Top Banner

of 35

Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

Jul 06, 2018

Download

Documents

Lê Công
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    1/35

    SPE-173776-MS

    Recent Advances in Viscoelastic Surfactants for Improved Production fromHydrocarbon Reservoirs

    Katherine L. Hull, and Mohammed Sayed, Aramco Research Center—Houston; Ghaithan A. Al-Muntasheri, Aramco Research Center—Houston, 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, 13–15 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, theacid 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.

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    2/35

    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.

    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 

    Table 1—A 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).

    2 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    3/35

    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).

    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 temper-

    ature 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 FormationWhen 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 hydro-

    carbon-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:

    Figure 1—Schematic 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.

    SPE-173776-MS 3

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    4/35

    (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.

    (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 

    Table 2—Contributions to the overall free energy of micellization are provided along with physical descriptions of their effects

    upon the aggregation behavior of the micelle.

    4 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    5/35

    qualitatively predicting the micelle structure and is still widely used (Israelachvili et al. 1976; Israelachvili

    1992). The parameter is defined as   0 /ael 0 where   0 is the volume of the surfactant tail, l 0 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. Asummary of the various micelle shapes and their corresponding packing parameters is provided in  Figure

    2.

    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.

    Figure 2—Schematic 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).

    SPE-173776-MS 5

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    6/35

    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, staticlight 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,  R g , 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,  l  p, is the length over which the micelle is considered to be rigid.

    As an example,  l  p 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,   l  p   was found to be 20 nm while the

    cross-section radius of the micelle,  Rcs, was determined to be 2 nm.

    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 k 1   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

    Figure 3—Wormlike micelle schematic showing the fundamental structural parameters that cover a wide range of length scales

    (adapted from Dreiss 2007).

    6 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    7/35

    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  G 0   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 character-

    ized by a zero-shear viscosity,   0     G 0  R. However, when the critical shear rate is exceeded, thewormlike 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 beendeveloped to “pseudo-crosslink” the surfactant micelles and enhance the viscosity of the fluids. Further-

    more, 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 polymer-

     based 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 7

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    8/35

    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.

    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

    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.

    8 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    9/35

    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 be-

    havior 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.

    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

    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.

    SPE-173776-MS 9

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    10/35

    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.

    Hydraulic fracturing VES-based fluids that are composed of anionic surfactants such as alkyl sarco-

    sinates 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 to6.5-10, 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 incorpo-

    rated 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.

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

    components of the tallow amido substituent.

    10 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    11/35

    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 deter-

    mined 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).

    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

    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.

    Figure 7—The chemical structures of (a) tallowiminodipropionate where Rtallow and (b) oleylamidopropyl betaine and (c) erucyl-

    amidopropyl betaine.

    SPE-173776-MS 11

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    12/35

    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 suspen-

    sion 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) pyroelectricnanoparticles 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.

    12 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    13/35

    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 widevariety 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 trietha-

    nolamine 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/CaBr 2, 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

    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).

    SPE-173776-MS 13

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    14/35

    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/CaBr 2 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/CaBr 2   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 ¼ 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 systemswere 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 borate-

    crosslinked 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 CaBr 2 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 pseudo-

    crosslinked VES-based fracturing fluids is given below in  Table 5.

    14 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    15/35

    Table 5—Summary of experiments performed with pseudo-crosslinkers. (R“regular” and PE”pyroelectric” nanoparticles)

    SPE-173776-MS 15

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    16/35

    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). Nanopar-

    ticles 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 pseudo-

    crosslink 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

    16 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    17/35

    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 treat-

    ments 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 solution’sionic 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

    SPE-173776-MS 17

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    18/35

    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 fluid’s 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, CO2is 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 Chenet 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 stimula-

    tions), 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 concentra-

    tions 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

    18 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    19/35

    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 three-

    dimensional) 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 averageconductivity 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.

    SPE-173776-MS 19

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    20/35

    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 a59% 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.

    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%.

    20 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    21/35

    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 concen-

    trations 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 CaBr 2. At high temperatures, around 5% methanolis 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 precipi-

    tation 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 200°F. 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

    SPE-173776-MS 21

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    22/35

    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 200°F and their ineffectiveness in plugging formationswith 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 reactwith 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 250°F and also up to 350°F 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

    350°F. At a temperature of 250°F, 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 100°C 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 200°F and 250°F 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 200°F. They noticed that

    22 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    23/35

    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.

    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 perme-

    ability in the range of 0.1 to 50 md at a temperature of 200 to 240°F 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 150°F 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 70°F) 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.

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

    VES CONC. (vol%) VISCOSITY Water, 200 °F (cp) VISCOSITY Water, 250 °F (cp) VISCOSITY CaCl2

    ,250 °F (cp)

    0.5 138 113 118

    1.0 152 - 141

    1.5 - 152 160

    3.0 433 312 -

    SPE-173776-MS 23

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    24/35

    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 325°F 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 190°F, 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 190°F for 4, 6, and 8 vol% VES

    concentration (Figures 3-5 of the manuscript). Similar behavior was noted at each surfactant concentra-

    tion.  He et al. (2013)  studied the effect of the hydrolysis of VES-based acids on the rheology and the

     performance of such systems at 250°F. 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 190°F. 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.

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

    24 SPE-173776-MS

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    25/35

    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.

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

    2013).

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

    LOCATION WELL INFO VES (vol%) HCl (wt%)

    PRODUCTION

    (bbl/day) REFERENCE

    Gulf of Mexico 198 °F, 14,500 ft - 10 5 Aceti c “Above expec ted  

    rates”

    Alleman et al. 2003

    Gulf of Mexico 150 °F, 9,600 ft - 10 5 Aceti c “Above opera tor  

    expectations”

    Alleman et al. 2003

    - Mainly calcite 6 20 6,000   Mohammed et al.

    2005

    Saudi Arabia   Water Injector*

    Horizontal

    3 20   75,000 (450% inc.)   Nasr-El-Din et al.

    2006aSaudi Arabia Tubing: 6,270 ft

    Vertical

    6 20 (150% inc.)   Nasr-El-Din et al.

    2006b

    Saudi Arabia - 6 20 3,700   Nasr-El-Din et al.

    2006b

    Saudi Arabia 200 °F, 9,310 ft

    3,270 psi

    4 20 -   Nasr-El-Din et al.

    2006b

     Niagaran Reef (MI) Mainly dolomite - 15-20 80 (400% inc.) 2X

    acid alone

    Kreh 2009

    Bombay High

    Oilfield, India

    5-500 md 3 20 800 (150% inc.)   Rawat et al. 2014

    SPE-173776-MS 25

  • 8/17/2019 Recent Advances in Viscoelastic Surfactants for Improved Production from Hydrocarbon Reservoirs.pdf

    26/35

    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 H 2S 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