You are on page 1of 35

SPE-173776-MS

Recent Advances in Viscoelastic Surfactants for Improved Production from


Hydrocarbon Reservoirs
Katherine L. Hull, and Mohammed Sayed, Aramco Research CenterHouston; Ghaithan A. Al-Muntasheri,
Aramco Research CenterHouston, Saudi Aramco

Copyright 2015, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE International Symposium on Oilfield Chemistry held in The Woodlands, Texas, USA, 1315 April 2015.
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
Viscoelastic surfactants (VES) are used in upstream oil and gas applications, particularly hydraulic
fracturing and matrix acidizing. A description of surfactant types is introduced along with a theoretical
description of how they assemble into micelles, what sizes and shapes of micelles can be formed under
different conditions, and finally how specific structures can lead to bulk viscoelastic solution properties.
This theoretical discussion leads into a description of the specific VES systems that have been used over
the last twenty years or so in improved oil recovery for upstream applications.
VES-based fluids have been used most extensively for hydraulic fracturing. They are preferred over
conventional polymer-based fracturing fluid systems because they are essentially solids-free systems
which have demonstrated less damage to the reservoir rock formation. Important advancements in VES
have been made by introducing pseudo-crosslinking agents such as nanoparticles to enhance the
viscosity. Fracturing fluid systems based on VES have also been improved recently by developing internal
breakers to lower their viscosity in order to flow back the well. The flexibility of VES-based fluids has
been demonstrated by their application as foamed fluids as well as their incorporation with brine systems
such as produced water.
A second key area that has benefited from VES-based systems is matrix acidizing carbonated-based
reservoirs. The viscosity of these VES-based fluids is mostly controlled by pH where, at low pH (low
viscosity), the acid system flows easily and invades pore spaces in the formation. During acidizing, the
acid is spent, and the pH and viscosity increase. Because the spent acid has higher viscosity, fresh acid
is diverted to low permeability un-contacted zones and penetrates the rocks to form wormholes. A number
of experimental studies and field applications to these effects have been performed and will be described
here.
In order for VES-based fluids to play a more prominent role in the field, inherent limitations such as
cost, applicable temperature range, and leak-off characteristics will need to continue to be addressed. If
we can efficiently and economically overcome these issues, VES-based fluids offer the industry an
excellent clean, non-damaging alternative to conventional polymer-based fluids.

SPE-173776-MS

Introduction
Surfactants have been used in a wide variety of industrial products including cleaning detergents, textiles,
cosmetics, paper production, food, mining, as well as fluids for the oil and gas industry. Their versatile
nature has allowed them to be utilized for as emulsifiers, wetting agents, and foaming agents. Surfactants
are amphiphilic organic molecules which consist of a component which, on its own, would be soluble in
a given liquid and a second component which, on its own, would not be soluble in the same liquid (Witten
and Pincus 2010). In aqueous environments, the hydrophilic head group interacts favorably with the
solvent medium while the hydrophobic tail has a more favorable free energy when away from the solvent,
concentrating at the liquid boundary. The surfactant molecules form an interface between two immiscible
liquids and larger quantities of surfactant lead to more interfacing of the two liquids until eventually they
are considered mixed. The specific chemical identity of the polar head groups and hydrocarbon tail groups
varies but surfactant molecules are typically broken into classes which include anionic, cationic, nonionic,
and zwitterionic species. Common examples of each surfactant class include carboxylate or sulfate polar
head groups (anionic), quaternary ammonium head groups (cationic), long chain alcohols (nonionic), and
betaines (zwitterionic). Table 1 shows each of the surfactant types, their charge, as well as names and
structures of specific examples.
Table 1A list of the types of surfactant molecules is provided along with specific examples that correspond with each of the
classes. Studies of the phosphobetaine (x10, n1-10) have been described (Chevalier 1988 & 1992).

Under certain conditions, surfactant molecules arrange into colloidal structures called micelles, where
the hydrocarbon tails of the surfactants orient towards each other while the polar head groups form an
interface with the surrounding aqueous media. Figure 1 provides a simple schematic to illustrate the
micellization process. The size and structure of these micelles is controlled by a variety of parameters and
will be described in later discussions. Surfactant micelles form spontaneously in aqueous solution when
the surfactant concentration, c, exceeds the threshold referred to as the critical micelle concentration
(cmc). Certain properties of aqueous solutions of surfactants such as osmotic pressure, turbidity, surface
tension, and electrical conductivity are observed to experience discontinuities at this critical concentration,
providing confirmation of the transition from discrete molecules to aggregates (Hargreaves 2003). For

SPE-173776-MS

example, the cmc of various solutions of cationic surfactants was determined by plotting the electrical
conductivity against the surfactant concentration. Two linear slopes were observed for all surfactants, and
the intersection of the two lines is the cmc (Gravsholt 1976).

Figure 1Schematic of a surfactant molecule depicting the polar head group and the hydrocarbon tail group, along with its
transformation into a colloidal structure when the surfactant concentration exceeds the cmc.

The effects of various factors on the cmc have been noted, including temperature, surfactant type,
hydrophobe chain length, and salts/cosolutes (Hamley 2007). The cmc is largely independent of temperature except in the case of nonionic surfactants, whose hydrophilic head groups are based on oxyethylene
groups, where increasing temperature causes the cmc to decrease. In the case of ionic versus nonionic
surfactants with equal hydrophobic chain lengths, the cmc is usually lower for nonionic surfactants
because the electrostatic repulsions between ionic headgroups are more difficult to overcome during
micelle formation. Similarly, as the size of the polar head group increases for ionic surfactants, the cmc
increases. However, as the length of the hydrophobic tail of the surfactant increases up to around 16
carbon atoms, the cmc decreases. Beyond this chain length, there is little change observed in the cmc.
Finally, salt addition to ionic surfactant solutions decreases the cmc because the ions reduce head group
repulsions.
A. Thermodynamics of Micelle Formation
When surfactant molecules are dissolved in aqueous solution, the attractive and repulsive forces between
the surfactant molecules and water cause micelle assembly, if the concentration of surfactant exceeds the
cmc. Initial work by Tanford (1974 & 1979) gave rise to a free energy model to describe the formation
and growth of these surfactant aggregates resulting from (1) the tail transfer, (2) the hydrocarbon-water
interface, and (3) the head group interactions. The free energy change associated with an aggregate of size
g (aggregation number) can be summarized with the following relationship in Equation 1 (Nagarajan
2002):
(1)
where k is the Boltzmann constant and T is the temperature in Kelvin. The tail transfer free energy (
arises from the transition of the hydrocarbon tail from the aqueous environment to the hydrocarbon-rich environment at the center of the micelle. This change provides a favorable contribution to the
overall free energy. By contrast, both the interfacial and head group interactions are unfavorable
contributions to the overall free energy. The interfacial free energy results from the residual contact that
the hydrocarbon tail has with water, while the head group free energy corresponds with the crowding
(repulsive forces) that takes place at the surface of the micelle. Both the interfacial free energy and the
head group free energy are dependent upon the surface area per molecule of the aggregate core, a. The
standard free energy change can be rewritten as in Equation 2:

SPE-173776-MS

(2)
where is the unit area for the residual contact between hydrocarbon and water and is the headgroup
repulsion parameter. For micelles in thermodynamic equilibrium, where a ae, the standard free energy
(g/kT) is minimized (Equations 3-5). Furthermore, Table 2 describes each contribution to the critical
micelle concentration in mole fraction units, Xcmc, and how each influences either the formation of
aggregates or the size of the aggregates.
Table 2Contributions to the overall free energy of micellization are provided along with physical descriptions of their effects
upon the aggregation behavior of the micelle.

(3)
(4)
(5)
The role of the surfactant hydrocarbon tail on micellization free energy has also been considered
(Nagarajan 2002). The hydrocarbon chains pack differently in the bulk as compared to the micelle where
they deform non-uniformly. Although the tail is constrained to a fixed location where it adjoins the polar
head group, the other end is free to position itself anywhere within the core of the micelle as long as the
core maintains a uniform density of hydrocarbon chains. The result is an additional contribution to the
overall free energy of micellization which accounts for the packing (Nagarajan and Ruckenstein 1991).
(6)
The packing free energy of the micelle core per surfactant molecule is then given for spheres, cylinders,
and lamellae, respectively:
(7)
where R is the radius of the micelle, L is the characteristic segment length, and N is the number of
segments. The numerical values 3, 5, and 10 correspond to the molecular packing differences in each of
the geometries (Nagarajan and Ruckenstein 1991).
B. Micelle Size and Shape
The size and shape of micelles is dictated by a wide range of factors including surfactant properties such
as charge, geometry, and concentration as well as solution conditions such as temperature, ionic strength,
type and concentration of salt, and shear rate. The molecular packing parameter is a classical method for

SPE-173776-MS

qualitatively predicting the micelle structure and is still widely used (Israelachvili et al. 1976; Israelachvili
1992). The parameter is defined as 0/ael0 where 0 is the volume of the surfactant tail, l0 is the maximum
length of the surfactant tail, and ae is the surface area per molecule at the water-surfactant interface as
described previously. When the molecular packing parameter is 1/3, the surfactant molecules are
predicted to assemble into spherical aggregates. When the packing parameter is between 1/3 and 1/2, the
surfactant micelle is expected to adopt a rodlike or wormlike shape. Finally, vesicles and lamellar
structures are expected when the packing parameter is between 1/2 and 1 or close to 1, respectively. A
summary of the various micelle shapes and their corresponding packing parameters is provided in Figure
2.

Figure 2Schematic diagrams of each micelle type and their respective qualitative prediction of shape based on the packing parameter
(adapted from Nagarajan 2002 and van Zanten 2011).

The packing model can be used to predict how changes to the surfactant molecule or solution
conditions will influence the micelle structure. Surfactant molecules with large polar headgroups (large ae)
are expected to promote the formation of spherical micelles, while surfactants with small headgroups
should encourage lamellae formation. For example, nonionic surfactants with small ethylene oxide
headgroups (small number of carbons, m) should favor bilayer/lamellae structures. However, nonionic
surfactants with larger headgroups (e.g., m10, 12, 14, 16) yield cylindrical or wormlike micelles (Jerke
et al. 1998; Bernheim-Groswasser et al. 2000; Imanishi and Einaga 2007). Similarly, the packing model
can be used to predict that salt addition to ionic surfactant solutions will cause a transition from spherical
to cylindrical micelles. The salt interacts electrostatically with the polar headgroups, thus reducing the
headgroup repulsion parameter, , thereby decreasing ae and increasing the molecular packing parameter.
For example, the cationic surfactant hexadecyltrimethylammonium bromide, which is also referred to as
cetyltrimethylammonium bromide (CTAB), forms spherical micelles in aqueous solution. Upon addition
of sodium nitrate (NaNO3) to the surfactant solution, the spherical micelles transform into wormlike
micelles (Kuperkar et al. 2008). Another example of using the packing parameter to predict micelle
structure involves surfactants which are composed of double rather than single hydrophobic tails. These
surfactants will have molecular packing parameters which are twice as large and, given the same
headgroup, they will be much more likely to form lamellar structures.

SPE-173776-MS

C. Wormlike Micelles
The micelle shape of particular interest to oilfield applications is wormlike micelles. These are elongated
rodlike micelles which are long and flexible and can entangle in solution, imparting viscoelastic properties
to the fluid. Detailed structural, chemical, and behavioral characteristics of wormlike micelles have been
recently reviewed in the chemical literature (Dreiss 2007). The review provides an extensive list of
surfactants systems that have been reported to form wormlike micelles and the corresponding methods
used to characterize these fluids such as Cryo-Transmission Electron Microscopy (TEM), rheology, static
light scattering (SLS), and small angle neutron scattering (SANS).
The structure of wormlike micelles can be described with a series of different length scales. Figure 3,
below, provides a schematic of a wormlike micelle and the various parameters that are used to describe
the structure. The radius of gyration, Rg, reflects the radius around which the micelle rotates. The total
length of the micelle, Lc, can vary from nanometers to microns and can be determined by direct imaging
via Cryo-TEM. The persistence length, lp, is the length over which the micelle is considered to be rigid.
As an example, lp was measured for 6 mM cetyl pyridinium bromide with 0.8 M sodium bromide added
(Witten and Pincus 2010; Cates and Candau 1990). At 35 C, lp was found to be 20 nm while the
cross-section radius of the micelle, Rcs, was determined to be 2 nm.

Figure 3Wormlike micelle schematic showing the fundamental structural parameters that cover a wide range of length scales
(adapted from Dreiss 2007).

