Sunteți pe pagina 1din 19

SPE/IADC 163492

Real-Time Evaluation of Hole Cleaning Conditions Using a Transient


Cuttings Transport Model
E. Cayeux, SPE, IRIS; T. Mesagan, SPE, S. Tanripada, SPE, M. Zidan, SPE, Statoil; K.K. Fjelde, University of
Stavanger

Copyright 2013, SPE/IADC Drilling Conference and Exhibition


This paper was prepared for presentation at the SPE/IADC Drilling Conference and Exhibition held in Amsterdam, The Netherlands, 57 March 2013.
This paper was selected for presentation by an SPE/IADC program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have
not been reviewed by the Society of Petroleum Engineers or the International Association of Drilling Contractors and are subject to correction by the author(s). The material does not
necessarily reflect any position of the Society of Petroleum Engineers or the International Association of Drilling Contractors, its officers, or members. Electronic reproduction, distribution, or
storage of any part of this paper without the written consent of the Society of Petroleum Engineers or the International Association of Drilling Contractors is prohibited. Permission to reproduce
in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE/IADC copyright.

Abstract
During a drilling operation, real-time analysis of surface and downhole measurements can give indications of poor hole
cleaning. However, it is not always intuitive to understand how and where the cuttings are settling in the borehole, because the
transportation of cuttings and the formation of cuttings beds are largely influenced by the series of actions performed during
the operation. Using a transient cuttings transport model, it is possible to get a continuously updated prognosis of the
distribution of cuttings in suspension and in beds along the annulus. This information can be of prime importance for taking
decisions to deal with and prevent poor hole cleaning conditions.
A transient cuttings transport model has been obtained by integrating closure laws for cuttings transport into a transient drilling
model that accounts for both fluid transport and drill-string mechanics.
This paper presents how this model was used to monitor two different drilling operations in the North Sea: one using
conventional drilling and one using MPD (Managed Pressure Drilling). Some unknown parameters within the model (e.g. the
size of the cuttings particles) were calibrated in order to obtain a better match with the top side measurements (cuttings flowrate, active pit reduction due to cuttings removal). Using the calibrated model, the prediction of cuttings bed locations were
confirmed by actual drilling incidents like pack-offs and overpulls while tripping out of hole.
Based on the calibrated transient cuttings transport model, it is thereby possible to evaluate what adjustments of the drilling
parameters are necessary to stop and possibly remove the cuttings beds, therefore giving the drilling team the opportunity to
take remedial and preventive actions based on quantitative evaluations, rather than solely upon the intuition and experience of
the decision makers.
Introduction
During drilling operations, ensuring proper hole cleaning conditions is of extreme importance. Otherwise serious drilling
problems can occur like stuck pipe incidents or pack-off situations, which can lead to fracturing of the formation and resulting
mud losses. The end result of bad cuttings transport is an increase in non-productive time. In order to predict how cuttings are
transported there has been performed a vast number of experimental work and different attempts on developing appropriate
cuttings transport models. An overview of some of the work that has been performed is given by (Pilehvari et al. 1999). It has
turned out to be quite complex to describe the cuttings transport process since transport is influenced by many different
parameters like for instance wellbore geometry, inclination, fluid density, rheology, rate of penetration, drill-string rotation,
flow patterns, flow-rate and cuttings size. One could divide the modeling approach into two main classes: i.e. the empirical
approach where correlations are developed based on experimental data (e.g. Larsen et al. 1997) or use of more fundamental
mechanistic approaches (e.g. Gavignet and Sobey 1989). The best approach may still not be decided. The models developed
are usually based on steady state considerations. However in a drilling operation, there are transient phenomena taking place
regularly that affect the cuttings transport in the well as well as cuttings bed evolution. Examples of such phenomena are
connections, temperature changes, variations in rate of penetration, flow-rate and drill-string rotation adjustments. Hence in
order to describe the cuttings transport process in an adequate manner it should be associated with the transient conditions
taking place in the well. In the following, we have utilized an empirical correlation for cuttings transport to build a transient

SPE/IADC 163492

drilling model which both accounts for fluid transport and drill-string mechanics. The chosen cuttings transport model is the
one developed by (Larsen et al. 1997) including the extensions proposed by (Jalukar 1993) and (Bassal 1995). One can also
note that inclusion of such a model will also improve the predictive capability of the transient drilling model, since the
physical processes taking place are interdependent. The combined model has been validated against two field cases and the
results of this will be presented in the paper. When comparing drilling models with real operational data, there is often a need
for performing some sort of calibration of the models to improve the predictive capability. In this work, the sensitivity of the
results to the cuttings particle size and a bed erosion factor have been analyzed to determine in which conditions the output of
the model did match the observations made during these drilling operations.
Cuttings Transport Model
Transient Drilling Hydraulics
Typically a drilling fluid is composed of a liquid phase (a solution like a brine or an emulsion of water in oil) with several
types of solid in suspension: low gravity solids (e.g. bentonite), high gravity solids (e.g. barite) and cuttings, the last
component is only relevant within the annulus. It should also be noted that gas could be present in the drilling fluid, but for the
sake of simplicity, the presented mathematical description will not include any gaseous phase. The hydraulic modeling of this
multi-phase fluid uses the drift-flux model, i.e. the fluid inside a control element is considered to be homogenous even though
the different phases of the fluid mix have relative velocities to each other. The average velocity of the mud in the axial
direction (approximation to a one dimensional flow) can be expressed as a function of the slip velocities of the various internal
phases:
(1)

where is a set of indices (lgs for low gravity solids, hgs for high gravity solids, c for cuttings), is the mean velocity of the
external phase (liquid), is the volume fraction of the i-phase, is the mean slip velocity of the i-phase compare to .
There are three balance equations that describe the interface exchange of mass, momentum and heat (Lage et al. 2003).
The mass balance (continuity equation) can be written:

(2)

where is time, is the curvilinear abscissa along the wellbore, is the cross-sectional area of a fluid element,
is the
averaged density,
is the average velocity, is the source term (a mass per length per time through the fluid element side
walls). Note that the source term includes the fact that cuttings particles can settle in a bed or return to suspension from a bed.
The presence of a cuttings bed changes the available cross-sectional area for the passage of the fluid in the annulus. Typically
it is considered that A changes with time because of the movement of drill-string elements. For instance, at any place along the
drill-string where there is a change of the outside diameter in a drill-string element, the axial movement of the drill-string
causes a variation of the annulus cross-sectional area at the position where this change of diameter in the drill-string element
occurs. In addition, it should be noted that the variation of cuttings bed size through time also affects the cross sectional area
(see Fig. 1).

Fig. 1: The mass exchange happens at the entry, exit but potentially also along the control volume through cuttings beds.

