Annular velocities of approximately 100 to 200 feet per minute are required to keep
holes clean with unviscosified liquids. It may not be possible to generate that high an annular
velocity in large diameter holes. If this is the case, the base liquid viscosity will have to be
increased to provide adequate cuttings transport capacity. A significant formation fluid inflow
may be exploited to improve cuttings transport, by allowing it to add to the annular flow rate.

2-172

Chapter 2

Underbalanced Drilling Techniques

Settling Velocity
Cuttings transport in a simple gasified liquid can be forecasted using an analysis presented by
Guo et al., 1993.12 Simple spreadsheet versions can be coded. The key assumption made is that
a gasified liquid can be treated as an homogenous mixture; this allows a cuttings settling
velocity to be determined by the average flow properties of the mixture. The settling velocity of
a particle of diameter, dc (inches), was taken to be:
f

v t = 92.6 d c c
f

(2.70)

where:
c....... cuttings density (ppg), and,
f ....... drilling fluids average density, at
the prevailing temperature and
pressure (ppg).
Equation (2.70), derived by Moore, 1974,17 incorporates a drag coefficient of 1.5. This is typical
of cuttings falling through a viscous liquid. Gray, 1958,18 found lower values for the drag
coeffcient, ranging from 0.8 for rounded particles to 1.4 for flat particles; for cuttings falling
through air, at elevated pressures. Since gasified liquids tend to have flow properties closer to
those of the base liquid than those of gases, Cd = 1.5 is assumed here. The settling velocities are
not very sensitive to the value of the drag coefficient. A cutting, with a diameter of 0.25 inches
and a density of 21 ppg (specific gravity 2.5), in a fluid with a downhole density of 5 ppg
(specific gravity 0.6), will have a terminal velocity of 83 ft/min. The annular velocity must
exceed this value if the cutting is to move uphole.

2-173

8000
7500
7000

Annular Pressure (kPa)

6500
6000

Hole Problems

5500

Hole Cleaning Inadequate


Hole Packed Off
Not inidicated by Surface Pressure

Wiper Trip

5000
4500
Bottomhole
Pressure

4000
3500
3000
2500

Surface Pressure

2000
1500
1000
500
0
0

100

200 300 400

500

600 700 800 900 1000 1100 1200 1300 1400 1500 1600 1700 1800 1900

Time (min)

Figure 2-43.

Comparison of downhole and surface pressures (after Saponja, 1995 7). To convert
from kPa to psi, multiply by 0.145.

Guo et al., 1993,12 assumed that the cuttings concentration in the annulus should not exceed
some critical value, Cc, if hole cleaning problems were to be avoided. The velocity, Vc, with
which cuttings must travel uphole to keep their concentration in the annulus below the critical
level was related to the penetration rate ROP by:
ROP
60 C c
where:
vc =

(2.71)

vc ........ critical velocity (ft/min),


ROP ... rate of penetration (ft/hr), and,
Cc ....... cuttings concentration (fraction).
Taking the critical concentration as 4 percent, cuttings would need to travel uphole with a
velocity twenty-five times greater than the penetration rate. For a penetration rate of 30 ft/hr, this
corresponds to 12.5 ft/min. In a vertical well, the minimum annular velocity required for this
cuttings velocity is the sum of the terminal velocity and the required cuttings velocity; 95.5
2-174

Chapter 2

Underbalanced Drilling Techniques

ft/min in this case, for vt = 83 ft/min. Larger cuttings, lower circulating fluid densities and higher
penetration rates will require higher annular velocities, as will highly deviated and horizontal
wellbores. For example, Saponja, 1995,19 reported hole cleaning problems in a horizontal well
drilled with nitrified water, when the annular velocity was 400 ft/min. A second horizontal well,
drilled with nitrified crude, showed adequate hole cleaning, with an annular velocity of nearly
600 ft/min.
Examples were given by Guo et al., 1993,12 relating the gas and liquid injection rates required for
good transport of different sized cuttings, for several different well geometries. Figure 2-44 is
one example; for a 6-inch diameter hole being drilled with 3-inch drillpipe, at a depth of
10,000 feet. Increasing the gas injection rate increases the size of cutting that can be lifted
uphole, whatever the liquid flow rate. In this example, a dry air injection rate of 2,400 scfm
would be sufficient for efficient hole cleaning with 0.25-inch diameter cuttings. A liquid
injection rate of less than 110 gpm (14.7 cfm) would actually reduce the efficiency of cuttings
transport. That is, it would decrease the size of cutting that could be lifted uphole efficiently.
The liquid rate would have to be above 110 gpm to maintain or increase cuttings transport
capacity.
The downhole density of the drilling fluid depends on the relative gas and liquid injection rates.
If the liquid injection rate is increased while the gas rate is held constant, the circulating fluid
density will increase. However, the average annular velocity of the drilling fluid downhole will
initially decrease, as the liquid injection rate is increased. At first, the higher downhole pressure
reduces the volume of the gas phase by more than the increase in the liquid injection rate. With
further increases in liquid injection rate, the rate of reduction in gas volume becomes smaller
than the increase in liquid volume, and the annular velocity begins to increase. As illustrated in
the example in Figure 2-44, this can cause the efficiency of cuttings transport to decrease if the
liquid injection rate is increased. This is analogous to drilling with mist rather than dry air; that
is, when injecting liquid, the air rate required to lift any given size of cutting increases. This
should be considered when designing a gasified liquid circulating program.
When designing the circulation program, it is possible to use formation fluid inflow to improve
cuttings transport efficiency. If the inflow occurs at the hole bottom, its effect on cuttings
transport can be determined by adding the inflow rate to the injected fluid rate(s). Gas inflow at
the bit will always improve cuttings transport. Liquid inflow at the bit may or may not improve
cuttings transport, depending on how the additional liquid flow affects bottomhole pressure and
annular velocity.
While this supplementary transport capacity can be a bonus, care should be exercised if drilling is
to continue below the inflow. Cuttings transport between the bit and the depth of the inflow may
be impeded by the additional fluid flow. As noted above, a gas inflow can either increase or
decrease the bottomhole pressure.

2-175

A liquid inflow will always increase the bottomhole pressure. If the bottomhole pressure
increases because of an inflow, the annular velocity between the bit and the inflow site will
decrease and the circulating fluid density will increase. An increase in bottomhole pressure
can reduce the efficiency of cuttings transport, even though the circulating fluid density
increases. Gas inflow may hinder cuttings transport between the bit and the inflow site. This
will probably not be a significant concern until the drill collars are below the inflow, since the
annular velocity is higher past the collars than it is past the smaller diameter drillpipe.
350
Cuttings Dimension
1/16 inch
1/8 inch
1/4 inch
1/2 inch
3/4 inch

Mud Flow Rate (gpm)

300

250

200

150

100

50

0
0

1000

2000

3000

4000

5000

6000

7000

8000

Air Injection Rate (cfm)

Figure 2-44.

Gas and liquid injection rates required for efficient cuttings transport (after Guo et al.,
199312).

The analysis of cuttings transport presented so far has assumed turbulent flow. With a large
annulus, it may not be possible to achieve an annular velocity sufficiently high for efficient
cuttings transport. In this case, the circulating fluid viscosity will have to be increased, in order
to reduce cuttings settling velocity. This will probably involve a substantial increase in the
liquid phases viscosity, bringing annular flow into the laminar or transitional regimes. A
cuttings settling velocity depends on the fluids effective viscosity.

2-176

Chapter 2

Underbalanced Drilling Techniques

A standard cuttings transport analysis17 gives the settling velocity, for a cutting falling through a
fluid in transitional flow, as:

( )
( )

0.667

v t = 175 d c

0.333

(2.72)

where:
vt ........ terminal velocity (ft/min),
dc ........ average cuttings diameter
(inches),
c....... cuttings density (ppg),
f ....... drilling fluids density (ppg), and,
........ fluids effective viscosity
(i.e. accounting for annular flow)
(cP).
For example, consider a 0.25-inch diameter cutting, with a density of 21 ppg, falling through a
fluid with a circulating density of 5 ppg. If the maximum attainable annular velocity is 15 ft/min,
Equation (2.72) indicates that the settling velocity would have to be restricted to 17.4 ft/min for
effective hole cleaning, at a penetration rate of 30 ft/hr. This would require an effective viscosity
of 160 cP. Equation (2.48) can be used to estimate the effective viscosity of the gasified liquid,
from the viscosity of the base liquid and the gas volume fraction. Remember that a high liquid
viscosity will make de-gassing the returning fluid more difficult. This can lead to significant
operational problems.
2.7.7

Equipment

Gas Injection System


Gas supply and compression equipment, appropriate to the selected gas, will be required; these
are described in Section 2.1 for air, Section 2.2 for nitrogen and Section 2.3 for natural gas.
Whatever gas is used, the injection system should be capable of delivering the required gas
volume, at the highest anticipated injection pressure. For drillstring gas injection, the gas supply
must be able to overcome the standpipe pressure. This can be of the order of 1,000 psi. For
parasite tubing or casing injection, the gas delivery pressure at the injection point must exceed
the hydrostatic pressure of a column of any anticipated wellbore liquid. This can also easily
exceed 1,000 psi, depending on the depth of injection. With air, on-site generated (membrane
filter) nitrogen, or natural gas, boosters will probably be needed. A cryogenic nitrogen pump unit
will probably have adequate delivery pressure for most normal drilling situations.
Whatever the type of gas and injection technique, the rate of gas injection must be monitored and
controlled more precisely, when drilling with gasified liquids, than when drilling with dry gases
or foams. Otherwise, the desired bottomhole pressure will not be maintained. Many compressors
provide little possibility for delivery rate control by varying operating speed. An adjustable

2-177

choke should be set up on the gas delivery line, to allow controlled venting of part of the
compressor output through a bypass line, direct to the flare pit. An orifice meter should be
installed downstream of the choke to measure the volume of air delivered to the well. The
metered flow rate can then be used to adjust the choke setting, as required.

Claytor et al., 1991,9 described using a backpressure valve, controlled by feedback from a
downstream flow meter, to regulate the volume of gas bypassed. This responds rapidly to
variations in standpipe pressure, such as those due to bit torque changes when drilling with a
downhole motor.
Cryogenic nitrogen pump units give much better control of delivery volume. The pump speed
can be adjusted to deliver the appropriate nitrogen volume and excess gas does not have to be
bypassed.
If natural gas will be injected down the drillstring, the gas supply should be connected to the
standpipe in the same way as for dry gas or mist drilling (refer to Sections 2.1 and 2.4).
Parasite Tubing String
The parasite gas injection string can be continuous (i.e. coiled), or jointed tubing.
Continuous tubing itself will be less costly than jointed tubing. It is more costly to rig up to run
continuous tubing, but once rigged up, it is usually quicker to run than jointed tubing. Only the
casing connections have to be made.
Whatever tubing is selected, it must have sufficient pressure integrity to survive any feasible
combination of internal and external pressures that it may experience during installation and
operation. The most stringent collapse condition will probably arise when cementing the casing
string it is attached to. Filling the parasite string with water or mud will offset the load trying to
collapse the string. The parasite tubing could be exposed to gas at full bottomhole pressure
minus the gas density gradient, or to gas at the formation pressure gradient. The parasite tubing
should be designed to the same burst rating as its host casing.
In many instances, 1-inch diameter coiled tubing will be adequate. If the parasite string is
jointed, 1.66-inch non-upset 10 threads per inch tubing is often used. When a high gas injection
rate is required, (for example, when drilling large diameter hole), or when the injection point is
very deep, the frictional pressure drop down the injection line may become significant. Under
these circumstances, larger diameter tubing (greater than one-inch) coiled tubing may be
required, in order to limit the gas injection pressure at surface. Clearance between the host
casing string and the annular openhole space through which the parasite string is run must be
considered when selecting the parasite tubing size. There has to be sufficient room for the
parasite tubing and whatever clamps are used to attach it to the host casing.

2-178

Chapter 2

Underbalanced Drilling Techniques

If the parasite tubing is jointed, the clearance must be sufficient for both the tubing outside of the
casing connection and the tubing connection outside of the casing body. Appropriate length pup
joints of the parasite tubing should be used to ensure that tubing connections are not made
opposite casing connections. This would increase the casing (drift) and openhole diameter
required to accommodate the combined casing-parasite tubing double string.

The lower end of the parasite string is normally ported into the casing through a custom casing
sub20 (refer to Figure 2-45). This is located in the casing, above the float collar. After cement has
been placed, water is circulated down the parasite tubing and into the casing to ensure that
cement cannot plug the parasite tubing. Alternatively, it is possible to use a gas-injection float
shoe,21 as shown in Figure 2-46. In that case, the lower end of the parasite tubing remains closed
until the float equipment is drilled out. This prevents cement from entering and allows the
parasite string to be pressure tested before the shoe is drilled out and the string is opened to the
annulus.
A jointed parasite string can simply be screwed into an appropriate box connection on the
injection sub or shoe. Consideration should be given to how a continuous parasite string will be
attached to the injection sub, since it is not possible to rotate the string during installation. It can
be welded onto a stub of tubing, previously screwed into the sub and projecting above the upper
end of the sub, or, with good access, it may be possible to weld the tubing directly to the sub. A
compression type fitting on the sub can also be used, but this will require more annular clearance
than a welded connection.
Pin End
2 1/6 -inch Parasite into
13 3/8-inch Casing

Box End

Figure 2-45.

2-179

Parasite tubing injection sub


(after Westermark, 1986 20).

Figure 2-46.

Parasite tubing injection


float shoe (after Comeau,
1995 21).

The connection of the parasite string to the casing must be strong enough to support tensile loads
which may be imposed when running and by the different pressure regimes that can be
experienced during cementing and subsequent operations.

The parasite string should be clamped to the outside of the casing, to avoid buckling of the tubing
under its own weight, if it is unsupported at any time during running. This can be done with
hinged, bolt-on clamps similar to modified casing centralizers.20 Ensure that there is sufficient
clearance between these clamps and the rotary table so that they will not snag when run or pulled
back through the table. Westermark, 1986,20 reported welding segments of clamp to the casing
connections, in order to minimze the outside diameter of the clamp. If this is done, consider the
impact of the welding operation on casing integrity.
A coiled tubing unit should be used when a continuous parasite string is run. Coiled tubing can
be pulled directly off an unpowered spool, by using the weight of the casing string being run.
But, there is no possibility for re-spooling the tubing if it is necessary to retrieve the casingtubing string. An injector head is not necessary. A coiled tubing gooseneck should be hung off
above the rig floor, out of the path of the block and elevators. The coiled tubing is run over the
gooseneck and through the rotary table with the casing. Ensure that it is possible for the slips to
support the casing without pinching the parasite tubing.
Continuous tubing can normally be pulled far enough to one side to clear collar-type elevators. It
may not be possible to bend the tubing sufficiently to clear slip-type casing elevators. With
jointed pipe, it is possible to keep the upper end of the parasite tubing below the casing elevators
and slips at all times,20 permitting any type of elevators and slips to be used. To do this, the first
joint of the parasite tubing has to be shorter than the first joint of casing above the injection sub.
 Make up the parasite tubing to the injection sub, and then make up two joints of casing.
 Pull back the casing until the parasite tubing is above the rotary table.
 Make up one joint of parasite tubing and clamp it to the casing.
 Lower the casing and hang off at the top connection.
 Pick up and make up the next joint of casing.
 Pull back the casing to expose the parasite tubing, add the next joint of parasite tubing and
repeat the process.
Often the parasite tubing will be run outside of the surface casing string, there will be no
wellhead and the parasite tubing will simply extend from the conductor pipe at the surface. In
this case, it is important that the casing is cemented back to surface. The parasite tubing and its
clamps will make it very difficult to run a top-up string into the annulus if cement returns are not
obtained and a top-up cement job is necessary.20 If there is a potential for lost cement returns,
consider running a top-up string with the casing, in addition to the parasite tubing string.
2-180

Chapter 2

Underbalanced Drilling Techniques

The parasite string will push the casing off-center in the conductor. When the wellhead, BOP
and rotating head or rotary BOP are installed on the casing that carries the parasite string, these
will not be aligned with the center of the rotary table. The degree of eccentricity can be sufficient
to cause rapid wear of the rotating head or RBOP sealing elements.20 Functionality of this
control equipment is essential. The rig can be skidded across to re-center the rotary table, after
setting the surface casing-parasite tubing string. If the conductor pipe will be set with the rig, it
is much easier for it to be pulled to one side before its cement sets. If this is done, the surface
casing-parasite tubing string will be centered below the rotary table when it is installed. If the
conductor pipe was driven before rigging up, the rig should be located so that the rotary table will
be centered on the (eccentric) surface casing-parasite tubing string and not on the conductor pipe.
If a parasite string is run outside the intermediate casing, it will have to penetrate the wellhead at
the surface and seal against it. Clearance for the parasite tubing inside the wellhead will be very
limited. It may be possible to obtain an eccentric casing hanger. If not, a reduced diameter
section of tubing may have to be used inside of the wellhead.
A check valve and a ball valve should be installed on the parasite tubing at the surface. These
prevent any fluid flow into the gas supply system and provide well control in the event that the
well has to be shut-in. The gas supply is connected to the parasite string through the ball valve.
Annular Gas Injection
Recognizing the wellhead complexities described above, it may be easier to temporarily install a
second casing string; inside the intermediate casing, to allow gas injection down the annulus
between these two strings; rather than using a parasite string on intermediate casing. If this is
done, the size of the intermediate casing may have to be increased to accommodate a temporary
string with sufficient drift diameter for the hole section to be drilled. This in turn may require
that the previous hole section is drilled to a larger diameter than would have otherwise been
done. The diameter of the lower hole section may also have to be reduced. Deis et al., 1995,22
gave an example of horizontal wells drilled with and without annular gas injection. 7 5/8-inch
intermediate casing was set and 5-inch casing was temporarily hung off to allow annular gas
injection while drilling a 4-inch hole. Without annular injection, the intermediate casing was
7-inch diameter and the hole below was drilled to 6 1/8-inch diameter.
In some instances, a special casing head has been used to hang off the temporary casing string,
inside the intermediate casing5,22 (refer to Figure 2-47). These heads typically have two pairs of
two-inch outlets, to give separate access to the intermediate-surface annulus and to the
temporary-intermediate annulus down which the gas was injected. It is advantageous for the
temporary strings hanger to have a sufficiently small outer diameter so that it can pass through
the BOP stack.5 This allows the temporary casing to be retrieved without having to remove the
BOP stack.
In certain cases, it may be possible to use conventional wellhead equipment. The temporary
string can be hung off inside a spool which is left empty when the temporary string is removed.
2-181

444.

76.2 mm
Cut-off Length for
244.5 mm Casing

60

6m
.

5 mm

20

1m
.

This will normally result in a larger gas injection annulus than the special bowl. It may also lead
to a wellhead and BOP stack-up so tall that it can be difficult to accommodate underneath the rig
floor.

339.7 mm
244.5 mm
177.8 mm

Figure 2-47. Casing head for micro-annular gas injection (after Teichrob, 1994 5).

As was the case with parasite tubing injection, it is necessary that the gas supply system can be
isolated from the casing annulus at the surface, if the well has to be shut-in. Connecting the gas
supply to the wellhead through a ball valve and a check valve allows this and also prevents any
liquid from flowing from the well into the gas supply system. Local well control regulations may
require a second ball valve for redundancy.
Some operators run two joints of casing, with the same diameter as the temporary string, at the
shoe of the intermediate string. This permits the well to be packed off before retrieving the
temporary casing, ensuring well control during this operation.
The temporary string is sometimes run with an anchor and pack-off tool, to locate it firmly in the
permanent casing inside which it is run. Injection ports are then milled in the temporary casing
above the pack-off. There has been concern that cuttings can get through these injection ports,
(for example, if the gas injection rate is reduced) and settle on the pack-off, preventing its release
when the temporary string is pulled. To avoid this, Teichrob, 1994,5 recommended running an
aluminum turbolator, between shear-pinned stop collars, at the base of the temporary string (refer
to Figure 2-48).
2-48

2-182

Chapter 2

Underbalanced Drilling Techniques

This should minimize the extent to which cuttings could build up around the temporary string. It
would also allow the collars to be sheared off and the casing retrieved, leaving the drillable
turbolator downhole, if an excessive buildup of cuttings did occur.
It is possible to run the temporary casing as a tie-back into a drilling liner. A slotted joint at the
bottom of the temporary string can be used to route the injected gas into the returning drilling
fluid.

1,000 mm

Stop Collar

1.6 mm Deep Groove

300 mm

10.3 mm

177.8 mm Casing

Turbolator

Figure 2-48.

Shear-mounted
Stop Collar Set
at 44,000 daN

Radius
End
12.5 mm

Concentric casing string shoe joint (after Teichrob, 1994 5).

Liquid Injection
The mud pumps on a conventional rig will normally be capable of injecting the liquid phase. In
some instances, the required liquid injection rate may be too low for the rig pumps. Then, it will
be necessary to install a smaller triplex pump, with the capacity to operate continuously at or
below the required rate. The delivery pressure capacity of this pump should exceed the highest
anticipated standpipe pressure.
Generally, the injection pumps will be fed from the rigs active mud tank. Returns will not
normally be taken directly to the rig tanks. The return system will be discussed below.
It is advisable to have an adequate supply of kill fluid, stored on site. Typically, this is 1.5 times
the well volume. In some instances, local regulations will dictate the volume required. When
drilling depleted reservoirs, the injected liquid itself may be sufficient to kill a well. If this is not
the case, storing the kill fluid may require additional tanks on site, over and above the rigs mud
tanks.

2-183

Drillstring
The BHA for a well drilled with gasified liquid should be kept as simple as possible.
Components such as stabilizers, spiral or square collars, and jars can prevent a pressure seal and
may damage the sealing element when run through a RBOP or rotating head.
As with other underbalanced drilling techniques, drillstring floats are required. Unlike the
techniques described so far, however, the string is often tripped with the well under pressure.
Normally, the float is retained in the string by the pin above it. If there is pressure beneath the
float, it may be forced out of the string when the connection above it is broken. This can be
avoided by machining the connection into which it is placed to accept a snap-ring above the
insert, to hold the insert in place. Alternatively, a short joint (pipe or collar, as appropriate) may
be run above the float.7
A float in the bottom stand of the string allows the entire string to be pulled with pressure
beneath it. Canadian practice is to use two flapper-type float valves in this last stand for
redundancy7 and to run a flapper-type upper string float every 330 feet (100 m). These floats
reduce the time taken for the drillstring pressure to bleed down on connections. They can also
limit separation of gas and liquid inside the string, when flow is shut down.
When tripping with pressure beneath the string, a special tool is used to release pressure trapped
beneath each float. This is installed on top of the float sub and a pin is screwed down to depress
the flapper valve. A side outlet directs the released, pressurized fluid flow laterally away from rig
floor personnel.
A custom designed hydrostatic control valve has been used in place of conventional or
modified drillstring float valves.21 This uses a second, spring-loaded seal beneath a float valve,
to prevent fluid from flowing downhole until the pressure above the valve reaches some
predetermined value. In this way, it is possible to keep the string above the hydrostatic control
valve filled with liquid during a connection, even though the annular pressure may be less than
the hydrostatic head of the liquid inside the drillstring. This in turn prevents air from being
drawn into the string at a connection, even if the string has been filled with liquid to allow a
conventional pressure pulse MWD system to transmit a survey uphole.
Finally, the influence of the strings geometry on circulating pressures should be considered. A
large diameter string decreases the frictional pressure down the string, reduces the standpipe
pressure, and increases the pressure drop up the annulus. In a prolific well, the increased annular
pressure drop may be desirable since it will tend to limit the surface flowing pressure. On the
other hand, it may make it difficult to achieve sufficient downhole pressure reduction for low
pressure reservoir applications. In these instances, it may be appropriate to use smaller diameter
drillstring components or even a tapered casing string.7

2-184

Chapter 2

Underbalanced Drilling Techniques

BOP Stack
The BOP stack design depends on anticipated well conditions and local regulations. A number
of different, successful configurations have been reported.
In the United States, many of the gasified liquid drilling applications to date have used aerated
water or mud to overcome lost circulation in hole intervals above known reservoirs, where
surface pressures are expected to be low. In these cases, rotating heads are used on top of
conventional BOP stacks, to seal around the drillstring and divert flow into the flow line.
In contrast, nitrified liquids are often used in Canada, to allow very productive wells to be drilled
underbalanced. Normally, a rotating blowout preventer (RBOP) is used rather than a lower
pressure rotating head. If the rig is equipped with a top drive it is possible to use dual annular
BOPs to give a high pressure seal around the string above the return line. Since an RBOP gives
better control of the closing pressure and has lower stripping friction, it is usually preferred.10
It has been recommended that blind rams should be installed at the bottom of the BOP stack.7
This maximizes the distance between the blind ram and the RBOP, allowing short, irregularly
shaped BHA components to be run into the well under pressure, without relying on the RBOP or
annular to seal around them.
A second set of pipe rams, below the blind rams, provides redundancy. They are normally only
used to shut-in the well, in the event that work is required on one of the elements higher up the
stack.23
Carefully consider how the BOP stack configuration will affect ram-to-ram stripping operations,
if there is any possibility of having to strip into the well under high pressure.
Clearance beneath the rig floor needs to be considered when designing the well cellar, wellhead
and BOP stack. BOP stacks, for underbalanced drilling, tend to be taller than those used in
conventional operations. Clearance is a greater problem if con-ventional wellhead equipment is
used to inject gas down a temporary casing string.
Return System (Low Surface Pressure and Limited Hydrocarbon Production)
The surface systems used to drill with gasified liquids vary considerably from application to
application. In areas where low surface pressures and limited hydro-carbon production are
anticipated, simple surface systems can be used.
If aerated water is used, the return line can be routed from the rotating head, out to a combined
return and flare pit; as would be done when drilling with dry air, mist or foam. No additional gas
separator is needed. A centrifugal pump is used to transfer water from the pit to the rigs mud
pumps, for re-cycling downhole. The suction line is fitted with a foot valve and filter cover, to
avoid having to re-prime the transfer pump and to prevent excessive solids pick-up. There
should also be a choke line, running from below the blind rams on the BOP stack, through the
rigs choke manifold, and out to the pit. This is shown in Figure 2-49.
2-185

When mud or brine is used as the liquid phase, cost limits the circulating volume. In this case,
returns will not normally be routed to the pit. Instead, the return flow line leads either directly to
the rigs shale shakers or to a mud/gas separator, mounted alongside the shakers (refer to Figure
2-50 The separator stand should be set so that the liquid discharge is high enough to run freely
2-50).
onto the shakers. Mud/gas separators are available, in a range of sizes, as rental items in most
parts of the United States. Both the flow line and the separator should be sized to handle the
maximum anticipated gas and liquid return rates; the sum of the produced and injected fluid
rates. The flow line is often six-inch diameter, although up to ten-inch diameter may be required.
It should be fitted with valves to allow flow to be sent either to the shaker or to the separator.
Typically, the inlet from the flow line to the separator will be six-inch diameter, as will the outlet
feeding from the separator to the shakers. A four- to six-inch diameter line should take gas from
the separator out to the flare/reserve pit. There should be an ignition source here, to permit
flaring, along with gas and hydrogen sulfide detectors; as are used in dry air, mist and foam
drilling.
Return System (High Hydrocarbon Production and/or Elevated Surface Pressure)
If there is any potential for significant hydrocarbon production or elevated surface pressures,
some means of closing the flow line is needed, adjacent to its connection to the BOP stack. An
emergency shutdown device (ESD) may be required in the flow line. Although not normally
used, it is possible to mount a manual choke between the flow line and the mud/gas separator, 24
if it is essential to apply a backpressure to the well.
There will normally be one set of choke and kill lines in these operations, with two valves
between the BOP stack and the choke line. Local regulations may dictate that one of these is
hydraulically controlled, and may also require that a check valve is used to prevent flowback into
the well. The choke line to the rigs manifold should have a pressure rating equal to the manifold,
BOP stack and wellhead arrangement. Two discharge lines, from the choke manifold, should
lead to the mud/gas separator and to the pit. These lines, and the choke line, should be able to
handle the maximum anticipated flow rate, without creating excessive backpressure anywhere
upstream.

2-186

Chapter 2

C o m p re sso r

Underbalanced Drilling Techniques

C o m p resso r

Booster

Flow Line
Rig Tanks
Pit
Mud
Pump

Ch oke
Manifold

Kill Line

Transfer
Pump

Figure 2-49.

A typical surface system, for drilling with aerated water, when significant oil
production and elevated surface pressure are not expected.

Gas (4-6-inch Outlet)

Mud and Gas


(6" Outlet)
Mud/Gas
Separator

Shaker

Butterfly
V alve

Mud and Gas


(6" Inlet)
Gate
V alve

R ig

Figure 2-50.

2-187

A typical surface layout, for drilling with aerated mud or brine, when significant oil
production and elevated surface pressure are not anticipated.

When significant liquid hydrocarbon production or elevated surface pressures occur - for
example, when planning to drill underbalanced through a known reservoir - a more sophisticated
surface separation system is probably required. This may be open or closed. Closed systems are
normally necessary, if there is any possibility for H2S in the produced fluids. It is not
advisable to use a closed system when drilling with an aerated liquid which could lead to the
formation of a combustible mixture in the separator. Section 2.8, Flowdrilling, describes
open system design and operation, and provides guidelines on choosing an open or closed
system. Details of the design and operation of closed systems are given in Section 2.9, Closed
Surface Systems.
Instrumentation
The amount of instrumentation, required to drill underbalanced with gasified liquid, depends on
the bottomhole pressure. In applications, where the gasified liquid is being used to overcome lost
circulation, there is no potential penalty in terms of formation damage if the well is
unintentionally drilled overbalanced for short periods. In that case, no extra instrumentation is
required beyond that used on an air drilling job. This includes, in addition to the normal rig
instruments, an orifice meter to gauge the air delivery rate, good air and liquid delivery pressure
gauges, an air delivery temperature gauge, and natural gas and hydrogen sulfide detectors on the
rig floor, at the shakers and at the flare/reserve pit.
More sophisticated instrumentation is necessary if the bottomhole pressure needs to be kept
within tightly specified limits; to avoid formation damage without compromising wellbore
stability or well control. Perhaps the best indication of the bottomhole pressure is direct
measurement downhole, using an MWD unit; several of which are able to measure annular
pressures. Conventional pressure pulse telemetry MWD systems cannot transmit information
from downhole to the surface, if the drillstring contains compressible fluid. Once the gas
concentration in the fluid, through which the signal is being transmitted, exceeds about 90,000
ppm, signal attenuation becomes too rapid for reliable transmission.21 In some drillstring gas
injection applications, it may be acceptable to fill the drillstring with a supposedly, nondamaging liquid, before sending MWD information uphole. This necessarily disturbs the
bottomhole pressure and normally results in a temporary overbalance downhole (a situation that
the pressure information being sent uphole is supposed to help avoid). The practical implication
is that mud pulse MWD systems can only effectively be used for real-time downhole pressure
monitoring when gas is only injected into the annulus using a parasite tubing or temporary
casing string.
MWD systems, with downhole (annular) pressure recording capability, do exist.25 These can be
interrogated at the surface, after a trip, to recover a record of downhole pressures seen during the
previous run. While this information can help adjust operating parameters for the next run, it
cannot be used for real-time control and may not be sufficient to avoid operating overbalanced
for some period of time.
Electromagnetic MWD can transmit downhole information to the surface, even when the
drillstring is filled with compressible fluid. There are some limitations to the reliability and
transmission capabilities of EMWD systems. They will, however, often operate satisfactorily in
many gasified liquid drilling applications.
2-188

Chapter 2

Underbalanced Drilling Techniques

One commercially available EMWD system does provide real-time indication of the annular
pressure,26 and others are under development.25
Several surface instrumentation packages have been developed, specifically for gasified liquid
underbalanced drilling with a closed surface system.25,26,27 These typically measure gas and
liquid injection pressures and rates, return fluid composition, pressure and rate, surface liquid
volume, measured depth, wellbore geometry and bit position. Other drilling data, such as hook
load, weight on bit, and rotary speed, may be included. From these measurements, it is possible
to compute the drilling fluid pressure, at any point in the well, using an hydraulics simulator.
Other outputs are often available, including the flow regime (laminar, turbulent), velocities of gas
and liquid, gas density, shear rates, viscosities, frictional and hydrostatic pressure gradients,
bottoms-up time (for sample depth location and for indicating system response time), and liquid
volumes.
These computed outputs can provide valuable information on the wells condition and guidance
for drilling operations. For example, the bottomhole pressure and pressure gradients can be used
to adjust the injection rates, to maintain a desired degree of underbalance. Figure 2-51 is one
example where bottomhole pressures indicated a progressive increase in annular pressure at the
bit and were used to adjust the injection rates to stay underbalanced.

1000
900

Bottomhole Pressure (psi)

800
700
600

Water: 57 gpm
Nitrogen: 950 scfm

Water: 40 gpm
Nitrogen: 950 scfm

500
Bottomhole Pressure at the Casing Shoe

400

Bottomhole Pressure at the Bit


Formation Pressure

300
200
100
0
0

500

1000

1500

2000

2500

Openhole Length (feet)

Figure 2-51.

2-189

Bottomhole pressure information, used to adjust injection rates, in order to stay


underbalanced (after Wilson, 1995 26).

Formation fluid inflows can also be detected and quantified.25,26 A qualitative indication of
permeability is possible, using the production data and an estimated borehole pressure. Water
producing zones can be identified. Changing gas production rates have been interpreted to
indicate the proximity to gas/oil contacts. There have been a number of horizontal wells in
Canada in which non-productive zones have been identified while drilling, leading to changes in
the planned trajectory away from regions of poor productivity.25 Wilson, 1995,26 cited an
instance where real-time analysis of production data, while still drilling, indicated that a
horizontal well had reached its economic productivity limit. This justified terminating the well
approximately 2,000 feet short of the originally planned horizontal length.
Determining annular pressure is comparatively straightforward, when gas is injected down a
parasite tubing or a temporary casing string.27 The flow geometry and the injected gass
properties will be well characterized. Measurements of the gas injection rate and pressure can be
used to accurately compute the pressure at the bottom of the injection string. The hydrostatic
head and the frictional pressure loss of the injected liquid (between this point and the hole
bottom) can also be predicted with reasonable confidence. Combining these with the annular
pressure at the gas injection depth indicates the bottomhole pressure with an accuracy that is
often acceptable.
2.7.8

Operating Procedures

Controlling Bottomhole Pressures


When drilling ahead, bottomhole pressure is principally controlled by varying the gas injection
rate. This is done by adjusting the bypass choke (for air or natural gas), or by adjusting the
nitrogen pump rate (for liquid nitrogen).
In applications where aerated liquid is being used to limit lost circulation, the air rate will often
be adjusted with reference to the pit volume. If the pit volume decreases, indicating losses are
occurring, the air rate will be increased to reduce the bottomhole pressure. Conversely, if the pit
volume increases, a liquid inflow is occurring, and the air injection rate will be decreased in
order to increase the bottomhole pressure and stop the inflow.
In most Canadian underbalanced drilling operations, gasified liquid is used to prevent formation
damage when drilling through a producing zone. The economic penalty for failing to maintain
underbalanced conditions is much greater when drilling through a reservoir which is prone to
formation damage, than when drilling other formations. Greater care is necessary to avoid
overbalanced conditions, and the benefit of real-time indication of downhole pressure is clear.
Since the downhole flow regime in these wells will usually be friction-controlled, the bottomhole
pressure will not normally decrease if the gas injection rate is increased. Instead, the liquid and
possibly the gas injection rate will have to be reduced in order to decrease the bottomhole
pressure. Determining new gas and liquid injection rates, to give a specific pressure, is not
trivial, and must be done using a flow simulator.
Since the circulating system has considerable compressibility, any adjustment to the injection
rates will cause the circulating pressures to respond and stabilize over a finite period of time.
2-190

Chapter 2

Underbalanced Drilling Techniques

There is a danger of over-controlling under these circumstances. It will be much easier to


achieve the desired pressure levels if new injection rates are first computed and used for guidance
in making the rate adjustments.
Connections
Connections can be slow when drilling with drillstring gas injection.
 Gas injection should be stopped.
 The liquid injection pumps should be then shut down.
 The pressure in the drillstring has to be bled down through the standpipe bypass line before
the joint can be broken open. This takes longer with gasified liquid than with dry air. Even
with an upper string float, it can take 5 to 15 minutes.7 Breaking open the joint before the
pressure has fully bled down should be avoided.
 If there is no check valve in the gas supply line, the gas supply should be shut off at the
standpipe manifold. The pressure in the supply line should be kept close to its operating
level, to avoid any possibility of liquid running into the gas supply equipment when injection
is resumed.10
 Making a connection can be speeded up by displacing the string as far as the upper float
valve, either to gas or to liquid. If the string is displaced to gas, the pressure can be bled
down rapidly and liquid should stay below the float valve. If liquid is completely displaced
from the string, liquid will be discharged at the floor when the joint is broken open.
Displacing to liquid is the quickest method of making a connection. The higher fluid density
in the string causes it to flow (U-tube) into the annulus, and the rig crew does not need to
wait for pressure to bleed down. This puts a slug of liquid into the circulating system that will
cause a pressure spike downhole when injection is resumed; possibly causing balanced or
overbalanced conditions.
Benion et al., 1995,28 demonstrated the variation of downhole pressure for different connection
methods (refer to Figure 2-52). There are larger pressure peaks when the string is displaced to
liquid, than when it is displaced to gas prior to the connection.
The drillstring does not have to be displaced to gas or liquid before a connection when annular
gas injection is used. Gas will migrate up the annulus and liquid will flow from the string into
the annulus. The liquid flowing into the annulus can be prevented from increasing bottomhole
pressure by continuing gas injection after the pumps are shut off. However, the annulus will
unload during the connection, whether or not injection is continued.
If connection times are short and the gas-to-liquid ratio is small, it is sometimes possible to keep
the annulus from unloading by stopping gas injection and shutting-in the BOPs.22 Extreme
caution should be used when doing this. The annulus pressure may build rapidly due to the
difference in drillstring and annular fluid densities and the migration of gas up the annulus. The
compressibility of the gas now trapped in the annulus, between the lower string float and the
BOP, will limit the pressure increase to less than the difference in the hydrostatic pressures of the
2-191

liquid in the drillstring and the gasified fluid in the annulus. Shutting-in the BOPs can cause
liquid to flow into the formation because pressures are greater than during circulation. The
pressure difference is larger at the casing seat than at the hole bottom. Applying surface pressure
to a gasified liquid gives a higher equivalent mud density at shallower depths than at greater
depths. Finally, the surface equipment will be subjected to very high instantaneous flow rates
when the BOPs are opened, unless the well is first flowed through the rigs chokeline (bleed off
line) to release built-up annular pressure.

Hydrostatic Pressure

Pressure

Reservoir Pressure

Pressure Spikes During Connections

N 2 Circulation Prior to Connections

Time

Figure 2-52.

The impact of the connection procedure on bottomhole pressure (modified after


Bennion et al., 1995 28).

Saponja, 1995,19 presented an example in which increasing the annular pressure actually sped up
making connections. The well was initially drilled with water, with nitrogen injection down the
drillstring. When injection was stopped to make a connection, a large oil influx killed the well. It
was not possible to regain circulation until the oil had been pushed back into the formation with
nitrogen pressure on the annulus. The connection took 45 minutes. An alternative procedure,
involving annular pre-charging, was devised. The annulus was charged with nitrogen, at
approximately 1950 scfm, for one or two minutes, to increase the annular pressure. This
restricted inflow. Circulation was easily re-established after making the connection, and the total
connection time came down to seven minutes. While connection time may be reduced,
formation damage may overwhelm any benefits, particularly if a water-based liquid is used. In
this example, the injected liquid was switched to oil before trying this annular pre-charging; there
was no detectable formation damage.

2-192

Chapter 2

Underbalanced Drilling Techniques

Trips
Tripping procedures depend on the extent and nature of any formation fluid inflow. Tripping is
easy if the well will not flow to surface, particularly if the well is not making any hydrocarbons.
First gas and then liquid injection should be shut-down. If there is uncertainty about lost
circulation and no potentially productive intervals are open to the wellbore, liquid injection can
be continued after the gas is shut-down, to determine if there are full liquid returns. The string is
stripped back through the rotating head or RBOP, as far as the BHA. At this point, the annulus
should have stopped unloading and the well should be dead. The rotating head or RBOP rubber
can now be removed and the trip completed. In general, it will not be possible to jet any produced
gas away from the rig floor, once the rotating head or RBOP rubber is pulled. This tripping
procedure cannot be used if the well is making any gas. If the well will not support a full
column of liquid, hole fill-up volumes cannot be used to confirm if the well is flowing.
With drillstring gas injection, establishing circulation, after a trip, may be difficult if there is lost
circulation. Many lost circulation zones can also produce substantial water volumes. Gas will
break out of the drilling fluid when the well is not circulated, leaving gas between the surface and
the top of the liquid column. After a short time, this liquid column will contain little entrained
gas and the hydrostatic pressure at the lost circulation zone will approach the formation pressure.
When gas and liquid injection are resumed, the drilling fluid will flow up the annulus to the lost
circulation zone, displacing the standing liquid uphole and increasing the pressure adjacent to the
loss zone. If the loss zone has significant permeability, which it usually will, much or all of the
gasified liquid can flow into the loss zone, rather than up the annulus.
It may be necessary to stage the drillstring into the hole, breaking circulation several times before
reaching bottom. Breaking circulation above the loss zone will gasify the liquid column in the
annulus and reduce the hydrostatic pressure sufficiently that circulation can be established from
below the loss zone.
With annular gas injection, it is possible to inject gas and gasify the liquid standing in the
annulus above the injection point, before any liquid is pumped into the well. If this does not
reduce the annular pressure enough for circulation to be established, the gas supply system can,
in principle at least, be re-configured so that gas can also be injected down the drillstring to
further reduce the annular pressure.
Tripping becomes more complex if the well is producing hydrocarbons or will flow to the
surface. It is possible to trip, as described above, if the well is making only oil with a very low
GOR, from a sufficiently underpressured zone that the well will load up and kill itself before oil
reaches the surface. In most cases, some oil will reach surface and other options need to be
considered. There are also concerns that repeated flow-cycles can lead to formation damage.2
If a well does flow to surface, the options are to strip out of the hole while allowing the well to
flow, or to kill the well before tripping. When feasible, stripping out of the hole is preferable,
since this should preserve the underbalance and avoid formation damage while tripping.
2-193

Drillpipe can be stripped through the rotating head or RBOP, while the well is allowed to flow
through the flow line. A rotating head or a normally configured RBOP will not seal on larger
diameter drillstring components. Cylindrical collars and other slick BHA components can be
stripped through the annular. If essential, it is possible to remove the kelly packer rubber from an
RBOP and to allow the inner packer element to seal around larger diameter string components;29
but, this practice is not recommended. It may be possible to stage irregularly shaped BHA
components, such as stabilizers, out of the well through an RBOP. To do this they have to be
small enough in diameter to pass through the RBOP sealing element29 and short enough for the
irregularly shaped portion to fit between the annular and the RBOP. In this case, the component
in question is stripped through the RBOP until the irregularity is above the annular. The
annular is shut around the string and the RBOP is opened to allow the irregularity to pass.
When stripping through the annular, the choke line below the blind rams should be opened and
the well flowed through an open choke to the separator system or the flare/reserve pit, as
appropriate. The blind rams should be closed below the bit before finally opening the annular
and removing the last stand.
Flowing the well through an open choke keeps the wellhead pressure to a minimum.
Nevertheless, it is almost inevitable that a pipe light situation will occur before the string is out
of the hole. This occurs when the pressure beneath the drillstring exerts an upward force on the
string that exceeds its effective weight (the sum of the buoyant weight of string below the
pressure seal and the weight in air above it). Without any restraint, the string is then forced out
of the well by the pressure beneath it. Snubbing calculations should be performed before
stripping the string from the well to determine when pipe light conditions will occur. These
should consider the probable surface pressure, the weight of the drillstring above the pressure
seal, its buoyant weight below the seal, the area of the string where it passes through the pressure
seal, and the frictional force of the sealing element on the string. When a pipe light situation
does occur, the string can sometimes be kept in the hole by the frictional force of the RBOP
sealing element. The frictional force on the string can be augmented by closing the annular.29
Increasing the closing pressure on the seal elements will increase the frictional force they exert
on the string. It is also possible to use reverse-acting slips. With high wellhead pressures, it may
be necessary to use a snubbing unit to control the string as it is withdrawn from the well.
With a BHA consisting of collars, the weight of the first stand may be sufficient to overcome
both the pressure thrust due to the flowing wells backpressure and the seal frictional force so
that it can enter the flowing well without additional assistance. The bit should be run just above
the closed blind rams, the RBOP (or annular) closed and the blind rams opened, before running
the string into the hole. It is possible to use a large diameter collar above drillpipe to push the
drillpipe into the well if the pipes weight alone is insufficient. A simple crossover between the
collar and the pipe can be difficult to handle, and a special connection tool has been used
instead.29 This resembled a pipe elevator, with a collar thread box on top. It was made up onto
the bottom of the collar and a single joint of drillpipe was picked up using the elevator portion.

2-194

Chapter 2

Underbalanced Drilling Techniques

This would be made up to the pipe held in the RBOP and the RBOP closing pressure reduced to
allow the string to run into the hole. Again, a snubbing unit may be required if the wellhead
pressure is still too high.
This procedure requires that the surface system can handle the unrestricted flow rate of the well.
The surface system should normally have been sized to do this. Well flow can be choked back, if
necessary, to restrict the flow rate. This will increase the pressure at the wellhead. If the well
flow has to be choked back, the pressure limitations of the BOP stack, and particularly of the
rotating head or RBOP, if these are being used to contain the annular pressure, must be
recognized. It may be necessary to kill the well if the flow cannot be handled safely.
Higher wellhead pressures will increase the snubbing loads. It may be necessary to use a
snubbing unit to control the string. Hydraulic Rig Assist (HRA) snubbing units have been
specifically developed to supplement drilling operations with a conventional drilling or workover
rig.30 A truck-transportable unit is shown in Figure 2-53. It incorporates a BOP system, a pair of
hydraulic rams to move the string, stationary and traveling slips to hold the string, a working
platform and a control panel. Both the stationary and the lower set of traveling slips hold the
string down against wellbore pressure when pipe light. The upper set of slips, the heavy slips,
hold the strings weight when necessary. The equalize/bleedoff spool, between the annular and
ram-type BOP units, is connected to the annulus below the blind rams and to the choke manifold.
It has hydraulically activated valves that are controlled from the work platform. It allows
pressures beneath different BOP components to be equalized (for example during staging
operations) or to be bled down as required.
The lower BOP is attached either directly to the drilling rigs BOP stack or to a suitably long
riser spool so that the snubbing unit is just above the drilling rigs floor. Figure 2-53 shows one
rig-up that has been used for drilling underbalanced with a conventional drilling rig and an
RBOP, with the HRA snubbing units BOPs above the rig floor and the rigs BOP stack below.
The BOPs are usually rated to 5,000 psi, and can often pass equipment up to about six inches in
diameter. This effectively limits these units to drilling operations inside seven-inch diameter or
smaller casing. A typical BOP stack for drilling underbalanced with a conventional drilling rig
and an HRA snubbing unit is shown in Figure 2-54.
The hydraulic cylinders, shown in Figure 2-53, provide maximum snubbing forces that typically
range from 45,000 to 80,000 lbf. These control the string as it is removed from the well and force
it back into the well against the wellbore pressure. Most hoisting operations are performed with
the rigs drawworks. Drillstring and tubing are racked back in its derrick when necessary.

2-195

Control Panel
Heavy Slips
Traveling Plate
Traveling Slips

Work Platform
Stationary Slips
4.25m
Typically

Hydraulic Cylinders
Annular
BOP
Equalizing Spool
Hydraulically
Activated Equalize/
Bleedoff Plug Valves
Structural Frame
Single Gate
Ram BOP

Figure 2-53.

An Hydraulic Rig Assist (HRA) snubbing unit (after MacPherson and Goertzen, 1995
).

30

Annular
BOP

Equalize/Bleedoff
Plug Valve
Bleedoff Line

Single Gate
Ram BOP

Riser Spool
Rig Floor
R BOP

Crossover Spool
Equalize Line
Annular
BOP

Double Gate
BOP

To 3 Stage Separator
Kill Line

Casing Bowl

Figure 2-54.

A typical BOP stack, for drilling underbalanced with a conventional drilling rig and
an HRA snubbing unit (after MacPherson and Goertzen, 1995 30).

2-196

Chapter 2

Underbalanced Drilling Techniques

Killing a well is not an attractive option if the well is being drilled underbalanced to avoid
formation damage. It may become necessary if the BHA configuration cannot be tripped from
the well without losing the pressure seal, if the string cannot easily be run into or out of the hole
in any pipe light state that can arise, or if the wells flow cannot safely be handled by the surface
system. The kill fluid should be chosen to minimize formation damage. Selecting nominally
non-damaging fluids is beyond the scope of this document. It will not be discussed in any detail.
An introductory discussion of the various factors involved can be found in many references; for
example, Ali et al., 1994.31 A clear brine is one possibility if sufficient density can be achieved
without undue cost or health and environmental concerns. Alternatively, liquids, to which
suitably sized calcium carbonate particles have been added, can be considered. The particles
should be sized to minimize fluid loss, ideally by conducting pilot tests on core samples from the
producing formation. Spotting an appropriate acid, after drilling, should remove the majority of
calcium carbonate particles and restore the permeability.
If possible, the kill fluid should be placed on top of produced crude oil, to reduce the probability
of contact between the formation and the kill fluid. To do this, the well should be allowed to
flow until all open productive formations are covered. The string should be stripped back above
the produced liquid, ideally inside the casing shoe. Kill fluid can then be spotted above the
produced liquid to kill the well,2 before continuing the trip. Alternatively, the kill fluid can be
pumped down the annulus while continuing to strip the pipe. In this case, the kill fluid should be
pumped at a rate equal to that at which the open displaced volume and internal volume of the
drillpipe are removed from the hole.32 The kill fluid density should be chosen so that the well is
dead before the top of the BHA reaches the surface. Whatever procedure is used to place the kill
fluid, conventional tripping practices, trip sheets, hole-fill, etc., should be followed once the well
is killed. As a warning, the higher density kill fluid and the lower density produced hydrocarbon
may flip over, due to their density differences, allowing the kill fluid to contact and potentially
damage the formation.2
Gas Inflows
When a gas inflow is detected, it may be possible to reduce the gas injection rate and cost.
However, if drilling continues until the BHA is below the inflow, the reduced annular velocity
close to the hole bottom may not be sufficient for good cuttings transport (if the gas injection rate
has been reduced substantially).
2.7.9 Limitations
Recent experience in Canada has shown that gasified liquids can be used to successfully drill
underbalanced, in a wide range of normally and sub-hydrostatically pressured reservoirs. There
are factors which limit the application of gasified liquids, including high formation pressure or
productivity, wellbore instability, inadequate pressure control, excessive produced water,
corrosion, and limited penetration rates.

2-197

Formation Pressure and Productivity


Gasified liquids give higher borehole pressures, in comparison to dry gases or other lightened
drilling fluids. Because of this, they can be used to drill higher pressure and productivity
formations, before the pressure and rate of hydrocarbon production while drilling approach the
limitations of the surface equipment. The limiting factors are the pressure capacity of the surface
equipment, particularly the BOP stack and diverter system and the flow rate capacity of the
surface separator system. The flow capacity of a suitably designed surface system can be very
high - natural gas production rates of as much as 40 MMcf/D have been handled safely in
Canadian underbalanced drilling operations.
Assessing formation productivity and pore pressure, along with the capacity of the surface
system, can indicate the largest drawdown that would be tolerable during drilling of a particular
interval. This effectively sets a restriction on the borehole pressure, below which it is unsafe to
drill. There are many overpressured reservoirs where the difference between borehole and
formation pressures would be too great for gasified liquids to be used safely, and, a single-phase
liquid drilling fluid with an appropriate density, could be used to generate underbalanced
conditions. Section 2.8, Flowdrilling, describes these oper-ations.
Wellbore Instability
Gasified liquids can be used in weaker, less competent rocks than dry gases and other lightened
drilling fluids. Their higher wellbore pressures reduce the effective stresses adjacent to the
borehole wall. Many horizontal wells have been drilled with gasified liquids in weak rocks.
Even so, borehole pressures will be lower with a gasified liquid than they would be with a
drilling mud and mechanically-induced wellbore instability remains more likely. Large
fluctuations in borehole pressure, that often occur when gasified liquids are used, may aggravate
wellbore instability, by disturbing rock fragments that might otherwise have remained more or
less stable under static loading.
Wellbore instability can be more of a problem when drilling through shales than through
sandstones or carbonates. For example, when drilling towards a depleted reservoir, overlying
shales may be overpressured, relative to the target formation. This increases the underbalance
pressure through the shale and makes them more prone to mechanical instability. Wellbore
instability provides another limit on how low the borehole pressure may be taken.
Water-sensitive shales can also swell and slough when exposed to aqueous liquids. As the liquid
phase of the drilling fluid is usually re-cycled, it may be economical to limit shale instability by
adding appropriate salt and/or polymer to a water-based liquid. Using hydrocarbon-based liquid
may eliminate concerns about chemically-induced wellbore instability in shales.
Pressure Control
In general, downhole pressures do not remain constant when drilling with gasified liquids. They
tend to fluctuate, particularly when tripping or making connections. Peak-to-peak pressure
2-55 shows
fluctuations of 500 psi are commonly recorded downhole.7,25 The example in Figure
Figure 2-55
significant downhole pressure fluctuations above and below an average value.
2-198

Chapter 2

Underbalanced Drilling Techniques

These pressure fluctuations will usually not cause major problems when drilling underbalanced
to avoid lost circulation, except possibly when re-establishing circulation. When drilling
underbalanced to avoid formation damage, the upper tolerable limit to borehole pressure is given
by the formation pressure. If the borehole pressure increases above the formation pressure, then
damage is probable. A lower tolerable borehole pressure limit will be set by productivity,
wellbore stability and the surface facilities. There can be great difficulty in drilling
underbalanced with a gasified liquid if this lower pressure limit is not at least 500 psi lower than
the formation pressure.
1000

Bottomhole Pressure (psi)

900

800
Large Liquid Slug at
Start of Circulation

700

600
Slug Flow Regime

500

174 psi Annular Friction

400
Drill String Connection

300
0

20

40

60

80

100

120

140

Time (minutes)

Figure 2-55.Bottomhole pressures, recorded while drilling with gasified liquid (after Saponja,1995 7).

Produced Water
Produced water can be a problem when gasified liquid is used to avoid lost circulation. The
equivalent circulating density changes with depth when drilling with a gasified liquid. The gas
volume fraction decreases with increasing depth due to the higher pressure, and this increases the
mixtures density. Depending on the variation of formation pressure with depth, this could mean
that a well can potentially be overbalanced near the hole bottom but underbalanced higher up
hole. Under these circumstances, it is possible to lose circulation at the same time that a zone
uphole is flowing into the well. It may be very difficult to achieve circulation if the zone uphole
is producing large quantities of water.

2-199

There have been instances in which uphole water inflows have kept the pressure in the wellbore
above the formation pressure in a lower loss zone, so that it was not possible to circulate any
fluid to surface even when injecting only dry air down the drilling string. In this case, the options
are to drill ahead without returns, to run casing or to abandon the well.
It will often be possible to maintain circulation with water flowing into the well, provided that
the inflow is not too rapid. If drilling continues for very long after the inflow starts, water
disposal can easily become a problem. If the well is making fresh water, surface discharge might
be approved. Otherwise, the produced water will usually have to be hauled away for disposal, off
site. This can be expensive. If it is an in-fill well, it may be possible to tie into an existing
disposal system.
Some wells will produce water at low circulating pressures, and then stop producing water and
instead lose circulation, if the borehole pressure is increased. In this situation, it may be possible
to re-inject the water downhole. This has been done in coalbed methane wells, in the Black
Warrior Basin.33 These were drilled with dry air until water production occurred and filled the
reserve pit. The drilling fluid would then be switched from dry air to produced water. Drilling
continued with total losses, until the reserve pit was emptied, at which time air drilling was
resumed. By repeating this process it was possible to reach TD with minimal amounts of
produced water for off-site disposal. Drilling without returns may not be possible in all wells.
Sloughing above the lost circulation zone can cause the pipe to become stuck. The potential for
wellbore instability and stuck pipe should be carefully considered before adopting this option.
Produced water entering the well cannot be separated from a water-based gasified liquid.
Gasified mud should not be used when large volumes of produced water are expected. The
produced water will dilute the mud, requiring continuous and expensive replenishment of mud
additives to maintain the liquids properties. If a mud has to be used to clean the hole, it should
be designed to minimize the cost of the consumables that will have to be added to counter
dilution by produced water. If water production is encountered when drilling with a gasified
mud, the borehole pressure should be increased, if possible (without risking formation damage or
losing circulation), in order to reduce the water inflow rate.
Corrosion
Corrosion of the drillstring and other downhole equipment, including any exposed casing in the
well, can be a very severe problem when drilling with aerated water or water-based liquids. The
rate of corrosion increases as the conductivity of the liquid phase increases. Using a saline liquid
phase, such as produced water, or adding salts to inhibit shale hydration, will increase corrosion
problems, as will high downhole temperatures and the presence of even very small quantities of
hydrogen sulfide.
Corrosion inhibitors have to be used with aerated, water-based drilling fluids. These are
expensive and can add considerably to well costs, particularly if water production dilutes the
circulated liquid.

2-200

Chapter 2

Underbalanced Drilling Techniques

Furthermore, some corrosion inhibitors become ineffective once the temperature exceeds about
100F and may even accelerate corrosion at temperatures of 300F or more.34 The injected
liquids pH should be controlled to be above about nine, if possible, since this will tend to slow
corrosion. Corrosion coupons should be used to monitor corrosion rates and to indicate if
protective measures are adequate. The cost of corrosion control when drilling with an aerated,
water-based fluid (if the downhole temperature is high and if the well is making high salinity
water) may be prohibitive.
Corrosion remains a problem if membrane-generated nitrogen or mixed air and nitrogen are used
to gasify the drilling fluid. It is not a concern when using natural gas, cryogenic nitrogen, or a
hydrocarbon-based liquid phase.
If there is any prospect of hydrogen sulfide production, all exposed equipment should be rated for
sour service.
Penetration Rates
The penetration rates, with a gasified liquid, will generally be lower than those for dry gas, mist
or foam. This is because of the higher borehole pressures. Because the borehole pressure is
lower than for a mud, the penetration rates will normally still be higher than for a mud. The
maximum gas delivery pressure is usually less than the maximum mud pump pressure.
Therefore, the standpipe pressures cannot be as high when injecting gas down the drillstring as
when drilling with mud. This limits the bit pressure drop that can be used with the gasified
liquid, and this can in turn restrict any penetration rate increase that might accompany the use of
the gasified liquid.
Gasified liquids are often used in weaker formations instead of other lightened drilling fluids. In
these formations, penetration rates can be high even when drilling overbalanced. The economic
benefits of penetration rate increases are less at these high penetration rates, when only a small
fraction of the total rig time is actually spent with the bit on bottom making hole. It is possible
that the time savings due to a penetration rate increase cannot cover the cost of the extra
equipment involved; drilling underbalanced with a gasified liquid may actually cost more than
conventional drilling. In this case, other benefits of drilling underbalanced have to be sufficient
to justify its use. This will often be the case.

2-201

References
1.

Bobo, R.A. and Barrett, H.M.: Aeration of Drilling Fluids, World Oil (1953) 137, No. 4,
145.

2.

Bieseman, T. and Emeh, V.: An Introduction to Underbalanced Drilling, paper presented


at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

3.

Eide, E., Brinkhorst, J., Volker, H., Burge, P. and Ewen, R.: Further Advances in Coiled
Tubing Drilling, paper SPE 28866, JPT (May 1995).

4.

Wodka, P., Tirsgaard, H. and Damgaard A.P.: Underbalanced Coiled Tubing Drilled
Horizontal Well in the North Sea, paper SPE/IADC 29359 presented at the 1995
SPE/IADC Drilling Conference, Amsterdam.

5.

Teichrob, R.R.: Low Pressure Reservoir Drilled with Air/N2 in a Closed System, Oil and
Gas J. (March 21, 1994) 80-90.

6.

Bennion, D.B. and Thomas, F.B.: Underbalanced Drilling of Horizontal Wells: Does It
Really Eliminate Formation Damage?, paper SPE 27352 presented at the 1994 SPE
International Symposium on Formation Damage Control, Lafayette, LA.

7.

Saponja, J:. Engineering Considerations for Jointed Pipe Underbalanced Drilling, paper
presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

8.

Wolke, R.M., Jardiolin, R.A., Suter, R.L., Moriyama, S., Sueyoshi, Y. and Kihara, Y.:
Aerated Drilling Fluids can Lower Drilling Costs and Minimize Formation Damage,
Geothermal Resources Council Bulletin (May 1990) 131-137.

9.

Claytor, S.B., Manning, K.J. and Schmalzried, D.L.: Drilling a Medium-Radius


Horizontal Well with Aerated Drilling Fluid: A Case Study, paper SPE/IADC 21988
presented at the 1991 SPE/IADC Drilling Conference, Amsterdam, The Netherlands.

10.

Curtis, F. and Lunan, B.: Underbalanced Drilling Operations: Correct Operating


Procedures Using a Closed Surface Control System to Drill for Oil and Gas, paper
presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

11.

Fried, S. and McDonald, C.: Nitrogen Supply Alternatives for Underbalanced Drilling,
paper presented at the 1995 1st International Underbalanced Drilling Conference, The Hague,
The Netherlands, October 2-4.

2-202

Chapter 2

2-203

Underbalanced Drilling Techniques

12.

Guo, B., Hareland, G. and Rajtar, J.: Computer Simulation Predicts Unfavorable Mud
Rate and Optimum Air Injection Rate for Aerated Mud Drilling, paper SPE 26892
presented at the 1993 SPE Eastern Regional Conference and Exhibition, Pittsburgh, PA.

13.

Hasan, A.R. and Kabir, C.S.: A Study of Multiphase Flow Behavior in Vertical Wells,
SPEPE (May 1988) 263-272.

14.

Poettmann, F.H. and Bergman, W.E.: Density of Drilling Muds Reduced by Air
Injection, World Oil (August 1955) 97-100.

15.

Poettmann, F.H. and Carpenter, P.G.: Drilling and Production Practice, API (1952) 257.

16.

Wang, Z., Rommerveit, R., Vefring, E.H., Bieseman, T. and Faure, A.M.: A Dynamic
Underbalanced Drilling Simulator, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

17.

Moore, P.L.: Drilling Practices Manual, PennWell Books, Tulsa, OK (1974) 234.

18.

Gray, K.E.: The Cutting Carrying Capacity of Air at Pressures Above Atmospheric, Pet.
Trans. AIME (1958) 213, 180-185.

19.

Saponja, J.: Comparing Conventional Mud to Underbalanced Drilling in a Depleted


Reservoir, paper presented at the 1995 1st International Underbalanced Drilling Conference,
The Hague, The Netherlands, October 2-4.

20.

Westermark, R.V.: Drilling with a Parasite Aerating String in the Disturbed Belt, Gallatin
County, Montana, paper IADC/SPE 14734 presented at the 1986 IADC/SPE Drilling
Conference, Dallas, TX.

21.

Comeau, L.: Underbalanced Drilling: Directional and MWD Experience, paper presented
at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

22.

Deis, P.V., Yurkiw, F.J. and Barrenechea, P.J.: The Development of an Underbalanced
Drilling Process: An Operators Experience in Western Canada, paper presented at the
1995 1st International Underbalanced Drilling Conference, The Hague, The Netherlands,
October 2-4.

23.

Hannigan, D.M. and Bourgoyne, A.T., Jr.: Underbalanced Drilling Rotating Control Head
Technology Increasing in Importance, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

24.

Shale, L.: Underbalanced Drilling Equipment and Techniques, presented at the 1995
ASME Energy Technology Conference, Houston, TX.

25.

Roy, R. and Hay, R.: Measuring Downhole Annular Pressure While Drilling for
Optimization of Underbalanced Drilling, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

26.

Wilson, J.: Optimizing Drilling of Underbalanced Wellbores with Data Acquisition


Systems, paper presented at the 1995 1st International Underbalanced Drilling Conference,
The Hague, The Netherlands, October 2-4.

27.

Taylor, J., McDonald, C. and Fried, S.: Underbalanced Drilling Total Systems Approach,
paper presented at the 1995 1st International Underbalanced Drilling Conference, The Hague,
The Netherlands, October 2-4.

28.

Bennion, D.B., Thomas, F.B., Bietz, R.F. and Bennion, D.W.: Underbalanced Drilling,
Praises and Perils, paper presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

29.

Cagnolatti, E. and Curtis, F.: Using Underbalanced Technology to Solve Traditional


Drilling Problems in Argentina, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

30.

MacPherson, L. and Goertzen, J.B.: Hydraulic Rig Assist Snubbing-Well Control for
Underbalanced Drilling, paper presented at the 1995 1st International Underbalanced
Drilling Conference, The Hague, The Netherlands, October 2-4.

31.

Ali, S., Burnett, D., McLeod, H., Peden, J. and Penberthy, W.L., Jr.: Experts Share Views
on Formation Damage Solutions, JPT (November 1994) 936-940.

32.

Joseph, R.A.: Underbalanced Horizontal Drilling - Conclusion Special Techniques and


Equipment Reduce Problems, Oil and Gas J. ( March 27, 1995) 41-47.

33.

Graves, S.L., Niederhofer, J.D. and Beavers, W.M.: A Combination Air and Fluid Drilling
Technique for Zones of Lost Circulation in the Black Warrior Basin, SPEDE (February
1986) 57-61.

34.

Scott, S.L., Wu, Y. and Bridges, T.J.: Air Foam Improves Efficiency of Completion and
Workover Operations in Low-Pressure Gas Wells, SPEDC, (December 1995) 219-225.

2-204

Chapter 2

2.8

Underbalanced Drilling Techniques

Flowdrilling

The term flowdrilling refers to drilling operations in which the well is allowed to flow to
surface while drilling. It was first coined by a group of Oryx Energy Company engineers (Stone,
unpublished), to describe the drilling technique that they had developed for the Austin chalk
reservoir in the Pearsall Field, in South Texas. Flowdrilling of this and other Austin chalk
reservoirs represents one of the most successful applications of underbalanced drilling to date in
the United States.1
The underbalanced drilling techniques, described previously in this manual, have all used drilling
fluids that are either completely gaseous or contain a gaseous phase. In the Austin chalk
application, however, underbalanced conditions were created with a liquid drilling fluid.
Although all underbalanced drilling is strictly flowdrilling, the term will be reserved here for
those underbalanced drilling operations that involve a liquid drilling fluid.
Flowdrilling occurs when a permeable formation is intentionally drilled with a drilling fluid that
encourages the formation to flow during drilling operations. Most commonly, the fluid influx
will be from a hydrocarbon-bearing formation, and the flow returning to surface will consist of
oil, natural gas and the drilling fluid. When flowdrilling, well control problems are handled at
the surface rather than downhole. Specific downhole and surface equipment are required for safe
and efficient flowdrilling operations.
Borehole pressures are necessarily higher for a liquid drilling fluid than for a gaseous or gascontaining drilling fluid. This allows flowdrilling to be successful in many applications where it
would not be technically or economically feasible to drill with other underbalanced techniques.
More productive formations, which may mean higher pore pressures, higher permeability
formations, or both, can be drilled underbalanced, before the limits of surface equipment are
reached. Higher borehole pressures may allow less competent formations to be drilled
underbalanced without serious wellbore instability. No gas supply system is required for
flowdrilling. This can lead to lower daily operating costs than other underbalanced drilling
techniques. Since gas is not flowed down the drillstring, conventional mud motors and MWD
units can be used for flowdrilling, making it particularly suited for drilling directional and
horizontal wells.
However, excessive productivity and wellbore instability prevent flowdrilling from being
appropriate for all oil and gas wells. These limitations will be discussed later in this section.
2.8.1 Creation of Underbalanced Conditions
In order for formation fluid to flow into the wellbore while drilling, the borehole pressure,
resulting from the hydrostatic pressure of the drilling fluid plus the frictional circulating pressure,
must be less than the pore pressure of the open formations. In other words, there must be a
pressure differential into the wellbore from the reservoir. The formation must also have
sufficient permeability, either matrix or fracture permeability, to adequately flow reservoir fluids.

2-205

There are two distinct situations where underbalanced conditions can be achieved with a liquid
drilling fluid. The most obvious of these occurs when the equivalent circulating density of the
drilling fluid is less than the pore pressure gradient of the formations being drilled. This is
directly analogous to the previously described underbalanced drilling techniques, that used
lightened drilling fluids to give a borehole pressure less than the pore pressure. There are many
localities where pore pressures are high enough that underbalanced conditions can be achieved in
this way, with a liquid drilling fluid.
In certain instances, it is possible to allow the formation to induce underbalanced conditions even
though the drilling fluid initially gives a higher hydrostatic pressure than the pore pressure. The
Pearsall Field is an excellent example of this.1 The native pore pressure gradient of this Austin
Chalk Formation is approximately 7.3 ppg. In this field, flowdrilling was achieved using a fresh
water system weighing 8.3 ppg! At first, this seems strange, yet it was here that horizontal
flowdrilling was developed. There had been significant oil production in the past, from those
vertical wells that had intersected one or more hydrocarbon-filled, natural fractures. The matrix
permeability is low. This history resulted in quite large variations in pore pressure, with some of
the previously-produced natural fractures having lower pressure than the rest of the reservoir.
The actual underbalanced condition occurred when the overbalanced fresh water system drilled
out from an impermeable formation into one of the low pressure, naturally fractured zones,
causing loss of circulation. This initiated a reduction in hydrostatic head in the wellbore and a
corresponding reduction in the hydrostatic drilling fluid pressure at all points downhole. This, in
turn, resulted in formation fluids flowing into the wellbore from permeable drilled intervals,
usually resulting in the introduction of oil and gas into the wellbore. These fluids have lower
density than the fresh water drilling fluid, so that they naturally lowered the hydrostatic pressure
of the wellbore and allowed more hydrocarbons to flow into the well. Thus, flowdrilling was
2-56
initiated. This process is represented in Figure 2-56.
Regardless of which method initiates flow, the flowdrilling process will perpetuate itself until the
hydrostatic pressure plus the equivalent circulating pressure equals the downhole pore
pressure(s). Once this occurs, influx of fluids into the wellbore from the formation will cease as
the two pressures equalize or as the hydrostatic head from the drilling fluid exceeds the
bottomhole producing pressure. Flowdrilling can also result in crossflow downhole between
higher pressured, permeable intervals and lower pressured fractures.
2.8.2 Drilling Fluids
The primary requirement for flowdrilling is that the wellbore pressures should be maintained
between a maximum pressure equal to the formation pressure and a minimum pressure dictated
by wellbore stability. The density of the drilling fluid should be chosen so that the circulating
pressure will normally be in this range.
The pressure limitation of the diverter should also be considered when selecting the drilling fluid
density. It is normally not advisable for the sum of the drilling fluid hydrostatic head and the
maximum tolerable annular pressure while drilling to be less than the formation pressure.2
2-206

Chapter 2

Underbalanced Drilling Techniques

Clear drill brine density less than or equal to 1.02


1.02g/cm
s/cc3

Oil, Gas, and Brine

9.5 ppg Brine

Pressure higher in HEEL of well


causing lost returns

Pressure lower in TOE of well


causes influx

Pore
PorePressure
Pressure==
3030
feet
20.9 psi
Mpaat@6234
1900m

Figure 2-56.

Flowdrilling a naturally fractured, horizontal well (courtesy of Signa Engineering


Corporation).

It should be remembered that, particularly when drilling for natural gas, the gas cut associated
with flowdrilling will reduce the density of the fluid in the annulus below that in the pits.
Tangedahl, 1995,3 noted that a 10 ppg mud can easily be reduced to an equivalent of 6 ppg in a
high gas/oil ratio well.
A major reason for drilling underbalanced is to avoid formation damage. It has to be recognized
that there may be periods when underbalanced conditions are not maintained. The drilling fluid
should damage the formation as little as possible. As a general rule, whenever possible, a clear
liquid should be used as the flowdrilling fluid.

2-207

The cost and health, safety and environmental concerns, associated with using heavy brines,
particularly those containing zinc bromide, will often make conventionally weighted muds more
attractive, if high densities are required.
Solids may be necessary in the drilling fluid to limit losses in those formations where lost
circulation can occur during flowdrilling. Experience in the Pearsall Field indicates that drilled
solids may be effective in limiting fluid losses.
Viscosifying the drilling fluid should be avoided if possible. It can lead to surface separation
problems once hydrocarbon production begins, by slowing the rate at which hydrocarbons can
separate from the drilling fluid. Many viscosifying polymers promote the formation of
emulsions, and these can increase separation problems.3 So too can lubricant additives that may
be required to control drillstring torque and drag with clear drilling fluids. Bentonite
concentrations should be restricted, if there is any possibility of the mud being lost into fractures
from which it is hoped to produce. Otherwise, and particularly if downhole temperatures are
high, the mud may gel up in the fractures and not flow back when the well is brought into
production.2
It will often be possible to achieve adequate hole cleaning with an unviscosified drilling fluid,
simply by keeping the annular flow turbulent and the annular velocity high enough. When
flowdrilling is established, that is once formation fluids are flowing into the wellbore, the annular
velocity will increase and this will help transport cuttings uphole. Calculations of terminal
velocity and mechanical energy balance equations are used to determine the annular velocity
needed for adequate hole cleaning. Generally, higher velocities are needed in deviated and
horizontal wells than in vertical wells.
With large hole sizes or extensive lateral sections, good hole cleaning may not be possible,
without viscosifying the drilling fluid. This may also be necessary in order to keep the surface
pressure within tolerable limits (by restricting migration of formation gas up the annulus).
A final point to consider when selecting the drilling fluid is the solubility of formation gas
downhole. If produced gas is in solution downhole, there may be high pressure surges when it
comes out of solution at the surface. This may exclude the use of an oil-based mud in a well
where gas production is anticipated.2
2.8.3 Surface Equipment
Producing oil and gas while drilling requires integrating drilling and production systems safely
and efficiently. When flowdrilling techniques were being developed, this was accomplished by
installing a rotating head on top of a conventional BOP stack with a high pressure choke
manifold, including a hydraulically-operated control valve (HCR) with a surface equipment
system for separating drilling fluids, cuttings, oil and gas. The rotating head diverts return fluids,
under pressure, through the surface system, while drilling continues.

2-208

Chapter 2

Underbalanced Drilling Techniques

The approximate maximum pressure capability of the conventional rotating head was 400 psi.
Introduction of higher operating pressure rotating blowout preventers (RBOP) has increased the
pressure rating to about 1,500 psi while drilling and 1,000 psi while stripping, providing a safer
operating margin. It is this level of underbalance which determines the amount of formation
influx that can be allowed. On drilling into a permeable or fractured zone, the driller diverts the
return fluids from the rig pit to a mud/gas separator with a hydrostatically-maintained fluid level.
The gas is separated from the total return stream, transferred to a remote flare system and burned,
while the fluids and cuttings are routed to special separation pits. There, the cuttings are
separated from this remaining stream, and, finally, drilling fluids are separated from any liquid
hydrocarbons. These drilling fluids are sent back to the rig pits, while all produced oil or
condensate is sent to remote storage. Figure 2-57 is a schematic of the surface equipment used
when flowdrilling.
Although surface equipment is important in any type of drilling, it is crucial to the success and
safety of flowdrilling.
An open, or atmospheric, system is commonly used in most flowdrilling applications. The Texas
Atmospheric System (TAS) is the most economical method of surface fluid separation, while
flowdrilling. The components of this system are illustrated in Figure 2-58. In very high gas
content formations or if hydrogen sulfide gas might be encountered, a Closed Loop, or pressured,
Separator system (CLS) should be used upstream of a vacuum degasser, for improved gas
management and safety. The CLS can cost about four times more than the TAS because of its
additional manpower requirements and the expense of having a nitrogen source available to
supply operating pressure for the separator system before the onset of natural gas production.
BOP Stack
A typical flowdrilling blowout preventer stack consists of a rotating head or a rotating blowout
preventer (RBOP), installed above the conventional rig blowout preventers (refer to Figure 2-59).
A single or double-ported drilling spool separates the two. One of these outlets leads to the
possum belly and shale shaker, for non-flow operations. The other outlet leads to the choke
manifold, for flowdrilling operations. Below this spool, the normal preventer stack is installed. It
consists of an annular, or Hydril, preventer flanged to a double ram-type BOP. This unit includes
a set of pipe rams above a set of blind rams. A second, double-ported drilling spool is flanged
next in the system. One outlet on this spool is used for the choke line and the other serves as the
kill line. Ideally, an additional, lower set of pipe rams is installed below this spool.
All components of the BOP stack should be tested to their rated working pressures prior to any
drilling operations. The drill crew and rig depend on the protection and isolation from formation
pressures that the rotating head or the RBOP can provide during drilling and stripping. The
pressure rating of the BOP stack should be selected to accommodate the worst case, or highest
possible surface pressure expected, in order to provide the maximum degree of drilling safety.

2-209

Higher pressure-rated units are now available.

In addition to the stack itself, ensure that the total choke system (choke and kill lines, gate or ball
valves and check valves, HCR valves and all manifold piping) is properly rated, tested and in
good working condition. A typical choke line should consist of a ball valve, an HCR valve and a
check valve allowing flow toward the choke manifold.

2-202

R BO P
MUD PITS

STACK
GAS/FLUID
SEPARATION
SYSTEM

CHEMICAL
INJECTION

UNDERBALANCE
DRILLING
MANIFOLD

57
Figure 2-56.

Schematic of surface equipment required for flowdrilling (courtesy of Signa


Engineering Corporation).
12 in. Flare
6 in. Flare

4-6 in.

4 in. Flare
Gas boot (Open on bottom)

Water to rig
Grade
Gas
Separater

Choke manifold

Gas
Separater

RBOP

Skimmer tanks

RBOP
AnnularPreventer
preventer
Annular
Pipe Rams

Oil tank

Blind Rams
Pipe Rams

Figure 2-57.
58

Oil to treatment
off
oil location

Atmospheric surface system for flowdrilling (courtesy of Signa Engineering


Corporation).

2-203

Chapter 2

Underbalanced Drilling Techniques

RBOP

Choke
Line
Choke

Figure 2-59.

Line

Typical flowdrilling BOP stack (courtesy of Signa Engineering Corporation).

The kill line should consist of dual ball or gate valves and a check valve allowing flow toward
the wellbore annulus. With proper valving, the return fluids (oil, gas, cuttings and drilling fluid)
can safely be routed to the choke manifold and beyond.
Rotating Head vs. RBOP
The rotating control head has been in use for many years. It provides a low pressure means of
diverting return fluids away from the rig, in the event of flow from the well. It has found
widespread acceptance in air, mist and foam drilling operations. It can also be used safely in low
pressure or partially depleted formation drilling, or in low gas/oil ratio (GOR) applications.
Caution always should be exercised in using a rotating head in drilling applications because gas
can migrate out of the returning annular fluids, bringing higher than expected pressures to the
surface.

2-204

The rotating control head uses one or two stripper rubbers, designed for interference between the
inside diameter of the rubbers and the outside of the drillpipe or kelly. This seal design operates
adequately at design pressures until these rubber elements become worn from use, and a low
pressure leak occurs, as shown in Figure 2-60.
2-60 The seal design also does not allow for
monitoring wear or predicting life expectancy of the rubber elements. Because of these
problems, the American Petroleum Institute (API) refuses to recognize the rotating head as a
blowout preventer and equipment manufacturers should not rate their equipment with regard to
pressure containment.
As more operators began using underbalanced drilling operations in the late 1980s, well control
became a major concern. More than a few incidents involving the use of a rotating head led to rig
accidents, fatalities and well cost overruns. The low operating pressure tolerance of the
conventional rotating head and its propensity to leak made it the weak link in the system, as
higher pressured formations were drilled. An example of correct use of a rotating head would be
horizontal flowdrilling of a partially depleted reservoir, which has a low bottomhole pressure and
predictable surface annular pressure.
Drillstring

Stripper
Rubber

Leak Path
Well Pressure
Oil and Gas

Figure 2-59.
60

Leaking worn rotating head rubber.

For higher pressure applications, an improved diverter system was needed. The rotating blowout
preventer (RBOP) was conceived, as shown in Figure 2-61
2-61. It solved the low pressure limitation,
by increasing the drilling pressure rating of the rotating head from 400 to 1,500 psi. The RBOP
uses hydraulically-actuated packing elements, supported on large roller bearings and isolated by
mechanical seals, inside a large pressure vessel. Hydraulic oil pressure activates the rubber

2-205

packing elements, which contact the drillpipe or kelly (refer to Figure 2-62
2-62). This activating
pressure can be varied automatically as the wellbore pressure varies. The packing elements are
designed to secure the wellbore with a minimum application of 300 psi hydraulic closing
pressure.
The roller bearings are cooled and lubricated with the same hydraulic oil, which is contained by
two mechanical seals. The seals isolate the rotating packer and bearings from the wellbore. The
internal, bag-type packer element is made in two sections, so that any split or leak in the inner
section does not result in a loss of actuating pressure. The inner section of the packer can be
replaced, without replacing the outer section. Also, the internal packer is engineered for fullopening operation, eliminating the need to disassemble the preventer stack in order to change
bits. This kelly packer element provides a positive pressure seal on any surface. As this element
wears during normal operations, automatic increases in hydraulic oil volume compensate for
rubber loss. This feature keeps leaks from developing throughout the entire wear life of the
packer element. Whenever the element eventually requires replacement, the kelly packer can be
easily pulled through the rotary table and exchanged for a new element, by simply releasing the
securing mechanism.
One of the best applications for the RBOP is drilling in highly fractured or vugular, high pressure
environments. In such areas, the potential for rapid loss of circulation and resulting influx of
hydrocarbons can cause significantly higher surface annular pressure. The higher working
pressure of the RBOP adds a margin of safety over that provided by a rotating head. The RBOP
is designed to meet the American Society of Mechanical Engineers (ASME) pressure vessel
codes and the APIs specifications for annular blowout preventers.

2-206

Chapter 2

Underbalanced Drilling Techniques





Figure 2-62.


 

Rotating blowout preventer


(RBOP).







Figure 2-61.

RBOP sealing elements.

Choke Manifold
The choke manifold is necessary to maintain a safe operating back pressure on the return fluids,
to control the rate of hydrocarbon influx into the wellbore, and also to maintain proper hole
stability. A typical flowdrilling choke manifold is shown in Figure 2-63. It should be designed
to handle the maximum expected volumes (4-inch minimum piping) and should be redundant,
with dual chokes that enable one to be quickly isolated and cleaned if plugged with cuttings,
while the backup choke is operating. Flanged fittings are recommended on the choke, in order to
expedite the replacement process, in the event of internal erosion during operations. After each
flowdrilling job, all ell and tee fittings should be visually or radiographically (x-ray) inspected,
for fluid or cuttings washout, and replaced if necessary. Systems for continuous monitoring of
critical flowline points are currently being developed. It is recommended that these fittings be
leaded to decrease the amount of erosion which can occur during extended flowdrilling.

2-207

Sometimes referred to as babbitt.

Hydraulic
Choke

63
Figure 2-62.

Manual
Choke

A typical flowdrilling choke manifold (courtesy of Signa Engineering Corporation).

If any hydrogen sulfide (H2S) gas is possible in the formation fluids to be produced, it is critical
that special trim and metallurgy be used in the choke manifold construction and that all safety
requirements and training for sour service are met. It is safer to use a closed, two- or threephase, low pressure separator system, ahead of an open atmospheric pressure separator, to
remove poisonous vapors from the return fluid stream. To counter H2S, the pH can also be
adjusted and scavengers added.
During flowdrilling, the chokes will be essentially fully-opened and will gradually be closed, as
necessary, to control fluid rates and pressures at the surface. It is imperative that the annular
pressure not exceed the maximum rated working pressure (MWP) of the rotating device, during
choking. From the choke manifold, the oil, gas, drilling fluid and cuttings move downstream to a
mud/gas separator.
A sample catcher is often connected at the choke manifold, so that formation cuttings may be
caught and analyzed from the return flow stream, during live flowdrilling operations when the
rig pits and shale shaker are bypassed to the choke manifold. However, cuttings from flowdrilling
operations can become contaminated with hydrocarbons, making mudlog analysis difficult.
Mud/Gas Separator
A large, vertical mud/gas separator, with a height-adjustable support frame, is necessary for
proper gas removal in the separator. This open system, or atmospheric-pressured vessel, should
be at least 6 feet in diameter by 12 feet high, with sufficiently large gas flare lines (6- to 12-inch)

2-208

Chapter 2

Underbalanced Drilling Techniques

and adequate liquid dump lines to handle the expected instantaneous flow rates. The flare stack,
with variable height adjusters for different flow or location conditions, must be equipped with an
automatic flare igniter system. In areas of high gas production, it is common to see a 50- to 100foot high flare, concurrent with an annular pressure of 1,000 psi or more. This instantaneous
production rate cannot be precisely calculated; rather it is empirically derived to be between the
expected future production rate and the absolute open flow (AOF) rate of the formation. The
operating fluid level inside this vessel, as well as the diameters of the gas and fluid lines,
however, can be calculated from this instantaneous rate.
Conventional production criteria cannot be followed in the design of flowdrilling surface
production equipment because of slug flow from the wellbore, caused by the relatively
unregulated vertical expansion of gas in the liquid system. A maximum instantaneous gas flow
rate must be selected and then that rate is used in calculations to properly size the flare line and to
optimize the liquid level in the vertical separator. The pressure drop down the flare line, when
flowing the maximum instantaneous gas rate, should be less than the hydrostatic head of liquid in
the U-tube between the separator and the skimmer tank gas boot. This prevents gas from
entering the skimmer tank.
Weymouths equation can be used to predict the pressure drop for a gas, in steady-state,
adiabatic, flow along a pipe:
To d 16/ 3 (P12 P22 )
Q = 4335
.
STLZa
Po
where:

(2.73)

Q ........ gas flow rate (scf/D),


d......... inside diameter of the pipe (the gas
flare line in this case) (inches),
To ....... standard temperature (520 R),
Po ....... standard pressure (14.7 psia),
S......... gas gravity (air =1),
T ........ flowing gas temperature (R),
L ........ pipe line length (miles), and,
Za ....... average compressibility factor
(Weymouth used Za = 1),
P1, P2 .. the inlet and outlet pressures
(psia).
Weymouths Equation (2.73) incorporates a friction factor, f = 0.032/d1/3.