Wormlike micelles are dynamic structures which differ from polymer solutions, in that the aggregates
are constantly breaking and reforming, versus polymer solutions where the lengths of the chains are fixed
at the point of quenching during synthesis (Cates 1990; Cates and Candau 1990). These so-called living
polymers are in thermal equilibrium and have molecular weight distributions which are normally quite
broad. The dynamics of breaking and reforming micelles can have an effect on the chain entanglements.
This in turn has an effect upon the diffusion properties of the chains as well as the viscoelasticity. Micelle
breakage is assumed to be a unimolecular process where the lifetime of the chain before breaking into two
segments is given by the parameter break (Cates 1990; Cates and Candau 1990):
(8)
where
is the average contour length of the micelles before breaking and k1 is the rate constant. A
second stress relaxation mechanism is reptation, a diffusion phenomenon also observed in entangled
polymer solutions. From Equation 8, when break rep, micelles break and reform faster than the time
scale for reptation. In the limit when break rep is much less than break rep, the viscoelastic surfactant

SPE-173776-MS

solution behaves as a Maxwell fluid with a single relaxation time given by Equation 9 (Cates 1990; Cates
and Candau 1990):
(9)
The frequency, , dependent elastic (storage) modulus G= and the viscous (loss) modulus G for a
Maxwell fluid are given by the following expressions (Schubert et al. 2003):
(10)
(11)
Where G0 is the high-frequency plateau modulus (the elastic modulus at infinite frequency or t0).
The moduli G= and G can be determined experimentally by oscillatory-shear measurements where the
sample is deformed sinusoidally and the response is measured. The inverse of the crossover frequency of
G= and G is the relaxation time R. This model has been applied to a large number of viscoelastic micelle
solutions and is generally considered a good indication that wormlike micelles are present (Dreiss 2007).
Under steady shear flow, wormlike micelles exhibit a low-shear Newtonian plateau, which is characterized by a zero-shear viscosity, 0 G0R. However, when the critical shear rate is exceeded, the
wormlike micelle solution undergoes strong shear thinning. The longest structural relaxation time R can
be estimated as the inverse of the critical shear rate. Measurement of G= and G can be used to determine
whether or not a fluid is viscoelastic at a given temperature. A simple observational technique has also
been described for determining whether or not a fluid is viscoelastic (Gravsholt 1976) i.e., bubbles that
appear during swirling of a sample will recoil when the swirling stops if the solution is viscoelastic.
D. Improved Oil Recovery Applications
Viscoelastic surfactant fluids have been used in a variety of applications in the upstream oil and gas
industry including gravel packing, frac packing, fracturing fluids, and matrix acidizing. Each of these
applications will be touched on throughout the course of this paper. Substantial recent developments in
the areas of hydraulic fracturing and matrix acidizing have been seen over the last decade and will be
discussed in greater detail. In the case of fracturing fluids, a variety of nanoparticle additives have been
developed to pseudo-crosslink the surfactant micelles and enhance the viscosity of the fluids. Furthermore, methods for breaking the gelled fluids in order to flow back the well have been developed and will
be described. Viscoelastic surfactants have also been used in more specialized cases such as foamed fluids
and heavy brine solutions. A detailed description of their use in matrix acidizing will also be given,
including a variety of experimental studies and several case histories of field treatments with VES-based
fluids and hydrochloric acid.

Discussion
A. Fracturing Fluids
VES-based fluids were first reported for upstream oil and gas applications in gravel-pack completions
(Nehmer 1988) and frac-packs (Brown et al. 1996; Stewart et al. 1994) and were later also developed into
fluids that were used for hydraulic fracturing (Samuel et al. 1997, 1999, 2000). Conventional polymerbased fracturing fluid systems incorporate a water-soluble polymer, crosslinker, and breaker among other
additives (Al-Muntasheri 2014). The fluid has viscosity which is sufficient for transporting proppant into
the fractures. Over time, the gel is broken by enzyme or oxidizer and the fluid is flowed back from the
formation to the surface. This process is operationally complex in that it requires polymer hydration and
a variety of additives such as biocides, crosslinkers, and breakers. By contrast, viscoelastic surfactants are
simpler to use in the field because there is no hydration step and fewer additives are required. A major
advantage over polymer-based systems is that VES-based fluids are essentially solids-free which means

SPE-173776-MS

that residue is not deposited in either the formation or the proppant pack, so these fluids are more efficient
in hydraulic fracture reservoir stimulation. This feature of VES-based fluids is significant since the
conductivity of the in-place proppant pack have a significant effect upon the overall well productivity
(Palisch et al. 2007). In addition, VES-based fluids have been observed to heal after exposure to shear,
and additives have been developed to improve the shear rehealing time (Lee et al. 2008; Chen et al. 2008).
There are some drawbacks to VES-based fluids relative to conventional polymer-based fracturing fluids.
Because they are relatively solids-free, they do not form a filtercake, so high leak-off has been observed
in many cases. Also, the temperature range over which VES-based fluids can operate is also lower, with
the highest temperatures observed falling less than 300 F. Table 4 provides a summary of the various
characteristics observed about the nature of VES-based hydraulic fracturing fluids.
Table 4 A summary of characteristics associated with VES-based hydraulic fracturing fluids is provided along with a list of the
limitations of these fluids.

VES-based hydraulic fracturing fluid systems have shown better performance in field applications as
highlighted by several reports (Samuel et al. 1997, 1999, 2000). In Alberta, five wells stimulated with
VES-based fluid were compared with five offsets that were fractured with 20 pptg borate-crosslinked
guar. For wells that were fractured with the VES-based fluid, the absolute open flow increase was 9%
greater (Samuel et al. 1997). Similarly, when the same VES-based fluid was used to fracture a well in
southwest Kansas, the initial production rate from this well was approximately 27% greater than the
production rate from the parent well in the same section. The rate was also 52% greater than the average
of 12 adjacent wells hydraulically fluid fractured. In Rock Springs, Wyoming, two identical offset wells
were fractured with either VES-based fluid or with 25 pptg guar (Samuel et al. 2000). The calculated
hydraulic fracture lengths for both the polymer and the VES-based treatments were similar. The fracture
height generated by the guar treatment was estimated to be more than double the height for the VES-based
fluid treatment due to the higher viscosity of the polymer fluid. It was determined that fractures occurred
outside the pay zone for the polymer treatment, resulting in propped fractures in nonproductive zones.
Analysis of the flowback fluid indicated that the VES-based fluid cleaned up faster than the polymer fluid.
The well stimulated with VES-based fracturing fluid had an initial production of 2.8 MMscf/D whereas
the polymer-stimulated well produced at 1.3 MMscf/D. Four wells in Armstrong County, Pennsylvania
were treated with either VES-based fluids or linear polymer gels, and the well treated with VES-based
fluids showed initial production levels that were three to four times higher than the linear polymer gel
treatments (Leitzell 2007). Likewise, of the four wells treated in Clearfield County, Pennsylvania with

SPE-173776-MS

either VES-based fluids or linear gels, the treatments incorporating surfactants showed better sustained
production. A higher temperature VES fluid application was implemented in the El Tordillo field which
is located in the central portion of the San Jorge Basin (Fontana et al. 2007). The well consisted of four
individual zones, and each was treated with a 6% VES-based fracturing fluid. Improved performance
relative to conventional treatments was noted in terms of reduced fracture height growth and, therefore,
reduced proppant volumes by 30-50% because fracture placement was concentrated within the zone of
interest.
A.1. VES-Based Fluid Chemistries
A variety of viscoelastic surfactant chemistries have been utilized for fluids used in hydraulic fracturing.
Examples of cationic, nonionic, zwitterionic, and anionic surfactants have all been utilized for hydraulic
fracturing applications primarily in aqueous systems, although some examples of gelled hydrocarbon
systems have been reported (Samuel 2009; Samuel et al. 2014). Beginning with cationic surfactants, early
studies of quaternary ammonium salts established their viscoelastic properties (Gravsholt 1976). Solutions
which incorporate water, a water soluble salt (electrolyte), and a quaternary ammonium salt have been
used for drilling fluids, completion fluids, hydraulic fracturing fluids, etc. (Teot et al. 1988). Figure 4(a)
shows the structure of one preferred surfactant system oleyl methyl bis(2-hydroxyethyl) ammonium
chloride. Quaternary ammonium salt chemistry was also utilized for developing a fluid system that is
stable up to 225 F for hydraulic fracturing applications (Norman et al. 1996). Erucyl bis(2-hydroxyethyl)
methylammonium chloride (EHAC), shown in Figure 4(b), was combined with ammonium chloride or
potassium chloride and, in some cases, also with an organic salt such as sodium salicylate, resulting in a
viscous fluid that can be used to fracture high permeability formations. This surfactant system is known
to form long wormlike micelles which entangle and impart viscoelastic characteristics to the fluid
(Raghavan and Kaler 2001). Further developments with the same surfactant system have also been seen,
showing insensitivity to a range of pH values as well as compatibility with seawater-based fluid systems
(Brown et al. 1999; Samuel et al. 2001). EHAC can alternatively be combined with a polymer such as guar
or modified guar in concentrations below 10-15 pptg to provide a dual system (Miller et al. 2003). The
polymer is crosslinked with boron, zirconium, or another metal to the extent that it forms a filter cake on
the formation face, thereby enhancing fluid loss control. Alternative cationic surfactants such as the
gemini surfactant dimethylene-1,2-bis(dodecyldimethylammonium bromide) display viscoelastic behavior at lower concentrations (0.7-1.7%) than conventional surfactants (Yang et al. 2013). The critical
micelle concentration of these dimeric surfactants is usually ~10 times lower than the conventional
monomeric surfactants.

Figure 4 The structures of specific quaternary ammonium salts that have been used for hydraulic fracturing applications are shown
including (a) oleyl methyl bis(2-hydroxyethyl) ammonium chloride, (b) erucyl bis(2-hydroxyethyl) methylammonium chloride (EHAC),
and (c) N,N,N, trimethyl-1-octadecammonium chloride.

Cationic surfactants, however, present potential problems to the formation in that they can oil-wet
formation rocks, which can lead to increased resistance to oil flow, so surfactant fluids composed of
anionic or nonionic surfactants may be preferred (Whalen 2000). In the case of nonionic surfactants,
amido amine oxides have been described for hydraulic fracturing applications. One specific example that

10

SPE-173776-MS

has been reported is tallow amido propylamido oxide, TAPAO (McElfresh and Williams 2007). The
general form of this compound is shown below in Figure 5 along with specific components of the tallow
substituent. A fluid composed of 6 vol% TAPAO and 3% KCl was prepared, and the viscosity was
measured at a shear rate of 100 s-1. The viscosity of the fluid was maintained above 100 cp in the
temperature range of 100-200 F and around 40 cp at 225 F.

Figure 5The general structure of the nonionic surfactant tallow amido propylamine oxide (TAPAO) is shown, along with the major
components of the tallow amido substituent.

Hydraulic fracturing VES-based fluids that are composed of anionic surfactants such as alkyl sarcosinates have been seen and can also be applied to gravel packing, frac packing, and water control issues
(Di Lullo Arias et al. 2002). The fracturing fluid includes a combination of a water soluble salt such as
potassium or sodium chloride, 0.5-6 wt% alkyl sarcosinate surfactant, a buffer for adjusting the pH to
6.5-10, and either 0.1-2 wt% carboxylic acid salt as an additional source of ions or, alternatively, 3-6 wt%
chloride or fluoride salts. The sarcosinate has been used for fracturing and has approximately 94% oleoyl
sarcosine as shown in Figure 6(a). Sufficient sarcosinate is present in solution to provide adequate
viscosity to the fluid in order to be able to transport proppant into the hydraulically created fractures. The
anionic surfactant methyl ester sulfonate (MES), which is shown in Figure 6(b), has also been incorporated as a treatment fluid (Welton et al. 2007b). 5 wt% MES was observed to gel (viscosity 20 cp at 511
s-1) in the presence of 5-10 wt% inorganic salt at pH4, ~7, and 10. 5 wt% MES was also observed to
gel in the presence of 5 wt% inorganic salt and 10% HCl, but not in the presence of 15% HCl. The MES
surfactant is generated by adding sulfur trioxide to the -carbon of a methyl ester and then neutralizing
with a base (Welton et al. 2007a). It is considered to be more environmentally friendly than other
surfactant gels because it is derived from renewable resources such as palm kernel oil, is biodegradable,
and is less toxic. Several fracturing treatments were performed in South America with an environmentally
friendly anionic surfactant (EHAC) at 4 vol% (Di Lullo et al. 2001) and were reported to improve
production in three case studies.

SPE-173776-MS

11

Figure 6 The chemical structure of (a) oleoyl sarcosine is shown, which constitutes around 94% of the sarcosinate product, along
with the structure of the surfactants (b) methyl ester sulfonate where R is an alkyl chain with 10-30 carbon atoms and (c) sodium xylene
sulfonate.

VES-based fluids composed of zwitterionic/amphoteric surfactants have also been reported, including
dihydroxyl alkyl glycinate, alkyl ampho acetate or propionate, alkyl betaine, alkyl amidopropyl betaine,
and alkylimino mono- or di-propionates derived from waxes, fats, or oils (Dahanayake et al. 2001, 2002,
2004). The surfactant is combined with either an inorganic salt or an organic compound such as phthalic
acid, salicylic acid, or their salts. Figures 7(a) and 7(b) show two specific surfactants that were reported
including disodium tallowiminodipropionate and disodium oleamidopropyl betaine. The viscosity of a
VES solution composed of 5% disodium tallowiminodipropionate and 2.25% phthalic acid was determined at variable shear rates. When the shear rate varied from 1 to 100 s-1, the viscosity ranged from
almost 100,000 cp down to 100 cp at both 25 and 50 C. Similar results were also obtained when
1.75-2.0% phthalic acid and 4% of ammonium chloride (NH4Cl) were added to 5% surfactant instead. The
viscosity versus shear rate was also reported for 4-5% disodium oleamidopropyl betaine, 3% KCl, and
0.5% phthalic acid. As the shear rate increased from 0.01 to ~30 s-1, the viscosity decreased from
approximately 100,000 to 1000 cp. In the Gulf of Mexico, 17 fracturing treatments utilizing zwitterionic
VES-based hydraulic fracturing fluids in concentrations ranging from 3.5-6.0% were reported by Sullivan
et al. (2006). The bottom hole static temperature ranged from 135-214 F, and the permeability of the
formation was as high as 170 md. Welton and Bryant (2011) furthermore describe another formulation
incorporating 7 wt% of the previously described oleylamidopropyl betaine, as in Figure 7(b), 0.14 wt%
potassium stearate (soap), and 2 wt% polyamide (nonaqueous tackifying agent). Steady shear viscosities
obtained in the temperature range of 175-225 F were enhanced relative to the same formulation that did
not contain the tackifying agent. Fluids incorporating a similar zwitterionic surfactant, erucylamidopropyl
betaine as shown in Figure 7(c), have been described for enhanced oil recovery (Morvan and Degre
2012).