The momentum balance derived from the Navier-Stokes equation is written as follow:

SPE/IADC 163492

t, s

t, s

t, s

(3)

where is the pressure, K is the friction pressure-loss term, is the average inclination of the control fluid element, is the
gravitational acceleration of the earth. Eq. 2 and 3 describe the fully transient behavior covering both pressure pulse
propagation and mass transport as illustrated in Fig. 2. In this figure the mud pumps are accelerated smoothly from 0 to 1500
l/min. One can see how the fluid velocity wave propagates first in the drill-string and then in the annulus. We can also see how
the downhole pressure gradient increases as a consequence of the fluid movement when the velocity wave propagates in the
annulus.

Fig. 2: Transient hydraulic effects during the acceleration of the mud pumps.

Finally, the energy conservation due to heat transfer can be written (Marshall and Bentsen 1982):

(4)

where is the enthalpy per mass unit, Q is the forced convective term, Q is the conductive and natural-convective term,
the heat generated by mechanical and hydraulic frictions.

Fig. 3: The different layers participating in the heat exchange in a control volume

The forced convective term can be expressed:


,

(5)

is

SPE/IADC 163492

The conductive and natural-convective term does not have a general expression. In the case of purely convective isotropic
material, we can use:
,

(6)

where is the thermal conductivity, T is the temperature. The heat exchange in a control volume is occurring between
different layers: fluid inside the drill-string, drill-string walls, fluid in the annulus, casing walls, cement, formation (see Fig. 3).
It is therefore a two-dimensional model in a curvilinear cylindrical coordinate system with an assumption of uniformity for the
angular dimension. Note that the thermal conductivity of the mud is affected by the concentration of cuttings. Furthermore, the
presence of cuttings beds is also changing the thermal conductivity in the annulus since heat transfer can happen between the
wall of the drill-string and the casing or formation through the cuttings bed instead of the annulus fluid.
The partial differential Eq. 4 describes the transient variations of the drilling fluid temperature in the drill-string and in the
annulus as illustrated by Fig. 4. Because of the slow evolution of heat transfer processes and the relatively rapid change in
drilling parameters, steady state conditions are, in practice, never reached.

Fig. 4: Evolution of the temperature when establishing circulation from geothermal conditions

Cuttings in Suspension
The presence of cuttings in suspension in the drilling fluid changes the properties of the original drilling fluid. This affects
the density, viscosity, specific heat capacity and thermal conductivity of the mud during the transport of the cuttings.
Drilling Fluid Density
In this model, the liquid phase will be considered compressible while the solid phase will be treated as incompressible. The
liquid phase can either be a solution (brine or diluted cesium formate) or an emulsion (a colloid system like water dispersed in
oil). The solid phase can have at most three components, namely high gravity solids, low gravity solids and cuttings.
The density of a drilling fluid is given by the combination of the densities of its components weighted by the volume
fraction of each element:
(7)
1
,

, , with

where m is the density of the mud, p the pressure, T the temperature, is the set of indices for the different components (g for
gas, w for water or brine, o for oil, lgs for low gravity solid, hgs for high gravity solid and c for cuttings), fi is the volume
fraction of the i-component and i is the density of the i-component.

Fig. 5: Effect of compression on the solid volume fraction.

SPE/IADC 163492

If we consider that the volume fraction of cuttings


is known at a given temperature
and pressure , then to
determine the volume fraction of solids at a different pressure and temperature, we should consider that the solid phase is
actually incompressible but the dispersion medium (oil/brine) is actually compressible and dilatable. With changes of pressure
and temperature, the volume occupied by the dispersion medium reduces or expands. As a consequence the distance between
the solid particles changes and therefore the volume fraction of the solids (see Fig. 5).
The volume fraction of solids can be expressed as a function of the initial volume fraction and the local temperature and
pressure as follow:
,

,
1

(8)
,

is the density of the mud at the reference pressure p1 and temperature T1. Fig. 6 shows how the local density of the
where
mud in the annulus is affected by the presence of cuttings. In addition, the local pressure and temperature at a given depth also
influences the density of the mud. It should be noticed that when there is circulation and the bit is off bottom, the cuttings
already under transport are still moving up while no new cuttings are produced at the bottom hole, therefore causing large
variations in the local mud density as a function of depth.

Fig. 6: The local density of the mud increases with the concentration of cuttings. During period of circulation without production of
cuttings, the local mud density is back to normal.

Rheology
Drilling muds are non-Newtonian fluids, they are more precisely shear thinning fluids with a yield stress. A typical
rheological model that describes satisfactorily such liquids is the one of Herschel and Bulkley. However a rheological
formulation has been proposed (Robertson and Stiff 1976) that has better properties than the one of Herschel and Bulkley to
describe the drilling hydraulic flow:
(9)
where is the shear stress, is the shear rate, A, B and C are the coefficients of the model. This is the model that is used in the
described transient hydraulic model. It should be noted that as with any other fluid, the rheology of drilling mud depends on
temperature. In addition, the mud viscosity increases exponentially with larger pressures, this being true at any temperature
(Houwen and Geehan 1986), following an Arrhenius type of law (see Fig. 7).

Fig. 7: The viscosity of drilling fluids decreases when temperature increases but increases when pressure increases.

SPE/IADC 163492

Furthermore, the cuttings in suspension modify the effective viscosity of the mud. The effect of solid particles in
suspension on the rheology of fluids was first studied by Einstein in 1906 (Einstein 1906). This analysis is based on an energy
balance to determine the viscosity of a suspension of solids in Newtonian liquids:
,

2.5

0, 0.54

(10)

where
is the viscosity of the drilling fluid with the cuttings,
is the original viscosity of the mud and is the volume
fraction of cuttings in the mud. Hastchek has proposed (Hastchek 1910) a modification of Einsteins equation based on Stokes
law for a slowly falling ball in an infinite continuum. One major advantage of this model is that it is valid for any
concentrations of particles:
,

,
,

1
,

4.5
1

,
,

0, 0.74

(11)

0.74, 1

The viscosity of the original mud can be derived from the rheological constitutive law of the drilling fluid using:
,

(12)

Specific Heat Capacity


The theoretical specific heat capacity of the multi component medium is the weighted average in terms of mass fractions of
the specific heat capacity of each component:

(13)

, with ,

is the specific heat of the drilling mud, , are the mass fractions of the components, , are the specific
where
heat capacities of each components and is a set of indices representing each component (w for water or brine, o for oil, lgs
for low gravity solid, hgs for high gravity solid and c for cuttings). The mass fractions are derived from the volume fractions
using the density of the components.
Thermal Conductivity
The calculation of the effective thermal conductivity in a heterogeneous multi-component medium is not simple. The first
solution to the calculation of the effective thermal conductivity of a suspension of solid particles in a homogeneous medium
was described by Maxwell and was based on the assumption that the particles can be assimilated to spheres and that their
concentration is small. For normal weighted drilling fluids, the concentration of solid particles cannot be considered as small.
Rayleigh has extended Maxwells model to higher order of concentration by considering spherical particles, yet non-touching,
but distributed on a regular cubic lattice. Rayleighs model being less and less accurate when the mass fraction of particles in
suspension reaches 0.5236, it has been modified by Churchill to circumvent that problem (Kandula 2011). The resulting
expression of the effective thermal conductivity for the mud is then:

where

2
1
2
1

6 3
4 3
6 3
0.409
4 3
0.409

3 3
4 3
3 3
0.906
4 3
2.133

(14)

is the solid to liquid thermal conductivity ratio.