2-209

Assuming a gas gravity of 0.6, and substituting for standard temperature and pressure, Equation
(2.73) becomes:
Q = 19,754

d 16/ 3 P12 P22


TL

(2.74)

Converting length, L, from miles to feet, and flow rate, Q, from scf/D to MMscf/D, the inlet
pressure, P1, is:
Q 2 TL
+ P22
16 / 3
2.06 d

P1 =

(2.75)

The pressure differential exerted by the U-tube head can be expressed as:
P1 P2 = 0.433h

(2.76)

where:
........ specific gravity of the fluid in the
U-tube or separator, and,
h......... height from the top of the gas boot
to the bottom of the U-tube (feet).
Equations (2.75) and (2.76) can be combined to solve for the U-tube height, in terms of the gas
flow rate, temperature, outlet (atmospheric) pressure, and flare line diameter:

h=

Q 2 TL
+ P22 P2
2.06d 16/ 3
0.433

(2.77)

The mud/gas separator and the various separation pits and oil transfer tank should all be
surrounded by earth dikes, to contain any possible spillage.
Primary Oil Separation Pit
The liquids (oil, water and drilling fluid) and the cuttings, with free gas already removed, are
carried out of the mud/gas separator in the liquids line, to an open-topped skimmer pit. This 400
bbl primary oil separation pit consists of an entry gas-scrubbing device, called a mud buster or
gas boot. It is designed to remove most of any retained gas, that was not previously separated in
the primary mud/gas separator. A gas boot is simply an open-ended vertical section of 20-inch
casing, extending from the inlet to the pit down to within two feet of the bottom of the pit. A
three- to four-inch atmospheric-pressured flare line will vent the gas boot and avoid
accumulation of free gas in this pit.

2-210

Chapter 2

Underbalanced Drilling Techniques

Skimming in this pit involves a bucket and weir type mechanism that depends on regulated
(manual) leveling of the fluids in the pit. The open-top design is necessary so that cuttings
accumulating in the first pit section can be removed by jetting, in much the same way that rig pits
are managed. After these cuttings gravity-settle out of the oil and drilling fluid, they are
periodically jetted out of the primary oil separation pit to a nearby, lined cuttings pit, where they
may be disposed of in an environmentally safe manner.
Secondary Oil Separation Pit
Drilling fluid from the primary oil separation pit is usually gravity flowed to another 400 bbl
settling pit, for additional retention and further removal of any cuttings and formation oil, carried
over from the first pit. The cleaned drilling fluid is then transferred to a drilling fluid return pit.
Both of these skimmer pits have baffles, partitions and valves to increase retention time. Longer
retention time improves oil recovery and allows any final cuttings to settle out of the drilling
fluid. Chemical injection treatments can be carried out in each of the skimmer pits, as necessary,
to help break oil-water emulsions and aid in oil recovery and drilling fluid cleaning.
Skimmer System Safety
An open-topped pit is a possible fire hazard and must be at least 400 feet downwind of the
wellhead. Vehicular traffic should be prevented from entering the skimmer area after
flowdrilling begins. It is not advisable to flow drill with an open-topped pit, if hydrogen sulfide
is anticipated or detected. Fire extinguishers should be placed near the skimming system pits, in
accordance with the requirements of the local regulatory authority.
All transfer pumps, lights, electrical connections and switchgear should be explosion proof. The
pits should have handrails, to prevent an operator from falling into them, and the exit stairways
should be located to permit quick evacuation of the pit area.
Drilling Fluid Pit
The drilling fluid pit serves as a reservoir for cleaned drilling fluid, which is automatically
returned to the rig pit for mud pump suction. The drilling fluid pit contains two or more
centrifugal pumps, with level controllers. These maintain levels adequate to ensure proper gravity
flow from the primary and secondary oil separation pits. Each centrifugal pump has its own level
controller and acts independently, to maintain pit level regardless of the return flow from the
well. Complete redundancy in these pumps is necessary, to ensure that the return flow of drilling
fluid will not be interrupted during critical flowdrilling periods.
Oil Transfer Tank
The oil transfer tank receives the oil recovered by the primary and secondary oil separation pits.
It is located at a safe distance from the drilling rig, and is constructed to allow the operators
representative to accurately estimate oil production rates during flowdrilling operations. This
information, along with flare data, provides the operator with fairly accurate production rate
estimates, for qualitative, real-time evaluation of downhole hydrocarbon-bearing intervals. The
oil transfer tank should be equipped with a high liquid level alarm and be positioned at least 100
feet upwind of the primary and secondary skimmer systems, for safe loading of oil transport

2-211

trucks. A manifold system and sufficient oil storage tanks are required to keep up with oil
production into the oil transfer tank.
2.8.4 Operating Procedures
The mechanical objectives during flowdrilling are to control the well at the surface (rather than
below the surface), to maintain underbalanced conditions downhole at all times, to minimize
differential sticking problems and to minimize drilling fluid losses caused by lost circulation.
Drilling operations should be designed to hold the maximum safe underbalance and to avoid all
unnecessary interruptions to circulation.
Initially, when drilling out the casing shoe, returns should be sent directly to the shale shakers.
As soon as formation fluid inflow starts and flowdrilling is established, returns should be
directed instead through the choke manifold into the surface separation system.
Maximum tolerable surface pressures should be established before drilling. These will depend
on the pressure bearing capacity of the rotating head or RBOP. The pressure limits of an RBOP,
are highest when the string is static, and are lower when stripping the string through the diverter
than when drilling ahead. If the surface pressure reaches these limits, steps have to be taken to
reduce it. This can be done by shutting the well in below the diverter and bullheading a higher
density liquid into the annulus. Alternatively, the influx can be circulated out through the choke.
Once the surface pressure has stabilized, the density of the drilling fluid should be increased to
reduce the underbalance pressure to a manageable level.
If the annular pressure increases during connections or trips, when circulation is suspended, this
probably indicates gas migration up the annulus. In this case, increasing the viscosity of the
drilling fluid can restrict the rate at which pressure increases. This can, however, cause problems
with separating gas from the drilling fluid at surface.
Tripping and logging operations should be planned to reduce the underbalance pressure, while
still maintaining underbalanced conditions.
Special procedures are used during flowdrilling, to make connections and for tripping. These
necessitate longer rig time than is required in overbalanced drilling. In making a connection,
pressure on the drillpipe must first be bled off above the top drillstring float, before breaking the
connection. One or more of these floats is installed in the drillstring to serve as check valves and
prevent reverse flow up the drillpipe due to the underbalanced pressure differential. One float,
installed every twelve joints, provides an ideal system for making connections and trips.
To make a trip, additional steps must also be taken to ensure safety. If a gas source is available,
injecting gas down the drillstring can displace the drilling fluid down to the deepest drill float, to
avoid pulling a wet string. On the annulus side, it is desirable to lower the pressure differential
before making a trip. This is generally done by bullheading down the annulus with a more dense
drilling fluid, to reduce the underbalanced pressure differential, but maintain, if possible,
underbalanced conditions downhole. The volume of fluid bullheaded is usually calculated to
2-212

Chapter 2

Underbalanced Drilling Techniques

displace annular fluids in the vertical, cased portion of the hole only, avoiding the curve and
horizontal lateral sections in directional holes.
An example is a well being horizontally flowdrilled with fresh water to achieve an underbalanced
condition downhole. With adequate pressure in the target formation, gas injection may not be
required for underbalanced drilling. The reservoir is immune to the negative affects of clay
swelling or other reactions caused by contacting it with fresh water. Surface annular pressures
during flowdrilling operations are sufficiently low to allow the use of the lower operating
pressure rotating head. Drilling costs are reduced because this equipment is significantly less
expensive to rent than a rotating blowout preventer.
Connections are easily made, following pressure bleed off above the shallowest drill float. To
minimize annular pressures before a trip, a 10.0 ppg saturated sodium chloride (NaCl) saltwater
is bullheaded down the annulus to the shoe at the bottom of the deepest casing set. Annular
pressures are then monitored during the trip, as the drillstring is removed under pressure from the
well. If the annular pressure increases to a safe level below the maximum working pressure of
the rotating head during the trip, for example 300 psi, additional brine must be bullheaded down
the annulus to lower the pressure differential before continuing the trip.
Once an underbalanced drilling operation is finished, the well should be completed using
underbalanced technology. Well impairment reductions achieved in the flowdrilling phase
can be lost or reduced if an overbalanced completion is made. If a slotted liner completion is
selected, the liner must be run below the slipping elements in the BOP stack while flooding the
annulus with clear, non-damaging fluid. Once the slotted line clears the BOP stack, flooding the
well can be discontinued and the liner running string can be stripped in to the liner setting depth.
If casing must be run and cemented, it is important that the pipe be run in the hole with the well
underbalanced. Again, annular pressures may be subdued with clear, non-damaging fluid, to
allow the safe running of pipe. In most cases, casing may be run through the rotating device by
removing the rubber packing element. In certain cases, a snubbing unit must be used, if annular
pressures cannot be sufficiently lowered. Foamed cement may be used to cement the casing,
while maintaining downhole underbalanced conditions.
2.8.5

Limitations

High Annular Pressures


As noted earlier, flowdrilling can allow more productive formations to be drilled underbalanced.
Correctly sized surface equipment should be able to handle any feasible instantaneous production
rate. It is possible that the limiting production rate used to design the surface system could have
been underestimated. In that case, once the production rate approaches the surface system limit,
the return flow can be choked back. The extent to which the flow can be choked back is limited
by the pressure capacity of the diverter. The annular pressure that is required to restrict
production to a manageable rate may reach the safe working limit of the diverter. Then, there are
two options; increase the drilling fluid density or change drilling technique. When lost circulation
occurs while flowdrilling, the cost and drawbacks of high density drilling fluids may make an
alternative drilling technique more attractive. Two possible techniques are mudcap drilling and

2-213

snub drilling (flowdrilling with a snubbing unit). These are described in Sections 2.9 and 2.10,
respectively.
Uncertain Formation Pressures
Flowdrilling is not particularly suited to areas where pore pressures are not well known, or where
the pore pressure gradient increases with increasing measured depth. Under these circumstances,
the well is likely to experience increasing surface pressures and/or crossflow between different
permeable zones downhole.
One situation that should be avoided, if at all possible, is flowdrilling a gas producing
formation when there is a lost circulation zone open uphole, that is known to have a
significantly lower pore pressure than the producing formation. It is quite possible that such a
well will have to be killed, to avoid charging the lost circulation zone uphole. If the producing
formation is to be drilled underbalanced, then it may be necessary to case the lower pressure
zone, before penetrating the producing formation. These concerns limit, but do not totally
preclude, the application of flowdrilling in vertical wells.
Generally, horizontal wells are drilled in regions of known formation pressures and the pay is
likely to have a more or less constant pressure gradient. As a result, flowdrilling is well suited to
horizontal wells.
Wellbore Instability
Flowdrilling involves higher borehole pressures than would be seen with lightened drilling
fluids. Weaker formations can be drilled with reduced risk of borehole collapse. Even so, the
borehole pressure may still be insufficient to prevent the onset of wellbore instability in weak
formations. Furthermore, the high annular velocities and high levels of turbulence during flow
can erode the borehole wall. Flowdrilling is not suitable for unconsolidated formations.

References
1.

Stone, C.R.: The History and Development of Underbalanced Drilling in the USA, paper
presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

2.

Joseph, R.A.: Underbalanced Horizontal Drilling-1 Planning Lessens Problems, Gets


Benefits of Underbalance, Oil and Gas J. (March 20, 1995) 86-89.

3.

Tangedahl, M.J.: Well Control Issues of Underbalanced Drilling, paper presented at the
1995 1st International Underbalanced Drilling Conference, The Hague, The Netherlands,
October 2-4.

2-214

Chapter 2

2.9

Underbalanced Drilling Techniques

Mudcap Drilling

Sometimes, uncontrollable loss of circulation occurs during flowdrilling operations. The driller
is faced with higher annular pressures than can safely be handled with the rotating head or RBOP
equipment. One technique, called mudcap drilling, can be used to overcome this situation. The
technique has been successfully used in drilling fractured carbonate formations in the United
States, in Western Siberia and in offshore drilling areas of Southeast Asia.1 It is not strictly an
underbalanced drilling technique, since the well does not flow to the surface. Rather, it is a
technique for managing severe lost circulation in an over-pressured environment, without losing
well control. Mudcap drilling does, however, use much of the equipment used for flowdrilling,
and, in some circumstances, it is a logical alternative, when the formation pressure is too high for
flowdrilling.
In mudcap drilling, the driller loads the annulus with a heavy, viscosified fluid, often saturated
brine, and shuts-in the annulus of the well. This is illustrated in Figure 2-64. The shut-in surface
pressure on the annulus, plus the increased hydrostatic pressure resulting from this viscous pad,
will equal the formation pressure. Viscosification of the pad should be designed to minimize gas
migration up the annulus. The annular column is held in place by its density and the bullheading
pressure of the rig pumps. It may be periodically necessary to add fluid to this mudcap, to offset
annular losses to the formation during connections or trips.
Drilling may then be resumed by pumping a clean fluid that is compatible with the formation
fluids down the drillpipe, while the choke is closed and the well remains shut-in. This blind
drilling approach with a sealed annulus results in bullheading all drilling fluid pumped with no
return flow. Obviously, the formation must be able to freely accept these fluids and the fluid used
must readily be available and relatively inexpensive. This process requires specialized well
control and circulating equipment; however, unlike flowdrilling, it does not require an extensive
fluid separation system, since the formation fluids are kept downhole.
Mudcap drilling is best applied in wells with:
 Sustained surface pressures in excess of 2,000 psi,
 Sour oil and gas production, and,
 Small diameter wellbores (3 7/8-inches up to 4-inches).
One additional difference between flowdrilling and mudcap drilling is that an RBOP, with its
high operating pressure limits, is essential for safe mudcap drilling. Flowdrilling is possible using
either an RBOP or a rotating head. A schematic layout is shown in Figure 2-65. In designing a
well plan using a mudcap drilling format, drilling engineers should carefully consider the high
standpipe pressures involved and the associated safety considerations, before finalizing their
recommendation.

2-215

Viscous Fluid Mudcap

Mudcap Interface
(Formation Fluid / Drillwater)
Water replacement in formation
fractures

Figure 2-63.
64

An example of mudcap drilling (courtesy of Signa Engineering Corporation).

Mudcap Drilling Example


A fractured carbonate formation, with an initial pore pressure gradient of 14.5 ppg, containing
sour oil and gas (H2S), was the drilling target below the 5-inch casing shoe, in a 4-inch hole,
at a true vertical depth (TVD) of 8,000 feet. Horizontal penetration, of the vertical fractures in
theformation, was expected to result in the loss of drilling fluid. Permanent formation damage
from the water-based mud was possible. The operator also wanted to keep the sour production
downhole while drilling underbalanced.
These challenges - high pressure, sour production, loss of circulation, formation damage, slim
hole - made this an ideal candidate for mudcap drilling.

2-216

Chapter 2

Underbalanced Drilling Techniques

Mudcap drilling can be used to control annular pressures and decrease the cost of lost drilling
fluid. In this example, a pre-calculated volume of fresh water was weighted to 14.5 ppg by
adding calcium carbonate chips and viscosified using XCD polymer. This fluid pad was
bullheaded into the annulus, with the return line shut-in. The driller would then start drilling with
the annulus closed in, by pumping down the standpipe, using either brine or fresh water, treated
with potassium chloride or lime for inhibition and a hydrogen sulfide scavenger. The 3-inch
motor used to drill a 4-inch hole requires approximately 110 gpm supply for efficient operation.
At this rate, a sixteen hour drilling day would require over 2,500 bbls of drilling fluid.

GAS
BUSTER

To flare pit

MUD
PITS
RIG FLOOR

HCR
Valve
(Closed)

MUD
PUMPS

Chemical
Injection
CHOKE
(Closed)

Section B
DIVERTER

Section A

Figure 2-65.

Schematic of equipment required for mudcap drilling (courtesy of Signa Engineering


Corporation).

If a 10 ppg NaCl brine was used as the drilling fluid in the above example, the standpipe pressure
would be 1,872 psi plus the drillstring friction pressure plus the formation injection pressure. If
fresh water was used, rather than the NaCl, the standpipe pressure would be 2,580 psi plus the
drillstring friction pressure plus the formation injection pressure. The relatively low pump rate
used in a slimhole application (low drillpipe friction pressure) and the high permeability of a
naturally fractured formation (low injection pressure) would minimize the increase in standpipe
pressure above the differential pressure exerted by the annular mudcap.

2-217

Slimhole drilling using the mudcap technique is logistically efficient. In the above example, the
volume of the viscous mudcap fluid is only 16 bbls in the annulus. This small volume allows the
driller to economically build and place the expensive mudcap fluid in place. The small annular
volume of a slimhole also decreases the bullheading volume required to place the mudcap in
position.
Bloys et al., 1994,1 discussed drilling protocols in certain carbonates, using mudcap drilling.
Currently [offshore South-East Asia] it is only possible to drill with a rotating head on rigs
equipped with a surface stack; this includes jackup and platform type rigs offshore. The
advantages of this method are that there is no guess work about the fluid level in the well and the
pump rate required on the annulus is dictated only by the need to prevent gas migration. In gas
filled carbonates the increasing over-balance would also be reduced by using a lower density
mud. Hence, the logistics of pumping mud down the annulus could be planned more precisely
based on the velocity required to prevent gas influx and migration (i.e. up to [120 ft/min]
depending on fluid viscosity). Hence, the use of a rotating head and lower density mud could
potentially reduce the volume of fluid required and the cost to drill a given massive lost
circulation interval. If substantial lost circulation is anticipated and if there is a potential for
mudcap drilling, Bloys et al., 1994,1 called out extra equipment and protocols for their offshore
operations, including:
 The subsea BOP stack must include two annular BOPs. An extra kill line connection
between them is useful.
 Kill and choke lines must be a minimum of three inches in diameter, to minimize friction.
 There must be a facility for displacing the choke line and the riser booster line to diesel and
periodically pumping diesel to keep the lines full.
 Supplementary monitoring equipment includes chart recorders to both the choke and kill
lines, as well as pressure gauges (0-500 and 0-1000 psi ranges).
 Pressure transducers (and backups) are required in both the booster line and the kill line at the
BOP and linked via the BOP control system to the surface pressure gauges and charts. This
gives direct measurement of pressure above and below the annular.
 The BOP and accumulator system must be tested, to permit ram controls to be left in the
activated open position. In the block position, the rams can close when the fluid level in the
riser falls.
 Two non-ported floats in the string are needed for mudcap drilling.
 For deep water, the riser should be equipped with a facility to choke out gas (up to 1000 psi)
that gets above the BOP and a line for injecting glycol from the surface directly into the BOP.
 Mixing and storage facilities are required. On the rig, mixing and storage for 10,000 BPD is
needed for up to two days.
 Storage and batch mix tanks are needed for 100 bbl pills.

2-218

Chapter 2

Underbalanced Drilling Techniques

 It must be possible to have diesel in all lines from the rig floor standpipe manifold to the
cementing unit and from the cement/gunk pill tank to the cementing unit.
 There needs to be an adequate diesel supply (200 gpm) to the cementing unit, for mixing and
pumping cement/gunk pills.
 There needs to be an adequate seawater supply (1000 gpm) to the mud pumps and
manifolding for seawater to be pumped down drillpipe and mud to be pumped down the kill
line or into the riser, simultaneously.
 There must be supply boat resupply of mud materials for 10,000 BPD.
 Contingency procedures and supplemen-tary crew training are essential.

References
1.

2-219

Bloys, B., Brown, J.D. and Tarr, B.A.: Drilling Safely and Economically in Carbonates:
Collective Experience of Arco, BP and Mobil, paper presented at the 1994 IADC Well
Control Conference of the Asia/Pacific Region, Singapore, December 1-2.

2.10

Snub Drilling

Snub drilling is simply an underbalanced drilling operation that involves the use of a snubbing
unit or a coiled tubing (CT) unit. The additional expense of this equipment can be justified if
very high formation pressure and uncontrollable loss of circulation are expected. Often,
personnel safety considerations provide the necessary reasons for snub drilling. Finally, if sour
gas is expected, there is additional motivation.
Both snubbing and CT units have BOP stacks that allow a drillstring (coiled tubing in the latter
case) to be run into or out of the hole, at much higher pressures (routinely up to 10,000 psi) than
can be tolerated by either a rotating head or an RBOP. Both units also allow the drillstring to be
pushed into a well under pressure, even when the weight of the string alone is insufficient to
overcome the pressure tending to push it out of the well. Snubbing and CT units can be used for
underbalanced drilling, at pressures that cannot be managed by conventional drilling rigs.
Stone, 1996 (unpublished), recently used a rig-assisted snubbing unit, to drill a horizontal hole
through an abnormally pressured (17.0 ppg), fractured carbonate formation at a TVD of 8,000
2-66 shows the well
feet. This well is used to illustrate the process of snub drilling. Figure 2-66
profile. The expected problems, which had been determined in preparing the well design plan
and cost estimate, were:
 Very high annular surface pressure,
 Vertical fractures,
 Severe loss of circulation,
 High pressure stripping operations,
 Expensive drilling fluid,
 Surface equipment limitations, and,
 Personnel and rig safety considerations.
Vertical offset wells, drilled with 17.0 ppg oil-based mud systems, were extremely expensive,
because of severe lost circulation. An evaluation was made to determine the feasibility of drilling
a well using a less expensive, lighter drilling fluid, with a viscous mudcap to reduce the surface
pressures during trips. The resulting recommendation called for setting 5-inch casing at the top
of the target formation and drilling a 4-inch hole using 2 7/8-inch tubing as the drillstring.

2-220

Chapter 2

Underbalanced Drilling Techniques

A snub drilling method was selected as the best approach. A standard 150,000 lbf snubbing unit
was mounted inside the derrick of the drilling rig, above a 15,000 x 10,000 psi snubbing BOP
stack. This combined snub drilling BOP stack is shown in Figure 2-67. The rig rotary table was
removed, to allow the top BOP element to extend through the rig floor. All circulating lines
(standpipe and kelly hose) were replaced with 10,000 psi equipment. Saturated 10.0 ppg sodium
chloride (NaCl) was selected as the drilling fluid. The expected 7.0 ppg underbalanced pressure
differential required extremely close coordination between all personnel involved in the
operation.
The vertical section of the well was drilled to 8,000 feet using 17.0 ppg oil-based drilling fluid
and was cased with 5-inch pipe. This oil mud then displaced with the NaCl brine after the
snubbing stack, snubbing jack and surface equipment were rigged up and pressure tested.

7" 26# @ 8128'

Pilot hole
dressed off
to 8285'

Top of productive interval @ 8157'

KOP @ 8302'

60 deg
6-1/8" Hole to 8550'

4-3/4" Hole

FORMATION DIP 6-8

N 82o E

8558'
(Secondary Target)

SHALE

SHALE

Target Center
(Primary Target)

8578'
8594'
8608'
8618'

Pilot Hole

Top of SHALE 8821'

Figure 2-66. Snub drilled well profile (courtesy of Signa Engineering Corporation).

2-221

Drilling Spool
7-1/16", 10M x 7-1/16", 5M

RIG FLOOR
STRIPPER

Cameron single
7-1/16", 10M

Annular Preventer
Cameron 7-1/16", 10M
Teflon & graphite impregnated rubber
STRIPPER

Cameron "U" double


7-1/16", 10M
PIPE
Gate Valves
HCR

Install companion flange


w/2" WECO 1502 thread

Drilling Spool
7-1/16", 15M x 10M
PIPE

Cameron "U" double


PIPE

7-1/16", 15M

DSA
7-1/16", 10M x 7-1/16", 15M
Frac Valve
7-1/16", 10M
TUBING HEAD
11", 5M x 7-1/16", 10M
Outlets
plugged

CELLAR

Outlet with (2) 1-13/16"


10M Gate Valve
SOW CASING HEAD
11", 5M x 9-5/8"

Figure 2-67. BOP stack (courtesy of Signa Engineering Corporation).

The intent was to flowdrill the well through the prospective formation, controlling the influx of
gas from this high pressure reservoir with the surface choke system. Mudcap drilling techniques
were only to be used for trips. The snubbing unit would be used for these trips. With no gas
cutting of the NaCl drilling fluid, the anticipated surface annular pressure would be 2,900 psi,
due to the high pore pressure of the formation. With the expected gas cutting of this fluid,
annular pressures approaching 5,000 psi were probable. One thousand barrels of 11.6 ppg
calcium chloride (CaCl2) brine and XCD polymer were kept on location for bullheading viscous
pills down the annulus during trips, to lower these surface pressures. Fine-to medium-ground
calcium carbonate was kept on location, as a weighting additive, to increase the density of
the viscous trip pills. To make a 100 bbl trip pill, weighing 14.5 ppg, 342 sacks (100 lbm
each) of fine calcium carbonate were mixed with the 11.6 ppg CaCl2 brine. Tripping with this
2-222

Chapter 2

Underbalanced Drilling Techniques

fluid in the annulus lowered the surface pressures to 1,050 psi or less. With this pressure on the
annulus, the rig crews had to pull wet strings during trips.
The drillstring was new 2 7/8-inch, 8.70 lbm/ft, P-105, PH-6 pipe. This tubing was pressure
tested at the surface, each time a joint was made up, as it passed through the top stripper rubber.
Each BOP stack element was fitted with a pressure bleed valve, to relieve trapped pressure
between elements, during trips. In addition to the surface BOP stack, a full-opening safety valve
and inside blowout preventer for the 2 7/8-inch drillstring were kept on the rig floor at all times.
Both upper and lower kelly cocks were installed.
The surface equipment system, shown schematically in Figure 2-68, included four choke
manifolds. Two of these handled annular flow when the fluid was primarily gas. They safely
eliminated erosion problems, caused by high pressure and velocity of a gas stream, laden with
drill solids. The remaining two chokes were designed to process the return stream when it
included liquids. The operator of these chokes would monitor the flare for liquids, to switch
from the gas chokes to the liquid chokes and vice versa. Both liquid choke manifolds were fourinch, 10,000 psi working pressure systems. Both the primary and the back-up chokes consisted
of one hydraulic and one manual choke. The command station for all four chokes consisted of
the controls for each manifold, as well as pressure gauges showing both the standpipe and
annular pressures. The chokes were adjusted to maintain constant standpipe pressure. Two
choke operators were on duty at all times to man the choke command station.
The skimmer pit received the drilling fluid and condensate from each of the liquid choke
manifolds and attached gas busters. The drilling fluid, skimmed of liquid hydrocarbons and
cleaned of all drill cuttings, was returned to the rig mud pits for circulation downhole. The
cuttings volume was quite small in this 4-inch slimhole application. The liquid hydrocarbons
were directed to the oil transfer tank. During drilling operations, two snubbing unit operators
worked on the rig floor. Their duties included maintaining control of the snubbing stack,
manipulating the pipe and assisting the driller. During drilling of the curve and lateral sections,
the drawworks was used to support and lower the drillstring to provide weight at the bit.
Rotation of this string was accomplished with a power swivel. As previously mentioned, the
snubbing unit was used only for tripping the pipe. The secondary BOP stack controls were
remotely positioned a safe distance away from the rig floor.

2-223

FLARE
PIT
6'' GAS

6'' GAS

LIQUID

GAS
BUSTER

SKIMMER

LIQUID

GAS
BUSTER

4'' GAS

DRILLING FLUID
RETURN

ALL
GAS

MUD PIT
ADJUSTABLE
MANUAL CHOKE

MANUAL
CHOKE

HYD.L
CHOKE

DRILLING FLUID
RETURN

GAS + LIQUID
SAND
SEPARATER

PREVAILI
NG
Prevailing
Wind
W IN D
Direction
D IR ECTIO N

WELLHEAD

Figure 2-67.
68 Snub drilling choke system (courtesy of Signa Engineering Corporation).

Communication between rig-site personnel becomes especially important when drilling with
these extreme surface pressures. A loudspeaker intercom was installed to interconnect the rig
floor, shaker pit, skimmer pit and all living quarters. Very high noise levels will normally
accompany fluids flowing through chokes that are dropping over 5,000 psi at times. Hearing
protection was necessary for all hands working on the rig floor and close to the choke manifolds.
The motorman, as usual, monitored and maintained the operation of the rig pumps and engines.
A pump truck was added to provide drilling fluid to the operation, when the standpipe pressures
exceeded 4,000 psi. The operation also required two fluid transfer operators, at all times during
drilling operations. Their primary responsibility was to mix brine and trip pills and maintain
working pit levels. The logistics of transferring and mixing brine were critical. A brine mixing
plant was recommended for use on location for rapid fluid building capability.
The project coordinator was responsible for training each man to do his job and to coordinate
verbal and hand signals among the team members. He had ultimate control of the project and
could shut down the entire operation, if necessary, for safety reasons. In the event of a shutdown,
the choke operators would monitor the surface pressures and report each 100 psi increase. Two
project coordinators were always on location, with only one having control of operations, at any

2-224

Chapter 2

Underbalanced Drilling Techniques

given time; the other served as a relief for the active coordinator. All critical team members were
housed on location so that any off-duty team member could be called on to assist, if needed.
Sunb drilling and coiled tube drilling are discussed further in Chapter 6. Adam and Berry, 1995, 1
summarized some of the considerations in an underbalanced well, drilled with coiled tubing, in
the Dalen field. These included custom modifications of standard well test equipment, and
separation/surge systems. Because of the potential for H2S, which was soluble in the drilling
fluid, a closed tank drilling fluid treatment and storage system were adopted. Drilled solids
were removed with sand filters and sand catchers, upstream of the choke manifold and separator.
This was sufficient because the volume of drilled solids was very low. They also developed a
sub-surface safety valve, which was installed in the completion string at 107 m and was used to
isolate reservoir pressures from the surface, effectively creating a downhole lubricator system.
While expenses were incurred, it was anticipated that there was economic merit since previous
wells had encountered severe mud losses, where the costs in terms of materials and time spent to
combat the problem were on the order of $2,500,000. Finally, there was the opportunity for
interactive well design. Underbalanced drilling was selected to overcome the lost circulation and
to allow gauging the production contribution from each fracture system intersected, as specific
fractures were penetrated. Thus it should be possible to tailor the length of the horizontal
section based on observed well potential rather than drilling to some nominal departure, which
may not provide any incremental benefit.

References
1.

2-225

Adam, J. and Berry, M.: Through Completion, Underbalanced, Coiled Tubing Side-Track
of Well Dalen 2, paper presented at the 1995 IADC Drilling Conference, Milan, Italy.

2.11

Closed Systems

Some of the underbalanced drilling techniques, described so far, have been categorized according
to the drilling fluid involved. This section describes an underbalanced drilling technique that
involves using a specific type of surface system, rather than a specific drilling fluid. The
distinguishing feature of this technique is the use of a pressurized, four-phase separator and a
fully closed surface system, to handle the fluids returning from the well.
The use of a closed surface system when drilling underbalanced was pioneered in Canada.1,2,3
These systems are finding applications elsewhere in the world; for example in the UK,4
Argentina, the Netherlands,5 and in Oman.6 At the time of writing, they are not extensively used
in the United States. They offer benefits over open surface systems, but they also have
limitations that have to be considered. The benefits arise because all fluids are contained within
flow lines, pressure vessels and closed holding tanks. These systems can safely manage natural
gas production containing hydrogen sulfide, prevent hydrocarbon vaporization from open pits,
(i.e. environmental benefits) and, with appropriate instrumentation, allow continuous
measurement of a wells productivity.
With planning, closed systems can be designed for high pressures, when drilling deep and overpressured reservoirs.
2.11.1 Surface Equipment
A closed surface system consists of a BOP stack connection, choke manifold, sample catcher, a
pressurized separator system which may consist of one or more component vessels), a flare stack,
production tank(s), and a clean drilling liquid return line. Figure 2-69
2-69 schematically shows a
closed system, typical of those used in Canadian operations. Many of these components are more
or less the same as those used in open surface systems. The distinguishing feature is that the
entire system is closed until cleaned drilling liquid returns to the mud pump suction tank.
Several Canadian service companies provide underbalanced drilling closed systems as rental
items. Often two or more of the components of the system will be combined on single skids.
Return Line
The return line will normally be taken from beneath the RBOP or rotating head outlet flange. If
dual annulars are used, it will be taken from a diverter spool immediately below the upper
annular. The line diameter should be sufficient to keep the pressure drop down the line to a safe
level, at the maximum anticipated return flow rate of injected and produced fluids. A 6-inch
return line may be necessary. Some means of closing the return line should be installed, adjacent
to its connection to the BOP stack. So too should an emergency shutdown (ESD) valve. Local
regulations may demand installation of this ESD. The ESD itself, and any piping between the
BOP stack and the ESD, should have a pressure rating that at least matches the rating of the BOP
immediately above the return line take-off.

2-226

Chapter 2

Underbalanced Drilling Techniques

The return line should be equipped with valving, downstream of the shut-in valves, to permit the
option of directing return flow to the shale shakers. This may be desirable if no hydrocarbons or
toxic gases are entering the well or anticipated to enter it. There will normally be a flow control
manifold and sample catcher arrangement between the return line and the separator system.

Ignitor

Flare Stack
S am ple Catcher

Production
Tank

Pressure Vessel

Choke Manifold

S tack

N2 Pumpers
Mix
Drilling Fluid Tank
Rig Pump

Figure 2-69.

Vaporizor

A typical closed surface system (modified after Lunan, 1994 2).

There may be one or two sets of choke and kill lines.3 As a minimum, a single choke line, from
below the blind rams, will be routed through the rig manifold to the separator system or directly
to the flare pit.7 The well can then be flowed while tripping or when the string is out of the hole
and the blind rams are closed. This will prevent excessive surface pressures from building up. A
second choke line may be installed, either between the annular and the upper pipe rams, or from
between the upper pipe rams and the blind rams. If installed, this will be valved to permit flow
to be routed through the return line and its flow control manifold to the separator or directly to
the flare pit.7 This system provides a degree of redundancy against blockage of the main return
line or problems with either of the choke manifolds.
Typical Canadian operations use 2,000 psi rated, hammer union coupled flow lines. The return
and choke lines should have pressure ratings matching those of the choke manifold to which they
are attached. These should be sufficient to contain any surface wellhead pressure that can
reasonably be anticipated. Since one of the major reasons for using a closed system is to handle
sour gas, the closed system components are normally rated for sour service. If there is any
potential for hydrogen sulfide being encountered, it should be confirmed that all components of
the surface system, that might be exposed to the well returns, are suitable for sour service. All
flow lines should be electrically bonded to prevent any sparking.

2-227

The lines, between the BOP stack and the separator, can erode because they are exposed to
solids-laden return flows. All elbows and tees should preferably be leaded, to restrict this
erosion. This becomes particularly important if consolidated sandstones or cherts are drilled.
Flow Control Manifold
Returns from the well should be routed through a flow control manifold at all times when drilling
with a closed system. A second choke manifold, in addition to the rig choke manifold, is
normally installed. Figure 2-70
2-70 is one example of a flow control arrangement. A flow control
manifold, designed specifically for underbalanced drilling, would normally be used for this
purpose. These tend to have larger diameter valves, (six-inch gate valves and four-inch ball
valves, in some instances7), than the conventional choke manifolds found on many rigs. This
allows them to handle higher flow rates, without generating excessive back pressure. The
underbalanced drilling flow control manifold will normally be used to control flow into the
separator. The rig manifold is only used to control flow to the separator when the primary
manifold requires maintenance or when excessive pressures are encountered.

Rotating Blow out


Preventer/Diverter

To Shale Shaker

RBOP
ESD

RBOP Height
1700 mm

Northland Manifold
6" Gate Valve

Annular Returns to Choke Manifold


and Separator

Sample Catchers

4" Globe
Valves

Annular Preventer

Willis Choke

127 mm (5") Pipe


Rams

Flare Stack

Choke Line Connected


to Northland Separator Manifold

Kill Line

Separator
200 psi Vessel

Rig Manifold

Shear/Blind Rams
127 mm (5") Pipe
Rams

Choke Line
Connected to
Rig Manifold

Kill Line
D.S.A.

Choke
Water Returned Oil Storage/
to Rig Tanks
Transport

HCR

Tubing
Spool

Choke

Flare Pit

Casing
Spool

Surface Casing
300-400m, 508.0 mm

Intermediate Casing
1300-1450m, 339.7 mm

Production Casing
1890 m, 244.5 mm

70
Figure 2-69.

Flow control arrangement (after Saponja, 1995 7).

2-228

Chapter 2

Underbalanced Drilling Techniques

Sample Catcher
With a closed system, cuttings are normally collected inside the separator vessel. Some form of
sample catcher has to be installed in
the flow line, between the BOP stack and the separator, if cuttings samples are to be taken for
geologic control. Usually, two sample catch vessels are used; one collects a sample while the
other is being emptied. The catch vessels may be mounted downstream of the flow control
manifold, or alternatively integrated with it. Figure 2-71 shows an arrangement that puts the
catch vessels on the flow control manifold. The two catch vessels normally contain screens
through which some of the well returns will flow, depositing cuttings on the screen. The catch
vessel is isolated from the return flow, and flow is routed through the other catch vessel while the
screen is removed from the first vessel and cuttings are collected. Screened vessels are not
necessarily used when drilling gas wells; instead the sample may just be allowed to collect in the
catcher pan.2
Separator
Many of the separators used in underbalanced drilling have been modified from production
separators, because of the different design requirements.8 A separator used to process
underbalanced drilling returns will need to provide four-phase separation of liquid hydrocarbon,
water or aqueous drilling fluid, gases (both produced and injected), and cuttings. Typically,
underbalanced drilling separators operate at 20 to 50 psig,9 but they can be rated for 200 to 500
psi maximum pressure. Separating gases from the returns is only efficient if the separator
pressure is low. Additional de-gassing equipment may be necessary downstream of the primary
separator - this will be discussed later in this section.
Frequently, underbalanced drilling returns are produced in slug flow, with intermittent slugs of
liquid at much higher instantaneous flow rates than the overall average liquid rate. These slugs
can overwhelm a conventional separator. At the very least, the internal separator design will
have to minimize splashing when a slug enters. Some underbalanced drilling separators have a
spiraled entry baffle for this purpose.2 Using a higher vessel pressure will reduce the tendency for
slug flow, but this will also decrease the efficiency of removing gas from the liquids.8

2-229

Flow
Direction

Output
Data
Header

Valve
#2

Sample
Catcher
#1

Sample
Catcher
#2
Valve #3
Full Bore
Valve #2

Full Bore
Valve #1
Valve
#1

Choke
Bypass
Input
Data
Well
Header
Effluents
Flow
Direction

71
Figure 2-70.

Valve
#4

Integrated flow control and sample catcher manifold (after Lunan and Boote, 1994 12).

The separator vessel should be able to rapidly dump significant liquid volumes, coming from a
large liquid slug. The pressure vessel may need to be fitted with several large, mechanically
activated dump outlets.
The vessel must be large enough to allow adequate retention time for the separation of the solids.
The larger the vessel the better. Originally, the separators used in Canadian underbalanced
drilling operations were sized so that all cuttings generated in the drilling operation could be
contained within the separator itself, without any need to discharge cuttings until total depth (for
the interval in question) was reached. If this is to be done, the solids capacity of the vessel
should be at least three times the volume of the hole to be drilled, to account for bulking of the
fragmented rock and for possible shale swelling.10 Newer separators, designed for underbalanced
drilling, incorporate sealed, auger, screw-type pumps, to remove cuttings from the separator
under pressure.2

2-230

Horizontal or vertical separators may be used. Vertical separators are more effective when the
returns are predominantly liquid, while horizontal separators tend to be able to handle higher gas
volumes more efficiently. It is possible to run vertical separators in parallel, to provide good
operating efficiency. Horizontal separation is better managed with one, large vessel.8
The Energy Resources Conservation Board of Canada (ERCB) has set out design and operation
guidelines for closed systems. These provide a sound basis for designing a closed, underbalanced
drilling system.11 A typical horizontal separator, used in Canadian operations, is shown
schematically in Figure 2-72
2-72. This cylindrical vessel is nine feet in diameter and 40 feet long,
with a working pressure rating of 200 psi. The maximum gas rate for this separator is 35
MMscf/D, with a total liquid capacity of 425 bbl. This pressure vessel is skid-mounted and is
enclosed and heated for cold climate operations.
In this horizontal separator, well returns enter and are slowed by the velocity-reducing baffle.
Solids settle principally in the first compartment, from where they can be removed by the solids
transfer pump. Liquids pass over the partition plate, into the second compartment, where further
solids separation occurs, and where the liquids begin to separate. The liquids spill into the third
compartment, where separation is completed. Liquid hydrocarbons and drilling liquid are
discharged from different levels in this compartment.
The separator vessel should be fitted with adequately sized pressure relief valves and an
emergency shutdown valve, triggered on high liquid level, and high or low vessel pressure. It
should also be fitted, at a minimum, with sight glasses to indicate liquid levels, and with some
means of observing solids levels without having to shut down or enter the vessel. Additional
instrumentation can be beneficial; this will be discussed below.
Well Effluents
In
Adustable Partition Plates
Velocity Reducer
Gas

Gas

Gas Out
Liquids
To Oil Tanks

Hydrocarbons
Drilling Fluid

To Wellhead

Liquids

Solids

Liquids

Soilds Transfer
Pump

Solids

Continuous Pressurized
Solids Transfer Pump

Figure 2-71
72.

A typical, horizontal, four-phase separator, for underbalanced drilling (after Lunan


and Boote, 1994 12).

2-231

Chapter 2

Underbalanced Drilling Techniques

If a closed system is to be used with a drilling fluid that contains any oxygen, it must be
possible to purge the separator vessel with an inert gas, such as nitrogen, in order to prevent
the buildup of a potentially explosive gas mixture in the separator.
Cuttings filter
It is possible to use a three-phase separator, when drilling small diameter holes where limited
volumes of cuttings are generated. This might be the case in coiled tubing drilling operations. To
allow this, a sand trap or filter is installed upstream of the separator vessel.5 This reduces erosion
of the flowline between the sand trap and the separator and of the entry region of the separator
itself.
Heater
Mixtures of low molecular weight hydrocarbons and water can potentially form hydrates at low
temperatures. These could interfere with flow into the separator or with the separators
operation. Normally, well returns will not warm enough to prevent hydrate formation. If there is
a large gas volume fraction in the returns and they are choked back, the decrease of the
temperature downstream of the choke may lead to hydrate formation. In this case, the returns
should be routed through a heater, upstream of the flow control manifold. The returns should
also be passed through a heater if large volumes of heavy oil (with viscosities at low
temperatures that prevent efficient gas separation) are being produced.
Degasser
Gas carryover from the separator to an open mudpump suction tank is potentially dangerous,
particularly if there is any possibility of H2S production. Gas in the re-injected liquid can cause
lower than anticipated pressures downhole, with consequent higher production and safety
concerns. It will also interfere with the efficiency of the mud pumps. All gas has to be
eliminated from any liquid to be re-injected into the well. In most circumstances, it will be
necessary to use an additional degasser, downstream of the primary separator.8 Secondary
degassing is particularly important if high pressures or significant gas dissolution are likely, as
they probably will be when drilling deep or overpressured reservoirs. An atmospheric separator
or vacuum degasser can be used for this second stage of degassing.
Adam and Berry, 1995,5 used closed settling tanks, vented to the flare line, to handle remnant
gas. A secondary degasser is not always used in Canadian closed systems. These wells typically
have low formation pressures. Here, the combination of low pressures, unviscosified liquids, and
long residence times allows an adequate portion of the entrained gas to escape so that the liquid
returned to the mud pumps from the separator vessel can be re-injected without problems.

2-232

Flare Stack/Pit
The flare stack must be large enough to handle the maximum anticipated gas rate (both produced
and injected gas). The height and location of the stack are likely dictated by local regulations
governing natural gas flaring. The flare stack should be equipped with an igniter and pilot flame
system to ensure combustion of any discharged flammable gases.
Saponja, 1995,7 described a separate flare stack and flare pit, with gas from the separator going to
the stack and gas vented from the wellhead casing annulus going to the flare pit.
Louison et al., 1984,13 used two separate flare lines; one twelve-inch diameter and one ten-inch
diameter, with separate six-inch diameter, twenty-foot tall flare stacks, to handle gas produced
while drilling the Midway and Navarro formations in South Central Texas. They reported flares
that were 25- to 30-feet high when drilling, and up to 120 feet high after a long shutdown in
drilling. They also recommended a flare pit at the base of the stack, for burning condensate,
which would not be required or appropriate with a closed system.
A backpressure regulating valve, in the flare line between the separator and the flare stack,
allows the separator pressure to be maintained when the gas production rate is low.
Produced gas does not have to be flared when drilling underbalanced. It is possible to route gas
from the separator to an export line, if there are local production facilities that can handle the
returning gas. This has been done successfully on an underbalanced coiled tubing drilling
operation, onshore in the Netherlands.5
Production Tank
The production tank should be closed to prevent vapor release. Any measurable gas, evolving
from liquid hydrocarbons in the production tank, can be vented to the flare stack.
Water Tank
When the drilling fluid has an aqueous phase, water from the separator will be discharged into
one or two settling tanks, before being transferred to the mud pump suction tank. Since
additional solids separation occurs in these tanks, they should have sloped bottoms and be fitted
with internal risers for water drawoff, to provide the cleanest possible water to the mud pumps.3
Solids Tank
There should normally be a cuttings storage tank to receive solids discharged from the separator,
unless the anticipated volume of cuttings is sufficiently low that the separator will only be
emptied at the end of the job.

2-234

Chapter 2

Underbalanced Drilling Techniques

Instrumentation
Closed surface systems collect all material returning from the well. With appropriate
instrumentation on the closed system and on the drilling fluid injection systems, it is possible to
determine the rates and pressures at which the well is producing hydrocarbons while drilling. It
is also possible to conduct a production test, using the closed system, at any time while drilling,
or before or after completion.
Much of the formation evaluation information, that can potentially be collected when drilling
underbalanced, requires that an integrated instrumentation package is used to record injection
rates, flow rates and pressures. Instrumentation packages for underbalanced drilling were
described in Section 2.7, Gasified Liquids. Separators, used for underbalanced drilling, can be
instrumented to measure gas pressure and temperature, liquid hydrocarbon and water levels (flow
rates), as well as methane, oxygen and hydrogen sulfide contents. The differential pressure,
across an orifice meter in the flare line, will give the gas flow rate.
If the drilling fluid contains any oxygen, measurement of oxygen and methane concentrations
inside the separator is essential. Teichrob, 1994,14 reported using a methane chromatograph,
along with a portable oxygen and lower explosive limit (LEL) meter, to continuously monitor the
flammability of the gas in a closed system separator, when drilling with an air/nitrogen mixture.
2.11.2 Drilling Fluids
Closed systems can, in principle, be used with any underbalanced drilling fluid. The majority of
applications to date have used nitrified liquids.
It is not normally appropriate to use a closed system when drilling with air, air mist, air foam or
an aerated liquid, due to the potential for forming an explosive mixture inside the separator
pressure vessel. It is possible to use a closed system, with a drilling fluid containing oxygen,
provided that the oxygen content of the injected gas is maintained at a level below which an
explosive mixture occurs. Steps necessary to establish a safe oxygen content were described in
Section 2.7, Gasified Liquids. It is necessary to monitor the gas composition inside the
separator and to purge it with inert gas, when necessary, to prevent combustion or explosion.
Using a viscosified drilling fluid, with a closed surface system, should be avoided whenever
possible. Viscosification slows the separation of gas from the liquid returns. This becomes
particularly important if there is any hydrogen sulfide in the produced gas.
Usually, the separator system has to be operated above atmospheric pressure. The
elevated pressure drives liquids from the separator, to the various secondary separation and/or
storage vessels downstream. Particularly for long, horizontal separators, this pressure promotes
efficient separation of the various phases within the vessel. If the well returns contain injected
and/or produced gas, the backpressure from gas flow through the flare line will normally provide
the required separator pressure. It there is little gas in the returns, it is possible to pressurize the
separator vessel directly with nitrogen.4 Alternatively, it is possible inject a small volume of
nitrogen in the drilling fluid at the standpipe, even though this may not be strictly necessary to
achieve underbalanced conditions downhole.7
2-235

It is possible to use a closed system for drilling underbalanced with foam. Defoaming will have
to take place inside the separator vessel(s). MacDonald and Crombie, 1995,15 described a closed
surface system that incorporated both mechanical (a polyurethane hydrocyclone) and chemical
defoaming inside a low pressure separator vessel. This operated satisfactorily, without any foam
carryover from the separator, while drilling nominally balanced with coiled tubing in a shallow
(1,500 ft), low bottomhole pressure (275 psi) gas well and a somewhat deeper (5,875 ft TVD),
higher pressure (740 psi), oil well. However, flow rates and surface pressures were low, in both
of these cases. The liquid discharged from the separator could not be re-injected, because it
contained defoamer. This prevented re-foaming, even with additional foamer.
2.11.3 Operating Procedures
This section describes only those operating procedures that are distinctive to closed systems.
Other procedures, that are specific to each type of underbalanced drilling fluid, are detailed in the
sections describing the techniques used for drilling with those fluids.
 Liquid and solid levels in the separator vessel should be monitored at all times that returns
are flowing into the separator. Liquids and solids should be discharged, to the appropriate
tanks, when necessary, to maintain separator efficiency.
 The liquid and solid levels should be recorded periodically, together with all liquid volumes
discharged, the gas pressure and temperature in the separator, and the gas flow rate down the
flare line. These measurements can be combined with drilling fluid injection rates to
estimate the variation in productivity with depth.
 The separator pressure should be monitored whenever returns are flowing into the separator.
If this pressure approaches the safe working limit of the separator vessel, the well should be
choked back using the flow control manifold, or the drilling fluid density should be increased
to reduce the rate of production. If this fails to keep the pressure sufficiently low, the well
should be shut-in and killed. An adequate volume of kill fluid, typically 1.5 times the hole
volume, should be kept on site, if the ungasified liquid phase of the drilling fluid is
insufficient to kill the well.
 The gas flow through the flare line can be choked back to increase the back pressure on the
separator. This can be
done if the separator pressure becomes too low for efficient liquid discharge or to avoid slug
flow into the separator. This should only be done if the gas flow rate is steady and there is
no possibility of a sudden increase in the gas flow rate. If that occurred, the separator
pressure would rise rapidly, potentially exceeding its safe working limits.
 As was noted in Section 2.7, Gasified Liquids, choking back the well returns should be
avoided if at all possible. This increases the gas compression and liquid injection power
requirements, and, more significantly, has an impact on the bottomhole pressure that is
difficult to determine. By manipulating the choke, it is very easy to inadvertently take the
well overbalanced and damage the producing formation.

2-236

Chapter 2

Underbalanced Drilling Techniques

 Watch for any signs of plugging inside the separator or flare line. If plugging occurs in a
redundant component, switch flow to the backup and clear the blockage. Otherwise, the well
should be shut-in and the plugging remedied immediately.
 If injecting a gas containing oxygen as a component of the drilling fluid, monitor the gas
composition in the separator vessel at all times. Purge the separator with inert gas if the
composition approaches the lower explosive limit.14
 Ensure that the flare is ignited at all times when natural gas is returning from the well. This
is especially important if hydrogen sulfide is present.
 Monitor all flowlines for erosion, paying particular attention at any changes in direction.
Ultrasonic pipe wall thickness meters make this an easy task.
 It may be necessary to enter the separator vessel after a job is complete, for example to clean
out cuttings. This should only be done if essential. All appropriate precautions should be
taken before and during any manned entry.
2.11.4 Limitations
Planning is required before using closed systems when high surface pressures are possible.
Precautions are also required if the drilling fluid is oxygenated. Other limitations include
availability of suitable equipment and personnel, and increased operating costs. Using a closed
system does not remove any of the specific limitations associated with the drilling fluid (other
than air) or the technique adopted.
High Surface Pressures
Properly designed and operated closed systems can handle high gas production rates, in excess of
35 MMscf/D. To date, closed systems have only been used in shallow or modestly deep, subnormally pressured reservoirs, that do not have the potential for very high surface pressures.
Deep, overpressured reservoirs can generate surface pressures that are much higher than available
separators can tolerate. The separators are normally exposed only to the back pressure from the
gas flow through the flare line. Plugging of the separator or flare line, for example, could expose
the separator vessel to pressures higher than its design rating.
Extreme caution should be exercised before using a closed system to drill a deep or overpressured gas reservoir, under underbalanced conditions. It is essential that a thorough safety
review is conducted before committing to the use of a closed system. In addition to the hazards
common to other drilling techniques, this review should evaluate all possible surface pressure
regimes, identify all hazardous conditions, and devise means to eliminate, or to handle, any
potential condition that might expose any surface system components to pressures in excess of
their safe working limits.
Oxygen Containing Drilling Fluids
The potential for creating an explosive gas mixture inside the separator, when there is oxygen in
the drilling fluid, has already been noted. Closed systems should not be used with dry air, air
mist, air foam, or aerated liquid drilling fluids.
2-237

Equipment and Personnel Availability


Currently, there are a number of customized, closed systems available for rental in Canada;
together with crews trained and experienced in their operation and maintenance. Few systems
and trained crews are available in the United States at this time.
Operating Costs
The additional surface equipment associated with a closed system, and, the additional personnel
required to operate and maintain that equipment, lead to higher daily costs than would be
incurred with other underbalanced or conventional drilling techniques. Additional costs will also
result with mobilization and demobilization of the equipment and experienced crews. Since
neither the equipment nor the crews for closed system underbalanced drilling are readily
available in the United States, these costs will probably be significant.
Any realistic economic evaluation of a closed system will be well specific. Closed system
drilling may offer real merit if the formation being drilled is sensitive to formation damage and is
likely to be very productive if it is drilled underbalanced.

Return to SC Manual

2-238

Chapter 2

Underbalanced Drilling Techniques

References

2-239

1.

Deis, P.V., Churcher, P.L., Turner, T. and Curtis, F.: Horizontal Underbalanced Drilling
Techniques used in the Mississippian Midale Beds of the Weyburn Field, paper 93-1105
presented at the 1993 CADE/CAODC Drilling Conference, Calgary, Alta.

2.

Lunan, B.: Underbalanced Drilling - Surface Control Systems, paper HWC94-20


presented at the 1994 SPE/CIM/CANMET International Conference on Recent Advances
in Horizontal Well Applications, Calgary, Alta.

3.

Deis, P.V., Yurkiw, F.J. and Barrenechea, P.J.: The Development of an Underbalanced
Drilling Process: An Operators Experience in Western Canada, presented at the 1995
International Underbalanced Drilling Conference, Amsterdam, The Netherlands.

4.

Crerar, P.: Underbalanced Re-entry Horizontal Drilling in the Welton Field Basal
Succession Reservoir Onshore UK, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

5.

Adam, J. and Berry, M.: Underbalanced Coiled Tubing Sidetrack Successful, Oil and
Gas J. (December 18, 1995) 91-98.

6.

Surewaard, J., de Koning, K., Kool, M., Woodland, D., Roed, H. and Hopmans, P.:
Underbalanced Operations in Petroleum Development Oman, paper presented at the 1995
1st International Underbalanced Drilling Conference, The Hague, The Netherlands, October
2-4.

7.

Saponja, J.: Engineering Considerations for Jointed Pipe Underbalanced Drilling, paper
presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

8.

Stone, C.R.: History and Development of Underbalanced Drilling in the USA, paper
presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

9.

Comeau, L.: Underbalanced Drilling: Directional and MWD Experience, paper presented
at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

10.

Northland Production Testing Limited, Underbalanced Drilling Division, Promotional


Literature (1994).

11.

Energy Resources Conservation Board: Interim Directive ID 94-3, Calgary, Alta (July
18, 1994).

12.

Lunan, B. and Boote, K.S.: Underbalanced Drilling Techniques Using a Closed System to
Control Live Wells - Western Canadian Basin Case Histories, presented at the 1994
SPE/CIM Annual One Day Conference, Calgary, Alta.

13.

Louison, R.F., Reese, R.T. and Andrews, J.P.: Case History; Underbalance Drilling the
Midway and Navarro Formations Successfully in Hallettsville, TX, paper SPE 13112
presented at the 1984 SPE Annual Technical Conference and Exhibition, Houston, TX.

14.

Teichrob, R.R.: Low Pressure Reservoir Drilled with Air/N2 in a Closed System, Oil and
Gas J. (March 21, 1994) 80-90.

15.

MacDonald, R.R. and Crombie, D.L: Balanced Drilling with Coiled Tubing, paper
IADC/SPE 27435 presented at the 1995 IADC/SPE Drilling Conference, Dallas, TX.

2-240

BENEFITS OF DRILLING
UNDERBALANCED

Underbalanced drilling can offer certain advantages over conventional drilling techniques:
 Increased penetration rate,
 Increased bit life,
 Reduced probability of differential sticking,
 Minimized lost circulation while drilling,
 Improved formation evaluation and reduced costs of testing,
 Reduced formation damage,
 Earlier production,
 Environmental benefits,
 Improved safety,
 Increased well productivity, and,
 Less need for stimulation treatments.
The extent to which it is possible to achieve any of these various benefits generally depends on
the specific application. Reasons why these benefits can occur are described. Field examples are
given; indicating, where possible, the significance of the various benefits that can be expected for
different applications.

3.1

Penetration Rate

One of the most significant and often quoted benefits of underbalanced drilling is increased
penetration rate. It is almost invariably true that switching from conventional, overbalanced
drilling to an underbalanced drilling technique will increase the instantaneous penetration rate.
The change in penetration rate will vary for different underbalanced drilling techniques and in
different rock types. Examples are given of the order of magnitude of change in penetration rate
that can be achieved with different underbalanced drilling techniques. If conventional drilling
gives high penetration rates or if the interval to be drilled is relatively short, drilling
underbalanced may not save enough rig time to pay for the additional equipment involved.
3-1

Chapter 3

Benefits of Drilling Underbalanced

Why Should Penetration Rate Be Higher?


It is often stated that reducing the pressure differential across the hole bottom increases the
penetration rate. Switching from overbalanced to underbalanced drilling changes the differential
pressure from positive to negative, so that a penetration rate increase should be expected. This is
an over-simplification. It does not address some fundamental aspects of how downhole pressures
influence drilling response in permeable rocks and shales.
Permeable Rocks
In permeable rocks, a positive pressure differential, between the drilling fluid in the wellbore and
the pore fluid adjacent to the workfront, reduces penetration rates in two ways.
 This differential pressure represents an effective confining stress acting on the rock being
fractured by the bits teeth.1,2 In most sandstones and carbonates, increasing the effective
confining pressure increases the shear strength of the rock. The assertion that higher
differential pressure increases the effective mean stress in permeable rocks needs to be
viewed with caution, depending on the specific poroelastic-plastic properties of the rock.
Regardless, shear stress, promoting failure, increases with underbalance.
 Cracks that form when the rock fractures will very rapidly fill with pore fluid at or near the
original pore pressure. The difference between this pressure and that of the fluid in the
borehole clamps the rock fragments against intact rock until communication with the
wellbore can equalize the pressures around these fragments and relieve the so-called chip
hold-down forces.2 If the bit teeth contact the workfront before the chip hold-down forces
have been overcome, work will be expended unnecessarily in re-grinding the rock fragments
that remain clamped to the workfront. Increasing the mud weight will increase the chip holddown forces, making the bits action less efficient and decreasing the penetration rate.
There is supporting physical evidence. Both micro-bit2,3 and full scale4 drilling experiments have
shown that (in overbalanced situations) the penetration rate in permeable rocks decreases as the
difference between the borehole pressure and the pore pressure increases. These measurements
have shown that the absolute value of the wellbore pressure in permeable rocks does not
influence penetration rate to any detectable extent; it is the difference that is important. Figure 31 shows one example of the effect of differential pressure on penetration rate, for a micro roller
cone bit, drilling Indiana limestone with an unweighted, water-based mud. With a positive
differential pressure of 1,000 psi, the penetration rate was only 30 percent of that seen at balance.
Figure 3-2 (Moore, 19745) is a composite of wells drilled with mud compared with a composite
of wells drilled with gas. The figure shows that, below a certain depth, drilling with gas was over
six times faster than drilling with mud.
When drilling at balance, there is no longer any (compressive) effective confining stress acting
on the rock being drilled nor any static chip hold-down force, favoring more rapid drilling.
Underbalance is hypothetically even more favorable.

3-2

It is the difference between the borehole pressure and the pore pressure, adjacent to the
workfront, that controls the penetration rate.4,6 This local pore pressure is not necessarily the
same as the undisturbed far-field pore pressure. Drilling fluid filtrate (and in some cases, whole
mud) will flow into the penetrated formations, if there is an overbalance. The hydraulic driving
forces for fluid flow will act from the formation
12
Indiana Limestone
Confining Pressure = 6000 psi
Bit Weight = 1000 lbm
50 rpm

Drilling Rate (ft/hr)

10

0
0

1000

2000

3000

4000

5000

Overbalanced Differential Pressure (psi)

Figure 3-1.

Micro-bit test data showing the influence of differential pressure on


3
penetration rate (after Cunningham and Eenink, 1959 ).

into the borehole, if it is underbalanced. In either case, fluid flow can change the pore pressure
close to the workfront from its original value. Figure 3-3 shows penetration rates (under
overbalanced conditions), measured at two different weights on bit, in laboratory drilling tests. A
7 7/8-inch diameter insert bit was used to drill 200 md Berea sandstone. The pressure
differential across the workfront was determined in these tests, and the penetration rates are
plotted versus the local differential pressure in Figure 3-3. An overbalanced pressure
differential of 500 psi reduced the penetration rate to between 30 and 45 percent of its value at
balance.
Pressure differentials, in normal drilling operations, vary from application to application. They
are usually on the order of a few hundred psi. At a depth of 6,600 feet, a mud overbalance of 0.5
ppg (an equivalent circulating density of 0.5 ppg higher than the formation pressure gradient),
corresponds to a differential pressure of approximately 165 psi. The laboratory drilling data
shown in Figures 3-1 and 3-3 indicate that the penetration rate might increase by 30 to 40 percent
in going from 165 psi overbalanced to balanced.

3-3

Chapter 3

Benefits of Drilling Underbalanced

As the borehole pressure is reduced below the pore pressure, i.e., as the level of underbalance
increases; the rock at the workfront moves into a state of effective axial tension. Chip holddown forces, in permeable rock at least, are eliminated. As a result, the penetration rate is
expected to increase, with increasing underbalance, but its sensitivity to pressure changes should
be lower then if it was overbalanced.
Drilling Days
0

20

40

60

80

100

120

1000

Drilled With Mud


Drilled With Gas

2000

3000

Depth (feet)

4000

5000

6000

7000

8000

9000

10000

Figure 3-2.

Gas drilling versus mud drilling (after Moore, 1974 ).

There is little experimental data, from laboratory drilling tests, to quantify the impact of
underbalance pressure on penetration rate. One set of experiments,3 with a micro (1.25-inch
diameter) roller cone bit, showed that the penetration rate in Berea sandstone increased by only
20 to 25 percent, in going from balanced to between 1,000 and 4,000 psi underbalanced.
Another set of micro-bit experiments,7 in Indiana limestone drilled at a 430 psi underbalance,
penetrated at a rate that was apparently 10 percent lower than when there was a 70 psi
overbalance.

3-4

Impermeable Rocks
The pore pressure in any fractures created during drilling in impermeable zones does not rapidly
equilibrate to reservoir pressure levels. There is an increase in effective porosity because of these
cracks. This is known as dilatancy. If the fluid pressure from the reservoir cannot rapidly enter
these cracks, suction occurs, and the pressure in the fractures can approach zero. In this situation
the chip hold-down forces will be proportional to the borehole pressure, compared to the
difference between the borehole and pore pressures for a more permeable medium. When fluid
from the wellbore has been able to flow into the chip-forming cracks, allowing the pressure in the
fractures to equalize with the drilling fluid above the workfront, the chip hold-down forces
decrease to zero. This argument suggests that penetration rates in hard shales, and other hard
impermeable rocks, should be determined by the absolute wellbore pressure and not by the
difference between wellbore and pore pressures. As with permeable rocks, effective radial
stresses will increase due to the stress concentration at the bottom of the hole, but these stresses
are relatively independent of the degree of balance. However, the variation in effective mean
stress is largely governed by the potential for dilatancy, depending more strongly on the
differential pressure.
The situation in other shales, particularly soft younger formations is more difficult to resolve.
Some measurements have shown that in certain softer or smectitic shales, ROP depends on
the differential pressure.
Zijsling, 1987,8 performed single cutter tests on medium-hard, illitic Mancos shale and on soft,
montmorillonitic Pierre shale. the total bottomhole pressure rather than the overbalance (and
thus the pore pressure in the sample) governs the cutting process of a PDC in Mancos shale. In
the Pierre shale tests, the magnitude of the pore pressure and the mud pressure were both
important. Similar observations were alluded to by Gray-Stephens et al., 1994.9 Regardless of
the specific mechanism, evidence indicates increased penetration rate with an increased
underbalanced pressure differential.

3-5

Chapter 3

Benefits of Drilling Underbalanced

100
90

15000 lbm
30000 lbm

80

ROP (ft/hr)

70
60
50
40
30
20
10
0
0

100

200

300

400

500

600

700

800

900

1000

Pressure Drop Through Filter Cake (psi)

Figure 3-3.

Penetration rate, as a function of the differential pressure (overbalanced)


across the workfront; from full-scale, laboratory, drilling test data (after
Black et al., 1985 4).

Maurer, 1966,10 showed the increase in drilling rate with reduced borehole pressure, during
microbit tests on two shales, a dolomite and a quartzite. Recent laboratory experiments 9 have
confirmed this, at least for hard shales. Figure 3-4 shows penetration rates for a milled tooth bit,
drilling a hard Jurassic shale, at several different combinations of borehole and pore pressures.
The penetration rate decreased with increasing borehole pressure, but it did not change if the pore
pressure was changed at constant borehole pressure. Compare Figures 3-4 (a) through (c). From
a mechanics point-of-view, this is completely consistent with a poroelastic constant approaching
zero, characteristic of a low permeability medium.
Figure 3-5 shows the results of several shale drilling experimental programs.9,11 In this figure,
the penetration rates are incorporated into a normalized drilling strength index, DSn:
WOB RPM

ROP

Pb
DS n =

WOB
RPM


ROP
Po

3-6

(3.1)

where:
WOB.............. weight on bit (lbf),
RPM .............. rotary speed (rpm),
ROP ............... rate of penetration (ft/hr),
P..................... pressure (psia),
subscript b .... indicates borehole
conditions, and
subscript o ..... indicates atmospheric
conditions.
This drilling strength is equal to one at atmospheric pressure. A value of five indicates that the
penetration rate, at a specific borehole pressure, is one-fifth of the penetration rate at atmospheric
pressure (i.e., the higher the index, DSn, the lower the penetration rate). The different shales in
this figure all showed similar increases in normalized drilling strength, with increasing borehole
pressure. At 2,900 psi borehole pressure, the normalized drilling strengths were all
approximately six; indicating that the penetration rate, at this borehole pressure, would be onesixth of the rate at atmospheric borehole pressure, with the same weight on bit and rotary speed.
These results show that a small reduction in borehole pressure only causes a small increase in
penetration rate, if the borehole pressure is significantly greater than atmospheric. Consider an
example:
 A well is being drilled with an unweighted mud (8.5 ppg), at a depth of 6,600 feet.
 The borehole pressure will be approximately 2,900 psi, neglecting the annular pressure drop.
 A reduction in effective mud weight to 7 ppg would reduce the borehole pressure by 500 psi,
to 2,400 psi. This would decrease the drilling strength, i.e. increase the penetration rate, by
less than 15 %.
 It would be necessary to reduce the borehole pressure to approximately 1,500 psi before the
penetration rate would be doubled. This degree of pressure reduction would be close to the
lowest achievable with a gasified liquid, and would be typical of stable foam drilling

3-7

Chapter 3

Benefits of Drilling Underbalanced

150
Pore Pressure ... 87 psi
Pore Pressure ... 508 psi

Rate of Penetration (ft/hr)

125

100

75

50

25

0
0

5000

10000

15000

20000

25000

30000

35000

40000

45000

50000

Downhole Weight on Bit (lbf)

(a)
150
Pore Pressure ... 580 psi
Pore Pressure ... 870 psi
Pore Pressure ... 116 psi

Rate of Penetration (ft/hr)

125

100

75

50

25

0
0

5000

10000

15000

20000

25000

30000

35000

40000

45000

50000

Downhole Weight on Bit (lbf)

(b)
Figure 3-4.

3-8

Penetration rate in a Jurassic shale, drilled with an 8-inch milled tooth


bit; as a function of weight on bit, for different pore pressures (Pp); (a) 440
psi borehole pressure; (b) 1450 psi borehole pressure; (after GrayStephens et al., 1994 9) (continued)

150
Pore Pressure ... 2320 psi
Pore Pressure ... 4930 psi

Rate of Penetration (ft/hr)

125

100

75

50

25

0
0

5000

10000

15000

20000

25000

30000

35000

40000

45000

50000

Downhole Weight on Bit (lbf)

(c)

Figure 3-4. Penetration rate in a Jurassic shale, drilled with an 8-inch milled
tooth bit; as a function of weight on bit, for different pore pressures (Pp); (c) 4800 psi
borehole pressure (after Gray-Stephens et al., 1994 9).

Normalized Rock Drilling Strength, DSn

10
DTM/Jurassic-2 Shale

SDM/Welsh Shale
Results from Cheatham et al.

SDM/Jurassic-2 Shale

7
6
5
4
3
2
1
0
0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

Bottomhole Pressure (psi)

Figure 3-5. The influence of bottomhole pressure on the normalized drilling strength in
hard shales (after Gray-Stephens et al., 1994 9).

.
3-9

Chapter 3

Benefits of Drilling Underbalanced

A borehole pressure of around 100 psi might be expected if dry air drilling. With such a low
borehole pressure, a penetration rate, as much as five times that seen when drilling with mud,
could be anticipated (on the basis of the laboratory data in Figure 3-5).
Additional Factors
As indicated, permeability influences how underbalance pressure can affect penetration rate. Bit
type, bit hydraulics and drilling fluid rheology are also important.
Roller cone bits rely on the bit hydraulics to remove cuttings from the workfront, before the bit
teeth again strike the same location on the workfront. Conversely, for bits that cut with a
shearing action, such as natural diamond, Thermally Stable Polycrystalline (TSP) diamond, and
Polycrystalline Diamond Compact (PDC) bits, the bits cutters push the cuttings ahead, across
the workfront. With different drilling action, individual bits may respond differently to varying
the magnitude of the differential pressure.
The specific influence of a bit also depends on the clearance between the bit body and the
workfront. When the bit is on bottom, many natural diamond and TSP bits have restricted
clearance between the bit body and the workfront. The drilling fluid pressure drops as it flows
across the workfront and up into the annulus. The pressure over the central portion of the
workfront can easily be more than 500 psi greater than in the annulus.12 This means that the
pressure differential experienced by the rock, as it is being drilled by a natural diamond or TSP
bit, may be greater (tending to be more overbalanced) than it would be when drilling with a roller
cone or PDC bit. In the extreme case, these bits might even generate locally overbalanced
conditions, even when the annulus pressure is underbalanced. How this affects penetration rate
is not specifically known.
Roller cone and most PDC bits provide minimal flow restriction to drilling fluid once it has
passed through the bit nozzles. For these bits, the drilling fluid pressure above the workfront is
more or less equal to that in the annulus.
All drill bits will not be characterized by the same relationship between underbalance pressure
and penetration rate. There are no laboratory data to confirm this. However, it has been shown
that natural diamond bits undergo a greater fractional reduction in penetration rate, with
increasing overbalance, than do roller cone bits.6 Drag bits showed slightly smaller reductions in
penetration rate than roller cone bits, with increasing overbalance.6
If the hydraulics of the entire system are considered, removing cuttings from the hole bottom is
not necessarily easier during underbalanced drilling. For example:
 Underbalanced drilling with gases, mist, and foam tends to require less bit hydraulic
horsepower than would be used in conventional drilling,
 If bit nozzles are used, sonic flow can modulate annulus conditions, and,
 Drilling fluid densities are lower.
3-10

Cuttings removal from the workfront may be less efficient with the lower density drilling
fluids. This might lead to smaller increases in penetration rates than would be expected from
the borehole pressure reduction alone.
Numerous drilling fluid properties can influence the penetration rate. It can be very difficult to
change one property of a drilling mud without changing other properties. Nevertheless, viscosity
likely has an important affect on penetration rate - higher viscosity results in lower penetration
rates, all other factors being equal.13,14 Drilling fluid viscosity will influence the rate at which
fluid can flow into any chip-forming cracks and the rate at which chip hold-down forces relax.6
It may even be possible that drilling fluid viscosity can favorably influence penetration rates in
low permeability rock that is drilled underbalanced. For example, gas or mist drilling might give
higher penetration rates in shales than techniques using foams or liquids, even if the borehole
pressures were similar. High viscosities for stable and stiff foams may cause lower penetration
rates than would occur for a lower viscosity, gasified liquid, at a similar borehole pressure.
There is no laboratory data to confirm or disprove this hypothesis.
Drilling fluid viscosity should have less influence on penetration rates in permeable formations.
Pore fluid from the undamaged rock will rapidly infiltrate chip-forming cracks and there will be
reduced driving force for drilling fluid to flow into these cracks.

3.1.2 Field Examples


Figure 3-6 shows depth versus time curves for three wells, all drilled in the same area. One well
was drilled exclusively with dry air, while the other wells had to be mudded up before reaching
total depth because of hole problems. This figure clearly shows that switching to mud gave
much lower penetration rates.
Figure 3-7 is a plot of bit weight and penetration rate versus depth for Well 1 in Figure 3-6. A
hammer bit was used to about 6,800 feet, with bit weights slightly greater than 5,000 lbf.
Hammer bits generally operate at weights much lower than those used in rotary drilling - hammer
bits are discussed in more detail in Chapter 6, Section 6.4, Percussion Drilling. The
penetration rate with the hammer was 55 to 65 ft/hr. Rotary drilling with air was used from
6,800 feet to total depth. Bit weight was increased to 20,000 lbf with the rotary bits, and the
penetration rate was about 30 ft/hr.
Figure 3-8 shows the equivalent weight and penetration rate data for Well 2. Hammer bits were
used to 6,000 feet. Rotary bits (still with dry air) were used to 8,100 feet. The well was mudded
up at 8,100 feet and the penetration rate fell from over 20 ft/hr, with 10,000 lbf bit weight, to
approximately 5 ft/hr, with 40,000 lbf bit weight.
Figure 3-9 is a plot of bit weight and penetration rate versus depth for Well 3. This well was
mudded up below 6,700 feet, due to hole problems. When drilling with mud, the penetration rate
dropped to about 5 ft/hr (as in Well 2), even with maximum bit weight. Despite a four-fold
increase in bit weight, the penetration rate with mud was less than one-third of the rate during air
3-11

Chapter 3

Benefits of Drilling Underbalanced

drilling. If the bit weight for air drilling is normalized to 40,000 lbf, a penetration rate of 80 ft/hr
might be expected. For this example, the penetration rate for air drilling could be predicted to
be as much as sixteen times that for mud drilling at the same weight on bit.
3000
Well 1
Well 2
Well 3

4000

Depth (feet)

5000
Switch to Mud

6000

7000

8000

9000
5

10

15

20

25

30

Days

Figure 3-6.

3-12

Depth versus days, for three wells drilled in the same area. Well 1 was air
drilled to TD. Wells 2 and 3 had to be mudded up because of hole
problems.

80
Hammer Bits (WOB)
Rotary Bits (WOB)
Penetration Rate

Bit Weight (lbf)

35,000

70

30,000

60

25,000

50

20,000

40

15,000

30

10,000

20

5,000

10

Penetration Rate (ft/hr)

40,000

5000

5500

6000

6500

7000

7500

8000

8500

Depth (feet)

Figure 3-7.

Bit weight and penetration rate for Well 1.

40,000

80
Hammer Bits (WOB)
Rotary Bits (WOB)
Penetration Rate

Bit Weight (lbf)

Rotary
Bits

Hammer
Bits

30,000

70
60

25,000

50

20,000

40

Switch to Mud

15,000

30

10,000

20

5,000

10

Penetration Rate (ft/hr)

35,000

0
0

1000

2000

3000

4000

5000

6000

7000

8000

9000

Depth (feet)

Figure 3-8.

Bit weight and penetration rate for Well 2.

3-13

Chapter 3

Benefits of Drilling Underbalanced

40,000
Hammer Bits (WOB)
Rotary Bits (WOB)
Penetration Rate

70

30,000

60

25,000

50

20,000

40
Switch to Mud

15,000

30

10,000

20

5,000

10

0
5000

Penetration Rate (ft/hr)

Bit Weight (lbf)

35,000

80

0
5500

6000

6500

7000

7500

8000

8500

Depth (feet)

Figure 3-9.

Bit weight and penetration rate for Well 3.

There are a number of published case histories, documenting penetration rate increases occurring
after switching from conventional to underbalanced drilling, or comparing wells drilled
underbalanced with adjacent wells drilled conventionally. For example:
 Pratt, 1989,15 compared penetration rates for air and mud drilled wells, in several fields in the
Canadian foothills and Rockies:

In the Jumping Pound Field, rotary air drilling from surface casing to 7,500 feet gave an
average overall penetration rate of 312 ft/day, whereas liquid drilling with water or mud
had given an average penetration rate of 115 ft/day over the same interval.

Two air drilled wells, in the Arrowhead area, averaged more than 600 ft/day over an
interval from 5,000 to 8,000 feet, where mud drilling was less than 220 ft/day.

Even higher penetration rates were reported when an air-percussion hammer was used.

 Lyons et al., 1988,16 reported on a shallow, 10 5/8-inch diameter hole, drilled in the Kirtland
Formation, in New Mexico. This was drilled with an insert bit; using mud from 273 feet to
323 feet and dry air from 323 feet to 373 feet. Various weights on bit were run, with both
drilling fluids. The average penetration rate with air was twice that with mud.
 Drilling with natural gas can result in penetration rate increases similar to those seen with dry
air. In the San Juan Basin, New Mexico, natural gas drilling, through Cretaceous sandstones
and shales, averaged from 40 to 75 ft/hr. Mud drilling resulted in between only 6 and 20
ft/hr.17

3-14

 Mitchell and Salvo, 1991,18 provided data comparing penetration rates for air mist and mud
drilling, through normally pressured and depleted sandstones and dolomites. These
formations were between depths of 2,700 and 4,100 feet, in the Eunice Monument Field, in
southeastern New Mexico. Air mist gave an average penetration rate of 57 ft/hr. Mud gave 28
ft/hr. Drilling with mist doubled the penetration rate seen with mud.
 Foam drilling penetration rates can also be higher than for mud drilling. Fraser and Moore,
1987,19 reported that, drilling through permafrost, stable foam gave penetration rates that
were twice those seen in adjacent mud drilled wells. Another permafrost application saw
stable foam giving penetration rates of up to three times those for unweighted mud.20
 Hutchinson and Anderson, 1972,20 discussed a well in Pecos County, Texas, where stable
foam drilling matched the best offset dry air penetration rates for 20-inch hole, to 1,000 feet.
Foam then gave an average penetration rate of 320 ft/day for 14-inch hole, to 5,800 feet.
Previously, the best average penetration rate over this interval, with an unweighted brine
mud, was 110 ft/day.
 Hutchinson, 1970,21 cited a number of fractured reservoir applications where stable foam
drilling gave penetration rates greater than those in adjacent wells, drilled conventionally
with mud.

Penetration rates of 30 to 75 ft/hr were seen in the Delaware limestone in Texas, where
brine penetration rates averaged 15 ft/hr.

A fractured schist, in Kern County, California, was drilled at rates from four to eight
times greater than that normally achieved with clay-based muds.

Miocene cherts, near Santa Maria, California, were drilled with stable foam at two to four
times normal mud penetration rates.

 Wolke at al, 1990,22 presented data that showed the instantaneous penetration rate increased
from around 18 ft/hr to more than 40 ft/hr when they switched from straight mud to aerated
mud, at a depth of 5,000 feet in a geothermal well.
 Comeau, 1995,23 reported that the instantaneous penetration rate, in the first horizontal well
drilled underbalanced in Canada, increased from 32 ft/hr to 137 ft/hr, when nitrogen injection
was started.
 Drilling with nitrified water gave penetration rates of up to six times those with mud, in a
horizontal well, in the Hussar field, South Central Alberta.24 The cost per foot with
underbalanced drilling was almost one-half of that for conventional mud drilling. This
glauconitic sandstone reservoir was depleted; the conventional mud gave a differential
pressure of 1,750 psi.
 Data from the Weyburn field in Canada25 showed that drilling horizontal wells underbalanced
with nitrified water (through a soft, chalky dolostone reservoir) gave smaller penetration rate
increases over mud drilling than those reported above. Horizontal sections drilled
underbalanced averaged 80 ft/hr, whereas a starch-based, freshwater mud averaged 60 ft/hr.
Typically, connections take longer with nitrified liquids than with mud if underbalanced
3-15

Chapter 3

Benefits of Drilling Underbalanced

conditions are to be maintained. The instantaneous penetration rate probably increased by


more than the average penetration rate.
These field examples show that it is possible to reduce, dramatically in some instances, the total
time to drill a full hole interval by drilling underbalanced. These improvements were mainly
achieved by experienced crews. An inexperienced crew will inevitably take longer to make
connections and trips and may not be able to achieve the same overall drilling time reduction that
can be achieved by an experienced crew.

3.1.3

Practical Significance

Instantaneous penetration rates will be higher underbalanced than overbalanced. Drilling with a
gas, mist or foam will give borehole pressures that are much lower than those with mud in the
hole. These underbalanced drilling techniques result in penetration rates that, in most
circumstances, are at least twice those for drilling overbalanced with mud. Higher penetration
rate increases, three-fold or more, can often be achieved when it is possible to drill with a dry
gas. Gasified liquids and flowdrilling have borehole pressures that are often controlled within a
few hundred psi underbalanced. The penetration rate increase, for these underbalanced drilling
techniques, may be less than two times, particularly when drilling shales or other low
permeability rocks. On the basis of the magnitude of the differential pressure alone, penetration
increases using underbalanced drilling should be relatively largest in depleted zones or
reservoirs.
The instantaneous penetration rate is the rate at which the bit advances when it is on bottom.
Connection times may be longer than when drilling conventionally. The difference in connection
times will be slight for an experienced crew, drilling with dry air, mist or foam. It may be
significant when using drillstring gasification with a gasified liquid drilling fluid. Round trip
times may also be increased when drilling underbalanced. When drilling with dry air, time is lost
after each trip if the wellbore has to be dried out before drilling can resume. Time will also be
lost if special steps have to be taken to limit the change in bottomhole pressure while tripping, as
is usually the case when drilling with a gasified liquid using drillstring injection. Any additional
time for connections or tripping reduces the benefit of increased instantaneous penetration rates.
Fewer trips, due to increased footage per bit, may more than offset increased tripping time.
The impact of extending time for connections or drilling on the overall drilling time depends on
the instantaneous penetration rate. Some simplistic examples illustrate this. In these examples,
the fractional changes in penetration rate, as well as the connection and tripping times, have been
chosen for illustrative purposes only. While they are not unrepresentative of what might actually
be seen in typical operations, they should not be arbitrarily used for economic analysis of a
potential underbalanced drilling application.

3-16

Suppose that a 3,000 foot long interval is to be drilled, from 6,000 to 9,000 feet.
 Drilling conventionally, connections take 5 minutes and the instantaneous penetration rate
averages 5 ft/hr. One hundred connections have to be made, which take 8.3 hours altogether.
The trip into the hole, at the start of the interval, will take 3 hours and the trip out, at the end
of the interval, will take 4.5 hours. Time on bottom will dominate this interval, even
assuming that no time is lost to unscheduled events. The 3,000 feet will require 600 hours on
bottom. The total time taken to drill the interval will be (600 + 8.3 + 3 + 4.5) = 615.8 hours.
 Suppose that the same interval is drilled with dry air and that the instantaneous penetration
rate increases to 10 ft/hr. But, suppose that both connection and trip times are increased by 50
percent. The net result is a substantial reduction in time to drill the interval; to 331.6 hours.
Under virtually all circumstances, this time reduction will more than repay for any additional
daily costs associated with underbalanced drilling.
Consider the same length of interval in a softer, more permeable formation.
 Presume that this formation can be drilled conventionally, at an instant-aneous rate of 100
feet per hour. In this case, the total time for the interval would be (30 + 8.3 + 3 + 4.5) = 45.8
hours.
 Drilling with a gasified liquid, the instantaneous penetration rate increases to 150 feet per
hour, but the connection and trip times are doubled. The total time for the interval is now (20
+ 16.6 + 6 + 9) = 51.6 hours. Since the penetration rate was so high when drilling
conventionally, the interval actually takes longer to drill underbalanced. A reduction in
formation damage may still repay the higher daily cost of underbalanced drilling, but in this
case the increased instantaneous penetration rate will not.

