Вы находитесь на странице: 1из 16

J179702 DOI: 10.

2118/179702-PA Date: 17-November-18 Stage: Page: 2202 Total Pages: 16

New Surfactants and Cosolvents Increase


Oil Recovery and Reduce Cost
Karasinghe A. Nadeeka Upamali, Pathma Jith Liyanage, Sung Hyun Jang, Erin Shook,
Upali P. Weerasooriya, and Gary A. Pope, University of Texas at Austin

Summary
Scientific understanding of how the molecular structures of surfactants and cosolvents affect microemulsion properties greatly speeds up the
process of arriving at optimal chemical formulations for enhanced recovery of a specific crude oil. With the main emphasis on reducing the
chemical cost of the formulations, novel surfactants and cosolvents have been developed and shown to have superior performance. We have
synthesized and tested surfactants with different hydrophobe sizes and structures varying from ultrashort to large to satisfy a variety of crude-
oil requirements over a wide range of reservoir conditions. The cosolvents and surfactants with ultrashort hydrophobes offer advantages such
as short equilibration time for the microemulsion formation and lower microemulsion viscosity. Chemical formulations developed using
these chemicals have shown excellent performance with very low cosolvent and surfactant retention in cores. Low retention means less
chemicals can be used to recover each barrel of oil from the reservoir. These chemicals can be made commercially at low cost. Through use
of these new developments, the chemical cost per barrel of oil is low enough to be economically viable, even at low crude-oil prices.

Introduction
Chemical enhanced oil recovery (EOR) (CEOR) offers the potential for recovering a large fraction of the remaining oil after waterflood-
ing. Significant breakthroughs in the recent past have made it possible to use CEOR over a wider range of reservoir conditions. These
advances have also led to higher efficiencies, lower surfactant retention, and lower chemical costs to recover incremental oil. Pandey et al.
(2016) reports the results of a recently completed and very successful alkaline/surfactant/polymer (ASP) pilot with new surfactants after a
polymer flood (Pandey et al. 2012; Prasad et al. 2014). Many of the recent advances in CEOR technology have come about because of the
development of new or improved surfactants and cosolvents (Iglauer et al. 2009; Adkins et al. 2010, 2012; Barnes et al. 2010, 2012; Yang
et al. 2010; Puerto et al. 2012; Ge and Wang 2015; Jürgenson et al. 2015; Liyanage et al. 2015a, b). New chemical structures are beneficial
in several ways, but especially in terms of increasing robustness and reducing surfactant retention and chemical cost. The emphasis in this
paper is on new hydrophobes with varying carbon-chain lengths and degrees of branching in both surfactants and cosolvents.
Different crude oils prefer different surfactant-hydrophobe types, so various classes of surfactants with different hydrophobe sizes
and structures, varying from very large to medium to very short, are needed (Arf et al. 1987; Puerto et al. 2012). Large-hydrophobe
Guerbet alkoxy sulfate and carboxylate surfactants have been shown to yield ultralow interfacial tension (IFT) and high oil recovery,
even for crude oils with a very high equivalent alkane carbon number (Adkins et al. 2010, 2012; Lu et al. 2014). The midpoint bent oleyl
hydrophobe of medium chain length (C18) has been found to be highly effective for many crude oils when an appropriate number of pro-
poxy (PO) and ethoxy (EO) units are incorporated in either the sulfate or the carboxylate surfactant molecule (Liyanage et al. 2015b).
By varying the number of POs and EOs for different hydrophobe structures, the surfactants can be fine-tuned for enhanced inter-
action at the oil/water interface (Gale et al. 1978; Aoudia et al. 1995; Miñana-Perez et al. 1996; Barnes et al. 2010; Solairaj et al. 2012;
Song et al. 2015). Light crude oils are more compatible with short hydrophobes such as tridecyl alcohol (C12–13) rather than large- or
medium-sized hydrophobes. Furthermore, the type and structure of the surfactant molecule as well as that of the hydrophobe is very
important for microemulsion formation (not macroemulsion), hardness tolerance, thermal stability, and favorable microemulsion rheol-
ogy. In this context, we have also revisited the anionic Gemini surfactant as a new class of large-hydrophobe EOR surfactant (Menger
and Littau 1993; Gao and Sharma 2013; Kamal 2016).
Many conventional cosolvents (e.g., simple alcohols and their low-numbered ethoxymers) improve aqueous stability, facilitate faster
equilibration of the microemulsion phase, lower the microemulsion viscosity, mitigate the formation of ordered/condensed structures,
and broaden the low-IFT salinity window, among other benefits (Salter 1977; Abe et al. 1986; Sanz and Pope 1995; Frank et al. 2007;
Levitt et al. 2009; Sahni et al. 2010; Fortenberry et al. 2015). However, cosolvents also tend to increase the IFT, although some
cosolvents with lipophilic and hydrophilic linkers can decrease the IFT (Acosta et al. 2004; Graciaa et al. 1993a, b). We show that
2-ethylhexanol-xPO sulfate (2-EHS), a surfactant with an ultrashort hydrophobe, has both surfactant and cosolvent properties. Our results
indicate that branching in the 2-EHS structure not only minimizes the formation of viscous phases and gels, but also enhances the rapid
equilibration of microemulsion and minimizes or eliminates the need for cosolvent. We have also studied the effect of the incorporation
of PO (more branching) followed by EO groups into previously developed cosolvents such as iso-butanol (IBA) and phenol ethoxylates.
Highly branched di-isopropyl amine ethoxylate could be another attractive cosolvent candidate that has not been tested before in CEOR.
In the following sections, we demonstrate how novel cosolvents in combination with surfactants can impart superior CEOR perform-
ance such as reduced chemical usage while maintaining high surface activity. This does not mean to suggest that cosolvents are always
needed or should always be used. There are exceptional cases where all the stringent performance requirements such as aqueous stabil-
ity can be met without using cosolvent. In such cases, the decision whether to use cosolvent in the chemical slug and/or the polymer
drive should be dependent on whether it lowers the chemical cost per barrel of incremental oil production by reducing the surfactant
use, or from its other performance benefits for the particular reservoir conditions.

Experiment Materials and Procedure


Materials. Anionic Surfactants. Oleyl, tridecyl alcohol, and 2-ethylhexanol alkoxylates were supplied by an industry manufacturer
and subsequent sulfation and carboxymethylation steps were performed in the laboratories of the University of Texas at Austin. Alcohol

Copyright V
C 2018 Society of Petroleum Engineers

This paper (SPE 179702) was accepted for presentation at the SPE Improved Oil Recovery Conference, Tulsa, 11–13 April 2016, and revised for publication. Original manuscript received for
review 23 June 2017. Revised manuscript received for review 1 February 2018. Paper peer approved 5 February 2018.