The background continuous medium, i.e. the liquid phase, can be a mixture solution (typically oil and brine) of different
fluids with different thermal conductivity. A first approximation of the effective thermal conductivity of the liquid mixture is:
(15)

where is a set of indices for the liquid components (typically o for oil and w for water/brine), is the mass fraction of the icomponent relative to the mass of the liquid mix, and
is the thermal conductivity of i-component of the liquid. Rowley
(1982) reported that experimental measurements showed that the weighted average method resulted in an excess thermal

SPE/IADC 163492

conductivity. This can be compensated by adding a correction term in the weighted average formula:
(16)

where

, ,

(17)

in which
,

, ,

(18)

are the NRTL parameters and R is the


is the non-randomness factor of the non-random two-liquid (NRTL) model and
universal gas constant. These parameters can be found in the literature for mixes of 2 liquids (see Nagata 1973; Gmehling et
al. 1977) or from the Dortmund Data Bank.
To calculate the effective thermal conductivity of the mud with several types of particles, one should, rigorously, use an
extension to the Churchill model to n-dispersed phases embedded in one continuous medium. Such an extension exists for the
Maxwell model (made by Eucken in 1938), but it does not exist for the Churchill model. Even though it is not exactly correct,
it is possible to use, as a practical approximation, the Churchill formula recursively taking one component at a time and using
the resulting thermal conductivity as the thermal conductivity of the background medium when adding the next component.
Cuttings Transport and Settling
A cuttings particle in suspension within a moving or dynamic drilling fluid is subject to several forces. The first set of
forces are the static ones: the gravity force (Fg) and the buoyancy force (Fb). In addition, there is a dynamic force: the
frictional force (Clark and Bickham, 1994). The frictional force is decomposed into a drag force (FD) following the flow
direction and a lift force (FL) perpendicular to the flow movement (see Fig. 8). The drag and lift forces depend on the local
velocity of the fluid around the cutting particle. If we consider only a bulk fluid velocity (averaged over the whole crosssection), then it is possible to calculate a terminal velocity, compared to the fluid, for a cuttings particle. If the component of
the terminal velocity in the direction of the wellbore axis is positive, then the cuttings particle is transported.

Fig. 8: Force acting on a cuttings particle in suspension

It should be noted that the fluid velocity field is not uniform in the cross-section of the annulus, and close to the walls or on
the low side when the drill-pipes are decentered (i.e., for an inclined section), the local fluid velocity may therefore not be able
to transport the cuttings even though it is sufficient on the high side of the annulus. As a consequence, the cuttings particles

SPE/IADC 163492

that do not enter the main fluid stream, are not transported, and they then settle into a bed. If the drilling fluid flow is turbulent,
there is the potential that such cuttings bed particles may be lifted in the main fluid stream and therefore be carried away. A
similar effect can be obtained by the stirring effect caused by the rotation of the drill-string. Such a rotation is causing a
secondary velocity field for the drilling fluid, perpendicular to the axis of the annulus, which should be combined to the axial
velocity field. This can lift up cuttings particles from the low side into the main stream. The rotation of the drill-pipe may
cause local turbulences that do not necessarily exist in the main stream. This turbulence effect can also help the particles to
move into the main fluid course where they can then be sent downstream. So a method based on terminal particle velocity is
only valid when the fluid velocity across the section is relatively symmetric around the wellbore axis (i.e., little effect of drillpipe eccentricity and drill-pipe rotation). These conditions correspond to vertical or low inclined part of the well. Experimental
studies indicate that the inclination limit is around 35deg (Tomren et al. 1986).
Experiments have confirmed that the eccentricity of the drill-pipe inside the annulus as well as the drill-string rotational
velocity play an important role in the actual cuttings transport for the more inclined part of the hole. Nazari et al. (2010) gives
an exhaustive overview of the dominant factors influencing the transport of cuttings, including the two mechanisms mentioned
above. Unfortunately, according to Ozbayoglu et al. (2008) a mechanistic relationship including both rotation and inclination
has not yet been published.
There are three possibilities to account for this complex phenomenon: use an empirical model based on relevant
experimental data, develop a physical law using statistical physics methods or synthetize the results of many simulations made
with a detailed computational fluid dynamics (CFD) model of the problem, like the work done on annular pressure loss around
stabilizers presented by Yao and Robello (2008). In a first approach, we have chosen a method based on empirical correlations
because of readily available results.
Indeed, starting in the early 1980s, several experiments have been performed to develop correlation models of the
minimum critical fluid velocity necessary to transport cuttings in various conditions. Ford and Peden of the Heriot-Watt
University, UK, were amongst the first to present their results (Ford et al. and Peden et al. 1990). Shortly after the University
of Tulsa, Oklahoma, USA, started an extensive and systematic series of experiments in a large scale flow-loop. From the
experimental data they extracted an empirical model first described by Larsen et al. (1997). This initial model has been
modified a first time to better account for the effect of hole size (Jalukar, 1993), and then a second modification has been made
to include the effect of the drill-pipe rotational velocity, since this effect was not accounted for in the initial study (Bassal
1995). A few years later, another independent study has been presented by the Bandung Institute of Technology, Indonesia
(Rubiandini 1999).
A systematic analysis of the models proposed by Larsen and Rubiandini has shown that Rubiandinis model, in all tested
configuration, provides larger critical transport fluid velocities (CTFV) than the one of Larsen (Ranjbar 2010). In the work of
Ranjbar, the mechanistic approaches of Gavignet and Sobey (1989), and Kamp and Rivero (1999) were also described. As
reported by Kamp and Rivero (1999), their model gave quite small/unreasonable CTFV. The Larsen and Jalukar model
predicted cuttings bed build up at a flow-rate which was ten times higher.
Practical use of the Larsens model in real drilling operations has shown that this model is usually conservative, but the two
extensions made by Bassal and Jalukar have not been accounted for either in the referred practical drilling operations or in the
analysis made by Ranjbar. However, those extensions are quite central to model the effect of drill-pipe rotation in various hole
sizes and should not be disregarded.
In this paper, we have used the Larsens model combined with the extensions of Jalukar and Bassal, to calculate the CTFV
along the annulus. The resulting correlation model accounts for the effects of: inclination, fluid velocity, fluid rheology, mud
weight, cuttings size, cuttings concentration, drill-pipe eccentricity, drill-pipe rotational velocity and ratio between hole size
and drill-pipe size. It should be noted that this model is based on approximately 2000 experiments. However, the inclination
used while acquiring the data has been varied between 55deg and 90deg, Furthermore, the correlation on the drilling fluid
apparent viscosity is only expressed through a Bingham Plastic rheology (all the experiments have been made with waterbased muds). With similar rheology, the type of mud has little impact on the CTFV (Hareland et al, 1993). However, cuttings
bed erosion behaves differently in oil-based mud, when compared with water-based mud. Since the extended Larsens model
is only used for calculating the CTFV, the lack of correlation with mud type should not impact the results too much. Even
though the extended Larsens model has been developed using data from experiments with an inclination above 55deg, the
estimation of the CTFV is considered to be valid from inclination angle as low as 35deg.
It should be noted that the CTFV value calculated by the extended Larsens model corresponds to conditions where there is
no cuttings bed forming at all. Therefore when the local fluid velocity is lower than the CTFV it can be expected that cuttings
will settle. In such conditions, for a given control volume, the concentration of cuttings in suspension shall be updated as well
as the height of the bed. In the opposite case, if the fluid velocity is greater than the CTFV, the cuttings that were trapped in a
cuttings bed return to suspension. This also has a consequence on the local (i.e., at each depth along the annulus) concentration
of cuttings in suspension and the cuttings bed height. In cases where several particle sizes are used, the depositing and resuspension of cuttings particles may concern some of the cuttings particles but not all of them. It is therefore necessary to
account for the local cuttings size distribution of the particles in suspension and those laying in a bed. Cuttings particles
continue to accumulate in a bed until the free cross-sectional area has been reduced in such a way that the fluid velocity is
larger than the CTFV. At that moment, the cuttings bed height limit has been reached (Clark and Bickham 1994).

