Sunteți pe pagina 1din 25

SPE 169552

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade


Ghaithan A. Al-Muntasheri, Aramco Research Centers-Houston & Saudi Aramco

Copyright 2014, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Western North American and Rocky Mountain Joint Regional Meeting held in Denver, Colorado, USA, 16 18 April 2014.
This paper was selected for presentation by an SPE 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 and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, 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 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 copyright.

Abstract
Hydraulic fracturing is a well-established process to enhance productivity of oil and gas wells. Fluids are used in fracture
initiation and the subsequent proppant and/or sand transport. Several chemistries exist for these fluids. This paper summarizes
the published literature over the last decade (90+ technical articles) and captures the advances in the design of water-based
fracturing fluids. Despite their old introduction, guar-based polymers are still being used in fracturing operations for wells at
temperatures less than 300oF (148.9oC). In order to minimize the damage associated with this class of polymers, the industry
attempted several approaches. These include the use of lower polymer concentration in formulating these fluids. Another
approach was to alter the crosslinker chemistry so that one can generate higher viscosity values with lower polymer loadings.
Moreover, the industry shifted towards the use of cleaner guar-based polymers. The reason is the fact that commercial guar
contains a minimum of 5 wt.% residues that cause damage to proppant packs. With fracturing deeper wells in hotter reservoirs,
the need arose for a new class of thermally stable polymers. Thus, the industry shifted towards polyacrylamide-based
polymers. These synthetic polymers offer sufficient viscosity at temperatures up to 232oC (450oF). Examples included 2acrylamido-2-methylpropanesulfonic acid (AMPS) and copolymers of partially hydrolyzed polyacrylamide (PHPA)-AMPSvinyl phosphonate (PAV). To address the challenge of high pressure pumping requirements on the surface, high density brines
have been used to increase the hydrostatic pressure by 30%. On the breakers chemistry, new breakers were introduced. These
breakers decrosslink the gel by reacting with the crosslinker. In order to minimize the environmental impact of using massive
amounts of fresh water and to minimize costs associated with treating produced water, the use of produced water in hydraulic
fracturing treatments has been reported. In addition, the paper captures the advancements in the use of slickwater where use is
made of drag reducing agents (PAM-based polymers) to minimize friction. The paper highlights the first use of breakers that
were introduced to improve the cleanup of these drag reducers. For foamed fluids, new viscoelastic surfactants (VES) that are
compatible with CO2 are discussed. The paper also sheds light on the use of emerging technologies such as nanotechnology in
the design of new efficient hydraulic fracturing fluids. For example, nanolatex silica was used to reduce the concentration of
boron used in conventional crosslinkers. Another advancement in nanotechnology was the use of 20 nm silica particles
suspended in guar gels. The paper provides a thorough review on all of these advancements.

Introduction
Fracturing oil and gas wells is a well-established practice for productivity enhancement in petroleum reservoirs and has been
reported more than 60 years ago (Clark 1949; Padgett 1951; Grossman 1951; Clark et al. 1952). At that time, despite their
safety concerns, fluids that were based on a hydrocarbon phase (kerosene, crude oil or gasoline) were the choice of use in
hydraulic fracturing. At a later stage, fatty acids were used to increase the viscosity of these oil-based fluids to initiate the
fracture and transport sand (Hurst 1953; Ghauri 1960). With more fundamental understanding of the rock-fluid interactions
especially in clay-rich reservoirs (Black and Hower 1965), the use of water-based fluids such as guar-based polymers emerged.
Guar is used to increase the viscosity of the fracturing fluid and help transport the proppant. In order to generate more
viscosity and minimize leakoff, crosslinked guar-based fluids were introduced (Dysart et al. 1969; Harris 1993). In recent
years, synthetic polyacrylamide polymers have also been reported as fracturing fluids for high temperature applications up to
232oC (450oF) (Holtsclaw and Funkhouser 2010 ). However, polymer-based fluids do require breakers. The breakers are used
to improve the retained conductivity of the fractures and minimize the damage associated with leaving residues of these fluids
(Small et al. 1991). As such, in order to minimize the damage associated with the polymer-based fluids, viscoelastic
surfactants (VES) were introduced (Samuel et al. 1999). However, their temperature stability is an issue for reservoirs at more
than 240oF (115oC) unless they are used at high concentrations. Moreover, these materials do not provide adequate leakoff

Al-Muntasheri

[SPE 169552]

control as will be discussed later in the manuscript.


In shale reservoirs (permeability in the nanodarcy range), the need for a high viscosity fluid to minimize leakoff is
eliminated. Thus, the industry made use of slickwater where the proppant transport is governed by the high velocity of the
injected water unlike the polymer-based fluids where the viscosity is the transporting mechanism. However, slickwater carries
a maximum of 1 lb/gal of proppant. The literature reported significant use of hybrid technologies where slickwater and
polymer fluids are combined (Handren and Palisch 2009; Coronado 2007; Robart et al. 2013). Energized fluids (fluids with
one compressible component such as nitrogen or carbon dioxide) were also used (Tan et al. 1995; Cawiezel and Gupta 2010).
The use of the gas component helps reduce the hydrostatic pressure and may eliminate the use of a gas lift after the fracturing
treatment. It also provides an easy-to-clean option from the formation after the hydraulic fracture is created. Polymer-based
fluids are still the most commonly used type of fracturing fluids. This is due to their versatile properties and the extensive
industry experience associated with their use. To the best of the authors knowledge, there is no literature report that captures
the advancements of the polymer-based fracturing fluids over the last decade. Therefore, the objectives of this paper are to: 1)
provide an overview of the available fracturing fluids, 2) discuss their different chemistries, 3) review their breakers packages
and 4) capture the new advancements reported in the literature for these fluids over the past decade.
Discussion
In hydraulic fracturing of low to moderate permeability reservoirs (non frac pack applications), a high viscosity linear polymer
(35 to 50 cp at shear rates of 40 to 511 sec-1) is injected under turbulent flow conditions where high friction and shear rate are
encountered. Once the gel leaves the perforations (higher shear rate), the gel contacts the formation face and starts building up
pressure. The pressure build up continues until the fracturing point is reached. A fracture is then generated extending away
from the wellbore into the formation. It propagates to provide maximum contact with the reservoir. A crosslinker is then mixed
with the linear gel resulting in a substantial increase in the gel viscosity (can be more than 1,000 cp). The high viscosity of the
crosslinked fracturing fluid is used for transporting proppants. At the end, once the proppant is successfully placed, the
pressure is released allowing the fracture to close. Then, the well flowed back for clean up while the proppant keeps an open
channel for the hydrocarbon production.
An ideal polymer-based fracturing fluid must possess certain key physical properties during the previously mentioned
process. First, it must have a high shear resistance where the gel does not lose viscosity upon high shearing in the pumping
equipment as well as in the piping system (well tubing and/or coiled tubing). At the same time, the gel should provide minimal
frictional losses during injection through the well tubing. Second, the gel must provide sufficient viscosity for proppant
transport. If the viscosity is too low, proppant screenout will result. Third, the gel should have minimal leakoff into the
adjacent matrix. This is to ensure that the fracture propagates deeper into the formation and that minimal damage takes place at
the adjacent matrix. Fourth, the gel should be compatible with the formation materials such as clays that are known to swell in
the presence of water and cause productivity loss. Fifth, the gel should be able to lose viscosity (break) and clean up efficiently
from the fracture face and the proppant matrix.
In the case of slickwater, it is used in very low permeability reservoirs (nanodarcy range). As such, leakoff is not an issue.
The proppant or sand is transported by the high velocity. In this case, friction becomes high and hence a drag reducer is added.
Most drag reducers are based on polymers and therefore, breakers have been added to improve the clean up of these fluids
from shale reservoirs as will be discussed later in the manuscript. The following sections will shed light on the available
polymer-based and slickwater fracturing fluids in light of the above requirements. A summary of all fluid systems reviewed in
this paper is given in Table 1.
Polymers
The use of bio polymers (like guar and cellulose) in fracturing fluids dates back to 1953 when they were used as fluid
thickeners in acid fracturing treatments (Hurst et al. 1953). Guar-based polymers are the most commonly used type of
polymers for hydraulic fracturing. Guar is a high molecular weight biopolymer consisting of a chain of polysaccharide
(mannose building block) with side chains of galactose (Jennings 1996). It is a naturally occurring material that is produced
from beans. Figure 1a shows the chemical structure of guar. In the presence of water, it hydrates generating a shear thinning
viscous fluid. In the field, it is prepared by dissolving solid guar in water at concentrations below 40 lb/1000 gallons (pptg).
For high concentrations (more than 40 pptg), hydration of dry guar can be difficult. As such, high concentration slurries (4
lb/gal) can be used as a source of guar instead of solid guar to minimize the formation of fish eyes (Sarwar et al. 2011). The
viscosity of this fluid (linear gel) is a strong function of temperature, shear rate and polymer content. For example, the
viscosity of linear guar gel can drop by 90% (from 103 to 10 cp) upon the reduction of guar concentration from 80 pptg to 20
pptg maintaining the same shear rate and temperature (at 511 s -1 and 77oF) (Ely 1989). Table 2 gives viscosity data of linear
gels reviewed in this paper. Despite the different sources of this data for guar viscosity, they still show good agreement. In the
subsections below, the two major polymer types used in hydraulic fracturing, namely, bio polymers and synthetic polymers
will be discussed.
A. Bio Polymers
Guar manufacturing involves the presence of insoluble residues (Sun and Qu 2011; Sarwar et al. 2011). These residues can
precipitate in the proppant pack causing loss of productivity. This is in addition to solid residues as a result of inefficient