2202 December 2018 SPE Journal

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2203 Total Pages: 16

alkoxylates used to synthesize the Gemini surfactants were supplied by the same manufacturer, while the C12–13 13PO-sulfate, C15–18
internal olefin sulfonate (IOS), C19–23 IOS, C20–24 IOS, and C19–28 IOS surfactants were obtained from two different manufacturing
facilities. Structures of the new surfactants and cosolvents used in this study are listed in Figs. 1A through 1I.
Cosolvents. IBA and triethylene glycol monobutyl ether (TEGBE) (Fig. 1K), diisopropylamine (DIPA) ethoxylate, and Phenol
alkoxylates and IBA alkoxylates were supplied by various industry manufacturers. The new phenol alkoxylate and IBA alkoxylate
cosolvent structures are shown in Figs. 1a and 1b, The addition of PO groups makes the structure more branched, which greatly mini-
mizes the formation of viscous macroemulsions. Moreover, the cosolvent properties can be easily tailored by changing the number of
POs or EOs to match different crude oils and reservoir conditions such as temperature and salinity. The cost to manufacture these
cosolvents is competitive with commonly used conventional cosolvents such as IBA, ethylene glycol monobutyl ether (EGBE), and TEGBE.
Polymers. Flopaam 3330S and 3630S polymers were provided by an overseas manufacturer. The chemical structures of Flopaam
3330S (8 million daltons) and Flopaam 3630S (18 million daltons) copolymers are shown in Fig. 1J.
Oils. The properties of oils used in this study are summarized in Table 1.
Electrolytes and Brines. Sodium chloride (NaCl), sodium carbonate (Na2CO3), and synthetic brine components were used as
received from the manufacturer. Brine compositions are summarized in Table 2.

Procedures. Microemulsion-Phase-Behavior Method. Surfactant and cosolvent formulations were evaluated using phase-behavior
tests. The detailed experimental procedures can be found in Levitt et al. (2009) and Flaaten et al. (2009). The microemulsion-phase-
behavior samples were made in borosilicate glass pipettes by injecting known volumes of oil, brine, surfactants, and cosolvents. Chelat-
ing agents might be added to the formulation when the brine has substantial divalent-cation content. The oil concentration used for
phase-behavior-salinity scans is typically 30 vol%. The pipettes were sealed using an open flame and then stored in an oven at the
experimental temperature. The pipettes were mixed vigorously and frequently to ensure good mixing and faster equilibration afterward.
The phase volumes were observed during and after equilibration. Solubilization ratios (r) are the volume of oil or water divided by the
volume of surfactant in the microemulsion. The IFT was not measured but can be estimated using the Huh (1979) equation:
C
c¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
r2
where c is IFT and C is approximately 0.3 mN/m. The surfactant formulations that produced ultralow IFT and other desirable behavior
such as fast equilibration, low microemulsion viscosity, and aqueous stability equal to or greater than the optimal salinity were selected
for coreflood experiments.
Microemulsion-Viscosity Measurements. The salinity range of interest for a given surfactant formulation was determined by first
evaluating the phase behavior. The microemulsion samples used for viscosity measurements were prepared in large-volume laboratory-
grade glass tubes at and around the optimal salinity. After the samples were fully equilibrated at the reservoir temperature, the micro-
emulsion phase was carefully extracted from the tubes. TA Instruments Advanced Rheometric Expansion System LS-1 rheometer was
used for shear-viscosity measurements of microemulsion samples at the corresponding reservoir temperatures (Walker et al. 2012;
Tagavifar et al. 2017).
Coreflood Procedure. All corefloods reported in this study were performed in either Berea or Bentheimer sandstone cores. The min-
eralogy was assessed by X-ray diffraction to identify the clays present in the cores. Bentheimer sandstone contained approximately
1.5 wt% total clay (mostly kaolinite), and the Berea sandstone contained approximately 12 wt% clays, including highly expansive
smectite clays. The cores were reduced before chemical flood using the solution of sodium dithionite, sodium bicarbonate, and tetraso-
dium ethylenediaminetetraacetate (Na4EDTA). The detailed procedure is reported by Rajapaksha et al. (2014). Before injecting the
chemical solutions, the cores were flooded with brine, oil, and then waterflood brine until the residual oil saturation was reached. All
the corefloods were conducted vertically at reservoir temperature. Coreflood procedures can be found in Jang et al. (2016). The oil
recovery, surfactant retention, and pressure drop, as well as the pH, salinity, and viscosity of effluent, were measured to evaluate the
chemical-flood performance. The procedure for measuring the surfactant and cosolvent retentions using high-performance liquid
chromatography (HPLC) can be found in Solairaj et al. (2012).

Results and Discussion


Phenol Alkoxylates. The new phenol alkoxylate cosolvents were evaluated with different crude oils (Oils A, B, and C). In the surfac-
tant formulations tested, these new cosolvents were able to replace conventional cosolvents with equal or lower concentration and still
provide excellent phase behavior and microemulsion viscosity, which means an overall reduction of chemical cost.
Oil A. Oil A is a light crude oil with a low total acid number (TAN) of less than 0.1 mg potassium hydroxide (KOH)/g oil, 28  API
and a viscosity of 12 cp at 45 C (see oil properties in Table 1). We refer to it as an inactive oil because it did not react with akali. Three
different cosolvents were tested with Oil A using the same surfactants. The surfactant formulation was 0.5 wt% C13-13PO-sulfate,
0.5 wt% C20–24 IOS, and either 1 wt% TEGBE (Formulation A1), 1 wt% phenol-4EO (Formulation A2), or 0.5 wt% phenol-1PO-5EO
(Formulation A3) cosolvent. The microemulsion-phase-behavior results are shown in Table 3 and Fig. 2.
A salinity scan using Formulation A1 was performed by adding Na2CO3 to 5,000 ppm NaCl brine. As shown in Fig. 2, microemul-
sion samples near optimal salinity [29,500 ppm total dissolved solids (TDS)] had a solubilization ratio of 24 and ultralow IFT. The
aqueous solution was clear up to a salinity of 35,000 ppm TDS. However, above optimal salinity in the Type II region, equilibration
was slow, macroemulsions were observed, and the interfaces were not fluid.
Next, the 1 wt% TEGBE cosolvent was replaced by 1 wt% phenol ethoxylate cosolvent while keeping the surfactants the same. The
phase-behavior results are shown in Fig. 2. The optimal salinity increased to approximately 40,000 ppm TDS and the optimal solubili-
zation ratio decreased to 16. The aqueous stability for the new formulation was approximately 42,000 ppm TDS. The phase behavior
was improved in terms of equilibration in the Type II region; gentle mixing of the equilibrium microemulsion samples showed very
fluid interfaces and no viscous macroemulsions were observed. The viscosities of Type I, III, and II microemulsions were measured at
different shear rates and the results are shown in Fig. 3a. The Type III microemulsion viscosity at optimal salinity (40,000 ppm TDS)
was 25 cp (approximately twice the oil viscosity), and the Type II microemulsion viscosity at a higher salinity (55,000 ppm TDS) was
approximately 35 cp. Both Type III and Type II microemulsions are Newtonian at low shear rates. Thus, replacing TEGBE with
phenol-4EO significantly improved the phase behavior and reduced the microemulsion viscosity. However, reducing the phenol-4EO
concentration from 1 to 0.5 wt% resulted in viscous macroemulsions and long equilibration times. In particular, Type II samples did not
reach phase equilibrium even after 6 months.