Figure 7The chemical structures of (a) tallowiminodipropionate where Rtallow and (b) oleylamidopropyl betaine and (c) erucylamidopropyl betaine.

In some cases, mixed surfactant systems have been developed. For example, quaternary amine
surfactant systems have been used in combination with anionic surfactants for fracturing applications. A

12

SPE-173776-MS

solution composed of N,N,N, trimethyl-1-octadecammonium chloride, shown in Figure 4(c), and sodium
xylene sulfonate shown in Figure 6(c), for example, forms a viscoelastic gel with viscosity in the range
of 20-500 cp at a shear rate of 100 s-1 (Zhang 2002). In a follow-up study of this mixed surfactant system,
12 different ratios of the cationic/anionic (C/A) surfactants were evaluated in terms of proppant suspension ability as well as elastic and viscous properties (Gomaa et al. 2011). A number of conclusions were
drawn about the formulations examined including (1) the gel viscosity was highly dependent upon the
total surfactant concentration and the temperature and (2) good proppant suspension was only observed
when the C/A ratio was higher than 1.5 and the total surfactant concentration was greater than 30 gpt. In
addition to mixed surfactant systems, mixed surfactant/polymer systems have been formulated and tested
(Gomaa et al. 2014). A low molecular weight associative polymer, with no added crosslinker, along with
surfactant was evaluated in 33 different formulations. A polymer-to-surfactant ratio of 1.4 produced the
broadest working temperature range with the highest elastic characteristics and the best proppant
suspension.
A.2. Pseudo Crosslinkers
Recent studies in the chemical literature have established that inorganic particles can be added to
wormlike micelles to enhance the solution viscosity. For example, barium titanate (BaTiO3) pyroelectric
nanoparticles increase the viscosity of solutions composed of sodium fatty acid methyl ester sulfonate
(Luo et al. 2012). Bandyopadhyay and Sood (2005) studied the rheology of semidilute solutions of the
cationic surfactant cetyl trimethylammonium tosylate (CTAT) in the presence of silica colloids with
diameters of 0.1 m. Upon addition of 1.3 wt% silica particles to 1.4 wt% CTAT solutions, the relaxation
time increases by 600%, the high frequency plateau modulus G0 increases by 37%, and the zero shear rate
viscosity 0 increases by 1600% when compared to solutions of pure CTAT. These changes in rheology
are attributed to the electrostatic interactions between the surfactant and the silica particles, namely
attractive forces between the surfactant headgroups and the surface of the silica. They describe these
interactions as resulting in the formation of bilayers where the silica particles form the center while the
surfactant headgroups form an outer layer.
Nettesheim et al. (2008) also proposed a model to describe the interaction between micelles and
nanoparticles in solution. When 30 nm diameter silica particles were added to solutions of cationic
cetyltrimethylammonium bromide (CTAB) and sodium nitrate, the relaxation time, the storage modulus,
and the zero shear rate viscosity were all observed to increase drastically. The nanoparticles are thought
to have an adsorbed layer of surfactant molecules on their surfaces with which the surfactant micelles can
interact. A micelle in solution can adsorb through its energetically unfavorable end-cap to a nanoparticle,
lowering the energy of the system by Ecap-Eads and causing the wormlike micelles to grow linearly. A
schematic of this proposed mechanism is shown in Figure 8. This interpretation is consistent with
Cryo-TEM micrographs as well as static and dynamic measurements. Helgeson et al. (2010) furthermore
described the viscoelastic behavior of solutions containing both wormlike micelles and nanoparticles as
a double network consisting of both micelle entanglements and particle junctions.

SPE-173776-MS

13

Figure 8 Schematic of the model for the association of the end cap of a worm-like micelle with the surface of a particle (adapted from
Nettesheim et al. 2008).

Oilfield researchers have capitalized on these developments over the last 5-10 years to improve the
viscosity and hence the proppant carrying capacity of VES-based hydraulic fracturing fluids. A wide
variety of colloidal particles have been developed to form links between surfactant micelles in order to
improve the viscosity of VES-based fluids. Sullivan et al. (2006) describe materials including silica,
aluminum oxide, antimony oxide, tin oxide, cerium oxide, yttrium oxide, and zirconium oxide. The
particles range in diameter from 8-250 nm. Colloidal particles ranging in size form 1 nm up to 2 m have
been used to associate or pseudo-crosslink VES micelles (Huang 2009). These particles may be
composed of any of the following materials such as zinc oxide (ZnO), berlinite (AlPO4), lithium tantalate
(LiTaO3), gallium orthophosphate (GaPO4), BaTiO3, along with a variety of other materials. These
particles are added to the viscoelastic treating fluid in the range of 0.1 to 500 pptg. Another type of
crosslinker reported for VES micelles comprises a transition metal complex (Reddy 2011). In particular,
the crosslinking agent used is in concentration of at least 0.15 wt%, is composed of zirconium triethanolamine glycolate, zirconium triethanlamine lactate, or zirconium ammonium lactate acetate, and is
combined with at least 3 wt% VES.
Several examples of pseudo-crosslinking VES micelles with nanoparticles for hydraulic fracturing
applications have been reported. For example, two VES fluids were prepared with 4 vol% VES mixed
with 13.0-ppg CaCl2/CaBr2, and one was loaded with 6-pptg of the 35 nm nanoparticles (Huang and
Crews 2008b). At 250 F, the VES fluid system with nanoparticles maintained its viscosity at 200 cp while
the VES fluid without nanoparticles was reduced from 200 cp to less than 40 cp in 80 minutes. Leak-off
tests also showed that VES fluids containing nanoparticles outperform fluids without the nanoparticles.
Base fluids with 3 wt% KCl, 4% vol VES, and 2 gptg internal breaker were prepared, and 15 pptg
nanoparticles were added to one of the fluids. Both fluids were tested at 150 F with 400-md ceramic discs
at 300 psi, and the fluid with nanoparticles showed significantly more control over fluid loss than the fluid
containing no nanoparticles.
Similarly, a fluid-loss control additive for VES-based fluids was then developed which consists of
small particles with an average diameter of less than 2 microns (Huang and Crews 2009a). These particles
associate the micelles into a stronger network and are shown to maintain the viscosity, improve proppant
suspension, and reduce fluid loss. Proppant suspension tests were performed both with and without the
fluid loss control agent. Samples were prepared that consisted of 2% VES in 13.0 ppg brine with one
pound of 20/40-mesh ceramic proppant in one gallon of liquid. To one of the samples was added 10 pptg
of the particles. For the VES fluid without any particle addition, all of the proppant settled out within 15
minutes. For the VES fluid with particle addition, there is no noticeable proppant settling even after 90

14

SPE-173776-MS

minutes. The viscosities of VES fluids with and without particle addition were determined at 250 F and
a shear rate of 100 s-1. Two fluids were prepared by combining 13.0 ppg CaCl2/CaBr2 with 2 vol% VES,
6.0 ppg stabilizer, and 1.0 gptg internal breaker, and 10 pptg of the particles were added to one of the
fluids. The fluid with the added particles maintained its viscosity at 230 cp while the fluid without
particles only maintained a viscosity of 200 cp. Finally, leak-off tests with 400 md ceramic discs at 250
F and 1,000 psi were performed. The fluids containing 5 gptg, 8 gptg, and 10 gptg particles showed
increasing degrees of fluid loss control over the base fluid.
Pseudo-crosslinked VES systems which incorporate pyroelectric (PE) nanoparticles such as ZnO were
developed and have been shown to delay crosslinking of VES micelles (Crews and Huang 2008; Huang
2009). The particles are described as developing charges on the faces of the crystal as the particles are
heated, so more association between the particles and the micelles will occur as the VES fluid is heated
by the reservoir. Studies of pyroelectric nanoparticles have been performed where base fluids were
prepared with 13.0 ppg CaCl2/CaBr2 and 2 vol% VES. Nanoparticles were added to two of the
fluids one with 15 pptg regular nanoparticles and one with 10 pptg pyroelectric nanoparticles. A fourth
fluid was prepared with 30 pptg borate-crosslinked guar in 2% KCl. The leak-off performance of these
four fluids was tested on 400 md inch thick ceramic discs at 150 F and 300 psi. The classical VES
system showed very high leak off over time because it was not wall-building. The other three systems
were wall-building and showed controlled fluid leak-off, with the best leak-off control coming from the
borate-crosslinked guar and the poorest leak-off control from the regular nanoparticles. The leak-off
control from the pyroelectric nanoparticle pseudo-crosslinked VES gel was comparable to the boratecrosslinked guar gel. The viscosity and proppant suspension characteristics of the conventional VES fluid
were compared with the VES fluid containing pyroelectric particles, and the latter was found to perform
better in both cases.
Further improvement of VES-based fluids was accomplished with the development of a system that is
thermally stable up to 275 F (Gurluk et al. 2013). Base fluids were prepared with either 2 or 4 vol%
amidoamine oxide surfactant in 14.2 ppg CaBr2 brine. Either 30 nm MgO particles or 30 nm ZnO particles
were added to these solutions. The viscosity of the 4 vol% VES solutions containing either ZnO or MgO
particles at 275 F and 10 s-1 were comparable at 100 cp, while the viscosity of the 4 vol% VES solution
without nanoparticles quickly decreased to zero cp. Solutions containing MgO and either 2 or 4 vol% VES
were also compared, the viscosity was found to increase when increasing from 2 to 4 vol% VES. An
additional observation was made of the difference between the effects of ZnO nanoparticles on 2% VES
solutions versus the effects of MgO nanoparticles on the same solution, namely that the viscosity of the
MgO solutions was higher than those containing ZnO. A summary of studies performed on pseudocrosslinked VES-based fracturing fluids is given below in Table 5.

SPE-173776-MS

Table 5Summary of experiments performed with pseudo-crosslinkers. (Rregular and PEpyroelectric nanoparticles)

15

16

SPE-173776-MS

VES technology developments suggest that polymer-based fracturing fluids may be replaced in some
regions (Crews et al. 2008a). Advances such as internal breakers and pseudo crosslinkers coupled with the
low molecular weight of VES have led to fluids that can simultaneously control leak-off, maintain high
viscosity, and controllably break and reduce viscosity in order to flow back the fluid. In a further
development, fluid loss in pseudo-crosslinked systems can be further reduced by adding 0.2 to 10 vol%
mineral oil (Huang and Crews 2009b; Huang et al. 2010a). The oil droplets collected in the filter cake over
time along with the nanoparticles and micelles, and the fluid loss continued to improve. The rheological
properties of the fluid were unchanged, which was unexpected since it was previously thought that
hydrocarbons disrupt micelles and reduce the viscosity of VES-based solutions. Pseudo-crosslinked VES
fluids have also been used to remove residual polymer from hydraulic fractures (Crews and Huang 2010).
Polymer breakers such as oxidizers or enzymes are combined with the nanoparticle pseudo-crosslinked
VES matrix during the initial fluid mixing and are believed to react more slowly than in the absence of
a VES network. The breakers are carried deeper into the fractures with the VES gel and can make better
contact with the polymer residue.
Nanoparticles have not only been used to enhance the viscosity of VES-based fluids but also dual
systems which incorporate both surfactants and polymers. For example, Fakoya and Shah (2003) studied
the following four systems to which 20 nm SiO2 nanoparticles were added: (1) 5% surfactant in 4% KCl,
(2) 33 lb/Mgal guar in 4% KCl, (3) 75 vol% surfactant and 25 vol% guar polymer and (4) 25 vol%
surfactant and 75 vol% guar polymer. Rheological data in the form of viscometry and frequency sweep
testing was collected in the temperature range of 75-175 F. The rheological properties of systems (1) and
(2) were both enhanced by particles added in concentrations of 0.24 wt% and 0.4 wt%. Dual systems (3)
and (4) were enhanced with nanoparticle concentrations up to 0.058 wt% and 0.24 wt%, respectively.
Dual purpose pseudo-crosslinking particles which improve the performance of the hydraulic fracturing
fluid but also control formation fines migration have been described (Huang and Clark 2013). Nanoparticles are added to the VES-based fluid along with an internal breaker. As the fluid system is pumped into
the formation to create fractures, a filtercake is developed on the fracture face by the pseudo-crosslinked
micelle networks. As the internal breakers are released, the worm-like micelles are collapsed into spherical
micelles. The nanoparticles are released and precipitate, attaching to nearby proppants and thereby
capturing formation fines. A successful field treatment was performed in the Gulf of Mexico utilizing
similar nanoparticle technology for fines control (Huang et al. 2010b).
As previously described, particles composed of ZnO, MgO, TiO2, or Al2O3 enhance the viscosity of
VES-based fluids. It has been shown, however, that particles composed of certain other materials have the
opposite effect upon these gels. For example, inorganic semiconductor particles such as cupric oxide,
cuprous oxide, silicon, silicon carbide, germanium, gallium arsenide, indium antimonide, and gallium
nitride are found to reduce the viscosity of the gelled aqueous fluid (Huang 2014). The reaction is
described as being either transition-metal-catalyzed or transition-metal-mediated, although the exact
mechanism is not reported. Organic semiconductors such as pentacene, anthracene, rubrene, etc. are also
observed to have similar effects upon VES gels as the inorganic semiconductors. Similarly, metal ions
present in the aqueous treating fluid are found to break, reduce, and/or digest the VES within the aqueous
treating fluid (Crews and Huang 2013). Addition of nanoparticles composed of materials that pseudocrosslink VES micelles along with a complexing agent such as ethylenediaminetetraacetic acid (EDTA)
was reported to prevent redox reactions between VES molecules and the metals ions in the fluid.
A.3. Internal Breakers
Hydraulic fracturing fluids rely on a breaking mechanism in order to reduce the viscosity of the carrier
fluid and flow back the well. Polymer-based systems utilize enzymes, oxidizers, acids or, more recently,
decrosslinking agents to break the fluid and reduce the viscosity. Initial VES-based fluids took advantage
of two different reservoir conditions in order to break the fluid (Huang and Crews 2008b). The first