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

breaking of the crosslinked gel as will be discussed in the coming sections. Moreover, thermal stability of the guar becomes
questionable at temperatures higher than 180oF. To address these issues, the industry improved the thermal stability of guar by
introducing other forms of guar-based polymers such as: hydroxypropyl guar (HPG) and carboxymethylhydroxypropyl guar
(CMHPG) shown in Fig. 1b and Fig. 1c, respectively (Ely 1989). Cellulose-based fluids are also used in hydraulic fracturing.
Examples are: carboxymethylhydroxyethyl cellulose (CMHEC), carboxymethyl cellulose (CMC) and hydroxyethyl cellulose
(HEC), with the chemical structures shown in Figs. 1d through 1f, respectively (Ely 1989).
In the early days, guar was used at concentrations as high as 80 lb/1000 gal (pptg). To minimize the damage associated
with the guar-based gels, the industry reduced the concentration of guar to levels ranging from 15 to 25 pptg at lower
temperatures of 140oF (Nimerick et al. 1997; Harris and Heath 1998). When allowed more time for hydration, these fluids
were found to give higher viscosity after crosslinking with borate. Jianshe et al. (2012) reported on another low polymer
content fracturing gel. The gel is based on HPG and an additional additive referred to as synthetic hydroxyl guar. The authors
did not provide details on the chemical nature of this material. Yet, the system gave, on average, 2.5 times more oil production
when applied in fracturing 47 oil producing wells in China in wells with bottom hole temperatures of 70oC (158oF).
Loveless et al. (2011) reported on a new system that is based on ingredients used in the food industry. It is the belief that
the system has a mixture of cellulose-based polymers and starch and is crosslinked with Al+3. Loveless et al. (2011) reported
viscosity data of this system at 140oF (60oC). Slot flow was used to investigate the ability of the system to transport sand
without settlement in the vertical direction. The slot was 12 ft. long and the sand concentration ranged from 1 to 5 ppg. The
performance of the system was compared to linear xanthan gel, a guar/borate gel and a VES-based system. The new foodbased system showed better results in terms of even distribution of the sand in the vertical direction. VES and the linear
xanthan were the least performing fluids. Static proppant carrying tests for 300 minutes showed better results for the new fluid
and xanthan gel. Holtsclaw et al. (2011) reported on the fluid loss and the retained conductivity measurements of the system.
The retained conductivity tests showed more than 80% retained conductivity. It should be mentioned that these tests utilized a
polymer concentration of 40 and 60 pptg at 120 to 200 oF (48.9 to 93oC).
Qun et al. (2008) reported the use of what they referred to as finely processed guar for high temperature applications of
150oC (302oF). The polymer loading was 0.45% (around 37.6 pptg). The finely processed guar is believed to have a low
insoluble content of 1.6% compared to 16 % in some types of guar. This linear gel maintained more than 150 cp for more than
2 hours at 150oC and a shear rate of 170 s-1. The system was implemented in gas producing wells where it gave 16 to 28 times
more gas after the treatments. It should be mentioned that, in addition to the use of the finely processed guar, the authors
attributed these successful treatments to a forced fracture closure technique where the pumping schedule of the proppant was
altered. In this technique, the well was shut-in allowing the proppant to flow from the low permeability to the high
permeability zones. In a similar approach, Williams et al. (2012) reported on the use of a high yield fast hydrating guar
polymer for use in fracturing applications. The polymer is thought to have less insoluble residues compared to existing
products. The authors observed higher viscosity with the linear gels of this polymer compared to those of standard guar
polymers. For example, after 90 minutes, the new polymer gave 580 cp compared to 420 cp (at 200oF and 100 s-1). The authors
realized a 33% reduction in polymer loading upon the use of this polymer. The data is summarized in Table 3. It should be
noted that this improvement in viscosity is partially due to the use of a size-designed, borate-based crosslinker.
Weaver et al. (2002a and b) reported a new approach to improve polymer clean up. The gel is mainly a low molecular
weight guar-based polymer crosslinked with borate. The HPG is depolymerized to give low molecular weight segments
(preferably between 100,000 and 250,000 Daltons). Then, the gel can form at pH values above 9. Upon contact with the
formation, the pH drops below 9. Thus, the borate crosslinker does not become functional. This fact combined with the low
molecular weight nature of the segments allows the gel to clean up without the need for a breaker. The system was applied in
gas and oil shale wells. The authors claimed successful placement of the proppant. Peles et al. (2002) reported another
application of the same system in the Springer Group reservoir of the Cement field in Oklahoma. The treated wells had
permeabilities ranging from 0.1 to 3.9 mD. Production data was not reported. However, data indicated that about 65% of the
fracturing fluid was recovered. Palmore et al. (2013) reported the application of this system in wells of the Means San-Andres
field in Texas. It is believed that the system provided better results compared to guar.
Walters et al. (2009) reported the use of a linear gel (based on a biopolymer) to carry proppants. The system requires
neither crosslinker nor a breaker. At 40 s-1 shear rate, it gave viscosity values of 352 cp and 216 cp at 96oF and 250oF,
respectively. The polymer concentration is believed to be between 40 and 70 pptg. The authors did not provide information
about the nature of the fluid. However, it is claimed that the system retains 90% conductivity compared to 14% of boratecrosslinked gels under the same conditions (200 oF and 6,000 psi closure stress).
Another advancement was the use of a mixture of non acetylated xanthan (NAX) gum and guar (Fischer et al. 2001).
NAX is a xanthan gum without the acetyl acid group. The chemical structure of NAX is shown in Fig. 2. The objective of
using NAX was to obtain higher viscosity compared to a solution of guar alone. Table 2 gives viscosity data at ambient
temperature of this system. Clearly, the introduction of NAX (at a ratio of 4:1) gave better viscosity results as opposed to using
guar or NAX alone. One drawback of this technology is its sensitivity to KCl. As such, alternatives to KCl such as tetra methyl
ammonium chloride (TMAC) should be used. There is neither conductivity data nor field application data of this system.
Tonmukayakul et al. (2011) reported the use of mixtures of cellulose-based and xanthan or guar polymers. To the best of the
authors knowledge, no data is reported for the performance of this system in terms of conductivity.

Al-Muntasheri

[SPE 169552]

In summary, the major advancements in the use of biopolymers in fracturing were the use of cleaner fluids where the
insoluble part of the guar was eliminated. At the same time, mixtures of different polymers were attempted. Moreover, use has
been made of lower molecular weight segments that can be used in low permeability reservoirs where leakoff is not an issue.
B. Synthetic Polymers
In recent years, the cost of guar beans increased dramatically and became very volatile. Moreover, guar is not stable at
temperatures close to 400oF. Gupta and Carman (2011) reported 360oF (182oC) as the upper limit of guar use while Harris and
Sabhapondit (2009) reported 375oF. With the cost increase trends and the need to fracture wells with temperatures more than
375oF (190oC), the oil and gas industry started developing alternatives. Synthetic polymers like polyacrylamide (PAM) have
been widely used as friction reducers in fracturing operations. They are usually added at concentrations of 1.25 to 2.5 pptg
(Gaillard et al. 2013). In 2010, Holtsclaw and Funkhouser reported the use of a PAM-based polymer as a fracturing fluid
where it can carry proppants. The system is based on a terpolymer of 2-acrylamdio-2-methylpropanesulfonic acid (AMPS) and
acrylamide (AM) with the structure shown in Fig. 3. Note that Funkhouser and Norman (2003) reported that the best
performance of this system is obtained upon the use of 60% AMPS, 39.5% amide and 0.5% acrylate. Moreover, for these high
temperatures, the authors found that sodium bromate is the best breaker. Holtsclaw and Funkhouser (2010) reported viscosity
data of this copolymer crosslinked with Zr+4 (66 pptg polymer and 3 gpt Zr+4) where the system gave a viscosity of 700 cp for
more than 1.5 hours at a temperature of 402 oF (204oC) under 40 1/s of shear rate. Under the same conditions, a gel based on
CMHPG containing the same concentration of polymer and crosslinker lost significant viscosity where after 22 minutes of
shear, its viscosity dropped to below 50 cp. The PAM-based polymer crosslinked with zirconium was applied in two wells in
south Texas at 450oF (232oC) and 23,300 ft. (Funkhouser et al. 2010). The data is summarized in Table 4. The authors
reasoned that the loss of production in Well 1 was due to the very low permeability of the reservoir (less than 10 micro darcy).
Gupta and Carman (2011) reported on the use of another high temperature fracturing gel which is based on a copolymer of
partially hydrolyzed polyacrylamide (PHPA)-AMPS-vinyl phosphonate (PAV). The authors did not report the content of each
monomer in the copolymer backbone. This copolymer is crosslinked by zirconium. The authors reported high temperature
viscosity data that is summarized in Table 5. The system maintained a 1,000 cp viscosity after shearing at 100 1/s for 2 hours
at 425oF. The system at 3,000 psi of closure maintained more than 65% of conductivity at the same temperature. The data is
also summarized in Table 6. It is worth mentioning that this system can also be energized by the use of carbon dioxide. Note
that energizing guar was limited to 260oF due to compatibility issues between guar and CO2. The authors did not report field
application data for this system. Acetic acid and sodium acetate buffers are used to maintain a low pH of 5. This is believed to
stabilize the gel.
Gaillard et al. (2013) reported the use of three PAM-based polymers for use in hydraulic fracturing. These polymers are
high molecular weight (5 to 15 million g/gmol) PAM-based friction reducers. They all contain at least 1.5 mol% of a
hydrophobe and have an acrylate content of 10 to 25 mol%. The third polymer had an AMPS content of 10 to 25 mol%. The
polymers exhibited good proppant settling properties. Yet, their viscosity at high shear rates and high temperatures are yet to
be investigated. Table 7 shows viscosity data of the three polymers and guar as a function of shear rate at 194oF (90oC).
High Density Brines
A formation can only be fractured if the bottom hole pressure exceeds the formation breakdown pressure. The below equation
can be used to calculate the surface treating pressure required for a fracturing operation (Pandey and Robert, 2002; Simms and
Clarkson 2008):
.......(1)
where
is the bottom hole treating pressure in psi,
is the friction pressure in psi and
is the hydrostatic pressure in
psi. Some unconventional reservoirs are located at depths up to 20,000 ft. and the reservoir temperature can be as high as
355oF (179.4oC) (Bartko et al. 2009). Fracturing such wells requires bottom hole treating pressures up to 20,000 psi. There are
limitations on the pressure rating of some bottom hole completion equipment. There could also be limitations on the pressure
rating of surface equipment and pumping equipment where they do not exceed 15,000 psi (Qiu et al. 2009). From Equation 1
above, increasing hydrostatic pressure of the fluid will result in a decrease in the surface treating pressure. To increase
in the above equation and address the 15,000 psi pumping pressure limitation, a high density brine (11.4 lb/gal) is used to
prepare the fracturing gels. The density of standard fracturing gels prepared in field brines is usually around 8.7 lb/gal (Simms
and Clarkson 2008). Thus, the use of high density brines can increase the hydrostatic pressure by 30%.
Bartko et al. (2009) reported the first field application of a high density brine in tight gas reservoirs in Saudi Arabia. This
system was based on a CMHPG polymer crosslinked with a delayed Zr +4-based crosslinker. The authors utilized NaBr having
a density of 12.3 lb/gal. Therefore, at a given depth, using this high density brine increases
by 41%. As a result, the
authors were able to triple the pumping rate (10 to 30 bbl/min) with a wellhead pressure increase of only 2,000 psi (from 9,500
to 11,500 psi). Nonetheless, the treatment was unsuccessful because of a mechanical failure where the liner has collapsed.
Bu-Khamseen et al. (2010) reported the first successful field application of this high density CMHPG/Zr +4 gel system in
Saudi Arabia. A well with a bottom hole temperature of 375 oF (190oC) was hydraulically fractured at an average depth of
18,300 ft. The treatment utilized a high strength 18/40 proppant. The lower part of formation (Sarah) was successfully treated.
The gas rate increased from 0.35 to 3.5 MMSCFD. Yet, a failure in one of the upper completion pup joints caused a change in

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

the direction of the treatment. This lower zone had to be isolated by the use of a composite plug. Then, the upper zone of the
same formation was fractured. After the second treatment, the plug was milled out. The total production rate was 2.8
MMSCFD.
Gupta et al. (2012) reported on the use of a different chemistry to address the same challenge. The authors utilized a
polyampholyte and a surfactant. The mechanism by which the viscosity builds up does not include a crosslinking reaction. The
nature of this viscosity buildup is believed to be an interaction between the polyampholyte with the salts in the high density
brine. The authors reported viscosity data and fracture conductivity testing of this system at temperatures up to 250oF (121oC).
The system was found to be compatible with various brines including CaBr2 and its mixtures with CaCl2 as well as NaBr and
its mixtures with NaCl. Conductivity data is shown in Table 8.