3.2

Bit Life

It is often claimed 23, 24, 26, 27 that bit life is increased by drilling underbalanced. If actual rotating
hours on a bit are increased, this has not been well documented. However, and perhaps more
importantly, it does appear well established that with underbalanced drilling, fewer bits are
required to drill comparable intervals underbalanced.
3.2.1 Roller Cone Bits
Bearing Wear
Before the introduction of sealed bearing bits, the number of hours that a bit could survive
downhole may well have been increased by drilling with dry air. The air would be more or less
free from the abrasive particles found in most drilling muds, and, open bearings could operate for
longer with air flowing through them, than with mud. Roller cone bits now almost invariably
have sealed bearings. Any benefit from the cleaner drilling fluid is lost. Using sealed bearings
has increased the number of rotating hours that can be accumulated before bearing wear becomes
excessive, irrespective of the drilling fluid in use.28

3-17

Chapter 3

Benefits of Drilling Underbalanced

Vibration levels downhole are usually higher when drilling with lightened drilling fluids.
Frequent failures of surveying equipment, such as MWD units and steering tools, attest to the
severe vibrations that are often experienced when drilling with dry air.29 As noted by Kelly,
1990,30 rough running causes momentary overloads that promote localized seizures of the bearing
surfaces, accelerating the wear process. As a result, bit bearing life, in hours or rotations to
failure, may actually be decreased slightly by drilling underbalanced, particularly for dry air
drilling in hard rock.
Cutting Structure
Bit life may be defined, either in terms of the hours downhole, or, in terms of footage drilled per
bit. Penetration rates are often higher when drilling underbalanced than when drilling
conventionally. A bit could last fewer hours underbalanced but still drill more footage than it
would if it was run under overbalanced conditions. The number of bits required to drill an
interval will be inversely proportional to the footage drilled by each bit. Consider the extent to
which drilling underbalanced can increase bit footage rather than just the number of hours
each bit can survive downhole.
The mechanical specific energy, MSE, is defined as the mechanical work that has to be done to
excavate a unit volume of rock.31 In rotary drilling, work is done both by the torque required to
rotate the bit and by the weight on bit. The work done by the bit is:
W = 2 +

WOB ROP
60 RPM

(3.2)

where:
W ....... work done by the bit (ft-lbf/
revolution),
......... torque (ft-lbf),
WOB.. weight on bit (lbf),
ROP ... rate of penetration (ft/hr), and,
RPM .. revolutions per minute.
The volume of rock excavated per revolution is:
V=

d 2b ROP
240 RPM

where:
V ........ volume of rock excavated per
revolution (ft3), and,
db ....... bit diameter (feet).

3-18

(3.3)

The mechanical specific energy is given by:


MSE =

480 RPM 4WOB


+
d 2b ROP
d 2b

(3.4)

where:
MSE .. mechanical specific energy (psi).
Bit torque does not normally change substantially with changing borehole pressures. As
discussed, penetration rates generally increase with decreasing borehole pressures. Mechanical
specific energies are, therefore, usually lower at lower borehole pressures. Laboratory drilling
data illustrate this trend. Gray-Stephens et al., 1994,9 showed an example, in a Jurassic shale,
where a roller cone bits torque increased by 40 percent, when the borehole pressure was
decreased from 4,800 to 450 psi. At the same time, the penetration rate increased by
approximately 525 percent. The net effect was that the mechanical specific energy decreased by
more than 75 percent.
In most cases, cutting structure wear rates (expressed in terms of the distance drilled) should be
inversely related to the mechanical specific energy. If the bit has to do less work to remove a
given volume of rock, its cutting elements should wear less and a bit should be able to drill
more footage (remove more rock), when drilling underbalanced than when drilling overbalanced.
There are few published examples of field data that provide unambiguous confirmation of this.
Mitchell and Salvos penetration rate comparison, for wells drilled with air mist and with mud, in
the Eunice Monument field,18 has already been mentioned. In that instance, the interval from
1,200 feet to 4,100 feet could be drilled with one bit using air mist, whereas mud drilling
required two bits. Comeau, 1995, 23 reported tricone bit runs in excess of 80 to 100 hours with
underbalanced drilling (two to three times the life in comparable overbalanced hole).
Gauge Wear
Roller cone bit gauge wear occurs when the outer surfaces of the bit (heel and gauge row teeth,
shirt-tails) slide against abrasive minerals in the borehole wall. This can be a problem when
drilling with lightened drilling fluids; particularly when dry gas or mist is used to drill hard,
quartzitic rocks.26 When measurable gauge wear occurs, it will normally be necessary to ream the
next bit to bottom. Both the cutting structure and the gauge of the next bit are likely to wear
when reaming, reducing the footage of new hole that the bit can drill, before it too becomes
excessively worn. In addition, reaming imposes side loads on the bits bearings that can increase
the rate of bearing wear.30

3-19

Chapter 3

Benefits of Drilling Underbalanced

Roller cone bits suitable for air drilling are available in a variety of diameters, including sizes
that are not normally used in oilfield applications. Reaming can be reduced, or eliminated, if it is
possible to run a slightly larger bit, at the top of an interval, than is required to complete the
interval. Cooper et al., 1977,26 cited an example of an 8-inch diameter mining bit being run
above a 7 7/8-inch diameter bit, allowing up to 0.125 inches of gauge wear on the 8-inch bit,
before reaming became necessary.
Underbalanced liquid drilling involves circulating a lower density fluid (than would be used in
conventional drilling). The reduced fluid density implies a lower solids content. This in turn
makes the under-balanced drilling fluid less abrasive than conventional drilling muds. As a
result, lower abrasive wear rates can be expected. This should make gauge wear less of a concern
than it would have been in a conventional drilling operation.
Dry gases have lower lubricity than liquid drilling fluids. Measurements of torque and drag in
deviated and horizontal wells indicate that dry gases have drillstring friction factors of
approximately 0.4 to 0.45.29 For water-based drilling muds,29 friction factors of 0.2 to 0.35 are
typical. Mist and foam will improve this somewhat. It is not really feasible to use lubricant
additives when drilling with dry gas, mist or foam. As long as the liquid volume fraction remains
low, the friction factor will probably be large. Higher friction usually equates to higher wear
rates. Gauge wear occurs more rapidly when drilling with gas, mist or foam as opposed to
liquid.
Gauge wear tends to be aggravated by poor hole bottom cleaning. Higher than minimum gas
injection rates can help keep gauge wear to a manageable level, when drilling abrasive
formations with dry gas, mist or foam.32
3.2.2 Diamond Bits
Bits with natural and synthetic diamond cutting structures have different wear characteristics than
roller cone bits. Both natural and synthetic diamonds exhibit considerable temperature
sensitivity in their wear resistance. Diamond is only metastable, at the ambient temperatures and
pressures that prevail on surface and downhole. Diamond reverts to graphite, once its
temperature exceeds a critical value (1472F), at pressures that are not dramatically higher than
atmospheric. As graphite is a very soft solid, the wear rate of a natural diamond bit becomes
catastrophic if the diamonds are allowed to overheat.
Conventional Polycrystalline Diamond Compact (PDC) cutters, used in PDC bits, show
detectable increased wear rates once their temperature exceeds 660F.33 Thermal stresses in the
polycrystalline diamonds can become so high that conventional PDC cutters have no effective
wear resistance once their temperature reaches 1,380F.34
Cooling diamond cutting structures, whether natural or synthetic, is essential in order to prevent
wear at unacceptably high rates. The thermal capacities of gases and mists are much lower than
those of conventional drilling fluids. This means that higher diamond temperatures are probable
(with gases or mist). In hard rocks, this could easily result in very rapid wear of natural and
synthetic diamonds. Thermal modeling led Glowka and Stone, 1986,34 to recommend that PDC
3-20

bits not be used for dry air drilling. Higher vibration levels, common in some lightened fluids,
could also lead to increased impact damage of both natural diamonds and PDC cutters. Despite
this, natural diamond bits have been used quite successfully in coring with dry air and mist; for
example, in shallow Arkoma Basin wells.
Natural diamond bits have also been used successfully in mist coring of the Devonian shale
with mist.35 There are no reports in the literature of successful use of PDC bits in full-hole
drilling with dry gas or mist.
Natural diamond and PDC bits may be better suited to drilling with foam than with other
gasified, aqueous, drilling fluids. Since diamond is not naturally water wet, thermal transfer
between the water phase and the diamond is poor. Foaming agents are surfactants and provide
better contact with diamond; potentially, foams offer better cooling than do other gasified
aqueous liquids. With reasonable foam qualities and circulation rates, it is possible that foam
could provide adequate cooling of natural and polycrystalline diamonds. The wear rate of the
diamonds, in terms of the volume of rock removed, is likely to be lower, and the footage higher,
than it would be if the drilling was overbalanced. There are a number of instances21,36,37 in which
natural diamond bits have been used in foam coring without excessive wear. To date, however,
there are no public domain field data to confirm that full-hole PDC bits can be successfully run
with foam, although one major operator is currently experimenting with this.
Drilling underbalanced with a liquid drilling fluid will give higher penetration rates than drilling
overbalanced. The solids content of the drilling fluid will be lower, while its thermal capacity
will be little changed. As a result, abrasive wear rates of both roller cone and PDC bits are
expected to be lower, and bit footages higher, in wells drilled underbalanced with a liquid
drilling fluid, relative to wells drilled conventionally through the same formations.
3.2.3

Summary

 Slightly lower rotating hours should be expected for a roller cone bit drilling with dry gas,
mist or foam than would be seen in mud drilling. Experience with liquids (e.g. nitrified
crude) suggests rotating hours can be increased by two to three times.
 A higher footage per bit is probable in the majority of cases where the instantaneous
penetration rate is increased significantly by drilling underbalanced.
 Gauge wear may be a problem in dry gas, mist and foam drilling of hard rocks. The overall
footage per bit is still likely to be higher than it is drilling over-balanced.
 Dry gas drilling hard rock with synthetic diamond bits could dramatically increase wear rates,
due to excessive diamond temperatures.
 Although both natural and synthetic diamond bits may have acceptable life in foam drilling,
there is no firm evidence confirming this.
 With a liquid drilling fluid, both roller cone and PDC bit footages should be higher when
drilling underbalanced than when drilling overbalanced.
3-21

Chapter 3

3.3

Benefits of Drilling Underbalanced

Differential Sticking

Differential sticking is a leading cause for a stuck pipe. When it occurs, it usually adds
considerably to drilling costs; through the rig time taken to recover as much of the string as
possible and the cost of equipment that may be left downhole. Additional rig time is incurred if
the hole has to be redrilled around lost equipment. Drilling underbalanced can prevent
differential sticking. Avoiding differential sticking in an area where it is prevalent can be
sufficient justification for drilling under-balanced.
In conventional drilling operations, a positive differential is maintained between the drilling fluid
pressure in the borehole and the pore fluid pressure in the formations that are open to the
borehole. Liquid from the drilling mud flows into exposed permeable formations, depositing a
cake of solids from the mud on the borehole wall. The thickness of this mudcake depends on a
number of factors, including the volume of liquid lost to the formation and the solids content of
the mud. The mechanisms of mudcake deposition will not be discussed. Although it is possible
to restrict the mudcake thickness by controlling the fluid composition and its properties, it will
form as long as the well is overbalanced and a permeable formation is exposed.
There will be a pressure gradient through the mudcake. The mudcake permeability is generally
lower than that of the host formation. As a result, the pressure at the interface between the
mudcake and the intact formation will usually be quite close to the undisturbed formation
pressure and the change in pressure through the mudcake thickness will be close to the total
overbalance pressure.
Mudcake is relatively soft. If the drillstring is forced against the mudcake, it can ultimately
contact intact rock. The contact area will increase with increasing length of permeable zone and
increasing mudcake thickness. With even a modestly long permeable interval, this contact area
can be considerable.
The pressure differential through the mudcake will impose a force on the drillstring, pushing it
against the borehole wall. The force, Fs required to pull the string uphole through the mudcake
is:
Fs =

A c P s
144

(3.5)

where:
Ac ....... area of contact (ft2),
P ...... pressure differential across the
mudcake (psid), and,
s........ coefficient of friction between the
string and the mudcake
(dimensionless).
3-22

Quite often, the force required to move the string exceeds the permissible overpull. The string is
then differentially stuck. This is a particular problem when depleted zones are drilled
conventionally, since the differential pressures through these zones tend to be high.
Various procedures can be implemented to reduce the severity of differential sticking. The
contact area can be reduced by using spiral or square drill collars, bladed stabilizers, and heviwate drillpipe in the BHA, or by reducing the mudcake thickness through improved fluid loss
and solids control. The coefficient of friction can be reduced by adding lubricants to the mud or
by spotting a pill of lubricating fluid opposite the site of sticking. An added benefit of
hydrocarbon-based lubricants is the destruction of WBM mudcakes, allowing pressure
equalization and freeing the string.
If there is no positive pressure gradient through the mudcake, then there will be no force tending
to stick the string and differential sticking will be impossible. Since the borehole pressure is
necessarily lower than the formation pressure when drilling underbalanced, it is not possible for
differential sticking to occur if the openhole portion of the well is underbalanced.
Saponja, 1995,24 gave an example of how underbalanced drilling can prevent differential
sticking. The first 330 feet of a horizontal well through a depleted reservoir were drilled
overbalanced. Differential sticking was a problem through this interval. The well was then
switched to underbalanced drilling, first with nitrified water and then with nitrified crude. There
were no more problems with differential sticking. The string was stuck once more when an
unstable shale was entered and hole cleaning became inadequate. Increasing the annular velocity
by 50 percent improved hole cleaning and no further problems were encountered.
A number of instances have been reported where a string has been freed by creating
underbalanced conditions after the string had become differentially stuck while drilling
conventionally. Cagnolatti and Curtis, 1995,38 described one such case, that occurred during a
well control operation. Reducing the mud weight by centrifuging would free the string, while
killing the well to prevent any further inflow would then re-stick it.
Differential sticking is possible as long as the string is exposed to a permeable zone with a
positive (overbalanced) pressure differential. If a well is being drilled underbalanced in order to
avoid differential sticking, it is important that the entire openhole be underbalanced. Louison et
al, 1984,39 reported drilling underbalanced through the Midway and Navarro formations in the
Hallettsville area of Texas, in order to reach a deeper but lower pressure productive zone, the
Edwards limestone. To do this, depleted Wilcox sands had to be penetrated. The pore pressures
in these sands were sufficiently low that a positive differential pressure existed across them even
though the Midway and Navarro were underbalanced. Differential sticking and lost circulation
were experienced in the Wilcox sands.

3-23

Chapter 3

3.4

Benefits of Drilling Underbalanced

Lost Circulation

Moore, 1974,5 characterized lost circulation as the most common drilling problem. It can
involve mud flowing into highly permeable zones, open fractures or vugs that intersect the
wellbore. In these thief zones, borehole pressure merely has to exceed the pore fluid pressure for
mud losses to occur. Losses can also occur if the borehole pressure is sufficiently high that the
formation is hydraulically fractured. Lost circulation can be very costly, particularly if expensive
mud systems are being used or if the well is in a location where water is costly. It can also be
dangerous, if it allows the well to kick when proper surface equipment is not available to handle
inflow.
Lost circulation can be used to create underbalanced conditions in flowdrilling. During
flowdrilling, however, the resultant production is anticipated and the surface facilities required to
handle it safely will be in place. Under these circumstances, underbalanced drilling can
legitimately be regarded as a safer procedure than drilling conventionally without the additional
surface equipment.
Whatever the underlying mechanism, reducing the circulating borehole pressure to less than the
pore pressure in exposed formations will prevent losses. Lost circulation should not, therefore,
occur when drilling underbalanced, in most cases. However, it still can, given the right
circumstances. Lost circulation can occur if the bottomhole pressure is extremely low. In a
Texas Panhandle well, the reservoir pressure was 30 psia at approximately 4,000 feet. Even
while drilling with air, the circulating pressure was greater than the reservoir pressure and the
well experienced partial lost circulation.
In two other wells, lost circulation occurred because of excess water production and low
bottomhole pressures. While drilling with mist, a high volume water zone was encountered
along with a depleted zone. The hydrostatic pressure of the wellbore could not be lowered
enough with air to circulate out the water from the water zone. The pressure required to circulate
the water to the surface was greater than the lost circulation zone could tolerate. Consequently,
circulation could not be established. There are very many instances where underbalanced drilling
has been successful in controlling this.17,18,37,39,40,41, 42,43,44 The technique of drilling with aerated
mud was developed primarily to combat lost circulation.45 As was the case with differential
sticking, it is entirely possible to lose circulation during what is supposed to be an
underbalanced drilling operation if part of the openhole section is over-balanced.

3-24

3.5

Formation Evaluation

Drilling underbalanced can improve formation evaluation, by providing an indication of


productivity while drilling, and by restricting or preventing filtrate invasion of exposed
formations. Specific techniques available for formation evaluation during underbalanced drilling
are discussed in Chapter 6, Section 6.7. Production potential and real-time formation
characterization are possible.
3.5.1 Production Measurement
Formation fluids flow into an underbalanced well whenever a permeable zone is penetrated. As
noted in Chapter 2, Section 2.7, Gasified Liquids, several different instrumentation packages
have been developed for use with closed surface systems.46,47,48 These can combine
measurement of formation fluid production rates and pressures, while drilling underbalanced.
A closed system is not necessary for measuring gas or liquid production rates. If the surface
system allows produced liquids to be separated from any liquid component of the drilling fluid,
liquid rates can be measured by recording tank levels. Gas flowing from the well can be routed
through the choke manifold to a critical flow prover or an orifice plate, before flaring.
If desired, the bit can be pulled off bottom and drilling fluid injection can be stopped so that only
formation fluids flow from the well, in effect performing an openhole production test. In a gas
producer, this can allow stable production rates to be determined within one hour.36 Limited (or
no) invasion of drilling fluid filtrate into the producing formation accelerates the development of
steady or pseudo-steady flow, as can the limited volume of liquid in the borehole, if the well is
drilled with a dry gas, mist or foam.
Holdup in horizontal sections and slug flow in vertical sections of wells can complicate the
interpretation of measured production rates.49 So too can production from multiple intervals. It
is generally not possible to determine the extent to which each producing zone is depleting with
time, if production from additional zones is added to the measured total production as the well is
deepened. Nevertheless, it is possible to obtain a qualitative indication of the permeability of
different exposed formations from the production rate measured while drilling underbalanced, by
comparing cumulative production and estimating thickness and formation pressures.23 This gives
an indication of potentially productive zones while drilling.
Canadian experience23 is that production rates observed when drilling usually give accurate
indications of the subsequent production capacity of the well. Drilling necessarily measures the
early production history only. In some instances, factors unrelated to the drilling technique can
prevent the observed production rate from being maintained when the well is put on production.
For example, sales production rates from some air mist drilled (San Juan Basin Mesaverde)
horizontal gas wells were reported50 to be lower the rates seen while drilling. This was thought
to reflect loading of the wellbore by produced liquids which restricted the flow of gas uphole.
Alternatively, fines migration during production can reduce the permeability of the near-wellbore

3-25

Chapter 3

Benefits of Drilling Underbalanced

region and cause the production rate to decrease with time.51 Different mechanisms of
permeability impairment and the extent to which underbalanced drilling can prevent them will be
discussed in the following section.
Measurements of production rates have allowed non-productive zones to be identified and the
planned well trajectory to be modified while drilling a number of horizontal wells. Changes in
the GOR can be inferred from the production measurements, and these have been interpreted as
indicating approach towards a gas-oil contact. Measurements of production made while drilling
can also be used to aid geosteering.23,25,49 Water producing zones can be identified. This may
assist in making completion decisions.
Very often, measurements of production rate made while drilling underbalanced will give
enough information that drillstem testing is not required. This reduces the cost of evaluating
the well. If it is necessary to perform an openhole test for gas production, this can be done very
rapidly when drilling with dry gas, mist or foam. Rig time is reduced by not having to make a
round trip, rig up and rig down the test tools, and, as noted above, the test time itself may be
reduced. Additional equipment is not required to conduct the test, except possibly the flow
prover. Reduced testing costs offer the possibility of more extensive evaluation than might
otherwise have been undertaken.
3.5.2 Invasion While Drilling
During conventional overbalanced drilling operations, whole mud, filtrate and solids may all
invade permeable zones while they are open to the wellbore. Liquid hydrocarbons and formation
water may be displaced. Filtrate may also increase water saturation close to the wellbore.
Invasion can complicate interpretation of wireline logging data. Understanding and managing
the effects of invasion are fundamental to quantitative log interpretation.52
Invasion can complicate testing by requiring lengthy flow periods or clean-up operations before
stable production rates are established. Difficulties in determining whether or not a well has
cleaned up prior to testing introduce uncertainty into the assessment of test results. In the Calling
Lake/Algar area of Northern Alberta, this uncertainty motivated one operator to switch to drilling
exploration wells underbalanced to improve evaluation performance,36 and avoid the cost of pretest clean-up operations and running DSTs.
Maintaining underbalanced conditions while drilling greatly restricts or prevents most forms of
invasion, by removing the physical force that drives material from the wellbore into exposed
formations. The quality of drillstem or production tests is generally improved if the well is
drilled underbalanced. Reduced invasion will normally help with the interpretation of logs, but
selected logging tools need to be run in air-filled holes. Formation evaluation is discussed in
Chapter 6. If at all possible, one should avoid filling the hole with mud to run logging tools since
this may damage the formation.

3-26

There are circumstances under which an aqueous phase of the drilling fluid can enter the
formation even though the pore pressure is higher than the drilling fluid pressure. This
invasion mechanism (spontaneous imbibition) will be discussed in Section 3.6, Formation
Damage. Also, underbalanced conditions may not be maintained throughout the drilling of an
entire interval, particularly when a gasified liquid is being used as the drilling fluid. Generally the
liquids used in these drilling operations do not have any fluid loss control, and can rapidly
penetrate the formation during even brief periods of overbalance, leading to significant invasion.49
Formation damage can occur when drilling underbalanced in lower permeability gas formations.
Even with air, the pressure at the sandface will fall until it is equivalent to the wellbore pressure.
While tripping, the well is allowed to flow and the pressure declines. After tripping back in the
hole, re-setablishing circulation can increase the wellbore pressure faster than the sandface
pressure can recover. Some formation damage can occur before the sandface pressure equals the
circulating pressure.
3.5.3 Invasion While Coring
Cores can be invaded as they are cut and while they are inside the core barrel downhole. Filtrate
invasion of cores can lead to inaccurate determinations of wettability, relative permeability, and
oil and water saturations.53
Successful techniques for overbalanced low invasion coring (Rathmel et al., 1990)54 have been
developed. An essential element of these procedures is that the coring penetration rate has to
exceed the filtrate invasion rate ahead of the bit. This may not always be possible, particularly in
certain hard, depleted reservoirs. To avoid significant invasion, underbalanced coring is an
option.
Cutting and collecting cores underbalanced has been done with air, air mist,35 foam and other
techniques.21,36,37 This can provide cores that are generally more representative of the reservoir.
The presence of oxygen in the coring fluid could influence formation wettability.18 It might be
appropriate to avoid using air in underbalanced coring fluids if wettability modification by
oxidation is a concern. Untreated lease crude may be a desirable base fluid.
3.5.4 Economic Impact
While the value of improved formation evaluation resulting from underbalanced drilling will
vary considerably from application to application, it can be substantial:
 Improved geosteering of horizontal wells can increase productivity or delay the onset of
coning.
 The cost of testing can be reduced or even eliminated totally.
 Underbalanced drilling will allow delineating additional, productive formations that might
have otherwise been bypassed or not tested, if they were drilled overbalanced. This would
include zones that might be difficult to interpret if they were significantly damaged or zones
difficult to recognize with normal logging (e.g. fractured zones or permeable, radioactive
dolomite).
3-27

Chapter 3

3.6

Benefits of Drilling Underbalanced

Formation Damage

Formation damage is the term used to describe reduction in the permeability of reservoir rocks,
induced while drilling, completing, stimulating, or producing/ managing a well.
Bennion et al., 1993,55 argued that formation damage seems to be a greater problem for
horizontal wells than for vertical wells through the same formations:
 The producing formation is exposed to the drilling fluid for longer periods of time, because a
greater distance is drilled through the pay zone and because drilling horizontally can give
lower penetration rates than drilling vertically.
 Many horizontal wells are left openhole or completed with slotted or pre-perforated liners.
Under these circumstances, it may not be possible to perforate through near-wellbore damage
or to do selected interval stimulation.
 Concerns about wellbore instability limit the drawdown that can be applied to some
horizontal wells. This in turn can restrict clean-up during early production.
 Finally, the greater length of exposed reservoir formation can render chemical clean-up
treatments unacceptably expensive.56
In many applications, drilling underbalanced can greatly reduce or eliminate drilling-induced
formation damage. This is the main driving force behind the recent, dramatic upsurge in
underbalanced drilling in Canada. Different mechanisms for formation damage are briefly
described below, distinguishing between those that can be prevented by maintaining
underbalanced conditions and those that cannot.
Laboratory studies of specific formations can indicate whether or not reduction of their
permeability can be avoided by drilling underbalanced. There is little public domain information
directly comparing wells drilled underbalanced and overbalanced. From the literature, it is
difficult, in most cases, to unambiguously establish the impact of underbalanced drilling on
productivity. Nevertheless, a number of field examples have been given showing how
productivity can be increased by drilling underbalanced and how, in some instances, it can
remove the need for primary stimulation. One case, in Canada, compared two wells drilled 328
feet apart; the well drilled underbalanced had a seven-fold increase over the first well which had
been drilled overbalanced.
There are applications in which drilling underbalanced may not yield economically valuable
productivity increases. These involve formations that can be damaged during underbalanced
drilling, during completion and subsequent production, and those that are not sufficiently
productive even when undamaged to be economically viable without stimulation.

3-28

3.6.1 Formation Damage Mechanisms


During drilling, formation damage can occur in a variety of ways.55,57 These include:
 Formation of scales, sludges or emulsions due to interaction between invading mud filtrate
and pore fluids,
 Interaction between aqueous mud filtrate and clay particles in the formation,
 Invasion of solids from the drilling fluid, restricting fluid flow through the pore structure,
 Invasion and entrapment of high oil or water phase saturations in the near wellbore region
(phase trapping/ blocking),
 Adsorption of drilling fluid additives, leading to permeability reductions or wettability
alteration,
 Migration of fines in the formation, particularly under highly overbalanced conditions, and,
 Generation of pore-blocking organic byproducts from bacteria entering the formation from
the drilling fluid.
These mechanisms all involve penetration of the drilling mud, filtrate or solids into the
formation. If the drilling fluid pressure is reduced below the pore pressure by drilling
underbalanced, the physical driving force for this penetration is removed. However,
underbalanced drilling causes higher effective stresses around the wellbore, potentially reducing
permeability.
3.6.2 Formation Damage While Underbalanced
Underbalanced drilling techniques do not completely eliminate the possibility of formation
damage. It can occur due to:

Temporary overbalance,

Spontaneous imbibition,

Gravity-induced invasion,

Wellbore glazing,

Post-drilling damage, and,

Mechanical degradation.

Temporary Overbalance
Overbalanced conditions may temporarily occur during drilling and completion. This may be
intentional; for example, to kill the well prior to tripping, to transmit MWD surveys when
drilling with gas injection down the drillstring, to log the well, or to perform various completion
and workover operations.58

3-29

Chapter 3

Benefits of Drilling Underbalanced

Overbalanced conditions may also inadvertently occur. Slug flow or liquid holdup in the annulus
can cause large fluctuations in downhole pressures. These are particularly likely when drilling
with a lightened liquid or when gas is flowing from the formation into the annulus.
 When drilling with gasified liquid by injecting gas down the drillstring (refer to Section 2.7),
fluctuations in downhole pressures, sufficient to create an overbalance, can arise during connections.
 With natural diamond and TSP bits, the drilling fluid pressure over the central portion of the
workfront is significantly higher than in the annulus. The pressure drop across the face of a
natural diamond bit can easily exceed 500 psi.12 If the target underbalance pressure (the
difference between the bottomhole and formation fluid pressures) is less than this, some
invasion is possible. Bennion et al., 1995,49 argued that annular gas injection can allow the
drilling fluid pressure (where flow from the bits nozzles impacts the workfront) to be
significantly higher than in the annulus.
 Flow from the formation into the wellbore after the bit has passed will reduce the pore
pressure adjacent to the wellbore, in effect locally depleting the reservoir. The locally
depleted regions will be less tolerant to fluctuations in downhole pressure before
overbalanced conditions are created.49
 Varying pore pressure regimes along the wellbore can allow one zone to be overbalanced
while another is con-currently underbalanced. An incorrect value of pore pressure may be
inadvertently used to select the target drilling fluid pressure.
 Temporary overbalance may occur when the drillstring is run in too fast after a bit change
(acts like a piston on fluids in the hole).
 Finally, there may be equipment malfunctions or procedural errors that prevent
underbalanced conditions from being maintained.
Very often, the drilling fluids used in underbalanced operations provide no fluid loss control. As
a result, there is nothing to reduce the rate at which they can invade the formation, if
overbalanced conditions occur. If a well is being drilled underbalanced principally to reduce
formation damage, the drilling fluid should be chosen to be compatible with the producing
formation. In this way, significant permeability impairment can be avoided if an over-balanced
situation is temporarily created. Laboratory screening tests, where core plugs are exposed to
candidate drilling fluids, can be very helpful in selecting a non-damaging fluid.59
Spontaneous Imbibition
It is possible for a liquid component of the drilling fluid to enter an exposed formation by
capillary pressure effects, even when there is an opposing underbalance pressure. Bennion et al.,
1993,55 indicated that this may allow formation damage to occur when drilling underbalanced.
Simplistically, capillary pressures arise because of the forces that act at liquid, gas, and solid
interfaces. These depend on the nature and composition of the fluids and the matrix. These
compositional variations regulate the magnitude of the capillary pressure and control which in-

3-30

situ fluids will preferentially wet (be in intimate contact with) the solids constituting the
formation. Some reservoir rocks are naturally water-wet, (Morrow, 1990),60 indicating that the
surface energy of the rock is lower when covered by water (or an aqueous fluid) than when it is
covered by a hydrocarbon. As a result, water will tend to be drawn into the rocks pore structure
if an initially hydrocarbon-filled pore is exposed to water. The capillary pressure, Pc, can be
thought of as the pressure that has to be applied inside a particular pore, to prevent the wetting
fluid (often water) from being drawn into it. The significance of this for underbalanced drilling
is that the underbalance pressure must equal or exceed the capillary pressure, if an exposed
water-wet formation is not to imbibe (draw in) water from a drilling fluid that contains an
aqueous phase.
In a porous medium, capillary pressure can be approximated as:
Pc =

2 cos
rp

(3.6)

where:
Pc ....... capillary pressure (psi),
......... interfacial tension between the
relevant fluids and/or gases
(lbf/in),
......... contact angle of the wetting fluid
on the rock matrix, measured
through the denser phase and
varying from 0 to 180, and,
rp ........ effective capillary radius (in).
This shows that the smaller the pore (rp indicates pore size), the higher the pressure that must
be applied to prevent water from being drawn into it. Taking typical values for quartz, at specific
methane and brine saturations61 the capillary pressure will be approximately 10 psi for a pore
diameter of 2 m, 100 psi for a pore diameter of 0.2 m, and 1,000 psi for a pore diameter of
0.02 m.
There is not a unique capillary pressure for various combinations of rock, water and hydrocarbon.
The capillary pressure is related to the degree of wetting phase saturation; that is to the fraction
of pores that are filled with the wetting phase. Presuming a very strongly water-wet formation,
the capillary pressure is initially very low at 100 percent water saturation, increases with
decreasing water saturation, and becomes very large (values exceeding 1,000 psi are not
infrequently measured) as the irreducible water saturation (Swirr) is approached. This is
illustrated schematically in Figure 3-10. The underbalance pressure necessary to prevent water
from being drawn into an exposed water-wet formation from an aqueous drilling fluid will
depend on the initial formation water saturation and the pore sizes.
3-31

Chapter 3

Benefits of Drilling Underbalanced

A number of water-wet gas reservoirs, notably in tight formations and in zones which have
experienced significant gas migration, have water contents below the irreducible water
saturation.49,62 It will be effectively impossible to impose an underbalance pressure that is high
enough to prevent spontaneous imbibition of water into these formations. If the water saturation
is above but close to the irreducible level, water may still be imbibed into the formation if the
underbalance pressure is low. As shown in Figure 3-10, this counter-current imbibition would be
possible into water-wet formations against an underbalance pressure of P, for initial water
saturations of up to Swc.
Equivalent imbibition of hydrocarbon-based drilling fluids into oil-wet (or mixed or neutral
wettability) formations is possible under analogous circumstances. Gas-bearing formations may
be oil-wet when they contain naturally oil-wet minerals, such as pyrobitumen, sulfur, asphalt or
residual heavy bitumen, or when a sub-irreducible oil saturation has been established by gas
displacing an original oil column.49

800

Capillary Pressure (psi)

700

Zone of Potential
Spontaneous
Imbibiton

Countercurrent Imbibition is Possible


for Initial Wetting Phase Saurations
Between 20 and 47% for the
Underbalance Pressure Shown in this
Example (200 psi).

600
500
400
300
Example Underbalance Pressure

200
100

S irr = 40%

S i = 20%

S c = 47% (Equilibrium)

0
0

20

40

60

80

100

Wetting Phase, , Saturation (%)

Figure 3-10. Schematic relationship between capillary pressure and water saturation for
water-wet reservoir rock (after Bennion et al., 1993 55).

3-32

Strongly water-wet formations will not, however, imbibe hydrocarbons, neither will strongly oilwet formations imbibe aqueous fluids. Once liquid is imbibed into the formation from the
drilling fluid, it may cause formation damage by any of the mechanisms that might have operated
if it had been forced into the formation by an overbalance pressure. For example, it may interact
with the formation itself or the formation fluids, or it may lead to phasetrapping and blocking.
Knowledge of wettability, initial saturation, and capillary pressure characteristics will facilitate
the identification of those regimes where spontaneous and counter-current imbibition may occur
during underbalanced drilling. Together with laboratory core-flood studies, this should assist the
selection of underbalanced drilling fluids and annular pressures that would minimize the
potential for formation damage in future wells.49,51,59,63
Gravity-Induced Invasion
The density of any liquid phase in the drilling fluid will almost invariably be greater than the
density of natural gas in a gas-bearing formation. Under normal circumstances, flow from the
formation into the borehole when drilling underbalanced will prevent movement of drilling fluid
into the formation, even if it has a higher density than the formation fluid. If, however, the
formation produces from natural fractures or large vugs, it is possible for liquid to flow into these
openings under the influence of gravity even though the well is underbalanced49 and natural gas
is flowing into the wellbore through the same apertures.
Wellbore Glazing
Particularly when drilling with a dry gas, high temperatures can be generated where the bit and
the BHA rub against the borehole wall. Since frictional coefficients between steel and dry rock
are high, more energy is dissipated in friction than if there was a significant liquid volume
fraction in the drilling fluid. Also, since the thermal capacity of dry gases is lower than for liquid
drilling fluids, they cannot cool the borehole as effectively. These high temperatures have long
been suspected of creating a thin, low permeability, glazed zone57 on the borehole wall.
Glazing has been noticed on the surface of whole core and rotary sidewall cores in air drilled
holes. In highly deviated and horizontal wells, there will almost certainly be beds of cuttings on
the hole bottom that will be reground between the BHA and the intact rock of the borehole wall.
Glazing has been noticed on the surface of whole core and rotary sidewall cores in air drilled
holes. This regrinding may create a paste of fine cuttings that can contribute to glazing.
Glazing tends to be localized in the immediate vicinity of the borehole wall, extending at most a
few tenths of an inch into the formation.57 It will not, therefore, inhibit the productivity of a
cased and perforated well. Similarly, glazing is unlikely to be a problem in formations that
produce from natural fractures or vugs, even if these wells are completed openhole. The size of
the productive features is generally too great to be obstructed by the glaze.49
Any glaze created in carbonates is likely to be acid soluble, and can probably be readily removed
by an acid wash. A silicate-based glaze might be expected in sandstones, and this could be more
difficult to remove.49 Mud acid or hydrofluoric acid may be effective.

3-33

Chapter 3

Benefits of Drilling Underbalanced

Damage After Drilling


There is little point in drilling a well underbalanced to avoid formation damage, if the well is
then intentionally damaged during completion, workover or production.
 The possibility of damaging the formation by killing the well to conduct completion or
workover operations has already been mentioned. If a separate completion fluid is used, care
should be exercised to ensure that this does not damage the formation. This is probably even
more important than would be the case for a conventionally drilled well since there will be
little or no protective mudcake on the wall of a well that was drilled underbalanced.
 Cementing can create damage by invasion of filtrate or cement solids. A cemented casing
will, however, be perforated and the perforations may penetrate beyond the damaged zone.
 There are several mechanisms that may cause the permeability of the near-wellbore region to
decrease during the productive life of the well. Fines may be mobilized from within the
producing formation by the flowing pore fluid and carried into the near-wellbore region.
Francis et al., 1995,51 conducted laboratory core-flooding measurements, in which plugs from
a reservoir that was a candidate for drilling underbalanced were exposed to the proposed
drilling fluid under both under- and over-balanced conditions. These experiments all showed
a significant permeability reduction, to 20 percent of the original level or less, irrespective of
the differential pressure. Thin section petrography and scanning electron microscopy showed
that the permeability impairment was caused by migration of intrinsic fines. The fines were
displaced both by flow from the formation, when underbalanced, and by flow into the
formation under conditions of high overbalance pressure. Drilling underbalanced would not
avoid formation damage by fines migration in this reservoir.
 It is also possible for liquids to be carried into the near-wellbore region during production, to
the extent that they reduce the permeability to gas. Drop-out of produced liquids has been
reported as the probable cause of formation damage in air mist drilled horizontal gas wells in
the San Juan Basin,50 and in a Canadian well drilled underbalanced59 with nitrified liquid.
Mechanical Degradation
The rock around a wellbore experiences a concentration of in-situ stresses, simply due to drilling
the well. As the wellbore pressure is lowered, particularly if there is no filter cake, the effective
stresses increase. With elevated stresses, porosity and available flow channels can decrease,
reducing permeability. This is analogous to reductions in permeability that may occur during the
productive life of a reservoir as the reservoir is depleted. Reductions in permeability may or may
not be recoverable, possibly less so in poorly consolidated formations.
Around the wellbore, stresses may become so large that failure occurs, in which case nearwellbore permeability may or may not increase.

3-34

3.6.3 Formation Damage Reduction


Beyond the drilling technique, many factors influence the productivity of wells drilled in the
same field. For example, there are inevitably variations in reservoir quality (porosity,
permeability, thickness ...) and in formation pressure support for production. Well trajectories
vary, as do their orientation and placement in the reservoir. These inherent variations must be
considered when comparing production data from wells drilled over- and underbalanced. Also,
the relative refinement of the different drilling techniques needs to be considered - it may be
inappropriate to compare the last overbalanced well with the first well drilled underbalanced.
Ideally, comparisons should be made from equivalent points on the learning curve.
Surewaard et al., 1995,58 recognized these uncertainties and developed a statistical model to
determine how many wells would have to be drilled underbalanced in order to establish, with
reasonable confidence, that a particular level of productivity improvement had been achieved.
They applied this analysis to the Nimr field in Oman, where conventionally drilled horizontal
wells showed a significant spread in productivity indices. The conclusion was that, for this field
at least, ten wells would have to be drilled underbalanced to establish (at an 80 percent level of
confidence) that a 50 percent increase in productivity had been achieved. Fewer underbalanced
wells would be required to establish higher productivity increases - one well would be enough to
prove benefit from drilling underbalanced if it gave more than 2.5 times the average productivity
of equivalent conventionally drilled wells.
Noting these various qualifications, a number of examples are presented to illustrate the sort of
productivity increases that drilling underbalanced can cause.
 In the United States, the Austin Chalk has probably seen the most underbalanced drilling
activity. Stone, unpublished, 1993,64 reported that many operators in South Texas had
consistently achieved higher productivity from Austin Chalk wells that were drilled
underbalanced (flowdrilled) than from wells that were drilled overbalanced.
 Joseph, 1995,65 described a deep, horizontal well that was drilled underbalanced into the
Austin Chalk, in central Louisiana. This achieved a gas production rate that was seven times
higher than the best rate from any horizontal well drilled conventionally in that area.
 Hutchinson, 1970,21 described a well drilled with stable foam in fractured Miocene cherts,
near Santa Maria, California. This well produced heavy oil at 500 BOPD, whereas the best
mud-drilled offset well produced at 150 BOPD.
 Recently, there has been considerable underbalanced drilling in Canada. Much of this has
been undertaken specifically to avoid formation damage. The Weyburn field, in southeastern
Saskatchewan, produces oil from a chalky dolostone at approximately 4,600 feet TVD. The
average porosity is 26 percent and permeability varies from 1 to 100 md. To date, over eighty
horizontal wells have been drilled underbalanced in this field. Despite this level of activity,
there remains uncertainty over the extent to which underbalanced drilling may impact
productivity. Springer et al., 1994,66 reported that drilling underbalanced in this field led to
productivity that was two to three times greater than for horizontal wells drilled
overbalanced. However, Mullane et al., 1995,67 suggested a more complex situation.
3-35

Chapter 3

Benefits of Drilling Underbalanced

Some of the early underbalanced wells were disappointing. It was suspected that they had been
damaged by allowing overbalanced conditions to occur when making connections. Acid
stimulation greatly improved production, possibly confirming damage. More recent
underbalanced wells have adopted better connection procedures that maintain the underbalance
pressure. The preliminary results from these wells indicated increased productivity, but it was
felt that further production would be required before this could be confirmed.
 The Westerose gas field, in Alberta, provides a more encouraging picture of the extent to
which underbalanced drilling can enhance productivity, while at the same time confirming
the difficulty in objectively comparing different drilling techniques. The productive
formations are glauconitic sandstones. They are prone to permeability impairment by
swelling of interstitial clays and by phase trapping. Four vertical wells were drilled
underbalanced, two using nitrified KCl and two using nitrogen/KCl foam. During drilling,
two of these wells flowed gas at rates that were typically three times those from DSTs in
equivalent wells drilled overbalanced.67 These increases were not sufficient for the wells to
be economic without further stimulation. In one well, the productive formation was damaged
by the water and condensate that were used to load the well for pre-frac testing and was
subsequently lost into the formation during testing. The remaining two wells that were
drilled underbalanced penetrated oil-bearing, shaley sands which did not flow either oil or gas
at economic rates.
 Lunan, 1995,68 presented production data from two vertical wells, drilled 330 feet apart, in
the Sinclair Field, in Northern Alberta. The productive formation in this field, the Paddy, is a
medium- to coarse-grained sandstone that is susceptible to formation damage when drilled
conventionally. The first well was drilled and completed conventionally and then fractured.
The fracturing treatment was not particularly successful, possibly because of high in-situ
stresses. The second well was drilled underbalanced using nitrified condensate. This well
produced at a rate that was seven times greater than for the first well. Production test data
indicated an effective permeability that was five times higher than for the first well. The well
drilled underbalanced had no skin, whereas the well drilled overbalanced and then fractured
had a skin of +43. The higher permeability in the second well negates this as an effective
comparison, although zero skin is encouraging.
 A horizontal well was drilled into oil-bearing glauconitic sandstone in the Hussar field, in
Alberta, using overbalanced and then underbalanced conditions along the same lateral
section, in an attempt to assess any benefits of drilling underbalanced.69 The first 330 ft of
the horizontal interval were drilled overbalanced with a starch/guar gum polymer water-based
mud. Damage sustained during overbalanced drilling prevented production from the first 330
feet of the hole when the well was switched to nitrified crude and a further 2,060 feet were
drilled underbalanced. This damage was subsequently removed by underbalanced abrasive
jetting. A rotating jetting tool was run downhole on coiled tubing. Oil was pumped down
the coiled tubing and nitrogen was pumped down the annulus between the coiled tubing and
the production tubing annulus. This effectively removed the mud-induced damage and
allowed that section to produce oil.

3-36

 Kitsios et al., 1994,70 described a vertical well drilled with air foam, in the Amal field in
Oman. This had a productivity index that was over twice the average for adjacent
conventionally drilled wells.
 It is not just production rate that can be improved by drilling underbalanced. The protocol
may accelerate hydrocarbon production from wells drilled in reservoirs prone to lost
circulation. Mitchell and Salvo, 1991,18 reported that wells air mist drilled in the Eunice
Monument Field, in New Mexico, would produce immediately. Conventional overbalance
wells typically took between one month and one year to begin producing oil, due to large
volumes of mud and water that had been lost into the producing formation.
3.6.4 Required Stimulation
In some wells, supplementary stimulation treatments are required; for example matrix acidizing
or hydraulic fracturing, to remove or bypass drilling-induced formation damage. Underbalanced
drilling can, in some instances, remove the need for primary stimulation, by restricting the
occurrence of formation damage.
Mullane et al., 1995,67 reported that horizontal wells drilled underbalanced in the Weyburn field
were cheaper overall than equivalent wells drilled overbalanced, principally because they did not
require the bleach clean-up or acidization treatments that were necessary to develop adequate
production in the overbalanced wells.
Hutchinson and Anderson, 1972,20 described an oil well drilled with air foam, into a carbonate
reservoir, in Colorado. This produced over 600 BOPD. Adjacent mud-drilled wells experienced
lost circulation in the pay and required massive acid treatments before they would produce.
3.6.5

Summary

 Drilling underbalanced will often reduce or avoid drilling-induced formation damage. The
resultant increase in productivity, earlier production and possible reduction in stimulation
costs often make underbalanced drilling economically attractive, even without considering an
improved penetration rate.
 There is no clear evidence that any one underbalanced drilling technique gives greater or
lesser productivity increases than other techniques.
 It is important that the well be kept underbalanced at all times, if formation damage is to be
minimized. A well that was drilled underbalanced can experience severe formation damage
if overbalanced conditions exist during completion, production or workover.
 There are mechanisms by which damage can occur during underbalanced drilling, even if an
underbalance is maintained continuously. The extent to which underbalanced drilling can
reduce formation damage will be strongly dependent on the specific target formations.
Laboratory evaluations can reveal the potential for reducing formation damage by drilling
under-balanced.

3-37

Chapter 3

Benefits of Drilling Underbalanced

 There is little point in drilling underbalanced (to avoid formation damage) if damage occurs
during subsequent production; for example, by fines migration. Increased ROP and other
factors in the total economics may dictate using underbalanced methods, regardless of
formation damage.
 Drilling underbalanced will not cause a negative skin unless near-wellbore failure causes
permeability increases (possibly undesirable from stability considerations). If an
underbalanced well is not economic with zero skin, it will still need stimulation. Avoiding
formation damage alone does not justify using underbalanced drilling techniques for a well
that will still have to be stimulated after drilling.

References

3-38

1.

Cheatham, J.B. Jr.: An Analytical Study of Rock Penetration By a Single Tooth, Proc.,
Eighth Drilling And Blasting Symposium, University of Minnesota (1958) 1A-22A.

2.

Garnier, A.J. and Van Lingen, N.H.: Phenomena Affecting Drilling Rates at Depth,
Trans., AIME (1959) 216, 232-239.

3.

Cunningham, R.A. and Eenink, J.G.: Laboratory Study of Effect of Overburden,


Formation and Mud Column Pressures on Drilling Rate of Permeable Formations, Trans.,
AIME (1959) 216, 9-17.

4.

Black, A.D., Dearing, H.L. and DiBona, B.G.: Effect of Pore Pressure and Mud Filtration
on Drilling Rates in a Permeable Sandstone, JPT (September 1985) 1671-1681.

5.

Moore, P.L.: Drilling Practices Manual, PennWell Books, Tulsa, OK (1974).

6.

Van Lingen, N.H.: Bottom Scavenging - A Major Factor Governing Penetration Rates at
Depth, paper SPE 165, JPT (February 1962) 187-196.

7.

Eckel, J.R.: Effect of Pressure on Rock Drillability, paper SPE 877G, Pet. Trans., AIME,
(1958) 213, 1-6.

8.

Zijsling, D.H.: Single Cutter Testing - A Key for PDC Bit Development, paper SPE
16529/1 presented at the 1987 Offshore Europe 87, Aberdeen, Scotland.

9.

Gray-Stephens, D., Cook, J.M. and Sheppard, M.C.: Influence of Pore Pressure on
Drilling Response in Hard Shales, SPEDC (December 1994) 263-270.

10.

Maurer, W.C. How Bottomhole Pressure Affects Penetration Rate, Oil and Gas J.
(January 10, 1966) 61-65.

11.

Cheatham, C.A., Nahm, J.J. and Heitkamp, N.D.: Effects of Selected Mud Properties on

Rate of Penetration in Full-Scale Drilling Simulations, SPE/IADC paper 13465 presented


at the 1985 SPE/IADC Drilling Conference, New Orleans, LA.
12.

Winters, W.J. and Warren, T.M.: Laboratory Study of Diamond-Bit Hydraulic Lift,
SPEDE (August 1986) 267-276.

13.

Moore, P.L.: How Drilling Rate is Affected by Drilling Fluid Properties, Oil & Gas J.
(October 1958) 141.

14.

Beck, F.E, Power, J.W. and Zamora, M.: The Effect of Rheology on Rate of Penetration,
SPE/IADC paper 29368 presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

15.

Pratt, C.A.: Modifications to and Experience with Air-Percussion Drilling, SPEDE


(December 1989) 315-320.

16.

Lyons, W.C., Miska, S.Z. and Johnson, P.W.: Downhole Pneumatic Turbine Motor:
Testing and Simulation Results, paper SPE 18040 presented at the 1988 SPE Annual
Technical Conference and Exhibition, Houston, TX.

17.

Cummings, S.G.: Natural Gas Drilling Methods and Practice: San Juan Basin, New
Mexico, paper SPE/IADC 16167 presented at the 1987 SPE/IADC Drilling Conference,
New Orleans, LA.

18.

Mitchell, R.K. and Salvo, G.S.: The EMSU Waterflood Project: A Case History of Infill
Drilling, Completions and Workovers, SPEDE (June 1991) 118-124.

19.

Fraser, I.M. and Moore, R.H.: Guidelines for Stable Foam Drilling Through Permafrost,
paper SPE/IADC 16055 presented at the 1987 SPE/IADC Drilling Conference, New
Orleans, LA.

20.

Hutchinson, S.O. and Anderson, G.W.: Preformed Stable Foam Aids Workover,
Drilling, Oil & Gas J. (May 15, 1972) 74-79.

21.

Hutchinson, S.O.: Stable Foam Lowers Production, Drilling and Remedial Costs,
presented at 17th Annual Southwestern Petroleum Short Course, (April 1970).

22.

Wolke, R.M., Jardiolin, R.A., Suter, R.L., Moriyama, S., Sueyoshi, Y. and Kihara, Y.:
Aerated Drilling Fluids Can Lower Drilling Costs and Minimize Formation Damage,
Geothermal Resources Council Bulletin, (May 1990) 131-137.

3-39

Chapter 3

3-40

Benefits of Drilling Underbalanced

23.

Comeau, L.: Underbalanced Drilling: Directional and MWD Experience, paper


presented at the 1995 1st International Underbalanced Drilling Conference, The Hague, The
Netherlands, October 2-4.

24.

Saponja, J.: Comparing Conventional Mud to Underbalanced Drilling in a Depleted


Reservoir, paper presented at the 1995 1st International Underbalanced Drilling Conference,
The Hague, The Netherlands, October 2-4.

25.

Deis, P.V., Yurkiw, F.J. and Barrenechea, P.J.: The Development of an Underbalanced
Drilling Process: An Operators Experience in Western Canada, paper presented at the
1995 1st International Underbalanced Drilling Conference, The Hague, The Netherlands,
October 2-4.

26.

Cooper, L.W., Hook, R.A. and Payne, R.R.: Air Drilling Techniques, paper SPE 6435
presented at the 1977 SPE Deep Drilling and Production Symposium, Amarillo, TX.

27.

Wilson, G.E.: Air Drilling and Crooked Hole Problems, paper SPE 9529 presented at the
1980 SPE Eastern Regional Meeting, Morgantown, WV.

28.

Brannon, K.C., Grimes, R.E. and Vietmeier, W.R.: New Oilfield Air Bit Improves
Drilling Economics in Appalachian Basin, PD-Vol. 56, Drilling Technology (1994),
ASME (1994).

29.

Carden, R.S.: Air Drilling has Some Pluses for Horizontal Wells, Oil & Gas J. (April 8,
1991) 76-78.

30.

Kelly, J.L., Jr.: Forecasting the Life of Rock-Bit Journal Bearings, SPEDE (June 1990)
165-170.

31.

Simon, R.: Energy Balance in Rock Drilling, SPEJ (December 1963) 298-306.

32.

Shale, L. and Curry, D.A.: Drilling a Horizontal Well Using Air/Foam Techniques, paper
OTC 7355 presented at the 1993 Annual OTC, Houston, TX.

33.

Glowka, D.A. and Stone, C.M.: Thermal Response of Polycrystalline Diamond Compact
Cutters under Simulated Downhole Conditions, SPEJ (April 1985) 143-156.

34.

Glowka, D.A. and Stone, C.M.: Effects of Thermal and Mechanical Loading on PDC Bit
Life, SPEDE (June1986) 201-213.

35.

Eaton, N.: Coring the Horizontal Hole, PD-Vol 27, ASME Drilling Technology
Symposium, Weiner, P.D. and Kastor, R.L (eds).

36.

Bentsen, N.W. and Veny, J.N.: Preformed Stable Foam Performance in Drilling and
Evaluating Shallow Gas Wells in Alberta, JPT (October 1976) 1237-1240.

37.

Cobbett, J.S.: Application of an Air-Drilling Package in Oman, paper SPE 9600


presented at the 1981 SPE Middle East Oil Technical Conference, Manama, Bahrain.

38.

Cagnolatti, E. and Curtis, F.: Using Underbalance Technology to Solve Traditional


Drilling Problems in Argentina, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

39.

Louison, R.F., Reese, R.T. and Andrews, J.P.: Case History; Underbalance Drilling the
Midway and Navarro Formations Successfully in Hallettsville, TX, paper SPE 13112
presented at the 1984 SPE Annual Technical Conference and Exhibition, Houston, TX.

40.

Sheffield, J.S. and Sitzman, J.J.: Air Drilling Practices in the Midcontinent and Rocky
Mountain Areas, paper SPE/IADC 13490 presented at the 1985 SPE/IADC Drilling
Conference, New Orleans, LA.

41.

Westermark, R.V.: Drilling with a Parasite Aerating String in the Disturbed Belt, Gallatin
County, Montana, paper IADC/SPE 14734 presented at the 1986 IADC/SPE Drilling
Conference, Dallas, TX.

42.

Claytor, S.B., Manning, K.J. and Schmalzried, D.L.: Drilling a Medium-Radius


Horizontal Well With Aerated Drilling Fluid: A Case Study, paper SPE/IADC 21988
presented at the 1991 SPE/IADC Drilling Conference, Amsterdam.

43.

Teichrob, R.R.: Low Pressure Reservoir Drilled with Air/N2 in a Closed System, Oil &
Gas J. (March 21, 1994) 80-90.

44.

Adam, J. and Berry, M.: Underbalanced Coiled Tubing Sidetrack Successful, Oil & Gas
J. (December 18, 1995) 91-98.

45.

Bobo, R.A. and Barrett, H.M.: Aeration of Drilling Fluids, World Oil (1953) 137, No. 4,
145.

46.

Roy, R. and Hay, R.: Measuring Downhole Annular Pressure While Drilling for
Optimization of Underbalanced Drilling, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

47.

Wilson, J.: Optimizing Drilling of Underbalanced Wellbores with Data Acquisition


Systems, paper presented at the 1995 1st International Underbalanced Drilling Conference,
The Hague, The Netherlands, October 2-4.

3-41

Chapter 3

3-42

Benefits of Drilling Underbalanced

48.

Taylor, J., McDonald, C. and Fried, S.: Underbalanced Drilling Total Systems Approach,
paper presented at the 1995 1st International Underbalanced Drilling Conference, The Hague,
The Netherlands, October 2-4.

49.

Bennion, D.B., Thomas, F.B., Bietz, R.F. and Bennion, D.W.: Underbalanced Drilling,
Praises and Perils, paper presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

50.

Dreiling, T., McClelland, M.L. and Bilyeu, B.: Horizontal and High Angle Air Drilling in
the San Juan Basin, New Mexico, The Brief (June 1996) 12-19.

51.

Francis, P.A., Patey, I.T.M. and Spark, I.S.C.: A Comparison of Underbalanced and
Overbalanced Drilling-Induced Formation Damage Using Reservoir Conditions Core
Flood Testing, paper presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

52.

Schlumberger: Log Interpretation Principles/Applications, Schlumberger Educational


Services, Houston, TX (1987).

53.

Pallatt, N., Stockden, I.L.M., Mitchell, P.S.H. and Woodhouse, R.: Low Invasion Coring
Gives Native Reservoir Water Saturations, European Society of Professional Well Log
Analysts, London (1991).

54.

Rathmel, J.J., Tibbitts, G.A., Gremley, R.B., Warner, H.R., Jr. and White, E.K.:
Development of a Method for Partially Uninvaded Coring in High Permeability
Sandstone, paper SPE 20413 presented at the 1990 SPE Annual Technical Conference
and Exhibition, New Orleans, LA.

55.

Bennion, D.B., Thomas, F.B., Bennion, D.W. and Bietz, R.F.: Formation Damage Control
and Research in Horizontal Wells, presented at the 1993 International Conference on
Horizontal Well Technology, Houston, TX.

56.

McLennan, J.D., Roegiers, J-C. and Economides, M.J.: Extended Reach and Horizontal
Wells, Reservoir Stimulation, Economides, M.J. and Nolte, K.G., (eds.) Prentice Hall,
Englewood Cliffs, NJ (1989).

57.

Bennion, D.B. and Thomas, F.B.: Underbalanced Drilling of Horizontal Wells: Does it
Really Eliminate Formation Damage, paper SPE 27352 presented at the 1994 SPE Intl.
Symposium on Formation Damage Control, Lafayette, LA.

58.

Surewaard, J., de Koning, K., Kool, M., Woodland, D., Roed, H. and Hopmans, P.:
Underbalanced Operations in Petroleum Development Oman, paper presented at the 1995
1st International Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

59.

Churcher, P.L., Yurkiw, F.J., Bietz, R.F. and Bennion, D.B.: Designing and Testing of
Underbalanced Drilling Fluids to Limit Formation Damage: Examples from the Westerose
Field, Canada, paper presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

60.

Morrow, N.R.: Wettability and Its Effect on Oil Recovery, JPT (December 1990) 14761484.

61.

Vavra, C.L., Kaldi, J.G. and Sneider, R.M.: Geological Applications of Capillary
Pressure: A Review, AAPG Bulletin (June 1992) 76, No. 6, 840-850.

62.

Katz, D.L. and Lundy, C.L.: Absence of Connate Water in Michigan Reef Gas
Reservoirs, AAPG Bulletin (January 1982) 66, No. 1, 91-98.

63.

Gray, R. and Bird, K.: Laboratory Evaluation of Underbalance Formation Damage


Compared to Neutral and Overbalance Conditions, paper presented at the 1995 1st
International Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

64.

Stone, C.R.: Horizontal Underbalanced Drilling, SPE Distinguished Lecture, 1993/1994.

65.

Joseph, R.A.: Underbalanced Horizontal Drilling - Conclusion. Special Techniques and


Equipment Reduce Problems, Oil & Gas J. (March 27, 1995) 41-47.

66.

Springer, S.J., Christie, D.S., Flach, P.D., Gust, D.A. and Porter, K.: A Review of the
First 1500 Horizontal Wells in Western Canada, presented at the 1994 SPE/CIM 4th
Annual One-Day Conference on Horizontal Wells, Calgary, Canada.

67.

Mullane, T.J., Churcher, P.L., Edmunds, A.C., Eddy, D.B., Martin, B.G. and Flach, P.D.:
Benefits of Underbalanced Drilling: Examples from the Weyburn and Westerose Fields,
Western Canada, paper presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

68.

Lunan, B.: Underbalanced Drilling - Surface Control Systems, paper presented at the
1995 1st International Underbalanced Drilling Conference, The Hague, The Netherlands,
October 2-4.

69.

Cheung, V., Ahearn, B., Scheidt, K.A., Saponja, J., Penrose, R. and Christensen, D.:
Underbalanced Horizontal Drilling in the Hussar Glauconitic A Pool, presented at the
SPE/CIM 4th Annual One-Day Conference on Horizontal Wells, Calgary, Canada.

70.

Kitsios, E., Kamphuis, H., Quaresa, V., Reynolds, E. and Rovig, J.: Underbalanced

3-43

Chapter 3

Benefits of Drilling Underbalanced

Drilling Through Oil Production Zones with Stable Foam in Oman, paper IADC/SPE
27525 presented at the 1994 IADC/SPE Drilling Conference, Dallas, TX.

3-44

4.1

SELECTING AN
APPROPRIATE TECHNIQUE
Introduction

Chapter 2 summarized various techniques for drilling a well underbalanced. This chapter
summarizes some of the con-siderations for:






Determining if underbalanced drilling is potentially applicable.


Evaluating wellbore stability. A controlling factor in selecting a drilling technique is the
allowable range of borehole pressures.
Selecting the potential underbalanced drilling methods, and,
Evaluating the economics to identify the most cost effective procedure.

Underbalanced drilling is technically feasible in almost all situations. There will be many
instances when it is also the most cost effective procedure. The same drilling technique does not
have to be used for an entire drilling program in a well. There are instances where it can be cost
effective to drill through the overburden conventionally and then switch to underbalanced drilling
to penetrate the productive zones. Conversely, in some instances, it may be advantageous to drill
the upper section of a hole underbalanced and then mud up to drill through a highly productive or
over-pressured pay zone.
There will often be uncertainty in carrying out technical and economic evaluations.
Underbalanced drilling is no different from conventional drilling in this regard. Local experience
with different drilling techniques often minimizes some of these uncertainties. Sensitivity
analyses and risk assessments can be helpful in selecting the optimum procedures. As experience
is gained with a specific drilling technique, operational efficiency increases as does the ability to
evaluate and forecast the associated costs and benefits.

4.2

Potential Applications

The key to successful exploitation lies in application of appropriate techniques. Underbalanced


drilling is a good example. This drilling protocol is not appropriate in all instances. The daily
spend rate for underbalanced drilling can be greater than for conventional drilling methods
because of the use of supplementary well control and surface production equipment. However,
4-1

Chapter 4

Selecting An Appropriate Technique

in many instances, total drilling costs and long-term return on investment are greatly improved.
Underbalanced drilling programs generally require more engineering and planning to ensure
safety and efficiency. The benefits may be downstream of the drilling operations.
Candidate Selection
It is necessary to economically screen underbalanced methodologies. Beyond reduction of
impairment, increased and earlier production, reduced fluid loss, better evaluation ... severe
drilling problems alone may dictate adopting underbalanced drilling methods. Regardless, each
situation must be evaluated individually.
Some of the situations where underbalanced drilling may offer advantages include:







Any formation subject to impairment, and particularly naturally fractured reservoirs drilled
with straight hole or horizontal laterals. Mud damage of the productive fractures can be
devastating when conventional drilling methods are adopted. Certain productive sands and
carbonates may also be severely impaired with conventional drilling.
Deeper reservoirs below depleted or underpressured zones can cause severe drilling
problems when conventional methods are used. These problems may include lost circulation
and stuck drillpipe.
Gas storage wells rely on high production rates during peak demand periods. Minimized
formation impairment is essential to guarantee adequate on-demand deliverability.
Produced water or other disposal wells, particularly where hydraulic fracturing is
environmentally unacceptable, depend on high injectivity potential associated with
minimized drilling impairment.
Better evaluation: Recognition of productive zones while drilling, should lessen the chances
of bypassed pay.
Any situation where ROP can be economically increased and fewer bits are required.

These are only a few of the generic situations where underbalanced drilling can offer significant
advantages. There are some situations where the opportunities are moderate to marginal. In each
case, there may be compensating factors such as higher ROP, reduced list circulation . .
Certainly, these situations require more careful evaluation. For example:




4-2

Low permeability, massive and relatively unfractured reservoirs which will likely need to
be hydraulically fractured.
Very high permeability formations which would produce significant volumes of formation
fluids and exceed surface equipment handling capacity. This is a solvable problem, but it
may not be economic. In this situation, non-invasive, clear fluids are preferred. Clear fluids
with acid-soluble solids are acceptable.

Macrofractured or vugular formations which are susceptible to gravity-driven invasion of


fractures or vugs or would produce volumes of formation fluid which cannot be
accommodated by surface equipment. These difficulties can be overcome by using clean
fluids with soluble diverters (i.e., calcium carbonate chips, rock salt ). In fact, there are
cases where underbalanced drilling is the only method that is appropriate.
Extremely high pressure formations which would require excessively costly fluids or
wellhead and surface equipment to ensure adequate safety. Even in mud drilling, the surface
equipment (BOPs, wellhead equipment, casing) must be designed for maximum shut-in
surface pressure. This design pressure is the same for conventional and under-balanced
drilling. Overpressured formations can be drilled with air mist and foam. Carden, 1997
(unpublished) reported deep drilling at a bottomhole pressure equivalent of 14.4 ppg (~0.75
psi/ft). As long as the production is being vented down the blooie line and there is no
backpressure, a rotating head may be acceptable. The problem with overpressured
formations arises with high production rates. It is more difficult (but not impossible) to kill a
high pressure well, flowing at high production rates. Even normally pressured wells, flowing
at high rates, are more difficult to kill. Either one of these situations may be drilled and
completed with proper design.

There are some situations where under-balanced drilling may not be appropriate:




Highly unconsolidated formations which require elevated wellbore pressure to maintain


hole stability.
Formations where swelling and significant hole size reduction or wellbore instability are
anticipated. Swelling and instability can also be serious problems with conventional muds.

These are generalities. In fact, under-balanced drilling may be attractive even in some of the
scenarios where stability is a real concern. The key to every under-balanced drilling project is
proper evaluation of the prospect and comparison of the cost to drill and complete, both
conventionally and underbalanced. In every case, a well must provide the maximum return on its
investment, including appropriate consid-eration of safety and the environment. Economics is the
final measure of success.
Drilling and production data from offset wells, either individually or together, facilitate
identification of most potential applications for underbalanced drilling. In some instances,
underbalanced drilling is adopted principally to improve drilling performance through one or
more intervals in a hole; overburden or productive zones. Alternatively, using underbalanced
drilling can be driven exclusively by the potential for reducing formation damage in known or
anticipated pay zones.

4-3

Chapter 4

Selecting An Appropriate Technique

Successful implementation of underbalanced drilling methods in offsets is an obvious indicator


of the viability of the technique. If offset wells were successfully drilled using a particular
underbalanced technique, that technique is a strong candidate for future drilling operations.
However, previous success does not guarantee that the technique will be the optimum for other
wells, even wells in the immediate locality. It does not necessarily imply that it was the best
technique for the wells already drilled. It merely demonstrates feasibility.
Penetration Rates
If offset well data indicates low penetration rates for mud (for example, less than fifteen ft/hr)
and the formations are old and hard, dry gas, mist or foam are potential alternatives to mud. The
lower the penetration rate, the more likely it is that these techniques will reduce the drilling times
sufficiently to make them economically attractive. Caution must be exercised since low
penetration rates do not always indicate hard rocks. For example, very soft Gulf of Mexico
shales are often characterized by penetration rates which are less than ten feet per hour, when
drilled with PDC bits and water-based muds. Air, mist or foam drilling would not be suitable for
these formations, principally because of the potential for wellbore instability. In addition, Gulf of
Mexico sandstones are often so permeable that formation fluid inflows, either water or
hydrocarbon, would probably occur at rates which could be difficult to manage on the surface if
they were drilled with dry gas, mist or foam.
Air, Mist and Dry Gas Drilling
Carden, 1993,1 reviewed drilling practices in the continental United States, indicating that air
drilling predominated in the Arkoma and Appalachian Basins, where target zones are
Pennsylvanian or older. Air drilling is also frequently used in the Paradox, Uinta-Piceance-Eagle
and San Juan Basins, where some of the formations are Cretaceous or even Tertiary. This
indicates that age alone is not a deciding factor in selecting air drilling. Inflow rate and
formation integrity are two other deciding factors.
Characteristic Formation Properties
While rate of penetration and formation age are not stand alone diagnostics of environments
where air, mist or foam drilling are appropriate, there are some general guidelines on identifying
appropriate formations.




4-4

The formations penetrated must be sufficiently strong (or at least competent) that the
wellbore will remain stable at the very low pressures associated with these techniques.
Water-sensitive shales are unlikely to be stable if drilled with dry gas, mist or foam.
Dehydration and associated desiccation (drying out and cracking) may occur with dry gas.
Clay swelling can occur with aqueous mists. Elevated pressure may be required to
compensate for borehole creep. Nitrified crude may help.
Wireline sonic logs can provide qualitative indications of integrity and strength. Sonic travel
times (the reciprocals of the velocities) are related to a formations bulk density and dynamic
elastic moduli (characterizing the ability of the formation to deform). However, high elastic
constants commonly correlate with higher compressive strength. Due to the many factors

which influence underbalanced drilling and sonic logging response, it is not possible to
specify a universal threshold for sonic travel time. As a rule-of-thumb, if the compressional
wave travel time is consistently less than 70 s/ft through formations which are known not to
be naturally fractured or water-sensitive, underbalanced drilling with dry gas, mist or foam is
probably an option.
Dry Gas Versus Mist Drilling
Chapter 2 indicated the different environments where these two techniques might be used. It
only takes a small liquid inflow to require that drilling is switched from dry gas to mist. Under
most circumstances, it is desirable to have the additional equipment required for mist drilling on
location, whenever drilling with dry gas. Even before drilling starts, the decision to use mist or
air drilling should be carefully considered.
When Is Foam Drilling Preferred?
Foamed drilling fluids can generally provide better hole cleaning than dry gas or mist. A foamed
drilling fluid should be considered when excessive water inflow has forced dry gas or mist
drilling to be terminated in offset wells, or if substantial water production is anticipated. At
significant depths, for example, below 10,000 feet, or when drilling large diameter hole,
unrealistically high gas injection rates may be required for good hole cleaning with dry gas or
mist. Foam would be a preferable fluid under these circumstances.
Stable foams have proven to be particularly successful for drilling in permafrost (Anderson,
1971,2 Fraser and Moore, 19873) and are a candidate drilling technique when permafrost will
need to be penetrated. Stiffened foams should be considered when a foamed drilling fluid is
indicated and where hole cleaning is particularly demanding or where water supply is severely
restricted. This is often the situation for large diameter holes.
As with other lightened drilling fluids, wellbore instability can be a problem when drilling
through weaker or less competent formations with foam. If penetration rates, when drilling
conventionally, were already acceptable, gasified mud should probably be used (instead of foam).
Lost Circulation
Major problems with lost circulation suggest using underbalanced drilling procedures. If the
interval to be drilled is sub-normally pressured, it will be necessary to use a lightened fluid to
minimize or avoid significant fluid loss. Since many lost circulation zones can also potentially
produce large volumes of water, stable foam is likely preferred over dry gases and mist.
Differential Sticking
Problems with differential sticking in offset wells are an indication that underbalanced drilling
may be desirable.

4-5

Chapter 4

Selecting An Appropriate Technique

Depleted Zones
Any well that will penetrate a depleted reservoir or formation, should be considered as a
candidate for underbalanced drilling. Underbalanced drilling will often reduce the occurrence of
differential sticking and restrict formation damage in these environments. The fractional increase
in penetration rate alone may not be adequate to justify drilling underbalanced. If the formation
pressure gradients are very low, foam may be necessary. Otherwise, a gasified liquid is probably
the preferred fluid.
Permeability Impairment
For target formations which are highly susceptible to drilling damage, and which will not be
stimulated by hydraulic fracturing, underbalanced drilling may offer substantial advantages. This
is particularly true for horizontal wells, which are more likely to benefit from underbalanced
drilling than vertical wells. This is even more true for naturally fractured reservoirs which are
highly susceptible to drilling damage and which are often most effectively produced by
horizontal wellbores, drilled to intersect the fractures.
Depleted and naturally fractured wells are primary candidates for underbalanced techniques.
However, normally pressured and over-pressured reservoirs can also be drilled underbalanced
using lightened drilling fluids (to minimize damage). If underbalanced conditions can be
achieved with a liquid drilling fluid, the additional cost of a lightened drilling fluid is likely only
justified when drilling a long interval of hard rock, where penetration rates would otherwise be
unacceptably low. Similarly, the additional cost for snubbing equipment will normally only be
justified if surface pressures are too high for other more conventional underbalanced drilling
techniques.
The chemistry and physics of formation damage are beyond the scope of this manual. For
additional information, the reader is referred to Allen and Roberts, 19824 (or many other
references).
Multiple Drilling Protocols
Just as it may be necessary to change from air drilling to mist drilling as the water cut increases,
it is not necessary for an entire well to be drilled underbalanced. It is possible for one or more
interval(s) in a well to be drilled underbalanced (with the associated benefits and costs) and the
remaining (or other) zones are drilled conventionally (underbalanced drilling in these other zones
is either uneconomic or technically impractical). There are many instances of depleted or
damage-prone reservoirs, below weak, water-sensitive or unconsolidated zones, where the
overburden has been drilled conventionally and cased off before drilling underbalanced through
the pay. This is common practice for horizontal wells in many fields in central and western
Canada. Also, there are many examples in the Rocky Mountains where intervals are drilled with
aerated mud, and then cased off to drill the pay zone conventionally.

4-6

4.3

Technical Feasibility

In evaluating the feasibility of candidate drilling techniques, a controlling factor is the range of
anticipated borehole pressures which will be required for each zone to be drilled. The upper limit
for underbalanced conditions is, by definition, the pore pressure. The lower limit will generally
be regulated by the lowest pressure at which wellbore stability is ensured and/or the pressure at
which produced fluids can be safely and economically handled at the surface. Often, determining
the anticipated pressure ranges will implicitly dictate the particular underbalanced drilling
technique(s) which could be adopted (refer to Table 4-1).
The first step in evaluating which underbalanced drilling technique to adopt is to determine
the anticipated pressures.
After the pressure ranges are delineated, determine which methods are functional within the
anticipated pressure window. Other wellbore and fluid specifics are then considered. These
include:






Will there be sloughing shales? Are aqueous fluids inappropriate?


Will water producing horizons be penetrated?
Will multiple, permeable zones, with dramatically different pore pressures, be encountered?
What is the potential for chemical formation damage, due to fluid/fluid or fluid/formation
interaction and is this an overwhelming problem, regardless of what wellbore pressure is
used?

Table 4-1.

Underbalanced Drilling Applications and Candidate Techniques

Reason for Drilling Underbalanced

Preferred Underbalanced Drilling Technique

Low ROP through hard rock.

(1) Dry air.


(2) Mist, if there is a slight water inflow.
(3) Foam, if there is heavy water inflow, if the borehole
wall is prone to erosion, or if there is a large hole
diameter.
(4) Nitrogen or natural gas, if the well is producing
wet gas and it is a high angle or horizontal hole.

Lost circulation through the overburden.

(1) Aerated mud, if the ROP is high (rock strength low


or moderate) or if water-sensitive shales are present.
(2) Foam is possible if wellbore instability is not a
problem.

Differential sticking through the overburden.

(1) Nitrified mud, if gas production is likely, especially


if a closed system is to be used.
(2) Aerated mud, if gas production is unlikely and an

4-7

Chapter 4

Selecting An Appropriate Technique

Reason for Drilling Underbalanced

Preferred Underbalanced Drilling Technique

open surface system is to be used.


(3) Foam is possible if the pore pressure is very low
and if the formations are very hard.
Formation damage through a soft/mediumdepleted reservoir.

(1) Nitrified brine or crude:

String injection, if the pore pressure is very low;

Parasite injection, if the pore pressure is high


enough and a deviated/horizontal hole needs
conventional MWD and/or a mud motor.
Temporary casing injection, if the pore pressure is
intermediate and a high gas rate is needed.

String and temporary casing injection, if the pore


pressure is very low and/or if very high gas rates are
required with a closed surface system.

(2) Foam, if the pore pressure is very low and an open


surface system is acceptable.
Formation damage through a normally pressured
reservoir.

Flowdrill (use a closed surface system if sour gas is


possible).

Lost circulation/formation damage through a


normally pressured, fractured reservoir.

Flowdrill (use an atmospheric system if no sour gas is


possible).

Formation damage through an overpressured


reservoir.

Snub drill (use a closed surface system if sour gas is


possible).





Is there a potential for sour gas production?


Are there features of the well geometry which dictate specific underbalanced protocols?
What is the local availability of suitable equipment and consumables (including liquids and
gases for the drilling fluids)?

These issues are considered below.


4.3.1

Borehole Pressure Limits

Pore Pressure
The pore pressure in formations that will be open to the borehole is the upper limit for the range
of borehole pressures which will give underbalanced conditions. The drilling technique adopted
must result in a borehole pressure which is less than the pore pressure in all open zones. This
restriction can only be relaxed for open zones that will not be influenced by overbalance.

4-8

In practice, the borehole pressure at any depth will fluctuate during drilling, principally when
circulation is shut down to make a connection or to trip the string. Chapters 2 and 5 describe
procedures for calculating the changes in borehole pressure that will accompany changes in
circulation. If there is no formation fluid inflow, borehole pressures with dry gas, mist, foam or
pure liquid drilling fluids will normally decrease when circulation is stopped. Therefore, if it is
established that underbalanced conditions will result during circulation, they will likely be
maintained if circulation is stopped, unless there is an overwhelming fluid influx.
As was discussed in Chapter 2, the borehole pressure can change significantly when gasified
liquid is used. It may either increase or decrease, depending on the connection and tripping
procedures. As a generalization, the borehole pressure will probably increase during a
connection when drilling with a liquid gasified by drillstring gas injection. In practice, if it is
generally not possible to maintain a circulating pressure that is 300 to 500 psi less than the
pore pressure, underbalanced conditions may not be maintained during connections, when
drilling with a gasified liquid.
As with all drilling techniques, pore pressure is a controlling element. The difficulties in
estimating pore pressure will not be discussed. It is inevitable that there will be some uncertainty
in the pore pressure estimates. In some cases, this is not a serious problem. For example, during
in-fill drilling in depleted, non-compartmentalized reservoirs, where underbalanced drilling is
commonly used, reservoir pressures are often well known. In other situations, formation pore
pressures will need to be inferred from offset drilling and wireline logging information. In these
instances, the pore pressure that is encountered may be very different from the estimated value.
Best practice is to use a reasonable lower bound estimate of the pore pressure when comparing
candidate drilling techniques (for stability and pore pressure). Wells may be mistakenly drilled
overbalanced if the pore pressure is lower than expected (Bennion, et al., 1995a5). Surface
equipment capacity and drilling specifics should be based on an upper bound estimate of the pore
pressure.
Wellbore Stability
Just as formation pore pressure is an upper limit to allowable underbalanced pressures, a lower
limit may be established by the minimum pressure required to maintain hole stability. In
underbalanced drilling:

When there is an aqueous component in the drilling fluid, water-wet formations may still
imbibe water when drilled underbalanced. This is summarized in Bennion, 1996,6 and
described in detail in Bennion, et al., 1995.5, 7 Precautions for inhibiting the base fluid should
still be taken (when drilling underbalanced) and exposure to diluted formation fluid from
greater depths should also still be considered. Because of adverse capillary pressure
relations, it is possible for the formation to imbibe water-based (and in some cases
hydrocarbon-based) fluids in the near wellbore region ... In an underbalanced drilling
operation, imbibition effects can cause phase trapping and damage problems in a number of
different reservoir scenarios. Formation damage is not the only risk. Instability can be a far
more serious concern.

4-9

Chapter 4




Selecting An Appropriate Technique

The effects of desiccation (drying out), when a formation is contacted with circulating dry
gas, are not well defined in the literature. Imbibition into the formation would seem to be
inhibited. If the water content reduces, the strength may increase (Hale et al., 199311).
However, desiccation cracking may occur during shrinkage and some sloughing may occur.
This may heighten imbibition if water does ultimately contact the shale. In sands, the
strength may also be increased because of increasing capillary forces. Horsrud et al., 1994,8
indirectly referred to the potential of desiccation cracking and its impact on exposure time
and extent.
Less dampening of vibrations may cause greater formation disaggregation uphole from the
bit.
When drilling in underbalance, pore pressure invariably leads to a less-stable situation at the
borehole wall (Bol et al., 1994).9. This is supported indirectly by Tan and Rahman, 1994,10
who stated When drilling under an overbalance condition without an effective barrier
present at the wellbore wall, mud pressure would penetrate progressively into the formation.
Due to the low permeability of shales, the mud pressure penetration would result in an
increase in pore pressure near the wellbore wall. The increase in pore pressure reduces the
effective mud support which leads to a less stable wellbore condition. This time-dependent
stability issue depends strongly on the mud filtrate and pore fluid properties and the rock
material composition. If the corollary is assumed to apply, underbalanced drilling should
reduce drilling problems in shale, provided of course, the effective stress conditions do not
lead to failure.

Failure Modes
Wellbore stability is an equal, if not greater, concern in underbalanced drilling as compared to
conventional operations. Wellbore failure can be caused by one of three generic mechanisms.
These include:
Tensile failure: This occurs if the maximum tensile stress acting around a wellbore exceeds the
tensile strength of the formation. Two types of tensile failure can occur. The first is associated
with increasing the wellbore pressure and generating an hydraulic fracture (lost circulation
situation). The second type of tensile failure is extensional, where concentric failure surfaces can
develop around the wellbore if the pressure in the wellbore is reduced and/or if there are
significant seepage forces due to inflow. This means that material can spall into the wellbore or
can be pushed or dragged into the wellbore by flowing formation fluids, in an environment
where frictional resistance is reduced because there are no compressive stresses acting (the radial
effective stresses, at least, are tensile).
Shear failure: In a frictional material, shear failure is governed by the mean stress (arithmetic
average of the principal stresses), the deviatoric stress tensor (representation of the differences in
magnitudes of the in-situ principal stresses) and the shear characteristics of the medium (often a
Mohr-Coulomb failure envelope, where the governing material parameters are the cohesion - the
inherent shear strength - and the angle of internal friction). Shear failure can occur in an active
or a passive mode. The first mode is associated with the lowest tolerable downhole pressure and

4-10

movement of the yielded material is inwards (towards the wellbore). The second mode requires
larger pressures and movement of the yielded material is outward (into the formation). Failure
planes and directions of sliding along these planes can occur in any direction, depending on the
orientations of the wellbore and the principal stresses and the material properties.
Matrix/Pore Collapse: This is structural collapse. It is governed by the spherical stress tensor
(stress average) and the matrix isostatic compressive strength (strength under hydrostatic stress
application). This failure is associated with a volume reduction and densification of the medium.
It can occur in chalks, weakly cemented sands and certain other high porosity materials, as the
effective hydrostatic stresses increase with a reduction in pore pressure.
The symptoms and results of these instabilities include:








The wellbore may slough because of high stresses (drilling the hole itself concentrates
stresses around the wellbore), causing shear failure and/or spalling. This can be accompanied
by inadequate hole cleaning, pipe sticking, mud rings ... Thermal stresses can mitigate the
effects of elevated in-situ stresses (cooldown reduces stresses) but in extreme cases cooling
itself can cause extensional fracturing.
Initiation and/or extension of hydraulic fractures if the borehole pressure becomes high
enough.
Local mobilization of pre-existing faults if the fluid pressure along the fault plane increases.
Sudden spalling may occur during swabbing or tripping due to transient reductions in the
wellbore pressure.
Rapid production of formation fluids/gas can occur if higher pressured zones are
encountered. Drag forces may cause rapid sand production.
Ductile formations such as salt, other evaporites and soft shales can squeeze and restrict the
hole.
Certain shales can swell, causing restricted hole gauge and accelerated sloughing.

The driller may not be able to identify what type of failure is occurring. For example, it may not
be possible to determine if it is shear failure or matrix collapse. The significance lies in
understanding how a material will fail due to pore pressure reduction. Shear and tensile failure
can jeopardize drilling. Matrix collapse may or may not jeopardize drilling but may lead to
significant skin and dramatically restrict future productivity. Understanding your reservoir is
extremely important.

4-11

Chapter 4

Selecting An Appropriate Technique

Shale Stability
Certain shales present additional problems - chemical instability.
The concerns for chemical stability in shale drilling are similar to those when conventional muds
are used (exposure time, filtrate chemistry, ionic transport...) with the added complication that
dehydration-induced desiccation may be a supplementary problem. In addition, some level of
wellbore integrity is provided by pressure drop through cake or by capillary forces in the
formation during mud drilling; less cake develops in underbalanced operations.
Borehole instability is related fundamentally to inadequate mud weight and to the influx of
water into the formation, which aggravates instability by increasing near-wellbore pore
pressure and by decreasing shale strength. Several mechanisms govern water (fluid)
movement in or out of the shale. The two most relevant mechanisms are the hydraulic
pressure difference, p, between the wellbore pressure (mud weight) and the shale pore fluid
pressure, and the chemical potential differences, , between the drilling and shale pore
fluids.
(Hale et al., 199311)
Smectitic Shales
These are often ductile and can creep. Reduced wellbore pressure will accelerate creep. They are
chemically sensitive to the filtrates ionic strength. True underbalanced drilling probably reduces
ionic degradation and associated swelling. In two-phase systems, it is reasonable to follow the
same procedures that would be used to minimize ionic diffusion when drilling balanced or
overbalanced. The rationale for this is that ionic movement is governed dominantly by
concentration gradients and base aqueous fluids will contact the wellbore surface regardless of
the pressure regime. Inflow will be usually restricted because of the pressure boundary
conditions, but this does not prevent chemical interaction, although transient conditions should
be periodically anticipated.
If the activity of water in the shales pore fluid is lower than the activity of water in the drilling
fluid, water will be imbibed by the shale. The shale will expand, with an associated degradation in
strength. Fragments of shale will slough off the borehole wall and fall into the wellbore. These
cavings can be much larger than typical cuttings; they are not readily lifted from the hole by many
underbalanced drilling fluids. High quality foam can be an exception. Unfortunately, the rate at
which cavings slough into the borehole often does not decrease with time. If they are too large to
be lifted from the borehole, they accumulate around the BHA and can stick the string, often leading
to a catastrophic loss of hole.
Increasing the boreholes hydrostatic pressure may return the hole to stable conditions, but only
temporarily. At this point, shale integrity has been compromised and ROP will be reduced by the
increased hydrostatic pressure.