SPE/IADC 163492

To calculate the size of the cuttings bed in the annulus at a given depth, we need to account for the actual packing
efficiency of particles in the bed (i.e., the ratio of the actual particle volume to the occupied volume). For the mono-dispersed
packing problem (a single size), the maximum packing efficiency is
0.74048, which is the highest possible density

amongst all possible lattice packing, as demonstrated by Gauss in 1831. In practice, when the spheres are added randomly, the
packing is irregular and the maximum achievable density is lower than the best lattice packing. It has been demonstrated that
with jammed packing, the packing efficiency cannot exceed the limit of 63.4% in the most compact way and 55% for loose
packing (Song et al. 2008). The poly-dispersed (n-components mixture) packing problem is extremely complex. In that case,
we consider that the different particle sizes are stacked on top of each other with the largest particle at the bottom and the
smallest at the top. However for the binary hard sphere packing problem (two sizes) with coarse and fine particles, there exist
solutions for calculating the packing efficiency of the mixture of particles. Zheng et al. (1995) have proposed:
(19)
1
ln
where
is the packing efficiency of the mix of coarse and fine particles,
is the packing efficiency of the coarse
is the packing efficiency of the fine particles,
is the volume fraction of fine particles, r is the size ratio of
particles,
coarse particles to fine particles and e is Eulers number.
All these packing efficiencies are for a very large region so that the effect of boundaries is insignificant. This hypothesis is
true as long as the cuttings size is small compared to the radius of the borehole. Thereafter, it is possible to determine the area
occupied by cuttings ( ) at a given depth by accounting for the packing efficiency. For a mono-dispersed system or a binary
dispersion with coarse and fine particle:
(20)
6
where n is the number of particles in the control volume, d is the diameter of the particles, PE is the packing efficiency and L
is the length of the controlled volume. For the case of a multi-dispersed cuttings bed, the area occupied by the cuttings is:
(21)

6
where k is the number of different particle sizes, is the number of particles of the i-size in the controlled volume, is the
diameter of the particles of the i-size and
is the packing efficiency for the particle of the i-size. Then it is possible to find
the free area in the annulus:
(22)

where

is the free area in the annulus cross-section,

is the radius of the wellbore and

is the radius of the drill-pipe.

Fig. 9: The different configurations for the cuttings bed and the drill-pipe

Finally, to find the height of the cuttings bed (see Fig. 9) one needs to solve the following piecewise equation:

acos

(23)

10

SPE/IADC 163492

acos
acos

acos

During transportation, the cuttings size changes as a result of mechanical interactions caused by grinding. This
phenomenon is clearly observable when using a dual concentric drill-string drilling method (Belarde and Vestavik 2011). With
this new drilling method, a dual concentric drill pipe is used where cuttings transport takes place inside the inner drill pipe
rather than the annulus. Alternating from conventional drilling to the dual-concentric drilling method within the same drilling
operation, significant variations in size and shape of the cuttings (see Fig. 10) were observed. This process is poorly
understood and we have not attempted to incorporate any modeling of that fact in the presented model, even though it probably
has important consequences on the cuttings transport and on the evolution of cuttings beds.

Fig. 10: Comparison of cuttings sizes from conventional drilling and Reelwell drilling method (courtesy Belarde and Vestavik 2011)

Cuttings Bed Erosion


When a cuttings bed is formed, it is subjected to erosional forces due to the flow of drilling fluid in the annulus. Laboratory
experiments have shown three possibly intertwined modes of streambed erosion (Ramadan et al. 2001): individual particle
entrainment at the surface of the bed (also called surface erosion), bed surface layer displacement (called mass erosion) and
entire bed fluidization resulting in a bulk movement of the bed. In the first mode of erosion, the movement of individual
particles at the surface of the bed is the result of forces acting on each particle. In that case, there are basically three
possibilities: the lift force, due to mud flow, on the particle is large enough to temporarily extract the particle from the bed; the
rotation of the drill-pipe creates an additional lift force that can initiate the temporary entrainment of the particle downstream;
the drag force, induced by the fluid flow, enables the particle to roll on top of the neighboring particles. Mass erosion occurs
when the entire top layer of the cuttings bed starts moving because the shear stress across the bed is larger than the cohesive
strength of the structure, therefore causing a plane of failure separating the top part of the bed from the rest. In the third mode
of erosion, the whole bed behaves like a fluid.
We will focus on the surface erosion mode. The mud flow induced by pumping or by the axial movement of the drill-pipe
creates a fluid velocity field in the axial direction of the wellbore that is referred as . The rotation of the drill-pipe creates a
fluid velocity field perpendicular to the wellbore direction, at least if circular Couette flow can be assumed. If the Reynolds
number is high enough, then Taylor vortices will exist, resulting in much more complex local velocities. For the moment, we
will consider a laminar circular Couette flow and denote the tangential component of the fluid velocity . The local fluid
velocity
is therefore the sum of the tangential fluid velocity and the axial velocity (
). The angle of the
wellbore axis is denoted and is in fact the inclination of the borehole at that location. The static forces acting on the particle
are: the gravity force, the buoyancy force, the plasticity force and 3 reaction forces
,
and
, since the particle
has three contact points with the particles in the layer below. The dynamic forces induced by the fluid movement are:
the
the lift force. The particle can be dislodged from its position either by lifting, i.e. the reactive forces become
drag force,
zero, or by rolling, i.e. the momentum of the resulting forces relative to an axis of rotation defined by two of the three contact
points is greater than zero. The positions of the points of contact are related to the separation distance between the particles
within a layer, and for the sake of simplicity it will be assumed that this distance is isotropic.