SPE-173776-MS

17

condition is a change in brine concentration. VES-based fluids are seen to be stable over a particular
concentration of salts and by diluting this concentration with produced fluids, the gel can be broken. The
second condition that will break VES-based fluids is their contact with reservoir hydrocarbons. When
hydrophobic substances such as oil or gas dissolve in the hydrocarbon core of the micelle the structure
swells and breaks into smaller spherical micelles. This loss of larger worm-like micelles results in the
viscosity of the solution being reduced. Since hydrocarbons from the formation have this effect upon the
micelles, an internal breaker was not utilized to reduce the viscosity in order to flow back the fluid.
However, depending upon these reservoir conditions is unreliable and, in some cases, additional treatments had to be performed in order to clean up the formation. Data indicates that 20% of the treatments
performed by these VES-based fluids in the 1990s required remedial actions (Crews et al. 2008;
Al-Muntasheri 2014). Hence, the need was established to develop internal breakers that would administer
a controlled break of the VES fluids in a manner that is comparable to conventional polymer-based
systems. However, in this case the breaking would not deposit residue into the formation and proppant
pack because of the solids-free characteristic of the VES.
A number of methods for controlling the viscosity of VES-based fluids have been observed. First, a
method for controlling or delaying the onset of gelation in VES-based fluids after the fluid has been mixed
has been seen (Hughes et al. 1999). Three different mechanisms can be used to control the fluid viscosity:
delayed release of a specific counter-ion, change in hydrogen bonding, or a modification of the solutions
ionic composition. Another method describes an intervention at the surface which can reversibly break the
viscosity of VES solutions that are used in a drilling application (Rose et al. 1988). When the fluid is
pumped into the well, it is viscous enough to carry cuttings to the surface, and the techniques developed
for breaking the fluid can be employed at the surface for solids removal. Some of these methods include
changing the temperature of the fluid, contacting the fluid with hydrocarbon, and adjusting the pH. These
treatments are reversible, and viscosity can be restored to the fluid. A third example of a system that can
be used to control the viscosity of the VES-based fluid involves the controlled addition of components that
decrease the viscosity (Nelson et al. 2005). The breaker can be either internal or external such as
precursors which release a component such as alcohol, salt or organic acid by one of the following
mechanisms: slow dissolution, melting, reacting with a compound that is present in the fluid or added to
the fluid, or breaking a coating.
Laboratory studies have shown that VES-based fluids can be broken by internal phase breaker
technology. The aqueous breaker solutions have shown controlled viscosity reduction in the temperature
range of 80 F to 225 F (Crews 2005). The break times observed for these fluids were comparable to
crosslinked-polymer hydraulic fracturing fluid systems, including times as short as 15 minutes. The
breakers were also functional over a wide range of salinities such as 3 wt% KCl, 12 wt% KCl, and ASTM
synthetic seawater, with increasing amounts of breaker required. A second study was performed which
compared the laboratory results of VES fluids with and without internal breaker (Crews and Huang 2007).
Berea core cleanup tests showed that little pressure or time is needed to initiate cleanup when the internal
breaker is used. By contrast, the unbroken VES fluid requires five times higher pressure than just the
partially broken VES fluid. Likewise, nitrogen gas core cleanup tests were performed where cores were
shut-in (left static) for 24 hours then displaced with nitrogen gas over a 48 hour period. The results showed
that the VES fluid that was internally broken was readily producible with gas, but the unbroken
VES-based fluid was very difficult to produce. Another laboratory study incorporating an internal breaker
was performed (Crews et al. 2008b). By varying the concentration of the internal breaker, complete
viscosity reduction can be achieved at 250 F. Core-regain permeability tests were performed where
VES-based fluid without any internal breaker showed just 2% regain permeability after 48 hours of low
pressure gas flow. The VES fluid with internal breaker, however, had a regain permeability of 141%.
The specific mechanisms that have been utilized for breaking VES-based gels have been described. A
microemulsion system has been developed which performs a variety of functions in addition to the

18

SPE-173776-MS

viscosity reduction of VES-based fluids, namely solubilization of by-products generated from breaking
and desorption/water-wetting to prevent VES residue from depositing in the formation (Crews 2010a). An
example of one such microemulsion system includes unsaturated fatty acid oil, glycol, a sorbitan
ester/ethoxylated sorbitan ester mixture, and an alkyl sulfonate. VES-based fluids can also undergo
viscosity reduction by a metal ion source and, optionally, either a second metal ion source, reducing agent,
or a chelating agent (Crews 2009). The reason that the fluids viscosity is reduced is described as one of
the following mechanisms: rearrangement of the micelle structure, deaggregating the VES micelle
structure, chemically altering the VES molecules, or a combination of these mechanisms. These potential
breaking mechanisms are also at play in systems that are broken by unsaturated fatty acids such as
monoenoic acid and/or polyenoic acid (Crews 2010b). For example, an amine oxide surfactant can be
broken with an oil that contains a large amount of unsaturated fatty acids such as soybean oil, fish oil, or
flax oil. The unsaturated fatty acids are described as auto-oxidizing into ketones, aldehydes, and saturated
fatty acids that break the VES fluid. Similarly, addition of specific types of bacteria such as Enterobacter
cloacae, Pseudomonas fluorescens, Pseudomonas aeruginosa have been observed to reduce the viscosity
of amine oxide surfactants such as TAPAO (Crews 2006). The mechanism for breaking may follow one
of two pathways: the VES micellar structure is directly rearranged or disaggregated or, alternatively, other
materials in the viscosified fluid may be degraded to form by-products that reduce the viscosity of the gel.
A.4. Foamed & Emulsified VES
Due to the desire to reduce the amount of water used in fracturing fluids, systems which incorporate both
viscoelastic surfactant and carbon dioxide to form a fracturing fluid have been developed in recent years.
These VES-CO2 fluid systems combine the benefits of VES fluids such as good proppant transport, low
formation damage, and low friction pressures with the enhanced cleanup and better hydrostatic pressure
of carbon dioxide. Historically, aqueous/CO2 mixtures have been characterized as foams. However, CO2
is pumped as a liquid at surface conditions, so it may be more appropriate to refer to these systems as
emulsions (Chen et al. 2005). The VES fluid-CO2 mixture can be thought of as two liquids with limited
miscibility but dispersed stably in each other (an emulsion). Since the two components remain separate
phases, the CO2 will not disrupt the worm-like micelles and hence the viscoelasticity. As shown by Chen
et al., foaming the VES with 70% CO2 roughly doubles the viscosity over that of 4% VES straight fluid.
The use of this system for field applications has been reported (Chen et al. 2005; Arias et al. 2008;
Al-Muntasheri 2014). Zwitterionic surfactants such as the betaines shown in Figure 7(b) and 7(c) and
carbon dioxide in a separate phase along with a cosurfactant such as C12 alkyl dimethyl benzyl ammonium
chloride have been described (Chen et al. 2007 and 2010).
There are several published case histories of wells that were fracture stimulated with VES-based
foamed fluids. VES foam is a significant unconventional fracturing fluid for tight gas reservoirs (Gupta
2009). Combination of an anionic surfactant such as sodium xylene sulfonate with a cationic surfactant
such as N,N,N-trimethyl-1-octadecammonium chloride in ratios of 1:4 to 4:1 will form a viscoelastic gel
that is combined with 53% to 96% or more of carbon dioxide by volume (Zhang et al. 2002). Foamed
surfactant gel treatments of this nature were used in 75 producing formations (3,100 treatments)
between 1998 and 2005 in the Western Canadian Sedimentary Basin (Gupta et al. 2005). Both CO2 and
N2 were used as the internal phase for the treatments, with N2 being used about 90% of the time. Proppant
concentrations over 800 kg/m3 (6.7 ppg) were used in about 35% of the wells (1,100 fracture stimulations), where some jobs placed up to 1,200 kg/m3 (10 ppg). In about 10% of the work, proppant
concentrations lower than 600 kg/m3 (5.0 ppg) were used because of either lower surfactant concentrations or the specific reservoir characteristics. Less than 20% of the CO2 foamed surfactant gel treatments
result in screen-outs, which is similar to non-energized surfactant gelled fluids or non-energized linear
guar gelled fluids. By comparison, surfactant gels foamed with nitrogen screen out 15% of the time. These

SPE-173776-MS

19

numbers are comparable to foamed conventional cross-linked gelled treatments which screen out 18% of
the time (Gupta et al. 2005).
Additional field cases describing the application of VES-CO2 based fluids have been reported (Table
6). VES-CO2 systems were used to improve the Olmos production in the Caterina SW field in Texas
(Semmelbeck et al. 2006). One test well was pumped with VES-CO2 then ten additional wells were
stimulated in this area. The VES-CO2 (70% quality) was pumped at 6 to 12 bbl/min, typically with 25,000
lbm of proppant (1 to 5 lbm proppant added per gallon of fluid). The propped fracture half-lengths
averaged 425 ft, with minimal height growth. The outcome from 11 treatments resulted in an average
production of 430 Mscf/D and 82 BOPD in zones that were previously bypassed. In another case, a
VES-CO2 based fluid was applied in the Frontier Formation, Big Horn Basin, Wyoming where two wells
have been reported utilizing fracturing fluids with high foam quality (Bustos et al. 2007). The fracturing
treatment in each well included the following: 2,000 gallons of a 10% surfactant solution, followed by a
PAD of 3% VES-CO2 at 70% foam quality and then 2.5% VES-CO2 as carrier fluid for all 20/40 Jordan
sand stages from 1 to 5 PPA, flushed with 1% VES-CO2. The total amount of proppant in the two
treatments was 61,000 lbs and 63,000 lbs, with an injection rate of 30 BPM. P3D (Pseudo threedimensional) fracture simulation demonstrated a fracture half-length of about 338 ft with an average
conductivity of 587 md-ft for the first well and a fracture half-length of about 511 ft with an average
conductivity of 553 md-ft for the second well. Furthermore, wells in Waltman field in Wyoming provide
another example of VES-CO2 fluid systems (Arias et al. 2008). Four of the wells that were treated with
VES-CO2 fluid (70% quality) were compared with nearby wells treated with linear hydroxypropyl guar
(LHG) polymer system with gel loading of 40 lbm/mgal. Initial production from the wells treated with
VES-CO2 was observed in the range of 5 to 7 MMcf/D, which is higher than the gas rates of 2 MMcf/D
seen from wells treated with the polymer-based fluid. Estimated fracture lengths for the two systems are
estimated to be similar because of the similarities in the viscosities and hence the proppant transport
characteristics. The difference in production is attributed to the clean nature of the VES-CO2 and its
minimal proppant pack damage versus the LHG polymer system. A final example of the application of
VES-CO2 in the field was in the Morrow Sands in Southeast New Mexico (Pandey et al. 2007). Three
Morrow completions covering eight stages were successfully pumped with VES-CO2. The fluid system
consisted of 4.5 vol% VES (for the pad; less for later stages) along with 70% liquid CO2 as the fracturing
fluid. In the first of the wells, 20/40 US Mesh size ceramic proppant was used in the treatments at 27,000
lbs (1 to 2.5 ppa). The pseudo 3D fracture simulator determined an average fracture length of 605 ft for
the four stages along with a propped fracture width of 0.05 inch and conductivity of 1385 md-ft.

20

SPE-173776-MS

Table 6 A summary of the published case histories of wells that were fracture stimulated with VES-based foam fluids. The foam
quality in each treatment is 70%.

A nitrogen-foamed viscoelastic surfactant-based system for use in unconventional natural gas wells has
been developed that is a coal/carboniferous shale-compatible solids-free (CCSF) fluid (Fredd et al. 2004).
The CCSF fluid showed greater than 70% retained permeability and cleanup factors less than 100
psi-min-K with formation coal samples from seven different basins in North America. By comparison,
conventional slickwater and polymer-based fracturing fluids exhibited significant coal pack damage with
retained permeabilities as low as 46% and 18%, respectively. Furthermore, the CCSF fluid exhibited a
59% one-year cumulative production increase relative to conventional polymer-based fluids in the
Devonian Shale formation.
A.5. Brines & Produced Water
VES fluids have been observed to be compatible with high-density brines which make them attractive
candidates for a variety of applications. As seen in Table 5, VES fluid formulations commonly contain
high salt content. An example of a high density fluid that is suitable for a broad range of applications
contains erucylamidopropyl betaine or oleylamidopropyl betaine, an alcohol such as methanol, and at least
12.5 ppg of a salt or mixture of salts of divalent cations such as chlorides or bromides of calcium and zinc
(Fu et al. 2006). The structures of these zwitterionic surfactants are shown in Figure 7(b) and 7(c). The
same surfactant in high-density brine fluid has been further applied as a perforation fluid (Samuel et al.
2011). Similar chemistry is described in the development of a fluid loss or lost-circulation-control pill
(Samuel et al. 2007). These particular surfactants, erucylamidopropyl betaine and oleylamidopropyl
betaine, have furthermore been combined with arylalkylsulfonate cosurfactant, and a polar solvent such
as water, alcohol, or glycol (Berger and Berger 2006). Viscosities are reported from 400-900 cp over the
temperature range of 20-100 C for surfactant solutions containing 30 wt% CaCl2. These fluids are
described as having applications in fracturing, acidizing, gravel packing, and other similar operations. In
a second example by Berger and Berger (2008), an amphoteric alkyl amido betaine surfactant is used to
viscosify an injection brine and reduce the interfacial tension between water and oil. The aqueous fluid
is pumped into an injection well, displacing hydrocarbons into the production well.