4-12

Salts, gilsonite additives or polymers can be added to the aqueous phase of a drilling fluid, to inhibit
shale hydration (that can lead to sloughing). This can be prohibitively expensive when drilling with
mist or foamed fluids (i.e., when the liquid is not re-cycled around the well.). Regardless, it may be
very difficult to match the activities of these inhibited drilling fluids with the activities of the
formation, considering dilution with formation water inflows from further downhole.
As a result, mist or foam drilling with an aqueous liquid phase should be avoided if water-sensitive
shales are to be penetrated. If there is no chance of water inflow and the shales are relatively strong,
it may be possible to drill them with dry gas. Alternatively, it may be economic to drill through
water producing zones with mist or foam and then case them off before switching to dry air drilling
to penetrate a deeper water-sensitive shale. An example of this is a well drilled in West Texas
(unpublished). In this well, it was cost effective to case off the water producing formations and
then drill with dry air through a water-sensitive shale that displayed unmanageable wellbore
instability when drilled with mist.
Illitic Shales
These formations, particularly those with significant quartz content, are frequently brittle and
more chemically inert. Creep and swelling are usually less of a problem. Fines migration can
occur if the drawdown is too large (probably more of a formation damage concern than a
wellbore stability issue). Potentially, massive instability can occur during drawdown because of
significant amounts of stored (a characteristic of brittle rock) energy. Natural fracturing may be
significant, providing opportunities for reducing lost circulation by drilling underbalanced.
Failure Criteria
Stability analysis is performed by determining the far-field principal stresses (using whatever
methodology is available; refer, for example, to Kunze and Steiger, 199112), calculating the
stresses which exist around the wellbore due to this stress field and then comparing the existing
stresses with a failure criterion (representing the strength of the material at particular in-situ
stress conditions). If the stress field concentrated around the wellbore exceeds the failure
criterion, yield can occur.
It is necessary to define a failure criterion in order to analyze shear failure conditions. There are
many classical failure criteria (e.g., von Mises, Tresca, Mohr-Coulomb, Drucker-Prager, Hoek
and Brown).13 Steiger and Leung, 1991,14 discussed various other failure/yield representations,
specifically for shales. It is important to choose one that represents the characteristics of the
medium in question. The Mohr-Coulomb relationship is probably the simplest to use. The basis
of this relationship is that a failure locus is developed from laboratory triaxial testing. The yield
or failure stress is characterized as:
= c o + n tan
where:

(4.1)

Lowry and Ottesen, 1993.13

4-13

Chapter 4

Selecting An Appropriate Technique

......... shear stress (one-half of the


difference between the maximum,
and minimum principal stresses),
n ...... effective normal stress (average of
the maximum and minimum
effective principal stresses),
co ........ cohesion (intercept on a - n plot),
and,
......... angle of internal friction (angle of
Mohr-Coulomb locus from the
n axis on a - n plot).
If the in-situ stress conditions are on or above this line (on a plot of shear stress versus normal
stress) failure is predicted.
Risnes et al., 1982,15 and Bratli and Risnes, 1981,16 provided simple analyses for evaluating
stress distributions and yielding of wellbores in saturated, elastic materials. Wang and Dusseault,
1991,17,18 presented a simple elastoplastic model. The analytical relationships in these
publications form a good basis for approximating the wellbore pressures required to maintain
mechanical stability. These computational approaches are described below.
4.3.2

Computational Approaches

Classical Elasticity
Linear elastic theory (Bradley, 1979;19 Zoback et al., 198520) overpredicts the stresses near the
borehole wall. It frequently indicates that almost any moderately competent sand or shale will
have yielded, although field and laboratory evidence suggests that stability is in fact still
maintained. Although these analyses are simple, they do not adequately represent non-linear
behavior and frequently yield and failure are not discriminated. They often provide pessimistic
estimates of stability - they are probably conservative.
Ramos, et al., 1994,21 demonstrated a simple application of elastic theory. This paper is
important because it showed basic elastic formulations, indicated how the basic equations can be
used on a well which is oriented in the direction of a principal stress which is not vertical (a
horizontal completion was modeled), indicated how fluid flow might be simply included and
described the role of core testing and log analysis. Ramos et al., 1994,21 used the following
elastic formulation:

The liquid, steady-state pressure distribution was characterized by:

p( r ) =

4-14

QB r
ln + p w
2 k a

(4.2)

where:
B ........ formation factor (bbl/sbbl),
p(r)
fluid pressure at a distance, r
(feet), from the center of the
wellbore (psi),
Q ........ flow rate per unit length (sBPD/ft),
......... viscosity (cP),
k......... permeability (md),
a ......... wellbore radius (feet),
pw ....... wellbore pressure (psi), and,
......... units conversion factor (221.8).

A critical flow rate was calculated, for steady state conditions, as:

kc o cos
1 sin

(4.3)

where:
........ units conversion factor
(6.806 x 10-3)




The advantage of this very simple formula is that it provides an indication of a critical
stability limit, based on flow (which can be measured) at the surface, rather than strictly on
pressure. Alternatively, it can be expressed in terms of pressure.
For a horizontal well, centrally positioned in a reservoir of thickness, H (feet), the tolerable
drawdown can be expressed as:

P =

2c o cos H
ln
1 sin d h

(4.4)

where:
P ...... tolerable drawdown (psi), and,
dh ....... wellbore diameter (feet). 22

22

This solution applies to slightly compressible fluids. Wang, et al., 1991, addressed gas.

4-15

Chapter 4

Selecting An Appropriate Technique

Wang and Dusseault, 1991,17,18 presented more complicated solutions for two wellbore
situations, accounting for plastic deformation. However, if their solutions are evaluated only up
to the limit of yield, the solutions are useful for providing conservative estimates of required
wellbore pressures. Wang and Dusseault, 1991,16,17 defined two wellbore situations. These were
active and passive yield. Active yield correlates with a reduction in wellbore pressure; the
reduced pressure causes the radial effective stress at the wellbore wall to be less than the
circumferential effective stress. Inward movement occurs and shear yield will eventually occur.
Passive yield (in shear) can occur when the wellbore pressure is increased so that the radial
effective stress at the wellbore exceeds the circumferential stress. Generally, tensile failure
(hydraulic fracturing) will occur before passive shear. Consider the following modified versions
of formulations proposed by Wang and Dusseault, 1991.17,18

For active loading (reduced wellbore pressure) analogous to an underbalanced drilling


situation, at the wellbore wall, with no filtercake:

In an impermeable medium:
= 3 HMAX HMIN p w

r = p w
In a permeable medium:
1 2

= 3 HMAX HMIN + 1 (p w p o ) p w

= p
w
r

(4.5)

where:
....... circumferential (hoop) total stress
(psi),
HMAX . far-field maximum total principal
stress (psi),
HMIN .. far-field minimum total principal
stress (psi),
........ Biots poroelastic constant
(effective stress, , is equal to
total stress, , minus p)
(dimensionless),
......... drained Poissons ratio
(dimensionless),

4-16

These formulations were written for a vertical wellbore. A horizontal wellbore can be represented by
incorporating the vertical stress.

p......... local reservoir pressure (psi),


po........ virgin reservoir pressure (psi),
pw ....... wellbore pressure (psi), and,
r........ radial total stress (psi).
 For an active state of stress, two other parameters were defined:
Sa =

2c o cos
1 sin

1 + sin
Na =
1 sin

(4.6)

Assuming a Mohr-Coulomb failure criterion (Equation 4.1), active failure can be predicted
when:
N a r + Sa = 0

(4.7)

For an impermeable medium:

p sa =

3 HMAX HMIN + Sa p o (1 N a )
1+ Na
(4.8)

where:
psa

is the minimum wellbore pressure


that is allowable before shear failure
is predicted (psi).
For a permeable medium (with =1):

1 2

p
3 HMAX HMIN + Sa + 2

1 o
p sa =
1 2
2
1
where:

(4.9)

po........ far-field reservoir pressure (psi).


If these calculations are used and psa > pw the wellbore is predicted to be at risk of shear
failure.

4-17

Chapter 4

Selecting An Appropriate Technique

Figure 4-1 is an example of tolerable pressures to avoid shear failure, based strictly on linear
elastic analysis (using Wang and Dusseaults formulations). The reservoir simulated was
assumed to have a depth of 5000 feet, HMAX = 4000 psi, HMIN varied, po = 2165 psi (0.433
psi/ft), = 35, and = 0.25. The cohesion varied. The figure indicates that cohesion is a
dominant parameter and every effort should be made to determine it. It also shows differences
between permeable and impermeable reservoirs. Wellbore pressures below the predicted value
(psa) indicate instability in shear. These predictions will generally be conservative. Similar
formulations were developed for passive failure. The reader is referred to the cited papers.
Detournay and Cheng, 1988,23 have shown poroelastic formulations for creation of hydraulic
fractures. This would be passive loading, with tensile failure.

In a porous, permeable medium, the breakdown pressure, Pb, can be approximated.


Hydraulic fracturing and associated lost circulation are predicted to occur if the wellbore
pressure exceeds this value.

pb =

To + 3 HMIN HMAX 2 p o
2(1 )

1 2
,
2(1 )

(4.10)

where:
pb ....... the breakdown pressure (psi), and,
To ....... tensile strength (psi).

4-18

Minimum Tolerable Wellbore Pressure (psi)

7000
Minimum Principal Stress 4000 psi

6000

Minimum Principal Stress 3500 psi

Permeable

Minimum Principal Stress 3000 psi


Minimum Principal Stress 2500 psi

5000

4000
Impermeable

3000

2000
Reservoir Pressure

1000

0
0

200

400

600

800

1000

1200

1400

1600

1800

2000

Cohesion (psi)

Figure 4-1.

Minimum tolerable pressures to avoid shear failure.

If no fluid penetrates the formation, the breakdown pressure can alternately be expressed as:

p b = To + 3 HMIN HMAX 2 p o (1 2 ) p i
(4.11)
where:
pi ........ pore pressure at the borehole wall
(behind a filtercake, if any).
Detournay and Cheng, 1988,23 also demonstrated the potential for time-dependent collapse
occurring in conjunction with equilibration of pore pressure, after drilling.
Simple elastic stability approximations are shown in Table 4-2.
Relationships such as these have been developed for deviated wells. (Bradley, 1979,19 Brudy and
Zoback, 1993;24 McLennan et al., 198925).

4-19

Chapter 4

Selecting An Appropriate Technique

Even if a material has yielded and is deforming plastically, stress is still carried by the postyielded media. Catastrophic collapse or spalling may not necessarily yet occur. The calculations
are further complicated by pore pressure changes due to transient flow. Wang and Dusseault,
1991b,18 presented quasi-analytical re-lationships which can be used to represent the additional
seepage forces that act during flow into the wellbore.
Table 4-2.

Possible Formulae for Estimating Mechanical Stability

Failure Mode

Permeable
Formation

Extensional
(Due to
Flow)

Active Shear

Impermeable
Formation

Equation

Q=

kc o cos
1 sin

Equation
Number

(4.3)

(4.9)
1 2

3 HMAX HMIN + S a + 2
p

1 o
p sa =
1 2
2
1

Active Shear

Passive
Tension
(Hydraulic
Fracturing)
Passive
Tension
(Hydraulic
Fracturing)

p sa =

3 HMAX HMIN + Sa p o (1 N a )
1+ Na

pb =

To + 3 HMIN HMAX 2 p o
2(1 )

(4.8)

(4.10)

p b = To + 3 HMIN HMAX 2 p o (1 2 ) p i (4.11)

Non-Linear Elastic Approaches (Pressure-Dependent Modulus and Damage Mechanics)


Classical predictions have usually assumed that the reservoir behaves as a linear elastic material,
with failure being predicted by comparing the stresses at the borehole wall, calculated using
elastic theory, with the peak strength of the rock. Recognizing the conservative character of these
analyses, Santarelli et al., 1986,26 developed a closed-form solution for the stresses and strains
around an axisymmetric wellbore in elastic rock, where the modulus was a function of the
confining pressure. This model indicated:
 The tangential stresses near the wellbore may be much lower than predicted by elastic
analysis, and,


4-20

The maximum tangential stress may not occur at the wellbore wall.

These imply that elastic analyses alone may be conservative. Non-linear elastic models and
damage mechanics models (assessing microstructural degradation due to stresses; refer for
example to Pellegrino et al., 199427) delimit a zone of reduced stresses around wellbores and can
even indicate extensional, circumferential failure. The basis of these models (Santarelli, et al.,
1986;26 Nawrocki and Dusseault, 199528) is to calculate the effective stress regime, with the
added stipulation that Youngs Modulus is a function of the effective secondary minimum
principal stress. The resulting stress field is compared to the failure criterion adopted to assess if
yield has occurred. Preferably, non-constant pore pressure distributions away from the wellbore
(at least transient and preferably steady-state as well) should be incorporated into the model to
represent seepage forces and more realistic radial variation of effective stresses.
Elastoplastic Approaches
These models predict a yielded zone around the wellbore, but they do impart some load-bearing
capability to these zones and the wellbore may remain intact. Elastic-plastic models are an
approximation of material behavior, but they do at least allow yield to occur. Numerical
modeling is generally required, unless restrictive assumptions are adopted. Finite element
models are usually used (Vaziri and Byrne, 199029; Wang, 199530). They have the advantage
that, pending adequate knowledge of material properties and an expert to perform the analysis,
various three-dimensional geometries, and assorted behavioral laws (creep, elastic-plastic, flow,
non-isothermal, ionic transport...) can be simulated. The relationships between the existing
stresses and the associated deformations are governed by constitutive relationships. This is
beyond the scope of this document. The reader is referred to Charlez, 1994;31 Ewy, 1991;32
Detournay, 1986;33 and, Veeken et al., 1989.34
Chemo-Mechanical Models
Some models have been developed, trying to integrate inflow, ionic diffusion and timedependent mechanical behavior. The calculations are usually complex, requiring numerical
modeling and a significant effort in determining material parameters. As Wong and Heidug,
1994,35 stated, the chemical effect results from modification of the strength of the structural and
hydration forces operating between closely spaced (<5 nm) and hydrated clay particles that
accompanies invasion of water-based drilling fluids. The model described by these authors used
an extension of Biots poroelastic theory to incorporate fluid chemistry and a modified version of
Darcys law to accommodate osmotic flow. A similar, but simpler, tactic (incorporating
hydration forces into the mechanical stress equations) has been proposed by Bol et al., 1994.9
Field Testing
Surewaard et al., 1996,36 discussed instability during underbalanced drilling in a heavy oil reservoir
in Oman. An attempt to drill a horizontal hole with foam had met with wellbore instability
problems in a different formation. Caliper data in the unstable shale above the target reservoir
showed how the hole diameter changed with different degrees of overbalance. These indicated that
the hole could enlarge considerably through this shale (they predicted a twenty-inch diameter if the
drilling was underbalanced). A field trial was performed, in which the proposed underbalanced
drilling fluid was circulated through a 100 foot interval of hole, drilled through the formation
(shale) with the most potential for wellbore instability. Caliper logs showed the hole evolving from
4-21

Chapter 4

Selecting An Appropriate Technique

8.2 inches to 20 inches in diameter, over that period. In order to drill the reservoir underbalanced,
an additional casing string would need to have been set across the unstable shale. Eventually, it was
decided that the incremental costs of this and other changes made underbalanced drilling
uneconomic.
The difficulties in forecasting stability can be significant. Field calibration of predictions is
often required.
An Example
Even though elastic analyses are approximate (usually overestimate the minimum wellbore pressure
for shear and underestimate the maximum wellbore pressure for hydraulic fracturing; except if
natural fractures are present), they are useful for preliminary scoping of stability situations.
Consider a vertical well with the following characteristics:
depth ............................................... 5000 feet
HMAX ...............................4000 psi (.8 psi/ft)
HMIN .............................................. unknown
co .......................................................1000 psi
................................................................35
po...................................2165 psi (.433 psi/ft)
To ........................................................500 psi
................................................................ 0.3
Figure 4-2 shows elastic stability limits for shear and tensile failure, as a function of different
values of the minimum horizontal stress. The intersection of the area beneath the hydraulic
fracturing line and above the shear failure line is predicted to be stable. Bottomhole pressures
should fall in this range for stable drilling.
Hydrocarbon Production Rate and Surface Pressures
In some circumstances, the maximum hydrocarbon production rate that can safely be handled at the
surface will enforce a lower limit on the borehole pressure that can be tolerated.
For gas production, the pressure inside the surface system will be controlled by the frictional
pressure loss upstream of the flare. If the surface system incorporates an atmospheric mud/gas
separator, the backpressure must not exceed the limit set by the height of the liquid seal. Otherwise,
gas may escape into the open liquids collection tanks, with potentially dangerous consequences.
With a closed surface system, the backpressure must not exceed the pressure limitation of the multiphase separator vessel.
There will be maximum tolerable liquid production rates for both atmospheric mud/gas separators
and for multi-phase, closed separators. The liquid return rate is the sum of the returning (drilling
fluid) liquid rate and the produced liquid rate. For either open or closed surface systems, an
excessive liquid return rate will cause inefficient gas separation, with gas carrying over into the
liquids collection tanks or liquid carrying over to the flare.
4-22

In principle, surface equipment can be designed to handle almost any conceivable production rate.
If the surface system can handle the production under absolute open flow rate conditions, combined
with the injected drilling fluids, the production rate or surface pressures will not place any lower
limit on the allowable borehole pressure while drilling. In practice, equipment availability will be a
factor. It is unlikely that it would be economic to construct an oversized surface system from
scratch.
5000
Shear Failure
Tensile Failure

Allowable Wellbore Pressure (psi)

4500
4000
3500

Shear Failure If Wellbore Pressure


Is Below This Curve

Stable

3000
2500
2000
1500
1000

Tensile Failure If Wellbore Pressure


Is Above This Curve

500
0
2000

2200

2400

2600

2800

3000

3200

3400

3600

3800

4000

Minimum In-Situ Stress (psi)

Figure 4-2.

Variation of allowable wellbore pressure, estimated using linear elastic theory, for an
example 5000 foot deep well.

Formation Damage
Another limit to the tolerable drawdown may be the onset of formation damage caused by fines
mobilization in the producing formation. Determining this limit is complex (and often highly
inaccurate); involving the reservoir rock mineralogy, petrology, permeability, strength and
formation fluid rheology, as well as the prevailing pressures. Laboratory experiments, using core
under simulated downhole conditions, are probably the only practical way to estimate what
drawdown can be tolerated before significant damage occurs.

4-23

Chapter 4

Selecting An Appropriate Technique

Backpressure
If the combined return rate of absolute open flow production and injected drilling fluid cannot be
handled by available surface equipment, it is possible to restrict the production rate by taking
returns through a choke before entering a separation system. This imposed backpressure will
increase the borehole pressure, which will in turn reduce the rate of influx. The allowable
backpressure is limited by the pressure ratings of the surface equipment upstream of the choke, the
choke itself, the wellhead and the surface casing ... The lowest working pressure limit is almost
invariably provided by the diverter system (i.e., a rotating head or RBOP) in use. RBOPs were
developed specifically to increase the ability to use surface pressure to control downhole influx
while drilling.
When using compressible drilling fluids, it is usually more cost effective to switch to a higher
density drilling fluid than to choke back the well. Applying a backpressure will increase the gas
injection pressure. It will also increase the gas injection rate required for acceptable hole
cleaning, when drilling with dry gas or mist. Both of these effects will increase the cost of the gas
supply; air compressors, natural gas or nitrogen ...
It is usually possible to increase the borehole pressure by reducing the gas injection rate when a
gasified liquid is used. Manipulating the surface backpressure is usually not necessary in this
situation.
When drilling with foam, backpressure may be necessary to maintain the foam quality within
acceptable limits throughout the wellbore.
The most benefit from choking back can be derived for liquid drilling fluids, in order to increase the
borehole pressure. This is often done to control the rate of production during the latter stages of
drilling a horizontal interval. It can permit continued use of simple drilling fluids, such as water,
without incurring the incremental costs for a weighted drilling fluid.
Once the maximum tolerable surface pressure is reached, production rate can only be further
reduced by increasing downhole pressure by increasing the effective density of the drilling fluid.

Implications of Drilling Technique Selection


Drilling pressure requirements
can be different at different
depths in a well.

4-24






Pore pressure gradients vary with depth.


Formation strength will vary with depth.
In-situ stresses will vary with depth.
The tolerable stresses, particularly from a stability point-of-view are affected by the
inclination and orientation of deviated, extended reach and horizontal wells. The
underbalance pressure at the onset of instability commonly de-creases as the hole angle
increases.

Due to these various factors, it is often not possible to represent the borehole pressure required to
prevent instability with a single, simple pressure gradient. Also, as was shown in Chapter 2, the
effective pressure gradient for a compressible drilling fluid increases with increasing well depth. A
thorough assessment of the potential for wellbore instability for different underbalanced drilling
techniques involves comparing predicted/ calculated circulating pressures and limiting borehole
pressures, at different depths throughout the interval(s) in question.
Production rates depend on the length of reservoir that is open to the wellbore and on the
underbalance pressure. The underbalance pressure (the drawdown) giving the maximum tolerable
production rate will be smallest at TD in the reservoir. In vertical wells, the amount of
underbalance will generally increase with increasing depth. This is not the case in horizontal wells.
In these wells, frictional pressure losses up the annulus will increase with measured depth, while
hydrostatic pressures depend on the true vertical depth. The circulating borehole pressure will be
higher at the bit than at the start of the horizontal section. Since the virgin pore pressure will be
nominally the same along the horizontal section, the underbalance pressure will generally be
smaller at the bit than further uphole. As a result, the rate of production while drilling a horizontal
well may not increase in direct proportion to the length of the horizontal section. Nevertheless, the
total production should be greatest when the hole is at TD. From the point-of-view of technique
selection, it is normally sufficient to confirm that the candidate drilling technique will not allow
excessive production at TD.
Once the borehole pressure limits, corresponding to wellbore instability and excessive
production rate, have been determined, a first pass evaluation of different drilling techniques
can be performed by using the ranges of equivalent circulating densities (ECD). If the range of
ECDs shown for a particular technique does not give a borehole pressure between the pore
pressure and the higher of the wellbore instability and excessive production limits, that
technique is unlikely to allow safe underbalanced drilling, without choking back the return flow.
This simple screening method is illustrated in Figures 4-3 through 4-5 (Examples 1 through 5), to
follow. These schematics show the approximate ranges of borehole pressure achievable with
various underbalanced drilling techniques, plotted against true vertical depth. These schematic
plots (Figures 4-3 through 4-5) are for depths up to 10000 feet. There is no reason why
underbalanced drilling should not be successful at depths well in excess of this. Note that the ranges
of achievable pressures shown are only approximate. With lightened fluids and gases, the borehole

4-25

Chapter 4

Selecting An Appropriate Technique

pressure, for a given combination of injection rates, will not increase linearly with depth. It will
increase more rapidly than this, due to the compressibility of the gaseous phase. The lower limit for
the borehole pressure when drilling with foam will almost certainly be higher than that shown for
depths in excess of 5,000 feet if the fluid is to remain a foam to the surface. A single limiting ECD
is used in these illustrations to avoid undue complexity. In practice, computer simulation of the
circulating pressures is essential for any potential foam application at depths much greater than
5,000 feet.
Example 1
Figure 4-3 shows a shallow, normally pressured (8.35 ppg or 0.433 psi/ft) reservoir. There are no
wellbore instability concerns, from the surface casing point, at 700 feet, to TD, at 4,000 feet. The
proposed surface system can handle the anticipated absolute open flow rate of the well. Minimal
water inflow is expected. Since stability is not a problem, any gaseous or lightened drilling fluid
could be used, and technique selection should be based principally on economic analysis (and
certain other considerations given in a Section 4.4). In practice, dry air drilling would probably be
the most economic drilling technique. Misting might be necessary to avoid mud-ring formation if
slight water inflows are encountered. If water inflow becomes excessive, a stable foam might be
needed.
4500
4000

Borehole Pressure (psi)

3500
3000
Liquid
Gasified
Liquid

2500
Foam With
Backpressure

Pore Pressure
(8.35 ppg)

2000
1500
1000

Foam

Tolerable
Borehole
Pressure

500

Mist

Gas

0
0

2000

4000

6000

8000

True Vertical Depth (feet)

Figure 4-3.

4-26

Stability regimes for the well described in Example 1.

10000

4500

4000

Borehole Pressure (psi)

3500

Pore Pressure Limit

3000

Gasified
Liquid

2500

Foam With
Backpressure

Liquid
2000

1500

Tolerable
Wellbore
Pressure

1000

Foam

Mist

Gas

500

Wellbore Stability Limit


0

2000

4000

6000

8000

10000

True Vertical Depth (feet)

Figure 4-4.

Stability regimes for the well described in Example 2.

4500
Pore Pressure Limit

4000

Borehole Pressure (psi)

3500
3000

Tolerable
Wellbore
Pressure

Liquid

Gasified
Liquid

2500
2000
Wellbore Stability Limit

1500

Truncated Area For


Tolerable Wellbore Pressure
DueTo A Production Rate Limit

Foam With Backpressure

1000
Foam
Mist

Gas

500
0
4000

5000

6000

7000

8000

9000

10000

True Vertical Depth (feet)

Figure 4-5.

Stability regimes for the wells described in Examples 3 through 5.

4-27

Chapter 4

Selecting An Appropriate Technique

Example 2
Figure 4-4 illustrates a hole interval from 2,000 to 6,000 ft which includes a depleted sandstone
from 3,000 to 4,000 feet. The pore pressure gradient in the depleted sandstone is 5 ppg (~0.26
psi/ft). The pore pressure gradient above and below the sandstone is 8 ppg (0.416 psi/ft). Drilling
conventionally with mud encounters significant problems
with lost circulation and differential sticking across this formation. The entire hole interval is
comparatively stable, and mechanical wellbore instability is not anticipated, if the ECD is
maintained above 2 ppg (~0.10 psi/ft). The depleted sand produces gas at low rates and the surface
system can comfortably handle the absolute open flow. This hole could be drilled with foam, either
stable or stiffened, or with a gasified liquid. In the latter case, high gas injection rates would
probably be required; air would probably be the only economic option for the gasifying medium.
Example 3
The next example is for an interval that encompasses overburden and a reservoir. The pore
pressure, gradient is equivalent to 8 ppg (0.416 psi/ft). This situation is shown in Figure 4-5. There
is a section of shale (from 6000 to 8000 feet) above the reservoir (from 8000 to 9000 feet) in which
it is estimated that a borehole pressure gradient equivalent to 7 ppg (0.364 psi/ft) is required to
avoid wellbore instability. The reservoir itself is competent, and wellbore instability is not expected
to be a problem unless the borehole pressure gradient falls below 5 ppg (0.260 psi/ft) equivalent.
The reservoir is, however, prolific. By the time the target TD (9,000 ft TVD) is reached, it is
estimated that the maximum tolerable drawdown would be 500 psi. Assuming that the pore
pressure at this depth would be 3,744 psi (i.e. 0.416 psi/ft x 9,000 feet), the minimum allowable
borehole pressure would be 3,244 psi. This corresponds to 6.93 ppg (0.360 psi/ft). If the
overburden is uncased when the reservoir is drilled, the complete section requires that the
equivalent circulating density of the drilling fluid should be in the range 7 to 8 ppg, using a low
density liquid, such as diesel or crude oil. Gasified water could also be used, but the target
pressures are at the upper end of those normally developed when drilling with gasified liquids. For
a gasified liquid, the volume fraction of gas required to maintain the borehole pressure gradient
above 6.93 ppg (0.36 psi/ft) would not be large. It might not be possible to operate in the frictiondominated flow regime. Operating in the friction-dominated regime is preferable for controlling the
borehole pressure. The possibility for drilling with a gasified liquid would require more
sophisticated evaluation of the circulating pressures.
Example 4
Suppose that the maximum tolerable drawdown in Example 3 was estimated to be only 100 psi.
This would set the lower borehole pressure limit at 3,644 psi (7.79 ppg) at TD. A drilling fluid
formulated with diesel or crude oil would give pressures lower than this. Plain water would be too
dense to give underbalanced conditions. This is little different from the situation in many Austin
chalk wells, that have been successfully flowdrilled with plain water. In these Austin chalk wells,
lost circulation has caused the onset of production and the hydrostatic pressure exerted by the
mixture of water and formation fluid is sufficiently low that production has continued thereafter. If
flow cannot be induced in this way, it may be possible to drill underbalanced with a light
hydrocarbon drilling fluid and to choke back flow, to keep the production rate below the limits of
the surface equipment. Any decision to drill like this would require detailed analyses of the

4-28

circulating pressures, incorporating produced fluids and the surface system capacity, to confirm that
this technique could be safely implemented. Experience with flowdrilling in offset wells might
provide sufficient confidence that production rates would not become excessive.
Example 5
A final example is a further modification of Example 3. Suppose that the reservoir had been
depleted to a 6.5 ppg (0.338 psi/ft) equivalent gradient. If the maximum tolerable drawdown
remained at 500 psi, the tolerable range for equivalent circulating density through the reservoir
would be 5.4 to 6.5 ppg. A gasified liquid would be required. This would not provide sufficient
support to avoid wellbore instability in the shale above the reservoir. It would be necessary to set
casing at the top of the reservoir to drill it underbalanced. This would certainly impact the drilling
economics.
When evaluating highly productive formations, more detailed numerical analyses of circulating
pressures are required than the simplistic ECD considerations described above.

Computation of borehole pressures and production rates, for different drilling fluid injection
schemes, is complex. Formation fluid inflows interact with the drilling fluid to change the
borehole pressure. This in turn leads to changes in the inflow rate. When circulation is
stopped to make connections or to trip the string, produced fluids will tend to lift the remnant
drilling fluid from the annulus. This changes the borehole pressure and the production rate.
Choking back the well returns further complicates calculation of borehole pressures and
production rates. If the drilling fluid is incompressible, the choked back circulating pressure
will be the sum of the circulating pressure at that depth, without the backpressure, and the
imposed backpressure. If the annular fluid contains any gas, compressibility normally causes
the borehole pressure to increase by more than the imposed backpressure. Determining
borehole pressure and production rates for different fluids, rates and backpressures normally
requires computer simulation.
It is inevitable that there will be uncertainty in simulator input parameters. Sensitivity of the
predictions to variations in these input parameters is required. In many circumstances, the
overall uncertainty in the input parameters can be so great that the effort required for repeated
computer-based simulations may not be justified, at least for technique selection.

Water Production
The flow of formation water into the
borehole can influence the selection
of a suitable drilling technique.

4-29

Chapter 4

Selecting An Appropriate Technique

Production of even small volumes of water can make dry gas drilling difficult. A mud ring can
form as damp cuttings collect, usually at the top of the BHA where the annular velocity is lowest. It
is common to switch to mist, or even foam, if a water inflow is encountered. When offset wells
indicate that formation water inflows are probable, the operator should not expect to drill below
the water producing zone with dry gas.
When misting, higher air injection rates are required to lift the water from the hole. The air rate
must be sufficiently high to prevent slug flow. Slug flow can damage the borehole and surface
equipment. The high air rate, in combination with the weight of water in the annulus, significantly
increases the standpipe pressure. Boosters are often needed to increase the gas delivery pressure
when substantial water inflows are encountered. More compressor power is required. If nitrogen
or natural gas are used as the gas phase, the gas supply cost will be greatly increased.
Certain wells can tolerate large water influxes and slugging, during mist drilling. Surface holes
have been drilled in Terrel County, Texas, with water flows greater than 1000 BPH. Since the
water was fresh, surface disposal was approved. In central Australia, fresh water flows into mist
drilled wells occurred at 2000 BPH. Surface disposal was also possible here. In both of these cases
the wellbores were stable enough that slugging was acceptable.
It may not be feasible to lift large volumes of water from the well by misting. At some stage, the
additional gas injection rate and pressure required become impractical. When this occurs, foam is
preferable. There is no unique water inflow rate above which foam drilling is preferable to misting.
The appropriate time for changing fluids depends on the availability of additional air compression
power when drilling with air mist, and on the cost of the gas supply when drilling with other gases.
Since other gases, such as nitrogen or natural gas, are much more costly than compressed air,
misting with these gases at lower rates is less economic than drilling with air mist.
Hole size also influences the impact of water inflow on required gas injection rate and pressure.
Increased cross-sectional area reduces annular velocity and hole cleaning efficiency, although large
holes can usually produce more water before the gas injection pressure becomes impractically high.
Carefully managed foam can lift considerable volumes of water from a well - as much as 500
BWPH has been reported by Shale, 1995.37 When large water inflows are anticipated, dry gas or
mist drilling may not be appropriate, even if wellbore stability and hydrocarbon production rates
indicate that these drilling fluids would be nominally acceptable.
In many locations, the cost of waste water disposal precludes mist drilling. Foam may or may not
be an alternative. Since borehole pressures will tend to be somewhat higher when foam is used
instead of mist or dry gas, the water inflow rate will probably be somewhat lower. The reduction in
inflow rate from normally pressured aquifers will not, however, be dramatic and disposal volumes
may not be substantially different. For example, the formation fluid pressure at 4,000 feet in a
normally pressured reservoir will be approximately 1,750 psi. Mist drilling might generate a
borehole pressure of 50 to 100 psi before any water inflow (assume 75 psi for illustrative purposes),

4-30

whereas a pressure in the range 500 to 1,000 psi might be developed when drilling with foam
(assume 750 psi for illustrative purposes). Using these wellbore pressures, the underpressure
(drawdown) with mist would be 1750 - 75 = 1,675 psi (not considering any borehole pressure
increase due to produced water) and 1750 - 750 = 1,000 psi for foam. This example suggests that
the water inflow rate with foam would be expected to be roughly 60% of that when drilling with
mist (presuming Darcy flow with the inflow rate proportional to the drawdown). However, higher
liquid injection rates are used in foam drilling. Considering these compensating factors (higher
pressure but greater base fluid load), the volume of waste water is often fairly similar for the two
different drilling techniques. When water disposal costs prevent mist drilling, it is unlikely that
foam drilling will be more cost effective.
In situations where water production rates are too high for mist or foam drilling, the main
alternatives are to drill underbalanced with aerated liquids or low density liquids, or to mud up and
drill overbalanced. In some instances, it is possible for a well to make water when drilled
underbalanced (for example, with dry air) and to lose circulation when drilled with liquid. In such a
case, it is worth considering drilling with dry air until the reserve pit is filled, switching to drilling
with the produced water until the reserve pits are depleted and then switching back to dry air. There
are only a limited number of applications where this may be possible. It has been cost effective in
the Black Warrior Basin (Graves et al., 199638). This practice should only be attempted when the
operator is confident that the uncertainties of drilling blind can be tolerated when water
circulation is lost.
Multiple Permeable Zones
More than one permeable zone
concurrently open to the wellbore can
complicate selecting the drilling
technique.

If all zones are to be drilled underbalanced, the circulating pressure must satisfy the borehole
pressure requirements for all open, permeable zones, simultaneously. Several factors can prevent
this from occurring. For example:





The equivalent circulating density of compressible drilling fluids increases with increasing
depth along the hole.
In vertical wells, it is possible for a permeable zone close to the bit to be overbalanced when
a permeable zone higher uphole, with the same pore pressure gradient, is underbalanced.
In highly deviated and horizontal wells, the effect may be more pronounced. The annular
pressure drop increases with distance along the hole whereas the pore pressure increases with
vertical depth. Particularly in horizontal hole sections, the borehole pressure at the bit can be
markedly higher than at the start of the horizontal section, even though the formation pore
pressure is the same throughout the section.

4-31

Chapter 4

Selecting An Appropriate Technique

There may be changes in the borehole pressure gradient along the wellbore, either due to
overpressure or due to production from another zone(s). Deis et al., 1995,39 cited an example
where a horizontal well in a compartmentalized reservoir, penetrated one portion of the
reservoir overbalanced and a second portion underbalanced.

The major concern when drilling underbalanced with more than one permeable zone open to the
wellbore is the potential for an underground blowout. In an underground blowout, formation fluid
flows uncontrollably from one permeable formation into another formation uphole, rather than up
to the surface and out of the well. If the borehole pressure is lower than the formation pressure in
all open, permeable zones, formation fluids will flow from all zones into the borehole and there will
be no crossflow from one zone to another. Care should be taken to ensure that changes in borehole
pressure, occurring when circulation is suspended for connections or tripping, or when the well is
shut-in, do not allow crossflow to occur. If this cannot be guaranteed, the consequences of charging
the formations uphole with formation fluids from downhole must be assessed carefully. The impact
is likely to be more serious when drilling through gas-producing formations. If crossflow cannot be
tolerated, the options include:





Using a different drilling technique that allows all permeable zones to remain underbalanced
(if such a technique can be identified),
Killing the well before suspending circulation (this may defeat the purpose of underbalanced
drilling for reducing formation damage), or,
Changing the casing scheme so that the upper formation(s) are cased off before penetrating
the lower zone(s).

Whether drilling under- or overbalanced, it must always be possible to mud up to kill the well,
without losing circulation or fracturing the formation. This may not be possible if the formation
fluid pressure in an open permeable zone is greater than the pressure at which circulation is lost into
another open zone or at which some portion of the openhole can be fractured. Correct selection of
casing points will prevent this situation. (It should only occur if the pore pressure or fracture
gradients are significantly different from those assumed when designing the well. If this does occur,
a well control program must be implemented). This problem does not occur exclusively when a
well is drilled underbalanced. The same situation can arise if a well is drilled conventionally, with
mud in the hole.
Sour Gas
The potential for producing hydrogen
sulfide to the surface always needs to be
considered when selecting a drilling
technique.

4-32

There must be no possibility for hydrogen sulfide to be released into the atmosphere while the well
is being drilled or completed. Beyond being extremely toxic, hydrogen sulfide (H2S) is also
combustible and very corrosive. If any is produced during drilling, it should be discharged into a
suitable flare. This can, in principle, be done safely if the drilling system is a dry gas or mist and if
all returns from the well are flared at the discharge end of the blooie line.
Gaseous hydrogen sulfide can become entrained in any liquid phase in the drilling fluid. At the
surface, it must be separated from produced and returned liquids Conventional mud/gas separators
and open liquid tanks should normally not be used if hydrogen sulfide bearing formations are to be
drilled underbalanced. The separation process must be completed in a closed vessel, with all
separated gas being routed to a suitable flare. Otherwise, hydrogen sulfide may be released into
the atmosphere. The requirement for controlling hydrogen sulfide release effectively mandates the
use of a closed surface system, if there is any possibility of producing sour gas while drilling.
Sour gas can also become entrained in a foamed drilling fluid. Complete separation of a foam into
its constituent gas and liquid phases can take a considerable time - hours - if no measures are taken
to kill the foam. Unless effective defoaming can be guaranteed, foams cannot be used with closed
systems and, as a result, foams should not normally be used when sour gas may be encountered.
Operators and all rig site personnel must adhere to all appropriate regulations. Appropriate
safety precautions are essential.
Drilling / Reservoir Fluid Incompatibility
From time to time, it can be difficult to avoid temporarily creating overbalanced conditions in a
well drilled with a gasified liquid, particularly when making connections and trips. Temporary
overbalance is also possible when drilling with a low-density liquid if, for example, a gas-producing
well is shut in and gas is allowed to migrate up the well.
This is a problem because steps are normally not taken to improve the fluid loss properties of
underbalanced drilling fluids. Mud cake will nominally not be deposited on the borehole wall when
the well is underbalanced. In fact, formation fluid flow, into an underbalanced wellbore, will tend
to remove any solids (for example drilled solids) remaining from previous transient overbalanced
interludes. Consequently, flow of liquid from the drilling fluid into the formation is unlikely to be
restricted when overbalanced conditions (unintentionally) occur. There is also the possibility for
counter-current imbibition of the liquid phase of the drilling fluid into the reservoir, if a water-wet
reservoir is below its irreducible water saturation or an oil-wet reservoir is below its irreducible oil
saturation. Presuming periodic overbalance or counter-current imbibition, it is important to ensure
that the drilling fluids liquid phase is compatible with the formation when a liquid or a gasified
liquid is used to drill through the reservoir (Bennion et al., 1995a;5 Surewaard et al., 1996;36 and
Deis et al., 199539).

4-33

Chapter 4

Selecting An Appropriate Technique

Laboratory testing can help to formulate an inert fluid system. Simple index tests (for example,
capillary suction testing) can be performed, or, alternatively, cores from target formations can be
exposed to candidate drilling fluids, under pressure regimes that are representative of conditions
that might be experienced while drilling and completing the well (Bennion et al.,1995a,5 Deis et al.,
199539).
It is very unlikely that overbalanced conditions will occur in a well that is being drilled with foam.
Counter-current imbibition could cause the aqueous phase of a foamed drilling fluid to enter the
formation, although there are no reports of this having happened. Therefore, there is probably less
need for comprehensive core evaluations, unless some form of foam block is possible.
Evaluating drilling fluid compatibility is probably unnecessary if dry gas is to be used as the drilling
fluid or if only the overburden is to be drilled underbalanced (unless chemo-mechanical degradation
is an issue).
Hole Geometry
Hole geometry can influence selection of an optimized drilling fluid in several different ways.

4-34

As has been discussed, the equivalent circulating density of compressible drilling fluids
increases with increasing depth. A gasified liquid, with an ECD of 2 ppg at 2,000 feet, can
develop an ECD of 5 ppg at 6,000 feet (refer to Figure 2-41). In other words, a fluid that
would give underbalanced conditions at one depth may not be able to provide comparable
underbalanced conditions at a greater depth, when the pore pressure gradient is the same.
Annular gas injection only reduces the density of the fluid column above the injection point.
Since the injection point is at or just above the shoe of the last string of casing, it is not
possible to develop a low ECD at the bottom of a long vertical section, with annular gas
injection. Drillpipe gas injection may be necessary if long vertical sections are to be drilled
with gasified liquid.
The increase in ECD with increasing depth means that it can be difficult to maintain
acceptable foam quality from the hole bottom to the surface, during deep drilling with foam.
Backpressure may be required to maintain the foam quality up the annulus within limits that
permit adequate hole cleaning. This increases the gas supply requirements, and can make
foam unattractive for drilling at great depth. However, foam has been used satisfactorily to
drill to 15,000 feet (Dupont, 198440) - the limitation is more economic than technical.
The effect of increasing hole diameter on hole cleaning efficiency is a major consideration.
The drilling fluid flow rate, required to achieve a given annular velocity, increases in
proportion to the annulus area. Good cleaning in a 17-inch diameter hole typically requires
circulation rates that are approximately four times greater those required for an 8-inch
diameter hole to the same depth. This geometric effect can restrict the use of those drilling
fluids that rely on high annular velocity for hole cleaning. This means that high viscosity
drilling fluids can become more attractive as the hole diameter increases. Foams, particularly
stiffened foams, and gasified muds are often the preferred fluids for drilling holes with
diameters of 17-inches and larger.

When drilling large diameter holes, the pressure capacity of diverter equipment, capable of
passing large diameter bits, may restrict the choice of drilling technique to those that do not
require any significant backpressure. RBOPs with through bore dimensions of eleven inches
(up to 6.75 inches with the kelly packer in place) are off-the-shelf items. Additional sizes
may be available.

When drilling with dry gas, reverse circulation can provide good cleaning in large diameter holes,
without unacceptably high gas injection rates. This technique is described in Chapter 2, Section
2.1.7. It does require special surface equipment and cannot, at this time, be regarded as an
established, routine, drilling technique. It may, nevertheless, be an option for drilling large
diameter holes with very low borehole pressures.
There are appreciable technical limitations on directional drilling equipment suitable for
underbalanced drilling. In particular, mud pulse telemetry MWD systems cannot operate with a
compressible drilling fluid in the drillstring. Liquids, or liquids gasified by annular injection, are
the only underbalanced drilling fluids that can be used with a mud pulse MWD system. Other
MWD systems or procedures can be used for liquids gasified through the drillstring.
Underbalanced drilling in directional and horizontal wells is discussed in more detail in Chapter 6.
Naturally Fractured Formations
Stable foam (stiff foam is even better) can be used at much lower annular velocities than other
underbalanced drilling fluids. Beyond cleaning performance, reduced wellbore erosion is
sometimes cited as a benefit of foam drilling. However, in competent rock, there are few
applications where significant erosion is likely if other underbalanced drilling fluids are used. This
may not be the case when drilling through naturally fractured formations. In fractured formations,
high viscosity drilling fluids, circulating at low annular velocities, can be beneficial in controlling
borehole enlargement (Santarelli et al., 199241). For this reason, stiff foam may be a preferred
candidate for drilling underbalanced through naturally fractured formations. The high viscosity and
good hole cleaning capacity of a stiffened foam will also help lift fragments of naturally fractured
rock that may have spalled off during drilling.
Logistics
Logistical issues, associated with certain locations, can often influence what fluid system is
selected. For example, in some desert locations, water supplies may be restricted and/or very
expensive. In these environments, a drilling technique that minimizes water requirements would be
preferred. A stiff foam may be more economic than its stable counterpart because (refer to Chapter
2, Section 2.6) it can be used at lower liquid injection rates. Drilling techniques that permit recirculation of water would also be desirable.
Availability and access also influence the choice of the gaseous phase for compressible drilling
fluids, as well as wellsite equipment. Cryogenic nitrogen is generally uneconomic in a remote
location. Natural gas can be attractive when drilling in a gas-producing field, particularly if local
gas gathering and processing facilities can recover some of the gas used.

4-35

Chapter 4

Selecting An Appropriate Technique

The specific location type is a consideration. Offshore operations can generally accommodate
much less supplementary surface equipment than land-based operations. The equipment that is
used in underbalanced drilling operations onshore may not be suitable for use offshore. New,
underbalanced drilling equipment is becoming available for offshore deployment (particularly
closed systems). Nesa et al., 1995,42 reported on some of these developments, designed to comply
with the relevant authoritys requirements for offshore application. These modular drilling systems
would be closed. Generically, the main components are an RBOP, a customized choke manifold, a
well fluid processing unit (separators, drilling fluid tank with degasser equipment ...), a nitrogen
injection unit and monitoring equipment.
The high cost of new offshore drilling operations and production facilities means that these fields
must be extremely productive if they are to be economic. This makes them unlikely candidates for
drilling with dry gas or mist. However, in mature, offshore fields, there are situations where the
infrastructure platform and pipeline systems are in-place, have been amortized and production is
declining. These will provide future opportunities for assorted, underbalanced drilling
technologies.

4.4

Economic Analysis

Introduction
The use of an underbalanced drilling technique is driven by different considerations in different
applications. For example, drilling cost reduction is the only justification for drilling underbalanced
through an interval of unproductive overburden. However, if there is a possibility of productive
zones within the overburden (that are not easily recognized on logs), underbalanced drilling can
identify these. Drilling a pay zone underbalanced will be justified if the total cost to bring that
interval onto production is lower than it would have been for conventional drilling methods. Even
if the well cost is increased by drilling underbalanced, in many instances, avoiding formation
damage can increase production rates sufficiently that the net present value of the well is increased
despite its higher cost. On the other hand, if the productive formation is not susceptible to drillinginduced damage and there are no other potential zones to be evaluated, underbalanced drilling must
reduce the well cost.
When evaluating drilling techniques that have not been used previously in a particular region, there
is uncertainty in predicting the factors that influence costs; such as penetration rates. It is often
useful to assess how sensitive any predicted benefit, i.e., reduced well costs or an increase in net
present value, is to the various assumptions made in its prediction. It can also be useful to
determine how much penetration rate or productivity has to increase if an underbalanced drilling
technique is to be economically beneficial, compared to conventional drilling; and then to decide
whether or not that represents an achievable target.

4-36

Variations in cost and productivity are inevitable, even when adjacent wells are drilled with
nominally identical techniques and procedures. Statistical analysis can help to compare the cost
performance of wells drilled using different techniques. If performance increases are not large, it
may be necessary to drill a number of wells with each technique before any firm conclusions can be
drawn about their relative merits.
Testimonials

Anderson, 1971,2 was an early proponent of foam drilling, in Arctic locations where foams
insulating properties minimized thermally-induced hole enlargement.

Cost of drilling in this remote area with mud was calculated to be $227/hr. Cost of
drilling with stable foam was $275/hr, but resulted in two and one-half to three times
more hole [per hour].

It is estimated that the use of stable foam on the second well resulted in a savings of
$12,000, even after allowing for the cost of transportation and standby time due to
weather.

Another operator drilled in this area using straight air. He had good penetration rates
without excessive hole enlargement, but air requirements were five times those used with
stable foam. On a 16-hr/day operating basis, the costs [daily] for air drilling would be
$915 as compared to $468 for stable-foam drilling. The transportation cost in and out of
this remote area would, proportionately, favor stable foam since it requires only two units
[compressors] as compared to six units for air drilling.

Anderson, 1984,43 advocated that recycling of foam on long-term foam drilling operations
can improve economics significantly. Savings of 25 to 50 percent in foaming agent and
additive usage were claimed. (Defoaming agents cannot be used if the foam system is to be
reclaimed). Anderson also cited that foams could improve the economics of recompletions in
situations where fluid loss is a severe problem in underpressured reservoirs. Surveys
indicate that foam recompletions showed a 30 to 50% savings over new well costs and
reduction in formation damage resulted in a 33% gain in oil production when compared to
new wells drilled with mud.

Bentsen and Veny, 1976,44 also found drilling pay with pre-formed, stable foam was
advantageous:

Mud costs were reduced.

Time-consuming conventional DST procedures and costs were eliminated by providing


a system that gave a continuous formation evaluation during drilling.

Immediate and accurate formation evaluations were possible by eliminating damage and
the need for lengthy, expensive formation cleanup operations.

Formation damage was less and completion costs were significantly reduced.

Comeau, 1995,45 presented information from carbonate reservoirs:

4-37

Chapter 4

The Welton and its satellite fields are exploited by 36 producing wells. Of its typical
daily production of around 3000 BOPD 40% of this is currently produced by the five
side-tracked wells drilled in the last two years. The two wells which were drilled
underbalanced produce 40% of the side-tracked wells production and 15% of the total
field production.

Cummings, 1987,47 discussed advantages of natural gas drilling in the San Juan Basin:

In carbonate reservoirs, treatments can be effective to dissolve solids that are plugging
the formation [from overbalanced drilling]. In sands, fewer effective treatment options
are available. The cost for such a service ranges from $100,000 to $300,000 Canadian
dollars depending upon the length of the well and the type of treatment required.

Crearar, 1995,46 stated:

Selecting An Appropriate Technique

Natural gas drilling in the Basin Dakota field near Blanco, New Mexico has eliminated
lost circulation potential and has increased penetration rates over conventional mud
drilling techniques. Refinements in drilling programs have helped reduce drilling costs
which are 20-30% less than mud drilling in lost circulation areas.

Deis et al., 1995,39 described drilling operations in western Canada. Underbalanced drilling
was definitely advantageous in the Weyburn area (refer to Table 4-3).

It has been observed, especially in the Weyburn project, that underbalanced drilling
methods can be employed not only cost effectively, but can even improve drilling
conditions and performance.

Penetration rates in the horizontal section averaged slightly below 20 m/hr in the
overbalanced wells and 25 m/hr in the underbalanced wells. Also, hole condition appears
superior in the underbalanced wells. This is apparent in that hole conditioning trips are
required on occasion to reduce drag only on the overbalanced wells.

The overbalanced wells require that each lateral be soaked in a bleach solution to
remove filter cake after drilling has been completed. This process requires re-entering the
first lateral. Re-entering laterals after a trip has not been a common problem on the
underbalanced wells.
Mullane et al., 1995,48 provided supplementary information on the Weyburn field. They
indicated that if acid stimulation is required to remove skin damage in recently drilled
overbalanced wells, these wells cost more on average than the underbalanced wells. They
also cited situations where the advantages of underbalanced drilling were not clear. In the
Westerose area, the incremental cost to drill underbalanced is approximately $100,000
(Canadian). The gas produced while drilling was found to be approximately three times that
produced during drill stem tests on offset wells that were drilled overbalanced using gel chem
or extended gel mud systems. However, these wells needed to be hydraulically fractured
anyway. If reduction of wellbore formation damage is the only motive for underbalanced
systems where fracturing will ultimately be required, the advantages may be marginal.

4-38

Table 4-3.
Comparison of Drilling Costs in the Weyburn/Lougheed Areas
39
(after Deis et al., 1995 )

Weyburn/Lougheed Well Type

Cost to Drill and


Case
($ Canadian)

Drilling Days

Single Laterals (16 wells)


Underbalanced

486,000

9.4

Single Laterals (2 wells)


Overbalanced

496,000

11.3

Dual Laterals (11 wells)


Underbalanced

721,000

11.6

Dual Laterals (4 wells)


Overbalanced

621,000

14.9

Incremental amounts of $16,000 and $30,000 were included to swab the wells clean before
pipeline tie-in, for the overbalanced single and dual laterals, respectively.

Russell, 1993,49 discussed significant improvements in drilling large diameter surface holes
through hard formations prone to severe lost circulation, in the Yemen Republic. Large
diameter holes were drilled through 1000 feet of hard, fractured, limestone. Modified
hammer bits were used with powerful industrial hammers. Polymer-enhanced foam systems
were used to remove cuttings. By changing foam properties, air injection rate, bit design,
hammer size and the depth at which the hammer drilling assembly was replaced with a rotary
drilling assembly, the time required to drill 17-inch surface holes was reduced by 65%.
Similarly, the time required to drill 26-inch surface holes was reduced by 35%.

Saponja, 1995,50 compared over- and underbalanced drilling in the Glauconitic A Pool of
the Hussar field, in south-central Alberta. The cost per foot for mud drilling was ~$200/ft;
the cost for underbalanced drilling was ~$115/ft.

Whitely and England, 1986,51 provided cost comparisons between air and conventional mud
drilling.

Most air drilling is done with three-cone carbide button bits and various types of
pneumatic HTs [hammer tools]. This method provides improved ROP over conventional
mud drilling. However, the recent introduction of a carbide-insert FPB/HT [flatbottomed bit/hammer tool] has dramatically improved ROP over the three-cone bits and
HTs previously used.

4-39

Chapter 4

Selecting An Appropriate Technique

The average cost per foot for a 12-inch hole was reduced from $35.48 to $18/ft - a
49% reduction.

Westermark, 1986,52 published figures for a situation where parasite string aeration was used
as the lightening technique. The average cost per foot without aerating was $145/foot. While
drilling with a parasite aerating string, the average cost was $67/ft.

Factors Affecting The Economics of Underbalanced Drilling


As the previous testimonials have shown, underbalanced drilling may be justifiable on the basis
of lower drilling costs, on higher drilling rate and lower completion cost, with increased
productivity and reserves. Each case needs to be evaluated individually to determine whether
underbalanced drilling is technically and economically feasible. The factors controlling the
overall drilling economics are summarized below. Many of these factors have been addressed in
previous chapters. Some are restated here for completeness.
Penetration Rate
The variables affecting penetration rate include the drill bit, the bit weight, the rotary speed, as
well as the bottomhole cleaning and mud properties.
Bit Selection
This will be discussed further in Chapter 5. Some general remarks are pertinent. Fear, 1996,53
discussed methodologies for optimizing bit selection and the drilling program in general. For
example, the influence of bit design on ROP, with PDC bits, probably only becomes significant
when the type of drilling fluid and rock lead to bit cleaning problems or when the cutters become
worn. For roller cone bits, the evidence suggests that varying the jet nozzle arrangement may or
may not influence ROP, depending on efficiency of bottomhole cleaning.
Generally, using air instead of conventional mud allows drilling higher footage per bit and
decreases the number of trips per well. In a fairly deep well, the rig time savings from this could
be substantial.
Bourgoyne, 1995,54 indicated that air or gas drilling can be used for intervals of a borehole that
have a high rock strength and a very low permeability, such that the borehole will not collapse
and the well cannot flow. The drilling rate possible with air or natural gas is usually at least
twice as fast as that with clear water and four times as fast as that with mud. The economic
counterbalance is that multiple compressors may be needed to provide the necessary air pressure
and flow rates. Also, small amounts of formation hydrocarbons mixed with compressed air can
be explosive and spontaneous combustion can occur down-hole. Further, if water is produced,
mist drilling may be required. Drilling rates with foam are generally less than with air but more
than with water or mud. De-pending on the capability of the formations to produce water, a misttype flow pattern could be more economical than foam.54

4-40

During air drilling, diamond-enhanced hammer bits, with diamond-coated tungsten carbide
inserts, offer improved cutting structure wear resistance. These bits have improved bit life and
penetration rate and have reduced the cost per foot in the Arkoma and Appalachian Basins
(Reinsvold et al., 198855).
Bit Weight and Rotary Speed
For many formations, an increase in bit weight and rotary speed will increase the drilling rate.
The relationships depend on specific bit types and the formations being drilled. In general, the
footage that a bit can drill tends to decrease with increasing rotary speed. Simple formulations
are available for assessing optimum conditions. Moore, 1974,56 provided a concise and understandable discussion. Beyond wear, increase in drilling rates must be tolerable, in order to
provide adequate hole cleaning. Main-taining adequate hole cleaning capabilities strongly affects
the logistics of under-balanced drilling. For adequate cleaning, sufficient bottomhole velocity
and/or viscosity is required and may require optimization of the bit type.
Mud Weight
Beyond formation damage issues, mud weight reduction probably is the greatest advantage of
underbalanced drilling systems. An increase in mud weight will decrease the drilling rate. There
may be limiting situations in some shales where increasing the mud weight beyond a certain level
does not cause significant change in the rate of penetration. Figures 4-6 and 4-7, from Moore,
1974,56 illustrate this premise. In both of the series of the field tests in Figures 4-6 and 4-7, the
degree of improved performance occurred primarily because the hydrostatic pressure was less
than the formation pore pressure.
Cost reduction, by minimizing lost circu-lation, is a very important consideration in evaluating
an underbalanced drilling program.
Completions and Stimulation
Underbalanced drilling is often merited (technically and economically) if it reduces formation
damage, hole problems, and reduces costs of stimulation or cleanup programs in moderate or
high permeability or fractured formations. If underbalanced drilling is used to avoid formation
damage, the total drilling and completion program should be designed so that the well will never
have to be killed, prior to being put on production. Mitigation of formation damage in
formations which will need to be hydraulically fractured (except if naturally fractured) may be a
poor and unnecessary economic decision.
Also, in certain circumstances, higher grade surface casing may be required, although the
potential for underground blowout in shallow situations is probably more strongly controlled by
the formation breakdown pressure. Pipe grade should also be a consideration in flowdrilling and
mudcap drilling or if corrosive fluids are expected. The opposite situation is also true, as was
discussed by Grace, 1975.57 In the Anadarko Basin of the Texas Panhandle and western
Oklahoma, the Morrow sands of Pennsylvanian age can be abnormally pressured while shallow
formations are almost always subnormally pressured. The routine practice in the early 1960s
was to set an intermediate string of casing on top of the Morrow interval in order that any
abnormal pressure might be properly controlled. This practice, combined with the stratigraphic
nature of the Morrow sands, resulted in unacceptable economics. Consequently, operators and
4-41

Chapter 4

Selecting An Appropriate Technique

drilling contractors adopted the practice of drilling into the Morrow interval without setting
protective pipe. Statistically, in approximately one in four Morrow tests, sands will be present,
and in approximately one in 10 Morrow sands, the productivity will be sufficient to cause a
drilling problem.57 This practice requires training and equipment for handling a controlled
blowout (flowdrilling) situation.
Drilling Days
0

20

40

60

80

100

120

1000

Drilled With Mud


Drilled With Gas

2000

3000

Depth (feet)

4000

5000

6000

7000

8000

9000

10000

Figure 4-6.

4-42

Gas and mud effect on drilling time (after Moore, 1974 56).

Rotating Time (hours)


0

10

20

30

40

50

60

70

80

90

100

500

Drilled With Air


Drilled With Water

Depth (feet)

1000

1500

2000

2500

3000

Figure 4-7.

Air and water effect on drilling time (after Moore, 1974 56).

Formation Evaluation
Real-time formation evaluation (refer to Chapter 6, Section 6.7) can offer significant advantages
in interactive well design. TD can be altered if a diminishing return on the investment is
diagnosed during drilling. Underbalanced coring is also possible. There is a potential for catching
more pristine core (at least without substantial saturation alteration). Core integrity in weak, high
GOR formations may be jeopardized by underbalanced recovery. Pressure transient analysis in
nominally pristine formations is also possible (i.e., reduced skin should improve measurement
quality and reduce time). Coring, logging and testing can be more cumbersome in underbalanced
situations, although they are generally possible. Bloys et at., 1994,58 provided a good discussion
of formation evaluation during mudcap drilling of offshore carbonates:




Wireline logging and conventional coring have been attempted in carbonates with massive
lost circulation when pressure control was maintained with a floating mudcap. However,
both these operations are considered extremely hazardous under these conditions.
A completely virgin formation production test of a mudcap drilled carbonate (to measure
productivity without acidizing to remove plugging materials) is only feasible if the whole
section of interest is mudcap drilled and then cased without incurring losses. Planning to
mudcap drill whole intervals to permit virgin formation production testing is only feasible
if a ware ship [supply ship] is planned to provide the volumes of mud required to mudcap
drill continuously.

4-43

Chapter 4

Selecting An Appropriate Technique

A future alternative to conventional coring with a floating mudcap is coring with a 9.5 m
long wireline retrievable coring system with an inner core barrel that can be run through the
drillstring The core is relatively small (1.71- to 2-inches) and there can be some diameter
restrictions for passing the barrel. Nevertheless, wireline coring with a floating mudcap
would be safer than conventional coring because no tripping of the drillstring is involved to
recover individual cores Until a combination LWD/wireline coring drilling assembly is
available for floating mudcap coring of massive lost circulation intervals, such intervals must
be drilled twice to get both logs and cores. The first time drilling with an LWD logging tool
in the string and the second time, after side tracking, coring with the wireline retrievable
coring system.

Environmental Savings
Pit construction and reclamation costs can make up a significant part of a drilling budget because
of increasingly stringent government regulations and surface owner concerns. Pitless drilling is
possible when a closed system is used (either over-or underbalanced) (refer to Chapter 2, Section
2-11). In assessing cost-benefit, the estimates of cost should include costs for ancillary
equipment, flocculation unit rental (for reconditioning, if necessary), labor and chemicals. The
cost (operational and rig-up) of additional rig equipment, such as the centrifuge, feed pump,
special tanks, and supplementary mechanical separators should be included. Estimates must also
include all labor, equipment and chemical costs (typically polymer, acids, bases and coagulants).
Astrella and Wiemers, 1996,59 indicated that the use of highly automated flocculation units
results in maximum cost savings to the operator in continuous drilling programs because the
equipment has a high utilization rate. The high utilization rate permits the capital cost to be
amortized at a reasonable daily rental rate. It is uneconomical for the expensive equipment to
stand idle. The intangible side of this is guaranteeing performance and equipment reliability.
Further, in selecting a closed system, there must be adequate capacity to keep up with maximum
anticipated penetration rates. Insufficient capacity can force a reduction in penetration rate or
require the digging of an emergency pit.
After a closed loop system has been designed, it is essential to compare its cost with the cost of
constructing and reclaiming reserve pits. The longer it takes to drill a well, the less
economically beneficial a closed system will be, unless disposal is an issue. There are other
considerations. State regulations may impact the use of centrifuged solids. It may be possible to
recycle these solids in the drilling operations (fluid). Also, recycling water reduces transportation
costs and may eliminate the need for a fresh water pit.
Water disposal is another consideration. As Carden, 1993,1 stated,
Ordinarily, water influx is not a drilling problem; rather it is a disposal problem. The reserve pit
will hold a limited quantity of water. When the pit is full air drilling must be discontinued or the
water must be hauled off location to a proper disposal site. Hauling water to disposal can be very
expensive. Disposal costs for reserve pit water can range anywhere from $1.00 to $10.00 per
barrel depending upon the solids content, salinity and the distance to the disposal. The operator
can determine the amount of water that can be hauled off each day. If it costs less per foot than
mud, it is economical.
4-44

Fluid Type
The bottom line controlling factor may be the specific fluid system adopted. Each fluid type has
technical and economic advantages and limitations (refer to Table 4-4).
Cost Comparisons
Some line item cost comparisons are available in the literature. For example, Allan, 1994,60
compared nitrogen and natural gas drilling, and Tag, 1995,61 compared the expenditures for
cryogenic nitrogen and membrane generation.
Case 1 (after Allan, 1994 60) Nitrogen versus Natural Gas
The nitrogen drilling system [membrane generated] eliminates the downhole fire risks
associated with air drilling in hydrocarbon producing formations while significantly reducing
costs as compared to pipeline gas (methane) drilling or trucked liquid nitrogen drilling.
Typically, wells that must be gas drilled through productive intervals rely on pipeline gas,
expensive trucked liquid nitrogen, or air-water injection (mist) systems. Water mist injection is
unacceptable in horizontal wells as the extended drilling time in producing intervals greatly
increases the probability of downhole fires.
Table 4-4.

Costs and Savings For Various Underbalanced Systems

Drilling Method or
Fluid System

Savings

Problems and/or Potential


Expenditures

Air

High penetration rates and


reduction in rig time.
Low bit cost.
Low water requirement.

Possible problems if water flow is


encountered.
Hole erosion, if poorly consolidated.
Possibility of downhole fire, if
hydrocarbons are encountered.
Supplementary equipment rental.
Is not suitable for H2S.

No mud removal.
Low additives cost.
Gas
(Nitrogen or
Natural Gas)

High penetration rates and


reduction in rig time.
Low bit cost.
Low water requirement.
No mud removal.
Low additives cost.

Problems if water flow is encountered.


Cost of gas and/or rentals.
Hole erosion if poorly consolidated.
Cost is high if a market for the gas exists.
Rig safety.
Supplementary equipment rental. If H2S
is expected, consider a closed system.

Mist

High penetration rates and


reduction in rig time.
Low bit cost.
Low water requirement.
No mud removal.

Problems if substantial water flow is


encountered. Gas costs if air not used.
Hole erosion if poorly consolidated.
Shale stability.
Disposal of waste water/gas and

4-45

Chapter 4

Selecting An Appropriate Technique

Drilling Method or
Fluid System

Savings

Modest additives cost.


Stable Foam

4-46

High penetration rates and


reduction in rig time.
Low bit costs.
Low water requirements.
High solids carrying capacity.
Good hole cleaning capability.
Compatible with oil, salt water,
calcium carbonate and most
formation contaminants.
Can safely entrain a considerable
volume of gas into aqueous foam,
rendering it non-flammable until
sumped.
Can handle large flows of water.

Problems and/or Potential


Expenditures
supplementary rental cost.
Air-mist not suitable if H2S is present.
Equipment rental.
Considerable foamer cost. Gas costs if
air not used.
Careful metering required.
Specialized metering equipment.
Defoaming.
Considerable cost.

Separation and disposal.

Water disposal.

Table 4-4.

Costs and Savings For Various Underbalanced Systems (continued..)

Drilling Method or
Fluid System

Savings

Problems and/or Potential


Expenditures

Stiff Foam

High penetration rates and


reduction in rig time.
Low bit cost.

Considerable mud and chemical costs.


Gas costs if air not used.
Fluid degradation possible if oil, salt
water or calcium chloride are
encountered.
Defoaming.

High solids carrying capacity.


Good hole cleaning capability.
Gasified Liquids

Higher bottomhole pressures.

Expense of running a parasite string or a


temporary casing string.
Higher gas rates are required.
Slow pressure response if a parasite
string is used.
Low underbalance pressure may cause
transient departures from underbalanced
conditions and advantages to impairment
reduction may be lost.
Tool problems with drillstring injection.

Improved directional drilling in


comparison to dry gases or mist
(refer to Chapter 6).
Reduced drillstring wear.
Supplementary surface equipment.
Reduced potential for downhole
Corrosion potential (and requirement for
fires in vertical holes with aqueous inhibitors62) if air is used.
fluids.
Flowdrilling

Higher borehole pressures reduce


the possibility of instability.
No gas supply system.
Conventional mud motors and
MWD units can be used.

Supplementary surface equipment and


safety measures.
Excessive production is possible.
Safety issues associated with oil and gas
on drill site.

Mudcap Drilling

Can be used in situations where


surface pressure is too high for
flowdrilling.

Supplementary equipment and safety


considerations.

Snub Drilling or CT

Can be used at pressures too high


for conventional units and
underbalanced drilling equipment.

Snubbing or CT unit.

Closed Systems

Environmental savings.
Can handle H2S. Better
monitoring of returns.

Equipment rental and operating costs.


Cannot be used with explosive mixtures.

4-47

Chapter 4

Selecting An Appropriate Technique

Natural gas drilling and membrane nitrogen drilling were compared for certain wells drilled in
the San Juan Basin, in New Mexico, where severe lost circulation was an issue.

General Assumptions
Flowrate................................................ 3000 cfm
Gas Price ............................................. $2.00/mcf
Trucking Distance .................50 miles (one way)
Drilling Hours/day ...........................................20
Average Gas Drilling Days/well ......................12
Diesel Usage/hour/unit..................... 10.7 gallons
Diesel Fuel Price ............................. $0.80/gallon
Standby Days (Equipment)/well ........................4

Nitrogen Drilling System Cost


Compressors (8) @ $135/unit/day ...........$12960
Boosters (2) @ $200/unit/day
(air use) ......................................................$4800
Membrane Skids (2) @$1500/unit/day
(1800 cfm/skid) ........................................$36000
Trucking/Transportation ............................$9200
Fuel (delivered)
25,680 gallons x $0.80/gallon ..................$20540
Mist Pump ..................................................$1500
Equipment Standby (4 days) ......................$1800
Total Nitrogen Drilling Cost/well ...........$88600

Pipeline Gas Drilling Cost


Pipeline gas 43.2 mmcf @
$2.00/mcf .................................................$86400
Booster (2) $300/unit/day (gas use) ...........$7200
Drill Gas Unit (installed on location).........$1000
Gas Line (2000 feet) ..................................$1800
Trucking/Transportation ............................$1800
Fuel (delivered)
5138 gallons @ $0.80/gallon .....................$4110
Mist Pump ..................................................$1500
Equipment Standby (4 days) ........................$700
Total Pipeline Gas Drilling
Cost/well.................................................$104510

Case 2 (after Tag, 199561) Liquid Nitrogen versus Membrane Nitrogen


Fried and MacDonald, 1995,63 compared options for nitrogen supply in underbalanced drilling.

4-48





Although operationally simple, the cost of the supply of the liquid nitrogen to the wellsite
can represent a significant expense to the total UBD program.
Under optimum circumstances, the use of natural gas can be the most cost effective method
for UBD programs with the only cost being the compression equipment. This can be
minimal if a high pressure feed supply is available at the wellsite.
The suitability of nitrogen membrane systems to high pressure, short duration applications is
not good. Only when the equipment is used over prolonged periods, at high utilization can it
be made cost effective on land. Offshore applications are more affected by support logistics
and obviously lend themselves to membrane technologies.
The systems largest operating expense is the cost of fuel for the air compression units
The advantage of the nitrogen membranes versus a liquid nitrogen system is the ample supply
of free nitrogen available in the air versus the cost of liquid nitrogen and the required
transportation to site.
The process of gas recycling can be cost effective with the previous systems in very specific
applications but is both technically and operationally challenging for most UBD programs.

Tag, 1995,61 provided estimates of the costs for liquid nitrogen and nitrogen generated on
location. These are indicated in Table 4-5.
Economic Analysis
Precise guidelines for economic analysis are too specific to individual operators. Some general
guidelines/examples are provided below.




On the basis of available technology, select the potential drilling systems to be evaluated.
Tabulate the tangible and intangible costs for each system.

Table 4-5.

61

Comparison of Liquid Nitrogen and Membrane Nitrogen (Tag, 1995 ).

Item

Liquid N2

Portable Nitrogen Generating


System

Drilling Program

90 days

90 days

Nitrogen Requirement

1500 scfm

1500 scfm

Duration of nitrogen
requirement

240 hrs (10 days)

240 hrs (10 days)

Nitrogen Purity

Minimum 95% (by volume)

Minimum 95% (by volume)

Nitrogen Pressure

5000 psi

5000 psi

Nitrogen Requirement

1500 scfm x 60 min/hr x 24 hr/day


x 10 days = 584,000 sm3

1500 scfm x 60 min/hr x 24 hr/day x


10 days = 584,000 sm3

4-49

Chapter 4

Selecting An Appropriate Technique

Item

Liquid N2

Portable Nitrogen Generating


System

= 834,000 liters liquid N2


= 139 tanks
Method of Nitrogen Supply

Trucked in Liquid N2
(equipment rental)

On-Site membrane
(equipment purchase )

Logistics

139 liquid N2 tanks, 1 evaporator


and 1 diesel skid (141 containers)

4 skids maximum, 14 tonnes each, 1


power unit, 14 tonnes (5 containers)

Cost of Utilities
(liquid N2, electricity,
diesel)

$1,284,000

Electrical Power: 1400 kW x 10


days x 24 hours @ $0.05/kWh =
$16800
(Power unit rental included in
capital costs)

Maintenance

None

10% of Interest and Depreciation =


$32,000

Capital Cost

None

Interest and Depreciation over 10


years = $324,000

Total

Approximately $1,300,000

Approximately $375,000

These units are also available on a rental basis.

When possible, rely on previous history and recognize the inevitability of statistical variation
(Fear, 1996;53 Surewaard, et al., 199564). For the particular situation evaluated by Surewaard
et al., 1996,36 it was determined that (if a log normal distribution was applied), at least ten
wells needed to be drilled underbalanced to prove with reasonable confidence (80%) that the
cutoff level of improvement (a 50 percent improvement in this case) would be achieved.
Perform basic cost/ft drilling evaluations. A convenient way to do this is to use the
formulation from Moore, 1974:56

CT =

B + C r ( t + T)
F

(4.12)

where:
CT............................................ total cost/foot,
B ........................................................bit cost,
Cr ........................................... hourly rig cost,
t.................................................rotating time,
T ................................... round trip time, and,
F.......................................footage per bit run.

4-50

If systems other than dry air are used, modify this relationship to include the cost of fluids.

Assess Drilling Costs


Carden, 1993,1 provided an example of using formulas such as this for direct evaluation of the
reduction in development costs. In any drilling operation, the development costs include the
drilling costs (footage and day work), the intangible costs (location and roads, coring, logging,
formation testing, fuel, water, drilling fluid, cementing, transportation, perforating, stimulating,
bits, rental equipment ) and the tangible costs (including surface and production casing,
tubing, the Christmas tree and surface connections, supplementary equipment ... ).
Underbalanced drilling can affect all of these areas. The example shown is for determining the
amount of water that can be hauled off each day. Strictly on the basis of development costs,
data from Carden, 1993,1 can be used to develop the comparison shown in Table 4-6.
Break-even water volumes are shown in Figure 4-8.
Accelerated Production
Drilling is an integrated part of the field development and exploitation process. In addition to
development costs, the potential for increased or accelerated production needs to be evaluated. A
criterion for success can be defined; such as a well inflow quality indicator (Bieseman and Emeh,
199565) or a production improvement factor, PIF, (Surewaard et at., 199564). The latter authors
evaluated the Nimr field; underbalanced operations were not considered to have an effect on the
ultimate recovery. The economic benefit in the Nimr field was considered to be earlier
(accelerated) oil production. Well pro-duction profiles were prepared for a range of PIF values,
but subject to a gross production ceiling, limited by the lift capacity of the currently used beam
pumps. Break-even well costs were then determined as a function of PIF, by evaluating the Net
Present Value (NPV) as an incremental oil acceleration project. Biesman and Emeh, 1996,65 also
discussed the impact of earlier production on the net present value (NPV).
Table 4-6.

Comparison Between Air and Mud Drilling

Item

Air Drilling

Mud Drilling

Interval

From 4000 to 7000 ft

From 4000 to 7000 feet

Interval Length (F) (ft)

3000

3000

Penetration Rate (ft/hr)

30

15

Rotating Time (t) (hr)

100

200

Bit Life (hr)

100

100

Bits Required

Unit Bit Cost

$4800/bit

$4800/bit

Bit Costs (B)

$4800

$9600

Trip Schedule

Trip in to 4000 ft

Trip in to 4000 ft

(unlike direct lifting expenses, these expenses are often capitalized in the first year of production),

4-51

Chapter 4

Selecting An Appropriate Technique

Item

Air Drilling

Mud Drilling

Trip out from 7000 ft

Trip out from 5500 ft


Trip in to 5500 ft
Trip out from 7000 ft

Total Trip Footage

11000 ft

22000 ft

Unit Trip Time


(hr/1000 ft)

1.5

1.5

Trip Time (T) (hr)

16.5

33

Hourly Operating Cost


(Cr)

$375/hr

$250/hr

Cost / ft =

4-52

B + C r (T + t )

9600 + 250(33 + 200)

3000
4800 + C r (16.5 + 100)

Competitive Cost for Air


Drilling

$22.62 =

Barrels of Water That


Can be Disposed of at
$1.00/bbl

= ($541.29-$375)/$1.00
= 166 x 24 = 3984 BWPD

Barrels of Water That


Can be Disposed of at
$5.00/bbl

= ($541.29-$375)/$5.00
= 33 x 24 = 798 BWPD

Barrels of Water That


Can be Disposed of at
$10.00/bbl

= ($541.29-$375)/$10.00
= 16.6 x 24 = 400 BWPD

3000
C r = $541.29 / hr

Including fluid, for illustrative purposes.

= $22.62 / ft

25
Mud Drilling

24
23
Intersection Indicates The
Economic Limit of Air Drilling

Cost ($/ft)

22
21
20

$1.00/BW
$2.50/BW
$5.00/BW
$7.50/BW
$10.00/BW

19
18
17
16
15
0

500

1000

1500

2000

2500

3000

Barrels of Produced Water Per Day

Figure 4-8.

NPV =

Economic water volume production (modified after Carden, 1993 1

1
t
t = (1 + DR )
(1 + DR )

(4.13)

where:
NPV ... Net Present Value (discounted value
of asset),
DR ..... Discount rate (%), and,
t .......... Discount time (years).

4-53

Chapter 4

Selecting An Appropriate Technique

Improved Production/Reserves
In addition to accelerated production, production at any time can be improved by reduction in
skin. To make this assumption, simple forecasts of production are required.

The absolute and relative increase in production should be calculated, or at least estimated.
For estimation, the productivity index (PI) for a vertical oil well is:
0.00708kh
(4.14)
PI =
r
1
Bo ln e + s
rw 2

The productivity index for an horizontal oil well is:

PI =

0.00708kL
2

1 + 1 L
L

2 re
h
+ ln
Bo ln
+ s
L
2 rw
h

2 re

where:
k......... reservoir permeability (md),
h......... reservoir thickness (feet),
......... oil viscosity (cP),
Bo ....... formation volume factor (bbl/sbbl),
re ........ external radius (of reservoir) (feet),
rw........ wellbore radius (feet),
L ........ length of horizontal reservoir
section (feet), and,
s ......... skin (dimensionless).


q=

For a vertical well, if the reservoir is considered to be radial, prior to pseudo-steady-state


conditions:
kh (pi p wf )

162.6Bo

log

4-54

kt
2 3.23 + 0.87 s ( oil )
c t rw

From Biesman and Emeh, 1996.

65

(4.15)

kh (pi2 p 2wf )
q=

1637ZT
1

kt
log
2 3.23 + 0.87 s ( gas)
c t rw

where:
q......... rate (BOPD, Mscf/D),
p i ....... average reservoir pressure (psi),
pwf ...... wellbore pressure (psi),
Z ........ real gas deviation factor
(dimensionless),
T ........ temperature, (R),
t.......... time (hr), and,
ct ........ total compressibility (psi-1).


q=

q=

At pseudo-steady state, for a radially flowing, vertical well:


0.00708kh (p i p wf )
r

Bo ln e 0.75 + s
rw

kh(p p
2
i

2
wf

(oil)

1424ZTln e 0.75
rw

(4.16)
(gas)

The Well Inflow Quality Indicator (WIQI) is the ratio of the PI for an impaired to that for an
undamaged well. PTA (pressure transient analysis) is preferable for determining skin. It can
be difficult and costly. Biesman and Emeh, 1995,65 argued that core flooding offered another
possibility.
Consider the following example for evaluating PI:

k.......................................................... 50 md,
h......................................................... 25 feet,
............................................................. 2 cP,
Bo .................................................. 1 bbl/sbbl,
re .................................................... 1980 feet,
rw......................................................... 0.411,
s ....................................................... variable,
orientation......................................... vertical,
4-55

Chapter 4

Selecting An Appropriate Technique

depth ........................................... 10,000 feet,


reservoir pressure ....................4330 psi, and,
BHPP............3000 psi (pseudo-steady state).
The relevant calculations are shown in Table 4-7 and Figure 4-9.
Simple analyses such as these can qualitatively show how production rate can be increased if
underbalanced drilling reduces skin. They may show that fewer wells are required and that the
producible oil or gas in place can be increased.
Abandonment pressure might be also reduced if the skin is reduced by drilling underbalanced.
This is because of the pressure drop through the skin. Consider the additional pressure drop due
to skin in an oil well, for radial steady-state flow.
Table 4-7.

4-56

Variation of PI and WIQI with Skin


Skin

Production Rate
(BOPD)

PI

WIQI

761

0.572

674

0.507

0.89

604

0.455

0.79

462

0.348

0.61

10

331

0.249

0.44

100

55

0.041

0.07

1.12
Production Rate
PI
WIQI

Production Rate (BOPD)

700

0.98

600

0.84

500

0.7

400

0.56

300

0.42

200

0.28

100

0.14

Well Inflow Quality Indicator


Productivity Index

800

0
0

10

100

Skin

Figure 4-9.

Ps =

Variation of PI, production rate and WIQI with drilling induced skin.

1412
. QBo
s
kh

(4.17)

where:
Ps
pressure drop due to skin, (psi),
Q ........ flow rate, (BOPD),
Bo ....... formation volume factor (bbl/sbbl),
......... viscosity, (cP),
k......... permeability (md),
h......... thickness (ft), and,
s ......... skin factor (dimensionless).

Consider damage only; not due to mechanical skin or partial penetration.

4-57

Chapter 4

Selecting An Appropriate Technique

This effect can be significant and should not be ignored in economic assessments.
An Example
One final example illustrates the impact of underbalanced drilling in one hypothetical five well
field. The analysis is very simplified and illustrative only. Consider the following scenario. It
was developed for an oil well. Similar analyses could be done for gas.
Revenue Interest ............................ R = 0.375
Working Interest..........................WI = 0.500
Gross Income (per net bbl)
Crude Price................................... $20.00/bbl
Gas Revenue...................................... $0.00
Total ............................................. $20.00/bbl
Less
Transportation ................................ $1.00/bbl
Production Taxes............................ $6.00/bbl
Leaves
Gross Income (per net bbl)................. $13.00
Estimated Operating Expenses
(per well month) .................................. $5000
Number of Wells ......................................... 5
Three cases were examined. The only differences were in the development costs and in
accelerated production.
Case 1: This is the base case. All five wells were drilled in the first year with a conventional
mud system. The operator has a 0.375 revenue interest and a 0.500 working interest. A cash
flow projection is shown in Table 4-8 and Figure 4-10. The appraisal value is equal to a fraction
of the present worth of the net cash flow, before federal taxes, computed at a safe rate of interest
(5%).
Case 2: This is the same as Case 1, with the exception that there is higher production due to
reduced formation damage from underbalanced drilling. Table 4-9 and Figure 4-10 show the
projections.
Case 3: This is the same as Case 2, with the exception that development costs for the five wells
are $150,000 less, due to improved drilling while underbalanced. The forecasts are shown in
Table 4-10 and Figure 4-10. The operator can assess both drilling costs and estimate the overall
economics on net present value. Doing this, a final decision on the drilling fluid system can be
made.

4-58

Similar calculations can be done for a gas producer.

Table 4-8.

Case 1 (Base Case)

Year

Total

Estimated Future

Operation

Units

(1) Gross Lease


Production

bbl

201,204 170,280 122,952 96,720 77,960 55,388 18,024 742,528

bbl

75,452

(2) Net Production R x (1)


to Operator

63,855

46,107

36,270 29,235 20,771 6,759

278,448

(3) Gross Income to (2) x $13.00 $


Operator

980,870 830,115 599,391 471,510 380,055 270,017 87,867 3,619,824

(4) Development
Costs

750,000 0

750,000

(5) Number of
Producing Well
Months

60

48

48

36

36

24

312

(6) Operating
Expense

(5) x $5000

300,000 300,000 240,000 240,000 180,000 180,000 120,000 1,560,000

(7) Capital
Expenditure

20,000

(8) Share of
Operating and
Capital
Expenses

WI x
$
[(4)+(6)+(7)]

535,000 160,000 130,000 130,000 100,000 100,000 70,000 1,225,000

(9) Cash Flow to


Operator

(3) - (8)

445,870 670,115 469,391 341,510 280,055 170,017 17,867 2,394,824

(10) 5% Annual
Deferment
Factor

0.9740

434,277 621,599 414,707 287,347 224,408 129,757 12,986 2,157,736

(11) Present Worth (10) x (9)


of Cash Flow

D CR

20,000

0.9276

20,000

0.8835

20,000 20,000 20,000 20,000 140,000

0.8414 0.8013 0.7632 0.7268 0.9010

(1 + i) (1 + i)
=
12[(1 + i) 1]
1 t

60

1/12

where:
DCR ...................annual deferment factors, applicable to equal payments at the end of each month
during a specific interval of one year between (t - 1) and t years from now,
i.........................................effective annual compound safe interest rate as a decimal fraction, and,
t.....................................................................................................................................time in years.

4-59

Chapter 4

Selecting An Appropriate Technique

Table 4-9.