December 2018 SPE Journal 2203

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2204 Total Pages: 16

H
O O X
O H B N
O O H
X
C X
A

O H
O O
X
O
O H E
O
X
D
O
O OH
a
b
O

c
O
OH
O S O–Na+
d
O
O
F
X

CH2 O C C17H34
a H3C CH2 CH CH CH2 CH2 OH
7 7
CH O CO C17H34
b,c,d
CH2 O C C17H34

O O

O O S O–Na+
H3C H2C 7 CH2 O O
G 8 Y O
X

OH

a,b,c
O

O O S O–Na+
O O
H X Y
O

OH OH
CH2 CH CH2 CH
O O a
O O m n
+
HO m OH C16H33 O C16H33 C=O C=O
J

C16H33 NH2 O Na
b,c

HO O
O O
O O
+Na–O S O O S O–Na+ K
O O y
O O
I O O
C16H33 O C16H33

Fig. 1—Structures of (A) IBA-XEO cosolvent; (B) phenol-XEO cosolvent; (C) DIPA-XEO cosolvent; (D) IBA-1PO-XEO cosolvent;
(E) phenol-1PO-XEO cosolvent; (F) synthesis of 2-ethylhexanol-XPO-sulfate from butanal, (a) aldol condensation, (b) dehydrogena-
tion, (c) hydrogenation, (d) PO addition and sulfation; (G) synthesis of C18-XPO-YEO-sulfate, (a) hydrolysis, and hydrogenation of
triglyceride (b, c, d) propoxylation, ethoxylation followed by sulfation reaction; (H) synthesis of C12–13-XPO-YEO-sulfate, (a, b, c)
propoxylation, ethoxylation followed by sulfation reaction; (I) synthesis of C18-2EO-C18-10PO-2SO4, (a) KOH, 110–1308C, (b, c) pro-
poxylation, followed by sulfation reaction; (J) chemical structure of Flopaam 3330S and 3630S; (K) structure of TEGBE.

2204 December 2018 SPE Journal

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2205 Total Pages: 16

Oil
A B C D E F G
Temperature (°C) 45 28 68 30 55 24 60
°API gravity 28 33.8 25 18.9 36 30 30
Viscosity at reservoir temperature (cp) 12 8 8 195 3 9 5
TAN (mg KOH/g oil) <0.1 <0.1 <0.1 1.9 <0.1 <0.1 <0.1

Table 1—Oil properties.

Oil
Ionic Species A B C D E F G
+
Na 1,966 126 8,520 – 539 202 1,966
+
K – – 66 – – – –
2+
Ca – – 60 – – – –
2+
Mg – – 58 – – – –

Cl 3,034 99 13,447 – 208 113 3,034

HCO3 – 96 - – 132 342 -
2–
SO4 – 20 56 – 740 – –
Total 5,000 341 22,207 – 1,619 657 5,000

Table 2—Electrolyte composition of brines (in ppm).

S* ASL
(ppm σ* (ppm μME,S*
3 3
Formulations Surfactants Cosolvents TDS) (cm /cm ) TDS) (cp) Observations
Oil A
A1 1 wt% TEGBE 29,500 24 35,000 - Slow equilibration
0.5 wt% C13-
A2 1 wt% phenol-4EO 40,000 16 42,000 25 Faster equilibration
13PO-SO4; 0.5
wt% C20–24 IOS 0.5 wt% phenol- Reduced cosolvent use;
A3 42,000 18 45,800 25
1PO-5EO faster equilibration than A2
Oil B
Large TEGBE
B1 2 wt% TEGBE 31,000 11 32,850 -
concentration
0.5 wt% C13-
Improved solubilization
B2 13PO-SO4; 0.5 2 wt% phenol-2EO 31,000 19 32,000 -
ratio
wt% C20–24 IOS
1 wt% phenol-
B3 20,000 21 23,000 - Reduced cosolvent use
1PO-2EO
Oil C
0.4 wt% C18-
C1 45PO-30EO- 0.25 wt% phenol- Low Newtonian
52,000 23 63,800 16
(Coreflood 1) COO 0.6 wt% 1PO-2EO microemulsion viscosity
C15–28 IOS

Table 3—Chemical formulations with phenol alkoxylate cosolvents. ASL 5 aqueous-stability limit.