Crosslinkers
Polymer-based fracturing gels are crosslinked using one of two major types of crosslinkers. Namely, borate esters and metallic
crosslinkers such as: zirconium (Zr+4) and titanium (Ti+4). In this section, light will be shed on the advances in crosslinkers for
polymer-based hydraulic fracturing fluids. Note that the advancements were only limited to borate-crosslinked gels. To the
best of the authors knowledge, the advancements in metallic crosslinkers were limited.
Sun and Qu (2011) reported on the development of a new class of boron-based crosslinkers. The attempt was aimed at
producing crosslinkers that can generate higher viscosity using less polymer content. The size of the crosslinker was the new
parameter to change. The authors utilized different sources of boron other than boric acid. These were: thiophenediboronic
acid (TDBA), benzenediboronic acid (BDBA) and biphenyldiboronic acid (BPDPA). The chemical structures of these fluids
are shown in Fig. 4. It has been found that increasing the size of the crosslinker gives higher viscosity. The data is summarized
in Table 9 where higher viscosity values were obtained with these larger-size crosslinkers compared to boric acid.
Unfortunately, the cost associated with making these crosslinkers makes them impractical. Thus, the authors reported on the
use of trimethylborates as a source of boron to react with amines. The amines that can be used include polyethyleneimine and
3-aminomethylbenzylamine which are shown in Fig. 5. The reaction yields cost-effective, large sized aminoborate compounds
to crosslink guar-based polymers. The best results were obtained by using tris(2-dimethoxyboronoaminoethylamine) which
gave a viscosity of 900 cp compared to 700 cp with a conventional borate crosslinker at 200 oF (93oC) and using 25 pptg of
polymer. Legemah et al. (2013) reported on the development of a crosslinking-delayed version of the technology. This is
covered in the crosslinking delay section.
Loveless et al. (2012) reported on a similar approach. The authors utilized a multifunctional boronic crosslinker. The said
crosslinker can crosslink at least 2 or more polymer molecules. Moreover, its chemical structure has spacers enabling its chain
to be large. In an experiment to measure the elastic modulus (G) of gel samples prepared using this crosslinker, the authors
realized a 10 times increase in G (sample 1) compared to a second sample with the conventional borate crosslinker. This is
despite the fact that only 2.5 micromoles of the multifunctional boronic crosslinker were used compared to 400 micromoles of
the conventional borate crosslinker in sample 2. Neither the conductivity data of this gel nor its breaking behavior was
reported.
Williams et al. (2012) reported on the use of a new crosslinking agent referred to as an efficient borate crosslinker. The
authors did not report on the chemical nature of the crosslinker. In addition, it was combined with another improving factor
which is the use of a high yield non derivatized guar. Yet, some viscosity data was reported where incremental viscosity
improvements were realized. The data is summarized in Table 3.
Delay Mechanisms
It is important to delay the crosslinking reactions of fracturing gels. This is done to reduce frictional losses encountered during
pumping of these fluids under turbulent flow conditions within surface pumping equipment and well tubing. Having less
frictional losses reduces the pumping pressure requirements on the surface. For metal-based crosslinked gels, another reason to
delay their crosslinking reaction is the fact that these gels are shear sensitive and do not recover their viscosity upon shear
reduction in the formation. Thus, the chances of proppant screen out increases. Delay agents are added to minimize viscosity
build up until the fluid leaves the high shear rate regions of the wellbore and the perforation. Hence, the problem of the limited
recovery of post-shear viscosity for zirconium-crosslinked gels is overcome. Ainley et al. (1993) and Legemah et al. (2013)
give a comprehensive list of patent literature resources on the delay methodologies used over the years: 1987 through 2000.
Over the last decade, several advancements were captured. For the new class of size-controlled borate crosslinkers that are
based on polyaminoborates, Legemah et al. (2013) reported the use of four ligands as delay agents. Namely, sorbitol,
pentaerythritol, dipentaerythritol and N-methylglucamine. These four ligands are used in the synthesis of the polyaminoborate
(PAB) crosslinker to produce the delayed version of the technology (DPAB). In order to quantify the delay of these ligands on
the crosslinking reaction of the guar with DPAB, the authors made use of the vortex time. This is done experimentally by
agitating the linear gel at 1,500 rpm using a three blade blender. Then, a vortex is created. Upon the addition of the crosslinker,
the viscosity of the mixture increases and the vortex closes. Vortex time data ranged from 1.5 to 4 minutes. The data is
summarized in Table 10. It is clear that adding more glyoxal enhances the crosslinking delay of this system. However, field
requirements for delay were not discussed.

Al-Muntasheri

[SPE 169552]

Putzig and Clair (2007) reported a new delay agent for metallic (zirconium and titanium) crosslinked gels. The delay
system is a chelating agent containing alcohol, carboxylic acid and amine functional groups. The system was tested at
temperatures up to 275oF (135oC). This delay agent was tested with various types of crosslinkers including non-aqueous
zirconium complexes like triethanolaminozirconate, zirconium lactate and triethanolaminotitanate. Some of the results from
this study are reported in Fig. 6. The data shows that adding more delay agent retards the crosslinking reaction further. Yet, the
effect of these materials on final gel viscosity is negative where the viscosity drops when more delay agent is added. It is
believed that the reason for this drop is the interaction between the chelating agent and the multivalent cation. This reduces the
concentration of the metal cations available for crosslinking the polymer. However, the authors suggest the use of more
crosslinker concentration to offset this observation.
Kalgaonkar and Patil (2012) developed a new shear stabilizing chemical for zirconium-crosslinked gels at high
temperatures of 300 to 375oF (149 to 191oC). The system is believed to be based on a combination of two buffers, namely,
potassium carbonate and sodium carbonate added at sufficient concentrations to raise the pH to 10.2. When sheared at 127 s-1
and 75oF (24oC), it took the stabilized-gel 75 minutes to crosslink where its viscosity remained below 100 cp. After that, the
temperature was increased to 350oF (176oC) and the gel maintained a viscosity of more than 400 cp for more than one hour.
On the other hand, at 275oF (135oC), a gel without the buffering system crosslinked in less than 5 minutes where its viscosity
reached more than 500 cp.
Breaker Packages
Upon the completion of the fracture initiation in the reservoir and the subsequent proppant placement, the used high viscosity
fluid must be removed from the proppant matrix. In addition, the filter cake it formed on the fracture face needs to be cleaned.
Thus, breakers are used to initiate a chemical reaction by which the viscosity is reduced. This chemical reaction is a function
of temperature, concentration of the breaker, polymer concentration and pH. Note that studies pointed out that due to fluid leak
off, the polymer concentration in the formed filter cake can be several times larger than the polymer originally used (Reddy et
al. 2013). Several breaker packages exist in the market with their own operating windows as shown in Table 11 (Sarwar et al.,
2011; Brannon et al., 1994; Crews 2013). Oxidizers (persulfates) degrade the guar-based polymer by a free radical reaction
with sulfur. This reaction defragments the polymer into shorter molecules, reducing its molecular weight and thus, its
viscosity. On the other hand, enzymes degrade the polymer by hydrolysis of certain side chains. Acids decrosslink the gel by
lowering the pH and are mainly used for borate-crosslinked gels that require high pH for crosslinking.
Sarwar et al. (2011) investigated the effectiveness of oxidizers such as: ammonium persulfate, sodium persulfate, sodium
and calcium peroxides and galactomannanase enzyme over a wide temperature range from 75 to 300oF (24 to 149oC). The
residue-after- break was used as one measure of the effectiveness of the breaker. All of the breakers generated residues from 5
to 7 wt.%. The enzyme was found to provide a more homogenous break leaving less residue. The authors did not indicate
whether the residue is a result of impurities within the guar or a residue of the unbroken gel. However, it is believed that
commercial guar contains up to 10 wt.% of impurities (Nasr-El-Din 2013).
Persulfates are too reactive at temperatures more than 140oF (60oC). Thus, at higher temperature applications, the chances
of a pre-mature polymer degradation and the subsequent proppant screenout are very high. In order to provide a better control
over the breaker activity and extend the working temperature range of persulfates to higher than 140oF (60oC), encapsulation
was introduced (Gulbis et al. 1992; Powell et al. 1999). The encapsulation takes more than 9 forms (Lo et al. 2002). For
example, it can include a coating that dissolves slowly releasing the breaker (Lo et al. 2002). Another method is the use of a
barrier (shell) that allows diffusion of the aqueous medium to the core containing the breaker. Then, the breaker is dissolved
into the fluid stream and diffuses back to the bulk fluid (Patil et al. 2013). A third type is the use of particles that rupture upon
the fracture closure where the stress increases significantly, thus, releasing the breaker afterwards (Nolte 1985; Gulbis et al.
1992). Table 12 shows breaker release data as a function of time for various systems.
A recent paper by Patil et al. (2013) reports the use of encapsulated persulfate oxidizers. The encapsulation was diffusion
controlled where the breaker was distributed throughout a polymer matrix. The nature of the polymer was not disclosed. Upon
the encapsulation, it took longer to decrease the viscosity of the gel to 50 cp. The data is summarized in Table 13.
Due to their safety concerns and their costs when encapsulated, alternatives to oxidizers were sought. Hanes et al. (2006)
introduced one of the new approaches where the breaker decrosslinks the gel by competing with the guar polymer to crosslink
the metallic crosslinker. The authors utilized polyaspartate and polysuccinimide. Reddy (2013) also suggested that these
materials can even displace the guar from the already existing guar/zirconium crosslinks. Reddy (2013) reported lab data of the
polysuccinimide testing as a breaker for CMHPG/Zr+4 gels. Using 5 micron discs, it was found that polysuccinimide did not
produce a filter cake while the oxidizing breaker did form filter cake with a high polymer concentration ranging from 183 to
300 pptg depending on shear. Note that these experiments utilized an initial polymer concentration of 38 pptg and they were
conducted at 200oF (93oC) by flowing the gel through the disc under pressures ranging from 200 to 800 psi. The author
suggested the use of ortho-substituted aromatic lignosulfonates. When incorporated in biopolymers, these materials are
thought to form a solid precipitate within the filter cake matrix. After the complete proppant placement, high pH conditions
can then trigger the dissolution of the solid breaker which then degrades the filter cake from within. The efficiency of these
new breakers using conductivity measurements at high temperatures is yet to be reported.
Gunawan et al. (2012) reported on the development of a new class of bio breakers that are non enzyme-based and that can
function at temperatures from 175 to 225 oF (79 to 107oC). The chemistry of these breakers was not published. Yet, they are

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

believed to degrade the polymer under alkaline conditions. These breakers were efficient for use with derivatized guar like
CMHPG which is difficult to break by enzymes. The breakers reported in this study showed a viscosity reduction to less than
50 cp for various polymers including: guar, high yield guar crosslinked with borate and CMHPG crosslinked with zirconium.
Re-use of Produced Water in Hydraulic Fracturing
A typical fracturing treatment requires the use of 4 to 6 million gallons of water. The transportation and processing of this
water is costly. Moreover, there are strict environmental regulations on the injection of produced water which can cost from
0.75 to USD 1.0 per gallon (LeBas et al. 2013). For these reasons, the industry started making use of water produced from
hydraulic fracturing treatments. In some parts of the world where water is scarce, this approach is even more needed.
Produced water contains high concentrations of salts (up to 270,000 of total dissolved solids (TDS)), may contain
hydrocarbons, suspended solids and bacteria. The bacteria is responsible for generating enzymes that degrade the guar-based
polymers significantly. Thus, when fracturing gels are prepared in produced waters, the resulting gels exhibit poor viscosity as
shown in Table 14. Biocides such as tetrakishydroxymethyl phosphonium sulfate (THPS) and quaternary ammonium
compounds are used to suppress bacterial growth (Li et al. 2010; Paulus 2005). These materials were successful in suppressing
bacterial growth in freshwater but not in produced water which can contain significant amounts of bacteria. This is because
bacteria can grow by more than 157 times in only one week (Li et al. 2009).
Huang et al. (2005) reported on the use of produced water in New Mexico. The study utilized a CMHPG/Zr +4 system
prepared in a produced water of 23,000 ppm TDS. In order to suppress the bacterial growth, the produced water was treated
with glutaraldehyde. The pH was found to play a key role in the viscosity of the final gel. The best results were achieved when
the pH was around 5.7. At a pH of 7.2, the weakest gel was obtained. A field application in the San Juan Basin of New Mexico
in a gas producing well resulted in 100% improvement in the gas production rate where the post treatment increased to 100
MSCFD from an initial rate of 50 MSCFD.
Li et al. (2009) developed a new fluid stabilizer that is based on organo-zirconium compounds such as: sodium zirconium
lactate, triethanolamine zirconium, zirconium tetrachloride and zirconium acetate. The technology was used in fracturing 31
wells in California. The produced water was first treated with 70 to 270 mg/L of the stabilizer for 20 minutes. Then, the water
was used to prepare the fracturing fluid. The reservoir had a temperature of 130oF (54oC). No gas production data was
reported. Yet, the use of produced water reduced the time it took to handle it by a factor of 6. Moreover, according to the
author, around USD 1.5 MM cost avoidance was realized. Another application at higher temperatures from 197 to 225 oF (92 to
107oC) in New Mexico was highlighted by Li et al. (2010). The system was applied in three vertical wells in New Mexico with
a total water requirement of 27,000 bbl. Another application was in 6 wells in Belmont Offshore Field in California with a
temperature of 190 to 210oF (88 to 99oC) and the total water requirement was 21,200 bbl. The water analysis for both
applications is reported in Table 15.
LeBas et al. (2013) reported the use of an electrocoagulation method for treating produced water. The process is based on
an electric current applied across a solution of the produced water where hydrocarbons, solids and other contaminants
coagulate leaving a clear water solution. Part of the chemical analysis of the produced water before and after this treatment is
shown in Table 16. Note that the major reduction is in the amount of total suspended solids and turbidity. The bacterial
activity was not reported as a concerning factor in this study. The treated water was used to prepare a CMHPG/zirconium gel
that was applied in 7 wells with 97 fracturing stages. The authors reported significant cost avoidance. The post fracturing
treatment data was not reported.
In a study by Gupta and Hlidek (2010), flowback water from fracturing treatments was recycled for use in other fracturing
treatments. The fracturing fluid was a foamed fluid that is based on N2 or CO2. CO2 is used for deeper wells while N2 was used
for shallower wells. Note that the use of N2 is more cost-effective than CO2. The surfactant used in these treatments was
composed of anionic and cationic components. When the charges become equal on both surfactants, lamellar structures form
which cause the viscosity to increase. The flowback water flowed into large tanks and was then left for 24 hours where the
solids settle to the bottom of the tank and the residual oil floats to the top of the tank. Then, water is extracted 3 ft. (1 m) from
the bottom of the tank. It is then filtered and used to prepare the fracturing fluid. This technique was tried in 5,000 treatments
in Canada. The reservoir temperature ranges from 158 to 212oF (70 to 100oC).
Although the use of produced water represents a viable environmentally-friendly approach, there are concerns with the
use of chemicals and their handling. It has been reported that up to 56% of the chemicals used in fracturing are hazardous to
different extents (Jacoby 2011). In addition, recent regulations in different states and debates about the effect of fracturing
fluids on groundwater raise public concerns (Arthur et al. 2011). Hepburn (2012) presents an example where a scoring system
was developed to assess the use of stimulation fluids in Eagle Ford formation in South Texas. The system examines the
Health, Safety and Environmental (HSE) issues of these chemicals. For example, in the health hazards areas, a score is given
to each chemical in 10 health hazard areas. Examples of these hazards areas include: toxicity, skin erosion and carcinogenicity.
The same is done with physical health hazards in 16 areas and 3 areas in environmental hazards.
In summary, the use of produced water represents a major change to the oil and gas hydraulic fracturing operations as it
will require use of better handling and treating technologies of produced water. Moreover, the use of produced water gives an
environmentally friendly approach to conserve water. However, the environmental regulations for use of these chemicals are
still not global in all fields. A concern still exists about the effect of these chemicals on the environment.