SPE/IADC 163492

11

Fig. 11: Forces acting on a single cuttings particle

Choosing a coordinate system with axes dictated by the wellbore direction (see Fig. 11), we can express the condition for
lifting as (note that the reaction forces are in this case 0):
(24)
0
and the condition for rolling:
(25)
0
where is the vector between the axis of rotation (the line defined by two of the three contact points) and the center of gravity
and is the cross product of vectors. Note that the reaction force at the opposite contact point to the rotation axis is zero when
the rolling starts and that the reaction forces at the contact points on the rotation axis do not contribute to the momentum
because of the zero length of the moment arm for those forces. The magnitude of each force has been defined in different ways
by several authors (Gavignet and Sobey 1989; Clark and Bickham 1994; Doron et al. 1997; Kamp and Rivero 1999; Ramadan
et al. 2001).
It is however difficult to evaluate the 3-dimensional fluid velocity around a particle in the general case. Using statistical
physic methods or CFD simulations as explained above, it could be possible to derive a general law for the three-dimensional
fluid velocity field as a function of the flow-rate, rotational velocity of the drill-pipe, geometrical parameters of the system and
the rheology of the drilling fluid, but that is an enormous task. We have adopted a simpler strategy. Noticing that the above
equations (24) and (25) take a much simpler form when there is no drill-pipe rotation since we know that the fluid velocity
field is parallel to the wellbore axis, we can deduce the erosion rate of the top layer of the bed (noted , with the dimension
L2T-1). This corresponds to the worst case scenario where there is no assistance from drill-pipe rotation. On the other hand, the
extended Larsens model can be used to find, for the given flow-rate, the necessary rotational velocity (noted
) of the drillpipe that will assure that particles cannot settle in a bed. Then an erosion rate as a function of the drill-pipe rotational
velocity can be defined as:
(26)
,
If the used flow-rate is insufficient to provoke any erosion of the top layer of the cuttings bed, we search for the minimum
flow-rate ( ) that causes either lifting or rolling of the particles on top of the cuttings bed while not rotating the drill-pipes.
Using the same procedure as described above, an erosion function is found for that flow-rate. The generated erosion function is
translated proportionally to the difference between the used flow-rate and the minimum flow-rate to initiate erosion:
(27)

,
where the constants A and B are functions of (

) and shall be chosen such that they are equal to 0 when

A Designer Well in the North Sea


We will review two cases of poor cuttings transport conditions in a well recently drilled in the North Sea. This well was
geometrically challenging and matched the definition given by Blikra et al. (1994) of a designer well. According to this
definition, a designer well is a well containing a combination of several of the following elements:
1) Significant turn in the horizontal plane (e.g. 30 to 150 degrees azimuth) at high deviation.
2) A combination of turns both right and left (i.e. snaky).
3) Turn not restricted by inclinations.
4) The ability to place the borehole, as dictated by geology.

12

SPE/IADC 163492

The well that we will use for those two examples was planned to be sidetracked from a mother well that was drilled in the
90s. The sidetrack was planned to be drilled in two sections (see Fig. 12).
12 x 13 Section: The first section was a 12 x 13 section. This section was kicked off from the mother well
using a previously installed whipstock in a sand-free formation. The whipstock was set in the 13 3/8 casing. A window was
milled through the 13 3/8 casing. This section was planned to be drilled in one run by using a rotary steerable system and a
PDC bit. An under-reamer was planned to be used to open up the hole to 13 .
The well path started with a build at a rate of 3/30 m to reach an inclination of 65, then continued as a tangent to the
section TD. Water based mud was used with a density of 1.55 sg at the beginning, later increased to 1.62 sg towards the end of
the section. At the end of the drilling operation, a 10 liner followed by a liner 9 5/8 was set inside the 13 3/8 casing.
8 x 9 Section: The second section was drilled with an 8 PDC bit and a 9 under-reamer. This was the last
section for this well and it was drilled in one run by using a rotary steerable system. In this section, the well path was steered
towards the targets with a maximum dogleg severity of 3/30 m. Geo-steering was used while drilling some intervals of this
section. The drilling fluid was an oil-based mud with a density of 1.74 sg.

Fig. 12: Trajectory and wellbore architecture of the test well

Summary of Operations: The cement plug was drilled, reaming back every meter. The BHA was pulled into the casing
shoe and a XLOT was performed. The formation was fractured during the XLOT and a cement stinger was run to set a new
cement plug.
A second attempt was made to take the XLOT but the BHA got stuck in the fresh cement. The drill pipe was worked out
and circulated until the string was freed. At the third attempt, multiple pack-offs were experienced before reaching the
formation. Eventually, the bottom of the cement plug was fractured and mud losses occurred. A LOT was nevertheless
performed, but it was uncertain of the quality of the results, therefore a new cement plug was set and another XLOT taken this
time without any doubt about the quality of the measurements. It was decided to set another cement plug and drill the section
using managed pressure drilling. Thereafter the 8 x 9 section was drilled successfully (see Fig. 13).

Fig. 13: Layout of the 8 x 9 BHA

First Use Case: Poor Cuttings Transport while Drilling a Cement Plug
Description of the Incident
During the third drill-out cement operation, there was a very poor cuttings transport rate, which was easily observed from
the differences between the expected active volume decrease and the real one (see Fig. 14). At 13:00, the active volume should
have dropped by at least 5m3, based on pure cuttings removal, and more reasonably by 10m3 if one accounts for the expected

SPE/IADC 163492

13

loss of mud around the cuttings particles. There were some cuttings coming to the surface, but not enough when compared to
the modeled well cuttings return. Therefore it was obvious that larger and larger cuttings beds were forming in the borehole.
(Cayeux et al. 2012)

Fig. 14: This time-based log shows that the expected drop in active volume (rainbow filled region) due cuttings removal never
happened (measured volume is displayed with blue solid line). One can also see that the expected cuttings flow-rate at surface (green
curve in cuttings flow-rate track) is much higher than the actually measured cuttings flow-rate (blue curve in the same track).

Analysis with the Transient Cuttings Transport Model


The aforementioned transient cuttings transport model was used to simulate the cuttings transport behavior during drilling
of the cement plug. The drill-out cement operation was performed using flow-rates around 1900 l/min, 80 RPM for the drillstring rotation, and an average drilling rate of 15 m/h.

Fig. 15: Wellbore conditions after nine hours drilling the cement plug.

Transport of different cutting particle sizes ranging from 2 mm to 6 mm was simulated in order to analyze the sensitivity
of the results on the cuttings size in this particular operation. The results of the simulations were matched with the observations
made during the actual drilling operations. From the simulations, it was clear that cuttings beds formed regardless of the size
of the particles but that the dimensions (heights, lengths) of the modeled cuttings beds varied with the depth position in the
annulus and the cuttings particle size. All the simulations resulted in no cuttings transport back to the surface with the actual
drilling parameters used in the drilling operation.