SPE-173776-MS

21

The development and first field application of these VES fluids for lost circulation applications have
been described (Samuel et al. 2003). Typical fluid-loss control systems are composed of high concentrations of polymers which are viscous and form a filter cake on the face of the rock. However, the
difficulty in cleaning up these systems led to the development of a solids-free fluid-loss pill. This
VES-based system is compatible with more completion brines that are used in well completions and was
demonstrated in the laboratory to be stable up to 375 F. Around 10-20% of zwitterionic viscoelastic
surfactant is mixed with heavy brine such as 12.5 ppg CaBr2. At high temperatures, around 5% methanol
is also added in order to stabilize the system. A second application of brine-based VES fluids includes
nonionic surfactants for frac packing (McElfresh et al. 2003). In the Adriatic Sea and in the Gulf of
Mexico, non-ionic VES gels based on brines up to 10.5 ppg CaCl2 were used in over 30 treatments.
Newer methods have also been developed to stabilize VES fluids in high-density brines. For example,
low molecular weight surfactant polymers can be used to control the curvature of surfactant micelles (van
Zanten 2011). Tethered polymers, which are nonionic surfactants, can be added to cationic/anionic
VES-based systems in order to stabilize the brine solutions and maintain the viscosity. Three different
CmEn alkyl poly(ethoxylate) nonionic surfactants (E10, 20, and 100) were examined for their effects on
five different cationic/anionic VES systems. The tethered polymer was observed to increase the viscosity
of the cationic/anionic VES systems in a variety of brine concentrations as well as prevent the precipitation of surfactants at high brine densities. Further experimental work on similar tethered polymer
systems was performed on cationic alkyl quaternary ammonium salt and cationic/anionic alkyl amine/
alkyl sulfate salt (van Zanten and Ezzat 2011). For the latter system, maintaining a viscosity of 1000 cp
in high-density brine had a temperature limit of 225 F.
Viscoelastic surfactants such as N,N,N, trimethyl-1-octadecammonium chloride and sodium xylene
sulfonate can also be used in combination with produced water (Gupta and Tudor 2005). A Case History
has been published from the Western Canadian Sedimentary Basin where 50 individual wells were treated
with VES-flowback water (Gupta and Hlidek 2010). It is preferred to use recycled water in which VES
was used in the first treatment. The concentration of chemicals needed for the second treatment is lower
than the first treatment, with the loading reduced from 210 to 120L for a typical job. It has been observed
that the anionic component of the VES fluid often remains in the flowback water while the cationic
component is believed to be adsorbed on the clay components of the formation. A cost savings of 12%
was realized from the project, and the well production performance was unaffected.
B. Matrix Acidizing
Matrix acidizing is extensively used to enhance the productivity and injectivity of wells drilled in
carbonate reservoirs. In matrix acidizing, acids are injected at pressures less than the fracturing pressure
of the formation in order to dissolve part of the rock, remove the damage and create open flow paths for
the reservoir fluids to flow through. Hydrochloric acid (HCl) is the main fluid used in matrix acidizing
both because of its inexpensive nature and because the reaction products are soluble in water and cause
no subsequent damage to the formation. The main issues with HCl are its incompatibility with some crude
oils and its fast reaction with rocks at temperatures above 200F. These short reaction times will result in
acid spending, face dissolution and failure of the treatment. Another issue with HCl is its inability to
stimulate heterogeneous zones with large changes in rock permeability. HCl will invade the high
permeability zone and little acid will be diverted to the low permeability zones and zones with high
damage. This problem is more pronounced in acidizing long horizontal wells. Acid diversion techniques
are necessary in such cases to guarantee the proper distribution of acid and to obtain better acidizing
results.
Several approaches can be employed to enhance the acid diversion to assure better acid distribution.
Injection rates can be increased during the treatment to increase the injection pressure and help to divert
acid in the low permeability regions (Alleman et al 2003). Acid diversion can also be accomplished using

22

SPE-173776-MS

mechanical diverters such as coiled tubing, ball sealers, rock salts and acid flakes, and using plugs and
packers for isolating different zones (Economides and Nolte 1989). The main disadvantage of these
techniques is their inability to work in large permeability contrast formations (Lynn and Nasr-El-Din
2001). Chemical diverting agents are also available for acid diversion. Use of foamed fluids like foamed
KCl solutions, foamed ammonium chloride solutions and gelled pills were reported by Zerhbouh 1993;
Zeilinger et al. 1995; Taylor and Nasr-El-Din 2001. The main limitation of these foamed and gelled
systems is their instability at temperatures above 200F and their ineffectiveness in plugging formations
with permeabilities above 500 md (Alleman et al. 2003). Gelled acids and cross-linked gelled acids are
used as chemical diverters to guarantee the proper distribution of acid in heterogeneous reservoirs (Gomaa
et al. 2010).
Viscoelastic surfactants have also been developed and applied for acid diversion (Chang et al. 2001a;
2001b). As mentioned before, viscoelastic surfactant based fluids develop their viscous nature through the
formation of micellar structures. Also, the pH value will control the rate of building the viscosity of
VES-based acid systems. At low pH values, the viscosity of VES-based acids is very low allowing the acid
system to flow and penetrate into the formation. Upon the reaction of acid with carbonates, the pH
increases and the concentration of divalent cations in solution increases and therefore the VES-based acid
system starts to build viscosity. At pH around 4, the viscosity of the spent acid will be high enough to
divert the fresh acid to low permeability un-contacted zones and fresh acid starts to penetrate and react
with these rocks to form wormholes (Nasr-El-Din et al. 2006b; Crews and Huang 2007; Crews et al. 2008;
Huang et al. 2008b; Yu et al. 2011). The viscosity of VES fluids can be reduced upon mixing with
hydrocarbons. This may make it unfavorable when used with dry gas wells. Surfactant based acids were
introduced in the petroleum industry by Chang et al. (2002) and Qu et al. (2002).
B.1. Experimental Studies
Alleman et al. (2003) developed a VES diverting agent with a vesicle structure type. The VES was stable
at temperatures up to 250F and also up to 350F by adding a material referred to as an intensifier to
interact with VES molecules and enhance the charge on the micelles. Adding this intensifier in quantities
of 0.2 to 0.3 wt% made the structure of the VES larger, stronger and more stable at temperatures up to
350F. At a temperature of 250F, the viscosity of VES diverting agents with 0.2 polyquat at 100 s-1shear
rate was 130 cp, while the viscosity increased to 330 cp when the polyquat concentration was increased
to 0.4 wt%.
Al-Ghamdi et al. (2004) studied the effect of different acid additives on the rheology of VES-based
acid systems. The acid system was 15 wt% HCl with 6.0 vol% VES. Additives like corrosion inhibitor,
iron, nonionic surfactant, anti-sludge agents and hydrogen sulfide scavengers were tested. They found
that, at 100C and 87 s-1 shear rate, the viscosity of spent VES based acid was increased from almost zero
to 130 cp by increasing the VES concentration from 1 to 6 vol%. Also, they noticed that addition of a
mutual solvent and iron (III) with concentrations higher than 1000 ppm caused reduction in the apparent
viscosity of the VES based acid systems. This finding is important since mutual solvent as well as
hydrocarbon fluids can be used to break down the remaining VES gel in the formation for better cleanup
and better performance. Care should be taken to prevent contact of the VES-based system with mutual
solvent during injection of the treatment in the well. Also, corrosion product addition needs to be
controlled to prevent effects upon the rheology of the VES-based acid systems which in turn will affect
the performance of the diversion system.
Nasr-El-Din et al. (2006a) studied the rheology and diversion ability of VES based acid systems. Table
7 shows the relationship between the VES concentration and the apparent viscosity of the formulation
measured at 40 s-1 and at 200F and 250F temperature. It is clear that the apparent viscosity increased
when increasing the VES concentration. A parallel core flow test was performed using 711 permeability
ration contrast core samples to study the acid diversion in such core samples at 200F. They noticed that

SPE-173776-MS

23

acid breakthrough occurred first in the low permeability core sample where one main wormhole was noted
by using the CT scanner. This occurred as a direct result of the effective diversion of acid that was
achieved by the VES formulation.

Table 7Rheology data measured by Nasr-El-Din et al. (2006a)


VES CONC. (vol%)
0.5
1.0
1.5
3.0

VISCOSITY Water, 200 F (cp)

VISCOSITY Water, 250 F (cp)

VISCOSITY CaCl2,250 F (cp)

138
152
433

113
152
312

118
141
160
-

Lungwitz et al. (2007) experimentally studied the use of VES based fluids as diverting systems in
acidizing treatments. They performed coreflood experiments using carbonate core samples with permeability in the range of 0.1 to 50 md at a temperature of 200 to 240F using 15 wt% HCl acid systems. Fluid
loss was examined using limestone samples of permeability in the range of 1 to 3 md at a temperature of
150F using 2 wt% KCl brines. It was found that both cross-linked based fluids and VES based fluids
showed similar leak off properties as well as the same initial viscosity (30 cp at 170 s-1 and 70F) Using
VES based acid systems, acid breakthrough was noticed in limestone and dolomite core samples after
injection of about 1 to 1.6 pore volumes of acid, while using cross-linked polymer acid systems did not
achieve breakthrough when tested under similar conditions. Also, VES based acid systems were compared
to 15 wt% plain HCl acids using the conductivity cells. The VES based HCl system created etched
surfaces which enhanced the conductivity of the fractures, while HCl alone caused face dissolution and
fewer enhancements in fracture conductivity.
Huang et al. (2008a) examined the use of VES-organic acid systems for acid treatments in carbonate
formations. The viscosity of the fresh mixture of 2.0 vol% VES 10 wt% organic acid 0.2 wt% internal
breaker at 100 s-1 shear rate and pH of 3 was almost zero. Upon acid spending, the pH increased to around
6, and the viscosity of the spent acid increased to 210 cp at 100 s-1. Also, they found that the use of
internal breaker is very important to break down the remaining VES structure to better clean up the
formation after the acid treatment.
Yu et al. (2011) studied the retention of VES surfactant in the porous media through using coreflood
experiments. The VES-based acid system used in their study consisted of 15 wt% HCl, 7 vol% VES and
0.3 vol% corrosion inhibitor. The acid injection rate was found to affect on the volume of acid
breakthrough and the wt% of VES retained inside the core sample. Figure 9 shows a relationship between
volume to breakthrough and the retained surfactant (wt%) as a function of the acid injection rate. As the
acid injection rate increased both the volume to breakthrough and the retained surfactant in core sample
decreased. They furthermore described that VES can be removed by the hydrocarbon fluids produced after
the treatment such that there is no need for a cleanup treatment to remove the remaining gel. In the event
that the cleanup after the treatment is not complete, they recommended the use of an internal breaker
system or a mutual solvent as a post flush to break down the remaining VES.

24

SPE-173776-MS

Figure 9 Change of volume of acid to breakthrough and retained surfactant (wt%) as a function of acid injection rate.

Wang et al. (2012) studied the rheology of VES based system designed for high temperature
applications. They found that the viscosity of the spent acid (at a pH value around 4.5) was around 200
cp at 325F for 4 vol% VES-based fluid at 10 s-1 shear rate. This will help the acid plug the high
permeability zones upon acid spending and divert the fresh acid to the low permeability zones.
Yu et al. (2012) studied the effect of hydrolysis on the rheology of VES-based acids systems. Under
high temperature and in acidic environment, the CO-NH bond (peptide bond) is broken and the system
is hydrolyzed. This bond breakage resulted in the formation of smaller molecular weight molecules which
were responsible for reducing the viscosity of the VES-based acids systems. At temperatures of 190F, an
increase in the apparent viscosity of the VES based acid system was observed for hydrolysis times less
than 2 hours. For hydrolysis times greater than 3 hours, the acid breaking rate is very high and phase
separation was noticed. This resulted in a large decrease in the acid viscosity. Yu et al. (2012) shows the
VES-based acid phase separation as a function of hydrolysis time at 190F for 4, 6, and 8 vol% VES
concentration (Figures 3-5 of the manuscript). Similar behavior was noted at each surfactant concentration. He et al. (2013) studied the effect of the hydrolysis of VES-based acids on the rheology and the
performance of such systems at 250F. As the degree of hydrolysis increases, more breaking takes place
and less damage occurred in the high permeability core samples. Figure 10 shows the effect of hydrolysis
on the apparent viscosity of a VES solution prepared using 4 vol% VES and hydrolyzed at 190F. The
results of these studies highlight the importance of proper design of VES-based acid systems in order to
prevent breaking the acid system before acid treatment is complete.

SPE-173776-MS

25

Figure 10 Effect of hydrolysis time and shear rate on the VES-based acid systems formulated using 4 vol% VES at 190F (He et al.
2013).

B.2. Case Histories


VES-based diverter was applied in several field trials, and a summary of the reported trials is provided
in Table 8. One of these cases was described by Alleman et al. (2003). The temperature of the well was
198 F, and the well depth was 14500 ft including an oil bearing formation that was 53 ft. The VES
diverter pill was mixed with 10 wt% HCl and 5 wt% acetic acid, and the treatment took place in stages.
The injection pressure increased from around 300 psi to 1,400 psi upon injection of the diverting VES pill
as a direct result of its high viscosity and its ability to plug the high permeability zones. In the final stage
of injection, the pressure increased by 200 psi upon the injection of the VES diverter pill. A second case
study was also reported by Alleman et al. (2003) with similar results.