Al-Muntasheri

[SPE 169552]

Surfactants
An additional type of fracturing fluid is viscoelastic surfactants (VES). These materials are believed to be less damaging to the
proppant packs as they leave less residues. In addition, they do not require crosslinkers which means less complexity in the
field. In water sensitive formations, the use of VES is preferred as it reduces the surface tension of the fracturing fluid and thus
helps flowing water out of the pore space of the rock. In fact, studies showed conclusive evidence that VES-based fluids give
less surface tension compared to borate-crosslinked gels (Nasr-El-Din et al. 2007). However, in high permeability reservoirs
(more than 200 mD), VES-based fluids are known to have high leakoff rate as they do not build a filter cake on the fracture
face (Crews et al. 2008). Other researchers refer to the permeability threshold for using VES-based fluids as 100 mD (Sullivan
et al. 2006). The reason for not building a filter cake is the lack of the crosslinking as will be discussed later. Moreover, their
thermal stability is questionable at temperatures more than 275 oF (135oC) unless a significant VES concentration is used (60
ptg) (Fontana et al. 2007). In this section, an overview of the available surfactant technologies and their limitations will be
presented.
Samuel et al. (1999 and 2000) reported on the use of a VES for hydraulic fracturing. The system is based on a quaternary
ammonium salt. that builds viscosity upon the interaction with inorganic salts. The VES is a low molecular weight molecule
having a hydrophilic head and a long hydrophobic tail. In the presence of salts such as potassium chloride, ammonium chloride
or ammonium nitrate, they form elongated micellar structures. With sufficient concentration of these micelles that is larger
than their critical overlap concentration, the micellar structures entangle to build viscosity. Because this entanglement is not
based on a chemical crosslinking reaction (unlike the case with crosslinked polymer gels), these fluids have high leakoff rates.
The high leakoff rate prevents their use in reservoirs with permeabilities larger than 100 mD. The authors reported case
histories of the mentioned system. It was applied in south Texas in wells with a bottom hole temperature of 180 oF (82oC) and a
reservoir permeability of 0.1 mD. The treatment utilized a maximum proppant content of 4 ppa. The well delivered almost 10
times more gas after the treatment. Another treatment was conducted in a lower permeability reservoir (0.03 to 0.05 mD) in
Wyoming. The reservoir temperature was 180 oF (82oC). A maximum of 5 ppa of proppant was placed. The treatment was
successful where the gas rate was 2.8 MMSCFD. The authors claim that a similar treatment in the same field with a borate
crosslinked gel gave 1.3 MMSCFD only.
VES-based fracturing fluids have a limitation when used in reservoirs with permeabilities larger than 200 mD. This comes
from the fact that their leakoff rate is high. The reason is the lack of structures unlike polymer-based gels where crosslinking
provides a three dimensional structure that forms a filter cake minimizing leakoff into the adjacent matrix. Crews and Huang
(2008), Huang and Crews (2008) and Huang et al. (2010) reported on an improvement of the leakoff rate of these fluids. The
idea was to use nanoparticles that are less than 100 nm in size (35 nm average particle diameter). These are zinc oxide
particles. The particles crosslink the VES which results in a crosslinked-like network. The network builds a filter cake which
reduces leakoff. The nature of the crosslinking between the particles and the VES is believed to be a result of van der Waals
and electrostatic forces. The surfactant was an amidoamine oxide. The particles were added at a concentration of 10 pptg. The
surfactant concentration in the final solution was 2 vol%. Upon the addition of the nanoparticles, the zero shear viscosity was
improved from 15 to 1,100 poise. It is claimed that the nano-VES fluid improved the proppant settling time from 15 minutes to
more than 90 minutes. To test the leakoff performance of the nano-VES system, the fluid was forced through ceramic disks
(400 mD, 0.25 in. thick) under 300 psi pressure drop. The addition of the nano particles improved the leakoff rate significantly
where the leakoff volume dropped from 425 ml in 5 minutes to 80 ml in 70 minutes. However, a 30 pptg guar crosslinked with
borate gave less leakoff (50 ml in 70 minutes). In summary, the addition of nano particles improved the performance of the
VES-based fluid. It is worth mentioning that the authors suggested the use of a hydrophobic breaker that is based on
unsaturated fatty acids to break the viscosity of the VES-nano fluids.
Sullivan et al. (2006) reported on the use of a zwitterionic surfactant in gravel packing operations of reservoirs with
permeabilities larger than 100 mD in the Gulf of Mexico. A gas well was hydraulically fractured to produce from a sandstone
reservoir having a temperature of 195 oF (90.5oC) and permeabilities ranging from 100 to 420 mD. The authors reported that
the proppant was successfully placed but no gas production data was presented. Another well with a bottom hole temperature
of 190oF (87.8oC) and a permeability of 45 to 170 mD was fractured with the zwitterionic VES system. No production data
was reported for this well. Moreover, the mechanism by which the leakoff control was achieved was not reported.
Although VES fluids can break efficiently in the lab, data indicated that 20% of field treatments utilizing VES-based
fluids in the 1990s required the use of remedial actions to revive wells fractured with these fluids (Crews et al. 2008). VESbased fluids can break upon contact with hydrocarbons or when the salt content is reduced in the mixing brine. These two
conditions may not be met in all cases. For example, when reservoirs do not produce hydrocarbons, then the VES does not
break. Crews (2005) reported the use of breakers with VES fluids. The VES system reported by Crews (2005) was a non ionic
VES. The author investigated 4 breakers that can be used over the temperature range from 80 to 225 oF (107oC). The chemistry
of these breakers is reported in patent literature (Crews 2009). The breaker is based on a metal ion (such as iron), a reducing
agent (such as ascorbic acid) and a chelating agent (such as sodium citrate). The metal attacks the micellar structure of the
VES and breaks it. The chelating agent and the reducing agents are used to fine tune the breaking time. Reducing agents tend
to reduce the required amount of metal (can be a factor of 10) needed to break the VES. The breakers gave good results in the
lab where VES viscosity values were always reduced to below 20 cp.
Crews and Huang (2007) reported another breaker technology for VES-based fluids. Two major mechanisms are utilized.

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

The first is by adding breakers that degrade the VES molecule and hence, preventing it from creating the micellar structures
and the subsequent viscosity generation. The second approach is by adding materials that generate breakers which destroy the
micellar structures. Crews (2010) reported the chemistry involved in this technology where use is made of oils (soybean or fish
oils) that contain unsaturated fatty acids. Upon heating, the unsaturated fatty acids produce fatty acids, aldehydes and ketones.
These materials alter the micellar structure and break the VES. The exact mechanism by which these materials function is still
not clearly understood.
In order to improve the cleanup of an amino oxide-based VES system, Crews and Huang (2007) reported a study on the
use of breakers for this class of VES. The breaker is thought to be part of the VES enabling it to be in both the matrix (as a
result of leakoff) or within the proppant pack. The nature of this bonding between the breaker and the VES is not reported.
However, it is claimed that the breaker does produce a hydrocarbon and thus initiates the VES breakage. Gomaa et al. (2011)
investigated the rheological behavior as well as the proppant carrying capacity of this VES system. It has been shown that
increasing the total surfactant concentration (TSC) improves the viscosity of the system where the viscosity increased from 10
to 100 cp upon increasing the TSC from 7.5 to 60.5 pptg. At the same time, it improved the proppant suspension capacity
from 3 to 90% after 40 minutes of suspension. Similarly, the elastic modulus was found to increase with adding more
surfactant. However, data on higher temperatures (more than 60oC) was not reported.
Foamed Fluids
The term: energized fluids refers to a fluid system consisting of one gas component and a liquid. In some cases, these fluids
are referred to as foams. Foamed fracturing fluids reduce the amount of water to be used in the fracturing treatment. These
fluids can facilitate fluid flowback and thus are used in water sensitive formations with high clay content. Riberio and Sharma
(2009) present a systematic approach on which energized fluid is to be used under which conditions. Although these
treatments can be costly, they can show better performance than non-energized fluids (Burke et al. 2011). The most commonly
used gases in these fluids are either carbon dioxide (CO2) or nitrogen (N2). At temperatures more than 88oF (31oC) and
pressures typical of those in fracturing treatments (above 1,000 psi), CO2 behaves as a supercritical fluid. In such a case, the
energized fluid is considered an emulsion. It is quite often that literature uses these two terms interchangeably (Gupta and
Hlidek 2010; Arias et al. 2008). Note that CO2 has more hydrostatic pressure than N2 and therefore, it is preferred in cases
where the formation breakdown pressure is high. This section will report the major advancements in this class of fracturing
fluids over the last decade.
Chen et al. (2005) reported on the development of a VES system that is compatible with CO2. Since CO2 is a nonpolar
fluid, it has properties that are similar to hydrocarbons and therefore, it disturbs the worm-like micellar structure causing a
drop in viscosity. The authors utilized a new concept where additional surfactants are added to the mixture and they stabilize
the CO2 droplet within the VES solution. In this case, the fluid is an emulsion rather than a foam. The system can gel with 2%
KCl and has a maximum temperature stability of 230oF (110oC). The best viscosity results were obtained with CO 2 qualities of
70%. The viscosity of the emulsion was found to be 125 cp at 200 oF (93oC) and 100 s-1. This compares to 50 cp with the
straight VES without the CO2. The fluid was tested in a reservoir in North Texas with an average bottom hole temperature of
150 to 170oF and a permeability of 0.01 to 0.1 mD. Gas rates after these treatments were not reported. However, for one of the
treatments, the authors reported that pressure transient analysis confirmed a 29% increase in productivity compared to an
offset well treated with a 40 pptg guar gel. Arias et al. (2008) reported the use of the same system in fracturing wells in
Wyoming. The fluid was used to transport sand at concentrations of 5 to 6 ppg. The authors reported almost a 30% increase in
cumulative gas production of these wells when compared to those fractured with a linear guar gel over a 6-month period.
Gupta and Leshchyshyn (2005) reported on the use of a VES that is based on a mixture of anionic and cationic surfactants
that form a lamellar structure (and hence increased viscosity) if the charges on both surfactants become equal. The VES is
foamed with supercritical CO2. The fluid was used in 3,100 oil wells in Canada where the maximum reservoir temperature was
212oF (100oC) and the permeability ranged from 0.1 mD to 10 D. Most of these treatments utilized nitrogen gas. The fluid
viscosity was sufficient to transport 5 to 6.7 ppg of proppant. No hydrocarbon data was reported in this paper to show the
effectiveness of these treatments.
In order to address the thermal stability issues of VES-foamed fluids, Gupta and Carman (2011) reported on the use of
associative polymers as an additional component to the VES-foamed fluid system. The polymer is thought to have an
advantage in treating clays. The viscosity increase is due to the thermosetting nature of the polymer. This happens when it is
heated to a certain temperature at which the thermo sensitive blocks aggregate. The system was able to maintain a viscosity of
150 cp at 260oF (126.7oC) under a shear rate of 100 s-1.