Using phenol-1PO-5EO cosolvent (Formulation A3) further improved the microemulsion-phase behavior. The phase-behavior
results are presented in Fig. 2. The optimal salinity was 42,000 ppm TDS and the optimal solubilization ratio was 18 after 20 days. The
IFT was approximately the same or slightly lower compared with Formulation A2. The aqueous stability was 45,800 ppm TDS. The
higher optimal salinities with Formulations A2 (phenol-4EO) and A3 (phenol-1PO-5EO) compared with Formulation A1 is attributed
to stronger hydrophilicity with more EO groups than TEGBE, and also possibly with phenol possessing higher electron density in its
aromatic ring. Therefore, the selection of cosolvents is subject to the requirements of the specific hydrophilic and lipophilic balance
(HLB) and desirable microemulsion-phase behavior. More examples regarding the effect of HLB of cosolvents will be provided later
for an alkali/cosolvent/polymer (ACP) formulation for Oil D. As shown in Fig. 3b, the microemulsion-viscosity data at optimal salinity
was approximately 25 cp and showed Newtonian behavior. The insertion of one PO group into the cosolvent and the resulting
branched structure resulted in superior performance compared with 0.5% phenol-4EO. The branching in the cosolvent resulted in
more-favorable microemulsion structures with fast equilibration time and favorable rheology and helped prevent the formation of
viscous macroemulsions.
Oil B. Oil B is an inactive light oil with low TAN of less than 0.1 mg KOH/g oil and viscosity of 8 cp at 28 C (Table 1). The surfac-
tant formulations for Oil B contained 0.5 wt% C13-13PO-sulfate and 0.5 wt% C20–24 IOS, which are the same surfactants used with Oil
A (Table 3). Salinity scans using Na2CO3 in synthetic soft brine (Table 2) were performed with 2 wt% TEGBE (Formulation B1),
2 wt% phenol-2EO (Formulation B2), and 1 wt% phenol-1PO-2EO (Formulation B3). The optimal salinity of Formulation B1 was
approximately 31,000 ppm TDS and the optimal solubilization ratio was approximately 11 after 31 days. The aqueous solution was
clear up to a salinity of 32,850 ppm TDS. The first comparison was made with the same amount of phenol-2EO (Formulation B2). The

December 2018 SPE Journal 2205

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2206 Total Pages: 16

phase-behavior results with 30 vol% oil are shown in Fig. 4. The optimal salinity in this case remained the same as with 2 wt% TEGBE
and the optimal solubilization ratio increased to 19 after 32 days. The aqueous solution was clear up to approximately 32,000 ppm
TDS. Therefore, the most notable improvement obtained with phenol-2EO is increased solubilization ratio, which indicates reduced
IFT. When the concentration of phenol-2EO was reduced from 2 to 1 wt%, the microemulsion-phase behavior showed a slower equili-
bration and viscous macroemulsion formation at high salinities. The phase-behavior results for Formulation B3 with 1 wt% phenol-
1PO-2EO are shown in Fig. 4. The optimal salinity decreased to approximately 20,000 ppm TDS and the optimal solubilization ratio
was 21 after 35 days. The aqueous stability was approximately 23,000 ppm TDS. The 1 wt% phenol-1PO-2EO cosolvent further
improved the phase behavior with ultralow IFT, shorter equilibration time, and no viscous-macroemulsion formation. The introduction
of phenol-1PO-2EO both improved the phase behavior and reduced the cosolvent usage by 50%.

60
Oil, A1 Water, A1
Solubilization Ratio (cm3/cm3)
Oil, A2 Water, A2
50 Oil, A3 Water, A3

40

30

20

10

0
0 10 20 30 40 50 60
Salinity (×1,000 ppm TDS)

Fig. 2—Phase behavior of 0.5 wt% C13-13PO-sulfate and 0.5 wt% C20–24 IOS with 1 wt% TEGBE (Formulation A1; green diamonds
and dashed lines), with 1 wt% phenol-4EO (Formulation A2; blue squares and dotted lines), and with 0.5 wt% phenol-1PO-5EO
(orange circles and solid lines) for Oil A at 458C.

1,000 1,000
Type I – 30,000 ppm TDS Type I – 30,667 ppm TDS
Type III – 40,000 ppm TDS Type III – 40,000 ppm TDS
Type II – 55,000 ppm TDS Type III – 43,000 ppm TDS
100 100 Type II – 60,000 ppm TDS
Viscosity (cp)

Viscosity (cp)

10 μo ~ 10 μo ∼
12 cp 12 cp

1 1

0.1 0.1
0.1 1 10 100 1,000 0.1 1 10 100 1,000
Shear Rate (s–1) Shear Rate (s–1)
(a) (b)

Fig. 3—Microemulsion viscosities of (a) Formulation A2 (with 1 wt% phenol-4EO) and (b) Formulation A3 (with 0.5 wt% phenol-
1PO-5EO) at 458C.

Oil C. Oil C is an inactive oil with low TAN of less than 0.1 mg KOH/g oil and viscosity of 8 cp at 68 C. Surfactant Formulation C
consisted of 0.4 wt% C18(oleyl)-45PO-30EO-carboxylate, 0.6 wt% C15–28 IOS (equal mass of C15–18 IOS, C19–23 IOS, and C19–28 IOS),
and 0.25 wt% phenol-1PO-2EO cosolvent. The salinity scan was performed by adding Na2CO3 to the synthetic formation brine that has
approximately 22,200 ppm TDS. Because the brine contained a total of 118 ppm Ca2þ and Mg2þ ions, 0.36 wt% Na4EDTA was added as
a chelating agent to prevent precipitation. Fig. 5 shows the phase-behavior results for Formulation C. The optimal salinity with 30 vol%
oil was 52,000 ppm TDS. The solubilization ratio at optimal salinity was 23 after 30 days. The aqueous-surfactant solution with
3,000 ppm Flopaam 3630S polymer was clear up to 63,800 ppm TDS. The microemulsion viscosity at optimal salinity was Newtonian
with a viscosity of 16 cp (approximately twice the oil viscosity) at 68 C. More experimental results using oleyl alkoxy sulfate and carbox-
ylate surfactants can be found in Liyanage et al. (2015b). Oleyl alcohol is produced from two naturally occurring triglycerides, palm oil
and tallow; the unsaturation marked by the olefinic double bond is an important precursor to biodegradable detergents. The olefin func-
tionality at the midpoint (ninth carbon atom) gives oleyl alcohol a bent structure. As is common with all natural oils, the oleyl segment has
a cis-configuration at the point of unsaturation, which makes the entire molecule appear to behave like a large hydrophobe, compared with
its linear analog.

2206 December 2018 SPE Journal

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2207 Total Pages: 16

60
Oil, B2 Water, B2
Oil, B3 Water, B3

Solubilization Ratio (cm3/cm3)


50

40

30

20

10

0
0 10 20 30 40 50 60
Salinity (×1,000 ppm TDS)

Fig. 4—Phase behavior of 0.5 wt% C13-13PO-sulfate and 0.5 wt% C20–24 IOS with 2 wt% phenol-2EO (Formulation B2; green dia-
monds and dashed lines) and with 1 wt% phenol-1PO-2EO (Formulation B3; orange circles and solid lines) for Oil B at 288C.