Table 8 Summary of field trials performed with VES-based acid systems. *The bbl/day rate for this well is the injection rate.
LOCATION

PRODUCTION
(bbl/day)

WELL INFO

VES (vol%)

HCl (wt%)

Gulf of Mexico

198 F, 14,500 ft

10 5 Acetic

Gulf of Mexico

150 F, 9,600 ft

10 5 Acetic

Mainly calcite

20

Above expected
rates
Above operator
expectations
6,000

Water Injector*
Horizontal
Tubing: 6,270 ft
Vertical
-

20

75,000 (450% inc.)

20

(150% inc.)

20

3,700

200 F, 9,310 ft
3,270 psi
Mainly dolomite

20

15-20

5-500 md

20

80 (400% inc.) 2X
acid alone
800 (150% inc.)

Saudi Arabia
Saudi Arabia
Saudi Arabia
Saudi Arabia
Niagaran Reef (MI)
Bombay High
Oilfield, India

REFERENCE
Alleman et al. 2003
Alleman et al. 2003
Mohammed et al.
2005
Nasr-El-Din et al.
2006a
Nasr-El-Din et al.
2006b
Nasr-El-Din et al.
2006b
Nasr-El-Din et al.
2006b
Kreh 2009
Rawat et al. 2014

26

SPE-173776-MS

Another successful treatment using VES-based acids was described by Mohammed et al. (2005). The
treatment was performed on a vertical oil-producing well where coiled tubing was used to inject the acid
to increase the efficiency of diversion in the wellbore section. The treatment included several steps: (1)
a preflush of 6% mutual solvent in diesel, (2) 20 wt% HCl with 0.4 vol% corrosion inhibitor and 0.4 vol%
H2S scavenger, (3) nitrified 20 wt% HCl with the same amount of corrosion inhibitor and H2S scavenger
with nitrogen at a ratio of 400 Scf/bbl, (4) 6 vol% VES, 20 wt% HCl, and 0.4 wt% corrosion inhibitor with
nitrogen at a ratio of 1300 Scf/bbl, and (5) a post-flush composed of 5% mutual solvent in diesel in order
to break the worm-like micelles from the VES treatment. A cationic surfactant was used and upon reaction
of the acid with carbonate, the pH increased from 0 to 2 and the viscosity started to build. The oil
production rate increased to 6,000 bbl/day after the treatment and no increase in the water cut for nine
months of production after the treatment.
Nasr-El-Din et al. (2006a) described another successful field treatment using VES-based acids. The
treatment was performed on an open-hole horizontal well designed to handle an injection rate of 80,000
bbl/day. Based on well testing analysis, the well was found to be severely damaged with a skin factor
ranging from 15 to 24. A VES-based system was used for acid diversion. The treatment involved a
multistage injection of 20 wt% HCl (where the acid system will dissolve the damage resulting from
drilling the well) and a diversion system to divert acid to the new location. The diversion system consisted
of 65% quality nitrogen foam solution of 3 vol% VES and water. Analysis of the well after the VES
acidizing treatment indicated that the treatment was successful where the injection rate increased to 75,000
bbl/day (from 17,000 bbl/day), and the injection pressure decreased to 1150 psi (from 1643 psi).
Nasr-El-Din et al. (2006b) furthermore described more than 60 field treatments performed using VESbased acid systems to stimulate wells drilled in carbonate reservoirs in Saudi Arabia. Approximately 95%
of these treatments were successful, and three case studies have been described where VES-based acid
was used to enhance the productivity of these wells.
Kreh (2009) studied the use of VES-based acid systems to stimulate wells drilled in the Niagaran reef
reservoir. Both the northern and southern trend parts of these reefs are heterogeneous carbonate, with
dolomite being the dominant lithology. Porosity falls in the range from 3.8 up to 19.6% in the northern
trend and from 2 to 14% in the southern trend. The reservoir thickness can range from 300 to 700 ft. Since
the lithology was mainly dolomite, it is preferred to use regular 15-28 wt% HCl. But as a direct result of
the heterogeneity in this reservoir, these regular acid treatments were alternated with 15-20 wt% HCl
VES-based treatment. This was done in many wells and it was proven as a successful treatment where
VES-based acid system was able to divert the regular acid to untreated zones. In one of the wells, the
VES-based acid treatment was able to achieve a 400% increase in the oil production rates compared to
100% increases in oil production rates achieved by using regular acid systems only.
Rawat et al. (2014) presented a case study in Bombay High Oilfield which is one of the major oilfields
in India. The reservoir rock is carbonate rocks with permeability in the range from 5 to 500 md. The
reservoir is heterogeneous and it possess several forms of porosity like vuggs, channels and molds besides
it has low thickness and all of this added to the low recovery factor. The field production rate decreased
from 130,000 BOPD in 1989 to 63,000 BOPD in 1993 and it has been fluctuating since that. VES-based
acid treatment was designed to treat one of the wells producing for this field. The VES-based acid
treatment was 15 wt% HCl and 3 vol% VES. The VES-based treatment was performed on multistages
with gelation time around 15 to 20 minutes to allow for acid spending and gelation of the VES-based fluid.
Then a retarding acid system (emulsified acid and gelled acid) was injected to stimulate the low
permeability zones. The injection pressure during the injection of the VES-based treatment increased from
300 psi to over 600 psi during the injection of the last VES-based fluid stage which indicates how this fluid
was able to build viscosity. Also, post treatment performance indicated that the oil production rate
increased from 550 BOPD to about 800 BOPD after the acid treatment.

SPE-173776-MS

27

Conclusions
Viscoelastic surfactants have been developed for upstream oil and gas applications, particularly hydraulic
fracturing and matrix acidizing. The chemical nature of the base surfactants spans a wide range of
surfactant types. The studies reported have primarily focused on the rheological properties of these
systems and the effects of adding organic or inorganic salts, nanoparticle pseudo-crosslinkers, or
components which can reduce or break the viscosity. Other characterization techniques for surfactant
micelle systems such as Cryo-TEM, SLS, and SANS have seen very limited use in our industry and may
provide an area of future exploration to develop a deeper understanding of structure-property relationships. In order for VES-based fluids to play a more prominent role in the field, inherent limitations such
as cost, applicable temperature range, and leak-off characteristics will need to continue to be addressed.
If we can efficiently and economically overcome these issues, VES-based fluids offer the industry an
excellent clean, non-damaging alternative to conventional polymer-based fluids.

References
Al-Ghamdi, A.H., Nasr-El-Din, H.A., Al-Qahtani, A.A., Samuel, M.M. 2004. Impact of Acid
Additives on the Rheological Properties of Viscoelastic Surfactants and Their Influence on Field
Application. Presented at the SPE/DOE Symposium on Improved Oil Recovery, Tulsa, Oklahoma,
17-21 April. SPE-89418-MS. http://dx.doi.org/10.2118/89418-MS.
Al-Muntasheri, G.A. 2014. A Critical Review of Hydraulic Fracturing Fluids over the Last Decade.
SPE Production Operations 29(4): 243260. SPE-169552-PA. http://dx.doi.org/10.2118/169552PA.
Alleman, D., Qu, Q., and Keck, R. 2003. The Development and Successful Field Use of Viscoelastic
Surfactant-based Diverting Agents for Acid Stimulation. Presented at the SPE International
Symposium on Oilfield Chemistry held on Houston, Texas, USA, 5-7 February. SPE-80222-MS.
http://dx.doi.org/10.2118/80222-MS.
Arias, R.E., Nadezhdin, S.V., Hughes, K., Santos, N. 2008. New Viscoelastic Surfactant Fracturing
Fluids Now Compatible with CO2 Drastically Improve Gas Production in Rockies. Presented at the
SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA, 13-15 February. SPE-111431-MS. http://dx.doi.org/10.2118/111431-MS.
Bandyopadhyay, R. and Sood, A.K. 2005. Effect of Silica Colloids on the Rheology of Viscoelastic
Gels Formed by the Surfactant Cetyl Trimethylammonium Tosylate. J. Colloid Interface Sci. 283
: 585591. DOI: 10.1016/j.jcis.2004.09.038.
Berger, P.D. and Berger, C.H. 2006. Viscoelastic Surfactant Mixtures. US Patent Application
2006/0084579.
Berger, P.D. and Berger, C.H. 2008. Process for Oil Recovery Employing Surfactant Gels. US Patent
7,373,977.
Bernheim-Groswasser, A., Wachtel, E., and Talmon, Y. 2000. Micellar Growth, Network Formation,
and Criticality in Aqueous Solutions of the Nonionic Surfactant C12E5. Langmuir 16(9): 4131
4140. DOI: 10.1021/la991231q.
Brown, J.E., Card, R.J., and Nelson, E.B. 1999. Methods and Compositions for Testing Subterranean
Formations. US Patent 5,964,295.
Brown, J.E., King, L. R., Nelson, E.B., Ali, S.A. 1996. Use of a Viscoelastic Carrier Fluid in
Frack-Pack Applications. Presented at the SPE Formation Damage Symposium, Lafayette, Louisiana, USA, 14-15 February. SPE-31114-MS. http://dx.doi.org/10.2118/31114-MS.

28

SPE-173776-MS

Bustos, O., Chen, Y., Stewart, M., Heiken, K., Bui, T., Mueller, P., Lipinski, E. 2007. Case Study:
Application of a Viscoelastic Surfactant-Based CO2-Compatible Fracturing Fluid in the Frontier
Formation, Big Horn Basin, Wyoming. Presented at the SPE Rocky Mountain Oil & Gas
Technology Symposium, Denver, Colorado, USA, 16-18 April. SPE-107966-MS. http://dx.doi.org/10.2118/107966-MS.
Cates, M.E. 1990. Nonlinear Viscoelasticity of Wormlike Micelles (and Other Reversibly Breakable
Polymers). J. Phys. Chem. 94: 371375. DOI: 10.1021/j100364a063.
Cates, M.E. and Candau, S.J. 1990. Statics and Dynamics of Worm-Like Surfactant Micelles. J. Phys.:
Condens. Matter 2: 6869 6892. DOI: 10.1088/0953-8984/2/33/001.
Chang, F.F., Qu, Q., and Frenier, W. 2001a. A Novel Self-Diverting-Acid Developed for Matrix
Stimulation of Carbonate Reservoirs. presented at the International Symposium on Oilfield
Chemistry held on Houston, Texas, USA, 13-16 February. SPE-65033-MS. http://dx.doi.org/
10.2118/65033-MS.
Chang, F.F., Acock, A.M., Geoghagan, A., Huckabee, P.T. 2001b. Experience in Acid Diversion in
High Permeability Deep Water Formations Using Visco-Elastic-Surfactant. Presented at the SPE
European Formation Damage Conference, The Hague, The Netherlands, 21-22 May. SPE-68919MS. http://dx.doi.org/10.2118/68919-MS.
Chang, F.F., et al. 2002. Fluid System Having Controllable Reversible Viscosity. US Patent #
6,399,546.
Chen, Y., Lee, J., and Pope, T.L. 2007. Carbon Dioxide Foamed Fluids. US Patent 7,291,651.
Chen, Y., Lee, J., and Pope, T.L. 2010. Carbon Dioxide Foamed Fluids. US Patent 7,803,744.
Chen, Y., Lee, J.C., and Samuel, M. 2008. Rheology Modifiers. US Patent 7,387,987.
Chen, Y., Pope, T.L., and Lee, J.C. 2005. Novel CO2-Emulsified Viscoelastic Surfactant Fracturing
Fluid System. Presented at the SPE European Formation Damage Conference, Scheveningen, The
Netherlands, 25-27 May. SPE-94603-MS. http://dx.doi.org/10.2118/94603-MS.
Chevalier, Y., Germanaud, L., and LePerchec, P. 1988. Micellar Properties of Zwitterionic Phosphobetaine Amphiphiles in Aqueous Solution: Influence of the Intercharge Distance. Colloid Polym.
Sci. 266(5): 441448. DOI: 10.1007/BF01457261.
Chevalier, Y., Mlis, F., and Dalbiez, J.P. 1992. Structure of Zwitterionic Surfactant Micelles:
Micellar Size and Intermicellar Interactions. J. Phys. Chem. 96(21): 8614 8619. DOI: 10.1021/
j100200a074.
Crews, J.B. 2006. Bacteria-Based and Enzyme-Based Mechanisms and Products for Viscosity Reduction Breaking of Viscoelastic Fluids. US Patent 7,052,901.
Crews, J.B. 2005. Internal Phase Breaker Technology for Viscoelastic Surfactant Gelled Fluids.
Presented at the SPE International Symposium on Oilfield Chemistry, Houston, Texas, USA, 2-4
February. SPE-93449-MS. http://dx.doi.org/10.2118/93449-MS.
Crews, J.B. 2009. Metal-Mediated Viscosity Reduction of Fluids Gelled with Viscoelastic Surfactants.
US Patent 7,595,284.
Crews, J.B. 2010a. Clean-Up Additive for Viscoelastic Surfactant Based Fluids. US Patent 7,655,603.
Crews, J.B. 2010b. Compositions and Use of Mono- and Polyenoic Acids for Breaking VES-Gelled
Fluids. US Patent 7,645,724.
Crews, J.B. and Huang, T. 2007. Internal Breakers for Viscoelastic-Surfactant Fracturing Fluids.
Presented at the SPE International Symposium on Oilfield Chemistry, Houston, Texas, USA, 28
February - 2 March. SPE-106216-MS. http://dx.doi.org/10.2118/106216-MS.
Crews, J.B. and Huang, T. 2013. Method to Complex Metals in Aqueous Treating Fluids for
VES-Gelled Fluids. US Patent Application 2013/0090270.