Nanotechnology
The advancements in the field of nanotechnology have been utilized by the oilfield to resolve some of the challenges related to
many applications. Nanotechnology provides an opportunity due to its well-designed shapes, high surface to volume ratios and
its small size allowing its particles to propagate through reservoir rocks. In this section, we present some of the recent
literature utilizing nanotechnology for hydraulic fracturing. In the surfactant section, some of the advancements were already
highlighted that are related to VES.

10

Al-Muntasheri

[SPE 169552]

Fakoya et al. (2013) studied the effect of adding silica particles to guar polymer gels, surfactants and mixtures of
surfactants and guar polymers. The authors utilized 20 nm silica particles. Upon the addition of more particles to a non
crosslinked guar gel, the solution viscosity increased until it doubled by the addition of 0.4 wt% particles. The data is
summarized in Table 17. The mechanism by which the particles increase the viscosity of the guar solution is believed to be by
adsorption of the polymer on the silica particles. Then, the polymer starts forming aggregates that increase its viscosity. More
work is needed to quantify the effects of these particles on viscosity of crosslinked gels and their temperature stability.
Lafitte et al. (2012) introduced a new concept by which nanolatex silica is synthesized with boronic acid to yield what is
known as boronic acid-functionalized nanoparticles. Fig. 7 shows the chemical structure of the new crosslinker where the blue
ball refers to the silica particle. The idea is to use these suspensions to reduce the concentration of boron used in conventional
systems. For example, upon the addition of 2 ppm of boron to the guar polymer, the resulting gel had a viscosity of 800 to
2,000 cp depending on shear rate. It was observed that the maximum viscosity difference between the nanolatex-crosslinked
gel and a conventional borate-crosslinked gel takes place at guar concentrations less than 16 pptg at low shear rates. It should
be mentioned that these particles have a size less than 30 nm with a specific area of 300 m 2/g. One more advantage of these
new crosslinkers is the independent nature of their resulting gels viscosity compared to conventional borates. The breaking
mechanism of these suspension-based fluids remains unanswered along with their behavior at high temperatures. Yet, they
show a window of opportunity.

Fracturing of Shale Reservoirs


A major advancement in horizontal drilling and hydraulic fracturing allowed economical production of shale reservoirs. These
reservoirs are characterized by nano darcy permeabilities. Moreover, some of them can contain water-sensitive clays such as
smectite, illite or kaolinite (Maley et al. 2013). These clays can cause fines migration upon contact with brine or can swell. In
either case, the permeability of the rock is compromised. Obviously, with these low permeabilities, crosslinked fluids can
seriously cause excessive productivity losses. Thus, use is made of water with a drag reducer as a fracturing fluid. It is usually
referred to as slickwater. It is worth mentioning that water fracturing is not new. Several papers report that it dates back
to1950s (Wang and Miskimins 2010) and 1970s (Grieser et al. 2003). Literature also reported hybrid treatments where
slickwater is combined with low concentrations of polymers (Handren and Palisch, 2009).
Unlike polymers, water has poor proppant suspension properties due to lack of structure and viscosity. As such, very little
proppant can be transported. This is usually at a concentration of 0.25 to 2.5 pounds of proppants added per gallon of
fracturing fluid (ppa) (Palisch et al. 2010). In order to place large volumes of proppants, slickwater treatments are injected at
very high rates that can be 120 bbl/min (Kaufman et al. 2008). Obviously, with low ppa of proppant, these treatments will
require very large volumes of water that can reach one million gallons. In addition to this disadvantage, slickwater treatments
present a challenge when it comes to proppant placement in vertical fractures. Due to the lack of suspension in water,
proppants tend to settle leaving large portions of fractures unpropped (Palisch et al. 2010). However, it is reported that seismic
when used to characterize hydraulically induced fractures, revealed more complex fractures upon the use of slickwater
compared to other fluids (Brannon and Bell, 2011). It is believed that this is due to the longer fracture length in the case of
slickwater. In view of the high rates encountered with these fluids where very turbulent flow conditions exist, friction reducers
are added. This section will shed light on advancements related to the fluids used in shale resources. These include new
polymers as friction reducers and new designs of slickwater treatments.
Sun et al. (2011) reported on the development and field application of a new friction reducer. The system is based on a
synthetic polymer that can break with oxidizers efficiently. It is believed that the polymer is PAM-based. It comes as a waterin-oil emulsion. Upon contact with the fluid stream during injection through the wellbore, the emulsion inverts where it
becomes oil-in-water emulsion. As such, the polymer is released and hydrated in water to reduce friction. This process is very
time sensitive. The authors reported an example where it only takes 3 minutes for the fluid to reach the perforation (based on
well tubing diameter of 5.5 in, length of 15,000 ft and an injection rate of 100 bbl/min). Field data showed that the system
requires a maximum of 0.75 gpt compared to 1 to 1.25 gpt of similar friction reducers. The system was used with an oxidizing
breaker package in oil and gas producing shales in Texas Panhandle field. The data indicated 47% increase in oil cumulative
production of the wells treated with the new polymer and the breaker. On the gas wells side, 30% increase was realized.
Sun et al. (2013) reported on the use of a synthetic polymer as a fracturing fluid in shale reservoirs. This was done to
address the challenges associated with the use of slickwater. Slickwater is used at high rates which translate into high
velocities (100 bbl/min through a 4 inches-diameter pipe translates into a velocity of 107 ft/s). The use of proppants at high
velocities leads to blasting of the pumping equipment. As such, polymers are used at low concentrations to substitute
slickwater. Guar has been used as a replacement, yet, its increased costs in the recent years posed a challenge. Therefore, the
authors reported the use of a synthetic polymer with an oxidizing breaker. Their work shows that the system can provide good
suspension at low concentration of less than 3 gpt of polymer. The authors reported field applications of the system in south
Texas in the Eagle Ford shale. The comparison data indicated that the system gives better results compared to treatments
utilizing guar-based polymers. Kostenuk and Browne (2010) reported a new approach by which the proppant settling is
reduced in slickwater fracturing. The idea is to coat the proppants with a material that helps create bubbles around each grain.
Thus, the buoyancy of the proppant is increased and its settling tendencies are reduced. The authors report around 23%
improvement in proppant settling with the new technique.

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

11

Conclusions
The paper presents a review of the available water-based fracturing fluids over the last decade. The review revealed the
evolution of several attempts. These include: use of cleaner guar-based polymers, use of synthetic polymers and the use of
larger sized crosslinker molecules. In addition, the use of high density brines as mixing fluids to reduce surface pumping
requirements is highlighted. Nanotechnology can be a promising tool to present better-performing fluids. The following major
conclusions can be drawn:
1. The true concept of a zero-damaging fluid is still nonexistent.
2. Guar-based polymers are still being used despite their old introduction to the oil and gas field. This is because of their
high shear stability and better clean up compared to other systems.
3. With fracturing hot (400oF) reservoirs, synthetic polyacrylamide-based polymers can be used. However, their cleanup
is still a question.
4. The use of larger-sized boronic-based crosslinkers helped reduce the concentration of polymers used in preparing
fracturing fluids. However, no field application data has been reported yet for this system.
5. Nano particles were used to improve leakoff of VES-based fluids. They were also used to produce crosslinkers at
lower concentrations. The use of these materials still requires more research. However, the use of the advances of
nanotechnology provides a window of opportunity in the area of fracturing fluids.
6. Produced water can present an opportunity to address environmental concerns about water preservation. This
becomes very important in shale reservoirs where millions of gallons of water are used. However, not all produced
waters are compatible with fracturing fluids.
7. The environmental impact of fracturing fluids chemicals remains to be a major public concern.

Acknowledgements
The author would like to thank Mr. Ahmed Al Sarraj for his help with the preparation of some of the figures presented in this paper.
Mr. Jim Crews, Dr. David Loveless and Prof. Hisham A. Nasr-El-Din are appreciated for the many useful discussions. Thanks to
Aramco Services Company and Saudi Aramco for permission to publish this work.

Unit Conversion
3
1 lb/1000 gal = 0.119826427 kg/m

Nomenclature

Quantity
Brine viscosity

Surface treating pressure


Bottom hole treating pressure
Friction pressure
Hydrostatic pressure

Symbol

Units

mPa.s (cp)
psi

psi
psi
psi

12

Al-Muntasheri

[SPE 169552]

Table 1: Summary of fracturing fluids reviewed in this paper


Fluid Type

PAM-Based/Zirconium
PAM-Based/Zirconium and
activator
Linear Biopolymer
Low Molecular Weight
HPG/Borate
AMPS-vinyl
phosphonate/Zirconium
Linear guar
Guar/ boronic acidfunctionalized nanoparticles
VES (Quaternary ammonium)
VES (Amidoamine oxide)
VES Zwitterionic
VES/CO2 Emulsion

Polymer
Concentration,
lb/1000 gal
66

Crosslinker
Concentration,
lb/1000 gal
3

87

3.4 to 3.7

40-70
Not Available

Not Available

40

Not Available

Working
Temperature
Range
269- 449.6oF
(132-232oC)
100.4- 269oF
(38-132oC)
<200oF (93oC)
200-275oF
(93-135oC)
425oF (218oC)

Reference

33
30

0
0.0166

<175oF (79oC)
100oF (37oC)

Fakoya et al. (2013)


Lafitte et al. (2012)

Not Applicable
Not Applicable
Not Applicable
Not Applicable

Not Applicable
Not Applicable
Not Applicable
Not Applicable

<180oF (82oC)
<250oF (121oC)
<195oF (90oC)
230oF (110oC)

Samuel et al. (1999, 2000)


Huang and Crews (2008)
Sullivan et al. (2006)
Chen et al. (2005)

Holtsclaw and Funkhouser


(2010)
Holtsclaw and Funkhouser
(2010)
Walters et al. (2009)
Weaver et al. (2002a and b)
Gupta and Carman (2011)

Table 2: Viscosity data of linear gels reported in this paper


Polymer Type

Concentration,
lb/1000 gal

Measurement Conditions

Viscosity, cp

Reference

Ely (1989)
Harris and
Heath. (1998)
Ely (1989)
Fischer et al.
(2001)
Fischer et al.
(2001)
Fischer et al.
(2001)

Guar
Guar

80
45

Temperature, oF
(oC)
77 (25)
77 (25)

Guar
Guar

20
25

77 (25)
77 (25)

511
50

10
50

Guar

20

77 (25)

50

15

Guar/Nonacetylyzed
Xanthan (NAX)

77 (25)

50

110

Finely Processed Guar

25
(with 4:1 ratio of
guar to XAN)
37.6

150 (65.6)

170

>150

Clean biopolymer

40-70

250 (121)

40

216

Clean biopolymer

40-70

96 (35)

40

352

Less residues polymer

25

150 (65.6)

100

580

Less residues polymer

25

200 (93)

100

500

Shear Rate,
1/s
511
511

103
42

Qun et al.
(2008)
Walters et al.
(2009)
Walters et al.
(2009)
Williams et
al. (2012)
Williams et
al. (2012)

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

13

Table 3: Effect of the use of high yield guar and new crosslinker on the viscosity of crosslinked gels (Williams et al. 2012)
Polymer Type

High Yield
Guar/New Borate
Crosslinker
Regular
Guar/Regular
Borate Crosslinker
High Yield
Guar/New Borate
Crosslinker
Regular
Guar/Regular
Borate

Polymer Loading,
pptg

Measurement Conditions

Viscosity, cp

Shear Rate, 1/s

25

Temperature, oF
(oC)
150 (65.6)

100

580

30

150 (65.6)

100

440

25

200 (93)

100

500

30

200 (93)