60
Oil, C1
Water, C1
50
Solubilization Ratio (cm3/cm3)

40

30

20

10

0
30 40 50 60 70
Salinity (×1,000 ppm TDS)

Fig. 5—Phase behavior of 0.4 wt% C18-45PO-30EO-carboxylate, 0.2 wt% C15–18 IOS, 0.2 wt% C19–23 IOS, 0.2 wt% C19–28 IOS, and
0.25 wt% phenol-1PO-2EO with Oil C at 688C.

These novel phenol alkoxylate cosolvents were tested in coreflood experiments with very promising results with high oil recovery
and low surfactant and cosolvent retentions in Bentheimer sandstone cores. The coreflood results are summarized in Table 4. A
0.30-pore-volume (PV) ASP slug was injected at 1 ft/D. The ASP slug was followed by a polymer drive with lower salinity in the Type
I region. The tertiary oil recovery was 90% of the waterflood residual oil and the final oil saturation after the chemical flood was 0.043.
The results are shown in Fig. 6. The surfactant and cosolvent retentions were measured by HPLC, and results are shown in Fig. 7. The
total surfactant retention was 0.072 mg/g rock, and the phenol-1PO-2EO cosolvent retention was 0.027 mg/g rock. No chromatographic
separation of surfactants and cosolvent was observed in the coreflood.
IBA Alkoxylates. For Oils D and E, IBA alkoxylate cosolvents showed good phase behavior and microemulsion fluidity in prelimi-
nary screening tests. The IBA alkoxylate cosolvents were tested with Oils D and E under variety of reservoir conditions. Similar to the
phenol alkoxylates discussed previously, these new IBA-based cosolvents also outperformed the conventional cosolvents. Chang et al.
(2016) studied cosolvent partitioning as a function of the EO content of cosolvents. The ideal cosolvent partitions equally between oil
and water. The cosolvents tested in this study were selected in part because of the nearly equal affinity for the oil and water at the salin-
ity and temperature of the various tests.
Oil D. Oil D is a viscous oil of 18.9  API and with viscosity of 195 cp at 30 C. Oil D is an active oil with a TAN of 1.9 mg KOH/g
oil and produces soap in the presence of alkali such as Na2CO3. Therefore, a new process known as ACP flooding (Fortenberry et al.
2015) that does not require any synthetic surfactant was tested.
IBA, IBA-3EO, IBA-10EO, and IBA-30EO cosolvents were tested in ACP formulations with Oil D. The effect of the number of
ethoxy groups on the microemulsion-phase behavior was observed from Na2CO3 scans using 1 wt% of each cosolvent. Table 5 summa-
rizes the phase-behavior results for 10, 30, and 50 vol% oil. The phase-behavior results with 1 wt% IBA cosolvent and 30 vol% oil
showed slow equilibration and formation of viscous macroemulsions. The optimal salinity was 10,000 ppm TDS. By increasing the
number of EOs to 30, the equilibration time decreased from 40 to 15 days and no macroemulsions were observed. Because the HLB of
natural soaps and cosolvents are not the same, the microemulsion-phase behavior depends on oil concentration. A plot of phase type vs.
oil concentration is known as an activity diagram. Examples can be found in Yang et al. (2010), Lu et al. (2014), and Fortenberry et al.
(2015). The activity diagram obtained with IBA-30EO is shown in Fig. 8. The small change in optimal salinity with oil concentration is
considered favorable for a robust ACP design. As shown in Table 5, the change in optimal salinity with oil concentration was greater

December 2018 SPE Journal 2207

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2208 Total Pages: 16

for the cosolvents with fewer EOs. Thus, the HLB for IBA-30EO and the soap generated from Oil D are similar compared with IBA-3EO
and IBA-10EO cosolvents.

Coreflood No.
1 2 3 4
Bentheimer Bentheimer Berea Bentheimer
Rock
sandstone sandstone sandstone sandstone
Temperature (°C) 68 25 25 60
Brine permeability (md) 2,400 2,373 148 2,266
3
PV (cm ) 78.6 134.0 121.0 81.8
Porosity 0.235 0.234 0.200 0.250
Initial oil saturation, Soi 0.77 0.784 0.569 0.67
Waterflood residual oil saturation, Sorw 0.45 0.420 0.30 0.29
Crude oil C F F G
ASP Slug
Formation-brine salinity (ppm) 22,206 66,000 66,000 50,000
Formulation C1 F2 F2 G2
Surfactant concentration (wt%) 1.0 0.75 0.75 1.0
Cosolvent concentration (wt%) 0.25 – – 1.0
PV injected 0.3 0.4 0.4 0.3
PV×C (%) 30 30 30 30
Polymer FP 3630S FP 3630S FP 3330S FP 3630S
Polymer concentration (ppm) 2,000 2,000 3,200 2,000
Salinity (ppm TDS) 61,656 58,180 58,180 40,500
–1
Viscosity (cp at 10 s ) 24.0 44 43.5 20.4
Velocity (ft/D) 1.0 1.0 1.0 1.0
Polymer Drive 1st 2nd
Polymer FP 3630S FP 3630S FP 3330S FP 3630S
Polymer concentration (ppm) 2,000 2,000 2,000 3,200 1,800
Salinity (ppm TDS) 61,656 23,456 45,680 45,680 28,500
–1
Viscosity (cp at 10 s ) 24 24 44 41 16.6
Velocity (ft/D) 1.0 1.0 1.0 1.0
Results
Oil recovery (%) 90.0 97.0 97.0 91.3
Final residual oil saturation, Sorc 0.043 0.013 0.009 0.025
Surfactant retention (mg/g) 0.072 0.082 0.11 0.10
Cosolvent retention (mg/g) 0.027 (Ph-1PO-2EO) 0.017 (2-EHS) 0.034 (2-EHS) –

Table 4—Summary of corefloods.

0.3 PV 0.2 PV
Polymer Drive 2
ASP Slug PD 1
100% 50%
Oil Recovery (%) Oil Cut (%)

80% Cumulative oil recovery 40%


Oil Saturation (%)

Oil cut
So
60% 30%

40% 20%

20% 10%

0% 0%
0.0 0.5 1.0 1.5
PV

Fig. 6—Oil recovery for Coreflood 1 with Oil C at 688C.