SPE-173776-MS

29

Crews, J.B. and Huang, T. 2008. Performance Enhancements of Viscoelastic Surfactant Stimulation
Fluids with Nanoparticles. Presented at the SPE Europec/EAGE Annual Conference and Exhibition, Rome, Italy, 9-12 June. SPE-113533-MS. http://dx.doi.org/10.2118/113533-MS.
Crews, J.B. and Huang, T. 2010. New Remediation Technology Enables Removal of Residual
Polymer in Hydraulic Fractures. Presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19-22 September. SPE-135199-MS. http://dx.doi.org/10.2118/135199-MS.
Crews, J.B., Huang, T., and Wood, W.R. 2008. New Technology Improves Performance of Viscoelastic Surfactant Fluids. SPE Drilling & Completion 23(01): 4147. SPE-103118-PA. http://dx.doi.org/10.2118/103118-PA.
Crews, J.B., Huang, T., and Wood, W.R. 2008a. The Future of Fracturing-Fluid Technology and Rates
of Hydrocarbon Recovery. Presented at the SPE Annual Technical Conference and Exhibition,
Denver, Colorado, USA, 21-24 September. SPE-115475-MS. http://dx.doi.org/10.2118/115475MS.
Crews, J.B., Huang, T., and Wood, W.R. 2008b. New Technology Improves Performance of Viscoelastic Surfactant Fluids. SPE Drilling & Completion 23(1): 4147. SPE-103118-PA. http://
dx.doi.org/10.2118/103118-PA.
Dahanayake, M.S., Yang, J., Niu, J.H.Y., Derian, P.J., Li, R., Dino, D. 2001. Viscoelastic Surfactant
Fluids and Related Methods of Use. US Patent 6,258,859.
Dahanayake, M.S., Yang, J., Niu, J.H.Y., Derian, P.J., Li, R., Dino, D. 2001. Viscoelastic Surfactant
Fluids and Related Methods of Use. US Patent 6,482,866.
Dahanayake, M.S., Yang, J., Niu, J.H.Y., Derian, P.J., Li, R., Dino, D. 2001. Viscoelastic Surfactant
Fluids and Related Methods of Use. US Patent 6,831,108.
Di Lullo Arias, G.F., Rae, P., Ahmad, A.J.K. 2002. Viscous Fluid Applicable for Treating Subterranean Formations. US Patent 6,491,099.
Dreiss, C.A. 2007. Wormlike Micelles: Where Do We Stand? Recent Developments, Linear Rheology, and Scattering Techniques. Soft Matter 3: 956 970. DOI: 10.1039/B705775J.
Economides, M., and Nolte, K. 1989. Reservoir Stimulation, second edition. Englewood Cliffs, New
Jersey, USA, Prentice Hall Inc.
Fakoya, M.F. and Shah, S.N. 2013. Rheological Properties of Surfactant-Based and Polymeric
Nano-Fluids. Presented at the 2013 SPE/ICoTA Coiled Tubing and Well Intervention Conference
and Exhibition, The Woodlands, Texas, USA, 26-27 March. SPE-163921-MS. http://dx.doi.org/
10.2118/163921-MS
Fontana, C., Muruaga, E., Perez, D., Cavazzoli, G., Krenz, A. 2007. Successful Application of a High
Temperature Viscoelastic Surfactant (VES) Fracturing Fluids Under Extreme Conditions in
Patagonian Wells, San Jorge Basin. Presented at the SPE Europec/EAGE Annual Conference and
Exhibition, London, United Kingdom, 11-14 June. SPE-107277-MS. http://dx.doi.org/10.2118/
107277-MS
Fredd, C.N., Olsen, T.N., Brenize, G., Quintero, B.W., Bui, T., Glenn, S., Boney, C.L. 2004.
Polymer-Free Fracturing Fluid Exhibits Improved Cleanup for Unconventional Natural Gas Well
Applications. Presented at the SPE Eastern Regional Meeting, Charleston, West Virginia, 15-17
September. SPE-91433-MS. http://dx.doi.org/10.2118/91433-MS
Fu, D., Chen, Y., Xiao, Z., Samuel, M., Sylvie, D. 2006. Viscoelastic Surfactant Fluids Stable at High
Brine Concentration and Methods of Using the Same. US Patent 7,148,185.
Gomaa, A.M., Cawiezel, K.E., Gupta, D.V.S., Nasr-El-Din, H.A. 2011. Viscoelastic Evaluation of a
Surfactant Gel for Hydraulic Fracturing. Presented at the SPE European Formation Damage
Conference, Noordwijk, The Netherlands, 7-10 June. SPE-143450-MS. http://dx.doi.org/10.2118/
143450-MS

30

SPE-173776-MS

Gomaa, A.M., Gupta, D.V.S., and Carman, P. 2014. Viscoelastic Behavior and Proppant Transport
Properties of a New Associative Polymer-Based Fracturing Fluid. Paper presented at the SPE
International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana,
26-28 February. SPE-168113-MS. http://dx.doi.org/10.2118/168113-MS
Gomaa, A.M. and Nasr-El-Din, H.A. 2010. New Insights Into the Viscosity of Polymer-Based
In-Situ-Gelled Acids. SPE Production & Operations 25(03): 367375. SPE-121728-PA. http://
dx.doi.org/10.2118/121728-PA.
Gravsholt, S. 1976. Viscoelasticity in Highly Dilute Aqueous Solutions of Pure Cationic Detergents.
Journal of Colloid and Interface Science 57(3): 575577. DOI: 10.1016/0021-9797(76)90236-8.
Gupta, D.V.S. 2009. Unconventional Fracturing Fluids for Tight Gas Reservoirs. Presented at the SPE
Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 19-21 January.
SPE-119424-MS. http://dx.doi.org/10.2118/119424-MS
Gupta, D.V.S. and Hlidek, B.T. 2010. Frac-Fluid-Recycling and Water Conservation: A Case History.
SPE Production & Operations 25(1): 6569. SPE-119478-PA. http://dx.doi.org/10.2118/119478PA.
Gupta, D.V.S., Leshchyshyn, T.T., and Hlidek, B.T. 2005. Surfactant Gel Foam/Emulsions: History
and Field Application in the Western Canadian Sedimentary Basin. Presented at the SPE Annual
Technical Conference and Exhibition, Dallas, Texas, USA, 9-12 October. SPE-97211-MS. http://
dx.doi.org/10.2118/97211-MS
Gupta, D.V.S. and Tudor, E.H. 2005. Method for Fracturing Subterranean Formations. US Patent No.
6,875,728.
Gurluk, M.R., Nasr-El-Din, H.A., and Crews, J.B. 2013. Enhancing the Performance of Viscoelastic
Surfactant Fluids Using Nanoparticles. Presented at the EAGE Annual Conference & Exhibition,
London, United Kingdom, 10-13 June. SPE-164900-MS. http://dx.doi.org/10.2118/164900-MS
Hamley, I.W. 2007. Introduction to Soft Matter: Synthetic and Biological Self-Assembling Materials.
Hoboken, New Jersey: John Wiley & Sons Inc.
Hargreaves, T. 2003. Chemical Formulation: An Overview of Surfactant-Based Preparations Used in
Everyday Life. Cambridge: The Royal Society of Chemistry.
He, Z., Wang, G., Nasr-El-Din, H.A., Holt, S. 2013. Hydrolysis Effect on the Properties of a New
Class of Viscoelastic Surfactant-Based Acid and Damage Caused by the Hydrolysis Products.
Presented at the SPE European Formation Damage Conference & Exhibition, Noordwijk, The
Netherlands, 5-7 June. SPE-165161-MS. http://dx.doi.org/10.2118/165161-MS.
Helgeson, M.E., Hodgdon, T.K., Kaler, E.W., Wagner, N.J., Vethamuthu, M., Ananthapadmanabhan,
K.P. 2010. Formation and Rheology of Viscoelastic Double Networks in Wormlike MicelleNanoparticle Mixtures. Langmuir 26(11): 8049 8060. DOI: 10.1021/la100026d.
Huang, T. 2014. Breaking Viscoelastic Surfactant Gelled Fluids Using Breaker Nanoparticles. US
Patent 8,778,852.
Huang, T. 2009. Viscosity Enhancers for Viscoelastic Surfactant Stimulation Fluids. US Patent
7,544,643.
Huang, T. and Clark, D.E. 2013. Improving Fracture Fluid Performance and Controlling Formation
Fines Migration with the Same Agent: Is It Achievable? Presented at the International Petroleum
Technology Conference, Beijing, China, 26-28 March. IPTC-17044-MS. http://dx.doi.org/
10.2523/17044-MS.
Huang, T., Crews, J.B. 2008a. Do Viscoelastic-Surfactant Diverting Fluids for Acid Treatments Need
Internal Breakers? Paper SPE 112484 presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA, 13-15 February. http://dx.doi.org/10.2118/112484-MS.

SPE-173776-MS

31

Huang, T., Crews, J.B. 2008b. Nanotechnology Applications in Viscoelastic-Surfactant Stimulation


Fluids. SPE Production & Operations 23(4): 512517.
Huang, T., Crews, J.B. 2009a. Fluid-Loss Control Improves Performance of Viscoelastic Surfactant
Fluids. SPE Production & Operations 24(1), 60 65. SPE-107728-PA. http://dx.doi.org/10.2118/
107728-PA.
Huang, T., Crews, J.B. 2009b. Use of Mineral Oils to Reduce Fluid Loss for Viscoelastic Surfactant
Gelled Fluids. US Patent 7,615,517.
Huang, T., Crews, J.B., and Agrawal, G. 2010a. Nanoparticle Pseudocrosslinked Micellar Fluids:
Optimal Solution for Fluid-Loss Control With Internal Breaking. Presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA,
10-12 February. SPE-128067-MS. http://dx.doi.org/10.2118/128067-MS.
Huang, T., Evans, B.A., Crews, J.B., Belcher, C.K. 2010b. Field Case Study on Formation Fines
Control with Nanoparticles in Offshore Wells. Presented at the SPE Annual Technical Conference
and Exhibition, Florence, Italy, 19-22 September. SPE-135088-MS. http://dx.doi.org/10.2118/
135088-MS.
Hughes, T.L., Jones, T.G.J., and Tustin, G.J. 1999. Viscoelastic Surfactant Base Gelling Composition
for Wellbore Service Fluids. UK Patent GB2332223.
Imanishi, K. and Einaga, Y. 2007. Wormlike Micelles of Polyoxyethylene Alkyl Ether Mixtures C10E5
C14E5 and C14E5 C14E7: Hydrophobic and Hydrophilic Chain Length Dependence of the
Micellar Characteristics. J. Phys. Chem. B. 111(1): 6273. DOI 10.1021/jp065317l.
Israelachvili, J.N. 1992. Intermolecular and Surface Forces. London: Academic Press.
Israelachvili, J.N., Mitchell, D.J., and Ninham, B.W. 1976. Theory of Self-Assembly of Hydrocarbon
Amphiphiles into Micelles and Bilayers. J. Chem. Soc., Faraday Trans. 2 72: 15251567. DOI:
10.1039/F29767201525.
Jerke, G., Pedersen, J.S., Egelhaaf, S.U., Schurtenberger, P. 1998. Flexibility of Charged and
Uncharged Polymer-Like Micelles. Langmuir 14(21): 60136024. DOI: 10.1021/la980390r.
Kreh, K.A. 2009. Viscoelastic Surfactant-Based Systems in the Niagaran Formation. Presented at the
SPE Eastern Regional Meeting, Charleston, West Virginia, USA, 23-25 September. SPE-125754MS. http://dx.doi.org/10.2118/125754-MS.
Kuperkar, K., Abezgauz, L., Danino, D., Verma, G., Hassan, P.A., Aswal, V.K., Varade, D., Bahadur,
P. 2008. Viscoelastic Micellar Water/CTAB/NaNO3 Solutions: Rheology, SANS and Cryo-TEM
Analysis. J. Colloid Interface Sci. 323: 403409. DOI: 10.1016/j.jcis.2008.04.040.
Lee, J., Chen, Y., Pope, T., Hanson, E., Cozzens, S., Batmaz, T. 2008. Viscoelastic Surfactant
Rheology Modification. US Patent 7,341,980.
Leitzell, J.R., Viscoelastic Surfactants: A New Horizon in Fracturing Fluids for Pennsylvania. 2007.
Presented at the SPE Eastern Regional Meeting, Lexington, Kentucky, USA, 17-19 October.
SPE-111182-MS. http://dx.doi.org/10.2118/111182-MS
Luo, M., Jia, Z., Sun, H., Liao, L., Wen, Q. 2012. Rheological Behavior and Microstructure of an
Anionic Surfactant Micelle Solution with Pyroelectric Nanoparticle. Colloid Surf., A 395: 267
275. DOI: 10.1016/j.colsurfa.2011.12.052.
Lungwitz, B.R., Fredd, C.N., Brady, M.E., Miller, M.J., Ali, S.A., Hughes, K.N. 2007. Diversion and
Cleanup Studies of Viscoelastic Surfactant-Based Self-Diverting Acid. SPE Production & Operations 22(01): 121127. SPE-86504-PA. http://dx.doi.org/10.2118/86504-PA.
Lynn, J.D., Nasr-El-Din, H.A. 2001. A Core Based Comparison Of The Reaction Characteristics Of
Emulsified And In-Situ Gelled Acids In Low Permeability, High Temperature, Gas Bearing
Carbonates. Presented at the SPE International Symposium on Oilfield Chemistry, Houston,
Texas, USA, 13-16 February. SPE-65386-MS. http://dx.doi.org/10.2118/65386-MS.