14

SPE/IADC 163492

Further analysis showed that the drilling parameters used during this operation were not sufficient to transport the cuttings
particles. The main factor that could have improved the transport capability was the flow-rate. However, it should be noted
that the fluid velocity around the BHA was sufficient, and therefore created enough lifting capability to transport the cuttings
properly in that particular area of the well bore (see Fig. 15). This is due to the very small area between the BHA and the 9 5/8
liner. But when the flow passed through a larger annular cross-section, the fluid velocity decreased and the cuttings deposited
in those sections of the well. The larger the cross-sectional area in the wellbore, the greater the height of the resulting cuttings
bed. From the start of the drill-out cement operation, the lack of lifting capability caused cuttings beds to accumulate in the
inclined part of the wellbore.
The evolution of the cuttings bed during drilling of the cement plug is presented in Fig. 16 for different cuttings sizes. The
simulations pointed out that smaller particles were more difficult to transport than the larger one. For small cuttings sizes the
cuttings bed was thick even inside the 10 liner. A more elongated and evenly distributed cuttings bed shape was observed
when simulating with medium and large cuttings particles.
Close to the end of the drill-out cement operation, a huge amount of cuttings was accumulated in the inclined part of the
wellbore. This situation most likely led to a sudden cuttings avalanche, which is likely to be the cause of the major drilling
pack-off incident that occurred during this operation.

Fig. 16: Evolution of cuttings bed development during the cement drill-out operation for different cuttings sizes.

Recommendations
Previous studies have shown that flow-rate is the key parameter affecting hole cleaning and cuttings transport performance.
The use of high flow-rate improves significantly the material transport, but at the same time it increases the risk of fracturing
the formation due to high downhole pressure. Since the drill-out cement operation was performed inside a closed annulus
without any exposure of the drilling fluid to any formation rocks, it would have been possible to increase the flow-rate well
above 1900 l/min without risking formation fracturing, at least until getting close to the end of the cement plug.
Furthermore, increasing the drill-string rotation could have improved the cuttings transport performance. High drill-string
rotation disturbs the cuttings bed and therefore contributes to redistributing the cuttings bed into suspension. However, this
would not necessarily be a good option since the operation was carried-out while the BHA was located inside a liner
Reducing the drilling rate would have been another option to improve the cuttings transport by counting on cuttings bed
erosion to transport the cuttings. Simulations made with the transient cuttings transport model show that using standard
parameters for the cuttings bed erosion factors, it would have been necessary to use extremely low ROP to achieve that effect.
This is due to the mud having a relatively high viscosity.

SPE/IADC 163492

15

Second Use Case: Hole Cleaning Problems during a MPD Operation


Indications of Poor Hole Cleaning
During the first days of the managed pressure drilling operation, indications of poor hole cleaning were visible. The
amount of cuttings that came back to the shakers was insufficient when compared with the modeled cuttings returns. This was
observed by comparing the actual active volume variations with the expected ones based on the arrival time of cuttings to the
surface. At the same time the downhole ECD steadily increased during continuous periods of drilling. However, at several
instances the drilling was interrupted for several hours to perform pore pressure tests while maintaining circulation. During
those periods, where no new formations were drilled, the downhole ECD decreased to normal values. After four days of such
variable drilling conditions, it was decided to reduce the ROP from 15m/h to 10m/h. Thereafter, no more problems with
cuttings transport were noticed.
Results from Transient Cuttings Transport Simulations
Three cuttings size distributions were studied with the transient cuttings transport model using the real-time drilling
parameters of the MPD operation. The accumulation of cuttings in the regions where the internal diameter of the wellbore
changed was confirmed by the simulations (liner hanger). As in the previous case, cuttings beds developed in all studied cases.
However, the shapes and lengths of those cuttings beds were different than in the previous case. This was attributed to the use
of a drilling fluid with a much lower viscosity than in the drill-out cement operation. With a lower viscosity, it is easier to
attain turbulence and therefore to increase the chance of redistributing the cuttings in the cross-sectional area of the annulus for
a better transport.

Fig. 17: Effect of flow-rate on hole cleaning and cuttings transport.

The effect of increasing the flow rate was seen in the simulations (see Fig. 17). There was an improved transition in the
hole cleaning index as the flow rate was increased. Similarly, the effect of increasing the pipe rotation was also seen in the
simulations (see Fig. 18). As a consequence the geometry of the cuttings bed is changed, as well as the hole cleaning index.

Fig. 18: Effect of hole flow rate on hole cleaning and cuttings transport

The sensitivity of the simulations to cuttings beds erosion factor (an internal calibration parameter of the model) was also
performed. Two simulation results are presented using a small factor and a factor five times larger (see Fig. 19). The larger
value of this calibration factor gave concordant results with the observed cuttings flow-rate during the drilling operation. The

16

SPE/IADC 163492

estimation of the actual cuttings flow-rate was done by analyzing the rate of change of the active pit volume during the
operation when the ROP was reduced to 10m/h.

Fig. 19: Effect of erosion rate on cuttings transport

Analysis
At the start of the MPD operation, a flow rate of 1800 l/min was used. After some time, it was increased to about 2100
l/min and there were improvements in relation to cuttings transport. From the simulations, it was seen that even at that flowrate, hole cleaning was not efficient. Since we had an MPD operation where pressure in the well was constantly monitored and
kept under control, it could have been possible to increase the flow-rate and compensate for the additional pressure losses in
the annulus by changing the settings of the MPD choke.
The pipe rotation was also critical. An increase in rotation from 60 rpm to 100 rpm showed an improvement, as most of the
cuttings in the beds were eroded and lifted up into the fluid flow. The increase in rotation coupled with an increase in flow rate
led to the disappearance of the cuttings bed within the 10 liner.
The reduction of the ROP was a viable choice as seen from the successful operation, however it substantially increased the
duration of the operation.
Conclusion
In the present work, a fully transient cuttings transport model was developed by integrating an empirical correlation for
critical conditions of cuttings transport into a transient drilling model which both accounts for fluid transport and drill-string
mechanics. The model was calibrated against real data by adjusting parameters such as cuttings size and cuttings bed erosion
factor using the measured cuttings rates at surface as the best fit criteria. A transient cuttings transport model makes it possible
to better predict the downhole conditions since it is able to represent phenomena that evolve over time such as cuttings bed
buildup or removal. Effects related to changing operational parameters are also taken into account so that the predicted well
status is as realistic as possible. By having a real estimate of the downhole conditions, it is possible to provide operational
recommendations that can be used to avoid pack-off and stuck pipe incidents. By adjusting the drilling parameters like the
flow-rate, drill-string rotational speed or the ROP, one can identify what adjustments are needed to avoid the formation of
cuttings beds or how to remove them.
The model has been used to analyze the cuttings transport conditions for two situations taking place in a Designer Well
in the North Sea. In the first case, a drill out cement operation was analyzed. The operation took place in an inclined part of the
well. During the operation, several pack off incidents occurred which were probably caused by cuttings bed avalanches that
led to a total obstruction of the annulus. Simulations verified that for the given drilling parameters, cutting beds were evolving
over time. It was seen that the annular velocity was too small in parts of the well, leading to a continuous buildup of a cuttings
bed that most probably developed into a cuttings avalanche, which then resulted in the subsequent pack off situations. The
main operational recommendation was that a higher flow-rate could have been used, since there was no risk of fracturing the
formation as the operation took place within a cased hole.
In the second situation, an MPD operation was analyzed. Here the rate of penetration was relatively high, leading to a
situation where cuttings were accumulating in the well. Simulations were performed to reproduce the well conditions and it
was seen that the flow-rate used was too low. Since the well was operating in MPD conditions, there was an opportunity to
increase the flow-rate and to compensate for the increased downhole pressure by changing the settings of the MPD choke. It
was also confirmed by simulations that an increase in RPM would have been beneficial, to decrease the risk of having cuttings
accumulations by actively stirring the cuttings beds. Simulations have also shown that material transport could happen by
cuttings bed erosion therefore justifying why the reduction of the ROP did stabilize the downhole conditions.