2208 December 2018 SPE Journal

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2209 Total Pages: 16

100 1
Surfactant injected = 0.37 mg/g of rock Oil cut
Surfactant retention = 0.072 mg/g of rock Surfactant
Cosolvent retention = 0.027 mg/g of rock Cosolvent
80 0.8

Oil Cut (%)


60 0.6

Cp /Ci
40 0.4

20 0.2

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
PV

Fig. 7—Effluent surfactant and cosolvent concentrations (Cp) relative to the injected concentration (Ci) for Coreflood 1.

Formulations Surfactants Cosolvents S*, 10 vol% oil S*, 30 vol% oil S*, 50 vol% oil Observations
Oil D (ACP formulation with active oil, TAN = 1.9 mg KOH/g oil)
Too-low optimal
D1 1 wt% IBA – 10,000 –
salinity
Steep slope in
D2 1 wt% IBA-3EO 20,000 35,000 45,000
activity diagram
Natural soap
Less-steep slope
D3 1 wt% IBA-10EO 35,000 47,500 50,000
than D2
Nearly flat slope,
D4 1 wt% IBA-30EO 50,000 55,000 60,000
fast equilibration

Table 5—ACP formulations with IBA alkoxylate cosolvents, S* in ppm TDS.

100,000

90,000
High IFT/Type II
Salinity (ppm TDS)

80,000

70,000

60,000 Ultra-low IFT

50,000
Low IFT/Type I
40,000

30,000
0 10 20 30 40 50 60
Oil Concentration (vol%)

Fig. 8—Activity diagram for Oil D with 1 wt% IBA-30EO at 308C after 15 days.

Oil E. Oil E is an inactive oil with low TAN of less than 0.1 mg KOH/g oil and viscosity of 3 cp at 55 C. Two formulations were
developed for Oil E with 0.4 wt% C13-7PO-sulfate, 0.4 wt% C19–23 IOS, and 0.2 wt% C18-35PO-10EO-sulfate surfactants, and
were tested with different IBA alkoxy cosolvents. Formulation E1 contained 1 wt% IBA-3EO and Formulation E2 contained 0.5 wt%
IBA-1PO-2EO cosolvent (Table 6). Salinity scans were performed by adding Na2CO3 in synthetic soft brine (see Table 2). The phase-
behavior results are shown in Fig. 9. The optimal salinity of Formulation E1 was 37,000 ppm TDS after 36 days, and the optimal solubi-
lization ratio was 17. The aqueous stability was 47,000 ppm TDS at 55 C. By replacing 1 wt% IBA-3EO with 0.5 wt% IBA-1PO-2EO,
the optimal salinity of Formulation E2 decreased to 33,000 ppm TDS, the optimal solubilization ratio increased from 17 to 24, and the
phase behavior showed faster equilibration within 2 to 3 days. The microemulsion-viscosity data are shown in Fig. 10. For Formulation
E1 with 1 wt% IBA-3EO, the microemulsion viscosity at optimal salinity was approximately 6 cp, or approximately twice the oil vis-
cosity, and Newtonian, which is extremely favorable for good transport in reservoirs at low-pressure gradients resulting in low surfac-
tant retention. For Formulation E2, the microemulsion viscosity at the optimal salinity is less favorable but still good compared with
most conventional formulations.

December 2018 SPE Journal 2209

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2210 Total Pages: 16

S* σ* ASL μME,S*
3 3
Formulations Surfactants Cosolvents (ppm TDS) (cm /cm ) (ppm TDS) (cp) Observations
Oil E
0.4 wt% Ultralow IFT at S*
C13-7PO-SO4 1 wt% Low Newtonian
E1 37,000 17 47,000 6
0.4 wt% IBA-3EO microemulsion
C19–23 IOS viscosity
Reduced cosolvent
0.2 wt% C18- 0.5 wt%
use with similar
E2 35PO-10EO- IBA-1PO- 33,000 24 37,000 8
properties of
SO4 2EO
microemulsion

Table 6—Chemical formulations with IBA alkoxylate cosolvents.

60
Oil, E1 Water, E1
Oil, E2 Water, E2
50
Solubilization Ratio (cm3/cm3)

40

30

20

10

0
10 20 30 40 50 60 70
Salinity (×1,000 ppm TDS)

Fig. 9—Phase behavior of 0.4 wt% C13-7PO-sulfate, 0.4 wt% C19–23 IOS, and 0.2 wt% C18-35PO-10EO-sulfate, with 1 wt% IBA-3EO
(Formulation E1) and with 0.5 wt% IBA-1PO-2EO (Formulation E2) for Oil E at 558C.

100 100
Tyep I – 30,000 ppm TDS Type I – 27,000 ppm TDS
Type III – 38,000 ppm TDS Type III – 32,000 ppm TDS
Type II – 57,000 ppm TDS Type II – 45,000 ppm TDS

10 10
Viscosity (cp)

Viscosity (cp)

μo ∼ μo ∼
3 cp 3 cp

1 1

0.1 0.1
0.1 1 10 100 1,000 0.1 1 10 100 1,000
Shear Rate (s–1) Shear Rate (s–1)
(a) (b)

Fig. 10—Microemulsion viscosity with (a) 1 wt% IBA-3EO (Formulation E1) and (b) 0.5 wt% IBA-1PO-2EO (Formulation E2), respec-
tively, with Oil E at 558C.

Novel Surfactants
Ultrashort Hydrophobe Surfactant 2-EHS. A novel class of extended ultrashort hydrophobe surfactant, 2-ethylhexanol-7PO-sulfate,
was developed in a commercially viable, cost-effective process in near-quantitative yield. The 2-ethylhexanol is a commodity chemical
and is produced by aldol condensation of n-butyraldehyde, followed by the hydrogenation of the resulting hydroxyaldehyde. It can also
be produced by dimerization of n-butanol in a Guerbet reaction. The PO and sulfate groups are combined with 2-ethylhexanol to create,
for example, 2-ethylhexanol-7PO-sulfate. The hydrophobe size of 2-EHS is smaller than that of common short hydrophobe surfactants
such as C13-7PO-sulfate but has substantially more-pronounced branching. Several surfactant formulations with 2-ethylhexanol-7PO-
sulfate indeed showed both surfactant as well as cosolvent properties. This makes it possible for 2-ethylhexanol-7PO-sulfate to replace
cosolvent without adversely affecting phase behavior, and thus reduces the chemical cost for CEOR processes. In this section, we show
the results for 2-ethylhexanol-7PO-sulfate surfactant without cosolvent in the surfactant formulations, and the corresponding
coreflood results.