Case 2

Year

Total

Estimated Future

Operation

Units

(1) Gross Lease


Production

bbl

221,324

187,308 135,247 106,392 85,756 60,927 19,826 816,781

ddl

82,997

70,241

(2) Net Production R x (1)


to Operator

50,718

39,897 32,159 22,848 7,435

306,293

(3) Gross Income to (2) x $13.00 $


Operator

1,078,956 913,127 659,330 518,661 418,061 297,018 96,654 3,981,806

(4) Development
Costs

750,000

750,000

48

48

36

36

24

312

(5) Number of
Producing Well
Months

60

60

(6)Operating
Expense

(5) x $5000

300,000

300,000 240,000 240,000 180,000 180,000 120,000 1,560,000

(7) Capital
Expenditure

20,000

20,000

(8) Share of
Operating and
Capital Expenses

WI x
$
[(4)+(6)+(7)]

535,000

160,000 130,000 130,000 100,000 100,000 70,000 1,225,000

(9) Cash Flow to


Operator

(3) - (8)

543,956

753,127 529,330 388,661 318,061 197,018 26,654 2,756,806

(10) 5% Annual
Deferment Factor

0.9740

0.9276

529,814

698,600 467,663 327,019 254,862 150,364 19,372 2,483,883

(11) Present Worth (9) x (8)


of Cash Flow

D CR =

0.8835

20,000 20,000 20,000 20,000 140,000

0.8414 0.8013 0.7632 0.7268 0.9010

(1 + i) (1 + i)
12[(1 + i) 1]
1 t

20,000

1/12

where:
DCR ...................annual deferment factors, applicable to equal payments at the end of each month
during a specific interval of one year between (t - 1) and t years from now,
i.........................................effective annual compound safe interest rate as a decimal fraction, and,
t.....................................................................................................................................time in years.

4-60

Table 4-10.

Case 3

Year

Total

Estimated Future

Operation

Units

(1) Gross Lease


Production

bbl

221,324

187,308 135,247 106,392 85,756 60,927 19,826 816,781

ddl

82,997

70,241

(2) Net Production R x (1)


to Operator

50,718

39,897 32,159 22,848 7,435

306,293

(3) Gross Income to (2) x $13.00 $


Operator

1,078,956 913,127 659,330 518,661 418,061 297,018 96,654 3,981,806

(4) Development
Costs

600,000

600,000

48

48

36

36

24

312

(5) Number of
Producing Well
Months

60

60

(6)Operating
Expense

(5) x $5000

300,000

300,000 240,000 240,000 180,000 180,000 120,000 1,560,000

(7) Capital
Expenditure

20,000

20,000

(8) Share of
Operating and
Capital Expenses

WI x
$
[(4)+(6)+(7)]

460,000

160,000 130,000 130,000 100,000 100,000 70,000 1,150,000

(9) Cash Flow to


Operator

(3) - (8)

618,956

753,127 529,330 388,661 318,061 197,018 26,654 2,831,806

(10) 5% Annual
Deferment Factor

0.9740

0.9276

602,864

698,600 467,663 327,019 254,862 150,364 19,372 2,551,458

(11) Present Worth (9) x (8)


of Cash Flow

D CR

0.8835

20,000 20,000 20,000 20,000 140,000

0.8414 0.8013 0.7632 0.7268 0.9010

(1 + i) (1 + i)
=
12[(1 + i) 1]
1 t

20,000

1/12

where:
DCR ...................annual deferment factors, applicable to equal payments at the end of each month
during a specific interval of one year between (t - 1) and t years from now,
i.........................................effective annual compound safe interest rate as a decimal fraction, and,
t.....................................................................................................................................time in years.

4-61

Chapter 4

Selecting An Appropriate Technique

700,000
Conventional (Total $2,157,736)

Present Worth of Cash Flow ($)

600,000

Underbalanced (10% Increase in Production) (Total $2,483,883)


Underbalanced (Add. 10% Reduction In Drilling Costs) (Total
$2,551,458)

500,000

400,000

300,000

200,000

100,000

0
1

Year

Figure 4-10.

Table 4-11.

Projections Over Seven Years

Summary of all Cases (Present Worth of Cash)

Year
Case

Total

434,277

621,599

414,707

287,347

224,408

129,757

12,986

2,157,736

529,814

698,600

467,663

327,019

254,862

150,364

19,372

2,483,883

602,864

698,600

467,663

327,019

254,862

150,364

19,372

2,551,458

Return to SC Manual

4-62

References
1.

Carden, R.S.: Technology Assessment of Vertical and Horizontal Air Drilling Potential
in the United States, U.S. DOE Report No. DOE/MC/28252-3514 (DE94000044)
(August 1993).

2.

Anderson, G.W.: Near-Gauge Holes Through Permafrost, Oil and Gas J. (September 20,
1971) 132-142.

3.

Fraser, I.M. and Moore, R.H.: Guidelines for Stable Foam Drilling Through Permafrost,
paper SPE/IADC 16055 presented at the 1987 SPE/IADC Drilling Conference, New
Orleans, LA.

4.

Allen, T.O. and Roberts, A.P.: Production Operations - Well Completions, Workover and
Stimulation, Volume 2, OGCI, Tulsa, OK (1982).

5.

Bennion, D.B., Thomas, F.B., Bietz, R.F. and Bennion, D.W.: Underbalanced Drilling,
Praises and Perils, presented at the First International Underbalanced Drilling Conference,
The Hague, The Netherlands, October 2-4, 1995a.

6.

Bennion, D.B.: Underbalanced Operations Offer Pluses and Minuses, Oil & Gas J.
(January 1, 1996) 33-40.

7.

Bennion, D.B., Thomas, F.B., Bietz, R.F. and Bennion, D.W.: Water and Hydrocarbon
Phase Trapping in Porous Media-Diagnoses, Prevention and Treatment, paper presented
at the 1995 Annual Technical Meeting of The Petroleum Society of CIM, Banff, Alta, May
14-17, 1995b.

8.

Horsrud, P., Holt, R.M., Sonstebo, E.F., Svano, G. and Bostrom, B.: Time-Dependent
Borehole Stability: Laboratory Studies and Numerical Simulation of Different Mechanisms
in Shale, paper presented at the 1994 SPE/ISRM Rock Mechanics in Petroleum
Engineering, Delft, The Netherlands (August 29-31) Eurock 94, Balkema, Rotterdam,
(1994) 259-266.

9.

Bol, G.M., Wong, S-W., Davidson, C.J. and Woodland, D.C.: Borehole Stability in
Shales, SPEDC (June 1994) 87-94.

10.

Tan, C.P. and Rahman, S.S.: The Mechanism of Mud Support Reduction Due to Mud
Pressure Penetration, paper presented at the 1994 SPE/ISRM Rock Mechanics in
Petroleum Engineering, Delft, The Netherlands (August 29-31) Eurock 94, Balkema,
Rotterdam (1994) 285-292.

11.

Hale, A.H., Mody, F.K. and Salisbury, D.P.: The Influence of Chemical Potential on
Wellbore Stability, paper IADC/SPE 23885 presented at the 1992 IDAC/SPE Drilling
Conference, New Orleans, LA, February 18-21.
4-63

Chapter 4

4-64

Selecting An Appropriate Technique

12.

Kunze, K.R. and Steiger, R.P.: Extended Leakoff Tests to Measure In Situ Stress During
Drilling, paper presented at the 1991 U.S. Symposium on Rock Mechanics, Norman, OK
(July 10-12) Rock Mechanics as a Multidisciplinary Science; Proceedings of the 32nd U.S.
Symposium, J-C. Roegiers (ed.), Norman, OK (1991) 35-44.

13.

Lowrey, J.P. and Ottesen, S.: An Assessment of the Mechanical Stability of Wells
Offshore Nigeria, paper SPE 26351 presented at the 1993 Annual Technical Conference
and Exhibition, Houston, TX, October 3-6.

14.

Steiger, R.P. and P.K. Leung: Critical State Shale Mechanics, paper presented at the 1991
U.S. Symposium on Rock Mechanics, Norman, OK (July 10-12) Rock Mechanics as a
Multidisciplinary Science; Proceedings of the 32nd U.S. Symposium, J-C. Roegiers (ed.),
Norman, OK (1991) 293-302.

15.

Risnes, R., Bratli, R.K. and Horsrud, P.: Sand Stresses Around A Wellbore, SPEJ
(1982) 22, 883-898.

16.

Bratli, R.K. and Risnes, R.: Stability and Failure of Sand Arches, SPEJ (April 1981)
236-248.

17.

Wang, Y. and Dusseault, M.B.: Borehole Yield and Hydraulic Fracture Initiation in
Poorly Consolidated Strata - Part I. Impermeable Media, Int. J. Rock Mech. Min. Sci. &
Geomech. Abstr. (1991a) 28, No. 4, 235-246.

18.

Wang, Y. and Dusseault, M.B.: Borehole Yield and Hydraulic Fracture Initiation in
Poorly Consolidated Strata - Part II. Permeable Media, Int. J. Rock Mech. Min. Sci. &
Geomech. Abstr. (1991b) 28, No. 4, 247-260.

19.

Bradley, W.B.: Failure of Inclined Boreholes, J. Energy Resources Tech., Trans., ASME
(1979) 101, 232-239.

20.

Zoback, M.D., Moos, D., Hastin, L. and Anderson, R.N.: Wellbore Breakouts and In Situ
Stress, J. Geophys. Res. (1985) 90, 5523-5530.

21.

Ramos, G.G., Katahara, K.W., Gray, J.D. and Knox, D.J.W.: Sand Production in Vertical
and Horizontal Wells in a Friable Sandstone Formation, North Sea, paper presented at the
1994 SPE/ISRM Rock Mechanics in Petroleum Engineering, Delft, The Netherlands
(August 29-31) Eurock 94, Balkema, Rotterdam, (1994) 309-315.

22.

Wang, Z., Peden, J. and Damasena, E.: The Prediction of Operating Conditions to
Constrain Sand Production From A Gas Well, paper SPE 21681 presented at the 1991
SPE Production Operations Symposium, Oklahoma City, OK.

23.

Detournay, E. and Cheng, A. H-D.: Poroelastic Response of a Borehole in a NonHydrostatic Stress Field, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. (1988) 25, 171182.

24.

Brudy, M. and Zoback, M.D.: Compressive and Tensile Failure of Boreholes Arbitrarily
Inclined to Principal Stress Axes: Application to the KTB Boreholes, Germany, paper
presented at the 1993 U.S. Symposium on Rock Mechanics, University of Wisconsin (June
27-30) Preprint Proc. (1993) 69-72.

25.

McLennan, J.D., Roegiers, J-C. and Economides, M.J.: Extended Reach and Horizontal
Wells, Reservoir Stimulation, M.J. Economides and K.G. Nolte (eds.), Prentice-Hall
Publisher, N.J. (1989) Vol. 2, 19.1-19.28.

26.

Santarelli, F.J., Brown, E.T. and Maury, V.: Analysis of Borehole Stresses Using
Pressure-Dependent, Linear Elasticity, Int. J. Rock Mech. Min. Sci. and Geomech. Abstr.
(1986) 25, 159-170.

27.

Pellegrino, A., Sulem, J. and Barla, G.: Nonlinear Effects in the Study of Borehole
Stability, paper presented at the 1994 SPE/ISRM Rock Mechanics in Petroleum
Engineering, Delft, The Netherlands (August 29-31) Eurock 94, Balkema, Rotterdam,
(1994) 231-237.

28.

Nawrocki, P. and Dusseault, M.: Addressing the Effects of Material Non-Linearities on


Wellbore Stresses Using Stress- and Strain-Dependent Elastic Moduli, Rock Mechanics,
J.J.K. Daemen and R.A. Schultz (eds.), Balkema, Rotterdam (1995) 819-824.

29.

Vaziri, H. and Byrne, P.M.: Analysis of Stress, Flow and Stability Around Wellbores,
Geotechnique (1990) 40, No. 1, 63-77.

30.

Wang, X.: Numerical Modeling of Wellbore Cavitation and its Influence on Fluid
Production in Coalbed Methane Reservoirs, Ph.D. Thesis, Technical University of Nova
Scotia, Canada (1995).

31.

Charlez, P.A.: The Impact of Constitutive Laws on Wellbore Stability: A General


Review, paper presented at the 1994 SPE/ISRM Rock Mechanics in Petroleum
Engineering, Delft, The Netherlands (August 29-31) Eurock 94, Balkema, Rotterdam,
(1994) 239-249.

32.

Ewy, R.T.: 3D Stress Effects in Elastoplastic Wellbore Failure Models, paper presented
at the 1991 U.S. Symposium on Rock Mechanics, Norman, OK (July 10-12) Rock
Mechanics as a Multidisciplinary Science; Proceedings of the 32nd U.S. Symposium, J-C.
Roegiers (ed.), Norman, OK (1991) 951-960.

4-65

Chapter 4

4-66

Selecting An Appropriate Technique

33.

Detournay, E.: Elastoplastic Model of a Deep Tunnel for a Rock With Variable
Dilatancy, Rock Mech, Rock Eng. (1986) 19, 99-108.

34.

Veeken, C., Walters, J.V., Kenter, C.J. and Davies, D.: Use of Plasticity Models for
Predicting Borehole Stability, paper presented at the 1989 ISRM-SPE/International
Symposium, Pau, France, Rock at Great Depth, Balkema, Rotterdam (1989) 2, 835-844.

35.

Wong, S.W. and Heidug, W.K.: Borehole Stability in Shales: A Constitutive Model for
Mechanical and Chemical Effects of Drilling Fluid Invasion, paper presented at the 1994
SPE/ISRM Rock Mechanics in Petroleum Engineering, Delft, The Netherlands (August 2931) Eurock 94, Balkema, Rotterdam (1994) 251-257.

36.

Surewaard, J., de Koning, K., Kool, M., Woodland, D., Roed, H. and Hopmans, P:
Approach to Underbalanced Well Operations in Petroleum Development Oman, paper
IADC/SPE 35069 presented at the 1996 IADC/SPE Drilling Conference, New Orleans, LA.

37.

Shale, L.: Underbalanced Drilling Equipment and Techniques, paper presented at the 1995
ASME Energy Technology Conference, Houston, TX, January 30 - February 1.

38.

Graves, S.L., Niederhofer, J.D. and Beavers, W.M.: A Combination Air and Fluid Drilling
Technique for Zones of Lost Circulation in the Black Warrior Basin, SPEDE (February
1986) 57-61.

39.

Deis, P.V., Yurkiw, F.J. and Barrenechea, P.J.: The Development of an Underbalanced
Drilling Process: An Operators Experience in Western Canada, paper presented at the 1995
1st International Underbalanced Drilling Conference, The Hague, The Netherlands, October
2-4.

40.

Dupont, J.: Foam Used to Drill, Gravel-Pack Deep Gas Well, Oil and Gas J. (May 7, 1984)
192-194.

41.

Santarelli, F.J., Dardeau, C. and Zurdo, C.: Drilling through Highly Fractured Formations: A
Problem, a Model and a Cure, paper SPE 24592, presented at the 1992 SPE Annual
Technical Conference and Exhibition, Washington DC.

42.

Nesa, D.O., Larsen V., and Birkeland, R.: Underbalanced Drilling Offshore, paper
presented at the 1995 1st International Underbalanced Drilling Conference and
Exhibition, The Hague, The Netherlands, October 2-4.

43.

Anderson, G.W.: Use of Preformed Stable Foam in Low Pressure Reservoir Wells, 5th
Offshore South East Asia Meeting, Singapore, February 21-24, 1984.

44.

Bentsen, N.W. and Veny, J.N.: Preformed Stable Foam Performance in Drilling and
Evaluating Shallow Gas Wells in Alberta, paper SPE 5712, JPT (October, 1976) 12371240.

45.

Comeau, L.: Underbalanced Drilling: Directional and MWD Experience, paper


presented at the 1995 1st International Underbalanced Drilling Conference and Exhibition,
The Hague, The Netherlands, October 2-4.

46.

Crearar, P.: Underbalanced Re-Entry Horizontal Drilling in the Welton Field Basal
Succession Reservoir Onshore UK, paper presented at the 1995 1st International
Underbalanced Drilling Conference and Exhibition, The Hague, The Netherlands, October
2-4.

47.

Cummings, S.G.: Natural Gas Drilling Methods and Practice: San Juan Basin, New
Mexico, paper SPE/IADC 16167 presented at the 1987 SPE/IADC Drilling Conference,
New Orleans, LA, March 15-18.

48.

Mullane, T.J., Churcher, P.L., Edmunds, A.C., Eddy, D.B., Martin, B.G. and Flach, P.D.:
Benefits of Underbalanced Drilling: Examples from the Weyburn and Westerose Fields,
Western Canada, paper presented at the 1995 1st International Underbalanced Drilling
Conference, The Hague, The Netherlands, October 2-4.

49.

Russell, B.A.: How Surface Hole Drilling Performance Was Improved 65%, paper
IADC/SPE 25766 presented at the 1993 SPE/IADC Drilling Conference, Amsterdam,
February 23-25.

50.

Saponja, J.: Comparing Conventional Mud to Underbalanced Drilling in a Depleted


Reservoir, paper presented at the 1995 1st International Underbalanced Drilling Conference,
The Hague, The Netherlands, October 2-4.

51.

Whiteley, M.C. and England, W.P.: Air Drilling Operations Improved by PercussionBit/Hammer-Tool Tandem, SPEDE (October 1986) 377-382.

52.

Westermark, R.V.: Drilling With a Parasite Aerating String in the Disturbed Belt,
Gallatin County, Montana, paper IADC/SPE 14734 presented at the 1986 IADC/SPE
Drilling Conference, Dallas, TX, February 10-12.

53.

Fear, M.J.; How to Improve Rate of Penetration in Field Operations, paper IADC/SPE
35107 presented at the 1996 IADC/SPE Drilling Conference, New Orleans, LA.

54.

Bourgoyne, A.T. Jr.: Rotating Control Head Applications Increasing, Oil and Gas J.
(October 9, 1995).

4-67

Chapter 4

4-68

Selecting An Appropriate Technique

55.

Reinsvold, C.H., Clement, J., Oliver, M., Witt, C. and Crockett, J.: Diamond-Enhanced
Hammer Bits Reduce Cost per Foot in the Arkoma and Appalachian Basins, paper
IADC/SPE 17185 presented at the 1988 IACD/SPE Drilling Conference, Dallas, TX,
February 28 - March 2.

56.

Moore, P.L.:
(1974).

57.

Grace, R.D.: Pressure Control in Balanced and Underbalanced Drilling in the Anadarko
Basin, paper SPE 5396 presented at the 1975 SPE Regional Meeting, Oklahoma City, OK
March 24-25.

58.

Bloys, B., Brown, J.D. and Tarr, B.A.: Drilling Safely and Economically in Carbonates:
Collective Experience of ARCO, BP and Mobil, paper presented at the 1994 IADC Well
Control Conference for the Asia/Pacific Region, Singapore, December 1-2.

59.

Astrella, L. and Wiemers, R.: Closed Loop Drilling Systems Can Eliminate Reserve Pit
Costs, Oil & Gas J. (May 27, 1996) 62-68.

60.

Allan, P.D.: Nitrogen Drilling System for Gas Drilling Applications, paper SPE 28320
presented at the 1994 Annual Technical Conference and Exhibition, New Orleans, LA.

61.

Tag, A.: Portable Prism Nitrogen (PPN): A New Concept to Replace Liquid Nitrogen,
paper presented at the 1995 1st International Underbalanced Drilling Conference and
Exhibition, The Hague, The Netherlands, October 2-4.

62.

Scott, S.L., Wu, Y. and Bridges, T.J: Air Foam Improves Efficiency of Completion and
Workover Operations in Low-Pressure Gas Wells, SPEDC (December 1995) 219-225.

63.

Fried, S. and MacDonald, C.: Nitrogen Supply Alternatives for Underbalanced Drilling,
paper presented at the 1995 1st International Underbalanced Drilling Conference and
Exhibition, The Hague, The Netherlands, October 2-4.

64.

Surewaard, J., de Koning, K., Kool, M., Woodland, D., Roed, H. and Hopmans, P.:
Underbalanced Operations in Petroleum Development Oman, paper presented at the
1995 1st International Underbalanced Drilling Conference and Exhibition, The Hague, The
Netherlands, October 2-4.

65.

Bieseman, T. and Emeh, V.: An Introduction to Underbalanced Drilling, paper presented


at the 1995 1st International Underbalanced Drilling Conference and Exhibition, The
Hague, The Netherlands, October 2-4.

66.

Allen, T.O. and Roberts, A.P.: Production Operations - Well Completions, Workover and
Stimulation, Volume 2, OGCI, Tulsa, OK, 1982.

Drilling Practices Manual, PennWell Publishing Company, Tulsa, OK

5.1

WELL ENGINEERING

Circulation Programs

Design of circulation programs for underbalanced drilling is fundamentally no different than for
balanced or underbalanced situations. The basis for hydraulics design is to guarantee adequate
hole cleaning, to ensure vertical transport of cuttings in annular zones where velocities are
reduced because of changes in annular area, to maintain wellbore stability and mitigate formation
damage and to operate within the pressure and rate constraints of the tubulars and the surface
equipment. Mathematically, most wellbore drilling hydraulics programs are based on continuity
of mass and solution of mechanical energy balance relationships (refer to Appendix D).
The total pressure drop over an interval is equal to the change in pressure due to hydrostatic head
plus the acceleration pressure drops (kinetic energy) for fluid and solid phases, almost always
negligibly small, plus the pressure drop due to frictional interaction with the drillpipe, casing and
openhole. The hydrostatic component includes the fluid(s) as well as the solids in the annulus.
As indicated, the kinetic energy effects of the solids are usually ignored, although they may play
some role in the region immediately above the drill collars where velocity is reduced, if extreme
overbalance is encountered. Kinetic effects for the fluids are usually only considered when
accounting for pressure drop through the bit nozzles.
Guo et al., 1993,1 showed relationships for pressure loss through a bit for an aerated mud.
Neglecting elevation change and energy loss due to friction, the energy balance equation through
a bit can be expressed as:
2
1 M 1
1
P1 P2 =

g c A F 2 F1

(5.1)

where:
P1 ....... upstream pressure (psf),
P2 ....... downstream pressure (psf),
gc ........ conversion factor
(32.17 lbmft/lbfs2),
M ....... mass flow rate (lbm/s),

5-1

Chapter 5

Well Engineering

A ........ nozzle area (ft2),


F2 ...... downstream density (lbm/ft3), and,
F1 ...... upstream density (lbm/ft3).
k

2 k 1
For air drilling, sonic Pa = Pb

k + 1
k +1
Ta R
1 k 1
k +
Sgk
k

Pb Pa = 0.471Pa

G
Pa =
An

(5.2)

2 k 1
For air drilling, sub - sonic Pa < Pb

k + 1
R ( k 1)G Tb
Pa = Pb 1 +

2gkSA 2n Pb2

k
k 1

(5.2 cont.)

where:
G ........ mass flow rate of air lbm/s,
An....... area of the bit nozzles (in2),
Ta ....... air temperature above the bit (R),
Tb ....... temperature beneath the bit (R).
R ........ the universal gas constant
(53.3 ftlbf/lbmR for air),
k......... ratio of the specific heat at
constant pressure to that at
constant volume (dimensionless),
S......... gas gravity (1 for air), and,
g......... gravitational constant (32.17 ft/s2).
Normally, in well executed drilling operations, the pressure loss through the bit will be
approximately fifty percent of the surface pressure. The flow regime through surface connections
and inside the drillstring is generally turbulent, accounting for between sixty and eighty percent
of the remaining component of the surface pressure. The flow regime in the annulus is generally
laminar for liquids, accounting for twenty to forty percent of circulating pressure losses
(excluding those through the bit). This is not necessarily the case for gas.

5-2

Turbulent Flow
Particularly in air and mist drilling, the majority of pressure losses in the circulating system are
associated with turbulent flow. These include the pressure losses in surface connections, drillpipe
and drill collars. In turbulent flow, the local velocity vectors are not ordered and the velocity
profile is flat, with the average velocity approximately equal to the maximum velocity (as
compared to parabolic distributions in laminar flow).
In frictional pressure relationships, the pressure drop is commonly incorporated by using the
Fanning Friction Factor, f. While there is a considerable uncertainty in values for f, approximate
relationships can be developed from relationships shown in Figure 5-1. Relative roughness, from
Moodys diagram, Figure 5-2, can be incorporated using Nikuradses formulations (refer to
Chapter 2).
Reynolds number is defined as:
R e = 15.47

Dv m

(5.3)

where:
Re ....... Reynolds Number (dimensionless),
D ........ equivalent hydraulic diameter (ft),
v ........ average cross-sectional velocity
(ft/s),
m....... mixture density (lbm/ft3), and,
......... mixture viscosity, cP.
In turbulent flow, viscosity has an unclear meaning. It may be represented as:
=

PV
3.2

(5.4)

where:
PV...... plastic viscosity (lbf/100 ft2).
The frictional pressure drop is then given by:
P =

m v 2 Lf
9.298 x 104 D

(5.5)

5-3

0.1
Laminar

Complete Turbulence (rough pipes)

Friction Factor, f

/ D = 0.05

/ D = 0.01
/ D = 0.005

64/Re
/ D = 0.0001

/ D = 0.0005
/ D = 0.0001

Smooth Pipe
/ D = 0.00005

0.01
100

1000

10000

100000

1000000

10000000

Reynolds Number, Re

Figure 5-1.

Friction Factor, f, as a function of Reynolds number.

1.00E+00

Relative Roughness, /D (dimensionless)

Asphalted Cast Iron


Galvinized Iron

1.00E-01
Cast Iron
Concrete

1.00E-02

1.00E-03

1.00E-04

1.00E-05

Drawn Tubing
Well Tubing

1.00E-06

Line Pipe
Commercial Steel or Wrought Iron

1.00E-07
1

10

100

Pipe Diameter, D (inches)

Figure 5-2.

Moody correlation for relative roughness.

5-5

Chapter 5

Well Engineering

where:
P ...... pressure drop (psi/ft), and,
L ...... length interval (feet).
Other pressure and velocity conditions are covered in Chapter 2. The following section shows
example calculations using some of the available analytical routines for predicting fluid (gas
and/or liquid requirements), specifically for air drilling. Depending on the particular project,
commercial wellbore hydraulics simulators should be used interactively with available or realtime drilling data.

5.2

Circulation Calculations (Air, Gas, Mist)

Dry Air Drilling


Order of magnitude calculations of required air rates can be determined using Angels analysis
techniques, as described in Chapter 2. Section 2.1. The problem can be approached in two ways.
The fundamental relationships can be specifically solved or approximations published by Angel
(Appendix C) can be used. Regardless of which analysis is used, the rates are commonly
underestimated. Some operators take Angels predictions and order out equipment to handle
twice these rates; further adjusting on the fly. Some improvements may be possible by
manipulating Angels parameter, b, to incorporate alternate friction factors.
Angels Approximate Method
 Collect the required information for the calculations. This includes:

Drilled hole size (inches),

OD of the drillpipe (inches),

Drilling rate (ft/hr), and,

Depth (thousands of feet).

 In the table in Appendix C, determine Qo and N. Interpolate values as required.


 Calculate the required circulation rate using:
Q = Q o + NH
where:
Qo, N.. parameters from Appendix C,
H ........ depth in thousands of feet, and,
Q ........ circulation rate (scfm).

5-6

(5.6)

Example 1
Using Angels table, determine the required circulation rate to air drill an 11-inch hole with 5inch drillpipe, at 90 ft/hr, at 11,000 ft.
 From the table in Appendix C:
Qo....... 1456,
N ........ 135, and,
H ........ 11.
Q = Q o + NH = 1456 + 135 11 = 2941 scfm
(5.7)
Solving Angels Equation
A somewhat more precise solution can be derived by programming Angels relationships into a
spreadsheet. The basic equations are:
6.61S (Ts + Gh)Q 2

(D

2
h

(P

D 2p v 2stp

2
s

=
(5.8)

+ bTav2 )e 2 ah / Tav bTav2

where:
S......... gas gravity, air = 1.0
(dimensionless),
Ts ....... surface temperature (F),
G ........ geothermal temperature gradient
(F/ft),
h......... depth below surface to any point
under consideration (feet),
Q ........ circulation rate (scfm),
Dh....... hole diameter (feet),
Dp....... pipe outside diameter (feet),
vstp ...... velocity of air at standard
temperature and pressure (Angel
recommended 3000 ft/min),
Ps........ pressure in the annulus at the
surface (including backpressure)
(psia),
Tav ...... average absolute temperature of the
flow stream (Ts+(Ts+Gh))/2, (R),

5-7

Chapter 5

Well Engineering

(to convert from F to R, add


459.67 to the temperature in F),
SQ + 28.8KD 2h
,
53.3Q

a .........

K ........ drilling rate (ft/hr), and,


10 6 Q 2
1625
.

b.........

( D h D p )1.333 D 2h D 2h

The only real difficulty in solving for this is the fact that Q appears on both sides of the equation.
Solve this by iteration.
Revising Angels Equation
One of the difficulties with Angels analysis is that it uses Weymouths equation for a vertical,
smooth-walled pipe and does not completely represent the contribution of cuttings to friction.
One resolution to this is modifying Angels parameter, b. It was defined as:
b=

(D

1625
.
106 Q 2
h Dp

) (D
1.333

2
h

D 2h )

(5.9)

Recognizing that this incorporates Weymouths friction factor, f = 0.014 (Dh-Dp)-.333, b can be
rewritten as:
b=

1167
.
104 Q 2 f

(D

)(

2
2
h Dp Dh Dp

(5.10)

Alternate values of b can be inferred for different friction factors (refer for, example, to Govier
and Aziz, 19822). For example, f can be varied to account for roughness and adjustments could
be made to include supplemental friction due to cuttings. Note that the simple formulation
presented here is for straight hole.
Simulation
Angels analysis can be modified for any gaseous drilling system. The advantage of Angels
approach is its simplicity. Additional discussion of using this technique is found in Johnson and
Cooper, 1993.3
Nitrogen Drilling
Angels method can also be used for nitrogen. The procedures are similar to those for dry air. As
a first order approximation, it can be assumed that the cuttings transport efficiency is similar to
that for air. The gas gravity, S, for nitrogen can be taken as S = 0.97, compared to S = 1 for air.

5-8

Natural Gas Drilling


Procedures for natural gas drilling circulation assessment, using Angels approach, are similar to
those for air and nitrogen drilling. Typically, the gravity will vary from 0.60 to 0.70 for dry
hydrocarbon gases. If PVT information is not available, a reasonable default value would be
0.65.
Mist Drilling
As indicated in Section 2.4, a simple approximation of the pressure and the required delivery
capacities during mist drilling can be found by using Angels analysis and artificially
representing added or produced liquid as cuttings. For order of magnitude predictions:
 Estimate the total liquid rate flowing up the annulus (BPH). This includes foamer and
produced water.

Determine the liquid added at the surface (BPH).

Determine the liquid inflow (BPH).

Determine the total liquid (BPH) by addition.

 Calculate an apparent, supplementary rate of penetration, to account for the liquids.


ROPe =

380Q L
D2

(5.11)

where:
ROPe .. apparent additional ROP due to water
(ft/hr).
QL ...... total liquid flow rate (BPH), and,
D ........ bit diameter (inches).
 Decide on the desired (actual) penetration rate, ROPd, ft/hr.
 Calculate an artificial penetration rate, ROP , to gas lift the cuttings and liquid.
ROP = ROPd + ROPe

(5.12)

 Use Angels methodology to determine the required air rate, Q, in scfm, and the bottomhole
pressure, Pb, using ROP .
Example 2
Estimate the required air circulation rate to mist drill at 11,000 feet at 90 ft/hr. Pipe diameter is
5-inches and the hole diameter is 11-inches. Foamer is added at the surface at 4 gpm and water
is produced into the annulus at 11,000 feet at 9 BWPH.

The injected liquid rate is 4 gpm, which is 5.71 BPH.

5-9

Chapter 5

Well Engineering

The produced formation fluid rate is 9 BPH.

The total liquid rate is 5.71 + 9 =14.71 BPH.

The apparent penetration rate to account for water is:

ROPe =

380 14.71
= 46.2 ft / hr
112

(5.13)

The desired penetration rate is, ROPd = 90 ft/hr.

The total artificial rate is ROP = 46.2 + 90 = 136.2 ft/hr.

From Appendix D:

Q o = 1456
N = 160 (Extrapolated)
Q = Q o + NH

(5.14)

= 1456 + 160 11
= 3216 scfm

5.3

Circulation Calculations (Gasified Liquids)

Approximate volumes and pressures, for gasified liquids, can be determined using the techniques
described previously. More precise predictions require added levels of sophistication. Some
examples of more refined circulating procedures are provided. These can be adopted for any of
the fluid systems discussed. The examples shown are for flow in the annulus. Drillpipe flow
calculations are similar, although cuttings are not included and the pipes internal dimensions are
used. Additional sophistication can be incorporated to account for pressure drop at the bit, and
changing head due to solids or water influx (refer to example Guo et al., 19931). The circulation
simulations described below indicate methodologies for estimating air requirements to lighten or
aerate muds. Although commercial simulators are readily available for drilling hydraulics, this
discussion is included so that the drilling engineer can understand the concepts of those models
and can do quick parametric comparisons.
Gasifying a liquid is intended to reduce the pressure within the wellbore and to minimize lost
circulation or formation damage. In either case, the pressure in the zone of interest must be
known in order to design the drilling system. The average density of the liquid above the zone of
interest must exert a pressure that is less than (or equal to) the anticipated pressure. The amount
of gas that needs to be injected into the drilling fluid can be estimated in two different ways;
using relationships presented by Poettmann and Bergman, 1955,4 or by doing first principle
calculations to determine volumes, hydrostatic pressures and friction losses in the annulus.

5-10

Poettman-Bergman Equations
A set of charts was developed by Poettmann and Bergman, 1955,4 for average annular
temperatures of 100, 150 and 200F. Figure 5.3 is a reproduction of the 150F chart (all three
charts are included in Chapter 2, Section 2.7). These nomographs can be used to quickly
estimate required air/gas volumes.
To Find The Air Requirements
1. Find the drilling depth in feet, on the bottom scale.
2. Go up to the appropriate curve of Desired Fluid Weight After Aeration (Wd).
3. Go across to the appropriate curve of the difference between the Actual Fluid Weight before
Aeration and the Desired Fluid Weight After Aeration (Wa -Wd).
4. Go up to the Cubic Feet of Air needed per Barrel of mud on the top scale.
Example (Follow the Dotted Lines)
1. Consider a 4200 feet deep hole (bottom axis).
2. Move up to 6 ppg. Desired Fluid Weight After Aeration.
3. Move over to 4.5 ppg (10.5 ppg Actual Fluid Weight minus 6 ppg Desired Fluid Weight).
4. Move up to 73 scf air/bbl of mud.
Alternatively, the relationship developed by Poettman and Bergman can be solved in a
spreadsheet, to calculate the volumes, rather than using a nomograph.

5-11

Chapter 5

Well Engineering

Figure 5-3.

Qa =

Nomograph for determining air rates to reduce the hydrostatic pressure of a drilling
fluid, at an average temperature of 150F (Poettman and Bergman, 1955 4).

42 h f 808 (P2 P1 )
P
4.071 Tavg ln 2 0.0764 h
P1

P2 = P1 + 0.052 hd

(5.15)

(5.16)

where:
h......... depth of interest, (feet (TVD)),
P1 ....... surface pressure (psia), (14.7 psia
or local barometric pressure if
there is no backpressure),
P2 ....... pressure at the depth of interest,
formation pressure plus barometric
pressure (psia),
Qa ....... gas rate (i.e. air, nitrogen) (scf/bbl of
5-12

drilling fluid),
Tavg..... average annular temperature (R),
d ....... desired average density of the
drilling fluid (ppg), and,
f ........ density of the drilling fluid (ppg).
Example 3 shows how the basic Poettman and Bergman, 1955,4 equation can be used. Equation
(5.15) does not account for friction losses in the annulus. Poettmann and Bergman, 1955,4
showed that friction losses make very little difference in most instances. However, there are
instances where it can be important, especially when using high rates of air or nitrogen.
Example 3 (Standpipe Injection)

Consider a well that is to be drilled in a formation at 6000 feet that has a pressure equivalent to
7.5 ppg (~2350 psi at 6000 feet). The average annular temperature will be 90F (550R). It is
planned to use a 9.0 ppg base drilling mud. Air must be added to reduce the hydrostatic head and
to develop underbalanced conditions. Determine how much air must be injected into the mud in
order to make the bottomhole pressure at 6000 feet equivalent to a 7.5 ppg mud.
 First, determine the pressure at the zone of interest (6000 feet in this case) using Equation
(5.16).
P2 = P1 + 0.052 hd
P2 = 14.7 + ( 0.052)( 6000 feet )(7.5 ppg) =
= 2354.7 psia
(assuming 14.7 psi is the local barometric pressure)
 Next, use Equation (5.15) to determine the number of standard cubic feet of air per barrel of
drilling fluid that must be injected into the standpipe.
Qa =

42 h f 808 (P2 P1 )
P
4.071Tavg ln 2 0.0764 h
P1

(42)( 6000)(9) (808)( 2354.7 14.7)


(4.071)(550) ln

2354.7
( 0.0764)( 6000)
14.7

= 34.59 scf per barrel


Therefore, 34.59 standard cubic feet of air must be added to every barrel of 9 ppg mud in order
for the pressure at 6000 feet to be equivalent to a 7.5 ppg mud.
5-13

Chapter 5

Well Engineering

Parasite String
The previous example implies uniform aeration of the column. If a parasite string is used, only
part of the column is lightened. Equations (5.15) and (5.16) can be modified to allow the same
calculation for injection down a parasite string. The drilling fluid will be aerated from the depth
of the parasite string to the surface. The drilling fluid below the depth of the parasite string will
not be aerated. The modified equations are:
Qa =

42 h p f 808(P2 P1 )
P
4.071Tavg ln 2 0.0764 h p
P1
(5.17)

P2 = P1 + 0.052 hd 0.052 h h p f
(5.18)
where:
h......... depth of interest (feet) (TVD),
hp ....... depth of the parasite string (feet)
(TVD), and,
P2 ....... pressure at the depth of the parasite
string (psia).
The average annular temperature is only considered over the depth interval from the bottom of
the parasite string to the surface.
Example 4 (Parasite String)
Given all the data in Example 3, determine the air volume that is required if a parasite string is
set at 2000 feet. The average annular temperature is only considered from 2000 feet to the
surface and not from 6000 feet to the surface. The average annular temperature was estimated to
be 70F (530R).
 First, calculate P2. This is the required pressure at the bottom of the parasite string.

P2 = P1 + 0.052 hd 0.052 h h p f
P2 = 14.7 psia + ( 0.052)(6000 feet )(7.5 ppg)
( 0.052)(6000 feet 2000 feet )(9 ppg)
= 482.7 psia

5-14

This indicates that the pressure where the air enters the annulus from the parasite string must be
482.7 psia, in order for the equivalent mud weight at 6000 feet to be 7.5 ppg.
 Now, calculate the required air volume per barrel of mud:
42 h p f 808(P2 P1 )
Qa =
P
4.071Tavg ln 2 0.0764 h p
P1

Qa =

( 42)(2000)( 9) (808)( 482.7 14.7)


( 4.071)(530) ln

482.7
( 0.0764)( 2000)
14.7

= 5120
. scf per barrel
Since all of the fluid in the annulus is not being aerated, more air is required to achieve the same
equivalent mud weight at 6000 feet when a parasite string is used. Figure 5-4 is a plot of the
annular pressure profile for this example well, using both aeration through the standpipe and
through a parasite string.
0

1000
Standpipe
Parasite

Depth (feet)

2000

3000

4000

5000

6000
0

500

1000

1500

2000

2500

Pressure (psia)

Figure 5-4.

Plot of pressure versus depth for the wells in Examples 3 and 4, injecting through the
standpipe or a parasite string at 2000 feet.

5-15

Chapter 5

Well Engineering

Annular Calculations
There is another, relatively simple way to determine pressures in a well drilled with gasified
fluid. It involves calculating volumes, hydrostatic pressures and friction losses in the annulus.
Pressures are incrementally calculated (from first principles).
 The pressure at the surface is assumed and the volumetric flow rates are calculated for a short
interval (usually around 100 feet), starting from the top.
 Based on the flow rates and fluid properties, the pressure at the bottom of the first interval is
calculated, using both friction losses and hydrostatic pressure.
 The pressure at the bottom of this first interval is used to recalculate the fluid properties and
flow rates. These are used to calculate the pressure at the bottom of the second interval. The
process is repeated until the pressure at the zone of interest is determined.
 If desired, you can keep going (incorporate pressure drop through the bit and work back up
through the drillstring) to get standpipe pressure.
Performing these calculations by hand can be tedious. Spreadsheets can be used to perform the
calculations efficiently. The calculations must be done over short intervals because the actual
volume, density and flow rate change with pressure. Since gas is compressible, the volume is a
strong function of pressure. The volume of the gas determines the flow rate and density of the
single phase (air) or composite (gas-liquid) mixture.

Calculating Gas Density and Volume

 The gas density and volume can be calculated, at any pressure, using the following equations:
g =

2.703SPi
Ti

(5.19)

Vg =

0.35Q a Q L
Pi

(5.20)

where:
Pi ........ pressure at the top of the first
interval (psia),
Qa ....... injection gas volumetric flow rate
(sft3/bbl of drilling fluid),
QL ...... base liquid volumetric flow rate
(gpm),
S......... gas gravity (air = 1),
Ti ........ temperature at the top of the first

5-16

interval (R),
Vg....... gas volumetric rate at pressure
(ft3/min), and,
g ....... gas density at temperature and
pressure (lbm/ft3).
 The density and volume of the liquid are determined. Equations (5.21) and (5.22) are then
used to convert liquid density (ppg) and rate (gpm) to units of lbm/ft3 and ft3/min.
L = 7.48f
VL =

QL
7.48

(5.21)
(5.22)

where:
QL ...... base liquid volumetric flow rate
(gpm),
VL ...... base liquid volumetric flow rate
(ft3/min),
f ........ drilling fluid weight (ppg), and,
L ....... drilling fluid density (lbm/ft3).
 The next step is to determine the combined density and volume of the mixture:
VT = Vg + VL
m =

g Vg + L VL
VT

(5.23)

(5.24)

where:
VT ...... volumetric flow rate of the mixture
of gas and liquid (ft3/min), and,
m....... density of the mixture of gas and
liquid (lbm/ft3).

5-17

Chapter 5

Well Engineering

 The hydrostatic pressure and friction losses for the discrete interval (annulus) are calculated
with the following equations, along with the pressure at the bottom of the interval (ignoring
cuttings) :
Phy =

m L tvd
144

(5.25)

(5.88 10 )
4

Pfr =

(D

h Dp

0 .81
m
3

VT1.81 PV 0.19 L md

) (D

h + Dp

1.81

(5.26)
Pi +1 = Pi + Phy + Pfr

(5.27)

where:
Dh....... diameter of the hole (inches),
Dp....... outside diameter of the pipe
(inches),
Lmd ..... measured length of the calculation
interval (feet),
Ltvd ..... true vertical length of the
calculation interval (feet),
Pfr ....... friction losses in the calculation
interval (psi),
Phy ...... incremental hydrostatic pressure
over the calculation interval (psi),
Pi+1 ..... pressure at the bottom of the
calculation interval (psi), and,
PV...... plastic viscosity (cP).
Using Equations (5.19) through (5.27), the annular pressures in the wellbore can be calculated at
any depth. Hydrostatic pressure and friction losses are both included. These equations can also
be used in a directional well since the hydrostatic pressure and friction pressures are calculated
separately. Hydrostatic pressure is a function of the true vertical depth and friction losses are a
function of measured depth.
The following example illustrates how these equations can be used to calculate the pressures in
the wellbore described in Example 3.

5-18

A solid phase can be added for cuttings.


incorporated.

In most cases this is second order.

Liquid inflow can also be

Example 5
Given the conditions from Example 3, determine the pressure at 6000 feet. The air injection rate
is 34.59 ft3/bbl of drilling mud. The drilling fluid rate is 250 gpm. The hole size is 8.5 inches
and the outside diameter of the drillpipe is 4.5 inches. There are 600 feet of 6.25-inch OD drill
collars at the bottom of the drillstring. Assume that the surface temperature is 60F and the
geothermal gradient is 1F/100 feet. The returns are being vented through a mud/gas separator
and the surface pressure is assumed to be 14.7 psia. The calculation interval length will be 100
feet. The plastic viscosity of the drilling fluid is 10 cP.
 Calculate the density and volume of each phase for the first interval. This interval is near the
surface. The temperature is ~60F which is (60+459.67) = 520R. The densities and flow
rates are:
g =

2.703SPi (2.703)(1)(14.7 psia )


=
=
Ti
520 R

= 0.0764 lbm / ft 3

Vg =
=

0.35Q a Q L
Pi

( 0.35)(34.59 ft 3 / bbl)(250 gpm)


14.7

= 205.89 ft 3 / min
L = 7.48 f = ( 7.48)(9.0 ppg) =
= 67.32 lbm / ft 3
VL =

QL
250 gpm
=
= 33.42 ft 3 / min
7.48
7.48

 Calculate the total flow rate and the density of the mixture of air and drilling fluid.
VT = Vg + VL = 20589
. + 33.42 =
= 239.31 ft 3 / min

5-19

Chapter 5

m =
=

Well Engineering

g Vg + L VL
VT

( 0.0764)(20589
. ) + ( 67.32)( 33.42)
239.31

= 9.47 lbm / ft

 Calculate the hydrostatic pressure, friction losses and pressure, at the bottom of the first
interval.
3
m L tvd (9.47 lbm / ft )(100 feet )
=
=
144
144
= 6.58 psia

Phy =

Pfr =

(588
. 104 ) 0m.81 VT1.81 PV 0.19 L md

(D

Dp

) (D
3

+ Dp

1.81

(5.88 104 )(9.47lbm / ft 3 )0.81

(8.5 in 4.5 in) (8.5 in + 4.5 in)


3

1.81

(239.31 ft / min) (10 cP) (100 ft )

(8.5 in 4.5 in) (8.5 in + 4.5 in)


3

1.81

0.19

1.81

= 171
. psia
Pi +1 = Pi + Phy + Pfr = 14.7 + 6.58 + 171
. =
= 22.99 psia
 The pressure at 100 feet is 22.99 psia. For the next interval, from 100 to 200 feet down the
annulus:
g =

( 2.703)(1)( 22.99)
520 R + 0.01o F / ft 100 feet
o

= 01191
lbm / ft 3
.
( 0.35)( 34.59)( 250)
Vg =
=
22.99
= 13165
. ft 3 / min

5-20

 The volume and density of the liquid remain approximately the same as for the first interval
since the base liquid is a slightly compressible fluid:
VT = 13165
. + 33.42 = 165.07 ft 3 / min
m =

( 01191
)(13165
.
. ) + ( 67.32)(33.42)
165.07

= 13.72 lbm / ft 3
Phy =
Pfr =

(13.72)(100)
144

= 9.53 psia

0.81
1.81
(588
. 10 4 )(13.72) (165.07)

(8.5 4.5) 3 (8.5 + 4.5)1.81

(10) 0.19 (100)


(8.5 4.5) 3 (8.5 + 4.5)1.81

= 118
. psia
Pi + 2 = Pi +1 + Phy + Pfr
Pi + 2 = 22.99 + 9.53 + 118
. = 33.70 psia
 The pressure at 200 feet is 33.70 psia. The procedure is repeated for the interval from 200 to
300 feet:
g =
Vg =

(2.703)(1)(33.70)
522

.
lbm / ft 3
= 01743

(0.35)(34.59)(250)
33.70

= 89.81 ft 3 / min

5-21

Chapter 5

Well Engineering

VT = 89.81 + 33.42 = 123.23 ft 3 / min


m =

( 01743
)(89.81) + (67.32)( 33.42)
.
123.23

= 18.38 lbm / ft 3
Phy =

Pfr =

(18.38)(100)
144

= 12.76 psia

(5.88 104 ) (18.38) 0.81 (123.23)1.81


(8.5 4.5) 3 (8.5 + 4.5)1.81

(10) 0.19 (100)


=
(8.5 4.5) 3 (8.5 + 4.5)1.81

= 0.88 psia
Pi + 3 = Pi + 2 + Phy + Pfr
Pi + 2 = 33.70 + 12.76 + 0.88 = 47.34 psia
 The pressure at 300 feet is equal to 47.34 psia.
 The procedure is repeated until the depth of interest is reached. In this case, the pressure at
6000 feet is calculated to be 2362 psia. This pressure is equivalent to a 7.57 ppg mud. This
implies that the friction losses in the annulus are equal to an equivalent mud weight of only
0.07 ppg (~0.36 psi per 100 feet). Given the fact that air rates do not remain absolutely
constant and that connections have to be made, the friction losses are insignificant. This
indicates that variations in actual bottomhole pressure are more a function of changes in air
volumes during connections than friction losses. Equation (5.17) is a good starting point for
any field operation. The actual volumes must be varied on location in order to accomplish
the project objectives.

5.4

Wellhead Design

Wellheads used in underbalanced drilling vary from crude, very simple equipment for very low
pressure operations to expensive, redundant systems designed for very high pressure operations.
For extremely low pore pressure drilling applications, a simple annular preventer alone might
suffice to contain wellbore pressures; however, a principal manufacturer of such equipment
strongly cautions that such use exceeds the design criteria of this equipment. Therefore, the
minimum setup for an underbalanced drilling system should consist of a rotating head
mounted above a two ram set of manually-operated blowout preventers, consisting of a pipe
5-22

ram and a blind ram. An improvement to this basic system would be installing the rotating head
above a set of hydraulically-operated blowout preventers. For slightly higher pressure operating
conditions, a system consisting of a rotating head, an annular preventer and a two ram set of
manually-operated preventers will probably work adequately. For added safety, hydraulicallyoperated preventers with a manual backup should be provided.
These basic systems all use a rotating head with a 400 psi (sometimes 500 psi) MWP (Maximum
Working Pressure) capability. Early day air drilling used this type of equipment, along with a
surface pit and a return flowline venting to this pit. Mist and foam drilling extended the depth
limits and maintained low surface pressures in this underbalanced system.
A rotating head is sufficient for almost all air and mist operations, in areas where formation
pressures are well characterized. The rotating head is used as a diverter only and is not used to
maintain a backpressure; therefore, there is no backpressure limitation. In fact, backpressure is
detrimental to air and mist drilling and should not be applied.
It is extremely unlikely that the flow rate through the blooie line would create enough
backpressure to exceed the pressure limitation of the rotating head. The flow rate would have to
be in excess of 100 MMscf/D to cause a 200 psi pressure drop in an 8-inch blooie line.
When accurate pressure data are available for an area, drilling with a rotating head is possible in
deeper wells, especially if the target contains oil rather than gas. Because of lower rental rates,
rotating head drilling applications are stretched to their pressure limits, before resorting to more
costly RBOP equipment. Many Pearsall and Giddings (Austin Chalk) field wells are still drilled
with this style of equipment, despite the safer MWP offered with the RBOP. Regrettably, cost
factors versus safety margin considerations often win out in favor of the rotating head. Today, a
typical well in these fields is drilled with a rotating head system, in combination with an annular
BOP and a hydraulically-operated three ram BOP stack, consisting of two sets of pipe rams and
one set of blind rams.
Return flow is diverted through the choke manifold and surface separation equipment to isolate
gas, oil, drilling fluids and cuttings. An automatic flare system safely burns off hydrocarbon or
hydrogen sulfide gases from the rig site. Even moderately deep wells are drilled in this manner
using nothing but fresh water as the drilling fluid. Heavier 10.0 ppg NaCl brine water is then
used to reduce surface pressures during trips.
To drill these deeper wells, many operators have added the additional set of hydraulic pipe rams,
to make a two pipe ram stack with blind rams. Above these BOPs, the annular preventer, with
the rotating head on top, completes the medium pressure wellhead equipment. The additional set
of pipe rams provides a higher level of operating safety, as well as making it possible to snub or
strip from the well. As higher pore pressures are encountered, sodium chloride brine fluids can
replace the previous fresh water system and provide lower surface operating pressures, in order to
stay in the pressure range of the rotating head.

5-23

Chapter 5

Well Engineering

The next higher level of underbalanced wellhead equipment usually involves drilling deeper
formations or gas-bearing reservoirs. Either one or both of these conditions can cause surface
pressures to exceed the working pressure limit of the rotating head. For these wells, the RBOP,
with a 1,500 psi MWP (some equipment now is functional at 2500 psi operating/rotating and
5000 psi static pressures), offers substantial depth capacity and safety advantages compared to
the rotating head.
If hydrogen sulfide (H2S) gas is expected or if formations with even higher pore pressure are
drilled, wellhead equipment design might call for either coiled tubing drilling (CTD) or snub
drilling operations.
Rotating Head
A typical rotating head and its components are shown in Figure 5-5. A schematic cross-section
of a rotating head is illustrated in Figure 5-6. In air and gas drilling operations, rotating heads are
also called air heads. Rotating heads are used to pack off the annulus, diverting the air and gas
flow down the blooie line. Without the rotating head, the air and gas would come up through the
rotary table and onto the rig floor.
Many adequate rotating heads are available. Newer-generation rotating heads usually have sealed
bearings, rather than external oilers. Most are driven by a kelly driver. This attaches to the kelly
and is mated to a machined piece on top of the bearing assembly. The kelly driver transfers
rotation of the drillstring to the sealing element in the rotating head. The bearing assembly
provides a seal and allows rotation of the stripper rubber while the bowl remains stationary. The
stripper rubber is designed to rotate with the kelly since rotating the kelly within the stripper
rubber would cause the stripper rubber to wear out much faster.
Hexagonal kellys allow for a better seal than do square kellys. These should be used whenever
possible for air drilling applications. The life of the sealing element will be increased by proper
lubrication and minimal tripping through it. The drillpipe can be tripped through the stripper
rubber when necessary but it will not last as long before it has to be replaced. If the well is
making smaller amounts of gas, the stripper rubber and bearing assembly can be pulled prior to
tripping. Gas can be jetted off of the rig by passing air through the primary jet in the blooie line,
as explained in Chapter 2.

5-24

Figure 5-5.
Components of a low pressure
5
rotating head (after Bourgoyne, 1995, source:
Williams Tool Company, Inc.).

Figure 5-6.
Cross-section of a rotating
head, showing how the stripper rubber seals
around the drillpipe or kelly, diverting flow
down the blooie line (after Cooper et al., 1977
1
).

In an air drilling operation, the rotating head is only a diverter and must not be thought of as
a replacement for a properly designed blowout preventer stack. Pressures on the sealing
element should be kept to a minimum. Most low pressure rotating heads are not designed to
handle much more than a few hundred psi. High pressure rotating heads or rotating blowout
preventers are available for underbalanced drilling.

5-25

Chapter 5

Well Engineering

The stripper rubber in a rotating head can be changed without tripping out of the hole. The driller
first pulls the pipe until a tool joint is at the bottom of the stripper rubber. The annular preventer
(or low rams) is closed and the pressure trapped below the rotating head is bled off. The bonnet
on top of the bowl is opened and the drillpipe is stripped from the hole. The first tool joint below
the rubber should pick up the stripper rubber and bearing assembly as the pipe is hoisted from the
hole. The tool joint is a larger OD than the pipe and will not slide through the rubber as easily.
When the stripper rubber and bearing assembly are above the rig floor, the slips are set and the
drillpipe connection just below the stripper rubber is broken. The stripper rubber and bearing
assembly can then be pulled off of the drillpipe. The old stripper rubber is released from the
bearing assembly and replaced with a new stripper rubber. The new stripper rubber and bearing
assembly are placed on top of the mouse hole and the drillpipe is stabbed through the rubber with
the aid of a spear. Once the new stripper rubber is above the tool joint, the tool joint is made up
to the drillstring and lowered into the hole. The stripper rubber and bearing assembly are again
placed in the bowl and the bonnet on top of the bowl is closed. Drilling can then continue with
the new stripper rubber.
Specific Considerations
Dry Air Drilling
A conventional open bell nipple will not direct the returning air flow away from the rig
substructure. To do this, additional equipment, a diverter, is required above the BOP stack.
Although it is possible to use various types of equipment as diverters, it is now normal to use
either a rotating head (Figure 5-6) or a rotating BOP (Figure 5-7).
Both of these use elastomeric elements to seal around the kelly and direct the return flow laterally
through the outlet and into the blooie line. The principal difference between these two types of
diverter is that the sealing element in a rotating head is actuated by the air pressure that it seals,
whereas the element in a rotating BOP is actuated hydraulically. Typically, rotating heads have a
pressure limitation of 400 psi, and new-generation rotating BOPs can seal higher pressures, up to
2,500 psi while drilling (5000 psi static).6 It is important that both are operated according to
their manufacturers recommendations. Un-acceptably rapid wear of the seal element and
mechanism will occur if the axis of the diverter is not aligned directly with the center of the
rotary table or if the lubrication is inadequate.

5-26

Quick-change packer
assembly

Kelly driver
assembly

Hydraulic fluid Inlet


Inner packer
Outer packer
Bearings

Mechanical seal

Hydraulic fluid return

Outlet flange

Casing flange

Figure 5-7.

Blowout preventer

With dry air drilling, the well should be stripped through the rotating head rubber and not the
primary BOPs. If the flow rate or pressures are too high to use the rotating head for stripping, the
well should be killed before leaving the bottom. After all, the pipe must be stripped out of the
well before it is stripped back into the well.
If the flow rates and pressures are high enough to require stripping ram to ram through the BOP
stack, then a snubbing unit will be required, with a snubbing stack.
The diverter system does not remove the need for a conventional BOP stack. This should
comply with local regulatory requirements. At a minimum, it should contain pipe and blind
rams, so that the well can be shut-in with the string in or out of the well. For gas wells at least,
the pipe and blind rams should be able to support the highest anticipated formation fluid
pressure. Where it can be accommodated beneath the rig floor, it is desirable to have a full stack
consisting, from the wellhead up, of pipe rams, blind rams, pipe rams and annular. This provides
operational flexibility. For example, it allows stripping pipe back into the well under pressure if
high pressure hydrocarbons are encountered that cannot be contained within the lower pressure
capacity of the diverter. The returning air flow is taken from the diverter to a flare pit through
the blooie line.

5-27

Chapter 5

Well Engineering

Nitrogen Drilling
The requirements are similar to those for dry air drilling.
Drilling with Natural Gas
Beyond the additional surface equipment, including gas monitoring instrumentation described in
Section 2.3, the required equipment is similar to that needed for air drilling. Cummings, 1987,7
indicated that a conventional rotating head could be used to divert gas flow into the blooie line in
certain situations (i.e. low permeability wells that need to be hydraulically fractured).
Particularly if it is anticipated that high formation gas pressures or production rates may be
encountered, the additional pressure capacity of a rotating blowout preventer is worth
considering.

Regardless, implement, follow and strictly enforce all safety requirements.


Mist Drilling
The requirements are similar to those for dry air drilling.
Foam Drilling
A choke should be installed in the blooie line, close to the rotating head or RBOP, to pressurize
the annulus, if necessary. Dupont, 1984,8 recommended a bladder-type choke. If the circulating
program indicates that annular backpressure may be necessary, then this additional pressure
should be considered when specifying the pressure capacity of the rotating head or RBOP. The
section of the blooie line between the choke and the rotating head should have a pressure rating
sufficient to support the highest back-pressure likely to be imposed.
In very cold conditions, foam returns may freeze and plug the blooie line. In these areas, Fraser
and Moore, 1987,9 recommended using an additional foam discharge line, with both the blooie
and foam discharge lines leading to the flare pit. This redundancy is not necessary for normal
operations, where there will be a line from the choke manifold to the flare pit that could be used
to continue circulation, but not drilling, if the blooie line were to plug. Small rigs may not have
the clearance below the floor for two diverters on top of a conventional BOP stack.
Gasified Liquids
In the United States, many of the gasified liquid drilling applications to-date (when surface
pressures are anticipated to be low), have used aerated water or mud to overcome lost circulation
in hole intervals above known reservoirs. In these instances, rotating heads are used on top of
conventional BOP stacks, to seal around the drillstring and divert flow into the flowline.
Nitrified liquids are often used in Canada, to drill very productive wells underbalanced. An
RBOP is normally used, rather than a lower pressure rotating head. If the rig is equipped with a
top drive, it is possible to use dual annular BOPs to give a high pressure seal around the string
above the return line. Since an RBOP gives better control of the closing pressure and has lower
stripping friction, it is usually preferred.10

5-28

It has been recommended that blind rams should be installed at the bottom of the BOP stack.11
This maximizes the distance between the blind ram and the RBOP, allowing short, irregularly
shaped BHA components to be run into the well under pressure, without relying on the RBOP or
annular to seal around them.
A second set of pipe rams, below the blind rams, will provide redundancy, and will normally
only be used to shut in the well, in the event that work is required on one of the elements higher
up the stack.12
The influence of the BOP stack configuration on ram-to-ram stripping operations should be
carefully considered, if there is any possibility of having to strip into the well under high
pressure.
The clearance beneath the rig floor should be considered when designing the well cellar,
wellhead and BOP stack. BOP stacks, for underbalanced drilling, tend to be taller than those
used in conventional operations. The situation is more difficult if conventional wellhead
equipment is used to inject gas down a temporary casing string.
Additional discussion of gasified liquids is found in Chapter 2, Section 2.7.
Flowdrilling
Although surface equipment is important in any type of drilling, it is crucial to the success of
flowdrilling. Wellhead equipment and operating procedures are discussed in detail in Chapter 2,
Section 2.8.
Mudcap Drilling
One significant difference between flowdrilling and mudcap drilling is that an RBOP, with its
high operating pressure limits, is essential for mudcap drilling. Flowdrilling is possible using
either an RBOP or a rotating head. A schematic layout for mudcap drilling is shown in Figure 5-8.
In designing a well plan using a mudcap drilling format, drilling engineers should carefully
consider the high standpipe pressure involved and the associated safety considerations, before
finalizing their recommendation.
Bloys et al., 1994,13 discussed offshore floating mudcap drilling, in carbonate zones in South
East Asia. Typically mud with a density only slightly higher than the pore pressure in the top of
the exposed section of carbonate is continuously pumped into the annulus at a rate that maintains
a constant fluid level sufficient to prevent any influx. (Mud pump rates depend on the loss rate to
the formation but rates of the order of 10,000 bbls/day are typical). Sea water is pumped down
the drillpipe at normal circulation rate for the hole size to clean and cool the bit and move
cuttings up the hole. To make connections, mud is pumped down the drillpipe to minimize the
differential across the non-ported drillstring float but it is the float that actually prevents any
influx entering the drillpipe.

5-29

Chapter 5

Well Engineering

With a ware ship [supply ship] mudcapping may be continued during bit trips and while running
casing but this involves taking extra risk as there will be times when the BOP cannot effectively
close off the well, e.g. when the BHA is across the BOP. The safe method for tripping is to stop
the losses in the open hole before tripping. Others have successfully stopped losses for tripping
by spotting a fast setting, viscous, cement slurry just inside the casing shoe as an alternative to
plugging the formation to cure losses but this is not recommended.

GAS
BUSTER

To flare pit

MUD
PITS
RIG FLOOR

HCR
Valve
(Closed)

MUD
PUMPS

Chemical
Injection
CHOKE
(Closed)

Section B
DIVERTER

Section A

Figure 5-8.

Schematic of equipment required for mudcap drilling (courtesy of Signa Engineering


Corporation).

Future refinements in mudcap drilling that have been considered , but not yet applied
offshore, would involve the use of a rotating head (or rotating BOP) and lower density mud being
pumped into the annulus . Currently it is only possible to drill with a rotating head on rigs
equipped with a surface stack; this includes jackup and platform type rigs offshore. The
advantages of this procedure are that there is no guess work about the fluid level in the well and
the pump rate required on the annulus side is dictated only by the need to prevent gas migration.
In gas filled carbonates the increasing over-balance would also be reduced by using a lower
density mud. Hence, the logistics of pumping mud down the annulus could be planned more
precisely based on the velocity required to prevent gas influx and migration (i.e. up to 2,200 m/hr
[120 fpm] depending on fluid viscosity).3 Hence, the use of a rotating head and lower density
mud could potentially reduce the volume of fluid required and the cost to drill a given massive
lost circulation interval.

5-30

One significant concern is to guarantee that riser collapse pressure is not exceeded. The
decision to use mudcap drilling in the operations described by Bloys, et al., 1994,13 is only
made after the first occurrence of total losses has been experienced and cured.
Snub Drilling and Coiled Tubing Drilling
Both snubbing and CT units have BOP stacks that allow a drillstring (coiled tubing in the latter
case) to be run into or out of the hole, at much higher pressures (routinely up to 10,000 psi) than
can be tolerated by either a rotating head or an RBOP. Both units also allow the drillstring to be
pushed into a well under pressure, even when the weight of the string alone is insufficient to
overcome the pressure tending to push it out of the well. Snubbing and CT units can by used
for underbalanced drilling, at pressures that cannot be managed by conventional surface
equipment. Chapter 2, Section 2.10, provides a description of required wellhead equipment and
operating con-siderations.
Closed Systems
Surface equipment commonly used in closed system drilling operations is discussed in Chapter 2,
Section 2.11.

5.5

Casing Design

Casing design for an underbalanced hole is not substantially different than for conventional
drilling. Normally, the casing is designed for tension, internal yield and collapse (biaxial casing
design), and a design factor is assigned for all three design parameters (using API Standard 5A,
operators company policies, regulatory agency requirements, whichever is more stringent).
For tension, some operators will calculate the string weights based on the buoyant weight of the
casing while others will calculate string weight based on air weight and use buoyancy as an
additional design factor. In an air hole, it is very common to run casing without filling the well
with fluid first. In this case, there is no buoyancy and the string weight should be based on air
weight rather than buoyant weight. Most operators use the same tension design factor for air as
they do for fluid.
The maximum internal yield pressure is usually based on the maximum anticipated shut-in
pressure. The maximum pressure at the depth of interest will be either the shut-in bottomhole
pressure minus a gas gradient or the fracture gradient at the shoe minus a gas gradient; whichever
value is lower. Since the design is the same for air or fluid, the same design factor is typically
used.
In designing for collapse resistance in fluid, it is common to assume that the pressure inside the
casing is zero and that the pressure in the annulus is equivalent to a full column of mud. In a
fluid-filled hole, the pressure inside the casing is rarely zero. Some operators do not assume that
the pressure inside the casing will be zero and there will always be some fluid inside the casing.
For this reason, the collapse design factor for fluid drilling is usually very close to one; however,

5-31

Chapter 5

Well Engineering

the pressure inside the casing will always be very close to zero in an air-drilled well. Most
operators will increase the collapse design factor when designing casing for an air-drilled hole. A
common design factor for a fluid-filled hole is 1.125 and a common design factor for an airdrilled hole is 1.20. Similar logic is followed if the drilling fluid is not dry air, although dry air
will likely be one of extreme situations for all of the design parameters.
There are other considerations for casing design. Even when designing casing for a fluid-filled
hole, the potential reduction in wall thickness due to corrosion and casing wear must be
considered. Casing wear must be considered in an air drilled hole.
Corrosion
For fluid-filled wells, corrosion is seldom a consideration while drilling. However, it must
usually be considered during the productive life of the well. Corrosion is not a factor when
drilling with dry air. If there is no water present, corrosion cannot occur. Corrosion can be a
problem with mist, foam and aerated fluids. Aerated fluids provide the most risk. In most air
drilling operations, corrosion is controlled with corrosion inhibitors. It is relatively easy to
control corrosion in a mist operation by adding corrosion inhibitors to the mist. As a result,
corrosion is seldom a problem when drilling with mist; if the mist contains salt for shale
stabilization, uninhibited corrosion rates will increase substantially.
Foams and aerated fluids will typically have higher corrosion rates, even when they are treated
with inhibitors. If a well is going to be drilled for an extended period of time with foam or
aerated fluids, it may be necessary to design the casing with a slightly greater wall thickness.
Spending a little extra money on corrosion control is also advantageous.
Casing Wear
Casing wear is caused by rotation of the drillstring inside the casing. If tool joints rub against the
casing while drilling, both the tool joints and the casing will wear. Casing wear is accelerated in
an air-drilled hole because there is no lubrication between the drillstring and the casing.
Fortunately, most air drilled holes are drilled rapidly. Spending less time drilling (fewer
revolutions of the drillstring) will result in less casing wear.

5-32

For substantial wear to occur, there must be doglegs in the wellbore trajectory. A tool joint
laying against the casing, in or near vertical well sections, will not apply sufficient force to cause
much wear. The force with which a tool joint presses against the casing is commonly called the
tool joint normal force. Placing the tool joint in a dogleg and adding tension will substantially
increase the tool joint normal force. Therefore, casing wear will preferentially occur in sections
of the hole that have doglegs and where the drillstring has sufficient tension. For this reason,
casing wear usually occurs closer to the surface and is normally not a problem near the bottom of
the well.
If a well is being drilled in an area where deviation and doglegging are problems, casing wear
should be considered in the casing design. If only a few days are spent drilling below each casing
string, wear will probably not be a significant problem and can generally be ignored, unless
severe doglegs are present. Casing wear will only be a problem when a well is drilled below the
casing string for an extended period of time. Bradley and Fontenot, 1975,14 presented a method
to predict casing wear, depending on the conditions in the wellbore, including dogleg severity,
tension in the drillstring and rotating hours below the dogleg.

5.6

Completion Design

One of the primary advantages of drilling wells underbalanced is the elimination or minimization
of formation impairment. In overbalanced situations, drilling fluid and solids can penetrate and
damage matrix porosity or fractures, reducing the permeability. If a well is properly drilled
under underbalanced conditions, but is completed using overbalanced methods, much if not
all of the impairment-reducing benefits might be permanently lost. Even if this completionrelated damage can be removed or bypassed, the associated expenses can be avoided if the
operator uses proper underbalanced completion procedures. These procedures, sometimes called
live well, underbalanced completion techniques, are described in this section. They include:
 Running production casing, liners, slotted liners and other tools under-balanced,
 Controlled cementing of production casing or liners,
 Running production tubing and down-hole completion assemblies, and,
 Perforating underbalanced.
Running Casing and Liners Underbalanced
Before drilling operations are completed, and the bottomhole assembly (BHA) is removed from
the wellbore under pressure, completion protocol must be determined. For example, will the
completion be barefoot (openhole) or will some type of casing or liner be run.
If the completion is not barefoot, it becomes necessary to run the casing or liner without killing
the well. In this scenario, surface pressures are usually increased to subdue exposed downhole
formations, without exceeding their pore pressures. This is done by replacing the lighter annular
fluid by bullheading a heavier fluid down the back side before tripping out of the hole.

5-33

Chapter 5

Well Engineering

To run production casing or an unslotted production liner in a live well, a float shoe and float
collar are usually used. The shoe and float collar are often separated by two joints of pipe, in
order to isolate contaminated cement and to prevent it from surrounding the lower portion of
casing in the open hole. Depending on the surface pressures, it may be necessary to flow the well
through the choke manifold while running pipe, to reduce the shut-in surface pressure. Even
flowing the well might not sufficiently reduce this pressure to permit passage of the pipe into the
well against underbalanced forces. If this is the case, a snubbing unit or a coiled tubing injector
head might be required to push the casing until it becomes pipe heavy.
On the other hand, a slotted liner does not restrict the flow of fluids into the liner (through the
slots). The slotted liner and liner hanger are run on the bottom of drillpipe or some other work
string. A drill float is generally run above an on/off tool, located immediately above the liner
hanger. Once the hanger is set, the on/off tool is released and the drillpipe or work string is
tripped out of the hole. The drill float provides back flow protection.
It may be necessary to flood the backside with drilling fluid to reduce the surface pressure and
enable tools or pipe to be run into the hole. Fluid is continuously pumped down the annulus to
overcome pressure resistance. If necessary, the rubber element or packer, inside a rotating head
or RBOP, can be removed to allow larger diameter pipe to be run through the wellhead stack.
Cementing Pipe Underpressured
Presuming that casing has been run underbalanced, underbalanced cementing should also be
considered. Formation impairment from the cement and associated filtrate fluids can be equally
or more damaging than drilling. Underbalanced cementing is not substantially different from
underbalanced drilling. The hydrostatic head of the slurry can be reduced by entraining gas,
usually nitrogen, or reduced-density additives. These technologies were originally developed to
avoid breakdown in weak formations.
The requirements of the cement remain the same as for conventional treatments. There must be
annular sealing to prevent flow and strength must be adequate to resist degradation of the cement
bond under the action of in-situ stresses. Other common considerations are permeability
elimination in the microannulus, compressive strength and drilling fluid displacement.
Nitrogen is often added to cement to reduce its density. This reduces the hydrostatic head
adjacent to target reservoirs and may further impede flow (into the formation, before setting)
because of its multi-phase characteristics. Nitrified cement was originally developed for
placement across lost circulation or underpressured intervals. It is commonly used for
underbalanced cementing.
Historically, operators have had some difficulty in guaranteeing adequate bond of lightened
cement to the casing and the formation, using conventional cementing methods, particularly
across gas-bearing zones. Formation gas would often channel into this cement and cause
contamination. This honeycombing could cause inadequate bond, resulting in crossflow (flow

5-34

of formation fluids in channels or a microannulus) between different formations. When one of


these reservoirs contained water, a channel could allow water flow into a hydrocarbon interval,
resulting in premature abandonment of the completion and loss of reserves. Todays high quality
foamed cements tend to reduce this problem. Compressible slurries can maintain cement pore
pressure throughout setting and hardening, particularly through the transitional phase where the
cement matrix is vulnerable to gas channeling.15
Alternatively, normal extenders are suitable for slurries with densities as low as 11.5 to 12 ppg.
Below this, the resulting water separation will affect slurry properties and continuity of the
annular cement column. Hollow microspheres have been used as a cement extender. Lightweight slurries using hollow spheres, enable slurries in the 9 to 12 ppg range, with no water
separation and a reasonable cost. Also, as indicated, foamed cement, using nitrogen as the
extender, has been accepted by many operators. The system requires alternatively formulated
base cement slurries and gaseous nitrogen to create an homogeneous, ultra light-weight slurry.
Running Tubing in Underbalanced Wells
Whether or not a well is designed as an openhole completion, a slotted liner completion or a
perforated casing com-pletion, production tubing is generally required to protect production
casing against excessive or concentrated pressures and to minimize corrosion and/or erosion.
Most regulatory agencies enforce the use of tubing in well completions, to protect shallow
freshwater aquifers against hydrocarbon or salt invasion and pollution.
If the underbalanced well above is an openhole or slotted liner completion, there will generally
be pressure at the surface. Cemented casing completions will have zero pressure until they are
perforated. Methods have been developed to run completion assemblies and tubing downhole
underpressured, in openhole and slotted liner wells. Unperforated, cased wells present no
problem and tubing is run into the well without special equipment.
If a well has surface pressure, it should preferably not be killed (formation impairment from
completion fluids and solids might occur). Since it is not possible to use permanent tubing string
floats similar to drill floats, temporary float systems have been developed. By placing a tubing
sub, containing a custom glass disk in the string, pipe and tools can be run in the well under
pressure without backflow of wellbore fluids. The glass or other similar material isolates the
inside of the tubing from pressure while it is run in the well along with a retrievable packer or
seal bore assembly for a permanent packer installation. Once the packer is set and the system
pressure tested, this glass disk is broken by dropping a sinker bar down the tubing and breaking
the disk. A catcher assembly is usually positioned inside a mud anchor, located below a
perforated nipple underneath the packer, to keep the sinker bar from falling out of the tubing into
the casing. This is a simple and effective technique.
There are other methods to isolate surface pressure and trip into a well. For example, a wirelineset permanent packer can be run with a pump-out or push-out plug assembly. Once the packer is
run and set in place, pressure above the packer can be bled off to zero to run the tubing. After the
tubing is stung into the packer and pressure tested, either pump pressure or a sinker bar is used to
open the well to the surface; communicating the formation below the packer with the surface
through the production tubing.
5-35

Chapter 5

Well Engineering

Another method to protect against surface pressure while running tubing into the well involves
the use of a pressure rupture disk located inside the tubing string. Again, after setting and
hydrostatic testing of the packer, this pressure disk is ruptured by pressuring up the tubing to a
preset limit. Of all of these methods, the most commonly used is the shear glass disk sub
because of its reliability and simplicity.

5.7

Bit Selection

Basic principals of bit selection for underbalanced drilling are essentially the same as those that
should be followed for conventional, overbalanced drilling oper-ations. There are some detailed
differences, relating to the properties of the drilling fluid in use and how these affect the drilling
process. The first section below outlines the bit selection process. Subsequent sections describe
those aspects of bit selection which are specific to drilling with air, mist and foam and to drilling
with gasified liquid and liquid drilling fluids.
The Bit Selection Process
The steps in developing the most cost-effective bit program are:
 Assemble offset well data,
 Develop a description of the well to be drilled,
 Review offset well bit runs,
 Develop candidate bit programs,
 Confirm that the selected bits are consistent with the proposed BHAs and,
 Perform an economic evaluation, to identify the preferred bit program.
The end result of this selection process should be a bit program, consisting of the sequence of
bits that will allow the well to be drilled for the lowest cost, while meeting trajectory and
wellbore quality targets.
Assemble Offset Well Data
Identify a number of offset wells, ideally in the same field and drilled through the same target
formations. Assemble as much information as possible about drilling these wells. When
possible, this should include bit records, mud logs, wireline logs, daily drilling reports, as well as
mud and directional drilling reports.
Develop A Description of the Planned Well
Characterize the proposed hole geometry - hole sizes, casing points, and trajectory. Using
available information and experience, outline the anticipated values of rock hardness and
abrasivity at all depths.

5-36

 Sonic travel time logs give qualitative indications of the formation hardness - low travel
times correspond to rocks with high compressive strengths. There are several proprietary
correlations that interrelate unconfined compressive strength with sonic travel time and
lithological information.16,17,18 There are methods that use such correlations for bit selection.
17

 Abrasivity is more difficult to quantify. This is not an intrinsic rock property. It is possible to
form a qualitative assessment of the rocks potential for abrasive bit wear. In general terms, a
rocks abrasivity is related to the hardness of its constituent minerals, its bulk compressive
strength, the grain size distribution and the grain shape. Quartz is the most abrasive mineral
commonly encountered in sedimentary rocks. The higher the quartz content, the more
abrasive the rock. Similarly, the higher the bulk compressive strength, the more abrasive the
rock is likely to be. An experimental correlation has been found between the angle of
internal friction of sandstones and the abrasive wear rate of drill bit materials sliding over
them.18
 Make note of any formations that may have a special impact on bit performance; for example,
a conglom-erate with a high proportion of chert nodules would normally not be PDC
drillable.
 To organize the bit selection process, divide the well into distinct zones. 19 Each zone would
correspond to any significant change in formation proper-ties or drilling conditions. For
example, a new formation usually means a new zone, as does a change in hole size or a
planned BHA change.
Review Offset Well Bit Runs
 From offset well bit records, determine what bits have been used to drill through each
formation likely to be penetrated by the planned well. Identify which bits gave the best and
worst performance, in terms of overall penetration rate and footage. Remember that many
factors can cause the penetration rate or bit run length to be less than what could have been
achieved with better operating practices or parameters. Consider the specific circumstances
when evaluating each bit run; past or planned.
 Consider the dull grading carefully. Why was each bit run terminated, what was the bits
wear state at the end of the run, and how did that wear affect the overall penetration rate?
 Use the observed bit performance to infer the apparent hardness and abrasivity of the
formations penetrated - low penetration rates often correspond to hard rocks; abrasive rocks
will give short bit runs, ending with severe cutting structure wear and undergauge bits, etc.
Identify Candidate Bits
The next step in the bit selection process is to identify candidate bits for each zone to be drilled.
Initially, consider both roller cone and fixed cutter bits.

5-37

Chapter 5

Well Engineering

Roller Cone Bits


The key design features to be considered for roller cone bits are the cutting structure, the bearing
and seal types, and gauge protection. The cutting structure involves tooth material, count, tooth
length, tooth shape, and cone offset. Roller cone bit teeth may be either steel (milled tooth bits)
or tungsten carbide hard metal (insert bits). The cutting structure should be matched to a
formations anticipated hardness and abrasivity. Bearing and seal selections are influenced by
the intended operating parameters and the required run duration. The degree of gauge protection
should be matched principally to the formation abrasivity and the nature of the drilling fluid.
Fixed Cutter Bits
Design features to be considered for fixed cutter bits include the cutting structure, the body
material and profile, gauge, and stabilizing (anti-whirl) features. For these bits, the cutting
structure involves the cutter material (PDC or Polycrystalline Diamond Compact, TSP or
Thermally Stable Polycrystalline diamond, impregnated diamond, and natural diamond), the size,
backrake, and density. As with roller cone bits, the cutting structure should be consistent with
the formation hardness and abrasivity.
 PDC cutters usually experience excessively rapid wear in very hard and abrasive formations
or formations that contain numerous hard inclusions, such as chert and pyrite.
 Impregnated and natural diamond bits can tolerate very hard and abrasive formations.
 The required gauge protection depends on a formations abrasivity.
 Requirements for directional drilling may influence the nature and length of gauge that is
suitable with a fixed cutter bit.
 To penetrate hard formations with a PDC bit, it may be necessary to specify anti-whirl
features for the bits cutter layout (to avoid vibrational damage).
Develop Bit Programs
At this stage of evaluation, the intent should be to develop several alternative bit programs.
These should consist of the bit type (or sequence of bit types) to be run, the start and end depths,
and anticipated penetration rates for each run.
For these remaining candidate bits, estimate their probable penetration rate and footage in the
different zones.
 Are there bits that could drill the entire hole section in one run?
 If so, the softest formation bit that can drill the interval without excessively rapid wear may
often be the most effective selection.
 If not, evaluate possible bits for different combinations of adjacent zones in each hole
interval. Make allowance for bit wear experienced in one zone on the penetration rate in
subsequent zones.

5-38

When considering candidate bits, remember the bit run objectives. In some instances, these may
be fixed by directional drilling considerations or by the casing program. It may not be
appropriate to select a bit that is capable of effectively drilling thousands of feet, if a trip is
planned to change the BHA after 500 feet or if the interval is only that long.
Confirm that the Bits are Consistent with the BHAs
Before proceeding any further with each candidate bit program, confirm that the bits are
consistent with the proposed BHAs. In other words, do the operating parameters of the proposed
BHAs inhibit bit performance? The BHA can limit bit performance if WOB is limited because of
restricted BHA weight or stabilization, or if the selected downhole motors operate at rotary
speeds that are too high for the chosen bit. It may be necessary to iterate the bit selection process
in order to develop a bit program that is consistent with all other aspects of the drilling operation.
Evaluate the Economics
By this time in the evaluation, several bits or sequences of bits should have been identified as
candidates for each hole interval. Use the estimated penetration rate and bit life to predict the
probable cost for each bit run:
C h i = C ri Ti + C b i
(5.28)
where:
C ri ..... the hourly cost of operating the rig
during that bit run, including the
rig rate, fuel, all special services
and rental items,
Ti ........ the duration of the run in hours.
C bi ..... is the cost of the bit.
The duration of the bit run should include all non-productive activities, such as making
connections, surveying, circulating, etc. Different amounts of non-productive time may be
required by different drilling techniques.
The predicted cost to drill the interval Ci is the sum of the costs predicted for each bit run in the
program.
Ci = C hi

(5.29)

Rank the alternative bit program according to the predicted cost to drill the hole interval.