32

SPE-173776-MS

McElfresh, P.M., Williams, C.F. 2007. Hydraulic Fracturing Using Non-Ionic Surfactant Gelling
Agent. US Patent 7,216,709.
McElfresh, P.M., Williams, C.F., Wood, W.R., Ruzic, M., Baycroft, P.D. 2003. A Single Additive
Non-Ionic System for Frac Packing Offers Operators a Small Equipment Footprint and High
Compatibility with Brines and Crude Oils. Presented at the SPE European Formation Damage
Conference, The Hague, The Netherlands, 13-14 May. SPE-82245-MS. http://dx.doi.org/10.2118/
82245-MS.
Miller, M.J., Samuel, M., Vinod, P.S., Olsen, T.N. 2003. Compositions and Methods to Control Fluid
Loss in Surfactant-Based Wellbore Service Fluids. US Patent 6,605,570.
Mohammed, S.K., Nasr-El-Din, H.A., Erbil, M.M. 2005. Successful Application of Foamed Viscoelastic Surfactant-Based Acid. Presented at the SPE European Formation Damage Conference,
Sheveningen, The Netherlands, 25-27 May. SPE-95006-MS. http://dx.doi.org/10.2118/95006-MS.
Morvan, M., Degre, G. 2012. Viscoelastic Composition with Improved Stability. US Patent 2012/
0086653.
Nagarajan, R. 2002. Molecular Packing Parameter and Surfactant Self-Assembly: The Neglected Role
of the Surfactant Tail. Langmuir 18: 18 38. DOI: 10.1021/la010831y.
Nagarajan, R. and Ruckenstein, E. 1991. Theory of Surfactant Self-Assembly: A Predictive Molecular
Thermodynamic Approach. Langmuir 7: 2934 2969. DOI: 10.1021/la00060a012.
Nasr-El-Din, H.A., Chesson, J.B., Cawiezel, K.E., De Vine, C.S. 2006a. Investigation and Field
Evaluation of Foamed Viscoelastic Surfactant Diversion Fluid Applied During Coiled-Tubing
Matrix-Acid Treatment. Presented at the SPE/ICoTA Coiled Tubing Conference & Exhibition,
The Woodlands, Texas, USA, 4-5 April. SPE-99651-MS. http://dx.doi.org/10.2118/99651-MS.
Nasr-El-Din, H.A., Chesson, J.B., Cawiezel, K.E., De Vine, C.S. 2006b. Lessons Learned and
Guidelines for Matrix Acidizing With Viscoelastic Surfactant Diversion in Carbonate Formations.
Presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, USA,
24-27 September. SPE-102468-MS. http://dx.doi.org/10.2118/102468-MS.
Nehmer, W.L. 1988. Viscoelastic Gravel-Pack Carrier Fluid. Presented at the SPE Formation Damage
Control Symposium, Bakersfield, California, USA, 8-9 February. SPE-17168-MS. http://dx.doi.org/10.2118/17168-MS.
Nelson, E.B., Lungwitz, B., Dismuke, K., Samuel, M., Salamat, G., Hughes, T., Lee, J., Fletcher, P.,
Fu, D., Hutchins, R., Parris, M., Tustin, G.J. 2005. Viscoelastic Reduction of Viscoelastic
Surfactant Based Fluids. US Patent 6,881,709.
Nettesheim, F., Liberatore, M.W., Hodgdon, T.K., Wagner, N.J., Kaler, E.W., Vethamuthu, M. 2008.
Influence of Nanoparticle Addition on the Properties of Wormlike Micellar Solutions. Langmuir
24: 7718 7726. DOI: 10.1021/la800271m.
Norman, W.D., Jasinski, R.J., Nelson, E.B. 1996. Hydraulic Fracturing Process and Compositions. US
Patent 5,551,516.
Palisch, T., Duenckel, R.; Bazan, L., Heidt, H.J., Turk, G. 2007. Determing Realistic Fracture
Conductivity and Understanding Its Impact on Well PerformanceTheory and Field Examples.
Presented at the SPE Hydraulic Fracturing Technology Conference, College Station, Texas, USA,
29-31 January. SPE-106301-MS. http://dx.doi.org/10.2118/106301-MS.
Pandey, V.J., Itibrout, T., Adams, L.S., Cowan, T.L., Bustos, O.A. 2007. Fracture Stimulation
Utilizing a Viscoelastic-Surfactant Based System in the Morrow Sands in Southeast New Mexico.
Presented at the SPE International Symposium on Oilfield Chemistry, Houston, Texas, USA, 28
February-2 March. SPE-102677-MS. http://dx.doi.org/10.2118/106301-MS.
Qu, Q., et alet al. 2002. Compositions Containing Aqueous Viscosifying Surfactants and Methods for
Applying Such Compositions in Subterranean Formations. US Patent # 6,435,277.

SPE-173776-MS

33

Raghavan, S.R. and Kaler, E.W. 2001. Highly Viscoelastic Wormlike Micellar Solutions Formed by
Cationic Surfactants with Long Unsaturated Tails. Langmuir 17: 300 306. DOI: 10.1021/
la0007933.
Rawat, A., Tripathi, A., and Gupta, C. 2014. Case Evaluating Acid Stimulated Multilayered Well
Performance in Offshore Carbonate Reservoir: Bombay High. Presented at the Offshore Technology Conference-Asia, Kuala Lumpur, Malaysia, 25-28 March. OTC-25018-MS. http://dx.doi.org/10.4043/25018-MS.
Reddy, B.R. 2011. Well Treatment Fluids Containing A Viscoelastic Surfactant and a Cross-Linking
Agent Comprising a Water-Soluble Transition Metal Complex. US Patent Application 2011/
0105369.
Rose, G.D., Teot, A.S., Doty, P.A. 1988. Process for Reversible Thickening of a Liquid. US Patent
No. 4,735,731.
Samuel, M. 2009. Gelled Oil with Surfactant. US Patent 7,521,400.
Samuel, M.M., Balabatyrov, Y., Chang, F.F., Griffith, M., Morris, L. 2011. Methods of Perforation
Using Viscoelastic Surfactant Fluids and Associated Compounds. US Patent 7,878,246.
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. SPE-59478-PA. http://dx.doi.org/10.2118/59478-PA.
Samuel, M., Card, R.J., Nelson, E.B., Brown, J.E., Vinod, P.S., Temple, H.L. Qu, Q., Fu, D.K. 1997.
Polymer-Free Fluid for Hydraulic Fracturing. Presented at the SPE Annual Technical Conference
and Exhibition, San Antonio, Texas, USA, 5-8 October. SPE-38622-MS. http://dx.doi.org/
10.2118/38622-MS.
Samuel, M.M., Dismuke, K.I., Card, R.J., Brown, J.E., England, K.W. 2001. Methods of Fracturing
Subterranean Formations. US Patent 6,306,800.
Samuel, M.M., Li, L., and Juel, J.E. 2014. Gelled Hydrocarbon System and Method with DualFunction Viscosifier/Breaker Additive. US Patent 8,653,011.
Samuel, M., Marcinew, R., Al-Harbi, M., Samuel, E., Xiao, Z., Ezzat, A.M., Khamees, S.A., Jarrett,
C., Ginest, N.H., Bartko, K., Hembling, D., Nasr-El-Din, H.A. 2003. A New Solids-Free NonDamaging High Temperature Lost-Circulation Pill: Development and First Field Applications.
Presented at the SPE 13th Middle East Show & Conference, Bahrain, 5-8 April. SPE-81494-MS.
http://dx.doi.org/10.2118/81494-MS.
Samuel, M., Marcinew, R., Xiao, Z. 2007. Non-Damaging Fluid-Loss Pill and Method of Using the
Same. US Patent 7,207,388.
Samuel, M., Polson, D., Kordziel, W., Waite, T., Waters, G., Vinod, P.S., Fu, D., Downey, R. 2000.
Viscoelastic Surfactant Fracturing Fluids: Application in Low Permeability Reservoirs. Presented
at the SPE Rocky Mountain Regional/Low Permeability Reservoirs Symposium and Exhibition,
Denver, Colorado, USA, 12-15 March. SPE-60322-MS. http://dx.doi.org/10.2118/60322-MS.
Schubert, B.A., Kaler, E.W., and Wagner, N.J. 2003. The Microstructure and Rheology of Mixed
Cationic/Anionic Wormlike Micelles. Langmuir 19: 4079 4089. DOI: 10.1021/la020821c.
Semmelbeck, M.E., Deupree, W.E., von Plonski, J.K., Mueller, F.A., Chen, Y., Lewis, J.W., Keto,
L.K., Fairhurst, D.L., Pope, T.L. 2006. Novel CO2-Emulsified Viscoelastic Surfactant Fracturing
Fluid System Enables Commercial Production from Bypassed Pay in the Olmos Formation of
South Texas. Presented at the SPE Gas Technology Symposium, Calgary, Alberta, Canada, 15-17
May. SPE-100524-MS. http://dx.doi.org/10.2118/100524-MS.
Stewart, B.R., Mullen, M.E., Howard, W.J., Norman, W.D. 1994. Use of a Solids-Free Viscous
Carrying Fluid in Fracturing Applications: An Economic and Productivity Comparison in Shallow
Completions. Presented at the SPE European Formation Damage Control Conference, Hague, The
Netherlands, 15-16 May. SPE-30114-MS. http://dx.doi.org/10.2118/30114-MS.

34

SPE-173776-MS

Sullivan, P., Christanti, Y., Couillet, I., Davies, S., Hughes, T., Wilson, A. 2006. Methods for
Controlling the Fluid Loss Properties of Viscoelastic Surfactant Based Fluids. US Patent
7,081,439.
Sullivan, P.F., Gadiyar, B., Morales, R.H., Hollicek, R., Sorrells, D., Lee, J., Fischer, D. 2006.
Optimization of a Viscoelastic Surfactant (VES) Fracturing Fluid for Application in HighPermeability Formations. Presented at the SPE International Symposium and Exhibition on
Formation Damage Control, Lafayette, Louisiana, USA, 15-17 February. SPE-98338-MS. http://
dx.doi.org/10.2118/30114-MS.
Tanford, C. 1974. Theory of Micelle Formation in Aqueous Solutions. J. Phys. Chem. 78(24):
2469 2479. DOI: 10.1021/j100617a012.
Tanford, C. 1979. Interfacial Free Energy and the Hydrophobic Effect. Proc. Natl. Acad. Sci.76(9):
41754176.
Taylor, K.C. and Nasr-El-Din, H.A. 2001. Laboratory Evaluation of In-Situ Gelled Acids for
Carbonate Reservoirs. Presented at the SPE Annual Technical Conference and Exhibition, New
Orleans, Louisiana, USA, 30 September 3 October. SPE-71694-MS. http://dx.doi.org/10.2118/
71694-MS.
Teot, A.S., Ramaiah, M., Coffey, M.D. 1988. Aqueous Wellbore Service Fluids. US Patent 4,725,372.
van Zanten, R. 2011. Stabilizing Viscoelastic Surfactants in High-Density Brines. SPE Drilling &
Completion 26(4): 499 505. SPE-141447-PA. http://dx.doi.org/10.2118/141447-PA.
van Zanten, R., Ezzat, D. 2011. Advanced Viscoelastic Surfactant Gels for High-Density Completion
Brines. Presented at the SPE European Formation Damage Control Conference, Noordwijk, The
Netherlands, 7-10 June. SPE-143844-MS. http://dx.doi.org/10.2118/143844-MS.
Wang, G., Nasr-El-Din, H.A., Zhou, J., Holt, S. 2012. A New Viscoelastic Surfactant for High
Temperature Carbonate Acidizing. Presented at the SPE Saudi Arabia Section Technical Symposium and Exhibition, Al-Khobar, Saudi Arabia, 8-11 April. SPE-160884-MS. http://dx.doi.org/
10.2118/160884-MS.
Welton, T.D., Bryant, J.E. 2011. Subterranean Treatment Fluids Comprising Viscoelastic Surfactant
Gels. US Patent 7,997,342.
Whalen, R.T. 2000. Viscoelastic Surfactant Fracturing Fluids and a Method for Fracturing Subterranean Formations. US Patent 6,035,936.
Witten, T.A.; Pincus, P.A. 2010. Structured Fluids: Polymers, Colloids, Surfactants. New York:
Oxford University Press.
Yang, J., Guan, B., Lu, Y., Cui, W., Qiu, X., Yang, Z., Qin, W. 2013. Viscoelastic Evaluation of
Gemini Surfactant Gel for Hydraulic Fracturing. Presented at the SPE European Formation
Damage Conference and Exhibition, Noordwijk, The Netherlands, 7-10 June. SPE-165177-MS.
http://dx.doi.org/10.2118/165177-MS.
Yu, M., Mahmoud, M.A., and Nasr-El-Din, H.A. 2011. Propagation and Retention of Viscoelastic
Surfactants Following Matrix-Acidizing Treatments in Carbonate Cores. SPE Journal 16(04):
9931,001. SPE-128047-PA. http://dx.doi.org/10.2118/128047-PA.
Yu, M., Mu, Y., Wang, G., and Nasr-El-Din, H.A. 2012. Impact of Hydrolysis at High Temperatures
on the Apparent Viscosity of Carboxybetaine Viscoelastic Surfactant-Based Acid: Experimental
and Molecular Dynamics Simulation Studies. SPE Journal 17(04): 1119 1130. SPE-142264-PA.
http://dx.doi.org/10.2118/142264-PA.
Zerhbouh, M. 1993. Product and Process for Acid Diversion in the Treatment of Subterranean
Formations. Patent US 5,203,413.
Zeilinger, S.C., Wang, M., Kibodeaux, K.R., Rossen, W.R. 1995. Improved Prediction of Foam
Diversion in Matrix Acidizing. Presented at the Production Symposium, Oklahoma, USA, 2-4
April. SPE-29529-MS. http://dx.doi.org/10.2118/29529-MS.

SPE-173776-MS

35

Zhang, K. 2002. Fluids for Fracturing Subterranean Formations. US Patent 6,468,945.


Zhang, K., Pierce, R., Litt, N.D., Gupta, D.V.S. 2002. Foam-Fluid for Fracturing Subterranean
Formations. US Patent No. 6,410,489.

You might also like