100

420

Table 4: Field application data of PAM-based polymer crosslinked with zirconium (Funkhouser et al, 2010)
Well
Number

Reservoir
Depth, ft

Reservoir
Temperature,
o o
F( C)

Proppant
Content,
lb/gal

Pad
Volume,
gallons

Number of
Frac
Stages

1
2

23,324
18,830

450 (232.2)
433 (222.7)

1 to 8
1 to 7.4

25,000
30,000

8
7

Post
Treatment
Gas Rate,
MSCF/day
Zero
4,933

Table 5: Viscosity data of PAM-based high temperature fracturing gels


(After shearing at 100 s-1 for two hours with no breaker)
Polymer
Content,
gal/1000 gal

Crosslinker
Content,
gal/1000 gal

40

Data Not
Available
Data Not
Available
Data Not
Available
0.5
1.5

40
50
66
66

Measurement Conditions

Viscosity,
cp

Reference

Temperature, oF (oC)

Shear Rate, 1/s

400 (204.4)

100

1,020

Gupta and Carman (2011)

425 (218.3)

100

1,000

Gupta and Carman (2011)

450 (232.2)

100

200

Gupta and Carman (2011)

399 (203.9)
399 (203.9)

100
100

200
450

Holtsclaw and Funkhouser (2010)


Holtsclaw and Funkhouser (2010)

14

Al-Muntasheri

[SPE 169552]

Table 6: Fracture conductivity data of proppant packs


Polymer/Crosslinker
Type

Polymer
Content,
lb/1000 gal
87

Breaker Content,
gal/1000 gal

Shut-in
Time,
hours
48

Temperature

Closure
Stress,
psi
5,000

PAM-Based/Zirconium
and activator

87

0.3 to 1.5

2,500

42-72

450oF
(232oC)

34 to 100

PAM-Based/Zirconium

66

4,000

71

399oF
(204oC)

43-73

PAM-Based/Zirconium

66

8,000

118

399oF
(204oC)

39-47

PAM-Based/Zirconium

66

10,000

165

399oF
(204oC)

37

40-70

Not Needed

6,000

NA

NA

6,000

200oF
(93.3oC)
200oF (93oC)

90

Guar/Borate
AMPS-vinyl
phosphonate/Zirconium
AMPS-vinyl
phosphonate/Zirconium

40

NA

3,000

Not
Reported
Not
Reported
24

64-69

40

NA

3,000

24

425oF
(218oC)
385oF
(196oC)

PAM-Based/Zirconium

Linear Biopolymer

399oF
(204oC)

Regained
Conductivity
,%
68

14

83

Reference

Holtsclaw and
Funkhouser
(2010)
Holtsclaw and
Funkhouser
(2010)
Holtsclaw and
Funkhouser
(2010)
Holtsclaw and
Funkhouser
(2010)
Holtsclaw and
Funkhouser
(2010)
Walters et al.
(2009)
Walters et al.
(2009)
Gupta and
Carman (2011)
Gupta and
Carman (2011)

Table 7: Effect of shear rate on viscosity of associative PAM-based polymers at 90oC (Gaillard et al. 2013)

Hydrophobic
Yes
Yes
Yes

Monomers Available
NaAMPS
Acrylamide
Yes
Yes
No
Yes
No
Yes
Guar

Viscosity at 10 s-1, cp

Viscosity at 100 s-1, cp

8,000
5,000
900
80

40
40
40
10

Table 8: Fracture conductivity data of high density polyamphoteric/surfactant gel proppant packs (Gupta et al. 2012)
Polymer
Content,
lb/1000 gal
50
50
60

Salt Type

Brine Density,
lb/gal

Closure Stress,
psi

Temperature,
o o
F( C)

Regained Gas
Conductivity, %

Regained Water
Conductivity, %

CaCl2
CaCl2
NaBr/NaCl

11
11
12

9,000
9,000
9,000

200
250
250

95
94
93

81
93
91

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

15

Table 9: Fann 35 viscosity data (300 rpm) showing effect of crosslinker size on viscosity (Sun and Qu, 2011)
Crosslinker Type

Polymer
Concentration,
lb/1000 gal

Viscosity,
cp

24
16
17
16

510
580
650
550

Boric acid (BA)


Thiophenediboronic aid (TDBA)
Benzenediboronic acid (BDBA)
Biphenyldiboronic acid
(BPDBA)

Table 10: Effect of glyoxal on vortex closure time for delayed polyaminoborates crosslinking 25 pptg guar at room
temperature (Legemah et al. 2013)
Glyoxal Loading, %

Buffer Loading, gpt

Crosslinker Loading, gpt

0
4
8
15

2.4
2.4
2.6
3

4.0
4.2
4.6
4.6

Vortex Closure Time,


seconds
150
176
207
240

Table 11: Various types of breakers with their operating conditions (Brannon et al. 1994; Sarwar et al., 2011; Parker and
Laramy, 1992; Crews 2013)
Temperature
pH
Polymer Breaker
Range, oF
Range
70-140
Low Temperature Enzymes
3-8.5
100-230
High Temperature Enzymes
4-9.5
Low Temperature Oxidizers: Sodium and
Potassium Persulfates
Mid Temperature Oxidizers: Calcium and
Magnesium Peroxides
High Temperature Oxidizers: Sodium and
Potassium Bromate
Encapsulated Low Temperature Oxidizers
(Persulfates)
Encapsulated Mid Temperature Oxidizers
(Metal Peroxides)
Encapsulated High Temperature Oxidizers
(Bromates)

80-180
150-225
220-310
100-200
150-230
220-340

2-12
8-13
4-12
2-12
7-13
4-12

16

Al-Muntasheri

[SPE 169552]

Table 12: Release data of various encapsulated breakers


Encapsulated Breaker
Type
Ammonium persulfate

Coating Type

Temperature, oF (oC)

Not reported

175 (79.4)

Potassium persulfate

Not reported

200 (93.3)

Sodium persulfate

Not reported

200 (93.3)

Ammonium persulfate
(20% coating)

Polyvinylidene
Chloride

160 (71.1)

Time,
hours
1
3
6
0.42
1.5
6
0.25
0.8
2.1

% Breaker Released

Reference

6
42
60
30
80
100
40
83
100

Norman et al. (1995)

0.5
1
1.5

40
60
90

Gulbis et al. (1992)

Patil et al. (2013)

Patil et al. (2013)

Table 13: Effect of encapsulation on sodium persulfate performance in breaking CMHPG/Zr gels at 200 oF (93oC) (Patil et al.
(2013))
Breaker Type
Neat sodium persulfate
Encapsulated sodium
persulfate
Encapsulated sodium
persulfate

Breaker Concentration,
lb/1000 gal
0.7
2.8

Time for the Gel to


Reach 50 cp, minutes
30
120

4.7

40

Table 14: Effect of stabilizers and produced water on viscosity of borate crosslinked gels (Li et al. 2009)
Sample

Shear Rate, 1/s

1
2

Viscosity after 2 hours,


cp
740
50

100
100

Temperature,
o o
F( C)
125 (52)
125 (52)

3
4

600
320

100
100

125 (52)
135 (93)

50

100

135 (93)

Comments
Guar/borate gel prepared in distilled water
Same as sample 1 but with an enzyme
breaker
Same as sample 2 but with a stabilizer
Guar/borate gel treated with stabilizer and
then prepared in produced water
Same as sample 4 but with no stabilizer

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

17

Table 15: Chemical analysis of produced water used in preparing fracturing gels (Li et al. 2010)

Component, mg/L

Produced Water Type

Sodium
Potassium
Magnesium
Calcium
Chloride
Bicarbonate
Sulfate

Offshore Oilfield in California

Oilfield in New Mexico

9,930
230
970
260
18,400
100
> 1,600

11,100
150
450
1,160
20,740
170
<400

Table 16: Chemical analysis of produced water (LeBas et al. 2013)


Parameter
Measurement before
Measurement after
Electrocoagulation
Electrocoagulation
Treatment
Treatment
Specific gravity
1.2
1.203
pH
4.83
8
Dissolved Oxygen
8.24
8.45
Chloride
163,637
164, 951
Sulfate
40
38
Aluminum
1.42
2.28
Barium
5.69
6.03
Calcium
29,222
28,845
Conductivity, (microS/cm)
257
258
Turbidity (NTU)
182
15.4
TDS (mg/L)
267,588
273, 552
TSS (mg/L)
10,623
92

Table 17: Effect of nanomaterial on viscosity data of 33 pptg guar linear gels (Fakoya et al. 2013)
Nano SiO2
Content, wt%

0
0.058
0.24
0.4

Measurement Conditions
Temperature, oF
(oC)
175 (79)
175 (79)
175 (79)
175 (79)

Viscosity, cp

Shear Rate, 1/s


50
50
50
50

24
30
37
50

18

Al-Muntasheri

[SPE 169552]

Figure 1: Chemical structures of polymers used in hydraulic fracturing (Ely, 1989)

Fig.1 a: Guar

Fig.1 b: Hydroxypropyl Guar (HPG)

\\\\

Figure 1c: Carboxymethylhydroxypropyl Guar (CMHPG)

Figure 1d: Caboxymethylhydroxyethyl cellulose (CMHEC)

Figure 1e: Carboxymethyl cellulose (CMC)

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

19

Figure 1f: Hydroxyethyl cellulose (HEC)

Figure 2: Chemical structure of the non acetylated xanthan, NAX (Fischer et al. 2001)

Figure 3: Chemical structure of the PHPA/2-acrylamido-2-methylpropanesulfonic acid (AMPS) copolymer (Holtsclaw and
Funkhouser, 2010)

20

Al-Muntasheri

[SPE 169552]

Figure 4: Chemical structures of larger-sized boronic acid-based crosslinkers (Sun and Qu, 2011)
Fig. 4a: Benzenediboronic acid (BDBA)

Fig. 4b: Biphenyldiboronic acid (BPDPA)

Fig. 4c: Thiophenediboronic acid (TDBA)

Figure 5: Chemical structures of polyethyleneimine (right) and 3-aminomethylbenzylamine (left) (Sun and Qu, 2011)

Figure 6: Delay time and maximum viscosity data for chelating delay agent
(175oF, 40 pptg CMHPG, 39 ppm Zr+4 from sodium zirconium lactate (Putzig and Clair, 2007))

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

21

Figure 7: Nanolatex silica synthesized with boronic acid (Lafitte et al. 2012)