5-39

Chapter 5

Well Engineering

Finally, determine the sensitivity of the predicted drilling cost, for each candidate bit program, to
uncertainties in the predicted drilling performance and to potential problems associated with each
candidate bit.
Bit Selection for Dry Gas, Mist and Foam Drilling
This section describes the main ways in which bit selection for drilling with dry gas, mist and
foam differ from bit selection for other drilling techniques. The effective hardness of the rock
will probably be somewhat lower than it would be for overbalanced drilling with mud. The rock
is subjected to lower confining stress due to lower borehole pressures. This may allow using bits
that are intended for softer formations than would have been appropriate when drilling with mud.
Roller Cone Bits
When drilling with mud, it is desirable to have as much of the hole bottom as possible contacted
by the bits teeth. This avoids leaving uncut rings of rock. The much lower borehole pressures
associated with dry gas, mist and foam cause more brittle rock failure. This creates a smoother
hole bottom, reduces the requirement for full hole bottom coverage, and allows bits with fewer
rows of teeth and longer teeth to be used. These features can improve penetration rates
somewhat.20 However, experience has shown that roller cone bit penetration rates are sometimes
not as strongly dependent on cutting structure design when drilling with dry gas, mist or foam as
they are when drilling with mud. When drilling underbalanced, selection of the bits cutting
structure may be less critical than it is for a conventionally drilled well.
Abrasive wear rates tend to be higher when drilling with dry gas, mist or foam than they are
when drilling with higher density fluids. This may be partly due to the reduced cooling and
lubrication capacities of low density drilling fluids.
Oilfield roller cone bits are usually designed with cones offset, in order to impart a scraping
motion to the bits teeth. This promotes good penetration rates when drilling with mud at
elevated borehole pressures, but seems to have little impact on penetration rates when drilling
with air.20,21 The increased scraping motion created by cone offset increases a bits susceptibility
to rapid tooth wear. When drilling with dry gas, mist or foam, excessive tooth wear can be
avoided by using bits with no cone offset. Specialized oilfield, air bits have been developed that
combine insert shapes suitable for relatively soft formations with zero cone offset.20 If there is
no cone offset, harder grades of tungsten carbide can be used for the inserts, further reducing
wear rates.
Brannon et al., 1994, reported that there were fewer problems with insert loss or rotation when
air drilling (in comparison to mud drilling). This makes it possible to use shallower insert holes
for air bits. These, in turn, can permit thicker cone shells, larger and more durable bearings, and
closer insert spacing on each row.20

5-40

When drilling with dry gas, mist or foam, it is almost always important to select a bit with good
gauge protection. Gauge wear is often a problem when drilling with dry gas, mist or foam.22 It is
frequently more severe than it would be during drilling the same formations with liquid. Gauge
wear can be particularly rapid if a low fluid injection rate causes inefficient removal of cuttings
from the workfront. Excessive gauge wear, accompanied by rounding of the heel region of the
cones, is often indicative of inadequate bottomhole cleaning.
If a bit is pulled undergauge, the next bit often needs to be reamed to bottom. Reaming can cause
wear of the new bits cutting structure and gauge even before it reaches the hole bottom. It also
imposes inward loads on the bearings that can reduce bearing life. If a new bit is run into an
undergauge hole without sufficient care, it is possible to pinch the bit and force the legs in
towards the bits center. Cone interference can be seen on a severely pinched bit, when teeth on
one cone strike those on the other cones or even the shells of the other cones. This can cause
catastrophic bit failure.
Older roller cone bits, designed for air drilling, had open roller bearings that were cooled by
diverting some of the air flow through the bearing. These bearings had much shorter operating
lives than more modern sealed bearings. Open bearing life can be very short if water inflows or
injected liquids enter the bearings and corrode the bearing surfaces. These bits should not be
considered for mist or foam drilling unless the drilled interval(s) are very short.
Both sealed ball and roller and sealed journal bearing bits are available. If sealed bearings are
used, journal bearings are often preferable to ball and roller bearings. Vibration levels can be
higher when drilling with dry gas, mist or foam than when drilling with mud. Journal bearings
are more tolerant of vibrations. Dry gas, mist and foam drilling is often done in smaller hole
sizes (less than 12-inches diameter), and at relatively low rotary speeds (less than 120 rpm);
both of these situations tend to favor journal bearings.
If wear of the shirt-tails becomes sufficient to expose the seal, gauge wear can contribute to
premature bearing failure. If the flow rate for efficient cuttings removal is inadequate, seal and
bearing failure is greatly accelerated, since this causes the shirt-tails and seals to rotate in a bed of
cuttings.
Seals that have been designed for mud drilling may not be optimal for dry gas, mist or foam
drilling, where the seal temperature can be higher. It has been suggested that larger crosssectional diameter seals, made of high temperature elastomers and installed with lower radial
squeeze, may help to extend the seal life (and therefore the bearings) in bits intended for dry gas,
mist and foam drilling.
Sealed bearing bits are usually more expensive than bits with open bearings. In most oilfield
applications, the reduced number of trips, associated with the longer life of sealed bearings, will
more than offset their additional cost. Regardless, the costs and benefits of using sealed bearing
bits should be evaluated for each application.
5-41

Chapter 5

Well Engineering

Mining bits have been successfully adapted for oilfield drilling with dry gas, mist or foam. These
have open roller bearings and their diameters are sometimes different from conventional oilfield
bits. This makes it possible to drill a tapered hole in situations where high gauge wear occurs;
for example, a 7 7/8-inch oilfield bit could be followed with an 8-inch mining bit, reducing the
amount of reaming required. This would also reduce the risk of pinching the new bit on the trip
to bottom.22 Mining bits also have different diameter tolerances than their oilfield equivalents. In
the past, they would be up to 0.25 inches over the nominal gauge. This may have led to problems
running downhole, if several mining bits were used in the same hole.
Fixed Cutter Bits
Under most circumstances, roller cone bits will be the most effective bit type with dry gas, mist
or foam. It is not advisable to use a PDC bit with these drilling fluids. Since their low thermal
capacities prevent adequate cooling of the PDC cutters, the polycrystalline diamond layer is
likely to overheat and to wear very rapidly.23 There are few, if any, public domain reports of
successfully using PDC bits with dry gas, mist or foam drilling. PDC bit applications are
normally successful when the increase in penetration rate and footage drilled (over roller cone
bits) and any resulting reduction in rig time cost offset the higher cost of the bit. When drilling
with dry gas, mist or foam, the penetration rates with roller cone bits are often so high that there
may be little economic advantage to PDC bits, even if they were capable of operating effectively
in these fluids.
It may be possible to use natural diamond bits with these drilling fluids; cores have been
successfully cut in shallow Arkoma Basin wells with dry air and mist, and in Devonian shales,
with mist. 24
Diamond is not naturally water-wet. Surfactants in injection water can reduce the diamond wear
rate for natural diamond bits. This does increase their penetration rate and overall life.25,26 The
surfactants used in drilling foam may permit cooling of the diamonds to rival that for water.
Natural diamond bits are more likely to be successful with foam than with dry gas. Natural
diamond bits have been successfully used to core with foam. 27,28,29
Bit size may influence the performance of natural diamond bits. Small diamond bits are more
likely to give acceptable footage when drilling with dry gas, mist or foam (in comparison to
larger bits in the same formations). With the smaller bit diameter, the cutting speed is slower for
a given rotary speed, the diamond temperature is reduced and abrasive wear is slowed.
Gasified Liquid and Liquid
Selecting a bit for underbalanced drilling with a gasified liquid or a liquid is not much different
than when drilling overbalanced. The only possible difference may be that the lower bottomhole
pressure may reduce the effective hardness of the target formation; a softer formation bit might
be effective where it would not be if drilling conventionally.

5-42

Gasified liquids and liquids have much better cooling capacities than dry gas, mist and foam.
PDC bits can be used to drill underbalanced with gasified liquids and liquid drilling fluids, if the
formation properties would allow these bits to be used in conventional drilling fluids.

5.8

Underbalanced Perforating

Wells which have been properly cased and cemented under pressure require no extra equipment
for running tubing and packers. Since there is no surface pressure until they are perforated, these
wells are handled in a conventional manner. However, perforating these wells does require an
underbalanced methodology.
Underbalanced perforating methods depend on the pressure conditions in a particular well. In
some cases, a low density completion fluid might be used to provide an underbalanced
hydrostatic pressure at the proposed perforation interval (if sufficient bottomhole pressure is
present in the reservoir). In other cases, a cushion of lower density fluid is placed in the tubing
before perforating. Typically, this involves removing some of the fluid column inside the tubing,
either by swabbing or displacing the completion fluid with a gas, such as nitrogen. The resulting
fluid column inside the tubing exerts lower pressure against the pore pressure in the reservoir,
maintaining underbalanced conditions.
Perforating itself can be done through tubing with a retrievable tubing gun (RTG), a strip-type
perforating charge or a disintegrating link-type charge, such as the Tornado Jet (T-J). A more
powerful tubing conveyed perforating gun (TCP) can be run with the tubing and positioned
below the packer. This type of perforator can be fired by several alternate methods. The most
popular approach is to drop a sinker bar to contact an explosive firing head on the tool. Another
method of activation is to use a timing device with a preset time interval before firing. The firing
mechanism can also be activated using pump pressure. TCP is an excellent approach for
perforating underbalanced. If desired, a tool dropoff assembly can be run, to remove the gun
assembly from the tubing after firing.
The merits of underbalanced perforating, and its requirements in wells that have been drilled and
cased underbalanced are less definitive than for drilling and casing. High overbalanced or
extremely overbalanced (ROPE) techniques, particularly if overpressure is generated by a
nitrogen cushion, may be acceptable or even preferred methods. The primary criteria are
adequate access to the formation, minimization or removal of chemically-related perforation
damage and minimization of mechanical skin or pseudo-skin.
Although tubing-conveyed perforating has been extensively implemented since the early 1970s,
definitive guidelines for the optimum underbalance and surge flow requirements, to achieve
maximum clean-up efficiency, have not been adequately defined. Rules-of-thumb may
recommend anywhere from between one-quarter of a gallon of surge flow per perforation to one
gallon per perforation. The actual amount, however, is very dependent on the number of
perforations actually open to flow. In reality, this may be as low as ten percent of the total
number of shots fired.

5-43

Chapter 5

Well Engineering

A number of recommendations have been made for the optimum underbalance for perforating
both oil and gas wells.
 Gas wells typically require a higher underbalance (differential), due to the more concentrated
amount of shock damage which occurs when perforating formations containing a highly
compressible fluid (Yew and Zhang, 1993).30
 Bell, 1994,31 recommended the following underbalances:
Permeability

Liquid
Reservoir

Gas Reservoir

High
(> 100 md)

200 to
500 psi

1000 to
2000 psi

Low
(< 100 md)

1000 to
2000 psi

2000 to
5000 psi

 King et al., 1986,33 used the productivity increases achieved from post-perforating acidizing
to quantify the effectiveness of underbalanced perforating procedures. In wells where
acidizing had no beneficial impact, it was felt that perforating underbalance was sufficient to
have removed the perforation-induced permeability damage. This underbalance pressure may
not actually have been optimum; but merely that acidizing was ineffective.
 Similarly, Crawford, 1989,32 interpreting the data of King et al., 1986,33 recommended using
permeability as an indicator for predicting the minimum underbalance, Pu(min), necessary to
achieve clean (zero skin) perforations in oil wells:
Pu (min) =

2500
k 0 .3

(5.30)

where:
P......... pressure (psi), and,
k......... permeability (md).
 Regalbuto and Riggs, 1988,34 showed that the final flow rate ratio through the perforations
was:

5-44

58% increased if the perforating was done with a 1000 psi underbalance, and,

50% greater if the perforating was done at balanced conditions and a 1000 psi surge
underbalance was applied after perforating.

In laboratory experiments with 100 md to 300 md Berea sandstone cores, flowing at a 40 psi differential pressure.

 Regalbuto and Riggs, 1988,34 also reported that, in comparison with balanced perforating, the
average perforation volume was about:

56% larger after perforating balanced and later surging with a 1000 psi underbalance,

61% larger after perforating with a 1000 psi underbalance, and,

140% larger after perforating with a 1000 psi underbalance and later surging with a 1000
psi under-balance.

 In other studies, Halleck and Deo, 1989,35 found that underbalance pressures of between 500
and 1000 psi were needed to obtain optimum flow efficiency in Berea sandstone. By
separating the perforation clean-up occurring from transient surge flow effects from that
occurring due to post-shot steady-state flow, they concluded that although continued clean-up
of damaged perforations (perforated at sub-optimal underbalance pressures) did occur, the
perforation flow efficiency did not recover to optimum levels.

5.9

Drillstring Design

Drillstring design for an air-drilled hole is very similar to that for a mud drilled hole. The
drillstring still consists primarily of drillpipe and drill collars. Stabilizers, reamers, jars and
shock subs can still be used in an air hole. There are a few subtle differences. These are
discussed below. Logical decisions on the string configuration can be made for various
underbalanced drilling fluids and configurations. Air drilling is used as an example.
Float Valve
Using a float valve is the primary difference between drilling with air and with fluid. It is not
common to run a float valve when drilling with fluid. A float valve is a requirement when
drilling with air. In an air hole, the drillstring should not be run without a float valve near the
bit. Air in the annulus contains cuttings, making it much more dense than the air inside the
drillstring. When air is vented from the drillstring to make a connection, air and cuttings will Utube into the drillstring from the annulus. As the differential pressures equalize, air will stop
moving and the cuttings will fall to the bottom. Inside the drillstring, the cuttings will settle on
top of the bit and plug the drillstring. The pipe will most likely have to be tripped out of the hole
in order to unplug the drillstring. Installing a float valve above the bit eliminates the possibility
of plugging the drillstring with cuttings while bleeding pressure off the drillstring. While the best
place for the float valve is immediately above the bit, sometimes it may have to be run
immediately above a downhole tool (such as a motor, hammer tool or stabilizer).
Safety is another reason for having a float valve above the bit. The float will prevent formation
gas from venting through the drillstring. While tripping or making a connection, gas will
continue to feed into the wellbore from the formation. Although most of the gas will flow up the
annulus and out the blooie line, some gas may flow to the surface through the drillstring. If a
float valve is present, gas cannot flow up the drillstring and all the gas will be vented through the
blooie line. Gas being vented from the drillstring onto the rig floor can be a safety hazard if it is
present in sufficient quantities.
5-45

Chapter 5

Well Engineering

As shown in Figure 5-9, two common types of float valves (check valves) are used. The flapper
style valve has a spring-loaded flapper that opens when air pressure is applied above it. When
flow stops, the spring closes the flapper. Any pressure below the flapper pushes against the
bottom of the flapper, keeping it closed. A piston (or dart) style float valve works in much the
same way. The flapper is replaced by a spring-loaded piston. When air pressure is applied above
the piston, the valve opens. When flow stops and the pressure differential on the piston reaches
zero, the spring moves the piston up and shuts off flow from below. Any pressure differential
from below the piston will help keep the valve closed. Either float valve works well above the
bit.

Flapper Style

Figure 5-9.

Dart (or Piston) Style

Flapper and piston style float valve.

In addition to the float above the bit, float valves are sometimes installed in the drillstring. When
a float valve is installed in the drillstring, it is commonly called a string float. String floats are
run to reduce the time required to bleed pressure off of the drillstring before making a
connection. As the drillstring gets longer and the capacity increases, it takes longer and longer
for the pressure to bleed off the drillstring during a connection; especially as the standpipe
pressure increases. A string float can be installed near the top of the drillstring in a float sub.
When the pressure is bled off during a connection, the only portion of the drillstring that has to
be depressurized is from the surface to the string float and it will depressurize much more
rapidly. Pressurized air-fluid below the string float will vent through the bit while making the
connection.
String floats can hinder wireline operations such as inclination surveys. The string float can
cause the survey tool to become stuck in the hole. The survey tool is heavy enough to open the
flapper style float and pass through it. When the survey tool is pulled from the hole, the float is
held partially open by the wireline but it is not open enough to pass the tool. The survey tool will
encounter the flapper and push it closed, causing the tool to become stuck. To prevent problems
with surveying, the string float is usually tripped out of the hole, laid down and the drillstring is
5-46

run back to bottom before surveying. After surveying, the string float is reinstalled in the
drillstring near the surface and drilling continues until the next survey. Each time the well is
surveyed, the string float is tripped out of the hole.
Some operators do survey through a string float. To keep the string float open while taking a
survey, the spring is removed from the flapper. When pressures are equalized, the flapper will
fall open and stay in the open position while taking the survey. Since the flapper is always in the
fully open position, the survey tool can generally be pulled back up through the float. There are
times when the survey tool will hang up on the float, but it can usually be worked through by
alternately pulling and slacking off the wireline. Regardless, care must be exercised when
pulling a survey tool through the float. If the survey tool hits the float valve too hard, the
wireline could part. In an air hole, the survey tool would then fall with a very high velocity and
usually would penetrate the bit. The survey tool, bottom float and bit would be ruined and the
drillstring would have to be tripped out of the hole. Survey tools should never be dropped in an
air hole.
Tripping the drillstring to remove the string float does not prevent problems that might occur if
the drillstring becomes stuck. There are times when freepoint tools and backoff shots have to be
run in the hole through the float. The string float can prevent running these tools; and, since the
drillstring is stuck, the string float cannot be tripped out of the hole. For this reason, operators
will use a flapper style string float with the spring removed to prevent problems running
wirelines even if they do not run surveys through the string float. Once pressure bleeds off the
drillstring, the flapper will fall open, allowing wireline tools to be run through the float valve.
The flapper valve still operates properly without the spring. Air rushing past the flapper will
cause it to close when bleeding pressure off of the drillstring.
Some operators drill a small hole in the center of the flapper. If the drillstring becomes stuck by
cuttings packing off in the annulus, the pressure in the drillstring below the flapper cannot bleed
off through the annulus. Wireline tools have to be run to open the flapper valve and equalize the
pressure. At times, the trapped air volume has been sufficient to blow wireline tools up the hole,
causing them to become stuck. In at least one instance, wireline tools have been blown
completely out of the hole when the string float was close to the surface. A small hole in the
flapper allows the pressure to equalize over an extended period of time. This hole must be
sufficiently small to avoid problems with making a connection or tripping pipe.
The piston style float precludes running any wireline tools. It should not be used as a string float.
As can be seen in Figure 5-9, the piston provides an obstruction in the drillstring that wireline
tools cannot pass through, even if the pressure is equalized.
Normally, the float is retained in the string by the pin above it. If there is pressure beneath the
float, it may be forced out of the string when the connection above is broken. This can be
avoided by machining the connection into which it is placed to accept a snap-ring above the
insert, to hold the insert in place. Alternatively, a short joint (pipe or collar, as appropriate) may
be run above the float.11

5-47

Chapter 5

Well Engineering

Downhole Tools
Downhole tools, such as jars and shock subs, can be used in an air drilled hole. It is best to use
mechanical jars rather than hydraulic jars. If the drillstring becomes stuck and the well cannot be
circulated, heat can build up in hydraulic jars, causing them to fail more rapidly. Unlike drilling
fluid, air does not readily conduct heat away from the jars. Excess heat buildup in the hydraulic
fluid can cause seal failures. This does not affect mechanical jars in an air hole. Hydraulic jars
can be run in air holes, but they may not perform as well over an extended period of time.
Hook et al., 1977,36 illustrated some typical bottomhole assemblies used in air drilling operations
(Figure 5-10). The assemblies shown may seem different than those typically used in fluiddrilled wells. This is because the assemblies in Figure 5-10 are used for deviation control. Air
drilling operations are commonly associated with deviation problems. Harder rocks, where air
drilling is particularly applicable, are often characterized by deviation problems, when dipping
bedding is encountered. The assemblies shown in Figure 5-10 are specifically designed to limit
dogleg severity or inclination. Air drilling is also conducted in areas where deviation is not a
problem and slick bottomhole assemblies (no stabilizers or reamers) are used to drill those wells,
just as are used in fluid drilled wells.
In Figure 5-10, the assembly on the extreme left uses a square drill collar to provide stiffness.
This drill collar is essentially a thirty-foot stabilizer. A square drill collar has a much higher
relative stiffness than a round drill collar and will reduce dogleg severity. The outside diameter
of this collar, along the diagonal, is only slightly smaller than the hole diameter; however, the
edges do contact the hole wall and will wear down. When the collar wears, it has to be built back
up to its gauge diameter. This is more expensive than rebuilding a stabilizer and has to be
considered in the economics, to determine if square drill collars should be used. Square drill
collars are not commonly used in air drilling but, they are much more common in air than in fluid
drilling.
The second assembly shown in Figure 5-10 uses reamers and stabilizers to stiffen the bottomhole
assembly. In an airhole, reamers are often used in place of stabilizers. Gauge problems can
occur in harder formations; the reamers are used to keep the hole in gauge. Often, formations
penetrated are hard enough that reamers alone (i.e., no stabilizers) are needed.
The third configuration in Figure 5-10 (on the right) is a pendulum assembly. The assembly
consists of a bit, shock sub, float sub, short drill collar, drill collar, stabilizer and additional drill
collars. This assembly is not stiff and will not minimize dogleg severity. A shock sub is not
required for this assembly. The shock sub and short drill collar can be replaced with a regular
drill collar. The stabilizer can be replaced with a reamer, without changing the bottomhole
assemblys effectiveness. The best position for the float sub would be below rather than above
the shock sub.

5-48

Drillstring Design
In any well, drillstring design starts with the bottomhole assembly. The bottomhole assembly
should be designed so that the top of the assembly remains in tension while drilling with the
maximum anticipated bit weight. A typical design may leave the top ten to fifteen percent of the
drill collars in tension.
There are times when drillpipe is run in compression; in some directional and all horizontal
wells. The drillpipe can be run in compression, provided that the compressive load does not
exceed the critical buckling load of the pipe. However, the critical buckling load in a vertical
well is near zero. Consequently, drillpipe should not be run in compression in a vertical well.
The drillpipe must remain in tension, otherwise it will buckle and may fatigue. Two examples
are provided, summarizing the basic com-ponents of BHA design.

Figure 5-10.

Typical bottomhole assemblies used in air drilling operations (after Hook et al., 1977
). The two assemblies at the left are referred to as packed hole assemblies; they are
designed to minimize angle building tendency. The assembly at the right (pendulum)
can be designed for building or dropping angle, depending on the applied WOB.

36

5-49

Chapter 5

Well Engineering

Example 6
Consider a planned well, where the maximum weight on an 8-inch bit will be 50,000 lbf, the
drill collar size will be 6- inches outside diameter by 2 13/16-inches inside diameter, the
drilling medium will be air and the excess collars should be ten percent to ensure that the
drillpipe remains in tension. Determine the number of thirty-foot drill collars that will be
required.
 The weight per foot of a drill collar can be determined from Equation (5.31).

Wf = 2.67 D p 2 D i 2

(5.31)

Wf = 2.67 (6.52 2.81252 ) = 92 lb / ft


where:
Di ....... inside pipe diameter (inches),
Dp....... outside pipe diameter (inches), and,
Wf ...... weight per foot in air (lb/ft).
 The length of the drill collars can be calculated using Equation (5.32). Since this well is to be
drilled in air, the buoyancy factor is one. It will not be one in other circumstances.
Lc =

W(1 + DF)
Wf B

(5.32)

where:
B ........ buoyancy factor (air = 1),
dimensionless,
DF...... design factor (decimal),
Lc ....... length of the bottomhole assembly
(feet), and,
W ....... bit weight, (lb).
 For a bit weight of 50,000 lb:
Lc =

50000 lb (1 + 010
. )

(92 lb / ft )(1)

= 598 feet

 The number of thirty-foot drill collars would be:


598 ft
= 19.93 or 20 drill collars.
30

5-50

 The total weight, Wtc , of twenty drill collars would be:


Wtc = 598 ft 92 lb / ft = 55,016 lb
To develop 50,000 lb of drilling weight, twenty drill collars are required. The total weight of the
drill collars will be approximately 55,016 lb, including the ten percent design factor.
Drillpipe is usually designed with both a design factor and an overpull. A common design factor
in tension is 1.10 (ten percent). The overpull usually ranges from 50,000 to 100,000 lbf. In
directional wells, it may be higher due to excess hole drag. Example 7 shows how overpull can
be incorporated in the calculations shown in Example 6.
Example 7
Using the data from Example 6, determine the drillstring configuration for a 12,000-foot deep
well. The drillpipe available is 5-inch, 19.50 lb/ft, Grade E and 5-inch, 19.50 lb/ft, Grade G.
The tensile capacity of the Grade E and G pipe are 311,000 lbf and 436,000 lbf respectively. All
the drillpipe is API Premium Class and the tensile strengths can be found in the API RP7G,
available from the American Petroleum Institute. Use a design factor of 1.10 and an overpull of
100,000 lbf.
From Example 6, the collar weight at the bottom of the Grade E pipe will be 55,000 lb. The
maximum pull on the Grade E, with the 1.10 design factor would be:
Pmax =

Tst
DF

(5.33)

where:
DF...... design factor (dimensionless),
Pmax ...... maximum pull on drillpipe (lbf), and,
Tst....... tensile capacity of drillpipe (lbf).
For this example:
Pmax =

311000 lb
= 283,000 lb
110
.

The tensile force that can be applied to five-inch diameter premium (used) drillpipe at minimum yield strength.

5-51

Chapter 5

Well Engineering

The maximum weight, Wmax, of Grade E that can be used with 100,000 lb overpull remaining is:
Wmax = 283000 55000 100000 =
= 128,000 lb
The maximum length, Lmax, that can be used is:
L max =

Wmax
Wf

L max =

128000 lb
= 6564 feet (Grade E)
19.50 lb / ft

(5.34)

The maximum pull, Pmax, on the Grade G, with the 1.10 design factor, would be:
Pmax =

436000 lb
= 396,000 lb
110
.

The maximum weight, Wmax, of Grade G that can be used with 100,000 lbf overpull remaining
is:
Wmax = 396000 55000 100000 128000 =
= 113,000 lb
The maximum length, Lmax, of Grade G drillpipe that can be used is:
L max =

113000
= 5795 feet
19.50

Since the length of Grade G is greater than that necessary to reach the surface, Grade G is
acceptable to the surface. The drillstring would consist of the following:

598 feet of drill collars (refer to Example 6),

6564 feet of 5-inch, 19.50 lb/ft, Grade E drillpipe, and,

4838 feet of 5-inch, 19.50 lb/ft, Grade G drillpipe.

In this example, the maximum force that can be pulled on the drillstring in the event it becomes
stuck is 100,000 lbf over the string weight, once all of the Grade E drillpipe is in the hole. The
weak point will be at the top of the Grade E drillpipe. If the drillstring is changed while fishing,
the new maximum pull must be calculated.
Return to SC Manual

5-52

5-41

Chapter 5

Well Engineering

References

5-42

1.

Guo, B., Hareland, G. and Rajtar, J.: Computer Simulation Predicts Unfavorable Mud
Rate and Optimum Air Injection Rate for Aerated Mud Drilling, paper SPE 26892
presented at the 1993 SPE Eastern Regional Conference and Exhibition, Pittsburgh, PA,
November 2-4.

2.

Govier, G.W. and Aziz, K.: The Flow of Complex Mixtures in Pipes, reprint, Robert E.
Krieger Publishing Company, Malabar, Florida (1982).

3.

Johnson, A.B. and Cooper, S.: Gas Migration Velocities During Gas Kicks in Deviated
Wells, paper SPE 26331 presented at the 1993 SPE Annual Technical Conference and
Exhibition, Houston, TX, October 3-6.

4.

Poettmann, F.H. and Bergman, W.E.: Density of Drilling Muds Reduced by Air
Injection, World Oil (August 1, 1955) 97-100.

5.

Bourgoyne, A.T., Jr.: Rotating Control Head Applications Increasing, Oil & Gas J.
(October 9, 1995).

6.

Cress, L.A., Stone C.R. and Tangedahl, M.: History and Development of a Rotating
Blowout Preventor, paper IADC/SPE 23931 presented at the 1992 IADC/SPE Drilling
Conference, New Orleans.

7.

Cummings, S.G.: Natural Gas Drilling Methods and Practice: San Juan Basin, New
Mexico, paper SPE/IADC 16167 presented at the 1987 SPE/IADC Drilling Converence,
New Orleans, LA.

8.

Dupont, J.: Foam Used to Drill, Gravel-Pack Deep Gas Well, Oil and Gas J. (May 7,
1984) 192-194.

9.

Fraser, I.M. and Moore, R.H.: Guidelines for Stable Foam Drilling Through Permafrost,
paper SPE/IADC 16055 presented at the 1987 SPE/IADC Drilling Conference, New
Orleans, LA.

10.

Curtis, F. and Lunan, B.: Underbalanced Drilling Operations: Correct Operating


Procedures Using a Closed Surface Control System to Drill for Oil and Gas, presented at
the 1995 International Underbalanced Drilling Conference, Amsterdam, The Netherlands,
February 28-March 2.

11.

Saponja, J.: Engineering Considerations for Jointed Pipe Underbalanced Drilling,


presented at the 1995 International Underbalanced Drilling Conference, Amsterdam, The
Netherlands, February 28-March 2.

12.

Hannigan, D.M. and Bourgoyne, A.T., Jr.: Underbalanced Drilling Rotating Control
Head Technology Increasing in Importance, presented at the 1995 International
Underbalanced Drilling Conference, Amsterdam, The Netherlands, February 28-March 2.

13.

Bloys, B., Brown, J.D. and Tarr, B.A.: Drilling Safely and Economically in Carbonates:
Collective Experience of ARCO, BP and MOBIL, paper presented at the 1994 IADC Well
Control Conference for the Asia/Pacific Region, Singapore, December 1-2.

14.

Bradley, W.B. and Fontenot, J.E.: The Prediction and Control of Casing Wear, JPT
(February 1975) 233-243.

15.

Dowell Schlumberger, Cementing Technology, Nova Communications Ltd., London,


(1984).

16.

Coates, G.R. and Denoo, S.A.: Mechanical Properties Program Using Borehole Stress
Analysis and Mohrs Circle, Trans. 1981 SPWLA Annual Logging Symposium.

17.

Mason, K.L.: Tricone Bit Selection Using Sonic Logs, paper SPE 13256 presented at
the1984 59th SPE Annual Technical Conference and Exhibition, Houston, Texas.

18.

Sparr, J., Ledgerwood, L., Goodman, H., Graff, R.L. and Moo, T.J.: Formation
Compressive Strength Estimates for Predicting Drillability and PDC Bit Selection,
SPE/IADC paper presented at the 1995 SPE/IADC Drilling Conference, Amsterdam, The
Netherlands, February 28-March 2.

19.

Fear, M.J., Meany, N.C. and Evans, J.M.: An Expert System for Drill Bit Selection,
SPE/IADC paper 27470 presented at the 1994 SPE/IADC Drilling Conference, Dallas,
Texas, February 15-18.

20.

Brannon, K.C., Grimes, R.E. and Vietmeier, W.R.: New Oilfield Air Bit Improves
Drilling Economics in Appalachian Basin, ASME J. Energy Tech. PD-Vol. 56, Drilling
Technology (1994) 79-87.

21.

Newman, E.F.: Design and Application of Softer Formation Tungsten Carbide Rock
Bits, IADC/SPE paper 11386 presented at the 1983 IADC/SPE Drilling Conference, New
Orleans, LA.

22.

Cooper, L.W., Hook, R.A. and Payne, R.R.: Air Drilling Techniques, paper SPE 6435,
presented at the 1977 SPE Deep Drilling and Production Symposium, Amarillo, Texas.

23.

Glowka, D.A. and Stone, C.M.: Effects of Thermal and Mechanical Loading on PDC Bit
Life, SPEDE (June 1986) 201-213.

5-43

Chapter 5

5-44

Well Engineering

24.

Eaton, N.: Coring the Horizontal Hole, PD-Vol. 27, ASME Drilling Technology
Symposium, Weiner, P.D. and Kastor, R.L. (eds).

25.

Selim, A.A., Schultz, C.W. and Strebig, K.C.: The Effect of Additives on Impregnated
Diamond Bit Performance, paper SPE 2387, SPEJ (December 1969) 425-433.

26.

Unger, H.F., Snowden, B.S. and Engelmann, W.H.: Diamond Drilling with Surfactants in
Upper Michigan Conglomerates Using Surface-Set Bits, paper SPE 4236 presented at the
1973 Conference On Drilling and Rock Mechanics, SPE AIME, Austin, Texas (January
1973).

27.

Hutchinson, S.O.: Stable Foam Lowers Production, Drilling and Remedial Costs,
presented at the 17th Annual Southwestern Petroleum Short Course (April 1970).

28.

Bentsen, N.W. and Veny, J.N.: Preformed Stable Foam Performance in Drilling and
Evaluating Shallow Gas Wells in Alberta, JPT (October 1976) 1237-1240.

29.

Cobbett, J.S.: Application of an Air-Drilling Package in Oman, paper SPE 9600


presented at the 1981 SPE Middle East Oil Technical Conference, Manama, Bahrain.

30.

Yew, C.H. and Zhang, X.: A Study of the Damaged Zone Created by Shape Charge
Perforating, paper SPE 25902 presented at the 1993 Low Permeability Reservoir
Symposium, April 26-28.

31.

Bell, W.T.: Perforating Underbalanced-Evolving Techniques, JPT (Oct. 1984) 16531662.

32.

Crawford, H.R.: Underbalanced Perforating Design, paper SPE 19749 presented at the
1989 SPE Annual Technical Conference and Exhibition, San Antonio, Texas, October 811.

33.

King, G.E., Anderson, A. and Bingham, M.: A Field Study of Underbalance Pressure
Necessary to Obtain A Clean Perforation Using Tubing-Conveyed Perforating, paper SPE
14321, JPT (June 1986) 38, No. 8, 662-664.

34.

Regalbuto, J.A. and Riggs, R.S.: Underbalanced Perforation Characteristics as Affected


by Differential Pressure, SPEPE (February, 1988) 83-88.

35.

Halleck, P.M. and Deo, M.: The Effect of Underbalance on Perforation Flow, JPT
(November 1989) 113-116.

36.

Hook, R.A., Cooper, L.W. and Payne, B.R.: Air, Mist and Foam Drilling: A Look at the
Latest Techniques: Parts I and II, World Oil (April and May, 1977).

6.1

SPECIAL CONSIDERATIONS

Safety in Underbalanced Drilling

Introduction
Underbalanced drilling differs from conventional overbalanced drilling in that flow of reservoir
fluids (brine, oil and gas) into the wellbore is intentionally promoted because of the differential
pressure. Appropriate surface equipment is available to process this returned fluid and to safely
separate hydrocarbons from brine, drilling fluid and cuttings. Since significantly greater
volumes of oil and gas are produced in underbalanced drilling (compared to overbalanced
drilling), and because these products are highly com-bustible, considerable attention must be
paid to safety procedures.
Mudcap drilling is one case where fluids are not produced to surface (Chapter 2, Section 2.9). In
this method, all reservoir fluids are contained below the surface and are not allowed to circulate
back to the surface. In mudcap drilling, drilling fluid, designed to have a smaller density than the
necessary kill weight fluid required by the formation pore pressure, is bullheaded down the
drillstring, while the annulus, which has been pre-loaded with a higher weight, viscosified fluid,
remains shut-in. Safety measures are essential.
For underbalanced drilling, safety procedures are required to isolate potentially explosive and
flammable hydrocarbons and to properly dispose of these liquid or gaseous hydrocarbons.
Onshore, oil or condensate is dispatched to storage facilities for removal while produced gas is
generally flared. Offshore, both oil and gas are usually burned due to space limitations for safe
storage at the drilling site. If a liquids pipeline outlet or oil tanker is available, it is possible to
pump produced oil or condensate and avoid burning these products.
Hydrogen Sulfide Gas Operations
Hydrogen sulfide gas (H2S) is extremely poisonous. It can be present in liquid hydrocarbon
vapors or as a free gas phase. H2S is particularly dangerous because it first attacks a persons
sense of smell and then renders the olfactory system incapable of noticing the distinctive rotten
eggs odor of the gas. H2S gas can kill at very low concentrations. The higher the concentration
of H2S, the more rapidly death occurs. Finally, H2S gas will attack high carbon content (high
strength) metals, causing them to become extremely brittle and fail.

6-1

Chapter 6

Special Considerations

Many states have regulations specifically for drilling in known H2S areas. If an operator
conducts underbalanced drilling operations in a known H2S area, or is drilling in a remote,
untested area, special precautions must be implemented to protect personnel and equipment.
These situations require:
 Providing necessary notice of the proposed operations and hazards.
 Adequate training.
 Special safety equipment, such as H2S sensors, warning alarms, wind direction socks, H2S
concentration measuring devices, portable and fixed air breathing respirators.
 An H2S emergency contingency plan with site specific information and detailed procedures.
 Hydrogen sulfide-resistant materials and training.
 Pressured surface separation vessels and auxiliary vacuum degassing equipment to isolate all
personnel from possible exposure to this poisonous gas.
Flaring Gas
Particularly where very high gas volumes will be produced to the surface, adequately sized flare
lines, leading to properly positioned flare stacks, equipped with automatic flame igniters, are
essential on underbalanced drilling sites. The prevailing wind direction must be taken into
consideration in placing these flare systems on location. Depending on the specific location
requirements, the height of the flare stack may be adjusted for optimum performance. Be certain
that flare lines are adequately anchored.
Separation and Storage
To avoid fire or explosion hazards, liquid hydrocarbon separation and storage facilities, again
with wind direction taken into account, must be positioned remotely. Adequate storage volume,
properly manifolded for transfer or loading, is necessary for safe, uninterrupted under-balanced
drilling.
Training
Personnel training and detailed, written, underbalanced drilling procedures are also required for
safe operations. Redundancy in critical manpower positions and in choke manifold equipment
helps to avoid or accommodate unforeseen emergencies, which can rapidly develop. Location
ingress and egress are critical during underbalanced operations. Emergency, back-up escape
access is very helpful in the event of wind direction changes or blocked primary entrance access.
Gas detection, fire extinguishing and other safety equipment should be placed at strategic
locations on the rig and on the location site.
Downhole Fire
Air, when used to lighten a drilling fluid column for underbalanced drilling, may lead to
equipment corrosion when it is mixed with water, or explosion or fire when contacted with
hydrocarbons. Extreme caution must be exercised when using air in drilling applications.
Extensive experiment-ation has been conducted to determine the maximum safe limits of

6-2

combining conventional air or reduced-oxygen content air with drilling fluids and hydrocarbons,
to avoid combustible mixtures (refer to Chapter 2).
Drilling with Natural Gas
Never underestimate the dangers of surface fires, in any underbalanced drilling operation. This is
particularly true for drilling with natural gas. In the United States, safety conformance on a
drilling rig using natural gas must at least meet the following guidelines.
 American Petroleum Institute (API) RP 500B: Recommended Practice for Classification of
Areas for Electrical Installations at Drilling Rigs and Production Facilities on Land and on
Marine Fixed Mobile Platforms - 1973.
 National Fire Protection Association (NFPA) 70: National Electric Code - 1990.
 NFPA 496: Purged and Pressurized Enclosures for Electrical Equipment in Hazardous
(Classified) Locations - 1988.
The cost savings while drilling with natural gas should offset any additional expenses associated
with regulatory compliance.
Backflow
To prevent flow back up the drillpipe, drillstring floats should be installed. Tools have been
developed to relieve pressure trapped below drillstring floats. This tool is installed above a drill
float and allows a pin to be screwed down to mechanically open the valve and release possible
trapped pressure. A side-port outlet on the tool safely releases any pressure.
Placement of drillstring floats is important for operational and safety reasons. Complete
redundancy of these valves ensures a reliable back-up in the event of a tool failure. The number
of float valves chosen for an underbalanced drilling job depends somewhat on the type of drilling
being used. If the drilling fluids are primarily liquids, most operators minimize the number of
drill floats, sometimes using only a single valve to save rig time and rental expense.
Underbalanced systems, where gas is used to reduce the drilling fluid density, generally use many
more float valves in the drillstring. Two floats should be placed close to the surface to minimize
the time required to bleed off pressure before making a connection. In the middle section of the
drillstring, valves are also important, to minimize downhole fluid separation and to prevent slug
flow. For maximum safety and control, two floats should also be installed in the last stand of the
drillstring assembly. For optimum conditions, a good rule-of-thumb in gas-assisted drilling is
to install a float every twelve joints. Be certain that floats are restrained when connections are
made (refer to Chapter 5, Section 5.9).
Well Control Procedures
Carefully develop contingency plans before the drilling operations start (refer to, for example,
Willis, 19951). The casing program and circulation design, as well as on-location quality control
and monitoring, are particularly important in situations where an underground blowout is
possible.

6-3

Chapter 6

Special Considerations

Equipment
Collins, 1994,2 provided a concise summary of certain important considerations for ensuring safe
operation of rig site equipment. The discussion was designed for flowdrilling but the
components and the philosophy can be adopted elsewhere. Some of the important aspects are:
 Operational and equipment testing procedures must be established, comprehended by all
personnel and enforced.
 Operations should not continue if pressures exceed the maximum limits established.
 In flowdrilling, emphasis is placed on monitoring pressure while drilling, tripping and
stripping, in addition to early kick detection if wells are killed. Like BOP drills, safety, fire
and environmental awareness should be included in routine practice.
 There are many BOP equipment failure causes. The only way to develop testing procedures
[to ensure safe equipment operation] is to prepare detailed BOP and manifolding flow
diagrams that show step-by-step testing for system parts. BOP stacks should be tested when
installed, each time they are reinstalled, once each week and following repair.
 Regularly inspect and monitor surface equipment (i.e. gas monitors, mud/gas separators,
diverter rubber elements and safety equipment).
 Stop flowdrilling when H2S is detected.
 Inspect mud/gas separators daily (check for cuttings plugging), or more frequently in areas
where ROP is high.
 Inspect diverter rubber elements several times a day.
 Check diverter alignment with the rotary table.
 Have developed contingency plans.

6.2

Regulatory Requirements

Introduction
As Eresman, 1993,3 stated, API Committee RP 53 proposed in draft Section 13 that
underbalanced drilling is any drilling operation where an influx of formation fluids is allowed
into the wellbore, circulated out and controlled at the surface. The influx can occur as a result of
severe and uncontrollable lost returns and/or by a conscious decision of the operator to drill with
an influx as a means of enhancing drilling performance.
In the United States, a survey of the primary oil and gas producing states indicated that there
were no special regulations written specifically for underbalanced drilling. In most cases, the
existing regulations could be broadly interpreted to cover underbalanced drilling. For instance,

Blowout Prevention Equipment Systems for Drilling Wells, American Petroleum Institute Recommended Practice
(1984) API RP 53.

6-4

in Louisiana, in reference to drilling fluids, the operator or company is required to use due
diligence in correcting any objectionable conditions. Most states will require adequate blowout
preventers; for underbalanced drilling as well as conventional drilling. In some cases, such as in
California, it may be sufficient to simply convince the State Division Inspector that operations
would be carried out in a prudent and safe manner.
In Texas, the Railroad Commission of Texas specifically recognized the possibility of
underbalanced drilling in clause (E) of Rule 13 (Casing, Cementing, Drilling, and Completion
Requirements-Amended August 13, 1991) of regulation 3.13. Clause (E) states that wells
drilling to formations where the expected reservoir pressure exceeds the weight of the drilling
fluid column shall be equipped to divert any wellbore fluids away from the rig floor. All diverter
systems shall be maintained in an effective working condition. No well shall continue drilling
operations if a test or other information indicates that the diverter system is unable to function or
operate as designed.
In planning an underbalanced well, it is always advisable to check with the local, state or federal
agency governing the wells location, to determine the latest changes in applicable laws or rules
which might apply. For example, in planning to drill a horizontal underbalanced well under the
jurisdiction of federal authorities, the limit on total emissions from the location probably will be
reached more rapidly than for a vertical well, due to the higher expected production and the
length of formation exposed in the lateral. An exception request to this limit, made in advance,
will more likely be approved than one made during or after drilling has started.
Regulations for Underbalanced Drilling (Canada)
In North America, the most detailed regulatory suggestions are outlined in Interim Directive ID
94-3,4 from the Energy Resources Conservation Board (ERCB) in Alberta, Canada. This
Directive covers recommended practices and contains sensible, well considered suggestions by
individuals experienced with underbalanced drilling. This document is available from the
ERCB. Of the recommended practices, the ERCB has mandated strict enforcement for three
areas, pertaining to:
 Blowout prevention system config-urations,
 Tripping procedures, and,
 Well control certification of key per-sonnel.
For the remainder of the practices, the ERCB accepted industry suggestions that these practices
place considerable legal and moral responsibility on operators and contractors involved in
underbalanced drilling to follow the recommended practices or to otherwise provide technically
equivalent or better practices.4

6-5

Chapter 6

Special Considerations

Regulations for Underbalanced Drilling (United Kingdom)


Outside of North America, other countries have developed regulations governing underbalanced
wells. In the United Kingdom, these rules have undergone significant changes over time.
Regulations are well summarized by Moore, 1995.5 The Department of Trade and Industry
(DTI), with authority under the Petroleum Production Regulations, sets specific requirements and
regulations pertinent to the drilling and completion of underbalanced wells. The DTI, in turn,
has delegated this authority to the Health and Safety Executive (HSE) to review operators
applications and detailed well plans, and to grant or deny permits for the proposed work.
Safety of drilling personnel and the general public is the overriding consideration of the HSE.
The work must be carried out in accordance with good oilfield practice, as defined by the HSE.
Safety at work is the primary subject of The Health and Safety at Work etc. Act of 1974 and the
Management of Health and Safety at Work Regulations of 1992, much the same as the
Occupational Safety and Health Act (OSHA) in the United States.
New regulations for the petroleum industry are based on a combination of these existing laws and
a review of safety incidents, notification of well operations, and follow-up inspections and audits.
Additionally, offshore requirements are established by The Offshore Installations (Safety Case)
Regulations of 1992, The Offshore Installations (Prevention of Fire and Explosion, and
Emergency Response) Regulations of 1995 and The Offshore Installations and Wells (Design
and Construction) Regulations of 1996.
These regulations are designed to protect the safety and health of offshore oilfield workers by:
 Identifying potential offshore drilling hazards,
 Properly assessing these risks,
 Ensuring that these risks are as low as reasonably practicable (ALARP),
 Correctly providing notice of proposed well operations,
 Defining the performance standards for all aspects of offshore operations, and,
 Establishing a clearly defined Safety Management System.
Three months prior to the beginning of operations, an operator must submit a safety case to the
HSE for their review. This must contain information identifying all hazards associated with the
proposed operations and ensure that appropriate countermeasures will be in place to handle these
problems and to keep the risks ALARP.
On November 30, 1995, Regulation 11 of the Safety Case Regulations became effective. It
requires offshore operators to notify the HSE twenty-one days before commencing well
operations, to avoid penalty. The Borehole Sites and Operations Regulations, which became
effective October 1, 1995, require onshore operators to follow similar notification procedures. It
is necessary to not only notify the HSE of proposed underbalanced operations, but also to update
them on all operational plans, drilling equipment to be used, procedures to be followed, other

6-6

specific materials required, and environmental considerations. Both the operator and the drilling
contractor must demonstrate that all identifiable risks are ALARP. The British legal system is
very familiar with interpretations of this law. They have established reasonable measures of
weighing an operators cost in money, time and effort to achieve these reductions in risks. If a
proposed operation is economically marginal or so excessively burdened with regulations as to
become intolerable, latitude is generally allowed.
Regulations for Underbalanced Drilling (United States)
In the United States, a survey of the primary oil and gas producing states indicated that there
were no special regulations written specifically for underbalanced drilling. In most cases, the
existing regulations could be broadly interpreted to cover underbalanced drilling.
If regulations are not clear or the burden is placed on the operator to adopt safe and prudent
practices, what should the operator do?
 Research specific regulations of the governing agencies.
 Interact with the relevant agencies to ensure that there will be no future repercussions.
 Adopt reasonable practices. For ex-ample, in Canada, surface BOP systems, as a minimum,
must consist of an ERCB Class III well servicing stack design (Energy Resources
Conservation Board, Alberta Oil and Gas Conservation Regulations Section 8.129 to 8.148
(including schedules 8, 10 and 11) with some type of diverter system, flaring capability, and
the capability to equalize pressure between the diverter line and a point below the lowest ram
type preventer (with some exceptions).
 For example, as Eresman, 1995,3 stated: When a well is being drilled with air, the licensee
shall install and maintain:

in addition to the blowout prevention equipment required in Schedule 8, a rotating head


that diverts the flow during the period the well will be drilled with air,

a diverter line not less than 50 m [164 feet] in length,

a reserve volume of drilling fluid equal to or at least 1.5 times the capacity of the hole,

when drilling formations that may contain hydrogen sulfide, a continuous hydrogen
sulfide monitor on the diverter line, and,

a continuous ignition device at the end of the diverter line.

Consider the following issues:


 Be certain that the BOP stack, with a diverter system:

permits drilling to proceed while controlling annular pressure,

allows connections to be made either with the well flowing or shut-in,

allows tripping of the drillstring under pressure to change bits or bottomhole assemblies,

6-7

Chapter 6

Special Considerations

provides for backup annular control in case of failure of the diverter,

provides for a choke manifold arrangement which allows annular pressure to be varied so
that it will not exceed related working pressure of the equipment,

provides a means to bleed-off pressure or to kill the well, independent of the diverter
system, and,

provides a means to quickly and safely shut-in the well.

 Use string float(s) and fire float(s), if air is used,


 If sour gas is present, drillpipe protection and blind shear rams are needed,
 Kill fluid is needed,
 Casing integrity needs to be guaranteed and full length cementing should be implemented as
regulated,
 Surface equipment spacing needs to adhere to appropriate regulations,
 Flaring must follow appropriate regulations,
 Appropriate separator equipment should be used, as required,
 Provide adequate provision for storage of produced fluids,
 Crews need to be appropriately certified and trained,
 Monitoring and alarms are essential for H2S environments, and,
 Adhere to all safety regulations.

6.3

Environmental Issues

Regulations vary significantly from


state-to-state. Check applicable
regulations carefully.

Land and Water Pollution


Underbalanced drilling provides some environmental benefits but also causes some unique
environmental complications which must be properly handled. For example, produced oil and
natural gas have to be safely and environmentally processed. Oil presents the greatest challenge.
If drill cuttings are oil-coated, surface handling is different than for regular, non-oiled material.
These oil-coated cuttings are commonly jetted to lined containers which are environmentally
disposed of following drilling operations. Designated, approved disposal sites are available for
this purpose. The cuttings are either trucked or barged to these sites.
Closed loop surface facilities provide extra environmental protection over drilling operations
which use earthen reserve and cuttings pits. With some exceptions (i.e. possibly supplementary
chemical additives for flocculation), cleanup and disposal requirements are simplified for closed

6-8

loop systems. Even more significantly, under-balanced drilling inherently involves lower drilling
fluid densities than are used in comparable overpressured drilling. Clear fluid systems are often
used. These fluids may be less expensive and may cause less formation damage if lost
circulation or seepage occurs. Heavier mud systems, in conventional operations, can contain
higher concentrations of heavy metals and other potentially harmful additives, commonly barite.
In Louisiana, the Department of Natural Resources Office of Conservation and the Department of
Environmental Quality both have jurisdiction over the petroleum industry. Statewide Order 29-B
sets limits on the allowable concentrations of nine heavy metals, sometimes found in drilling
fluids. These are arsenic, barium, cadmium, chromium, lead, mercury, selenium, silver and zinc.
The concentration limit for barium is set at 20,000 ppm for wetland areas and 40,000 ppm in
upland areas of the state. When significant amounts of barite are in a liquid mud system, the
fluid becomes contaminated. This mat-erial requires environmental disposal or dilution
treatment on location (land farming) in order to restore the surface locations (as part of cleanup
during post-drilling operations). Soil samples must be taken before and after drilling operations
to ensure proper compliance with these rules. This expense can be minimized or eliminated
entirely by using underbalanced drilling methods and clear drilling fluids.
Heavier drilling fluids also tend to effectively entrain produced oil or condensate, without
breaking out of the mud at the surface. On the other hand, clear fluids separate oil fairly easily.
Some liquid muds can contain as much as thirty percent oil by volume. These oil-saturated
drilling fluids, as well as oil-based mud systems, are an environmental consideration for storage
or disposal, if they cannot be reused on another well.
Since underbalanced drilling incorporates new generation wellhead equipment, environmental
(as well as safety) considerations are improved. The rotating head provides low to medium
pressure containment, in addition to the rigs conventional set of blowout preventers. The RBOP
provides safer, medium to high pressure containment, as redundancy to the regular blowout
preventers. Both systems offer improved benefits over the rig preventers alone.
Surface production handling equipment, including dual choke manifolds, gas/liquid separators,
liquid skimmer and retention tanks and liquid hydrocarbon storage tanks, all serve to provide a
more environmentally-safe operation for underbalanced drilling.
Air Pollution Considerations
Burning of hydrocarbons, produced during underbalanced drilling, can become an environmental
concern. Air pollution, especially when liquid hydrocarbons are burned, can exceed allowed
emissions limits for a particular drill site. This problem can be more severe when the location is
in a populated or smog-prone area. Federal and state limits on total emissions originating from a
particular location must be known in advance, in order to properly plan the well and to comply
with these limits (or request a special waiver for an exception). Exceptions may be granted.
Dust clouds may be a problem during air drilling. Water sprays can quench the dust. With air
drilling, noise pollution may be a special concern near populated areas.

6-9

Chapter 6

Special Considerations

Produced Water Disposal


In addition to hydrocarbons, produced formation water can be a disposal issue. Produced water
can be supplemented by drilling water. This becomes more of an issue with foams. Defoaming
is an important surface procedure. Options are available for disposal. They are all impacted by
specific regulations and associated costs. Disposal operations can include:
 Disposal into surface water drainage systems. This is dependent on volumes, water quality
and supplementary surface processing (such as aeration), regulations and monitoring
facilities.
 Reinjection. Subject to regulations, reinjection into permeable zones in designated disposal
wells may be permitted. It may or may not be required to remove cuttings. Cuttings
reinjection down the annulus may be possible, if pressures are acceptable, and zone
permeability is adequate. Avoiding aquifer contamination and ensuring containment integrity
are primary considerations in permitting disposal wells.
 Approved land disposal sites.
 Overboard offshore disposal. This is highly regulated and it should be anticipated that
regulations will become even more stringent in the future.
 Reserve pits. As indicated by Collins, 1994,2 Reserve pits should be built with integrity to
retain wastes and prevent overflow during rains. In addition, pits and drilling pads should be
designed to collect rig wash, spills and leaks from equipment, stormwater runoff, and should
keep non-drilling pad rainwater from entering pits. Lined pits, barrier walls or closed mud
systems are required in environmentally sensitive areas. Collins, 1994,2 provided further
summary discussions regarding pits, i.e. liners for pits for oil-based muds and brines must be
compatible with these fluids. Operator responsibilities were summarized.
Any pollution resulting from drilling must be a concern to the petroleum industry. Before drilling
the operator must be aware of pertinent regulations.

6.4

Directional Drilling

Introduction
There is no reason why directional or horizontal wells should not be drilled underbalanced. The
design and operational issues are the same as for conventional direction drilling - directional
control, surveying, hole cleaning and drillstring friction. The same downhole equipment can
generally be used as for overbalanced drilling. Many high angle, extended reach and horizontal
wells have been drilled with gasified liquids, with the attendant advantages.
Drilling high angle, extended reach or horizontal wells with air as the circulating medium is not
yet common practice; however, air has some distinct advantages over drilling mud. These
include:
 Penetration rates are significantly increased, leading to shorter drilling times and lower costs,
 Lost circulation problems are eliminated (or moderated) especially in areas with very low

6-10

bottomhole pressures,
 Drillstem testing of potential producing formations is continuously possible, and,
 Formation damage is minimized.
Along with the advantages of underbalanced operations, using compressible drilling fluids can
complicate direction drilling. Because compressible fluids may pose more operational problems
than liquids, this section concentrates on their use in directional drilling. Some of the problems
that could be expected are:
 Conventional downhole motor life is shorter and conventional motors do not operate as
efficiently,
 Conventional Measurement While Drilling (MWD) systems do not work in an environment
with compressible fluid,
 Hole cleaning can be a problem at inclinations above 50, with dry gas, mist and foam,
 The horizontal section length is reduced because of the increased drillstring friction (drag),
and,
 Not all formations and lithologies are suitable for drilling with dry gas, mist or foam.
However, horizontal drilling with air, mist and foam systems can be, and has been, successful.
With careful planning, all of these problems can be minimized or eliminated.
Horizontal wells have been successfully drilled with air or foam since 1986. At a minimum,
operators are drilling the horizontal section with air or foam to eliminate lost circulation
problems in low pressure (partially depleted) reservoirs, and to reduce formation damage due to
drilling fluid invasion. Problems have been en-countered in drilling these horizontal wells. Not
all of these problems are unique to air drilling. They are sometimes exaggerated by the
conditions in an air hole. By changing the conventional method of operations in an air hole and
anticipating potential problems, air drilling can be cost-competitive and possibly even less
expensive than conventional drilling.
Bottomhole Assemblies
The United States DOE has led efforts in horizontal air drilling.6 The main issues for directional
drilling underbalanced are similar to those for conventional directional drilling:
 Directional control,
 Surveying,
 Hole cleaning, and,
 Drillstring friction.

6-11

Chapter 6

Special Considerations

Forces Acting at the Bit


A bottomhole assembly is the arrangement of the bit, stabilizer, reamers, drill collars, subs and
special tools, used at the bottom of the drillstring. Anything that is run in the hole to drill, ream
or circulate is a bottomhole assembly. Hole inclination can be controlled by adjusting the
bottomhole assembly stabilization, exactly as it is in rotary drilling with conventional fluids. The
simplest assembly is a bit, collars and drillpipe. This is often called a slick assembly. Use of this
assembly in directional drilling is very limited and is usually confined to the vertical section of
the hole, where deviation is not a problem.
Bottomhole assemblies will deviate a hole. To understand why this happens, consider the slick
assembly. It is the simplest and easiest to understand. The deviation tendency in this assembly is
a result of the flexiblility of the drill collars and the forces which act on the assembly, causing the
collars to bend. Even though drill collars seem to be very rigid, they will bend enough to cause
deviation.
The point at which the collars contact the low side of the hole is called the tangency point (Figure
6-1). The distance, L, from the bit to the tangency point, depends on the collar size, the hole size,
the applied bit weight, the hole inclination, and the hole curvature. Generally, L is less than 150
feet. Above the tangency point of the slick assembly, the remainder of the drillstring has no effect
on deviation. As weight is applied to the bit, the tangency point will move closer to the bit.
Because of the bending of the drill collars, the resultant force applied to the formation is not in
the direction of the hole axis. It is in the direction of the drill collar axis. This is shown in
Figure 6-2. The angle between the hole and the collar axis is denoted as . As bit weight is
applied, the tangency point moves toward the bit, increasing the angle . An increase in bit
weight leads to an increase in and a resulting increase in deviation tendency.
Unfortunately, the resultant force incorporates additional components (Figure 6-3). The primary
component is the drilling force in line with the axis of the borehole. In addition, F B is the bit side
force caused by bending of the collars, acting perpendicular to the axis of the borehole. FP is the
force due to gravity (acting on the unsupported section of drill collars) resolved in the opposite
direction to and counteracting FB. The net deviation force is then equal to the summation of FB
and FP. Ideally, if FP is greater than FB, the hole angle will drop. If FP is equal to FB, the hole
angle will remain the same; and, if FP is less than FB, the hole angle will increase.
The deviation tendency can be controlled by changing the bit weight. Increasing the bit weight
will lower the tangency point, increasing the angle . Since FB is proportional to sin, an
increase in bit weight increases the bit side force and ultimately the deviation tendency.
Conversely, a decrease in bit weight will decrease the deviation tendency.

6-12

Figure 6-1.

Schematic representation of a bottomhole assembly in an inclined hole.

Figure 6-2.

The resultant force applied to the formation is not in the direction of the hole axis.
The force is in the direction of the drill collar axis because of bending of the drill
collars.

6-13

Chapter 6

Figure 6-3.

Special Considerations

Schematic showing deviating and restoring force components.

This may all sound significant, but the deviation force for a slick bottomhole assembly is very
small. Figures 6-4
6-4 through 6-8 show the calculated resultant force for a slick assembly under
various conditions. In most instances, a slick assembly will have a negative resultant force
(dropping tendency) at inclinations above 2.
Drill Collars
The stiffness of the drill collars also affects the deviation tendency. Stiffer collars will bend less.
This increases the length, L, to the tangency point. If the tangency point moves up the hole, the
deviation tendency will be reduced. The relative stiffness of a drill collar is proportional to the
collar radius to the fourth power. As an example, assume that the relative stiffness of a six-inch
OD drill collar is one. An eight-inch or a ten-inch OD collar would be approximately three or
eight times stiffer, respectively (Table 6-1). Therefore, small OD collars (with less stiffness)
increase the tendency for deviation.

6-14

Table 6-1.

Relative stiffness of various drill collars.

Collar Diameter
(inches)

Relative Stiffness

12

16

10

0.2

100
80

Resultant Force (lbf)

60
40
20
0
-20

1 o Inclination

-40
2 o Inclination

-60
-80
-100

3 o Inclination

-120
0

10000

20000

30000

40000

50000

60000

70000

Bit Weight (lbf)

Figure 6-4.

Resultant force for a slick assembly in a 7 7/8-inch diameter hole.

6-15

Chapter 6

Special Considerations

50
40

Resultant Force (lbf)

30
20
10

2 o Inclination

0
o
-10 1 Inclination

-20
-30
-40
3 o Inclination

-50
0

5000

10000

15000

20000

25000

30000

35000

Bit Weight (lbf)

Figure 6-5.

Resultant force for an assembly with 4-inch drill collars in a 6-inch hole. Note the
reduction in the resultant force, in comparison to Figure 6-4.

250
200

Resultant Force (lbf)

150

12 1/4-inch hole

100
9-7/8 inch hole

50
0

7-7/8 inch hole

-50
-100
-150
0

10000

20000

30000

40000

50000

60000

Bit Weight (lbf)

Figure 6-6.

6-16

Resultant force in a hole, inclined at 2, drilled with 6-inch collars. The resultant
force is shown as a function of bit weight and hole size.

100

50

Resultant Force (lbf)

12 1/4-inch hole

-50
9-7/8 inch hole

-100

-150

-200
0

10000

20000

30000

40000

50000

60000

70000

Bit Weight (lbf)


Figure 6-7.

Resultant force in a hole, inclined at 2, drilled with 8-inch collars. The resultant force is shown as a
function of bit weight and hole size. Note the influence of the larger diameter drill collars.

Resultant Force (lbf)

-50
12 1/4-inch hole

-100
9 7/8-inch hole

-150

-200

-250
0

10000

20000

30000

40000

50000

60000

70000

Bit Weight (lbf)


Figure 6-8.

Resultant force in a hole, inclined at 2, drilled with 9-inch collars. The resultant force is shown as a
function of bit weight and hole size.

6-17

Chapter 6

Special Considerations

Stabilizers
Adding a stabilizer above the bit significantly affects the deviation tendency of a bottomhole
assembly. The stabilizer acts as a fulcrum, around which the unsupported section of the
bottomhole assembly reacts. The addition of the moment arm between the bit and stabilizer
increases the bit side force. In fact, a single stabilizer assembly is a very strong building
assembly.
Addition of multiple stabilizers to an assembly makes determination of side forces at the bit
much more complicated. Analyzing these types of bottomhole assemblies is beyond the scope of
this manual.
Formation Anisotropy and Dynamic Conditions
Assuming that a formation is uniform and that the bit can drill in any direction, the bottomhole
assembly would drill in the direction of the vectorial sum of the forces at the bit. Unfortunately,
a bit's side cutting and forward cutting ability are not equal. Also, anisotropic failure of the rock
can cause deviation in a direction other than the simple vectorial sum of the forces at the bit.
The side cutting ability of a bit is proportional to the side force exerted at the bit (the vectorial
sum of FB and FP). Under static conditions, the side force on the bit can be calculated using
computer modeling. When the entire bottomhole assembly is considered, it can be shown that the
stabilizers in the assembly also exert a side force. Therefore, stabilizers also have a side cutting
ability. It would seem that the deviation tendency could then be calculated. However, the side
forces will change under dynamic conditions. Both the bit and the stabilizers cut sideways,
reducing the side force on each, until equilibrium is reached.
Under dynamic conditions, the relative side cutting of the bit and stabilizers becomes
complicated. This makes it very difficult to calculate the deviation tendency. The relationship
between the bit and stabilizer side cutting depends on the type of bit, the type of stabilizer, the
penetration rate, the rotary speed, the lithology, the hole size, and the bottomhole assembly type.
Amoco test data, published by Millheim and Warren, 1978,7 proved that both the bit and
stabilizers will cut laterally under dynamic conditions. Figure 6-9 shows the results of some of
those tests. The insert at the top shows the influence of the magnitude of the side force on a bit.
The rate of side cutting decreased after this force was first applied, until an equilibrium was
reached; after this, the rate of displacement remained constant. The slope of the line would be the
side cutting ability of the bit under ideal conditions. A significant difference in side cutting was
caused by a small change in the magnitude of the side force.
6-9 demonstrates how the side cutting ability of a bit changes with
The middle insert (in Figure 6-9),
the penetration rate. The cutting rate was lower at higher penetration rates. The bottom insert,
shows the influence of side cutting, when drilling with a stabilizer and a constant side force.

6-18

Note that with 1,500 lbf side force, the blade stabilizer cut laterally at approximately the same
rate as the bit with 800 lbf side force.
These tests indicated that even though the side cutting ability of a bit is small compared to the
forward cutting ability, it is enough to drill in the direction of the vectorial sum of the forces at
the bit under dynamic conditions. Since the angle, , is usually less than 1 in normal drilling
operations, an operator could expect horizontal displacements similar to those in the Amoco
tests. Also, the side cutting ability of soft formation bits is generally considered to be better than
for hard formation bits. Diamond bits have a greater side cutting ability because they are
designed with more of a cutting structure along the lateral face of the bit.
An additional factor affecting deviation tendency is formation anisotropy. In isotropic
formations, equal chip volumes are formed on each side of the bit tooth and the bit will drill
straight ahead (Figure 6-10). However, bedding planes and oriented structural weaknesses make
most rocks anisotropic. In such an anisotropic situation, relatively large chip volumes are formed
on one side of the bit tooth, causing the bit to deviate (Figure 6-10).
The magnitude and direction of deviation, attributable to formation anisotropy, depends on the
bedding dip. Generally, a bit will walk updip when beds are dipping from 0 to 45 and downdip
when beds are dipping from 65 to 90. Bedding plane dip between 45 and 65 can cause either
an updip or downdip walk. Bedding strike can cause the bit to walk left or right.
Basic Assemblies
There are three basic types of assemblies used in directional drilling. These are:

building assemblies,

dropping assemblies, and,

holding assemblies.
Bit Side Force Test Data - Variable Side Force
0
9-7/8 inch Series 1-1-1 Bit
Carthage Marble
100 rpm
40 ft/hr

0.5

Depth (feet)

1.5

2
1000 lbf

800 lbf

2.5

3
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Horizontal Displacement (inches)

6-19

Chapter 6

Special Considerations

Bit Side Force Test Data - Variable Penetration Rate


0
9-7/8 inch Series 1-1-1 Bit
Bedford Limestone
100 rpm
800 lbf Side Force

0.5

Depth (feet)

1
100 ft/hr

1.5

2.5

150 ft/hr

3
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Horizontal Displacement (inches)

Stabilizer Side Force Test Data


0
9-7/8 inch Spiral Blade Stabilizer
Bedford Limestone
100 rpm
100 ft/hr
1500 lbf

0.5

Depth (feet)

1.5

2.5

3
0

0.1

0.2

0.3

0.4

0.5

0.6

Horizontal Displacement (inches)

Figure 6-9.

6-20

Results of Amoco tests assessing the side cutting ability of bits and stabilizers (refer to
Millheim and Warren, 1978 7).

ISOTROPIC
Equal Chip
Volume Formation

ANISOTROPIC
Unequal Chip
Volume Formation

Figure 6-10.

Chip volume formation in isotropic and anisotropic rock (refer for example, to
Millheim, 1979 8).

A building assembly is intended to increase hole inclination; a dropping assembly is intended to


decrease hole inclination; and, a holding assembly is intended to maintain hole inclination. It
should be noted that a building assembly may not always build angle. Formation characteristics
may cause the assembly to drop or to hold angle. The building assembly is intended to build
angle. The same is true for the dropping and holding assemblies.
Building Assemblies
As indicated, the building assembly uses a stabilizer, acting as a fulcrum, to apply a side force to
the bit. The magnitude of this force is a function of the distance from the bit to the tangency
point. An increase in bit weight and/or a decrease in drill collar stiffness will increase the side
force at the bit, increasing the rate of build.

6-21

Chapter 6

Special Considerations

The strongest building assembly consists of one stabilizer, placed 3 to 6 feet above the bit face,
with collars and drillpipe above the stabilizer. This assembly will build in the majority of
situations. The rate of the build will be controlled by formation character-istics, the bit and
stabilizer types, the lithology, the bit weight, the drill collar stiffnesses, the drillstring rpms, the
pene-tration rate, and the hole geometry.
Another strong to moderate building assembly consists of a bottomhole stabilizer, placed 3 to 6
feet from the bit face, 60 feet of collars, a second stabilizer, collars, and drillpipe. This is the
most common assembly used to build angle. The second stabilizer tends to dampen the building
tendency. This assembly can be used when the previous assembly builds at an excessive rate.
Other building assemblies are shown in Figure 6-11.
Dropping Assemblies
A dropping assembly is sometimes referred to as a pendulum assembly. Wilson, 1980,9 provided
a good discussion, relevant to air drilling. In this assembly, a stabilizer is placed at 30, 45 or 60
feet from the bit. The stabilizer produces a plumb-bob or pendulum effect; hence the name
pendulum assembly. The purpose of the stabilizer is to prevent the collar from touching the wall
of the hole, causing a tangency point.
An increase in the effective length of the bottomhole assembly (the length below the tangency
point) results in an increase in the weight. Since the force, FP, is determined by that weight, it is
also increased, exceeding the force, FB, due to bending. The net result is a side force on the bit,
causing the hole to drop angle.
Addition of bit weight will decrease the dropping tendency of this assembly because it increases
the force due to bending, FB. If enough bit weight is applied to the assembly to cause the collars
to contact the borehole wall (between the stabilizer and the bit), the assembly will act as a slick
assembly. Only the section of the assembly below the tangency point affects the bit side force.
If an increase in dropping tendency is required, larger diameter or heavier collars should be used
below the stabilizer. This increases the weight of the assembly and causes an increase in the
dropping tendency. As an example, suppose that a dropping assembly, with seven-inch OD
collars, is being used in a 12-inch hole. By substituting nine-inch OD collars for the seven-inch
OD collars, an increase in dropping tendency can be achieved.
Dropping assemblies will have a higher rate of drop as hole inclination, I, increases. The force,
FP, (which causes the dropping tendency) is calculated using the following equation:
FP = 0.5W sin I
where:
W ....... buoyant weight of the collars
below the stabilizer (lbf), and,
I.......... inclination angle (degrees).

6-22

(6.1)

An increase in hole angle results in an increase in FP, resulting in an increase in dropping


tendency. Additional dropping assemblies are shown in Figure
Figure 6-12.
6-12
Holding Assemblies
Holding the inclination in a hole is much more difficult than building or dropping angle. Under
ideal conditions, most assemblies either have a building or dropping tendency. Most straight
hole sections of a directional well will have alternating build and drop sections. When holding
inclination, these build and drop sections should be minimized and spread out over a large
interval.

Strength

Bottomhole Assembly
90 feet

High

High

60 feet

30 feet

High
60 feet

High
To
Medium

45 feet

30 feet

Medium
To
Low

Bit

Figure 6-11.

Stabilizer

Collar

Schematics of various building assemblies (refer for example, to Millheim and


Apostal, 1980 10).

6-23

Chapter 6

Special Considerations

Strength

Bottomhole Assembly
60 feet

High

60 feet

30 feet

High

45 feet
Medium

30 feet
Low
This Assembly Can Build Angle Under Certain Circumstances.

Bit

Figure 6-12.

Stabilizer

Collar

Schematics of various dropping assemblies (refer for example, to Millheim and


Apostal, 1980 10).

Amoco has statistically compared the performance of various holding assemblies. Figure 6-13
6-13
shows three of the most commonly used holding assemblies. The assembly at the top proved to
be the most successful, even though it maintained inclination only 60 percent of the time. The
assembly in the middle maintained inclination less than 50 percent of the time, and the bottom
assembly was even less effective.
When selecting a holding assembly, research the well records in the area, to assess which
assembly works best for the formations being drilled. If no formation-specific data are available,
use the assembly at the top of Figure 6-13 and adjust it as necessary.

6-24

Downhole Motors
Downhole motors play a major role in drilling directional and horizontal wells. Kicking off from
vertical requires a downhole motor with a bent sub or bent housing. All directional wells require
steering during initial kickoff, correction runs, sidetracks and redrills. High build rates (greater
than say 2 or 3/100 ft) also require a motor and bent housing.
The increasing availability of steerable systems has dramatically changed the number of wells
that are drilled directionally and horizontally. These systems include a downhole motor (or
turbine) fitted with a bent sub, as well as some means of surveying while drilling. Orienting the
motor and drilling with no string rotation allows the bit to drill (more or less) in the direction in
which it is pointed, providing the ability to change inclination and azimuth, as required.
Adjusting the bent housing angle controls the build rate.

Strength

Bottomhole Assembly
15 - 20 feet
30 feet

30, 60 or 90 feet

Medium
5 - 15 feet
30 feet

30 feet

Medium

30 - 40 feet

30 or 60 feet

30 or 60 feet

Low

Bit

Figure 6-13.

Stabilizer

Collar

Schematics of various holding assemblies (refer for example, to Millheim and Apostal,
10
1980 ).

Rotating the drillstring, when drilling with a motor and bent sub, allows the bit to drill ahead at
more or less constant angle.
Since drillstring friction can restrict weight transfer if the string is not rotating, the penetration
rate tends to be lower when drilling in oriented mode as opposed to rotating the string. Using a
bent housing means that the hole is drilled slightly over-gauge, whenever the pipe is rotated.
Rotating the pipe while drilling with a motor and a high angle bent sub can lead to tracking, poor
penetration rate and rapid off-center wear with a roller cone bit. These same problems are

6-25

Chapter 6

Special Considerations

experienced whether the well is drilled under- or overbalanced; with a compressible or


incompressible fluid. However, drillstring friction can be a more significant issue if drilling
underbalanced with a compressible fluid.
Conventional positive displacement mud motors can be run on compressible fluids. As indicated
by Shale, 1991,11 there are major disadvantages to doing this.
 Conventional motors are designed to run with low volumetric flow rates and high pressure
drops. This leads to high inlet pressures and low efficiency when the motor is run on
compressible fluids such as air, mist or foam.
 With compressible fluids, these motors are prone to stalling and can be difficult to restart
after connections.
 The high inlet pressure means that there is a large amount of energy stored in the compressed
fluid in the drillpipe above the motor. If the bit is pulled off bottom so that the resistance to
rotation is removed, the motor can speed up rapidly and be damaged by the overspeed, before
the energy stored in the drillpipe is dissipated. The high motor pressure drop often requires
that a booster is run in the compression system.
 Typically, the air volume required to clean the hole is three times greater than the
recommended flow rate for the conventional mud motor.11 There is a way to reduce the
flow through a motor. Some of the air can be diverted before passing through the motor
section. This can be accomplished by placing a jet sub above the motor. This will allow
some of the air to escape from the drillstring without passing through the motor. In motors
with a hollow rotor, the bypass valve can be replaced by an orifice. In each case, the orifice
is a predetermined size and diverts the necessary air to the annulus without it passing through
the motor section. Jets should also be placed in the bit to provide adequate bottomhole
cleaning and extra cooling for the motor. The jets can be designed for bottomhole cleaning,
using the method presented by Lyons, 1984,12 although, a pressure drop of 200 to 300 psi is
usually sufficient. Expansion of the air through the bit nozzles provides the cooling. Friction
between the rotor and the stator within the motor causes heat to build up. Cooling at the
bottom of the hole helps the motor to run longer.
 Mud motors are hydrostatic, they can use only the displacement work, and not the expansion
work of the compressed air.13
Downhole motors have been developed to run on compressible fluids. These are designed to run
at higher volumetric rates and lower pressure drops than conventional mud motors. The
advantages of these new generation air motors are:
 Boosters are not needed,
 Efficiency is improved,
 These motors do not stall as easily as conventional motors,
 Overspeed is less likely, and

6-26

 They can be used with compressible (dry and mist) and slightly compressible fluids (gasified
liquids, liquids).
Shale, 1991,11 provided a summary of the flow rate and pressure drop calculations required to
assess motor performance. He showed power outputs for one air motor type. The power outputs
of these air motors are typically 1.5 to 2 times those of similar size mud motors. As a result of
the higher efficiency, these motors give higher penetration rates than conventional mud motors
run under the same conditions. They are capable of penetration rates that are essentially the same
as those for rotary drilling with the same fluid.
These motors are recommended for directional drilling with compressible fluids.
This type of improved performance depends on proper lubrication. Mist or foam can provide
lubrication. When water (included in the mist or foam) is introduced into the wellbore, enough
must be used to completely wet the borehole and the generated cuttings. Otherwise, a mud ring
will form and the drillstring will become stuck. As a rule, a minimum rate of 10 BWPH is used
even though the motor does not require this volume for lubrication.
Injection of water into a wellbore sometimes causes shale stability problems and a dry hole may
be desirable. Injecting a small quantity of oil into the air stream can also provide effective motor
lubrication. 0.15 to 0.25 gallons of motor oil per hour should be added to dry air (Shale, 199111).
Foamer, graphite powder, and bentonite have also been used successfully. Injection rates of 5
gallons/hour will provide ample lubrication. Too much oil will cause the drill cuttings to become
slightly wet; this can stick the drillstring. Therefore, oil injection rates should be limited.
Dreiling, et al., 1996,14 provided some field data.
Because dry air has an extremely deleterious effect on motor and steering probe
performance, a compromise using an air/mist mixture was reached. Initially, a mixture
of 2-3 percent potassium chloride with small amounts of a dual ionic polymer and
inhibitor plus surfactant as required to clean the hole was utilized. This mixture was
injected at 5-15 gpm into the 1,500 to 2,500 cfm of air. Although the polymer was not
required to maintain hole stability unless the Mancos Shale was exposed, it performed
extremely well in decreasing torque and drag.
It was later determined that the lobe configuration should not be the predominant factor
for motor selection; rather that the number of stages has a greater effect. With lesser
stages, the air undergoes fewer compression/decompression cycles as it is forced through
the motor, resulting in less vibration. Various combinations of bit jetting and flex nonmagnetic drill collars were also tried experimentally to dampen the vibration amplitude
with poor success. Jetting the bit did help but reduced the motors power output.

6-27

Chapter 6

Special Considerations

Lyons et al., 1988,15 outlined the principles and performance of a prototype air turbine drill.
Numerical simulations indicated the feasibility of using a turbine motor to convert the internal
energy in a gas stream to mechanical energy at the bit. Two field tests have been performed.
The drilling rates obtained compared favorably with those from conventional air drilling.
Surveying
MWD
When drilling directionally, it is essential to know the hole inclination and azimuth. In
conventional directional drilling, this is normally accomplished with an MWD (MeasurementWhile-Drilling) system in the BHA. This periodically transmits magnet-ometer and
accelerometer measurements uphole, from which inclination and azimuth (or toolface
orientation) can be calculated. Data transmission normally involves pulsing the mud flow, to
create encoded pressure pulses in the drilling fluid in the drillpipe. These pulses are detected at
the standpipe. Mud pulse telemetry does not work with compressible fluids - the signal is
attenuated too rapidly for it to reach the surface. Conventional mud pulse MWD systems can
only be used in underbalanced drilling if the drilling fluid inside the drillstring is incompressible.
This will be the case in flowdrilling, when drilling with a parasite string or when using temporary
casing gasification.
Roy and Hay, 1995,16 discussed some pressure measurement systems. The Pressure-WhileDrilling system measures the annulus pressure by means of a pressure sensor that is located in a
special sub with a port to the annulus located below the MWD. The pressure data are either
recorded down-hole or transmitted in real-time to the surface using MWD transmission.16
EMWD
One option with compressible fluids is electromagnetic MWD (EMWD). This
electromagnetically transmits a signal through the formation (the casing and/or drillstring will
serve as a wave guide). It requires adequate formation conductivity for the signal to reach the
surface. Transmission can be interrupted by high resistivity formations. The transmission range
can be limited even when the formations are suitable. More details on these systems are given by
Soulier and Lemaitre, 1993.17 In the past, reliability has been an issue with EMWD systems.
Recent reports are encouraging, although vibration levels downhole can be particularly severe
when drilling with dry air, mist or foam. Shale, 1994,13 indicated that vibrational damage and
inadequate signal transmission may become problematic at depths below 4000 to 5000 feet.
Inherent limitations relate to signal attenuation.
 Younger, uncompacted rocks with high apparent porosity are characterized by extremely high
attenuation.
 There is no difference of attenuation in cased hole sections for boreholes filled with air or
water-based muds but a noticeable reduction of attenuation has been observed in the
openhole zone.17

6-28

EMWD systems have been used in air holes with mixed results. Signals do get back to the
surface and the information is correct. Reliability has been the main problem with EMWD.
Frequent failures have been experienced in air holes. Drilling conditions in an air hole are more
extreme than those experienced in a mud-filled hole. There is no fluid in the hole to dampen
vibration. EMWDs are not yet durable enough to work consistently in an air hole; but with
experience and improvements, EMWD continues to improve.
Steering Tools
In most air wells, a steering tool is used to get survey and tool face data while drilling with a
downhole motor. The steering tool is different from EMWD in that information is sent back to
the surface using a single conductor wireline rather than electromagnetic waves. The system is
cumbersome and more time consuming because the wireline extends back up through the
drillstring to the surface. Depending on the system used, the wireline must be removed in order
to add more drillpipe as drilling progresses. This adds time and expense to the drilling operation.
Until recently, the drillstring could not be rotated with the steering tool in the hole. Therefore, a
steerable system could not be used with a steering tool. A steerable system requires that the
drillstring is rotated for at least a portion of the drilling process. In the past few years, some
service companies have developed what is called a quick- or wet-connect. The tool allows the
drillstring to be rotated with the steering tool still in the bottomhole assembly. A sub is run in
the vertical portion of the well and the wireline is connected from the steering tool to the sub.
The wireline is connected into the top of the sub when a survey is taken or when drilling without
rotation. The wireline is disconnected from the sub and pulled into a custom swivel when the
pipe is rotated. Now operators can run a steerable system in air.
Currently, the motors are oriented with a steering tool that sends information to the surface using
a single conductor wireline. Some steering tools have the same problem as EMWD. Vibrations
in the well cause frequent failures. In fact, all steering tools will experience more failures in an
air hole. Selection of the right steering tool is essential. Placing jets in the bit also reduces the
vibration experienced by the steering tool.
There are two methods that have been used to survey the horizontal section. Both involve
tripping the drillstring. The first method uses an electric line with a side entry sub. The
drillstring is pulled from the hole until the bit is at an inclination of 70. A side entry sub is
required along with a latch-in assembly for the steering tool. Above an inclination of 70, the
steering tool will no longer fall in the hole and cannot be pumped down with air. Unlike mud,
the air passing by the steering tool does not generate enough drag to carry a heavy tool down the
hole. The steering tool must be installed through a side entry sub above 70 and tripped to
bottom with the drillstring. A side entry sub is installed and the survey tool (single-shot or
steering tool) is run to the bit on an electric line. Then, the drillstring is tripped back to bottom
with the remainder of the wire on the outside of the drillstring. After reaching TD, a survey is
taken and the drillstring is tripped back out of the hole to the side entry sub. The survey tool is

6-29

Chapter 6

Special Considerations

removed and the drillstring is run back to bottom to continue drilling. As is evident, surveying
the horizontal section is time consuming and expensive.
An alternative method has been used to reduce the time and associated costs for surveying.
Tripping the drillpipe to 70 is still required; but an electric line and side entry sub are not used.
A single-shot (a steering tool cannot be used) is run on slickline with a releasing overshot. When
the single-shot enters the monel collars, a monel sensor activates the releasing overshot,
disconnecting the single-shot from the slick line. The slickline is removed from the hole and the
drillstring tripped to bottom. The survey is taken and the pipe is tripped back to 70. The singleshot is retrieved by using a standard overshot on the slickline. Costs are reduced because tripping
is much quicker without an electric line on the outside of the pipe, and the cost of the electric line
is eliminated. The cost of the releasing overshot is only a fraction of the electric line costs.
Without an MWD system for air holes, it is much more difficult to use a steerable system in the
horizontal section. The steerable system has to be oriented by a steering tool, and the drillstring
must be tripped to 70 in order to install the steering tool. The cost savings associated with
steerable systems are derived from reducing the amount of tripping necessary to drill the well. In
an air hole, tripping is required anyway, so steerable systems are frequently not cost effective,
especially since motor life is shorter and less predictable. Ordinarily, rotary assemblies are used
to drill the horizontal sections using motor corrections as necessary.
Other alternatives to EMWD include using either a continuous wire cartridge link to a steering
tool that stays downhole with the drilling BHA, or a wet-connect to the steering tool. The former
is described by Shale and Moberley, 1992,18 while the latter is described by Shale and Curry,
1993.19 The cartridge system is similar to the wet-connect but, instead of having a single wetconnect, the wireline is connected back to the surface with a series of disposable cartridges. It is
connected into a slip ring sub at the top of the kelly to allow electrical connection while rotating.
The system is relatively new and its reliability is not yet proven.
 A cartridge data transmission (CDT) system uses a steering tool that allows orientation of the
drilling motor in a particular direction, while still allowing drilling straight ahead with
drillstring rotation. ... the CDT system comprises three main subsections; the downhole
probe, which is connected to a transition assembly; the wireline cartridge as-sembly and
cartridge landing assembly; and, a rotating slip ring assembly which connects to the surface
processing unit and drillers remote display. Data is transmitted to the computer interface
unit via a simple conductor wireline.
 The time taken for a survey with the wet-connect system partially depends on hole geometry,
but it is on the order of 10 minutes.
Dreiling, et al., 199614 discussed using a wet connect system. In this system, the steering probe
is inserted in its receiving non-magnetic collar while the drillstring is in the vertical hole. The
wireline is cut and a connection made to a wet-connect sub inserted into the drillstring at the
surface. The trip is then completed and drilling operations begun. When a survey is desired, the
drillstring is blown down to evacuate the air and the female end of the connection is then
6-30

lowered on wireline to the wet-connect sub. These operations were so time-consuming and
there was so much trouble time that subsequent wells were drilled with EMWD.
Regardless of which method is used with the steering tool, failures in air are still a problem. Like
EMWD, steering tools fail much more frequently in air than they do in mud. Both tools need to
be hardened to work consistently in an air environment.
Hole Cleaning
It is more difficult to lift cuttings effectively from a highly deviated or horizontal hole than it is
from a near-vertical hole. The cuttings settling direction is still vertical while the drilling fluid
flow is not. As a result, the cuttings tend to settle on the hole bottom. There is a strong tendency
for cuttings beds to form at hole angles of from 45 to 60. If accumulation is left unchecked, this
deposition can lead to difficulties in tripping or even to stuck pipe. Also:
Experience has shown that the volume of air that will clean the hole while drilling with
a rotary assembly will not necessarily clean the hole during downhole motor drilling
with no drillstring rotation.
In high inclination or horizontal wells, cuttings fall to a maximum inclination. Thus,
poor hole cleaning will be evidenced by excessive drag while the BHA is pulled through
that section, and by bridges encountered during a trip in the hole.
Shale, 199520
It is possible to clean deviated and horizontal holes drilled underbalanced, whether using dry air
(Lattimore, Carden and Fisher, 198721), foam (Shale and Curry, 199319), gasified liquids (many
horizontal holes are drilled with gasified liquids in Canada) and flowdrilling (Pearsall and
Giddings fields of South and Central Texas).
 Even when dry air is used, the rate needed to clean the hole is more than for a vertical well.
As a rule-of-thumb, the rate should be twice that recommended by Angel, 1957.22 Even more
rate is required when drilling with mist or foam. The exact volume is difficult to predict. For
this reason, it is often desirable to run downhole motors using oil as the lubricant.
 Drillpipe rotation aids cleaning in an air hole. Experience has shown that the volume of air
that will clean the hole while drilling with a rotary assembly, will not clean the hole while
drilling with a downhole motor (no drillstring rotation). The cuttings are agitated and ground
finer by the rotation of the drillpipe, allowing the air to carry them out of the hole.
Many studies of cuttings transport from inclined holes have been done for mud. Relatively little
has been done with compressible fluids. There are theoretical analyses of cuttings transport for
dry air,23 for foam,24 and for gasified liquids.25 These models have the same limitations as they
do in vertical hole applications - for example, the cuttings size is rarely known with any
accuracy.