2210 December 2018 SPE Journal

ID: jaganm Time: 13:53 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2211 Total Pages: 16

Oil F. Oil F is an inactive, light oil with low TAN of less than 0.1 mg KOH/g oil and viscosity of 9 cp at 24 C. Surfactant
Formulation F1 consisted of 0.55 wt% C13-13PO-sulfate, 0.2 wt% C20–24 IOS, and 0.75 wt% phenol-2EO cosolvent. The formulations
for Oil F are presented in Table 7. The optimal salinity for this formulation is 33,500 ppm TDS and the optimal solubilization ratio is
20 after 36 days (Fig. 11). The aqueous-surfactant solution was clear up to approximately 38,380 ppm TDS. The microemulsion viscos-
ity at the optimal salinity was 27 cp, but at a slightly higher salinity of 41,000 ppm TDS, the microemulsion viscosity was much higher
(approximately 60 cp) and showed shear-thinning behavior (Fig. 12b). Therefore, a new formulation using 2-ethylhexanol-7PO-sulfate
was developed. Formulation F2 consisted of 0.1 wt% C13-13PO-sulfate, 0.2 wt% C20–24 IOS, 0.2 wt% C13-45PO-10EO-sulfate, and
0.25 wt% 2-ethylhexanol-7PO-sulfate (no cosolvent). The 0.2 wt% C13-45PO-10EO-sulfate was added to enhance the aqueous stability
of the new formulation. The optimal salinity with 30 vol% oil was 58,000 ppm TDS and the optimal solubilization ratio was 29 (com-
pared with 20 for Formulation F1). The aqueous stability was 63,400 ppm TDS. The phase-behavior results are shown in Fig. 11. The
microemulsion viscosity (Fig. 12b) is significantly lower than for Formulation F1 with cosolvent, the total chemical concentration of
Formulation F2 is lower, and the IFT is lower. Thus, using a short hydrophobe surfactant rather than cosolvent is highly advantageous
for this oil.

S*
(ppm ASL μME,S*
σ* (cm /cm )
3 3
Formulations Surfactants Cosolvents TDS) (ppm TDS) (cp) Observations
Oils F and A (short hydrophobe 2-EHS surfactant)
Approximately
50 cp
0.75 wt%
0.55 wt% C13-13PO-SO4 33,500 20 38,380 27 microemulsion
F1 phenol-2EO
0.2 wt% C20–24 IOS – – – – viscosity at

41,000 ppm
TDS
0.1 wt% C13-13PO-SO4 – – – – – No cosolvent
0.2 wt% C20–24 IOS uses near-
F2 – – – – –
0.2 wt% C13-45PO-10EO- Newtonian
(Corefloods – 58,000 29 63,400 17
SO4 behavior at low
2 and 3)
0.25 wt% 2-ethylhexanol- shear rates less
7PO-SO4 – – – – – than10 s
–1

– – – – – Similar
0.3 wt% C13-45PO-10EO-
46,000 24 50,800 25 microemulsion-
sulfate
– phase behavior
A4 0.3 wt% C20–24 IOS
and viscosity
0.4 wt% 2-ethylhexanol- – – – – with less total
7PO-SO4 – chemical use
Oil G (Propoxylated Gemini surfactant)
– – – – Favorable
0.5 wt% C12–13-7PO-SO4
G1 microemulsion-
0.5 wt% C20–24 IOS 27,000 17 30,000 – phase behavior
1 wt%
DIPA-5EO – – – – Improved
0.5 wt% C12–13-7PO-sulfate
G2 microemulsion-
0.5 wt% C18-2EO-C18- 32,000 25 40,300 –
(Coreflood 4) phase behavior
10PO-disulfate
over G1

Table 7—Chemical formulations with new surfactants.

70
Oil, F1 Oil, F2
60 Water, F1 Water, F2
Solubilization Ratio (cm3/cm3)

50

40

30

20

10

0
10 20 30 40 50 60 70
Salinity (×1,000 ppm TDS)

Fig. 11—Phase behavior of 0.55 wt% C13-13PO-sulfate, 0.2 wt% C20–24 IOS, and 0.75 wt% phenol-2EO (Formulation F1; green dia-
monds and dashed lines) and 0.1 wt% C13-13PO-sulfate, 0.2 wt% C20–24 IOS, 0.2 wt% C13-45PO-10EO-sulfate, and 0.25% 2-ethylhex-
anol-7PO-sulfate (Formulation F2; orange circles and solid lines) with oil F at 248C. Photograph on the right is for Formulation F2.

December 2018 SPE Journal 2211

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2212 Total Pages: 16

100 100

10 10
μo ∼

Viscosity (cp)

Viscosity (cp)
μo ∼
9 cp
9 cp

1 1
Type I – 55,900 ppm TDS
Type I – 32,000 ppm
Type III – 57,500 ppm TDS
Type III – 38,495 ppm
Type II – 62,500 ppm TDS
Type II – 41,000 ppm
Type II – 44,000 ppm Type II – 70,000 ppm TDS
0.1 0.1
0.1 1 10 100 1,000 0.1 1 10 100 1,000
Shear Rate (s–1) Shear Rate (s–1)
(a) (b)

Fig. 12—Microemulsion-viscosity data for Oil F at 248C: (a) Formulation F1, (b) Formulation F2.

Coreflood 2 was performed in a Bentheimer sandstone core to evaluate Formulation F2 (Table 4). A 0.40-PV ASP slug with
0.75 wt% surfactant was injected. The oil-recovery results and effluent surfactant concentration measured by HPLC are shown in
Fig. 13. The tertiary oil recovery was 97% of the waterflood residual oil, and the final oil saturation was 0.013. The pressure gradient
during the flood was low and favorable (less than 1 psi/ft) in part because of the low microemulsion viscosity. The total surfactant reten-
tion was 0.082 mg/g rock (76% of the injected surfactant recovered). The 2-ethylhexanol-7PO-sulfate retention was 0.017 mg/g rock
(85% recovery). No chromatographic separation of 2-ethylhexanol-7PO-sulfate from the other surfactants was detected within
measurement uncertainty.

0.4 PV
Polymer Drive
ASP Slug
100 1.0
Total surfactant retention = 0.082 mg/g
Total surfactant injected = 0.347 mg/g Cumulative oil
Oil Recovery (%) Oil Cut (%) So (%)

Total 2-EHS retention = 0.017 mg/g Oil cut


Total 2-EHS injected = 0.116 mg/g
So
80 Total surfactant 0.8
2-EHS surfactant

60 0.6

Cp /Ci
40 0.4

20 0.2

0 0.0
0.0 0.5 1.0 1.5 2.0
PV

Fig. 13—Oil-recovery results and effluent surfactant concentrations (Cp) relative to the injected concentrations (Ci) for Bentheimer
sandstone Coreflood 2 with Oil F at 248C.