References
Ainley, B.R., Nimerick, K.H., and Card, R.J. 1993. High-Temperature, Borate-Crosslinked Fracturing Fluids: A Comparison
of Delay Methodology. Paper SPE 25463 presented at the Production Operations Symposium, Oklahoma City, OK,
U.S.A, 21-23 March.
Arias, R.E., Nadezhdin, S.V, Hughes, K., and Santos, N. 2008. New Viscoelastic Surfactant Fracturing Fluids Now
Compatible with CO2 Drastically Improve Gas Production in Rockies. Paper SPE 111431 presented at the SPE
International Symposium and Exhibition on Formation Damage Control, Lafayette, LA, USA, 13-15 February.
Arthur, J.D., Hochheiser, H.W. and Coughlin, B.J. 2011. State and Federal Regulation of Hydraulic Fracturing: a Comparative
Analysis. Paper SPE 140482 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas,
24-26 January.
Bartko, K., Arocha, C., and Mukherjee, T.S. 2009. First Application of High Density Fracturing Fluid to Stimulate a High
Pressure and High Temperature Tight Gas Producer Sandstone Formation of Saudi Arabia. Paper SPE 118904 presented
at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 19-21 January.
Black, H.N., and Hower, W.E. 1965. Advantageous Use of Potassium Chloride Water for Fracturing Water-sensitive
Formations. Paper API 65-113 Drilling and Production Practice, 1965.
Brannon, H., and Bell, C.E., 2011. Eliminating Slickwater Compromises for Improved Shale Stimulation. Paper SPE 147485
presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, USA, 30 October-2 November.
Bu-Khamseen, R., Khakimov, A., Sierra, L., Young, D., and Machala, M. 2010. First Succesful Tight Gas Reservoir
Completion and Fracture Stimulation in Sarah Formation, Rub Al-Khalil Empty Quarter of Saudi Arabia. Paper SPE
130722 presented at the SPE Deep Gas Conference and Exhibition, Manama, Bahrain, 24-26 January.
Burke, L.H., Nevison, G.W., Peters, W.E. 2011. Improved Unconventional Gas Recovery With Energized Fracturing Fluids:
Montney Example. Paper SPE 149344 presented at the SPE Eastern Regional Meeting held in Columbus, Ohio, 17-19
August.
Cawiezel, K.E., and Gupta, D.V.S. 2010. Successful Optimization of Viscoelastic Foamed Fracturing Fluids With
Ultralightweight Proppants for Ultralow-Permeability Reservoirs. SPE Production and Operations 25 (1): 80-88.
Chen, Y., Pope, T.L., and Lee, J.C. 2005. Novel CO2-Elumsified Viscoelastic Surfactant Fracturing Fluid System. Paper SPE
94603 presented at the SPE European Formation Damage Conference, 25-27 May, Sheveningen, The Netherlands.
Clark, J.B. 1949. A Hydraulic Process for Increasing the Productivity of Wells. Journal of Petroleum Technology 1 (1): 1-8.
Clark, J.B., Fast, C.R., and Howard, G.C. 1952. A Multiple-fracturing Process for Increasing the Productivity of Wells. Paper
API 52-104 Drilling and Production Practice, 1952.
Coronado, J.A. 2007. Success of Hybrid Fracs in the Basin. Paper SPE 106758 presented at the Production and Operations
Symposium, Oklahoma City, Oklahoma, USA, 31 March-3 April.
Crews, J. 2013. Private communication. 13 November.
Crews, J. 2005. Internal Phase Breaker Technology for Viscoelastic Surfactant Gelled Fluids. Paper SPE 93449 presented at
the SPE International Symposium on Oilfield Chemistry, Houston, Texas, USA, 2-4 February.
Crews, J. 2009. Metal-Mediated Viscosity Reduction of Fluids Gelled with Viscoleastic Surfactants. US Patent 7,595,284.
Crews, J. 2010. Composition and Use of Mono and Polyenoic Acids for Breaking VES-Gelled Fluids. US Patent 7.645,724.
Crews, J. and Huang, T. 2007. Internal Breakers for Viscoelastic-Surfactant Fracturing Fluids. Paper SPE 106216 presented at
the SPE International Symposium on Oilfield Chemistry, Houston, Texas, USA, 28 February-2 March.
Crews, J. and Huang, T. 2008. Performance Enhancements of Viscoelastic Surfactant Stimulation Fluids With Nanoparticles.
Paper SPE 113533 presented at the SPE Europec/EAGE Annual Conference and Exhibition, Rome, Italy, 9-12 June.
Crews, J., Huang, T. and Wood, W.R. 2008. The Future of Fracturing-Fluid Technology and Rates of Hydrocarbon Recovery.
Paper SPE 115475 presented at the SPE Annual Technical Symposium and Exhibition, Denver, Colorado, 21-24
September.
Dysart, G.R., Spencer, A.L., and Anderson, A.L. 1969. Blast-fracturing. Paper API 60-068 Drilling and Production Practice,
1969.
Ely, J.W. 1989. Fracturing Fluids and Additives. Vol. 12 of Recent Advances in Hydraulic Fracturing (SPE Henry L Doherty
Monograph Series). SPE, Richardson, Texas.
Fakoya, M.F., and Shah, S.N. 2013. Rheological Properties of Surfactant-Based and Polymeric Nano-Fluids. Paper SPE
163921 presented at the 2013 SPE/ICoTA Coiled Tubing and Well Intervention Conference and Exhibition, The
Woodlands, TX, USA, 26-27 March.
Fischer, C.C., Navarrete, K., Constien, V.G., Coffey, M.D., and Asadi, M. 2001. Novel Application of Synergistic Guar/Non-

22

Al-Muntasheri

[SPE 169552]

Acetylated Xanthan Gum Mixtures in Hydraulic Fracturing. Paper SPE 65037 presented at the SPE International
Symposium on Oilfield Chemistry, Houston, Texas, USA, 13-16 February.
Fontana, C., Muruaga, E., Perez, D., Cavazzoli, G. and Krenz, A. 2007. Successful Application of a High Temperature
Viscoelastic Surfactant (VES) Fracturing Fluid under Extreme Conditions in Patagonian Wells, San Jorge Basin. Paper
SPE 107277 presented at the SPE Europec/EAGE Annual Conference and Exhibition, London, UK, 11-14 June.
Funkhouser, G.P., Holtsclaw, J. and Blevins, J. 2010. Hydraulic Fracturing Under Extreme HPHT Conditions: Successful
Application of a New Synthetic Fluid in South Texas Gas Wells. Paper SPE 132173 presented at the SPE Deep Gas
Conference and Exhibition, Manama, Bahrain, 24-26 January.
Funkhouser, G.P. and Norman, L. 2003. Synthetic Polymer Fracturing Fluid for High-Temperature Applications. Paper SPE
80236 presented at the SPE International Symposium on Oilfield Chemistry, Houston, Texas, USA, 5-7 February.
Gaillard, N., Thomas, A., and Favero, C. 2013. Novel Associative Acrylamide-based Polymers for Proppant Transport in
Hydraulic Fracturing Fluids. Paper SPE 164072 presented at the 2013 SPE International Symposium on Oilfield
Chemistry, The Woodlands, TX, USA, 08-10 April.
Ghauri, W.K. 1960. Results of Well Stimulation by Hydraulic Fracturing and High Rate Oil Backflush. Journal of Petroleum
Technology 12 (6):19-27.
Gulbis, J., King, M.T., Hawkins, G.W., and Brannon, H.D. 1992. Encapsulated Breaker for Aqueous Polymeric Fluids. SPE
Production Engineering 7 (1): 9-14.
Gunawan, S., Armstrong, C.D., and Qu, Q. 2012. Universal Breakers with Broad Polymer Specificity for Use in Alkaline,
High-Temperature Fracturing Fluids. Paper SPE 159396 presented at the SPE Annual Technical Conference and
Exhibition, San Antonio, Texas, USA, 8-10 October.
Gomaa, A.M., Cawiezel, K.E., Gupta, D.V.S., Carman, P. and Nasr-El-Din, H.A. 2011. Viscoelastic Evaluation of A
Surfactant Gel for Hydraulic Fracturing. Paper SPE 143450 presented at the SPE European Formation Damage
Conference,Noordwijk, The Netherlands, 7-10 June.
Grieser, B., Hobbs, J., Hunter, J. and Hubter, J. 2003. The Rocket Science Behind Water Frac Design. Paper SPE
80933presented at the SPE Productions and Operations, Oklahoma City, Oklahoma, 23-25 March.
Grossman, W.L. 1951. Hydrafrac Operations in the Spraberry Production, West Texas. Paper SPE 128 presented at the Fall
Meeting of the Petroleum Branch of AIME, Oklahoma City, Oklahoma, USA, 3-5 October.
Gupta, D.V.S. and Carman, P. 2011. Fracturing Fluid for Extreme Temperature Conditions is just as Easy as the Rest. Paper
SPE 140176 presented at the SPE Hydraulic Fracturing Technology Conference and Exhibition, The Woodlands, Texas,
USA, 24-26 January.
Gupta, D.V.S., Carman, P., and Venugopal, R. 2012. High-Density Brine-Based Fracturing Fluid for Ultra-Deep Fracturing
Stimulation. Paper SPE 159822 presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas,
USA, 8-10 October.
Gupta, D.V.S. and Hlidek, B.T. 2010. Frac-Fluid Recycling and Water Conservation: A Case History. SPE Production and
Facilities Journal 25(2): 65-69.
Gupta, D.V.S. and Leshchyshyn, T.T. 2005. CO 2 Energized Hydrocarbon Fractruing Fluid: History and Field Application in
Tight Wells in Rock Greek Gas Formation.
Hanes, Jr., R.E., Weaver, J.D., and Slaubaugh, B.F. 2006 Methods and Compositions for Reducing the Viscosity of Treatment
Fluids. US Patent 7,082,995.
Handren, P. and Palisch, T. 2009. Successful Hybrid Slickwater-Fracture Design Evolution: An East Texas Cotton Valley
Taylor Case History. SPE Production and Operations 24 (3): 415-424.
Harris, P.C. 1993. Chemistry and Rheology of Borate-Crosslinked Fluids at Temperatures to 300F. Journal of Petroleum
Technology 45 (3): 264-269.
Harris, P.C. and Heath, S.J. 1998. Rheological Properties of Low-Gel-Loading Borate Fracture Gels. SPE Production and
Facilities 13 (4): 230-235.
Harris, P.C. and Sabhapondit, A. 2009. Chemistry Applied to Fracture Stimulation of Petroleum Wells. Paper SPE 120029
presented at the SPE Middle East Oil and Gas Show and Conference, Bahrain, Bahrain, 15-18 March.
Hepburn, K. 2012. Development and Practical Implementation of a Chemical Hazard Rating System. Paper SPE 160548
presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, USA, 8-10 October.
Holtsclaw, J. and Funkhouser, G.P. 2010. A Crosslinkable Synthetic-Polymer System for High-Temperature HydraulicFracturing Applications. SPE Drilling and Completion 25 (4): 555-563.
Holtsclaw, J., Loveless, D., Saini, R., and Fleming, J. 2011. Environmentally Focused Crosslinked-Gel System Results in High
Retained Proppant-Pack Conductivity. Paper SPE 146832 presented at the SPE Annual Technical Conference and
Exhibition, Denver, Colorado, USA, 30 October-2 November.
Huang, F., Gundewar, R., Steed, D., and Loughridge, B. 2005. Feasibility of Using Produced Water for Crosslinked Gel-Based
Hydraulic Fracturing. Paper SPE 94320 presented at the SPE Production Operations Symposium, Oklahoma City,
Oklahoma, USA, 16-19 April.
Huang, T. and Crews, J. 2008. Nanotechnology Applications in Viscoelastic Surfactant Stimulation Fluids. SPE Production
and Operations 23 (4): 512-517.
Huang, T, Crews, J and Agrawal, G. 2010. Nanoparticle Pseudocrosslinked Micellar Fluids Optimal Solution for Fluid-Loss