6-31

Chapter 6

Special Considerations

A rule-of-thumb that seems to work is that adequate cleaning of a horizontal hole will
probably be achieved by using 2.5 times the volumetric rate that would have been required for
a vertical hole of the same measured depth. In the absence of field experience, this is probably
a reasonable starting point.
Torque and Drag
Torque and drag are significant issues in drilling with compressible fluids. They are caused by
frictional interaction between the drillstring and the wellbore wall. Drag is a function of the
frictional coefficient. In mud-filled holes, the coefficient of friction is regulated by the lubricity
of the mud and the relevant additives. In holes drilled with compressible fluids, the drag can
only be impacted by modifying the frictional coefficient (may be possible with foam or mist),
changing the profile of the well (building angle at higher rates and to low inclinations will yield
a greater drag than buildup at low rates and to high inclinations26) or changing the string weight
(reducing the tension in a dogleg reduces the torque and drag in the wellbore26).
The friction coefficient in an air-drilled hole can be three to four times that expected in mudfilled holes (Shale, 1995).26 Carden, 1991,27 gives typical numbers for friction coefficients with
dry air and with oil- and water-based muds. String torque and drag must be carefully considered
before drilling a deviated or horizontal well with a compressible fluid, to ensure that the
proposed BHA can be tripped in and out of the hole and rotated safely, and can supply the
required weight on bit. Higher string friction with air (or mist particularly) can limit the ultimate
horizontal departure that can be attained to much less than that achievable with a liquid drilling
fluid. It will also reduce the penetration rate that can be achieved when drilling in oriented mode.
Horizontal Section Length
The length of horizontal hole that can be drilled with air will be less than with mud. At some
point, drag will prevent the drillstring or casing from falling in the hole. The drag is a function of
the friction coefficient between the pipe and the wall of the hole. In a mud-filled hole, the
friction coefficient is affected by the lubricity of the mud, which can be controlled with additives.
There are no friction reducing additives that can easily be added to air. Foam or mist can
increase lubricity, but the attendant hole cleaning problems can often nullify the effect. Cuttings
will create additional drag.
A typical friction coefficient for an air hole is 0.45. Frictional coefficients in mud-filled holes
range from 0.2 to 0.35. Figure 6-14 is a plot of hook load versus horizontal hole length for 5inch, 20 lbm/ft casing at 2,600 feet TVD, with various friction coefficients. When the hook load
falls below zero, the pipe will no longer fall into the hole by itself, limiting the amount of
horizontal hole section that can be drilled.
Lithology and Target Constraints
The types of lithologies that can be drilled with air are limited. Generally, older, consolidated
rocks are applicable for air drilling. Softer rocks will have a tendency to slough since there are
reduced fluid pressure forces to support the borehole wall.

6-32

The amount of water that can be accommodated in an air hole is limited. If the formations above
the target reservoir produce significant amounts of water, that portion of the well would have to
be drilled with fluid. The horizontal section can still be drilled with air, if casing is set through
the water producing strata. A cost analysis would have to be performed in order to determine if
an extra casing string would be economical.
60000

Hook Load (lbf)

50000

Friction Coefficient 0.2

40000

30000

Friction Coefficient 0.3

20000

10000
Friction Coefficient 0.4
Friction Coefficient 0.5

0
0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

Horizontal Section Length (feet)

Figure 6-14.

Horizontal section length versus hook load.

Air drilling cannot be continued when excessive oil and/or gas rates are realized from the
producing formation. The gas presents a fire hazard while tripping. Normally, up to 5 MMcf/D
can be kept off the rig floor by using a blooie line with proper jetting configurations. Large
volumes of oil must be picked up from the pit, stored in tanks, cleaned and sold.
Without a true steerable system for an air hole, thin reservoirs cannot be drilled efficiently. A
target thickness of 50 feet or more is required when using rotary assemblies. The build and drop
tendencies of rotary hold assemblies are difficult to maintain below 0.25/100 feet. In thin
targets, too many motor correction runs would usually be required to make air drilling cost
effective.

6-33

Chapter 6

Special Considerations

Summary
Underbalanced drilling of directional and horizontal wells is essentially no different than if these
well were drilled overbalanced with mud. Drilling directional and horizontal wells
underbalanced is now a practical proposition and can be a cost effective alternative to drilling
with mud. The operator needs to be aware of the limitations and advantages of air drilling in
order to optimize drilling operations. Some of the standard practices used in drilling vertical air
holes have to be modified. As with any horizontal drilling operation, careful planning is one of
the keys to a successful well. Underbalanced directional surveying is evolving so rapidly, that
service companies should be consulted for the status of current technology.
Specialized air motors are available. These have largely overcome many of the restrictions on
using downhole motors with compressible fluids. Even so, motor life can still be shorter in an
air-drilled hole than when drilling with mud, and motor operating procedures need careful
attention. Surveying the horizontal section is difficult and time consuming. A reliable
electromagnetic MWD system would greatly simplify survey programs and would make a
steerable system cost effective. Horizontal drilling with air is more effective in thicker target
intervals because of the limitations of steerable systems. Hole cleaning is more difficult above
inclinations of 50 and is complicated even further when mist or foam is used. The formations
that are applicable to air drilling are older, consolidated formations that do not produce excessive
amounts of gas, oil, or water.

6.5

Percussion Drilling

Background
In percussion drilling, rock is broken by causing the bit to repeatedly strike the workfront,
without imparting any significant shearing component to its action. In oil and gas drilling, this
action is achieved by incorporating a hammer tool in the BHA, immediately above the bit. The
hammer is actuated by the drilling fluid flowing through it. It takes energy from the fluid flow
and uses this to drive the bit down to impact the rock, lift the bit, drive it down again, and so on.
Typical hammer operation rates are 10 to 20 Hz. Noise may or may not be a problem; the
operating frequencies are in the low audible range (particularly when the bit is near the surface).
Percussion drilling is normally only used when drilling with dry gas, mist or foam. The main
reason for this is that reliable hammers are not available for drilling with liquid (such as mud).
One can imagine that a liquid would effectively dampen the hammers impact. With water
inflow, the increased annular pressure can also dampen percussion effectiveness. Modest
pressure increases (or differential pressure in permeable rock) over formation pressure may
reduce penetration rates more than would happen in rotary drilling. This pressure rapidly
suppresses tensile failure at the workfront that occurs in a dry hole being percussion drilled.
Similar dampening can occur with foam.28

6-34

Impacts between teeth (inserts) of a percussion bit and the rock cause crushing at the points of
impact. In addition, impact loading sends compressional stress waves into the rock being drilled.
These reflect off internal boundaries in the rock (grain boundaries, bedding planes, etc.) as tensile
waves. These tensile waves spall fragments from the rock surface. Since the rock breaks in
tension, the fracturing process is very efficient. Percussion drilling can be one of the most
energy-efficient drilling techniques. It removes the largest volume of rock for a given quantity of
energy input. Percussion drilling is not appropriate in soft rock, particularly soft, ductile
shales. It should not be used if even a short interval of ductile rock will be penetrated.
One of the main benefits of percussion drilling is a high penetration rate in hard rock, with a low
weight on bit. It can be very difficult to get sufficient weight in the string above the bit to drill
hard rock near surface with a roller cone bit. In this situation, a hammer can be very beneficial.
Penetration rates with percussion drilling can be significantly higher than those for rotary drilling
with the same drilling fluid, even when it is possible to get full weight on the roller cone bit.
Details of penetration rate increases associated with percussion drilling are given by Finger,
1984,28 and Pratt, 1989.29
Percussion drilling requires low weight on bit. Only a few thousand pounds is required, even for
large diameter bits. The weight only needs to be enough to ensure that the bit is in contact with
the workfront and to actuate the hammer. The hammer action provides the energy to break the
rock, not the string rotation or the weight on bit. Whitelely and England, 1986,30 suggested
drilloff tests to find the optimum weight for each application.
Another potential benefit of percussion drilling (over rotary drilling) is that the reduced rotary
speed and weight on bit can increase the life of drillstring components. Abrasive wear is much
reduced, and the fatigue life can be improved because of the reduced number of string
revolutions to drill a given distance.
It has been claimed that hammer drilling will reduce hole deviation, because the hammers are
operated at lower weights on bit (Reinsvold et al., 198831). Pratt, 1989,29 noted that a stabilized
hammer is required if percussion drilling is to give dogleg severities less than those for rotary
drilling and to avoid spiraled or ledged hole.
Equipment
Oilfield air percussion hammers, using reinforced three cone bits, were introduced in the early
1960s. Roller cone bits are not used with current generation hammers since higher impact energy
levels will lead to fatigue failure of the bit body. Industrial hammers, with a solid head and a
relatively flat-bottomed bit, with tungsten carbide inserts, supplanted the original oilfield
hammers. As Finger, 1984,28 indicated, testing established that industrial hammers, with less
WOB and about the same air supply, drill much faster than oilfield hammers. Oilfield hammers
are designed with lower piston velocities, to protect the bearings. Also, energy transfer through
the one-piece industrial bit is more efficient than through the bearings and threaded joints in the
roller bit.

6-35

Chapter 6

Special Considerations

Industrial hammers with flat-bottom bits were used as early as the 1960s. The operating pressure
of these hammers was typically 100 psi and their effectiveness was limited. Hammers, with an
operating pressure range of 250 psi, were introduced in the 1970s and penetration rates
improved. By the 1980s, the trend toward higher pressure hammers continued, with operating
pressures as high as 350 psi; however, the higher pressures can cause problems with the flatbottom bits. Hammer pressure drop typically adds 150 to 200 psi to the standpipe pressure. An
operator will usually need to run a booster, in addition to air compressors. Otherwise, no special
requirements are needed, beyond those for rotary drilling with the same fluid. Finger, 1984,28
indicated that it has been theoretically shown that the power input to the rock should increase as
the 3/2 power of the fluid supply pressure. Figure 6-15 shows laboratory data correlating ROP to
supply pressure.
Figure 6-16 shows one type of industrial hammer. This figure indicates that an industrial
hammer transfers kinetic energy through the bit to the rock by using a reciprocating piston within
the hammer. The rock fails in compression and the hammer drills ahead. In most cases, weight
on bit is maintained at less than 5000 lbf and the hammer is rotated at 15 to 30 rpm. The
percussion hammer provides high penetration rates at low bit weights. Figure 6-17 shows
common flat-bottomed bits used with an air percussion hammer.
The device in Figure 6-16 is commonly called an internally ported hammer. High pressure air is
contained within the control rod by the choke at the end of the rod. The air pressure forces the
piston to travel up and down. The hammer in Figure 6-16 is in the open position, as it would be
while being run into the hole. Once the bit is placed on bottom, the bit pushes the piston up,
until these upper ports align with the upper control rod window. High pressure air will pass
through the upper piston ports and into the bottom air chamber, causing the piston to move up.
Once the piston moves above the foot valve, the high pressure air is allowed to bleed off through
the bit. As the piston continues to travel up, the lower piston ports pass by the three lower
windows in the control rod, allowing high pressure air to pass into the upper air chamber. The
high pressure air above the piston will cause the piston to reverse direction and the piston is
forced down. As the piston hits the strike face of the bit, the high pressure air above the piston is
vented below the choke, through the lower piston ports. Simultaneously, the upper piston ports
will again be aligned with the upper control rod window, repeating the cycle.
Figure 6-18 shows another hammer available to the industry. This is an externally ported
hammer. The tool works in the same way as the internally ported hammer except that the high
pressure air travels on the outside of the piston rather than through a central tube. Components
of this type of tool are shown in Figure 6-19.
The rate of penetration of a hammer tool in hard rock is proportional to the operating pressure.
The higher the pressure differential between the upstream and downstream side of the choke, the
more impact energy that is imparted to the bit. A common hammer will cycle at 10 to 20 Hz
(higher pressure is associated with a higher frequency).

6-36

If a rocks unconfined compressive strength, Co, is known, it is possible to use the high efficiency
of percussion drilling to estimate the maximum penetration rate that will be achieved. The
mechanical specific energy, MSE, for drilling, is the energy input required to excavate a unit
volume of rock. In hard rock percussion drilling, this can be close to the unconfined compressive
strength of the rock (Teale, 1963,32 and Simon, 196333). An approximation of the ROP can be
made:
 Assume that the MSE = Co.
 Determine the hammer manufacturers power output value. The penetration rate is related to
the rocks unconfined compressive strength, the hammer power output and the hole area by:

70
15 rpm
20 rpm
25 rpm

Penetration Rate (ft/hr)

60

50

40

30

20

10

0
50

100

150

200

250

300

350

Air Supply Pressure (psi)

Figure 6-15.

ROP for three different percussion tools with 8- to 8-inch solid-head bits in Sierra
White Granite, for a WOB of 5000 lbf. The flow rate was between 600 and 1100 scfm
28
for each hammer (after Finger, 1984 ).

6-37

Chapter 6

Special Considerations

Pin up Connection

Backflow Valve

Control Rod
Piston Case
Control Rod Windows (4)

Piston

Foot Valve

Choke

Piston Strike Face


Upper Bearing Surface
(lock rings attach here)

Lock Rings
Rotational Drive Spines

Driver Sub

Lower Bearing Surface


Fishing Threads
Head Section
Diamond-Enhanced Inserts

Cutaway view of a Downhole


Industrial Hammer

6-38

Hammer Bit Profile


with Specific Features

Figure 6-16.

Internally ported hammer and a flat-bottom bit (anon.).

Figure 6-17.

Flat-bottom bits, used in conjunction with an air percussion hammer (anon.).

Figure 6-18.

Externally ported hammer (anon.).

6-39

Chapter 6

Special Considerations

Piston
O-Ring
Top Sub

Bit Retainer
Ring
O-Ring

O-Ring

Driver Sub

Dart

Spring

Foot Valve

Check
Valve Guide
Make-Up
Ring

Drill Bit

Choke

Rigid Valve

Piston Case
O-Ring

5" - 8"

Piston Retainer Ring

Figure 6-19.

ROP = 2.52 106

P
Co D h2

Components of an externally ported hammer (anon.).

(6.2)

where:
ROP ... rate of penetration (ft/hr),
P ....... hammer power output (hp),
Co ....... unconfined compressive strength
(psi), and ,
Dh....... hole diameter (inches).

6-40

The pressure differential or operating pressure is a function of the choke size, the air rate and the
bottomhole pressure. The operating pressure is the difference between the pressure upstream of
the choke and the bottomhole pressure. The deeper the well gets, the higher the bottomhole
pressure will be. This reduces the pressure differential. As a well gets deeper, smaller choke
sizes are required to maintain a constant operating pressure if the flow rate is not increased.
When water is encountered, wellbore pressure can rise dramatically and hammers often quit
working.
The injection flow rates should meet the manufacturers specifications and should be consistent
with rates required for good hole cleaning. If hole cleaning requires a higher flow rate than the
hammer can tolerate, there may need to be a bypass port in the hammer or in the string just above
the hammer. This port is adjusted (using a changeable choke/nozzle) to ensure that the hammers
specifications are not exceeded.
Hammers require lubrication for trouble-free operations. Inadequate lubrication is a major cause
of hammer wear and failure. Most hammer manufacturers recommend using rock drill oil,
because it has a high film strength and will adhere to the piston. The oil is usually injected into
the standpipe with a positive displacement pump. The minimum recommended injection rate is
approximately 0.2 quarts per 100 scfm per hour. Alternatively, oil can be dumped down the
drillstring during connections. When first running a hammer, several gallons of oil should be
poured down the drillstring, to coat its inside; otherwise, very little oil will reach the hammer
during initial injection.
A dirty drillstring can lead to premature hammer failure. Rust, mud and dirt in the drillstring can
flake off while drilling with the hammer and this debris will be carried down to the hammer with
the air. To work effectively, hammer tools have small clearances between the piston and other
parts. Dirt or rust between moving parts will cause excessive wear. To avoid problems, the
inside of the drillstring should be cleaned, especially if it has been previously used in mud
drilling. Many operators will hit the drillpipe with a hammer as it is being picked up, to knock
rust and mud loose. The best method for cleaning is to use a high pressure wash inside the
drillstring, similar to that used to clean paraffin from tubing. Avoid excessive doping of
connections; dope can accelerate hammer wear.
One of the primary uses of hammers has been to drill larger diameter surface holes in hard rock.
When drilling with the rotary close to the surface, very little weight can initially be applied to the
bit, due to the small number of drill collars. Not enough weight can be applied to cause the rock
to fail in compression (the primary drilling mechanism). A hammer will drill with very low bit
weights because the energy is supplied by the piston. Surface hole can be drilled in a matter of
hours rather than days.

6-41

Chapter 6

Special Considerations

Hammers have also been used to control hole deviation. The literature indicates that a well
drilled with a hammer will have less deviation than a well drilled with the rotary, although no
specific reason can be given. In any event, penetration rates with the hammer are much higher
than with a rotary assembly fanning bottom. The operator should prove to himself that the added
expense of the hammer is offset by lower total drilling costs. Deviation problems still exist, even
with hammers.
Directional drilling is possible with a hammer in the string, using stabilizer placement to control
bit inclination. The hammers stiffness and proximity to the bit can limit possible stabilizer
placement, restricting directional control.
Percussion hammers have been used in place of drilling with a typical rotary assembly, but they
are not economical when maximum bit weights can be run with the rotary assembly. Even
hammer bits should be rotated at the slowest rotary speed possible, without stalling the rotary
drive (15 rpm, if possible). The string is rotated only to ensure that the bit inserts do not fall into
cavities created by the previous impact and to pulverize the formation over the full crosssectional area. Rotation does not impart additional energy to the rock. Pratt, 1989,29 stated that
We generally kept the rotary table speed below 45 rev/min because bit life above this value is
short. Excessive rotary speed increases gauge wear. Although penetration rates will be
comparable in both instances (hammer versus rotary), the extra cost of the hammer and potential
failures makes the cost per foot for percussion drilling higher. If maximum bit weight cannot be
run, then the hammer will outperform the rotary assembly. Whiteley and England, 1986,30
provided some indication of the influence of WOB for the bit-hammer combinations that they
used (refer to Figure 6-20).
Finger, 1984,28 also addressed the issue of WOB, emphasizing that solid heads are vulnerable to
excessive weight because they are relatively weak in shear and that maintaining relatively small
weight control can become difficult at greater depth.
Hole Cleaning
Before we used the industrial air hammer, we used Angels model to calculate minimum air
volumes as a function of depth. Fill on bottom and long circulation times necessary to clean the
hole indicated that the air volumes used were inadequate The problems with the Angel model
are that cuttings slip velocity is neglected, cuttings are assumed to be dust size, and 3,000 ft/min
air velocity is assumed to be adequate for hole cleaning.29 Pratt, 1989,29 indicated that an
alternative, more fundamental cleaning model was adopted and modified by:
 Using a revised air prediction module inside the drillstring where the friction factor was
calibrated from actual measurements.
 Exit boundary conditions were modified as an input parameter and exit chip velocity was
fixed at zero.
 The influence of the BHA and changing hole size were incorporated, and,
 Chip size change was built into the model.

6-42

Misting is usually required in larger-diameter holes because of freshwater sands encountered


near the surface. Soaps, clay stabilizers, and corrosion inhibitors mixed with water are injected
into the supply air to facilitate hole cleaning, to avoid clay swelling, and to protect the drillstring.
The volume of chemicals and water injected into the supply air is usually negligible and does not
affect the operation of the FPB/HT [flat-bottomed percussion bit/hammer tool] at typical
injection rates and supply air volumes.30
Inhibited mist is a mixture of KCl water and soap to clean the hole of cuttings and prevent mud
rings from forming. This formula also works in protecting water sensitive shales from sloughing.
Usually a concentration of 3 to 4% KCl will provide some inhibiting qualities to eliminate
hydration of clays; however, higher concentrations may be required depending on the degree of
sensitivity. An anionic polymer can be used in conjunction with the KCl water and soap to
provide additional inhibiting qualities. The liquid volume fraction [for foam] must be
maintained between 2 to 5% to possess a high solids carrying capacity [polymer base fluid]. Too
much liquid causes the foam to collapse under its own weight, whereas too much air causes large
air pockets to form.34

Recommended Weight on Bit (lbf)

7000

6000

5000

4000

3000

2000

1000

0
2

10

12

14

16

Bit Diameter (inches)

Figure 6-20.

FPB/HT [flat-bottomed percussion bit/hammer tool] tandem recommended WOB


versus hole size (after Whiteley and England, 1986 30).

6-43

Chapter 6

Special Considerations

Russell, 1993,35 also advocated misting and/or foaming to prevent air slugging and cleanup. To
prevent air slugging, water and foaming surfactant were sometimes used in concentrations that
were higher than necessary. This increased the hydrostatic pressure of foam in the annulus,
which reduced the frequency of the air hammer and, as a result, also reduced its power . It
became apparent that, to optimize air hammer performance, a foam with high bubble stability
and low density was required . The desired foam properties were achieved by adding polymers
to the water and foaming agent. The optimum formula, per barrel of water, consisted of:

1/8 lb polyanionic cellulose polymer (PAC)

1/8 lb xanthan gum polymer (XC)

1% foaming agent by weight water.35

Later CMC (carboxymethyl cellulose) was used instead of XC and PAC. It was less expensive,
easier to mix, resistant to calcium and chloride ion contamination, and the liquid volume was
reduced.35
Gauge Wear
Hammers are not without problems. The most significant problem has been gauge wear. In
some formations, flat-bottomed bits can loose gauge quickly. Higher bit weights have a tendency
to increase this wear (Sheffield and Sitzman, 1985;34 Whiteley and England, 198630). Gauge
problems are particularly prevalent when drilling through abrasive formations, such as those with
a high quartz content (Reinsvold, et al., 198831). Undergauge hole is very difficult to ream with a
flat-bottomed bit and hammer. Separate reaming runs with a rotary assembly are required. A
solution is to continually downsize subsequent bit runs. By running a smaller bit on the next run,
reaming can be avoided. Downsizing is always an option in larger surface holes but may be a
limited option in smaller holes and where long intervals are drilled.
The diamond-enhanced hammer bit was an advancement in hammer bit technology, designed to
combat gauge problems. The inserts are diamond-enhanced by fusing polycrystalline diamond to
the tungsten carbide. This bit resists gauge wear and will generally last longer than a standard
hammer bit. With the diamond-enhanced bit, gauge problems are reduced but not eliminated.
Since these bits are much more expensive than standard hammer bits, their application should be
economically justified. Figure 6-21 is a schematic of a diamond-enhanced insert. As can be
seen, there is a layer of polycrystalline diamond on top of the tungsten carbide core of the insert,
with transition layers to prevent spalling of the diamond. These bits have been found to be much
more effective in some horizontal sections with extreme side loading and abrasive sands
(Dreiling, et al., 199614).
Smooth Hole
Pratt, 1989,29 reported that hammer bits do not necessarily drill a smooth wellbore. Both spiral
holes and ledges were observed in test holes, drilled in granite blocks using an industrial hammer
and a flat-bottomed bit. Figures 6-22 and 6-23 show these effects. As a result, full-gauge rotary
bits and stabilizers will not follow the hammer bit without reaming to bottom.

6-44

Whenever possible, a slightly larger hammer bit should be used, prior to drilling with a rotary
assembly. Pratt used a stabilized body and driver sub with gauge-protected bits to eliminate the
problem. It is appropriate to use a near-bit stabilizer, or preferably a hammer with a stabilized
body. Otherwise, the BHA can be very light and simple - one or two collars and only enough
heaviwate to avoid fatigue problems at the transition from the collar to the heaviwate to the
drillpipe.
Polycrystalline Diamond

Diamond/Carbide Composite

2 Transitional Layers of

Figure 6-21.

Diamond-Enhanced Insert (after


Reinsvold, et al., 1988 31).

Do not ream down a hammer bit. If the hammer actuates without sufficient rock in front of the
bit, the bit can break off the rock ledge before all of the blow energy is absorbed. The hammer
piston then hits its stops and a tensile wave is created in the bit shank. This can rapidly lead to
the bit shanking and the head falling off Similarly, do not actuate the hammer if there is any
possibility of the bit not meeting intact rock.

Figure 6-22.

Spiral hole drilled with an industrial hammer and a flat-bottomed bit (after Pratt, 1989
).

29

6-45

Chapter 6

Figure 6-23.

Special Considerations

Ledges drilled with an industrial hammer and a flat-bottomed bit (after Pratt, 1989 29).

Fatigue
Over a period of time, a hammer will fail and extra trips will be required to replace it. To
prevent excess wear, the inside of the drillpipe should be clean. Dirt, rust and scale will lead to
premature failure.
Flat-head bits are prone to fatigue cracking where the head attaches to the shank (shanking the
bit). This is one of the main problems with percussion drilling. Some of the energy in the
compressional stress wave, that is sent down through the bit by the hammer, is reflected at the
bit-rock interface and travels back up through the bit as a tensile wave. Concentration of these
repeated tensile stresses at the transition from the bit head to the shank can lead to fatigue
cracking. It is important to avoid large differences between the bit head and shank diameters.
This can be difficult for some hole sizes (notably 17-inch hole) as the shank size is dictated by
the hammer to be used, and there are some large gaps between available hammer diameters.
Avoid using smaller diameter hammers than the hole can accommodate. This minimizes the
risk of bit shanking and provides more hammer power, leading to higher penetration rates.
Some flat-head bits are provided with a fishing thread on top of the head. Fishing for the head
can be difficult if there is any hole instability - debris often covers the thread. An alternative is to
use a catching ring that prevents the bit head from being left downhole when the hammer is
tripped, even if the head has separated from the shank.
Performance
Pratt, 1989,29 cautioned that acoustic analysis indicated that hammers may strike more slowly
than indicated by the manufacturers specifications, confirming that the manufacturers air
volume consumption curves are highly inaccurate in deep oilfield applications. These curves are
commonly developed for atmospheric pressure discharge from the tool. Air volume through
the hammer is reduced considerably by backpressure. Choke sizes should be adjusted to keep
the on- versus off-bottom pressure difference at about 100 psi. Choke sizes can be calculated by
6-46

using the equations for calculating bottomhole pressures.


Summary
 Maintain proper WOB.
 Rotate as slowly as possible.
 Provide an air bypass if necessary.
 Keep the threads clean and use recommended lubricant. Dope the pins only.
 When changing out bits in unfinished hole, be sure that the next bit is no more than 0.25inches larger than the bit being removed.
 Stabilize as required, particularly if the bit size is more than two inches larger than the outer
diameter of the hammer tool.
 Monitor the compressors and the standpipe gauge.
 Blow the hole clean periodically.
 Never run on downhole junk.

6.6

High Pressure Drilling

In previous chapters, various high pressure, underbalanced drilling techniques were described.
These techniques included flowdrilling, mudcap drilling and snub drilling. Each of these
methods will be discussed with emphasis on their application in high pressured environments.
Because of the high pressures, flowdrilling with coiled tubing (CTD) will also be discussed and
compared to jointed pipe flowdrilling. Special attention must be given to the higher surface
pressures because of the additional forces required to oppose them when running tools and pipe
into the hole. Stringent safety considerations are required because of the higher surface
pressures.
Flowdrilling in High Pressured Formations
Flowdrilling may be performed using jointed pipe or coiled tubing (CT). With the trend towards
drilling higher pressured en-vironments, more and more work is being done with coiled tubing.
Improvements in CT manufacturing and CTD (coiled tubing drilling) technology have facilitated
this trend. CTD provides a safer method of drilling than jointed pipe drilling, primarily due to
the elimination of connections.
Coiled tubing provides a safe, rigless way to perform underbalanced drilling and significantly
reduce formation damage in horizontal wells, lateral drainholes and deepenings. Coiled tubing
units are typically used for live well interventions. Coiled tubing equipment isolates wellbores
from the atmosphere during the entire well work or drilling operation and avoids killing wells
before tripping. Coiled tubing, however, is also limited by higher cost compared to fully
depreciated conventional rigs, and cannot perform jointed pipe running for casing and tubing
completion operations.36

6-47

Chapter 6

Special Considerations

Regardless of the increasing use of CTD, a great amount of high pressure, underbalanced drilling
is performed using jointed pipe. Most of this drilling also involves drilling in horizontal laterals
or curves. This includes using drillpipe in the string as well as using tubing drillstrings. Without
exception, all surface well control equipment pressure ratings must be increased to a safe
level. The maximum working pressure (MWP) of standard BOPs must be upgraded to exceed
the highest possible surface pressures. Rotating blowout preventer equipment should replace
the rotating head to achieve safer, higher pressured operations. All other surface equipment
should also be reviewed, to ensure maximum operational safety.
In many high pressure drilling applications with jointed pipe, using clear drilling fluids could
result in surface pressure that would exceed even the higher rated maximum working pressure of
the RBOP. For additional discussion on RBOPs, refer to Chapter 2 or to Hannegan and
Bourgoyne, 1995.37 Unfortunately, the permeability impairment-reducing properties of clear
drilling fluids must be sacrificed in changing to a more conventional heavier mud system. The
increased hydrostatic pressure resulting from these heavier muds would still be below the pore
pressure in the target, high formations. Heavier muds would reduce surface annular pressures to
a safe operating range and formation permeability impairment would still be minimized by
drilling underbalanced.
An example of flowdrilling very high pressured formations using jointed pipe is a fractured
limestone target at 13,000 feet, with a pore pressure in excess of 9,100 psi; a pressure gradient of
0.70 psi/ft. At this pressure level, the rig BOPs, must be rated for 10,000 psi service. Mud
weights of approximately 13.0 ppg could be used to achieve underbalanced drilling conditions,
and still reduce surface annular pressures to a safe operating limit. In this example, a static
surface pressure of greater than 300 psi would be seen if a full column of drilling fluid opposed
formation pore pressures. Reduction in fluid density, caused by hydrocarbon influx and gas cut
mud, would result in higher surface pressures. Extreme care should be used to ensure that these
pressures remain below the MWP of the RBOP. Before making trips in such pressured
environments, a higher density mud could be bullheaded down the annulus to subdue the well
and allow for lower surface pressures during tripping operations.
As even higher pressure areas are drilled, coiled tubing drilling, mudcap drilling or snub drilling
may be required (refer for example, to Bloys, et al., 199438).
Coiled Tubing Drilling
Coiled tubing (CT) is ideal for deployment in live well conditions. Its very design, using a reel
of continuous drillpipe, eliminates the need for making connections, required in jointed pipe
operations. Coiled tubing has a uniform outside diameter. This eliminates tool joint and tubing
upsets. Leising and Newman, 1993,39 summarized some typical CTD applications (refer to Table
6-2). Vertical deepening, with a pendular assembly to keep the hole straight, is probably the
most straightforward application. A long BHA is used to provide WOB without buckling the
CT. The neutral point is always in the BHA so that the CT is always in tension.

6-48

Coiled tubing operations in Canada have encompassed underbalanced deepening of vertical wells
and horizontal sidetracks, after rotary drilling to the top of the target formation. In Alaska, coiled
tubing operations have been coupled with larger workover rigs. Snubbing jacks have also been
used for completion and running tubing after drilling. Since CTD operations rely on downhole
tool movement with a mud motor, rather than drillstring rotation, stripper rubbers may be used to
contain surface annular pressures. Typical CTU blowout preventer stacks include a four ram
unit with tubing rams, blind rams, cutter rams (shear rams), and slip rams. Above the rams is a
riser. The riser must be of sufficient length to lubricate tools in or out of a live well. At the top
of the riser a hydraulic pack-off is installed to facilitate an annular seal while sliding pipe in or
out of the well.3 All rams are replaced with an annular for shallow gas. Snubbing CT into the
well with hydraulic jacking force is accomplished with an injector head, rather than a snub unit.
The injector head is an integral part of a coiled tubing unit. Its function is to inject CT pipe into,
or remove it from, the well during live well conditions.
Madigan, 1993,40 in discussing slimhole CTD, indicated that:

Openhole drilling with motors on coiled tubing has several key advantages over
conventional rotary drilling techniques. First, a smaller, much more mobile rig can be
used. Second, considerable trip time can be saved since connections do not have to be
made up and broken every thirty feet. Third, coiled tubing offers a considerable safety
advantage in that drilling can proceed underbalanced with relative safety.40

Another benefit is that the addition of a circulating sub above the motor allows the
drilling fluid to be circulated out of the well immediately upon completion of the well.
The well can be immediately tested, and the tubing stripped out of the well under
pressures as high as 5000 psi.40

Drilling engineers often worry whether relatively flexible coiled tubing and the lack of
drill collars will allow drilling straight holes. High speed PDMs and small fixed cutter
bits, such as thermally stable diamond (TSD) bits, allow much higher rate of penetration
at lower bit weights than possible with rolling cutter bits, affording straight hole. TSDs
apparently also produce smaller cuttings.40

Table 6-2.

Summary of Typical CTD Applications (after Leising and Newman, 1993 39)
Vertical

Deviated

Re-entry Drilling

Deepening of existing wells

Lateral drainholes

New Well Drilling

Disposal exploration wells

Steam injection

Observation and delineation wells

Environmental observation

Slimhole production/injection wells

6-49

Chapter 6

Special Considerations

Coiled tubing design and manufacturing technology continues to advance. Similarly, slimhole
tools run in the hole with CT have improved in recent years, with the increased use and demand
for these operations. With that has come improved field experience and training. Higher
strength alloy steels, combined with sophisticated and rigorous quality assurance measures,
provide safer, more dependable CT products. Field testing and computer modeling have also
helped to determine fatigue limits for every string of CT pipe currently in use.
Design Considerations for Coil Tubing Drilling
Leising and Newman, 1993,39 is a good reference on CTD design. The protocol should
(iteratively) entail:
 Select the CT size, the hole size, the drilling fluid, and the BHA.
 Calculate the reel weight and size. Be sure it can be transported and will fit on location.
 Calculate the tubing forces and stresses, to ensure that they never exceed 80% of the yield
strength and that the minimum acceptable WOB can be provided at TD. Include friction
associated with bending the BHA around any curves.
 In Vertical Wells: If a constant wall thickness with depth is assumed, the maximum depth
of the CT in the drilling fluid, without exceeding 80% of the yield strength of the material, is
given by:
D max =

y
4.245 0.06493Wdf

(6.3)

where:
Dmax ... maximum depth (feet),
Wdf ..... drilling fluid weight (ppg),
and,
y ....... yield stress (psi).
Tapered strings can be used for greater depths.
 In Deviated Wells: A tubing forces model should be used because of the potential for helical
locking in the hole. The force that can be applied to the CT to push the BHA around the
curve and into the deviated section may also be limited by the maximum force that can be
applied in the vertical section.

6-50

Ensure that the injector can supply the necessary pull/push (if kickoff is shallow).

Calculate the drilling fluid pressure drop in the CT, BHA and annulus at 100% motor
flow capacity and determine the absolute pressure in the CT during drilling. Be certain
that pressure limitations are not exceeded.

Assess torsional limitations. The downhole motor-stall torque should be no larger than
the maximum working torque for the CT. When the string is picked up, torque will be
near zero. During drilling, the torque again will be at the maximum operating torque.
Torsional cycling compounds bending cycling and reduces fatigue life .39 In general,
torque is not a significant limitation unless large hole drilling is performed with small
diameter CT. This is most relevant to shallow, vertical-hole drilling.

Calculate the fatigue life using the parameters from above. Large diameter and
continuous application of drilling pressure can increase fatigue wear. Some life
prediction models are available.

Assess any hydraulic limits (assume drilling at 80% of the maximum motor flow rate).
Consider hole cleaning in vertical, inclined and horizontal sections.

Be certain that methods for directional control are possible. If the tubing is too limber,
small WOB and torque changes will change the toolface angle. An orienting tool is used
to make toolface changes during CTD. This tool can be controlled from the surface by
means of pressure, weight, or an electric wireline and causes drilling to proceed in a
different direction BHA analysis is the same as for conventional drilling. Because the
CT cannot be rotated, to drill straight ahead, either the natural tendency of the formation
must be countered by the BHA, or the orienting tool must be actuated periodically to
prevent undesired uniform build.39

High Pressure Mudcap Drilling


Mudcap drilling, described in Chapter 2.9, is another technique commonly adopted for high
pressure operations. This method uses a high density, viscosified mud in the annulus, while
drilling without returns; a less expensive drilling fluid is pumped down the drillpipe. In lower
pressure environments, this drilling fluid is often a non-damaging clear fluid such as brine water.
As higher pore pressure formations are encountered, the maximum working pressure of the
RBOP is reached with this relatively low density fluid. Mudcap drilling in these areas requires
higher density drilling fluids.
Consider, for example, a formation at 15,000 feet, with a pore pressure of 11,900 psi. The mud
weight necessary to balance this pressure is approximately 15.2 ppg. Mudcap drilling in this
formation could be performed with a 14.0 ppg viscosified annular fluid closed in at the surface,
with approximately 1,000 psi pressure. The drilling fluid pumped through the drillpipe must
weigh more than saturated NaCl brine water, due to the 4,100 psi resulting underbalanced
pressure. That pressure, plus the frictional pressure losses, would put too great a strain on
surface pumping equipment. An 11.7 ppg calcium chloride water (40.0% calcium chloride by
weight of solution) would lower surface pumping pressures by 1,326 psi, still resulting in
excessive conditions. More costly brines such as calcium bromide or zinc bromide would be cost
prohibitive. A 12.8 ppg lignosulfonate mud could be used to mudcap drill the zone, with surface
pump pressures of 1,916 psi plus friction loss.

6-51

Chapter 6

Special Considerations

In selecting fluids to be used for mudcap drilling projects, a close evaluation of cost and possible
formation impairment must be factored into the equation. The maximum operating limits of the
rigs pumps, as well as its BOPs and the RBOP must also be considered. The optimum design
for the job requires a balance between these pressure limits and the cost of the necessary fluids.
Bloys et al., 1994,38 provided additional discussion of required equipment for offshore
operations.
Snub Drilling in High Pressure Formations
Conventional snubbing units, as well as hydraulic rig assist (HRA) snubbing, allow entering a
well under pressure. In underbalanced drilling, snubbing enables a well to be safely drilled, by
allowing the removal and insertion of all downhole equipment, including tools and pipe, under
live well conditions. HRA snubbing was introduced in Canada roughly fifteen years ago. About
eight years ago, it began to be used in the United States. It offers a lower cost alternative to the
conventional snub unit.
Equipment for HRA snubbing is designed to be portable and easy to rig up and rig down. It
consists of a set of hydraulic ram-type BOPs, connected directly above the rig BOPs or RBOP.
A pair of hydraulic cylinders provides the jacking force for pipe movement. This equipment,
along with a working platform and control panel, is transported on a tandem axle truck which
also carries the hydraulic power source. Risers are used to position the unit just above the rotary
table. As part of this riser spool, ports with hydraulically-operated valves are added to bleed off
and equalize pressures between the ram preventers and an annular preventer mounted on top of
the snub stack. Above this annular, a set of inverted slips, or stationary snubbers, are installed, to
hold the pipe while the snubbing operator extends or retracts the traveling head of the unit. In
this manner, the operator is able to move the pipe into or out of the well one step at a time. The
inverted positioning of these slips prevents the pipe from being pushed out of the well by the
pressure. A second set of inverted slips, called traveling snubbers, is mounted on the bottom of
the traveling head, along with a conventional set of slips, mounted on top of the traveling head.
This traveling head is fixed on top of the hydraulic cylinders, to mechanically push or pull the
pipe from the wellbore under pressure.
HRA snubbing is so named because it assists the drilling rig, while using the rig drawworks and
derrick for pulling and racking pipe during trips. The HRA snub operator works on the unit
platform, operating the hydraulic cylinders, slips and BOPs at the control panel. This operator
must work very closely with the driller in coordinating all rig activities. In snubbing pipe into the
hole, the hydraulic cylinders provide the physical force to push the pipe downhole under pipe
light conditions. When sufficient pipe has been run to allow enough weight in the string, the
pipe becomes pipe heavy, and the cylinders are no longer required to push the pipe into the
hole. The rig drawworks are then used to run the pipe into the hole or pull it from the well, until
a pipe light condition again exists.
A typical HRA snubbing unit crew consists of two men per twelve hour tower and one twentyfour hour supervisor, for a total compliment of five men. Close coordination between the

6-52

toolpusher and the snubbing supervisor is very important. The rig crew and the snub crew must
work together to safely achieve the desired drilling goal. Due to the technical nature of snubbing
equipment and because of the advanced training and experience of the snubbing supervisor, the
snubbing supervisor has final authority whenever disputes arise. Aside from the very high daily
expense of the snubbing unit, there are a number of advantages to snubbing, including the
following:
 Pressure control is handled with the snubbing units BOPs, with its annular preventer as a
primary system and its rams as secondary preventers. The snubbing unit also has its own
accumulator system. The rig BOPs and accumulator provide backup to these systems.
 Snubbing crew personnel are highly trained and experienced in working with high pressure at
the surface; rig personnel are more accustomed to controlling and preventing such pressures.
 In the event of pipe light conditions, the snubbing unit can provide the necessary control to
inject or pull pipe and tools under pressure.
Snubbing offers a versatile and safe, yet expensive, solution in underbalanced drilling operations.
To offset this expense, it is possible to minimize the time on location for these services.
Snubbing provides the most positive method for pipe movement and pressure control through its
independent use of hydraulic cylinders, pipe slips and BOPs.

6.7

Cementing

Introduction
Slotted or pre-drilled liners are commonly run in high permeability environments. There are
situations where cementing and perforating are carried out; for example, in lower permeability
environments, in certain naturally fractured formations, or where a gas cap needs to be cemented
off, etc.
Remedial procedures to overcome incomplete cement columns due to formation fracturing or
excessive loss of returns, are expensive and historically are relatively ineffective. Lightweight
cement slurries were designed to overcome gradient-driven loss of cement to in-situ formations
(either from hydraulic fracturing in formations with low in-situ stresses and low tensile strength
or flow into naturally fractured or cavernous formations). These slurries complement other
underbalanced operations performed on a well; there is less chance of fracturing weak formations
and the number of stages required can be reduced. This section briefly highlights some of the
design considerations for underbalanced cementing.
Normal extenders can be used to provide slurries with densities as low as 11.5 to 12 ppg. For
densities less than this, water separation becomes a significant problem, unless alternate
methodologies are adopted. These alternatives include using hollow microspheres as extenders or
using cement foamed with nitrogen.

6-53

Chapter 6

Special Considerations

Extremely lightweight cement can be used to:41


 Provide primary cementing in formations that will fracture easily. Foamed cement density
can be reduced to as low as 6 ppg in order to minimize, or completely eliminate, losses
during single-stage cementing of long intervals. Cements with light-weight solid extenders
are commonly used for primary cementing of conductor and surface pipes, where washouts
and low fracturing pressures are common.
 Cure Lost Circulation in Cavernous Vugs: Remedial plugging of large vugs is not always
effective with conventional systems, since gravity causes the slurry to slump to the bottom of
the cavity. Thixotropic characteristics of foamed cement can reduce this gravitational
segregation.
 Squeezing Depleted Zones: Squeezing is carried out to place cement in a specific location to
achieve an hydraulic seat. This includes shutting off perforations, fractures, channels and
other undesirable voids. Accurate placement and controlled dehydration is difficult in
formations that cannot tolerate con-ventional cement densities (13 to 16 ppg). Light-weight
systems are appropriate.
 Zonal Isolation: Gas influx during hydration and setting may be reduced if foamed cements
are used.
 Heat Insulation: As is the case for foamed drilling in permafrost, the entrained voids in
foamed cements provide good insulating characteristics.
Properties of Foamed Cement
Nitrogen is incorporated directly into the slurry. Some modifications of the base cement system
are usually adopted to guarantee relatively homogeneous distri-bution of the gas phase in the
slurry. Conventional extenders, such as perlite, will crush as hydrostatic pressure of the column
increases. Nitrogen concentrations, on the other hand, can be increased to overcome volumetric
reductions with increasing depth. On location, foamed cement is generated by providing gas from
a compressor or a liquid nitrogen vaporizing unit; surfactant(s) are added to stabilize the foam.
The specific surfactant system used is a key to cement stability, particularly in the alkaline
environment of the water phase of the slurry.
As with foamed drilling fluids, rheology is relatively complex, depending on the gas content,
bubble size, water-to-cement ratio ... Pumping is usually not a problem and these foams can be
generally characterized as pseudoplastic fluids. The flow behavior index, n, generally decreases
as the quality of the foam increases. Retarders and accelerators can be used, as in conventional
systems, in accordance with service company recommendations, and preferably in conjunction
with supporting laboratory information.

6-54

The questions many operators ask are:


 Will the strength of the system be adequate and will the sheath be completely destroyed by
perforating?
 The compressive strength of the foam slurry will generally be higher than a comparable nonfoamed slurry of the same density. Compressive strength will vary with the base slurry
density and the foam quality. Consider the specifics of the well and consult with the service
company. The degree and impact of perforation damage has probably not been adequately
quantified.
 Will there be gas migration through the cement itself after the system has set? Possibly.
Consult with the service company to design an appropriate system.
 Will bond be any different than for conventional systems? The same precautions apply as
for any cement system, the most important being effective mud removal.
Systems With Low Density Particulate Additives
The service companies also have various competitive products where hydrostatic head of the
cement column is reduced by the addition of lightweight particulates, for example hollow
microspheres. Using these as extenders facilitates developing lower density (9 to 12 ppg) with
only modest sacrifices in strength, permeability and free water segregation. There are upper
limits of applicability associated with hydrostatic particle crushing; but these limits are often
quite high. Extra caution is required when mixing slurries with densities below 9 lb/gal. It is
important that free water is carefully controlled. Hollow glass spheres have been used to lighten
drilling fluids, with the advantage that the system is relatively incompressible. These drilling
systems have been described by Medley et al, 1995.42
Design Considerations
Foam Quality
As with foamed drilling fluids, foam quality indicates the volumetric concentration of gas in the
total cement system. In general, depending on the specific application, the quality should not
exceed 40 to 44%.41 Recent publications suggest that higher qualities may be used, depending on
the required strength and permeability of the sheath. Good foam stability is essential in order to
maintain a pore structure which is not highly interconnected when the cement sets. This
[interconnectivity] is caused by unstable nitrogen bubble walls that rupture upon contact with
other nitrogen bubbles, then bubbles coalesce forming large gas pockets. This results in a
sponge-like structure with possible density inhomogeneity due to gravity drainage of the base
slurry. A good rule-of-thumb for an initial design is to mix the base slurry at its optimum water
ratio ... The same mechanisms that produce free water will contribute to segregation in foamed
slurries. 43

6-55

Chapter 6

Special Considerations

Foamers and stabilizers are selected on the basis of stability, efficiency, compatibility (elevated
temperature, pressure, highly alkaline calcium containing water phase), adequate strength and
permeability control, cost, safety, ease of handling ... Pure surfactants are never used without
foam stabilizers, preventing/inhibiting bubble coalescence. de Rozires and Ferrire, 1990, 43
found it difficult to produce stable foams with qualities greater than 70% (except in certain high
shear environments where they reached 80%).
PVT Behavior
Nitrogen requirements are affected by the system response to temperature and pressure; the
pressure after displacement and the temperature prior to setting. Cement wellbore hydraulics
simulators are available for designing the nitrogen requirements. Cement systems with nitrogen
are commonly designed to yield densities in the range of 7 to 11 ppg. System design is similar to
conventional systems, taking into account the density variations and the in-situ fracture and
formation pressure gradients.
Cement System
Nitrogen requirements are also affected by the formulation of the basic system. The base slurry
can be extended itself and this will reduce nitrogen requirements (higher water content). This
will however lead to sacrifices in strength and permeability. In foamed cementing, it is essential
to have a reliable caliper log since the hydrostatic head strongly influences the systems densities.
Finally, non-nitrifed caps (on the order of 500 feet) of mud, spacer or cement are commonly used
when circulating to the surface.
Free Water
Excess free water is an extremely critical issue in deviated, particularly horizontal sections.
This is of particular importance in nonconventional wells, since any free water present may
migrate to the high side of the annulus and create a channel. 44 This may severely restrict the
use of some foamed or solid extended systems; regardless, great care is required. Requirements
for horizontal well cementing are well summarized by Wilson, 1991.45 Ryan et al., 1995,46
provided general guidelines for mud clean-up in horizontal wells.
Backpressure
With modern simulation computer codes, the hydrostatic and density profile can be characterized
during all of the circulation period. Accurate backpressure control during circulation to the
surface is extremely important (de Rozires and Ferrire, 199043).
Permeability
Permeability increases with quality. When the bubbles are small and the distribution uniform,
this increase is very progressive. If the bubbles are non-uniform, the permeability remains fairly
low and then increases sharply. This transition occurs between [10 and 10.85 ppg] on the class G
system and between [9.2 and 10 ppg] on the class C system. Therefore it seems that in case of
non-uniform bubbles, a percolation threshold occurs around 35% quality. 43

6-56

Compressive Strength
Below 5 ppg, foams are not solid. Above this density, strength and permeability depend on the
bubble size distribution, the cement type, the additives and many other parameters. Table 6-3
shows some controlled laboratory data from de Rozires and Ferrire, 1990.43 Measurements by
these authors also supported using the thickening time of the base slurry to estimate the
thickening time of the complete system.
Fluid Loss
In general, increasing the gas content reduces the fluid loss. This is shown in Figures 6-24 and 625, from Chmilowski and Kondratoff, 1992.47 System permeability can also be affected by
temperature. Increased temperature can reduce permeability due to preferential growth of certain
crystals.

6-57

Chapter 6

Special Considerations

Table 6-3.
Example unconfined compressive strengths and permeabilities for various foamed
cements (after de Rozires and Ferrire, 1990 43)
Class G foamed cement (curing time 72 hours at 81F)
Broad Bubble Size Distribution

Narrow Bubble Size Distribution

Compressive
Strength
(psi)

Permeability
(md)

Compressive
Strength
(psi)

Permeability
(md)

1.89

4915

0.025

1.45

1960

0.039

1480

1.27

1450

0.160

1000

14.5

1.12

910

44

770

173

0.85

465

12022

410

676

0.61

390

5321

275

3730

Density
(g/cm3)

Class C foamed cement (curing time 72 hours at 81F)


Broad Bubble Size Distribution

Narrow Bubble Size Distribution

Compressive
Strength
(psi)

Permeability
(md)

Compressive
Strength
(psi)

Permeability
(md)

1.77

3015

0.050

1.46

2030

0.0075

1335

1.25

1.22

1365

0.017

985

5.7

1.08

1075

0.054

825

25.7

0.90

680

89

465

199

0.61

220

58890

230

6300

Density
(g/cm3)

6-58

160
140

Fluid Loss (cm )

120
100
80
60
20% Foam Quality

40

37% Foam Quality


50% Foam Quality
Base Slurry

20
0
0

10

15

20

25

Time (minutes)

Figure 6-24.

Fluid loss values for varying foamed cement qualities (after Chmilowski and
Kondratoff, 1992 47).

120
20% Foam Quality
37% Foam Quality

100

50% Foam Quality

Fluid Loss (cm )

Base Slurry

80

60

40

20

0
0

10

15

20

25

30

Time (minutes)

Figure 6-25.

Fluid loss values for 37% quality foamed cement with varying amounts of fluid loss
additive (after Chmilowski and Kondratoff, 1992 47).

6-59

Chapter 6

Special Considerations

Basic Design
The design of a foamed cement job should consider the pressure and density profiles at various
stages during circulation and not just at the end of the job. Simulators consider, for each stage,
the depth to the middle of the stage, and calculate the pressure at the middle of the stage. The
overall job design optimizes:
 The number of stages,
 The nitrogen ratio for each stage,
 The value of the backpressure (critical because the magnitude of the back-pressure strongly
affects the amount of base fluid in the well) and,
 The number of feet of cap fluid and its method of placement. The cap can be circulated
ahead of the foam or can be pumped when foam first reaches the surface (bullheading down
the annulus). While more difficult operationally, the second option may be preferable to
avoid U-tubing of the first foam behind the cap. Also, if the cap precedes the foam, the
hydrostatic pressure of the column may exceed the fracturing pressure when the cap turns
around the shoe. For bullheading, after foam has reached the surface and the correct
backpressure is applied, the tail is pumped and followed with mud (avoid free fall). When
the tail has turned the shoe, pumping stops and the well is shut in, keeping backpressure on
the return line. The cap is then pumped down the casing.
The governing constraints are:
 The hydrostatic profile of the column should always remain below the fracturing pressure,
 In certain environments, it may be possible to cement completely under-balanced by
cementing below the formation pressure. This is only possible if it can be guaranteed that the
slurry quality will not be adversely affected (gas pockets, supplementary liquids) and that
channeling and/or microannuli will not develop. Work closely with the service company.
 The density profile at the end of the job must provide the cement qualities to support casing
and provide adequate isolation,
 The procedures need to be as operationally simple as possible,
 Cement loss must be minimized and,
 Nitrogen usage should be minimized.
Purvis and Smith, 1994,48 discussed some operational issues and presented case studies.
Chmilowski and Kondratoff, 1992,47 provided observations on designing foamed cement squeeze
operations in low pressure, highly permeable formations (e.g., the Ellenberger and San Andres in
West Texas, the Sadlerochit in Alaska, the Keg River, Wabamum and Leduc D3 in Canada and
the Baturaja in Indonesia). They advocated more effective use of information from pre-squeeze
injectivity data.

6-60

One service company uses the bottom of the stage.

6.8

Formation Evaluation

Evaluation of Underbalanced Holes


Evaluation of Formation Fluids While Drilling
The information that can be acquired while drilling underbalanced holes is commonly more
diagnostic than that obtained while
drilling conventional holes. Since under-balanced drilling implies that the pressures in the
borehole are less than formation pressures, formation fluids may be produced into the borehole.
The circulating drilling fluids will carry these fluids to the surface where these produced fluids
may be evaluated.
Evaluation at the surface can be qualitative or quantitative. For instance, an experienced driller
can infer something about the produced fluids by simply watching the flare or effluent at the end
of the blooie line in an open drilling system. Increased amounts of natural gas will, of course,
lead to a larger flare. Oil would cause a darkening of smoke from the flare while an influx of
water will result in a noticeable change in the color of the flare, if not extinguishing it altogether.
Mud logging should be used to help identify the formations, assess the quality of the reservoir
rock, and pinpoint the types and depths of formation fluid influx into the wellbore.
In a closed system, flares may still be monitored but liquids can be quantitatively measured,
using sight tubes attached to the separator at various levels. Other instru-mentation can be used
to keep track of fluid types and rates. In addition, the various liquids produced can be collected
in separate tanks after separation, for recycling, sales or disposal at a later time.

6-61

Chapter 6

Special Considerations

Evaluation with Logging Tools 49,50,51,52


The choice of logging tools run into a borehole, whether drilled underbalanced or not, will
largely depend on what liquids are in the borehole at the time of logging. If a liquid was used to
drill the hole and covers the zones of interest during logging, the logging program would be no
different than for a conventionally drilled hole. Depending on the pressure balance at the time of
logging, the wireline tools might have to be run through a lubricator. However, if the hole was
drilled with gas (air, methane, nitrogen ...) and gas remains over the zones of interest at the time
of logging, a careful selection of logging tools is required. Any, or all, of the following tools
could be useful for evaluation of such holes:










Gamma Ray - For formation or bed definition (e.g. distinguishing sands from shales).
Spectral Gamma Ray - Quantitative definition of the gamma ray spectrum; to define
clay content, clay and mineral type, and to aid in fracture detection.
Epithermal Neutron - To identify porosity of liquid-filled zones.
Induction Resistivity - To help distinguish hydrocarbons from saline formation water
and to help quantify the water saturation.
High Resolution Density - To quantify porosity.
Temperature - To indicate liquid level in the borehole and to delineate zones where
fluids are actually being produced.
Production - Production logs, such as borehole spinners, can help to quantify the
relative amount of production from each interval.
Nuclear Magnetic Resonance - To help quantify the permeability of formations.

Gamma Ray Log


The Gamma Ray Log measures the natural radioactivity of the formation and usually correlates
with the Spontaneous Potential, SP, (which cannot be run in gas-filled boreholes). Like the SP,
the gamma log can distinguish sands from normally more radioactive shales. Also, it provides a
good correlation curve, defines bed thicknesses, and aids in interpretation of environmental
deposition. It is usually a more diagnostic indicator of beds in a carbonate sequence than the SP.
The gamma log can be run in any liquid- or gas-filled hole, either cased or uncased. A gamma
log run in a gas-filled borehole may have a slightly higher radioactive count than the same log
run in a hole filled with mud, particularly if the mud has a high concentration of barite.
Spectral Gamma Ray Log
Like the conventional Gamma Ray Log, the Spectral Gamma Ray Log records gamma ray
intensity; however, additional curves identify potentially productive zones. These curves
(uranium, potasium and thorium concentrations) also provide for more detailed correlation with
nearby wells that have similar information. Sometimes, naturally fractured zones or carbonates
will have small quantities of uranium, precipitated by subsurface water flow. This can be
identified by the uranium curve as a potentially productive zone whereas it might be

6-62

misinterpreted as a shale interval from the gamma ray log. In an underbalanced hole, this would
provide important auxiliary information to help pinpoint flow intervals.
Epithermal Neutron Log
Neutron logging instruments contain a radioactive source that bombards the formation with fast
neutrons. These neutrons are slowed primarily by hydrogen atoms in formation fluids and can be
counted by detectors in the tool which are scaled in porosity units. In clean or shale-free
formations, the neutron log does not recognize gas-filled porosity. It can be compared with
another porosity device, such as a density log, to detect gas-filled zones (crossover). It can also
be used in combination with other porosity logs for interpretation of lithology, including shaly
sands.
In an air-filled or mixed fluid (such as nitrified crude oil) hole, the normal Compensated Neutron
Log will not respond properly; detrimental effects of the borehole fluid can be minimized with a
Sidewall Epithermal Neutron instrument. In such a tool, both the neutron source and the detector
are mounted in a pad shoe that is pressed against the borehole wall. Intermediate speed
(epithermal) neutrons can then accurately reflect the liquid-filled porosity. Log interpretation
charts are available to distinguish sandstone and dolomite porosity from limestone porosity.
Induction Resistivity Log
The induction log is designed for medium to deep investigation of formation conductivity. It is
probably the most commonly run logging tool. This tool consists of one or more coils that
generate an alternating electromagnetic field. The induced currents, which depend on the
conductivity of the formation, are detected by receiver coils. Multiple coils are used to focus the
measurement deep into the formation. This minimizes effects of the borehole, the invaded zone,
and nearby zones. The induction log works best when the borehole fluid is relatively resistive, as
would be the case with freshwater, oil-based muds or empty boreholes. Focused resistivity tools,
such as the Laterolog, are recommended for conductive borehole fluids, but are not suitable for
air-filled holes where air (or other gases) remains in the borehole at the time of logging.
With a measurement of formation resistivity, Rt, (the reciprocal of conductivity) and knowledge
of porosity, , from other logs or core measurements, formation water resistivity, Rw, the
cementation exponent, m, and the saturation exponent, n; an estimation of water saturation, S w,
may be made from the classic Archies equation.
R

Sw = w m
Rt

1/ n

(6.4)

Where significant amounts of conductive minerals, such as clay, are also present in the
formation, allowance must be made for their influence on the total resistivity. As with logging in
conventional holes, consideration must also be given to borehole geometry and bed thickness.

Mark of Schlumberger.
assuming a = 1

6-63

Chapter 6

Special Considerations

If a Dual Induction Focused Log is run in an air hole, the SP and focused curves are not
presented (because a conductive borehole liquid would be required) but the medium and deep
induction curves do an excellent job, due to the lack of invasion and reduced borehole fluid
effects.
High Resolution Density Log
Under ideal conditions, the density log can provide excellent measurements of porosity in a gasdrilled hole. This tool, which is firmly pressed against the borehole wall, emits a beam of
gamma rays into the rock, from a source, such as Cs137. These gamma rays interact with
electrons in the formation through Compton scattering. The resulting, lower-energy gamma rays
are sensed by two detectors above the source. Although the tool responds to electron density, for
most formation rocks the apparent bulk density is practically identical to the actual bulk density.
For a few rocks or minerals, such as sylvite and halite (and to a lesser extent gypsum, anhydrite,
and coal), small corrections are needed to arrive at true bulk density values. In addition to bulk
density, the tool will record an index of photoelectric absorption, P e. This value is useful in
determining formation lithology. Porosity is calculated from:
=

ma b
100
ma f

(6.5)

where:
......... porosity (percent),
ma ..... density of the rock matrix (g/cm3),
b ...... bulk density of the formation
(g/cm3), and,
f ....... density of the formation fluid
(g/cm3).
One of the difficulties in this formulation is selecting f. To illustrate the importance of selecting
the correct density for the formation fluid, consider the following example. If a gas zone is
drilled with an aqueous mud, f could be taken as1 g/cm3. If the same hole is air-drilled, a
possible assumption for f would be 0.5 g/cm3. Assuming b = 2.1 g/cm3 and ma = 2.64 g/cm3,
porosity would be calculated as:
=
or
=

6-64

2.64 2.1
= 33% for an aqueous mud
2.64 10
.
2.64 2.1
= 25% for an air drilled hole
2.64 0.5

Conditions that can cause difficulties with density log interpretation include rugose or caved
boreholes and lack of information on matrix or fluid density. Environmental corrections must be
made for the borehole size and the fluid in the well.
Temperature Log
Because cooling occurs as gas expands into the wellbore from the formation, a temperature log
can assist in pinpointing zones of gas entry. Estimates may be made of the percent contribution
from each interval. Temperature logging can also locate zones of lost circulation or fluid entry
into the formation.
Either absolute or differential temperature can be measured. A logging tool that measures
differential temperatures will quantify the temperature difference between one sensor and
another, located a short distance away. The differential temperature curve allows observation of
very small temperature changes without the need to frequently adjust the borehole temperature
curve scale. Temperature tools are usually combined with gamma ray tools for correlation and
readings are taken as the tool is run into the borehole.
Nuclear Magnetic Resonance Log
Although Nuclear Magnetic Resonance (NMR) logs are less commonly used in underbalanced
holes, they can provide useful information for liquid-filled formations. Borehole fluid will not
affect these measurements. Parameters such as effective porosity, initial water saturation, and
permeability (especially for sandstones) can be estimated without prior knowledge of lithology or
calibrations with cores. Other information, such as oil viscosity, pore size distribution, and
residual oil saturation, may also be estimated.
NMR tools evaluate the magnetic moments of hydrogen nuclei, abundantly present in both water
and oil. Magnets in the tool orient the magnetic moments of the hydrogen nuclei parallel to the
field. The magnitude of the magnetization is directly related to the number of hydrogen nuclei
and consequently the liquid-filled porosity. By disturbing the magnetic field with an RF pulse
from the tool, the rate of return of the nuclei to their original configuration can be measured.
Short relaxation times correspond to small pores, long relaxation times indicate large pores. The
variance in relaxation times can be related to pore size distribution, which can be correlated to
capillary pressure and permeability. Fluid flow should be not be permitted while these
measurements are made.
MWD
If an MWD (Measurement While Drilling) system is used, further evaluation may be made, while
drilling with integrated logging devices that sense natural radiation, resistivity and porosity. If
gas or gasified liquid is used as the drilling fluid, conventional mud pulse telemetry cannot be
used to transmit the signals. Alternatively, as described previously, other data transmission
systems, such as electromagnetic or hard-wired tools, may be considered.

6-65

Chapter 6

Special Considerations

Coring Underbalanced
Underbalanced coring has been carried out, to obtain relatively pristine samples and still
avoid formation damage. Reducing coring fluid invasion allows for careful determinations of
formation properties where wettability alteration has been minimized.
Permeability And Deliverability Assessments
Butler et al., 1996,53 emphasized the value of pressure monitoring for real-time evaluation of
reservoirs; to identify geologic anomalies (fractures, tight zones, pinchouts, discontinuities, water
zones ). Computer modeling may be used during drilling to evaluate sandface drawdown,
develop inflow performance relationships, and, in combination with mass balance monitoring,
indicate wellbore productivity. Effective monitoring of production rates permits real-time
decisions regarding changes in drilling depth, wellbore orientation, and overall section length.
In some cases, hori-zontal lengths have been reduced 75% from that planned when economic
production rates have been achieved, saving the operator significant drilling costs. 53 These
authors emphasized the importance of an adequate sampling interval (take measurements as
frequently as needed), to detect any and all significant fluid slugs.
Each drillpipe connection introduces an annular fluid slug into the system. Annular fluid
slugging occurs primarily during pump off periods. The severity of annular slugging depends on
fluid concentration in the injection mixture as well as the type of hydrocarbon produced and the
capacity of the well to flow under its own energy. If a well produces relatively small amounts of
gas, for example, and there is a relatively large fluid concentration, a definite fluid slug will form
in the bottomhole.53
If the underbalanced ROP is sufficiently high and bottoms up (along with the connection slugs)
are not being circulated out, controlled drilling should be considered.
Consider a low-energy gas well. Precise production data monitoring after a connection shows
decreased gas production followed by a significant gas peak. As the fluid slug, formed by
separation of injected fluids moves to surface, gas production is reduced. Eventually, the slug is
purged from the well, and the resultant compressed gas behind the slug expands at the surface
and manifests itself as an increase in production.
A well that is producing liquid hydrocarbons would continually exhibit fluid slugging due to
formation loading. Prolific fluid producers can actually kill themselves during times of pump
off. Restarting gas/fluid injection down drillpipe can be problematic under these circumstances.
If there is insufficient time for each fluid slug, containing increased concentrations of drilled
solids, to circulate through the system and back to surface before another connection is made, the
slugs can accumulate and create enough backpressure to overbalance the well. Penetration rate
must therefore be adequately controlled. If only one fluid slug is observed at the surface

Periodically circulate bottoms up before making a connection.

Return to SC Manual

6-66

between connections, the system is in equilibrium with injection, production and penetration rate.
If more than one slug occurs, the system is not in equilibrium. The effective rate of penetration is
defined by the overall pumping time necessary to purge a slug of fluid from a well.

References
1.

Willis, R.B.: Well Control Practices for Air-Drilled Holes, paper SPE 14497 presented at
the 1995 Eastern Regional Meeting, Morgantown, WV, November 6-8.

2.

Collins, G.J.: Proper Planning Improves Flowdrilling, World Oil (October 1994) 43-50.

3.

Eresman, D.: Underbalanced Drilling - A Regulatory Perspective, paper 93-306


presented at the 1993 CADE/CAODC Spring Drilling Conference, Calgary, Alberta,
Canada, April 24-16.

4.

Recommended Practices for Underbalanced Drilling, Energy Resources Conservation


Board Interim Directive ID 94-3 (July 1994) Calgary, Alberta, Canada.

5.

Moore, B.: The Regulation of Underbalanced Drilling in the UK Sector, paper presented
at the 1995 International Underbalanced Drilling Conference, Amsterdam, The
Netherlands.

6.

Yost, A.B.: Horizontal Gas Well Promises More Devonian Production, The American
Oil and Gas Reporter (July 1988).

7.

Millheim, K.K. and Warren, T.M.: Side Cutting Characteristics of Rock Bits and
Stabilizers While Drilling, paper SPE 7518 presented at the 1978 SPE Annual Fall
Technical Conference and Exhibition, Houston, TX, October 1-3.

8.

Millheim, K.K.: Hard-Formation Directional Drilling Calls for Special Care, Oil and
Gas J. (February 12, 1979).

9.

Wilson, G.E.: Air Drilling and Crooked Hole Problems, paper SPE 9529 presented at the
1980 SPE Eastern Regional Meeting, Morgantown, WV, November 5-7.

10.

Millheim, K.K. and Apostal, M.C.: The Effect of Bottom-Hole Assembly Dynamics on
the Trajectory of a Bit, paper SPE 9222 presented at the 1980 SPE Annual Technical
Conference and Exhibition, Dallas, TX.

11.

Shale, L.: Development of Air Drilling Motor Holds Promise for Specialized Directional
Drilling Applications, paper SPE 22564 presented at the 1991 SPE Annual Technical
Conference and Exhibition, Dallas, TX.

12.

Lyons, W.C.: Air and Gas Drilling Manual, Gulf Publishing Company, Houston, Texas
(1984) 46-25.

6-67

Chapter 6

6-68

Special Considerations

13.

Shale, L.: Underbalanced Drilling: Formation Damage Control During High-Angle or


Horizontal Drilling, paper SPE 27351 presented at the 1994 SPE International
Symposium on Formation Damage Control, Lafayette, LA, February 7-10.

14.

Dreiling, T., McClelland, M.L. and Bilyeu, B.: Horizontal and High Angle Air Drilling in
the San Juan Basin, New Mexico, The Brief, Murphy Publishing Inc. (June 1996).

15.

Lyons, W.C., Miska, S.Z. and P.W. Johnson: Downhole Pneumatic Turbine Motor:
Testing and Simulation Results, paper SPE 18040 presented at the 1988 SPE Annual
Technical Conference and Exhibition, Houston, TX, October 2-5.

16.

Roy, R. and Hay, R.: Measuring Downhole Annular Pressure While Drilling for
Optimization of Underbalanced Drilling, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

17.

Soulier, L. and Lemaitre, M.: E.M. MWD Data Transmission Status and Perspectives,
paper SPE/IADC 25686 presented at the 1993 SPE/IADC Drilling Conference,
Amsterdam, The Netherlands, February 23-25.

18.

Shale, L. and Moberley, G.T.: Development of a Cartridge Data Transmission System for
Use with Air Drilling Motor, paper SPE/IADC 23907 presented at the 1992 SPE/IADC
Drilling Conference, New Orleans, February 18-21.

19.

Shale, L. and Curry, D.A.: Drilling a Horizontal Well Using Air/Foam Techniques, paper
OTC 7355 presented at the 1993 Annual Offshore Technology Conference, Houston, TX.

20.

Shale, L.: Underbalanced Drilling Equipment and Techniques, PD (1995) 65, Drilling
Technology, ASME (1995).

21.

Lattimore, G.M., Carden, R.S. and Fisher, T.: Grand Canyon Directional Drilling and
Waterline Project, paper SPE/IADC 16169 presented at the 1987 SPE/IADC Drilling
Conference, New Orleans, LA, March 15-18.

22.

Angel, R.R.: Volume Requirements for Air or Gas Drilling, Pet Trans., AIME (1957)
210, 325-330.

23.

Guo, B., Miska, S.Z. and Lee, R.: Volume Requirements for Directional Air Drilling,
paper IADC/SPE 27510 presented at the 1994 IADC/SPE Drilling Conference, Dallas, TX,
February 15-18.

24.

Guo, B., Miska, S.Z. and Hareland, G.: A Simple Approach to Determination of
Bottomhole Pressure in Directional Foam Drilling, paper presented at the 1995 ASME
Energy and Environmental Expo 95, Houston, TX, January 29-February 1.

25.

Guo, B., Hareland, G. and Rajtar, J.M.: Computer Simulation Predicts Unfavorable Mud
Rate and Optimum Air Injection Rate for Aerated Mud Drilling, paper SPE 26892
presented at the 1993 SPE Eastern Regional Conference and Exhibition, Pittsburgh, PA.

26.

Shale, L.: Underbalanced Drilling With Air Offers Many Pluses, Oil and Gas J. (June 26,
1995) 33-39.

27.

Carden, R.S.: Air Drilling Has Some Pluses for Horizontal Wells, Oil and Gas J. (April
8, 1991) 76-78.

28.

Finger, J.T.: Investigation of Percussion Drills for Geothermal Applications, JPT


(December 1984) 2128-2136.

29.

Pratt, C.A.: Modifications To and Experience With Air-Percussion Drilling, SPEDE


(December 1989) 315-320.

30.

Whiteley, M.C. and England, W.P.: Air Drilling Operations Improved by PercussionBit/Hammer-Tool Tandem, SPEDE (October 1986) 377-382.

31.

Reinsvold, C.H., Clement, J., Oliver, M., Witt, C. and Crockett, J.: Diamond-Enhanced
Hammer Bits Reduce Cost per Foot in the Arkoma and Appalachian Basins, paper
IADC/SPE 17185 presented at the 1988 IADC/SPE Drilling Conference, Dallas, TX,
February 28 - March 2.

32.

Teale, R. : The Concept of Specific Energy in Rock Drilling, Int. J. Rock Mech. and
Mining Sci. (1965) 2, 57-73.

33.

Simon, R.: Energy Balance in Rock Drilling, SPEJ (December 1963) 298-306.

34.

Sheffield, J.S. and Sitzman, J.J.: Air Drilling Practices in the Mid-continent and Rocky
Mountain Areas, paper SPE/IADC 13490 presented at 1985 SPE/IADC Drilling
Conference, New Orleans, LA.

35.

Russell, B.A.: How Surface Hole Drilling Performance Was Improved 65%, paper
SPE/IADC 25766 presented at the 1993 SPE/IADC Drilling Conference, Amsterdam, The
Netherlands, February 23-25.

36.

Teel, M.E.: Whats Happening in Drilling-Pseudo-underbalanced Drilling and Beyond,


World Oil (April 1995) 29.

37.

Hannegan, D.M. and Bourgoyne, A.T.: Underbalanced Drilling - Rotating Control Head
Technology Increasing in Importance, paper presented at the 1995 1st International
Underbalanced Drilling Conference, The Hague, The Netherlands, October 2-4.

6-69

Chapter 6

6-70

Special Considerations

38.

Bloys, B., Brown, J.D. and Tarr, B.A.: Drilling Safely and Economically in Carbonates:
Collective Experience of ARCO, BP and Mobil, paper presented at the 1994 IADC Well
Control Conference for the Asia/Pacific Region, Singapore, December 1-2.

39.

Leising, L.J. and Newman, K.R.: Coiled-Tubing Drilling, SPEDC (December 1993) 227232.

40.

Madigan, J.: Applications of Slimhole Technology - A Service Company Perspective,


presented at the 1993 DEA 44/67 Forum on Horizontal and Slimhole/Coiled Tubing
Technology.

41.

Dowell Schlumberger:
1984.

42.

Medley, G.H., Maurer, W.C. and Garkasi, A.Y.: Use of Hollow Glass Spheres for
Underbalanced Drilling Fluids, paper SPE 30500 presented at the 1995 SPE Annual
Technical Conference and Exhibition, Dallas, TX, October 22-25.

43.

de Rozires, J. and Ferrire, R. F.: Foamed Cement Characterization Under Downhole


Conditions and Its Impact on Job Design, paper IADC/SPE 19935 presented at the 1990
IADC/SPE Drilling Conference, February 27-March 2.

44.

Reiley, R.H., Black, J.W., Stagg, T.O., Walters, D.A. and Atol, G.R.: Improving Liner
Cementing in High-Angle/Horizontal Wells, World Oils Handbook of Horizontal
Drilling and Completion Technology, Gulf Publishing Company, Houston, TX (1991) 5458.

45.

Wilson, M.A.: Cementing Horizontal Wells in Preparation for Stimulation, World Oils
Handbook of Horizontal Drilling and Completion Technology, Gulf Publishing Company,
Houston, TX (1991).

46.

Ryan, D.F., Browne, S.V. and Burnham, M.P.: Mud Clean-Up in Horizontal Wells: A
Major Joint Industry Study, paper SPE 30528 presented at the 1995 SPE Annual
Technical Conference and Exhibition, Dallas, TX, October 22-25.

47.

Chmilowski, W. and Kondratoff, L.B.: Foamed Cement for Squeeze Cementing LowPressure, Highly Permeable Reservoirs: Design and Evaluation, SPEDE (December
1992) 284-290.

48.

Purvis, D.L. and Smith, D.D.: Real-Time Monitoring Provides Insight Into Flow
Dynamics During Foam Cementing, SPEDC (June 1994) 124-132.

49.

Halliburton Log Interpretation Charts, Halliburton Energy Services, Houston, TX (1994).

Cementing Technology, Nova Communications Ltd., London,

50.

Log Interpretation Principles/Applications, Schlumberger Educational Services, Houston,


TX (1991).

51.

Wireling Services Catalog, Schlumberger Educational Services, Houston, TX (1991).

52.

Western Atlas Services Catalog, Western Atlas, Atlas Wireline Services, Houston, TX
(1984).

53.

Butler, S.D., Rashid, A.U. and Teichrob, R.R.: Monitoring Downhole Pressures and Flow
Rates Critical for Underbalanced Drilling, Oil and Gas J. (September 16, 1996) 31-39.

6-71

7.1

CASE STUDIES

Introduction

There are numerous published and unpublished case studies demonstrating the feasibility and
practice of underbalanced drilling. This chapter selects some illustrative examples. The case
studies provided include.
 Case Study 1
Controlling Bottomhole Pressure
This case study emphasizes one of the most important elements in underbalanced drilling.
This is dealing with reservoir influx. An example vertical well is shown, from Saponja,
1996.1 The discussion is supplemented by methods for forecasting and the value of
monitoring bottomhole pressure.
 Case Study 2
Barrolka 3
This case study was provided courtesy of Santos Ltd. It demonstrates successful
implementation of air drilling.
 Case Study 3
Swan Lake-1 ST
This is another air drilling example, also supplied courtesy of Santos Ltd. This well showed
production ten times that of conventionally drilled wells. Sloughing difficulties were
encountered. However, drilling was finally stopped because it was not advisable to continue
with the high gas flow rate.
 Case Study 4
Karwin-1 ST
This air drilling example was provided courtesy of Santos Ltd. While sloughing was a
problem, air drilling proved to be more successful than conventional techniques.
 Case Study 5
Unloading A Hole From The Bottom
This is an hypothetical example showing how a well can be unloaded from the bottom after
setting and cementing casing. It is included to demonstrate the diversity of the applications

7-1

Chapter 7

Case Studies

for air drilling and to emphasize how an understanding of wellbore hydraulics can lead to
cost effective solutions. The authors are grateful to R. Graham (Reuben L. Graham, Inc.) for
this demonstration case study.
 Case Study 6
Husky Wainwright 15B-31-44-4W4M
Creative engineering, with extensive preplanning and safety evaluations, led to successful
drilling of this well, with water, air and nitrogen, and a concentric string gasifying
configuration, into a low pressure reservoir. This is a clearly and comprehensively
documented case (Teichrob, 1994).2
 Case Study 7
Deep, High Pressured Re-entry
Overbalanced re-entry had failed to access a target gas sand at 16,500 feet, in three previous
sidetrack attempts. Casing had collapsed at 15,450 feet. Flow-drilling was ultimately
successful (Stone and Cress, 1997).3
 Case Study 8
Cotton Valley Lime
Using an RBOP and determining an acceptable trip rate (to prevent surging and swabbing
pressures from exceeding critical downhole limits) substantially reduced lost time in this well
(Stone and Cress, 1997).3
 Case Study 9
Depleted Fractured Carbonate
A target carbonate was overlain by overpressured shales and underlain by normally pressured
permeable water sands. Significant lost circulation problems had been encountered in the
past. Flowdrilling reduced the time to drill the section from about 60 days to 2 days (Stone
and Cress, 1997).3 Although not included here, two additional papers, by Joseph, 1995a,b,4,5
are important reading for anyone considering high pressure drilling.
 Case Study 10
Dalen-2
A considerable amount has been published about the coiled tubing drilling of this well in the
Netherlands. It demonstrates the value of effective planning and coiled tubing technology.
The case study is accompanied by a further description of some of the methods available for
underbalanced completions. Two of the papers des-cribing the procedures in this well are by
Adam and Berry, 1995,6 and Wang et al., 1995.7
 Case Study 11
Cementing The Dakota Formation
Purvis and Smith, 1994,8 provided a description of some of the methodologies for
underbalanced cementing (whether used with under-balanced drilling or not).

7-2

 Case Study 12
The Friction Controlled Regime
Saponja, 1995,9 clearly demonstrated the importance of effectively manipulating flow rate of
each phase, during multiphase, underbalanced drilling.

7.2

Case Study 1
Controlling Bottomhole Pressure

Reference
Saponja, J.: Challenges With Jointed Pipe Underbalanced Operations, paper SPE 37066
presented at the 1996 SPE International Conference on Horizontal Well Technology, Calgary,
Alberta, Canada, November 18-20.
Background
Saponja, 1996,1 described a vertical well, drilled underbalanced with nitrogen and water, at 706
cfm and 3.8 BPM, respectively. Gas inflow started at 6970 feet and rapidly increased to 2.3
MMcf/D. Concurrently, the bottomhole pressure dropped from 2030 to 1305 psi, since this
system was in the hydrostatically-dominated regime. Drilling continued but it was inefficient
since the formation gas inflow would have allowed lower nitrogen rates. Figure 7-1 shows how
reducing the nitrogen injection rate by 75 percent, and increasing the water injection rate,
impacted the pressure behavior.
This example clearly indicates the importance of forecasting a reservoirs inflow performance.
In this instance, bottomhole gas inflow would be reduced by almost 60 percent if the
bottomhole pressure were maintained at 2030 psi. How should formation inflow be
anticipated? Offset well information can be used as can real-time monitoring. Pre-drilling
simul-ation can be of great value.
Before Drilling
Develop a typical inflow performance curve (Vogel IPR curve, Figure 7-2). This will allow for
sizing of drillpipe and surface facilities. For example, separator sizing, to avoid blow-through,
depends on the peak gas rate (refer, for example, to MacDougall, 1991).10
Most operators specify hole and tubing size prior to deciding on specific fluids and operational
protocols. Hole and tubular size may prevent or at least severely limit successful drilling and
unloading of solids and produced fluids. This is more of a problem in re-entry situations where
the hole and casing sizes often require compromises.
During Drilling
Reservoir fluid inflow is particularly important for horizontal wells that penetrate long sections
of productive pay or intersect natural fractures. As Saponja, 1995,9 stated:

7-3

Chapter 7

Case Studies

Maximizing use of the natural energy available from the reservoir can reduce circulation system
requirements, such as nitrogen injection rates, but reservoir liquid inflow during connections can
result in large liquid slugs, large pressure spikes and problems with regaining circulation (lost
circulation). In highly permeable reservoirs, inflow of liquid during connections can kill the
well, making circulation impossible to regain without imposing overbalanced pressures.
Connection time must be minimized to reduce the amount of liquid inflow from the reservoir.
For horizontal wells, as more horizontal length is drilled, it can become progressively more
difficult to regain circulation after a connection. Saponja, 1995,11 has discussed a procedure,
referred to as Annular Pre-Charging, where the annular pressure and GLR are increased prior to a
connection, reducing the formation drawdown, as well as the liquid inflow. This protocol also
reduces the liquid volume in the wellbore and inhibits liquid slugging and inertial acceleration
effects. Operationally, the procedure requires precise timing of the annulus closure followed by a
period of continued gas injection and caution to avoid causing overbalanced conditions. In
horizontal wells particularly, different inflow conditions along the length of the well must be
considered; for example to prevent excessive inflow at the heel and/or fluid loss at the toe,
because of frictional effects (Comeau, 1995).12
Comeau, 1995,12 discussed using a Hydrostatic Control Valve during drilling with gasified
liquids with a parasite string. The valve maintains an hydrostatic head while making a connection
and prevents the drillpipe from going on vacuum. It also prevents backflow up the drillpipe
during tripping and making connections.

7-4

Chapter 7

Case Studies

Annular Bottomhole Pressure (psi)

3500
Corrected Circulation Rates:
Water: 6.3 BPM
Nitrogen: 177 cfm
Formation Gas: 989 cfm

3000

Circulation Rates:
Water: 6.3 BPM
Nitrogen: 0 cfm

2500

2000
Start Nitrogen Injection
Circulation Rates:
Water: 3.8 BPM
Nitrogen: 706 cfm

1500

1000
Formation Gas Inflow
2.3 MMcf/D

Circulation Rates:
Water: 3.8 BPM
Nitrogen: 706 cfm
Formation Gas: 2.3 MMcf/D

500

0
6300

6400

6500

6600

6700

6800

6900

7000

7100

7200

Depth (feet)

Figure 7-1.

Controlling bottomhole pressure in a vertical well drilled with nitrogen and water
(after Saponja, 1995 9).

Bottomhole Pressure/Reservoir Pressure

1
0.9
0.8
0.7
0.6

Higher Bottomhole Pressures Lead To


A Lower Rate of Flow From The Reservoir

0.5
0.4
0.3
0.2
Lower Bottomhole Pressures Lead To
A Higher Rate of Flow From The Reservoir

0.1
0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Flow Rate/Maximum Flow Rate

Figure 7-2.