SPE/IADC 163492

17

The simulations performed have confirmed that a transient cuttings transport model is able to recreate the downhole well
conditions, and that it can be a valuable tool for providing real time operational support and recommendations with respect to
avoiding poor cuttings transport conditions. Future work should focus on performing more field studies to verify the
applicability of the model, as well as on improving the calibration of the model. Surface measurements of cuttings size and
surface cuttings flow-rate could be integrated and linked more directly to the model in a real time environment for calibration
purposes. Further model improvements could also involve inclusion of mud gelling effects and incorporation of cuttings
avalanches to make it possible to predict potential pack off situations.
Acknowledgement
The presented work has been performed in the Centre for Drilling and Wells for Improved Recovery (SBBU). This is a
cooperation between the International Research Institute of Stavanger (IRIS), Sintef, the University of Stavanger and the
Norwegian University of Science and Technology (NTNU). SBBU is funded by grants from the Research Council of Norway,
Statoil, Total, ConocoPhillips, Det Norske, Talisman Energy and Wintershall.
The last author wants to thank Statoil for funding his position at the University of Stavanger through the Akademia
program.
Nomenclature
BHA-Bottom Hole Assembly
CFD - Computational Fluid Dynamics
CTFV Critical Transport Fluid Velocity
CTV- Cuttings Transport Velocity
ECD Equivalent circulating density
ID-Internal Diameter
IRIS - International Research Institute of Stavanger
LOT-Leak Off Test
MPD - Managed Pressure Drilling
NTNU - Norwegian University of Science and Technology
NRTL Non-Random Two Liquid
PDC- Polycrystalline Diamond Compact
PVT- Pressure Volume Temperature
ROP Rate of Penetration
RPM- Rotation Per Minute
SBBU - Centre for Drilling and Wells for Improved Recovery
TD-Total Depth
XLOT- eXtended Leak-Off Test
Symbols
cross sectional area
specific heat capacity of the drilling fluid
specific heat capacity with i= w for water/brine, o for oil, lgs for low gravity solid, hgs for high gravity solid, c for
cuttings, m for mud, l for liquid and s for solid
d
diameter of the cuttings particles
diameter of the cuttings particles of the i-size
buoyancy force on a particle
drag force on a particle
gravitational force on a particle
lift force on a particle
plasticity force on a particle
,
,
reaction forces on a particle
volume fraction with i= w for water/brine, o for oil, lgs for low gravity solid, hgs for high gravity solid, c for cuttings,
m for mud, l for liquid and s for solid.
gravitational acceleration of the earth
enthalpy per mass unit
thermal conductivity with i = w for water/brine, o for oil, lgs for low gravity solute, hgs for high gravity solid, c for
cuttings, m for mud, l for liquid and s for solid
L
length of the controlled volume
n
number of particles in the controlled volume
number of particles of the i-size in the controlled volume

18

SPE/IADC 163492

p
p1
Q

pressure
reference pressure when cuttings are coming in suspension
conductive and natural-convective term
forced convective term
heat generated by mechanical and hydraulic frictions
packing efficiency of the coarse particles
packing efficiency of the cuttings particles of the i-size
packing efficiency of the fine particles
packing efficiency of the mix of coarse and fine particles
R
universal gas constant
r
size ratio of the coarse particles to fine particles
s
curvilinear abscissa
T
temperature
T1
reference temperature when cuttings are coming in suspension
volume of phase i, with i= w for water/brine, o for oil, lgs for low gravity solid, hgs for high gravity solid, c for
cuttings, m for mud, l for liquid and s for solid
axial velocity of mud
slip velocity of the i-phase compare to
velocity of the liquid phase
tangential velocity of mud
velocity of mud
volume fraction of fine particles
Greek Letters
NRTL parameters

non-randomness NRTL model


shear rate

separation between two particles


viscosity of the drilling fluid with cuttings
viscosity of the drilling fluid
density with i= w for water/brine, o for oil, lgs for low gravity solid, hgs for high gravity solid, c for cuttings, m for
mud, l for liquid and s for solid
mud density at reference pressure p1 and temperature T1
mass fractions of the liquid phase with
cuttings erosion rate at current flow-rate Q and drill-pipe rotation 0
ratio of the solid thermal conductivity to the liquid thermal conductivity
critical rotational velocity to clear cuttings beds at flow-rate Q

set of indices {w, o, lgs, hgs, c}

set of indices {w, o}

set of indices {lgs, hgs, c}


average inclination
shear stress
References
API Recommended Practice 13D, 2006. Rheology and Hydraulics of Oil-well Drilling Fluids. Fifth Edition, June 2006.
Bassal, A.A., 1995. A Study of the Effect of Drill Pipe Rotation on Cuttings Transport in Inclined Wellbores. MS Thesis, University of
Tulsa, Oklohama, USA.
Belarde, M. and Vestavik, O., 2011. Deployment of Reelwell Drilling Method in Shale Gas Field in Canada. Paper SPE 145599-MS
presented at the SPE Offshore Europe Oil and Gas Conference and Exhibition, Aberdeen, UK, Sep. 06-08, 2011
Blikra H., Drevdal K.E., Aarestad T.V., 1994. Extended Reach, Horizontal and Complex Wells: Challenges, Achievements and CostBenefits. Paper SPE 28005 presented at the University of Tulsa Centennial Petroleum Engineering Symposium, Tulsa, Oklahoma,
USA, Aug. 29-31, 1994.
Cayeux E., Daireaux B., Dvergsnes E., Slevik G., Zidan M., 2012. An Early Warning System for Identifying Drilling Problems: An
Example From a Problematic Drill-Out Cement Operation in the North-Sea. Paper SPE 15942 presented at the SPE Drilling Conference
in San Diego, California, USA, 6-8 March, 2012.
Clark, R. K., Bickham, K.L., 1994. A mechanistic Model for Cuttings Transport. Paper SPE 28306 presented at the SPE Annual Technical
Conference and Exhibition, New Orleans, Louisiana, USA, Sept. 25-28, 1994.
Doron, P., Simkhis, M., Barnea, D., 1997. Flow of Solid-Liquid Mixtures in Inclined Pipes. Paper published in International Journal of
Multiphase Flow, vol. 23, No. 2, pp. 313-323.
Einstein, A., 1906. Eine Neue Bestimmung Der Molekldimensionen. Paper published in Annalen Der Physik, vol. 19, No 5, 1906.