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

23

Control With Internal Breaking. Paper SPE 128067 presented at the International Symposium and Exhibition on
Formation Damage Control, Lafayette, Louisiana, USA, 10-12 February.
Hurst, W. 1953. Establishment of the Skin Effect and its Impediment to Fluid Flow into a Well Bore. The Petroleum Engineer.
Petroleum Engineering, Dallas. 25 (Oct.): 36-38, B6 through B16.
Jacoby, D. 2011. Global Trade Restrictions and Related Compliance Issues Pertaining to Oil and Gas Production Chemicals.
Paper OTC 22005 presented at the Offshore Technology Conference, Houston, Texas, USA, 2-5 May.
Jennings, A.R. 1996. Fracturing Fluids - Then and Now. Journal of Petroleum Technology 48 (7): 604-610.
Jianshe, G., Zhou, B., Xia, A., Lu, H., and Li, Z. 2012. A New Fracturing Fluid of Low Concentration. Paper SPE presented
at the IADC/SPE Asia Pacific Drilling Technology Conference and Exhibition, Tianjin, China, 9-11 July.
Kalgaonkar, R. and Patil, P. 2012. Performance Enhancements in Metal-Crosslinked Fracturing Fluids. Paper SPE 152040
presented at the North Africa Technical Conference and Exhibition, Cairo, Egypt, 20-22 February.
Kaufman, P., Penny, G.S., and Paktinant, J. 2008. Critical Evaluations of Additives Used in Shale Slickwater Fracs. Paper SPE
119900 presented at the SPE Shale Gas Production Conference held in Forth Worth, Texas, USA, 16-18 November.
King, M.T., Gulbis, J., Hawkins, G.W., and Brannon, H.D. 1990. Encapsulated Breaker For Aqueous Polymeric Fluids. Paper
PSOC 90-89 presented at the Annual Technical Meeting, Calgary, Alberta, Canada, 10-13 June.
Kostenuk, N. and Browne, D.J. 2010. Improved Proppant Trasnport System for Slickwater Shale Fractruing. Paper SPE
137818 presented at the SPE Canadian Unconventional Resources and International Petroleum Conference, Calgary,
Alberta, Canada, 19-21 October.
Lafitte, V., Tustin, G., Drochon, B., and Parris, M. 2012. Nanomaterials in Fracturing Applications. Paper SPE 155533
presented at the SPE International Oilfield Nanotechnology Conference, Noordwijk, The Netherlands, 12-14 June.
LeBas, R., Lord, P., Luna, D., and Shahan, T. 2013. Development and Use of High-TDS Recycled Produced Water for
Crosslinked-Gel-Based Hydraulic Fracturing. Paper SPE 163824 presented at the 2013 SPE Hydraulic Fracturing
Technology Conference, The Woodlands, TX, USA, 04-06 February.
Legemah, M.U., Guerin, M., Sun, H., and Qu, Q. 2013. Novel High-Efficiency Boron Crosslinkers for Low-Polymer-Loading
Fracturing Fluids. Paper SPE 164118 presented at the 2013 SPE International Symposium on Oilfield Chemistry, The
Woodlands, TX, USA, 08-10 April.
Li, L., Eliseeva, K., Eliseev, V., Bustos, O.A., England, K., Howard, P.R., Boney, C.L., Parris, M.D., and Ali, S.A. 2009. Well
Treatment Fluids Prepared with Oilfield Produced Water. Paper SPE 124212 presented at the SPE Annual Technical
Conference and Exhibition, New Orleans, Lousiana, USA, 4-7 October.
Li, L., Ezeokonkwo, C.I., Lin, L., Eliseeva, K., Kallio, W., Boney, C.L., Howard, P., and Samuel, M.M. 2010. Well Treatment
Fluids Prepared with Oilfield Produced Water: Part II. Paper SPE 133379 presented at the SPE Annual Technical
Conference and Exhibition, Florence, Italy, 19-22 September.
Lo, S., Miller, M.J., and Li, J. 2002. Encapsulated Breaker Release Rate at Hydrostatic Pressure and Elevated Temperatures.
Paper SPE 77744 presented at the SPE Annual Technical Conference and Exhibition, 29, San Antonio, Texas, USA, 29
September-2 October.
Loveless, D., Holtsclaw, J., Saini, R., Harris, P., and Fleming, J. 2011. Fracturing Fluid Comprised of Components Sourced
Solely from the Food Industry Provides Superior Proppant Transport. Paper SPE 147206 presented at the SPE Annual
Technical Conference and Exhibition, Denver, Colorado, USA, 30 October- 2 November.
Loveless, D., Saini, R., and Weaver, J. 2012. US Patent application, 12/0000659 A1.
Maley, D., Farion, G., Gabriela, Giurea-Bica and ONeil, B. 2013. Non-Polymeric Permanent Clay Stabilizer for Shale
Completions. Paper SPE 165168 presented at the SPE European Formation Damage Conference and Exhibtion held in
Noordwijk, The Netherlands, 5-7 June.
Nasr-El-Din, H.A., Al-Mohammed, A.M., Al-Aamri, A.D., and Al-Fuwaires, O.A. 2007. A Study of Gel Degradation, Surface
Tension, and their Impact on the Productivity of Hydraulically Fractured Gas Wells. Paper SPE 109690 presented at the
Annual SPE Technical Conference and Exhibition held in Anaheim, California, USA, 11-14 November.
Nimerick, K.H., Temple, H.L., and Card, R.J. 1997. New pH-buffered Low-Polymer Borate-Crosslinked Fluids. SPE Journal 2
(2): 150-156.
Nolte, K.G. 1985. Fracturing Fluid Breaker System which is Activated by Fracture Closure. US Patent 4,506,734.
Norman, L., Vitthal, S., and Terracina, J. 1995. New Breaker Technology for Fracturing High-Permeability Formations. Paper
SPE 30097 presented at the SPE European Formation Damage Conference, The Hague, Netherlands, 15-16 May.
Padgett, J.C. 1951. Information on Hydrafac Process. Paper WPC 4142 presented at the 3rd World Petroleum Congress, The
Hague, the Netherlands, 28 May -6 June.
Palisch, T.T., Vincent, M.C., and Handren, P.J. 2010. Slickwater Fracturing: Food for Thought. SPE Production and
Operations Journal, 25(3):327-344.
Palmore, L., McKenzie, and Foster, R., 2013. High-Performance Fracture Fluid Outperforms Conventional Fluids. World Oil,
224(6):42-46.
Pandey, V.J. and Robert, J.A. 2002. New Correlation for Predicting Frictional Pressure Drop of Proppant-Laden Slurries using
Surface Pressure Data. Paper SPE 73753 presented at the International Symposium and Exhibition on Formation Damage
Control, Lafayette, Louisiana, USA, 20-21 February.

24

Al-Muntasheri

[SPE 169552]

Patil, P., Muthusamy, R., and Pandya, N. 2013. Novel Controlled-Release Breakers for High-Temperature Fracturing. Paper
SPE 164656 presented at the 2013 North Africa Technical Conference and Exhibition, InterContinental Citystar, Cairo,
Egypt, 15-17 April.
Paulaus, W., and Ed 2005. Directory of Microbiocides for the Protection of Materials: A Handbook. The Netherlands,
Springer.
Peles, J., Wardlow, R.W., Cox, G., Haley, W., Dusterhoft, R., Walters, H., and Weaver, J. 2002. Maximizing Well Production
with Unique Low Molecular Weight Frac Fluid. Paper SPE 77746 presented at the SPE Annula Technical Conference and
Exhibition, San Antonio, Texas, 29 September-October 2.
Powell, R.J., McCabe, M.A., Slabaugh, B.F., Terracina, J.M., Yaritz, J.G., and Ferrer, D. 1997. Applications of a New,
Efficient Hydraulic Fracturing Fluid System. Paper SPE 38956 presented at the Latin American and Caribbean Petroleum
Engineering Conference, Rio de Janeiro, Brazil, 30 August-3 September.
Putzig, D.E., and St. Clair, J.D. 2007. A New Delay Additive for Hydraulic Fracturing Fluids. Paper SPE 105066 presented at
the SPE Hydraulic Fracturing Technology Conference, College Station, Texas, USA, 29-31 January.
Qiu, X., Martch, E., Morgenthaler, L., Adams, J., and Vu, H. 2009. Design Criteria and Application of High-Density BrineBased Fracturing Fluid for Deepwater Frac Packs. Paper SPE 124704 presented at the SPE Annual Technical Conference
and Exhibition, New Orleans, Lousiana, USA, 4-7 October.
Qun, L., Jiang, T., Yun, X., Ding, Y., Lu, Y., Bo, C., Shu, Y., and Duan, Y. 2008. The Study and Application of Low-Damage
and Massive Hydraulic Fracturing Technique in Tight Gas Formations With High Temperature and High Pressure. Paper
SPE 114303 presented at the IPC/SPE Gas Technology Symposium 2008 Joint Conference, Calgary, Alberta, Canada, 1619 June.
Reddy, B.R. 2013. Laboratory Characterization of Gel Filter Cake and Development of Non-Oxidizing Gel Breakers for
Zirconium Crosslinked Fracturing Fluids. Paper SPE 164116 presented at the 2013 SPE International Symposium on
Oilfield Chemistry, The Woodlands, 8-10 April.
Riberio, L.H and Sharma, M. 2013. Fluid Selecion for Energized Fracture Treatments. Paper SPE 163867 presented at the SPE
Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 4-6 February.
Robart, C., Ruegamer, M., and Yang, A. 2013. Analysis of U.S. Hydraulic Fracturing Design Trends. Paper SPE 163875
presented at the 2013 SPE Hydraulic Fracturing Technology Conference, The Woodlands, TX, USA, 04-06 February.
Samuel, M., Card, R.J., Nelson, E.B., Brown, J.E., Vinod, P.S., Temple, H.L., Qu, Q., and Fu, D.K. 1999. Polymer-Free Fluid
for Fracturing Applications. SPE Drilling and Completion 14 (4): 240-246.
Samuel, M., Polson, D., Kordziel, W., Waite, T., Waters, G., Vinod, P.S., Fu, D., and Downey, R. 2000. Viscoelastic
Surfactant Fracturing Fluids: Application in Low Permeability Reservoirs. Paper SPE 60322 presented at the SPE Rocky
Mountain Regional/Low Permeability Reservoirs Symposium and Exhibition, Denver, Colorado, USA, 12-15 March.
Sarwar, M.U., Cawiezel, K.E., and Nasr-El-Din, H.A. 2011. Gel Degradation Studies of Oxidative and Enzyme Breakers to
Optimize Breaker Type and Concentration for Effective Break Profiles at Low and Medium Temperature Ranges. Paper
SPE 140520 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 24-26
January.
Simms, L., and Clarkson, B. 2008. Weighted Frac Fluids for Lower-Surface Treating Pressures. Paper SPE 112531 presented
at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA, 13-15
February.
Sullivan, P.F, Gadiyar, B., Morales, R.H., Hollicek, R., Sorrells, D., Lee, J., and Fischer, D. 2006. Optimization of a
Viscoelastic Surfactant (VES) Fracturing Fluid for Application in High Permeability Reservoirs. Paper SPE 98338
presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA,
15-17 February.
Sun, H., and Qu, Q. 2011. High-Efficiency Boron Crosslinkers for Low-Polymer Fracturing Fluids. Paper SPE 140817
presented at the SPE International Symposium on Oilfield Chemistry, The Woodlands, Texas, USA, 11-13 April.
Sun, H., Wood, B., Stevens, D., Cutler, J., Qu, Q., and Lu, M. 2011. A Nondamaging Friction Reducer for Slickwater Frac
Applications. Paper SPE 139480 presented at the SPE Hydraulic Fracturing Technology Conference and Exhibition, The
Woodlands, Texas, USA, 24-26 January.
Sun, H., DeBenedicts, F., Zhou, J., Culter, J., Royce, T., Nelson, S., and Qu, Q. 2013. Field Case Histories of a Non-Damaging
Guar Alternative for Linear Gel Application in Slickwater Fracturing. Paper SPE 165130 presented at the SPE European
Formation Damage Conference and Exhibition, Noordijk, The Netherlands, 5-7 June.
Tan, H.C. and McGowen, J.M. 1995. Field Experiment Gathers Friction-Pressure Data for CO2-Energized Fluids. SPE
Production and Facilities 10 (4): 213-218.
Tonmukayakul, N, Saini, R.K., Loveless, D.M., Funkhouser, G.P., Liang, F., Fitzgerald, R.M., Holtsclaw, J., Harris, P.C., and
Norman, L.R. 2011. Clean Viscosified Treatment Fluids and Associated Methods US Patent Application, US
20110214860.
Walters, H.G., Stegent, N., and Harris, P.C. 2009. New Frac Fluid Provides Excellent Proppant Transport and High
Conductivity. Paper SPE 119380 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands,
Texas, USA, 19-21 January.
Wang, Y., and Miskimins, J.L. 2010. Experimental Investigations of Hyraulic Fracture Growth Comlexity in Slickwater

[SPE 169552]

A Critical Review of Hydraulic Fracturing Fluids over the Last Decade

25

Fracturing Treatments. Paper SPE 137515 presented at the SPE Tight Gas Completions Conference, San Antonio, Texas,
USA, 2-3 November.
Weaver, J., Schmelzl, E., Jamieson, M., and Schiffner, G. 2002a. New Fluid Technology Allows Fracturing Without Internal
Breakers. Paper SPE 75690 presented at the SPE Gas Technology Symposium, Calgary, Alberta, Canada, 30 April-2 May.
Weaver, J.D., Wilson, S.F., Bowles, B.K., Slabaugh, B.F., Parker, M.A., Barrick, D.M., Heath, S.J., Walters, H.D. and Cole,
R.C. 2002b. Subterranean Formation Treating Fluid Concentrates, Treating Fluids and Methods. US Patent 6,488, 091.
Williams, N.J., Kelly, P.A., Berard, K.G., Dore, E.L., Emery, N.L., Williams, C.F., and Mukhopadhyay, S. 2012. Fracturing
Fluid with Low-Polymer Loading Using a New Set of Boron Crosslinkers Laboratory and Field Studies. Paper SPE
151715 presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette,
Louisiana, USA, 15-17 February.

S-ar putea să vă placă și