7-6

A typical representation of Inflow Performance Relationships showing the


interrelationship between bottomhole pressure and flow rate (modified from Saponja,
9
1995 ). For additional information, refer to Brown and Lea, 1985,13 or Vogel, 1968.14

Numerical Predictions
Butler et al., 1996,15 stated:
 Multiphase flow modeling software should incorporate a choice of several distinct fluid
flow models to commingle injected and produced fluids, ultimately defining flow regimes
and velocities.
 It is extremely difficult to predetermine the inflow from a hori-zontal wellbore during
underbalanced drilling operations. Therefore, a range of inflow rates is designed into the
program to ensure successful comple-tion. Typically, inflow at 50%, 100%, and 150% of
the anticipated production rate is modeled and evaluated. It is critical to ensure that the
pressure drop calculations across the horizontal section conform to equilibrium of
production, drawdown, and reasonable backpressure held at the surface.
 A test matrix, comprised of differing flow rates at specified drawdown is performed. The
objective is to construct an operational envelope under which the probability of success is
maximized. The envelope should contain enough end points in pressure drawdown and
corresponding flow rates to encompass actual well site conditions. This analysis ensures
effective contingency planning is in place should reservoir inflow markedly differ from
that anticipated.
 As drilling proceeds, reservoir inflow will likely change. On site engineering personnel may
have to respond immediately with an injection blend ratio change to maintain optimum
underbalanced drilling conditions. Injection blend changes are based on updated multiphase
flow model results. Computer modeling may be used during drilling to evaluate sandface
drawdown, develop inflow performance relation-ships and, in combination with mass balance
monitoring, provide wellbore productivity information.
 Due to the nature of underbalanced drilling, the well control during underbalanced drilling is
changed from primary pressure control by the hydrostatic pressure of drilling fluids to flow
control where the bottomhole pressure and thereby the production during drilling is
monitored and controlled. The critical parameters are the gas injection rate (if applicable),
well head pressure, bottomhole pressure, and reservoir drawdown. It is essential that the
operation is in a stable operating range where the bottomhole pressure is not sensitive to the
changes in normal control parameters, especially gas injection rate and reservoir drawdown.
Because of the added level of complexity, underbalanced drilling numerical simulators have been
developed, affording organized pre-planning. For example, Misselbrook et al., 1991,16 discussed
a simulator for evaluating circulation conditions for multiple phases and for forecasting coiled
tubing stress states. These authors gave an example:
Computer modeling allows the user to calculate whether or not a job is possible. It is also an
economical means of optimizing particular job programs. For example, consider a sand cleanout
in a deviated well with a water sensitive formation and low bottomhole pressure. An effective
cleanout will require that the return velocity in the annulus be above certain minimums in order
for sand transport to take place. As the pump rate is increased to accomplish this, it is possible
7-7

Chapter 7

Case Studies

that the combination of hydrostatic and friction pressure may become greater than the bottomhole
pressure can support and the returns will then be lost to the formation. Knowing this, the model
can be used to simulate different approaches. Foam or various ratios of liquid and nitrogen slugs
may be tried to reduce the hydrostatic pressure and smaller diameter coiled tubing can be
modeled to ascertain the effects of reduced annular flow restriction.
Wang et al., 1995,7 described a dynamic (as opposed to static or steady-state simulators which
represent constant flow rates for drilling fluids, gas injection and reservoir production, with no
significant drillstring movement) underbalanced drilling simul-ator. It incorporated transient
multiphase hydraulics in a realistic well geometry and well-reservoir interaction. Reservoir
influx (matrix and fractured) was considered.
Downhole Monitoring
Operators can use downhole monitoring to comprehend the interactive processes which occur
during underbalanced drilling. Adjustments made to surface parameters can be monitored so that
the operator can make procedural modifications. Simulations can also be run during planning
stages so that correctly sized tubulars and surface equipment, such as separators, can be used.
High frequency data acquisition can allow identification of downhole pressure fluctuations.
Figure 7-3, from Wilson, 1995,17 is an example of pressure fluctuations discernible with rapid
data acquisition, showing:
(1) ................................................ pipe purge,
(2) ...................................... connection made,
(3), (4)...................... an annular slug forming
and traveling to the surface, and,
(5) ................re-establishment of equilibrium
with injection fluids, gases and produced gas.
Figures 7-4 and 7-5, from Wilson, 1995,17 show the increase in gas production rates and the
variations in pressure with increasing openhole length. Intersection of fractures was evidenced.
Figure 7-5 shows that while the casing shoe pressure remained relatively constant, the bit
pressure increased constantly until liquid and nitrogen injection rate adjustments were required to
maintain an underbalanced condition.
Various pressure monitoring systems have been described in this manual. Roy and Hay, 1995,18
described a system for measuring pressure while drilling (PWD) and provided example data
showing the potential for deviating from underbalanced conditions. This system for pressure
measurement monitors the annulus pressure with a sensor that is located in a sub below the
MWD. The data is either recorded or, when appropriate, it can be transmitted in real-time to the
surface using MWD protocols. The well described was drilled with drillpipe N2 injection, with
misting to lubricate the air motor. Data were recorded only. Real-time manipulations of flow to
maintain underbalanced conditions were not possible since the pressure data were recorded only.
Figure 7-6 shows the measured pressures. It indicates that much of the well was drilled
7-8

underbalanced although overbalance was much more common during the latter part of the
drilling program. This figure demonstrates the occurrence of overbalance as inflowing fluid fills
the annulus, particularly during connections when nitrogen was not injected.
3.5

Production Rate (MMcf/D)

2.5
3

1.5
2

4
5

0.5

0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Time (hours)

Figure 7-3.

Variation of production rate, as measured at five minute sampling inter-vals, showing


17
characteristic pressure signatures (after Wilson, 1995 ).

7-9

Chapter 7

Case Studies

3650

7000

Encountered
Fracture

3750

Gas Rate (cfm)

5000
Gas Rate
True Vertical Depth

3800

4000
Encountered
Fracture

3000

3850

2000

3900

1000

3950

0
3500

4000

4500

5000

5500

6000

True Vertical Depth (feet)

3700

6000

4000
6500

Drilled Depth (feet)


Figure 7-4.

The variation of gas production with drilling and the indications of intersecting fractures (after, Wilson,
1995 17).
1200

Bottomhole Pressure (psi)

1000

800
Water: 57 gpm
Nitrogen: 950 scfm

Water: 40 gpm
Nitrogen: 950 scfm

600
Bottomhole Pressure at the Casing Shoe
Bottomhole Pressure at the Bit

400

Formation Pressure

200

0
0

500

1000

1500

2000

Openhole Length (feet)


Figure 7-5.

7-10

The variation of bottomhole pressure with drilling and demonstration of the maintenance of
underbalanced pressure by manipulating injection rates. (after Wilson, 199517).

2500

2200
Bottomhole Annulus Pressure
Reservoir Pore Pressure

2100

Annulus Pressure (psig)

2000
1900
1800
1700
1600
1500
1400
1300
1200
0

10

12

14

16

18

20

22

Time (hours)

Figure 7-6.

Downhole annular pressure during drilling (after Roy and Hay, 1995 18).

Summary
Saponja, 1996,1 stated:
Underbalanced drilling has been unsuccessful in some reservoirs wells believed to be drilled
underbalanced were found to have formation damage or positive skin. However, review of
operating procedures and circulating systems revealed that overbalance pressures occurred
during drillstring connections and incompatible drilling fluids were used.
 A properly designed circulating system does not allow extreme reservoir inflow rates or high
annular flowing pressures at the surface.
 Multi-phase flow can allow drilling at any pressure gradient. Computer models simulate the
multi-phase flowing system, including reservoir inflow and pressure-volume-temperature
relationships. Most available models are static and interpret steady state flow. The dynamic
simulators that have been developed consider the impact of starting, stopping and changing
circulation, fluid effects during connections and drillstring movement.
 At the pre-planning phase, multi-phase flow modeling is required to determine circulation
system parameters. Injection fluids must be analyzed in conjunction with produced reservoir
fluids, at a variety of conditions, to determine bottomhole pressure operating limits.
Operating limits are determined prior to execution so that contingency plans for the
circulating system are in place. Actual well conditions, reservoir pressure variations, and

7-11

Chapter 7

Case Studies

The reservoir inflow performance will determine the optimal circulating system.

7.3

Case Study 2
Barrolka 3

Reference
This unpublished case study was provided courtesy of Santos Ltd.
Objective
The primary objective was to air drill the Toolachee 85.3 sand to minimize formation damage
and to evaluate the undamaged potential of this reservoir. This was a gas exploration well.
Figure 7-7 is a well schematic.
Operational Summary
Sixteen-inch conductor pipe was set at forty feet below ground level.
12-inch hole was drilled to 1754 feet RT, without problems. A mill-tooth bit was run on a
slick BHA and pre-hydrated bentonite spud mud was used. The bit drilled to 1754 feet RT in
16.0 hours, at an average ROP of 109.6 ft/hr.
9 5/8-inch casing was run and cemented to the surface, with the shoe at 1748 feet.
8-inch hole was drilled with few problems. A new, milled-tooth bit was run in the hole on
a pendulum BHA, with stabilizers at 45 feet and 75 feet and jars located two drill collars
below the heavy-weight drillpipe. A leakoff test was performed after drilling out the shoe
track with water. Breakdown occurred at 14.9 ppg (0.77 psi/ft).
The hole was displaced to 8.9 ppg KCl/PHPA mud. Drilling continued through the CadnaOwie, Murta and Namur formations to a depth of 5188 feet, at which point the bit was pulled
due to increased torque. The bit drilled a total of 3434 feet in 45.5 hours, at an average ROP
of 75.5 ft/hr.

7-12

Formation

Depth (feet)
0

1000

2000

3000

4000

5000

Cadna-Owie
Murta
Namur
Westbourne

6000

Birkhead
Hutton

7000
Nappamerri

8000
Toolachee
Patchawarra

9000

Figure 7-7.

Formations drilled and the casing program for Barrolka 3.

 A new 8-inch bit was run in the hole on a pendulum BHA assembly, similar to that used
before. The bit was reamed/washed to bottom from 5063 feet and drilling then continued
through the Namur, Westbourne, Adori, Birkhead and into the Hutton formation, to 6645
feet. The mud weight had been gradually increased with KCl to 9.4 ppg by 6600 feet. The bit
was pulled due to an increase in torque. Tight spots were encountered on the way out and
these were worked (from 6000 to 5900 feet and from 5500 to 5450 feet). This bit drilled a
total of 1457 feet in 52 hours, at an average ROP of 28.0 ft/hr.
 A new bit, on a similar assembly, was run into the hole. Reaming was required from 5754 to
5980 feet and from 6252 to 6645 feet. At 7129 feet, a 700 psi pressure drop was observed.
After pressure testing surface equipment, the rig pulled out of the hole, looking for a washout
in the drillstring. This bit drilled 484 feet in 32.5 hours, at an average ROP of 14.9 ft/hr. The
pressure drop was due to a lost nozzle.

7-13

Chapter 7

Case Studies

 The bit used to drill to 6645 feet was run back into the hole with a junk sub. The hole was
worked and drilled to 7132 feet and the lost nozzle was recovered.
 A new 8-inch bit was run into the hole on a BHA with a junk sub, near-bit reamer and a
stabilizer at 60 feet. It was necessary to ream hole from 4971 to 5501 feet and from 5739 to
7132 feet. Drilling continued to 7150 feet, at which time the bit was pulled because of
concern over having reamed an excessive amount of hole. The average ROP over the 18 feet
drilled was 18 ft/hr.
 A new bit was run in on a pendulum assembly and drilling continued through the Nappamerri
formation to 7993 feet. The bit was pulled due to a decrease in the penetration rate. 843 feet
had been drilled, at an average rate of 15.5 ft/hr.
 A new bit was run into the hole on a similar pendulum BHA. The hole was reamed from
7891 to 7993 feet. Drilling continued to 8420 feet, with the top of the Toolachee formation at
8298 feet (casing depth).
 Logging showed enlarged hole at 8406 feet.
 Seven-inch casing was set and cemented. After waiting on cement, the tubing spool and 7
1/16-inch BOPs were installed and successfully pressure and function tested to a high of
3000 psi and a low of 200 psi.
 The cement and shoe track were drilled out with mud. A new, six-inch bit was used on a
slick assembly (one drill collar, two float subs, 12 drill collars and one set of drilling jars).
The bit drilled the cement, float shoe and collar and three feet of new hole using 9.4 ppg
KCl/POLY/PHPA drilling fluid to 8423 feet. A pressure integrity test was performed. The
formation held at an equivalent mud weight of 14.0 ppg (0.73 psi/ft). The bit was tripped and
a bond log was run. The actual cement top was logged at 4100 feet, with the casing shoe at
8418 feet.
 As air drilling was a new procedure in this region, before any air drilling started, all
wellsite personnel were trained on air drilling procedures and safety aspects. There were
eight, one-half hour sessions.
 The well was rigged up to convert to air drilling operations. The same six-inch bit was rerun.
The drilling fluid was displaced with water and the water was displaced by air/mist. Air
drilling the six-inch hole began at 8423 feet RT.
 The six-inch hole was drilled from 8423 to 8481 feet in 3 hours, when the flow from the
blooie line ignited. Open flow tests were inconclusive. Drilling contin-ued to 8488 feet and a
second flow test was performed. The results were inconclusive. At TD (8531 feet), the well
was flowed through a test choke manifold and successfully flow tested.
 Because of wet conditions on location, the crane needed for the snubbing unit could not be
used. The drillstring was stripped out of the well using the primary jet and the rotating head
rubber, without incident. 108 feet were drilled in 5 hours, at an average ROP of 21.6 ft/hr.
 During all flow testing and drilling operations, exposed coals appeared to be sloughing.
Several attempts were made to stabilize the formation. The entire Toolachee 85.3 sand

7-14

reservoir and approximately ten feet of the lower coal seam were drilled. There were hole
cleaning problems due to sloughing, during the air/mist operations. Future air drilled wells
might use lower air rates to prevent formation erosion and switch to higher air rates if hole
cleaning becomes more difficult. Water rates will be minimized in the future.
The well was logged under lubricator and wireline BOPs.
A hydraulic set retrievable production packer with blanked off tailpipe on 2 3/8-inch tubing
was snubbed in the well and set at 8340 feet RT. Drilling operations lasted for 34.44 days
from spud.
Highlights of the Air Drilling Program
The air drilling operations went ex-tremely well, particularly considering that this was the
initial well in the project.
Time (days)
0.0

10.0

20.0

30.0

40.0

1000
Run 9 5/8-inch Casing and
Air Drilling Training for Crews

2000

Planned
Actual

3000

Depth (feet)

4000

5000

6000
Fish for Nozzle and
Ream Undergauge Hole

7000

8000

9000

Run 7-inch Production Casing


and Rig Up to Air Drill
Air Drill and Snub
in Completion

10000

Figure 7-8.

Drilling schedule (Barrolka 3).

7-15

Chapter 7

Case Studies

During air drilling, there were several gas shows and open flow tests were performed. The
air drilling eliminated the cost of conventional openhole testing.
It is difficult to compare rates of penetration between mud and air drilling. Some comparison
is possible from Table 7-1.
Pre-drilling safety and training were effective.
The final pressure test showed subs-tantial production potential.
Although sloughing was a continual problem, air drilling in the coal seams may have
identified supplementary gas reservoirs. This has been the case in certain coalbed methane
reservoirs in the United States.

7.4

Case Study 3
Swan Lake 1 ST

Reference
This unpublished case study was provided courtesy of Santos Ltd.
Objective
The primary objective was to sidetrack this well to test the Toolachee, Epsilon, Patchawarra and
Tirrawarra formations. The proposed total depth was 10,092 feet. The formations had been
tested in the original well, where DSTs indicated a combined rate of 0.83 MMscf/D.
A permanent packer was set at 9622 feet with 3-inch tubing. The tubing was pulled, the packer
was plugged off, and a retrievable whipstock was set at 8195 feet. The sidetrack was to be drilled
using air/mist in order to minimize formation damage and to identify productive zones while
drilling.
Operational Summary
Pressure was bled to 100 psi and 9.94 ppg brine was pumped down the tubing to kill the gas
flow. A BPV was installed in the tubing hanger and the tree was removed. The BOP stack
was nippled up and pressure tested. The tubing was pulled.

7-16

Table 7-1.

Drilling Summary

Formation

Depth In
(feet)

Depth Out
(feet)

Feet Drilled

Hours

Cumulative
Hours

Toolachee

8420

8423

0.5

233.0

Toolachee

8423

8481

58

3.0

236.0

Patchawarra

8481

8531

50

2.0

238.0

ROP
(ft/hr)

WOB
(lbf)

Standpipe
Pressure
(psi)

Mud Weight
(ppg)

Annular
Velocity
(ft/s)

Jet Velocity
(ft/s)

6.0

15,000

1650

9.4

416

47

19.3

15000

600

~0

3371

385

25.0

15,000-17,000

600

~0

3371

385

 A DR latch plug was run on tubing and set at 9614 feet. The casing and packer, with the DR
plug, were pressure tested to 3000 psi.
 Wellhead equipment was installed and pressure tested.
 A six-inch bit and casing scraper were run on a BHA (twelve 4-inch drill collars and 3inch drillpipe) to the sidetrack depth. The bottom collar was plugged with scale. The BHA
had to be pulled and a casing scraper was run. The hole was displaced with 9 ppg KCl and
the tubing was pulled.
 A retrievable whipstock assembly was set at 8195 feet. A casing window was milled from
8178 to 8186 feet and new formation was drilled to 8193 feet. The casing seat was tested to
13.7 ppg EMW.
 A six-inch hole was drilled with mud from 8193 to 8206 feet. The drilling mud was
displaced with water and the water was displaced with air/mist. Six-inch diameter air drilling
began at 8206 feet.
 The hole was drilled to 8329 feet without difficulty. An openhole flow test was carried out.
183 Mcf/D was measured for three hours, through a -inch choke, at 18 psi.
 Air drilling continued without difficulty to 8452 feet, at which point the standpipe pressure
increased by 600 psi and the mud loggers had an increase in gas. The average ROP was 20
ft/hr. Flow testing indicated 4.74 MMcf/D through a -inch choke, at 857 psi.
 Air drilling continued to 8505 feet. A third flow test indicated 8.20 MMcf/D, through a inch choke at 1535 psi.
 Air drilling continued to 8578 feet when the standpipe pressure increased to 1000 psi for 15
minutes. The pipe was free but it was difficult to get back to bottom. The pressure fell to 710
psi and the drillstring was worked several times through the tight hole. After making a
connection at 8578 feet, the pipe became stuck without reciprocation or rotation. A single had

7-17

Chapter 7

Case Studies

to be laid down and the kelly reconnected. Air pressure was 750 psi and appeared to be
circulating. The drillstring was worked, without the jars hitting, and air was circulated for
two hours to clean the hole. The drillstring started to move down and the jars started hitting.
The drillstring started rotating. White smoke, indicating water, appeared at the blooie line.
The string was worked to bottom while the well unloaded fluid.
 The drillstring was pulled inside of the casing window. Singles were run back into the hole,
while circulating air and mist. The blooie line discharge cleared up quickly. The hole was
reamed and washed 40 feet to bottom.
 After circulating for 30 minutes to clean the hole, the connection was made and air drilling
continued to 8671 feet. A flow test was conducted to evaluate the Epsilon formation (8531 to
8556 feet). A flow rate of 8.00 MMcf/D was measured through a -inch choke at 1485 psi,
after a flow period of 289 minutes.
 A final flow test was performed, indicating a rate of 7.685 MMcf/D, through a -inch choke,
at 1418 psi, after a flow period of 196 minutes.
 The drillpipe was pulled using the primary jet and the rotating head rubber. While pulling the
string, it became stuck at 5600 feet. The string was jarred out to 5038 feet, where it pulled
free. This was caused by casing collapse from 5027 to 5050 feet.
 The drillstring was run in the hole and the well was dynamically killed with 9.3 ppg mud.
 The drilling parameters for this sidetrack are shown in Figure 7-9.
Highlights of the Air Drilling Program
 Air drilling was initiated through the target interval. Stress related sloughing of the
carbonaceous shales and coals caused some hole cleaning problems while drilling.
 The major difficulty encountered while drilling was the 7-inch casing collapse. Drilling was
not terminated because of these problems. Drilling was stopped because it was not
advisable to continue with the high gas flow rate.
 The DSTs performed in the original straight hole had indicated a combined rate of 0.83
MMcf/D. The final flow test in the sidetracked, air drilled section indicated rates almost ten
times higher than this. It would seem that the air drilling program dramatically reduced
formation damage.
 During drilling, a leak had developed in the rotating head. It was fixed with no difficulties.
The entire air discharge system would be inspected before future use, to ensure that the
components had not experienced excessive wear.

7-18

Parameter
0

10

20

30

40

50

60

70

80

90

100

8150
8200
8250
8300

Depth (feet)

8350
8400
8450
Weight on Bit (1000 lbf)
ROP (ft/hr)
RPM
Rotating Time (hours)

8500
8550
8600
8650
8700

Figure 7-9.

Drilling parameters, during sidetracking of Swan Lake-1 ST, with air mist.

 To avoid future casing collapse problems, surface pressure could be monitored in the
production-intermediate casing annulus and pressure buildup prevented. On re-entries of
existing wells, it was recommended that a casing inspection log be run to determine if
corrosion has reduced the wall thickness and weakened the pipe.

7.5

Case Study 4
Karwin-1 ST

Reference
This unpublished case study was provided courtesy of Santos Ltd.
Objectives
The Karwin-1 ST well was sidetracked below 7753 feet in order to test the Patchawarra
Formation. It had been tested in the original well, resulting in gas flows that were too small to
measure. The sidetrack hole was to be drilled with air and mist, in order to minimize formation
damage and increase productivity.

7-19

Chapter 7

Case Studies

Operational Summary
 Cement was drilled out to the sidetrack point and a bond log was run. An anchorstock
whipstock assembly was run in the hole and the well was sidetracked at 7753 feet. The
casing seat was tested to an EMW of 13.6 ppg.
 The mud was unloaded from the hole and the Patchawarra Formation was drilled with
air/mist. The hole was drilled from 7767 to 7878 feet without difficulty. An openhole flow
test was conducted (79 Mcf/D).
 A second flow test was conducted at 7938 feet (177 Mcf/D).
 Hole cleaning problems became evident at 7985 feet. The formation was sloughing and the
hole was becoming more difficult to clean. A third flow test was conducted (311 Mcf/D).
 Drilling continued and hole cleaning problems increased. The pipe became stuck while
drilling at 8182 feet. It was jarred free. The remainder of the kelly was drilled down to 8205
feet. While circulating to clean the hole, the pipe became stuck again, due to sloughing.
Sloughing problems continued. Since most of the drilling targets had been reached, drilling
was terminated.
 A four-arm caliper run could not pass 8110 feet. An elliptical hole was demonstrated. There
was a large wash-out in a carbonaceous shale from 7956 to 7962 feet, and there were massive
washouts below 8000 feet.
 A final flow test indicated a flow rate of 182 Mcf/D.
 A completion packer was run into the hole and set at 7640 feet. The annulus was filled with
2% KCl brine and the tree was installed.
Air Drilling Highlights
Problems were encountered because the formations were sloughing. The sloughing problems
were responsible for considerable time being spent to circulate and clean the hole, stuck pipe and
fishing. Of all the options available, dusting has been the most successful in reducing sloughing
in this well. It has been recommended that future wells be dusted until there is an indication that
the wellbore needs to be misted.
In future operations, if water is encountered and the hole has to be misted, the hole would be
drilled with as little water as possible and at lower annular velocities. If the well sloughs and
drilling can continue, the viscosity of the mist should be increased by adding more surfactant. If
the well still does not clean up, the air volume should be increased up to the maximum rate
available. If the hole cannot be cleaned by running all of the compressors, air drilling will need to
be terminated. Otherwise, the BHA will eventually become stuck. Figures 7-10 and 7-11
indicate the drilling history of this well.

7-20

Parameter
0

10

15

20

25

30

7600

7700

Weight on Bit (1000 lbf)


Cumulative Time (hours)

Depth (feet)

7800

7900

8000

8100

8200

8300

Figure 7-10.

Weight on bit and rotating time while air mist drilling the sidetrack in Karwin-1 ST.
Parameter
0

20

40

60

80

100

120

140

7600

7700

Depth (feet)

7800

7900

8000

8100

8200

ROP (ft/hr)
RPM

8300

Figure 7-11.

ROP and RPM while air mist drilling the sidetrack in Karwin-1 ST.

7-21

Chapter 7

7.6

Case Studies

Case Study 5
Unloading the Hole From the Bottom

Reference
This hypothetical situation was provided courtesy of R.L. Graham, Rebuen L. Graham, Inc.
Introduction
Unloading a hole from the bottom is a technique that can be used to remove water or mud from
the hole after setting and cementing casing, so that air drilling can be resumed. This operation is
a simple process that is not widely used or understood, but can be implemented at any depth (it
has been used at depths greater than 17,000 feet) with a relatively modest booster capacity (1000
psig). This is accomplished by reducing the equivalent mud density of the liquid by adding air.
The maximum available booster operating pressure is used to counteract the hydrostatic head and
friction. This relationship determines the amount of air that can be injected into the circulating
system. Typically, liquid circulation is established at 2 to 3 BPM and air is then injected at an
appropriate rate. To determine the proper air injection rate, check the circulating pressure with
just liquid (as above) and subtract this pressure from the maximum booster operating pressure.
The result is the maximum difference in hydrostatic head which can be utilized for unloading the
liquid from the hole. A calculation must be made to determine the proper volume of air to be
injected. One simple method is to determine the average pressure in the well by adding the
surface circulating pressure (mud gauge) and the liquid hydrostatic pressure at the bottom of the
hole [0.052 (psi/ft/ppg) depth (feet) mud weight (ppg)] and divide by two. This gives an
average pressure which can be used to estimate the volume of air in the drillstring for a first
unloading cycle. Consider the following hypothetical case.
Suppose that 8 5/8-inch, 36 lb/ft casing has been set and cemented at 7000 feet. The cement
plug, cement and shoe have been drilled out using water and it is desired to go back to air
drilling. The compressor system consists of two 1050 scfm primaries and a two-stage booster
which can handle about 1400 scfm at 1000 psig. The drillpipe is 4-inch, 16.6 lb/ft and there are
fifteen 6-inch x 2-inch collars. The temperature gradient is 1F/100 feet and the surface
temperature is 60oF (520oR).
First Injection Cycle
The first step is to calculate the volume in the drillstring and in the annulus.
 The drillstring capacity is the drillpipe volume plus the drill collar volume.

Drillpipe Internal Diameter......3.826 inches


Drill Collar Internal Diameter .....2.5 inches
Drillpipe Length............................. 6550 feet
Drill Collar Length .......................... 450 feet
7-22

 The capacity of the drillpipe is:


2

.
3826


12
= 523 ft 3
6550
4

(7.1)

 The capacity of the drill collars is:


2

2.5

2
= 15.3 ft 3
30 15
4

(7.2)

 The total drillstring capacity is 538 ft3.


 For the specified casing, the internal diameter is 7.825 inches. The annular volume is:
( 4.5 / 12) 2
( 7.825 / 12) 2
7000

6550
4
4

( 6 / 12) 2
3
450
= 1525.94 ft
4

(7.3)
Calculate the average volume of air.
 For the equipment on location, it is known that the circulating pressure at 3 BPM (126 gpm)
is 250 psig as fluid is pumped down the drillpipe. This pressure can be up to 900 psi (this
provides a cushion of 100 psi below the booster capacity).
 The hydrostatic pressure with water in the hole is:
7000 feet 0.052 psi / ft / ppg
8.33 ppg = 3021.2 psig

(7.4)

 The average downhole pressure is the arithmetic average of the surface pressure plus the
bottomhole hydrostatic pressure (ignoring friction at this low rate):

( 30212
. + 900) / 2 = 1960.6 psig

(7.5)

7-23

Chapter 7

Case Studies

 The average temperature is determined by dividing the total depth by two, multiplying this by
the temperature gradient and adding the surface pressure.

( 7000 / 2) 0.01 + 60 = 95 F = 555 R

(7.6)

 The average volume occupied by 1 scf of air is:


T P
555 14.6

V2 = V1 2 1 = 1
=
520 (14.6 + 1960.6)
T1P2
= 7.89 10 3 ft 3

(7.7)
where:
T1 ....... temperature at standard conditions
(oR),
T2 ....... average downhole temperature (oR),
P1 ....... pressure at standard conditions
(psia), and,
P2 ....... average downhole (psia).
The density of this unit surface volume of air at average downhole conditions is:
T P
520 (14.6 + 1960.6)
2 = 1 1 2 = 0.0764
=
555 14.6

T2 P1

(7.8)

= 9.68 lbm / ft 3 = 129


. ppg
where:

1 ....... density of air at standard temperature and pressure (0.0764 lbm/ft3),


and,
2 ....... density of air at the average downhole temperature and pressure
(lbm/ft3).
 Since the density of the air is less than the density of water, the hydrostatic pressure in the
column can be reduced by aeration. Recall that the available booster pressure is 1000 psi
minus a 100 psi safety factor minus 250 psi which is expended in circulating water at 3 BPM.
This gives an additional 650 psi to work with.

7-24

First Cycle
 Knowing that the excess booster capacity is approximately 650 psi, calculate the volume of
air that will reduce the hydrostatic pressure by 650 psi when the lightened fluid reaches the
bit:
P = 7000 (feet) .052 (psi / ft / ppg)
8.33 (ppg) water 7000 (feet) .052 (psi / ft / ppg)
AF (ppg)

(7.9)

aerated fluid

650 = 3032 3.64AF AF = 6.54 ppg

 The required density of the aerated fluid is 6.54 ppg. With water (8.33 ppg) circulated at 3
BPM (126 gpm) and air, at the average downhole pressure, (1.29 ppg), the required air rate is:
AF ( ppg) Q AF (gpm) =
= w ( ppg) Q w (gpm) + A ( ppg) Q A (gpm)
6.54{126 + Q A } = 8.33 126 + 129
. QA

(7.10)

Q A = 43.2 gpm = 5.77 cfm


520 (14.6 +1960.6)
Q A = 5.77
= 713 scfm
555 14.6

 Since this is a gross approximation (frictional changes have not been incorporated), a smaller
air rate, for example, 500 scfm, would be chosen and pumping started. To begin, one compressor unit is brought on line, making every effort to control the air rate (with the bleed
choke) to 500 scfm (4.05 cfm). Since the total drillstring capacity is 538 ft 3 and the
approximate downhole rate is 20.89 cfm (126 gpm = 126/7.48 = 16.84 cfm plus 4.05 cfm), it
should take approximately 538/(20.89) 26 minutes of pumping for the light fluid to reach
the bit. The circulating pressure should rise from 250 psi to about 900 psi.
 This is an approximate method that can be used on location for estimating the initial required
air volumes.
Second Cycle
 The next step is to more accurately determine the time and circulating pressure when the light
fluid (aerated) turns the bottom of the hole. Simple approximations, shown above, indicate
that this will occur after ~26 minutes. The circulating pressure could change this time.
Observe the circulating pressure and determine the largest circulating pressure and when this
peak occurs. Using these observed pressures and time data, recalculate the average pressure
7-25

Chapter 7

Case Studies

and the average volume occupied by one sft3 of air. As the circulating pressure drops, more
air can be added. It is important not to increase the air rate too aggressively. Therefore, the
next step is to evaluate an interim scenario after pumping for about ten minutes (this is an
arbitrary time) longer than required to get the aerated fluid to the bottom.
Evaluate The Pressures and Volumes After 36 Minutes
 Because of mass balance considerations, pumping for ten minutes longer than needed to get
aerated fluid to bottom leads to a volume of aerated fluid in the annulus that is approximately
equal to the sum of the volume of water pumped in the ten minutes plus the downhole
volume of air injected in the same ten minutes.
 The volume of water moving into the annulus in the ten minutes is:
Vw =

126 gpm 10 minutes


= 168.5 ft 3
3
7.48 gallons / ft
(7.11)

 The volume of air moving into the annulus in the ten minutes is:
VA = 500 scfm 10 minutes
555 14.6

= 39.44 ft
(
)

+
.
.
520
1960
6
14
6

(7.12)
 The total volume moving into the annulus in the ten minutes is:
VT = 168.5 + 39.44 = 207.95 ft 3

(7.13)

 The annular capacity around the drill collars (ft3/ft) and (ft3) is:
2
2
7.825 6
= 0.1376 ft 3 / ft

4 12 12

450 feet 0.1376 ft 3 / ft = 61.93 ft 3


(7.14)

7-26

 Since 207.95 ft3 has moved into the annulus, 207.95 61.93 = 146.02 ft3 is behind the
drillpipe. The annular capacity around the drillpipe (ft3/ft) and (ft3) is:
2
2
7.825 4.5
= 0.2235 ft 3 / ft

4 12 12

6550 feet 0.2235 ft 3 / ft = 1464.02 ft 3

(7.15)
 The height of aerated liquid in the annulus between the drillpipe and the borehole wall is
146.02 ft3/0.2235 ft3/ft = 653.3 feet. The total height of aerated fluid in the annulus is 450 +
653.3 feet = 1103.3 feet.
 The hydrostatic head at this point is (7000-1103.3) feet x 0.052 psi/ft/ppg x 8.33 ppg = 2554
psig. Using the pressures of 2554 psig and 3021.2 psig (refer to Equation (7.4)), to calculate
an average pressure for the aerated fluid in the annulus, results in an average pressure of 2788
psig. Using this average pressure and an average temperature of 584oR (temperature at the
midpoint of 6446 feet) the corrected air volume in the annulus is:
Vcorrected = 500 scfm 10 minutes
584 14.6

= 29.25 ft
520 ( 2788 +14.6)
(7.16)
 The total volume moving into the annulus in the ten minutes is then:
VT = 168.5 + 29.25 = 197.75 ft 3

(7.17)

 The volume behind the drillstring is 197.75 61.93 = 135.82 ft3. This corresponds to 135.82
ft3/0.2235 ft3/ft = 607.7 feet. The total height of aerated fluid in the annulus is 450 + 607.7 =
1057.7 feet.

7-27

Chapter 7

Case Studies

 Calculating the bottomhole pressure:


PBH = ( 7000 1057.7) 0.052 8.33 +
+ 1057.7

0.052

7.48 ( 29.25 + 168.5)

520 ( 2788 + 14.6)


{29.25 .0764
+
584 14.6

+ 168.5 7.48 8.33} =


= 2573 + 404.3 = 2978 psi

(7.18)
 This indicates that the bottomhole pressure has decreased from 3021 to 2978 psi. Very little
pressure drop has occurred. This pressure drop may not be apparent on a typical mud gauge.
Continue the pumping program in the same manner until a pressure decrease is apparent.
Perform additional interim pressure calculations, for example, after another ten minutes (i.e.,
46 minutes after the start of pumping).
Evaluate The Pressure and Volumes After 46 Minutes
As a starting point, take a first estimate of the average downhole pressure in the aerated column
on the backside as the pressure at the top of the column in the annulus after 36 minutes of
pumping. This would be for a depth of water of 7000 1057.7 = 5942.3 feet. The
corresponding hydrostatic pressure would be:
PBH = 5942.3 0.052 8.33 = 2574 psi

(7.19)
 Approximate the temperature as:
T = 5942.3 0.01 + 520 = 579  R

(7.20)

 The volume of water moving into the annulus in this additional ten minutes is:
VW =

126 gpm 10 minutes


= 168.5 ft 3
3
7.48 gallons / ft

(7.21)

7-28

 The volume of air moving into the annulus in this additional ten minutes is:
VA = 500 scfm 10 minutes
579 14.6

= 31.40 ft
(
)
520 2574 +14.6

(7.22)

 The total volume moving into the annulus in this additional ten minute period is:
. = 199.9 ft 3
VT = 168.5 + 3140

(7.23)

 The additional footage of annulus filled is behind drillpipe only. This is 199.9 ft3/0.2235
ft3/ft = 894.4 feet. The total height of aerated fluid in the annulus is approximately 1057.2 +
894.4 1952 feet.
 The bottomhole pressure is now approximately:
PBH = ( 7000 1952) .052 8.33 +
+ 1952

0.052

7.48 ( 31.40 + 168.5)

520 ( 2574 + 14.6)


{3140
. .0764
+
579 14.6

+ 168.5 7.48 8.33}


= 2925 psi

(7.24)
 The measured pressure drop (from initial conditions) is now 3021 - 2925 = 96 psi. This is
probably detectable on the mud gauge. This suggests that there is now a possibility to
increase the air rate and decrease the water rate to further lighten the fluid.
 Since it is not known exactly how much to increase the air, increase the air rate nominally
and cut the water rate to offset the increase in the volume of air. The rate changes can be
estimated. Suppose that the air injection rate is increased from 500 to 800 scfm. The
previous calculations have indicated that there is approximately 96 psi in excess booster
capacity because of the reduced head in the annulus. The average downhole pressure is
determined from the circul-ating pressure of 900 psig and the previously estimated
bottomhole pressure of 2925 psig [(900+2925)/2 = 1913 psig. The temperature can be
averaged over the full well depth (555R).

7-29

Chapter 7

Case Studies

 Increasing the surface air rate from 500 to 800 scfm (by 300 scfm) increases the average
downhole air rate by:
555 14.6

Q A = 300 scfm
=
(
)
520 1913 + 14.6
= 2.425 cfm = 18.1 gpm

(7.25)
 Therefore, if the air rate is increased by 300 scfm, reduce the water rate by 18 gpm, from 126
to 108 gpm. Presuming that the surface circulating pressure is 900 psig, the bottomhole
conditions, after an additional 15 minutes of pumping (61 minutes total air injection time)
can be approximated.
Evaluate The Pressure and Volumes After 61 Minutes
As a starting point, take a first estimate of the average downhole pressure in the aerated column
on the backside, as the pressure at the top of the column in the annulus after 46 minutes of
pumping. This would be for a depth of water of 7000 - 1952 = 5048 feet. The corresponding
hydrostatic pressure would be:
PBH = 5048 0.052 8.33 = 2187 psi

(7.26)

 Approximate the temperature as:


T = 5048 0.01 + 520 = 570 R

(7.27)

 The volume of water moving into the annulus in this additional fifteen minutes is:
VW =

108 gpm 15 minutes


= 216.6 ft 3
3
7.48 gallons / ft

(7.28)
 The volume of air moving into the annu-lus in this additional fifteen minutes is:
VA = 800 scfm 15 minutes
570 14.6

= 87.23 ft
(
)
520 2187 +14.6

(7.29)

 The total volume moving into the annulus in this additional fifteen minute period is:
VT = 216.6 + 87.23 = 297.83 ft 3

(7.30)

 The additional footage of annulus filled is behind drillpipe only. This is 297.83 ft3/0.2235
ft3/ft = 1332.6 feet. The total height of aerated fluid in the annulus is approximately 1952 +
1332.6 3285 feet.
7-30

 The bottomhole pressure is now approximately:


PBH = ( 7000 3285) 0.052 8.33 +
+ 3285

0.052

7.48 ( 216.6 + 87.23)

520 ( 2187 + 14.6)


{87.23.0764
+
570 14.6

+ 216.6 7.48 8.33} = 2772.4 psi

(7.31)
 It is apparent from this calculation that the adjustments are very reasonable and it is justified
to proceed with pumping the new volumes.
Additional Cycles
At this point, nearly half of the non-aerated water is out of the hole and the emphasis should be
on cutting back on the water rate.
Use the pressure gauge to govern additional pumping. Cut the water volume by 0.5 bpm and
pump for ten minutes, continuously monitoring the pressure. If there is no change (increase or
further decrease) in pres-sure, reduce the water rate by another 0.5 bpm and continue until all
water injection has been stopped. If the pressure increases at any time, try to continue
pumping at the same rate, unless the pressure exceeds the pre-established pressure limits, in
which case, increase the water rate by the last 0.5 bpm decrement until pressure decreases.
The final air volume is sufficient to unload the hole, but it can be increased after the non-aerated
water has been displaced.
Rigup for this operation (presuming the depth shown and water initially in the hole) should
consider that the water injection will last for about two hours (nearly 360 bbls of total water). If
possible, rig up so that the displaced water can be reused. Also, the rigup should take into
account that there will be a substantial kick when the aerated water nears the surface. Be sure
that the flowline is staked down and be prepared for the kick.
These calculations are for demonstration only. More sophisticated calculations can be used and
easily programmed into a spreadsheet. The major point of this hypothetical example is to
emphasize the time savings that are possible using this method, as compared to staging into the
hole.

7-31

Chapter 7

7.7

Case Studies

Case Study 6
Gasified Liquid (Con-centric String Injection)

Reference
Teichrob, R.R.: Low-Pressure Reservoir Drilled With Air/N2 In A Closed System, Oil & Gas
J. (March 21, 1994) 80-90.
Background
A well, with an 1800 ft lateral section, was proposed in the Camrose reservoir, in Alberta (Husky
Wainwright 15B-31-44-4W4M). This is a low-pressure reservoir with a history of high fluid
losses during conventional drilling. The reservoir proper-ties were:
Current reservoir pressure............... 210 psi,
Original reservoir pressure.............. 580 psi,
GOR .......................................... 12.57 scf/ft3,
Reservoir top..................................2264 feet,
Solution gas ................................ 88-90% C1,

5% N2, 5-7% C2, C3 and C4+, and,

Reservoir temperature ....................... 70F.

Preparation

Since oil production in a closed system was expected, the potential for a flammable mixture was
assessed using lease crude, solution gas, mud and air. Ignition due to compression was not
considered a problem (because of the low reservoir pressure). Test results showed that a safe
mixture was (by volume) 85% air, with 9% drilling fluid and 6% hydrocarbons. 70% air was
ultimately decided as a safe limit. However, to meet underbalanced conditions, 95% air would
have been required. To meet this, further testing showed that a 60:40 air/nitrogen gas mixture
would be acceptable.
Surface Facilities
The BOP stack, including an RBOP, allowed for commingled gas injection on the casing side
(9 5/8-inch x 7-inch) and recovery of the 7-inch concentric string after the well had reached
TD.
Flow line diameter from the RBOP flange through the separation system was sized at 4inches because of the low reservoir pressure and the relatively high injection rates.
A three-phase, horizontal, skid-mounted separator (rated at 50 psi) was sized to handle return
fluids.
Produced/drilling water was recycled. Produced oil was transferred to a 400 bbl tank and then
trucked off location.
Produced gas (air, nitrogen and hydrocarbon) was run through a 4-inch line to a 39-foot flare
stack.
Drilled solids were contained in the separator during the horizontal drilling portion.
7-32

 Because of combustibility concerns, a 1-inch N2 blowdown line was run to an inlet upstream
of the separator. A methane chromatograph and a portable O2 and lower explosive limit
meter were incorporated in the surface system.
Operations
A parasite string was an undesirable option in this horizontal well. Also, a water hydrostatic
from the kickoff point to the TVD would kill the well. A system had to be devised to get
injection gas to TVD that could eventually be removed and reused. A concentric string
application was selected (this could be surveyed with conventional MWD and unloaded from the
casing side simultaneously).
Operations went as follows (refer to Figure 7-12):
 17.5-inch surface hole was drilled to approximately 490 feet TVD (mud and native clays
were used for the mud system).
 13 3/8-inch casing was cemented to surface.
 12-inch hole was drilled to the top of the Nisku formation (planned build angle of 810/100 feet). The upper Colony gas would be squeezed.
 There was severe lost circulation on drilling into the Nisku. Following a squeeze, drilling
continued and 9 5/8-inch casing was set at 2375 feet MD, at a 64 inclination.
 Drilling continued with an 8 3/4-inch bit to intermediate TD at 2927 feet MD and an
inclination of 90. A 7-inch drilling liner was run to intermediate TD, set and foam
cemented.
 A 4-inch motor and MWD system were used for drilling. Flow rates were 1.25 BPM of
liquid with 805 scfm (initially) air/nitrogen down the drillpipe and 505 scfm air/nitrogen
down the casing string. With the kelly down, liquid was injected at 3.1 to 3.8 BPM and
casing injection was increased to up to 706 scfm. Nitrogen and air services were on line once
the 6-inch bit was in the 7-inch shoe.
 Water and nitrogen were circulated to the surface.
 Air was then injected and commingled (60/40 air/nitrogen). This supply was then split and
fed to the casing and the drillpipe.
 When the well was blown down and circulating rates had stabilized (and water added to the
drillstring stream), the shoe was drilled out at stabilized annular pressures of 21.8 to 29 psi
(standpipe pressures of 725 to 1015 psi).
 With the kelly down, prior to the next connection, all gas was directed to the casing and the
water rate was increased, as previously indicated. The drillpipe was circulated with water
and the connection was made.
 After a survey, initial rate conditions were re-established. Injection pressure typically
increased (annular pressure of 189 psi) until the fluid column was displaced from the well.
7-33

Chapter 7

Case Studies

With the resulting decrease in annular pressure, injection pressure also declined.
 After drilling approximately 656 feet, it was determined that if the survey frequency was
increased, injection and annular pressure did not have a chance to return to minimum levels.
Eventually overbalanced conditions resulted, the well was killed and the pipe was stuck.
 To avoid this, injection rates on the backside were increased. Oil returns possibly suggested
the success of this procedure. Eventually, backside rates were increased to 1515 scfm.
 With increasing oil production, nitrogen purging through the separator blowdown line had to
be initiated (and maintained) at 230 scfm.
 At TD, air and water pumping were stopped and the well was displaced with nitrogen (to
purge oxygen).
 The well was displaced with native crude.
 A bridge plug was run and set in the 7-inch liner, the casing above the liner top was displaced
with water and the bridge plug was pressure tested.
 The BOP stack was laid down and the 7-inch tie-back concentric string was recovered.
 A 9 5/8-inch packer was set at the surface and the rig was released.
Results
 Initial production rates from this well were between about 2.5 and 6 times higher than typical
production rates from vertical wells in the area.
 Pre-drilling expenditures, to evaluate required nitrogen content, resulted in substantial
savings, since decreased volumes of nitrogen were used.
 Underbalance would not have been possible without the concentric string (problems with
bending of a parasite string, holiday in the intermediate casing, economics of recycling,
required flow rates ).

7-34

17.5-inch hole diameter


13 7/8-inch surface casing
492 feet TVD
3 1/2-inch drillpipe
9 5/8-inch intermediate casing
12 1/4-inch hole diameter
7-inch tie back liner
39 foot slotted joint
Foamed cement
6 1/4-inch diameter hole
N2/air/water
N2/air
N2/air/water/oil

Figure 7-12.

7.8

Schematic configuration of Wainwright 15B-31-44-4W4M (after Teichrob, 1994 2).

Case Study 7(Underbalanced Re-Entry)

Reference
Stone, C.R. and Cress, L.A.: New Applications for Underbalanced Drilling Equipment, paper
SPE 37679, manuscript under review, 1997.
Background
This was a re-entry candidate. 5-inch, 26.80 lb/ft casing had been set at 16,500 feet, but had
collapsed at 15,450 feet. Originally, attempts were made to whipstock at 15,378 feet in order to
go around the obstruction. Three attempts failed. Flow-drilling was then attempted and was
successful.
The First Re-entry Attempts
 A whipstock was set and a sidetrack was initiated in the 5-inch casing. The BHA
incorporated 3 1/8-inch drill col-lars, without stabilization. 4 3/8-inch hole was drilled to
16,100 feet. Shale and gas problems at this depth led to increasing the mud weight to 17.5
ppg. There was increasing evidence of torque and drag. Swab and surge pressures led to
more gas problems during tripping. The string became stuck while in several partially
depleted sands. The pipe was freed with oil spotting agents. Eventually, a fish was left in the
hole at 16,207 feet.

7-35

Chapter 7

Case Studies

 A cement plug was set above the fish and dressed off. Drilling proceeded until the pipe again
became stuck at 16,100 feet, due to sloughing shale. A second fishing job resulted in a
second sidetrack around a fish left in the hole.
 The drilling fluid was changed to an oil-based mud. A cement plug was set above the fish
and dressed off. A mud motor was used to sidetrack off the plug. The water-based mud was
displaced by oil-based mud at the window in the casing. A kick resulted because of problems
in the displacement. The kick was circulated but the drillstring failed and another fishing job
was required. This was also unsuccessful.
The Successful Re-entry
After the previous failures, certain modifi-cations were made.
 Another whipstock was set at 14,500 feet, to eliminate problems with the original window.
A schematic of the well is shown in Figure 7-13.
 Another problem had been swabbing gas during trips It was determined that underbalanced
conditions would be maintained if the mud weight did not exceed approximately 16.0 ppg.
 Differential sticking and hole stability had also been identified as problems, the stability
being related to the water-based mud system.
 The decision was made to change to an oil-based mud, to decrease the mud weight to 16.5
ppg and to modify the bottomhole assembly. The BHA was changed to consist of a positive
displacement motor, one 3 1/8-inch monel drill collar and 2 7/8-inch PH-6 tubing as a
drillstring. This assembly was designed to minimize surging and swabbing problems, by
reducing the length of the BHA. Not using stabilizers helped avoid sticking problems across
the depleted sands.
 An RBOP was used. The swab pres-sures were high due to the tight clearances between the
hole size and the bottomhole assembly size. Calculations showed a 1.2 ppg hydrostatic
pressure loss due to these tight clearances. Regardless of how slowly the drillstring was
pulled out of the hole, the hole would not fill up with the correct amount of fluid. To
overcome this problem the RBOP was used as a stripping device. The annulus was pressured
up to 100 psi. The drillpipe was stripped through the RBOPs kelly packer until the pressure
was reduced to 0 psi. In effect, the fluids were being forced past the bottomhole assemblies
by the additional pressure that was applied at the surface.
 Many wells in this area experience pressured shale or tight gas sand stringers. On penetration
of these zones, gas cut mud has indicated that mud weight might not be adequate. In this
area, increasing mud weight to overcome this had led to a chain reaction of required
additional weighting. Before raising the mud weight, bottoms up were circulated several
times. The pressure was allowed to decrease to as little as 16.2 ppg. Sticking was not
encountered and the well was successfully drilled.

7-36

Justification
 The savings of weighting up from 16.2 to 17.5 ppg were calculated to be $42,000.
 Maintenance of the additional mud weight would have cost $1800/day.

14500
feet

4 3/8-inch Hole
Oil Mud at 16.2 ppg
RBOP
15378
feet

Water-Based Mud
17.5 ppg
Conventional Drilling
Unstable
Hole
Fish
Fish
Shale

Gas Target Sand

5-inch, 26.80 lbm/ft at 16500 feet

Figure 7-13.

Schematic of the re-entry drilling history described in Case Study 7 (mod-ified from
3
Stone and Cress, 1997 ).

 Trip time was dramatically reduced.


Other References
The reader is also referred to two papers by Joseph, 1995a,b,4,5 which provide compre-hensive
descriptions of underbalanced drilling operations in deep, high temp-erature, abnormally
pressured Austin chalk, in Louisiana.

7-37

Chapter 7

7.9

Case Studies

Case Study 8 Controlled Tripping

Reference
Stone, C.R. and Cress, L.A.: New Applications for Underbalanced Drilling Equipment, paper
SPE 37679, manuscript under review, 1997.
Introduction
 During vertical drilling of the Pinnacle Reef structures of the Cotton Valley Lime, at depths
of up to 16,000 feet, operators would typically drill and set 9 5/8-inch casing at 9500 feet and
drill ahead with 8-inch hole through the Cotton Valley Sand into the top of the Cotton
Valley Lime for a 7-inch drilling liner seat. The target is then drilled with a 6-inch bit,
using a tapered drillstring (4 x 3-inch). In offset wells, the mud weight was typically 15.0
ppg, increasing to 17.5 ppg at the top of the Cotton Valley reef. Mud weights of up to 18.5
ppg had often been used through the reef. A typical section is shown in Figure 7-14.
 It was determined that the maximum pore pressure in the reef was typically 13 ppg and many
previous drilling operations had been severely over-balanced.
 In drilling through the Cotton Valley Sand, at approximately 12,000 feet, stuck pipe had
previously been encountered, as well as gas seepage, particularly CO2.
 Mud rheology problems had been diagnosed because of difficulty in breaking circulation and
excessive wall cake buildup across the tight sands. There were difficulties in pulling
stabilizers out of the hole across these sands as well as sticking of logging tools.
 Lost time across the sands had also been extensive because of swabbing high pressure gas.
 Tripping back into the hole, circulating bottoms up and conditioning the mud would be
common, time-delaying requirements because the holes would not accept the proper amount
of mud.
Solution
How were these problems resolved in a new well?
An RBOP was installed on the top of the BOP stack, with a pressurized flowline and an HCR
valve to prevent swabbing gas into the wellbore and to avoid kicks during tripping out of the
hole. Tripping was performed as follows:
 The RBOPs activating pressure was set to automatic mode so that its sealing rubber would
automatically exert 300 psi closing force greater than the annular pressure.
 The driller would close the flowline and the choke valve to isolate the annulus.
 As the bit was pulled off bottom, annular pressure was monitored to determine the magnitude
of the swab pressure at the BHA.
 Tripping speed was optimized so that the assembly could be pulled without exceeding the
trip margin at the bit.
7-38

 Using this information, the RBOP was opened and tripping was carried out at this particular
rate. It was determined that the pipe speed must not cause an annular pressure of more than
390 psi. In this well, the swab pressure around the BHA decreased considerably after the first
three stands had been pulled and the mud gel properties had been overcome.
This technique should be used with caution if there are weak under-pressured formations
open to the wellbore above the bottomhole assembly (BHA). All sections of the open hole
above the BHA would be subjected to the inverse of swab pressure which is surge pressure.
Surge pressure can cause well control problems by breaking down a weak formation and
possibly initiating loss of circulation.
Outcome
The well was successfully drilled with minimal lost time due to contaminated mud and swabbing
on trips.
Water Based Mud
CO 2 Contamination
pH and Viscosity Variations

Trip Speed Calculations


(Bit at 15000 feet)
(Unknown Mud Properties Downhole)
1 Trip Margin 0.5 ppg 390 psi
2 Close RBOP, Flowline and Choke
3 Pull Pipe at Given Speed
4 Monitor Annulus Pressure
5 Pipe Speed Must Not Produce Annular
Pressure > 390 psi
6 Open Annulus and RBOP
7 Trip Out at Prescribed Rate

CO 2

Tight Cotton Valley Sands

Cotton Valley Reef

Figure 7-14.

Wellbore schematic for Case Study 8 (after Stone and Cress, 1997 ).

7-39

Chapter 7

7.10

Case Studies

Case Study 9
Flowdrilling

Reference
Stone, C.R. and Cress, L.A.: New Applications for Underbalanced Drilling Equipment, paper
SPE 37679, manuscript under review, 1997.
Introduction
The primary producer in this field is a thick (600 feet), highly-fractured, carbonate, at
approximately 8000 feet TVD. Gas production contains up to 3% H2S and a small amount of
CO2. The zone is partially depleted (approximately 5 ppg EMW). There are overpressured
shales above and normally pressured, permeable water sands below. The thickness of the target
and the over- and underlying formations prevented drilling the carbonate with any other
formation exposed. In the past, significant problems had been encountered:
 Typically, 9 5/8-inch casing was set near the top of the carbonate and drilling proceeded with
an invert emulsion oil mud.
 Within 100 feet of the casing seat, massive lost circulation would occur.
 With this loss of circulation, the annular fluid column would fall below equilibrium and a
kick would lead to severe well control problems and pressures up to 2500 psi on the surface
annulus.
 At the point where drilling was discontinued because of well control, operations would be
suspended. The wells were commonly completed with only partial interval coverage or were
plugged and abandoned.
Solution
These problems were significant and were compounded by the motivation to reach a deeper
target, seismically forecasted at approximately 11,000 feet. The protocol used to successfully
drill to this new zone was as follows (refer to Figure 7-15).
 9 5/8-inch intermediate casing was set above the depleted zone.
 Drilling proceeded towards the deeper target (6 1/8-inch hole), using flow-drilling with 7.4
ppg invert emulsion oil mud.
 Surface equipment consisted of an RBOP on top of a three-ram stack and one annular
preventer with twin gas separators (capable of each processing 40 MMscf/D), vacuum
degasser tanks and chemical injection pumps.

7-40

Annular Surface Pressures Up To 2500 psi

13 3/8-inch

7.4 ppg Oil Mud (90/10)


N 2 Injection As Needed
H 2 S Scavenger In The Mud And
Injected At The Choke Manifold
9 5/8-inch
At 8202 Feet
N 2 Foamed Cement

Highly Fractured Carbonate


Depleted Gas Zone (5 ppg EMW)
Severe Loss Of Circulation
3% H 2 S + CO 2

7-inch Liner At 8858 Feet

Figure 7-15.

Wellbore schematic for Case Study 9 (after Stone and Cress, 1997 ).

 Twin vacuum degassers removed entrained gas from the drilling fluid. Produced gas was
flared through a vertical flare boom with an automatic igniter. H2S monitors were located
upstream of the gas separators and in the vacuum degasser tanks. The drill fluid was
pretreated with liquid H2S scavenger. Any sour gas detected at the choke manifold was
further treated using a liquid injection pump downstream of the choke manifold.
Drilling
 During flowdrilling of the depleted interval, the maximum surface casing pressure was 2500
psi. Minor seepage losses were encountered.
 At one point in the drilling of this section, the kelly packer needed to be changed during a
connection. The well was alive with about 1800 psi on the annulus. The annular preventer
was closed, and the volume between the annular preventer and RBOP contained trapped
pressure. Assuming H2S laden gas, we bled the pressure to the gas separators and purged the
stack section with N2 prior to opening the RBOP. The RBOP kelly packer was quickly
changed and drilling resumed. The entire section was drilled in two days.
 Historically, the operator had struggled with this section for as much as 60 days using
conventional well control procedures.

7-41

Chapter 7

Case Studies

 Full circulation of the drill fluid in the hole followed by a shutdown of the pumps to monitor
fluid column stability preceded any attempt to trip out of the hole. The well had an ability to
bubble and come alive at any moment with H2S gas at the surface with no conduit (drillpipe)
in the hole to circulate kill fluid.
Logging
With a wireline lubricator on top of the RBOP, the section was logged, without incident, with
less than 500 psi wellhead pressure.
Running The Liner
 A conditioning trip was made. There were only minor seepage losses.
 The drilling team desired to minimize well control risk when the drillpipe was out of the
hole or when running casing.
 With the kelly packer removed, the RBOP had an 11-inch open bore with the inner packer
relaxed.
 Since it was anticipated that the well would be surged and would kick while running casing,
facilities to strip the 7-inch liner into the wellbore were rigged up as a contingency.
 If the surface pressure was low (less than 2500 psi, as was expected), the casing couplings
would be stripped between the RBOP and the annular preventer.
 For surface pressure in excess of 2500 psi, the casing couplings would have been stripped
between the upper pipe rams (7-inch rams installed) and the RBOP.
 With the well static, casing was run slowly without incident and cemented with a nitrogenfoamed cement. The RBOP was used (a pipe ram and annular preventer could have been
used) because it allowed rotation. Rotation can be particularly important in a hole with
unknown rugosity or in a depleted formation. In this case, it was important to allow
reciprocation and rotation to prevent the casing from becoming stuck off bottom with the
depleted zone uncovered.

7.11

Case Study 10
Coiled Tubing Drilling

References
Wang, Z., Rommetveit, R., Vefring, E.H., Bieseman, T. and Faure, A.M.: A Dynamic
Underbalanced Drilling Simulator, presented at the 1995, 1st International Underbalanced
Drilling Conference and Exhibition, The Hague, The Netherlands, October 2-4.
Adam, J. and Berry, M.: Underbalanced Coiled Tubing Sidetrack Successful, Oil & Gas J.
(December 18, 1995), 91-98.

7-42

Background
Underbalanced, horizontal drilling was selected as a desirable methodology for developing the
Dalen field, a sour gas play in the eastern part of the Netherlands. Despite production success in
this field, previously drilled wells, including an horizontal well, had suffered severe mud losses
to the formation when natural fractures were encountered, leading to unacceptable costs for time
and materials as well as stuck pipe. The Dalen 2 well had previously been shut-in when
production declined to approximately 1.06 MMcf/D. It was desired to sidetrack the original hole
and drill horizontally in the fractured carbonate, underbalanced, for a target delivery of 4.6
MMcf/D. The procedures included:
 Abandon the lower section of the original hole (remove the old production tubing, plug back
the perforated interval, cut and retrieve the existing 7-inch production casing and replace it
with a new tapered string (7 5/8-inch x 7-inch) from surface to 3294 feet MD to
accommodate the five-inch outside diameter SSSV (sub-surface safety valve), control lines
and the side pocket mandrel.
 Sidetrack conventionally to the anhydrite at the top of the reservoir, from below the 7-inch
casing shoe. During actual operations, about 66 feet of the reservoir were penetrated. Mud
losses (fracture communication) required plugging back and dressing off the cement to the
anhydrite.
 Run and cement a five-inch liner. The shoe was approximately 10 feet above the reservoir.
The hole angle at the shoe was approximately 60.
 The five-inch liner shoe track was drilled out prior to running and stabbing the completion
string into the top of the five-inch liner tie-back polished bore receptacle.
 The Christmas tree was installed.
 Drill 3-inch hole into the reservoir through the completion, underbalanced, with coiled
tubing. This is discussed below.
Drilling Fluid
The underbalanced drilling fluid was form-ulated according to:
 The expected bottomhole reservoir pressure was 4860 psi at 9482 feet (9.86 ppg EMW or
0.51 psi/ft).
 8.83 ppg NaCl brine was selected to provide the underbalance and to provide a margin of
tolerance if backpressure was applied.
 Annular nitrogen injection was also possible at 3232 ft MD.
Surface Facilities
Flaring produced gas had to be minimized. To deal with produced gas, returned drilling fluids
and solids, a standard well test equipment configuration was modified for underbalanced
operations:

7-43

Chapter 7

Case Studies

 Gases and liquids would be separated with a conventional three-phase separ-ator.


 Gas would be directed to existing gas-handling facilities or flared if necessary.
 Condensate and drilling fluid would be directed to their respective surge tanks.
 Condensate would be trucked off location.
 Drilling fluids would be reconditioned.
 Since H2S could occur and was soluble in the drilling fluid, a closed system was needed.
 Drilled solids were removed by a set of sand filters and catchers upstream of the choke
manifold and separator.
Testing
Before drilling out the cement, circulation computer models were calibrated with a nitrogen lift
test. A new two-inch diameter coiled tubing string was successfully tested for fatigue life.
Operations and Deployment
 Rigging up for CT drilling was carried out, along with training of site personnel.
 The CT BHA was 75.5 feet long. A sub-surface safety valve was used to deploy this
assembly. It was installed in the completion string at 350 feet MD and isolated the reservoir
pressures from the surface a downhole lubricator system. This allowed running the BHA
into and out of the live, but isolated well.
 The first assembly consisted of two pieces, in addition to a 3-inch barracuda mill for
drilling out cement below the liner and drilling to the reservoir top. Pressure testing failed
and the BHA was reconfigured. A short Monel drill collar was run directly above the
orienting tool, allowing the three-inch inverted rams to be closed above the orienting tool.
 The top of the cement was located at 9770 feet. Drilling was slow. Circulating pressures were
high. Addition of friction reducer decreased the circulating pressure by 30 percent. Drilling
rate was very variable. The target formation was encountered at 9846 feet. The nitrogen gas
lift test was conducted at this time.
 Drilling was discontinued at 9849 feet and the assembly was pulled. The mill was 50% worn
and O-ringed.
 22 bbl of 15X HCl were spotted with NaCl brine. This was an attempt to communicate with
the previously encountered (and cemented) natural fracture. No significant gas was encountered.
 At 9872 feet, the electrical connection to the BHA was lost. It was rebuilt after retrieval.
Difficulties with isolating the orienting tool when the SSSV was operated were encountered
and overcome.

7-44

 A reverse circulation junk basket was run to recover metal debris. It was determined that the
SSSV had been dislodged. Components were retrieved and replaced.
 On a subsequent run, the BHA was lost and fished.
 Drilling progressed at approximately 6.5 ft/hr to 9902 feet where the tool became stuck while
pulling back for a survey. Eventually the entire string was displaced with N2. It was still not
possible to work free and there was no influx of gas even when the wellhead pressure was
bled to zero.
 Eventually, the assembly was retrieved and the hole was plugged back above the fish (the top
of the fish was approx-imately 36 feet below the liner shoe).
 There was speculation that debris had contributed to the sticking. The sidetrack proceeded
with a higher drilling fluid gradient. The cement was tagged at 9721 feet and drilling
proceeded to 9829 feet where the motor stalled.
 When the next assembly was run in, the well started to unload brine and produce gas at 0.53
to 1.06 MMcf/D at an FTP of 638 psi, at a depth of 9823 feet.
 Bottomhole pressure was maintained with the choke at approximately 4130 psi. The
inclination of the well had dropped (instead of the planned build) from 57 to 40 in 98 feet,
because of a computational error. After this, it was impossible to build angle and steer.
Because the well was producing, high differential pressures between the inside and outside of
the drilling assembly acted on the internal holding slips of the orientation tool, seriously
impacting its effectiveness. The effect of the extreme borehole geometry on the bent BHAs
could not be overcome by the orientation tool, and the hole tended to spiral uncontrollably.
 The well was drilled to 10,046 feet where a drilling break was encountered. Pressures
increased rapidly at 10,053 feet and production went to 10.6 MMcf/D. At 10,128 feet, gas
production varied from 8.8 to 21.2 MMcf/D.
 The hole angle had recovered to 70 but directional control was very difficult. No further
progress could be made and operations were terminated.
Results
Despite difficulties in steering, this well was successful. Its production (over approx-imately
three hundred feet in the pay) is second in the field, only exceeded by the previously drilled
horizontal well, which accessed approximately 1360 feet of pay. Important operational
considerations and major cost saving techniques were developed.
Related Considerations
A downhole lubrication system, such as the configuration used in Dalen 2, is an important
component in drilling operations. In most instances, it is also important (if not essential) to log
and complete under-balanced. Deployment systems have been developed which allow inserting
any length of BHA under pressure (both into and out of wells). In addition, a Down Hole Swab
Valve (DHSV) has been developed, permitting insertion of long BHAs into live wells (Sharman
and Pettitt, 199519). These techniques (new generation deployment systems, etc.) mean that

7-45

Chapter 7

Case Studies

lubricator height is no longer the limiting factor for the length of a BHA. Applications include:
 Long perforating guns,
 Long logging tools, and,
 Running screens or other completion components into live wells.
The normal constraint on the length of a BHA which can be run in a live well is the distance
between the top of the lubricator and the swab valve of the Christmas tree. Occasionally,
extremely long BHAs are called for, or headroom limitations severely restrict lubricator height.
In the past, techniques used to circumvent this problem have included lubricator valves (not
suitable for production wells), using the riser (in a subsea well) or well kill. Since there are
certain situations where none of these options are appropriate, alternate deploy-ment systems
have been developed.
One system uses breechlock connectors, with a hollow core. These connectors can be
manipulated by a deployment actuator while inside the riser and allow latching and/or unlatching
of segments of the BHA while under pressure. The lower section of the BHA is hung off from a
no-go shoulder inside the deployment actuator, while the upper section is pulled higher into the
lubricator. The deployment actuator manipulates the connector with dual rams in a converted
BOP body. Standard gate valves are then closed between the hung-off BHA sections and the
upper broken-out section. After breaking the lubricator, the upper BHA section is extracted. The
sequence is repeated to add or remove more BHA sections. This particular deployment system
may be used with either wireline or coiled tubing. Usually coiled tubing will be used because of
the weight of the assembly.
A Down Hole Swab Valves (DHSV) has also been developed. This surface-controlled downhole
valve creates a subsurface lubricator. This is a second downhole barrier, in addition to the tubing
retrievable subsurface safety valve. It eliminates the need for a well kill. For example:
 Presume that the well has been drilled and cased. A conventional completion is run to the
required depth. A downhole swab valve (DHSV or DSV) and control line are added to the
string approximately 100 feet above the typical tubing retrievable downhole safety valve.
 Tubing conveyed perforating guns, logging tools or other BHAs are made up. Both the DSV
and the downhole safety valve are closed.
 Pressure is bled off from the tree. CT BOP equipment and the injector head are nippled up to
the tree. Safety checks are performed.
 The well is opened and the tools are run (adding components as required).
 After the assembly is in the hole, typical coiled tubing procedures are followed. Pressure is
bled off from the DSV, allowing it to open. Pressure is then increased to open the downhole
safety valve.
 The assembly is deployed to the desired depth and the completion performed.
7-46

 The assembly is retrieved.


In addition to drilling, development of these tools emphasizes the importance of maintaining
underbalanced conditions at all times during the life of the well. Other underbalanced
completions protocols can include:
 Flowing The Well: For example, a liner can be run into the openhole while the well is
flowing. This technique has been commonly used in the Austin chalk in Southeastern Texas.
 Two-Stage Drilling: This common method entails drilling to the top of the critical horizon,
running and cementing intermediate casing, drilling out the shoe and to TD underbalanced
and running a liner across the production interval. As always, safety precautions are followed
and appropriate wellhead equipment is used.
 Inflatable Packer Isolation: This entails setting an inflatable bridge plug to serve as a
temporary barrier. After underbalanced drilling, the drilling assembly is pulled and an
inflatable bridge plug is run and set in the intermediate string at a depth great enough to allow
the liner (or TCP guns, sand screens, etc. ) to be run above it. The casing above the packer
is bled off (or pressurized and then bled off) to check integrity. The liner is run into the hole
and the bridge plug is engaged and released with an overshot. The liner and bridge plug are
moved to the desired depth and cemented in place as required. Thru-tubing inflatable bridge
plugs have also been used and retrieved after the liner is temporarily suspended in the casing.
Other References
Walker, T. and Hopmann, M.: Under-balanced Completions Improve Well Safety and
Productivity, World Oil (November 1995), 35-39.
Sharman, D.M. and Pettitt, A.J.: Deploy-ment Systems and Down Hole Swab Valves, paper
SPE 30406 presented at the 1995 Offshore Europe Conference, Aberdeen, Scotland, September
5-8.

7.12

Case Study 11
Cementing

Reference
Purvis, D.L. and Smith, D.D.: Real-Time Monitoring Provides Insight Into Flow Dynamics
During Foam Cementing, SPEDC (June 1994), 124-132.
Background
Purvis and Smith, 1994,8 reported several examples of cementing across the Dakota formation, a
corrosive saltwater zone. Two to three salt zones and three to four potential lost-circulation
zones can be encountered, depending on the area, during drilling and completion. The number
of stages and casing collapse problems have been reduced by using foamed systems.

7-47

Chapter 7

Case Studies

Compatibility with the drilling fluid is an important consideration. Invert-emulsion oil-based


muds have become common in vertical and horizontal wells in the Williston basin. Although
these muds provide a more uniform hole size, additional problems occur when using foam
cement. The oil phase of the muds destabilizes the leading edge of the foam slurry, which
releases the nitrogen and results in free gas migration To prevent this problem, a 100- to 200sack cap slurry is pumped to contain the nitrogen In wells where salt-saturated muds are used,
foam stability is usually not a problem. On these wells, the elimination of the cap slurry has
become common. Lower concentrations of foam stabilizer and foaming agent may also be used.
An Example Bakken Shale Well
This well had been drilled with an invert-emulsion oil-based mud and 5-inch casing was to be
set to 10,000 feet (planned top at approximately 3200 feet). The bottomhole circulating
temperature was 185F. The cementing protocol was as follows.
 Pump 20 bbl of oil-based flush.
 Pump 20+ bbl of turbulent flow spacer.
 Pump 350 bbl of saltwater spacer.
 Pump 20 bbl of freshwater flush.
 Mix 150 sacks of 15.8 lbm/gal cap cement.
 Mix 16.2 lbm/gal cement, foamed to 9.5 lbm/gal downhole. This foamed slurry covered
approximately 1900 feet from the uppermost salt to approximately 500 feet above the Dakota
interval.
 Mix 16.2 lbm/gal cement, foamed to 12.5 lbm/gal downhole. This was placed in the 1600
feet between the Charles salt and the upper salts (Pine members) to prevent collapse.
 Mix a 15.6 lbm/gal Class G tail, to extend from the stage collar to above the Charles salt.
 Bump the plug.

7-48

7.13

Case Study 12
The Friction Dominated
Regime

Reference
Saponja, J.: Challenges With Jointed Pipe Underbalanced Operations, paper SPE 37066
presented at the 1996 SPE International Conference on Horizontal Well Technology, Calgary,
Alberta, November 18-20.
Background
This was a re-entry to drill a 4-inch horizontal well. Figure 7-16 shows the predicted annular
bottomhole pressure (at the bit) versus nitrogen injection rate, for a fixed surface backpressure of
51 psi. 38 API gravity oil was used as the liquid phase in the drilling fluid. The figure
demonstrates that if the liquid rate is increased, the gas-to-liquid ratio decreases and the optimum
nitrogen rate needs to be increased to accommodate underbalanced conditions. This is the crux
of multi-phase underbalanced drilling. Saponja, 1995,9 illustrated it conceptually in Figure 7-17.
7-17
Friction Versus Hydrostatic
 As the nitrogen rate is changed, at a constant liquid injection rate, the bottomhole pressure
changes because of a change in hydrostatic pressure. At low gas injection rates, the annular
friction does not affect the bottomhole pressure as strongly as does the hydrostatic head; with
a decrease in gas rate, bottomhole pressure increases substantially.
With an increase in the nitrogen rate, frictional effects become more significant and the rate
of change of bottomhole pressure decreases.

7-49

Chapter 7

Case Studies

Predicted Annular Bottomhole Pressure


at 6890 feet (psi)

2500
Oil Injection at 1.25 BPM
5-1/2" Intermediate Casing
4-3/4" Main Horizontal Hole
I ndicates Optimum N 2 Rate

2000
1.89 BPM

1500
1.25 BPM

1000
0.63 BPM

500

0
0

200

400

600

800

1000

1200

1400

1600

1800

Nitrogen Rate (scfm)

Figure 7-16.

Calculated annular bottomhole pressure for an horizontal well with varying rates of
9
nitrogen and oil injection (modified after Saponja, 1995 ).

Predicted Annular Bottomhole Pressure


at 6890 feet (psi)

2500
Oil Injection at 1.25 BPM
5-1/2" Intermediate Casing
4-3/4" Main Horizontal Hole

2000

OPTIMUM
Minimum Achievable Bottomhole
Pressure for a Specfied Rate

1500

1000

FRICTION-DOMINATED
Nitrogen is Wasted
More Stable System
Increasing Nitrogen or Influx Increases BHP

HYDROSTATIC-DOMINATED
Unstable
500
Large Pressure Changes
Gas Influx Reduces BHP

0
0

200

400

600

800

1000

1200

1400

1600

1800

Nitrogen Rate (scfm)

Figure 7-17.

7-50

Variation of annular bottomhole pressure, for an horizontal well with oil pumped at
1.25 BPM and varying nitrogen rates (modified after Saponja, 1995 9).

 In a representation of bottomhole pres-sure versus nitrogen (or other gas) injection rate (refer
to Figure 7-17), rates less than a critical value indicate an hydrostatically-dominated regime.
This minimum is the optimal circulating point. It is reached when the reduced hydrostatic
head is balanced by the increased annular friction.
 If the nitrogen (gas) injection rate is increased beyond this optimum value, the system
becomes friction-dominated.
 In planning and performing drilling operations, it is essential to determine whether the
circulating system is hydrostatically- or friction-dominated. An increase in gas injection rate
may not always cause a decrease in bottomhole pressure and excessive (additional expense)
gas volumes may be pumped.
 The impact of fluid inflow on the circulation system must be considered for its proper
design, operation limits, and control of annular bottomhole pressure. It is important to
review a large variety of reservoir fluid inflows to determine their impacts on the
circulating system and the limitations of the circulating system.9
 The consequences of operating in the hydrostatically-dominated regime in-clude:

Small changes in the gas injection rate or formation inflow can cause dramatic changes in
the bottomhole pressure.

Gas inflow can decrease the bottom-hole temperature, increasing the pressure gradient
into the well, causing the production of more formation gas, and so on.

Drastic pressure changes can occur with small changes in the gas rate downhole (injected
and/or form-ation). These pressure fluctuations impact all open zones and can cause
overbalance and/or instability.

Liquid holdup and slugging can more readily occur in the hydrostatically-dominated
regime.

 In the friction-dominated regime:

The circulation system is more stable.

Changes in gas injection rate cause smaller pressure fluctuations.

If formation gas is inflowing, the increased bottomhole gas rate will modestly increase the
bottomhole pressure, impeding inflow. Reservoir fluid inflow is minimized and
controlled in the friction-dominated regime.

At higher liquid rates, stability of the circulation system may be less of a concern and it
may not be necessary to be in the friction-dominated regime (refer to Figure 7-17).

7-51

Chapter 7

Case Studies

References

7-52

1.

Saponja, J.: Challenges With Jointed Pipe Underbalanced Operations, paper SPE 37066
presented at the 1996 SPE International Conference on Horizontal Well Technology,
Calgary, Alberta, November 18-20.

2.

Teichrob, R.R.: Low-Pressure Reservoir Drilled With Air/N2 In A Closed System, Oil
& Gas J. (March 21, 1994) 80-90.

3.

Stone, C.R. and Cress, L.A.: New Applications for Underbalanced Drilling Equipment,
paper SPE 37679, manuscript under review (1997).

4.

Joseph, R.A.: Planning Lessens Problems, Gets Benefits of Underbalance, Oil & Gas J.
(March 20, 1995a) 86-89.

5.

Joseph, R.A.: Special Techniques and Equipment Reduce Problems, Oil & Gas J.
(March 27, 1995b) 41-47.

6.

Adam, J. and Berry, M.: Underbalanced Coiled Tubing Sidetrack Successful, Oil & Gas
J. (December 18, 1995) 91-98.

7.

Wang, Z., Rommetveit, R., Vefring, E.H., Bieseman, T. and Faure, A.M.: A Dynamic
Underbalanced Drilling Simulator, presented at the 1995, 1st International Underbalanced
Drilling Conference and Exhibition, The Hague, The Netherlands, October 2-4.

8.

Purvis, D.L. and Smith, D.D.: Real-Time Monitoring Provides Insight Into Flow
Dynamics During Foam Cementing, SPEDC (June 1994) 124-132.

9.

Saponja, J.: Engineering Considerations for Jointed Pipe Underbalanced Drilling,


presented at the 1995 1st International Underbalanced Drilling Conference & Exhibition,
The Hague, The Netherlands, October 2-4.

10.

MacDougall, G.R.:
1991) 279-284.

11.

Saponja, J.: Comparing Conventional Mud Drilling To Underbalanced Drilling In A


Depleted Reservoir, paper presented at the 1995 Calgary DEA-44 Horizontal,
Slimhole/Coiled Tubing International Technical Forum, June 28-30.

12.

Comeau, L.: Underbalanced Drilling: Directional and MWD Experience, paper


presented at the 1995 1st International Drilling Conference and Exhibition, The Hague,
The Netherlands, October 2-4.

Mud/Gas Separator Sizing and Evaluation, SPEDE (December

13.

Brown, K.E. and Lea, J.F.: Nodal Systems Analysis of Oil and Gas Wells, J. Pet. Tech.
(October 1985) 1751-1763.

14.

Vogel, J.V.: Inflow Performance Relationships for Solution-Gas Drive Wells, J.Pet.
Tech. (January 1968) Trans., AIME 243 (83-92).

15.

Butler, S.D., Rashid, A.U. and Teichrob, R.R.: Monitoring Downhole Pressures and Flow
Rates Critical for Underbalanced Drilling, Oil & Gas J. (September 16, 1996) 31-39.

16.

Misselbrook, J., Wilde, G. and Falk, K.: The Development and Use of a Coiled-Tubing
Simulation for Horizontal Applications, paper SPE 22822 presented at the 1991, 66th
Annual Technical Conference and Exhibition, Dallas, TX, October 6-9.

17.

Wilson, J.: Optimizing Drilling of Underbalanced Wellbores with Data Acquisition


Systems, paper presented at the 1995, 1st International Underbalanced Drilling
Conference and Exhibition, The Hague, The Netherlands, October 2-4.

18.

Roy, R. and Hay, R.: Measuring Downhole Annular Pressure While Drilling for
Optimization of Underbalanced Drilling, paper presented at the 1995, 1st International
Underbalanced Drilling Conference and Exhibition, The Hague, The Netherlands, October
2-4.

19.

Sharman, D.M. and Pettitt, A.J.: Deployment Systems and Down Hole Swab Valves,
paper SPE 30406 presented at the 1995 Offshore Europe Conference, Aberdeen, Scotland,
September 5-8.

7-53

APPENDIX A
Normal Atmospheric Pressure at Different Altitudes
Altitude
(ft above Sea Level)

Pressure
(psia)

Altitude
(ft above Sea Level)

Pressure
(psia)

14.70

5500

11.99

500

14.42

6000

11.77

1000

14.16

6500

11.55

1500

13.91

7000

11.33

2000

13.66

7500

11.12

2500

13.41

8000

10.91

3000

13.16

8500

10.70

3500

12.92

9000

10.50

4000

12.68

9500

10.30

4500

12.45

10000

10.10

5000

12.22

10500

9.90

A-1

Underbalanced Drilling Manual

APPENDIX B
Orifice Factors, Fb, for Different Meter Run Diameters and Pressure
Tap Types
Meter Run Diameter
(inches)
Taps

Flange

Pipe

Flange

Pipe

Flange

Pipe

0.25

0.217

0.218

0.217

0.217

0.216

0.217

0.375

0.485

0.495

0.484

0.489

0.483

0.487

0.50

0.861

0.895

1.340

0.874

0.856

0.868

0.625

1.352

1.434

1.936

1.378

1.337

1.361

0.75

1.963

2.131

2.645

2.012

1.928

1.973

0.875

2.702

3.019

3.470

2.784

2.630

2.710

1.00

3.584

4.148

4.416

3.709

3.444

3.579

1.125

4.632

5.592

5.490

4.802

4.370

4.588

1.25

5.884

7.467

6.702

6.089

5.412

5.746

1.375

7.391

9.956

8.063

7.584

6.573

7.067

1.50

9.245

9.591

9.376

7.856

8.565

1.625

11.31

11.47

9.268

10.26

1.75

13.25

13.96

10.82

12.18

1.875

15.45

16.95

12.51

14.34

2.00

17.94

20.55

14.36

16.79

2.125

20.85

24.98

16.38

19.58

2.25

18.58

22.75

2.375

21.00

26.38

2.50

23.65

30.56

2.625

26.57

35.40

2.75

29.78

41.04

2.875

33.34

3.00

37.42

Orifice Diameter
(inches)

B-1

APPENDIX C
Data for calculating approximate circulation rates required to produce a minimum
annular air velocity which is equivalent in lifting power to a standard air velocity
of 3,000 ft/min (Angel, 1957)
Air
Value of N
Drilling Rate (ft/hr)
Hole
Size
(in)
17 1/2

15

12 1/4

11

9 7/8

8 3/4

7 7/8
7 3/8
6 3/4
6 1/4
4 3/4

Pipe
OD
(in)
6 5/8
5 1/2
4 1/2
6 5/8
5 1/2
4 1/2
6 5/8
5 1/2
4 1/2
6 5/8
5 1/2
4 1/2
5 1/2
5
4 1/2
5
4 1/2
3 1/2
5
4 1/2
3 1/2
4 1/2
3 1/2
3 1/2
3 1/2
3 1/2
2 7/8
2 7/8
2 3/8

Qo
(scfm)
4,209
4,428
4,588
2,905
3,124
3,285
1,700
1,918
2,079
1,237
1,456
1,616
1,079
1,163
1,240
898
975
1,103
827
903
1,032
670
798
676
535
430
494
229
271

0
82.2
79.8
78.0
71.7
68.7
66.0
62.3
58.0
55.3
60.6
54.8
50.6
53.0
50.3
47.8
49.1
46.1
41.5
49.0
46.0
40.8
44.7
39.2
38.5
37.3
37.0
32.8
31.6
27.8

30
131
126
123
112
107
103
97.8
89.5
83.6
94.5
83.8
76.9
80.3
75.5
71.7
73.0
68.5
61.0
72.7
67.8
60.0
65.0
56.7
55.0
52.8
51.5
46.0
41.3
37.2

60
177
171
166
151
143
137
130
119
111
124
110
101
104
98.7
93.3
94.4
88.5
79.0
93.2
87.3
77.3
82.7
72.5
69.8
66.1
63.6
57.3
49.5
44.8

Gas (Gravity 0.60)


Value of N
Drilling Rate (ft/hr)

90
221
213
207
188
178
171
160
146
136
151
135
124
126
120
114
113
107
95.5
112
105
93.7
98.3
86.9
83.2
78.0
74.7
67.7
56.5
51.6

Qo
(scfm)
5,434
5,716
5,924
3,751
4,033
4,241
2,194
2,477
2,684
1,597
1,880
2,087
1,393
1,502
1,600
1,160
1,258
1,424
1,068
1,166
1,332
865
1,031
873
690
555
638
296
350

0
66.3
61.8
58.0
64.2
58.6
54.0
63.0
56.3
50.8
64.5
55.5
50.0
56.4
52.3
48.8
53.0
49.0
42.0
53.5
49.1
41.8
50.1
41.6
41.6
41.5
42.0
37.0
37.0
32.3

30
128
119
113
118
108
100
112
97.7
88.2
112
95.4
84.4
94.7
87.7
81.6
87.1
80.3
68.9
87.0
80.0
68.3
78.8
66.3
65.3
63.8
63.1
55.1
51.3
45.6

60
186
174
165
167
154
144
155
137
124
152
131
116
128
119
111
116
108
93.1
115
107
92
104
87.8
85.5
82.3
80.0
71.4
62.6
56.3

90
240
226
215
214
197
185
194
172
157
188
163
146
157
147
138
141
132
115
140
130
114
125
107
104
99.0
94.7
85.4
72.2
65.5

Q (scfm) = Qo + N H (depth, thousand ft)


Example: Calculate the circulation rate required to air drill an 11-inch hole with 5-inch
drillpipe, at a rate of 90 ft/hr, at 11,000 ft.

Q = Qo + N H = 1,456 + 135 x 11 = 2,941 scfm


C-1

Underbalanced Drilling Manual

APPENDIX D
vA = = constant
2
2

1
g
M F v F dv F
M v dv
PT =
+ S S S + Pf
( M F + M S ) z +
Q F + QS
gC
Fg C
SgC
1
1

PT = ( F F + S S )

2 M F v F dv F 2 M S vS dvS
1
g
z +
+

+ Pf
gC
Q F + QS 1 F g C
SgC
1

where:
PT..........................total pressure change over an interval from Point 1 to Point 2 (in the drillpipe
through the nozzles, up the annulus and through any flow restrictions at the surface),
QF................................................................................. volumetric flow rate of fluid (gas or liquid),
QS ......................................................................................volumetric flow rate of solids (cuttings),
MF ............................................................................. mass flow rate of fluid (gas or liquid ... vA),
MS ............................................................................................... mass flow rate of solids (cuttings),
v..................................................................................................... cross-sectional average velocity,
A ........................................................................................................................cross-sectional area,
g..............................................................................................................acceleration due to gravity,
gC ..........................................................................................conversion factor (32.17 lbmft/lbfs2),
z................................................................................. change in elevation from a specified datum,
vF ............................................................................................ average fluid (liquid or gas) velocity,
vS ................................................................................................. average solids velocity (cuttings),
F ................................................... velocity profile correction factor (fluid) in kinetic energy term,
S ................................................. velocity profile correction factor (solids) in kinetic energy term,
Pf ............................................................................................................... frictional pressure drop,
F .................................................................................................................................. fluid density,
S ................................................................................................................................ solids density,
F................................................................cross-sectional average in-situ volume fraction of fluid,
S...................................................... cross-sectional average in-situ volume fraction of solids, and,
..........................................................A/A, A is the areal cross-section occupied by the phase
and A is the overall cross-sectional area.

D-1

S-ar putea să vă placă și