SPE/IADC 163492

19

Ford, J.T., Peden, J.M., Oyeneyin, M.B., Gao, E., Zarrough, R., 1990. Experimental Investigation of Drilled Cuttings Transport in Inclined
Boreholes. Paper SPE 20421 presented at the Annual Technical Conference and Exhibition of the SPE, New Orleans, Louisiana, USA,
Sep. 23-26, 1990.
Gavignet, A.A., Sobey, I.J., 1989. Model Aids Cuttings Transport Prediction. Paper SPE 15417 published in the Journal of Petroleum
Technology, vol. 41, No 9, pp. 916-921, Sep. 1989.
Gmehling, J, Onken, U., 1977. Vapour-Liquid-Equilibrium Data Collection. Vol.1, Part 2a, Verlag & Druckerel Friedrich Bischoff,
Frankfurt, 1977.
Hareland, G., Azar, J.J., Rampersad, P.R., 1993. Comparison of Cuttings Transport in Directional Drilling using Low-Toxicity Invert
Emulsion Mineral-Oil-Based and Water-Based Muds. Paper SPE 25871 presented at the Low Permeability Reservoirs Symposium,
Denver, Colorado, USA, Apr. 26-28, 1993.
Hastchek, E., 1910. Die Viskositt Der Dispersoide. Paper published in Kolloid-Zeitschrift, vol. 8, pp. 34-39, 1910.
Houwen O.H., Geehan T., 1986. Rheology of Oil-Based Muds. Paper SPE 15416 presented at the SPE Annual Technical Conference, New
Orleans, Louisiana, USA, October 5-8, 1986.
Jalukar, L.S., 1993. Study of Hole Size Effect on Critical and Subcritical Drilling Fluid Velocities in Cuttings Transport for Inclined
Wellbores. MS thesis, U. of Tulas, Oklahoma, USA
Kamp, A.M., Rivero, M., 1999. Layer Modeling for Cuttings Transport in Highly Inclined Wellbores. Paper SPE 53942 presented at the SPE
Annul Technical Conference and Exhibition, Caracas, Venezuela, Apr. 21-23, 1999.
Kandula, M., 2011. On the Effective Thermal Conductivity of Porous Packed Beds with Uniform Spherical Particles. Paper published in
Journal of Porous Media, Vol. 14, No 10, pp. 919-926, 2011
Lage, A., Fjelde, K.K., Time, R., 2003. Underbalanced Drilling Dynamics: Two-Phase Flow Modeling and Experiments. Paper SPE 83607
published in the SPE Journal, vol. 8, No 1, pp. 61-70, March 2003.
Larsen, T.I., Pilehvari, A.A., Azar, J.J., 1997. Development of a New Cuttings-Transport Model for High-Angle Wellbores Including
Horizontal Wells. Paper SPE 25872 published in the SPE Drilling & Completion Journal, vol. 12, No. 2, pp. 129-136, June 1997.
Marshall D.W., Bentsen R.G.,1982. A Computer Model to Determine the Temperature Distributions in a Wellbore. Paper published in the
Journal of Canadian Petroleum Technology, vol. 21, No 1, pp.63-75, Jan.-Feb. 1982.
Nagata, I., 1973. Prediction Accuracy of Multicomponent Vapour-Liquid Equilibrium Data from Binary Parameters. Paper published in
Journal of Chemical Engineering Japan, vol. 6, No 1, 1973.
Nazari, T, Hareland, G., Azar, J.J., 2010. Review of Cuttings Transport in Directional Well Drilling: Systematic Approach. Paper SPE
132372 presented at the SPE Western Regional Meeting, Anaheim, California, USA, May 27-29, 2010.
Ozbayoglu, M.E., Saasen, A., Sorgun, M., 2008. The Effect of Pipe Rotation on Hole Cleaning for Water-Based Drilling Fluid in Horizontal
and Deviated Wells. Paper SPE 114965 presented at the IADC/SPE Asia Pacific Drilling Technology Conference and Exhibition,
Jakarta, Indonesia, Aug. 25-27, 2008.
Peden, J.M., Ford, J.T., Oyeneyin, M.B., 1990. Comprehensive Experimental Investigation of Drilled Cuttings Transport in Inclined Wells
Including the Effects of Rotation and Eccentricity. Paper SPE 20925 presented at Europec, The Hague, Netherlands, Oct. 22-24, 1990.
Pilehvari, A.A., Azar, J.J., Shirazi, S.A., 1999: State of the-Art Cuttings Transport in Horizontal Wellbores, SPE Drilling & Completion
14 (3), September 1999.
Ramadan, A., Skalle, P., Johansen, S.T., Svein, J., Saasen, A., 2001. Mechanistic Model for Cuttings Removal from Solid Bed in Inclined
Channels. Paper published in Journal of Petroleum Science and Engineering, vol. 30, No. 3-4, pp. 129-141, Sep. 2001.
Ranjbar, R., 2010. Cuttings Transport in Inclined and Horizontal Wellbore. MS Thesis, University of Stavanger, Norway.
Robertson R.E., Stiff H.A., 1976. An Improved Mathematical Model for Relating Shear Stress to Shear Rate in Drilling Fluids. SPE Journal,
February 1976, vol. 16, No 1, pp. 31-36.
Rowley, R., 1981. A local Composition Model for Multicomponent Liquid Mixture Thermal Conductivities. Paper published in Chemical
Engineering Science, vol. 37, No 6, pp. 897-904, 1982.
Rubiandini, R.S., 1999. Equation for Estimating Mud Minimum Rate for Cuttings Transport in an Inclined-Until-Horizontal Well. Paper
SPE 57541 presented at the SPE Annual Technical Conference and Exhibition, Abu Dhabi, UAE, Nov. 8-10, 1999.
Song, C., Wang, P, Makse, H.A., 2008. A Phase Diagram for Jammed Matter. Paper published in Nature, 453, pp. 629-632, May 2008.
Tomren, P.H., Iyodo, A.W., Azar, J.J., 1986. Experimental Study of Cuttings Transport in Directional Wells. Paper SPE 12123 published in
SPE Drilling Engineering Journal, vol. 1, No 1, pp. 43-56, Feb. 1986.
Yao, D., Robello, S., 2008. Annular Pressure Loss Predictions for Various Stand-off Devices. Paper SPE 112544 presented at the SPE/IADC
Drilling Conference, Orlando, Florida, USA, Mar. 4-6, 2008.
Zheng, J., Carlson, W, Reed, J., 1995. The Packing Density of Binary Powder Mixtures. Paper published in the Journal of the European
Ceramic Society, vol 15, No 5, pp. 479-483.

S-ar putea să vă placă și