Formulation F2 was also tested in a Berea sandstone core with a high clay content (approximately 12 wt% clay). The tertiary oil
recovery was 97% of the waterflood residual oil, with a final oil saturation of 0.009 (Fig. 14). Surfactant retention was 0.11 mg/g rock
with 64% of the injected surfactant recovered. The retention of 2-EHS surfactant was 0.034 mg/g rock. Even though the clay content is
approximately eight times higher in the Berea core, the surfactant retention was only slightly higher than in the Bentheimer sandstone.
There was no indication of preferential retention of surfactant in the core. This very favorable behavior of the formulation containing
2-ethylhexanol-7PO-sulfate is in part because of the favorable microemulsion-phase behavior and rheology, which correlate with low
surfactant retention. The surfactant blend appears to be synergistic and not require cosolvent for favorable behavior.
Oil A With 2-EHS. The 2-ethylhexanol-7PO-sulfate was also used to eliminate cosolvent from the formulation shown in Fig. 2
with Oil A. In the new Formulation A4, both the sulfate and sulfonate concentrations were reduced and 2-EHS-7PO-sulfate replaced
the cosolvent. In place of C13-13PO-sulfate, C13-45PO-10EO-sulfate was used to enhance the aqueous stability. The new formulation
consisted of 0.3 wt% C13-45PO-10EO-sulfate, 0.3 wt% C20–24 IOS, and 0.4 wt% 2-ethylhexanol-7PO-sulfate. Fig. 15a shows that the
optimal salinity is 46,000 ppm TDS and the optimal solubilization ratio is 24. The aqueous stability is 50,800 ppm TDS at 45 C. The
microemulsion viscosity at optimal salinity is approximately 25 cp (Fig. 15b), which is similar to that of Formulation A3 shown in
Fig. 3b. Thus, desirable phase behavior and microemulsion viscosity can be achieved without the use of cosolvent for Oil A if 2-ethyl-
hexanol-7PO-sulfate is used in the formulation.

2212 December 2018 SPE Journal

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2213 Total Pages: 16

0.4 PV
Polymer Drive
ASP Slug
100 1.0
Cumulative oil

Oil Recovery (%) Oil Cut (%) So (%)


Oil cut
So Total surfactant retention = 0.111 mg/g
80 Total surfactant Total surfactant injected = 0.313 mg/g 0.8
2-EHS surfactant Total 2-EHS retention = 0.034 mg/g
Total 2-EHS injected = 0.104 mg/g

60 0.6

Cp /Ci
40 0.4

20 0.2

0 0.0
0.0 0.5 1.0 1.5 2.0
PV

Fig. 14—Oil-recovery results and effluent surfactant concentrations (Cp) relative to the injected concentration (Ci) for Berea sand-
stone Coreflood 3 with Oil F at 248C.

50 100
Oil, A4
Solubilization Ratio (cm3/cm3)

Water, A4
40

10
μo ∼
Viscosity (cp)

30
12 cp

20
1

10 Type I – 38,000 ppm TDS


Type III – 45,000 ppm TDS
Type II – 60,600 ppm TDS
0 0.1
20 30 40 50 60 70 0.1 1 10 100 1,000
Salinity (×1,000 ppm TDS) Shear Rate (s–1)
(a) (b)

Fig. 15—(a) Phase behavior of Formulation A4, with 0.3 wt% C13-45PO-10EO-sulfate, 0.3 wt% C20–24 IOS, and 0.4 wt% 2-ethylhexa-
nol-7PO-sulfate. (b) Microemulsion viscosity of Formulation A4 with Oil A at 458C.

Propoxylated Gemini Surfactants. Propoxylated anionic Gemini surfactants are composed of two hydrophobes, a diethyleneglycol-
linking spacer group with two PO chains, and two anionic groups, as shown in Fig. 1I. With two hydrophilic head groups, the propoxy-
lated anionic Gemini surfactants show high water solubility in high-salinity brines. This highly branched Gemini surfactant was used in
combination with the new DIPA-5EO cosolvent. The special feature of this cosolvent structure (Fig. 1), in addition to the two-isopropyl
branching, is the presence of the amine functionality that can impart alkalinity. With two isopropyl groups constituting the hydrophobe,
this structure has a high degree of branching in a short hydrophobe.
Oil G. Oil G is an inactive light oil with low TAN of less than 0.1 mg KOH/g oil and viscosity of 5 cp at 60 C. Initial Formulation
G1 developed for this oil consisted of 0.5 wt% C20–24 IOS, 0.5 wt% C12–13-7PO-sulfate, and 1 wt% DIPA-5EO. A Na2CO3 scan in
5,000 ppm NaCl brine showed an optimal salinity of 27,000 ppm TDS and optimal solubilization ratio of 17. The aqueous stability is
30,000 ppm TDS. C20–24 IOS in Formulation G1 was replaced with C18-2EO-C18-10PO-disulfate while keeping other components the
same. New Formulation G2 is 0.5 wt% C18-2EO-C18-10PO-disulfate, 0.5 wt% C12–13-7PO-sulfate, and 1 wt% DIPA-5EO. The phase-
behavior results are shown in Fig. 16. The optimal salinity was 32,000 ppm TDS and the optimal solubilization ratio increased to 25.
The aqueous stability is 40,300 ppm TDS.
Formulation G2 was tested in a Bentheimer sandstone (Coreflood 4). An ASP slug of 0.30 PV with 1 wt% surfactant was injected,
followed by a polymer drive at a lower Type I salinity. The oil-recovery results and the effluent surfactant-concentration data are pre-
sented in Fig. 17. The tertiary oil recovery was 91.3% of the waterflood residual oil, and the final oil saturation after the chemical flood
was 0.025. The total surfactant retention was 0.10 mg/g rock. The C18-2EO-C18-10PO-disulfate retention was 0.04 mg/g rock, and the
C12–13-7PO-sulfate retention was 0.06 mg/g rock. Within measurement uncertainty, there was no evidence for chromatographic separa-
tion in the core.

Synergy and Stability of Surfactants. Alkoxy sulfate surfactants are thought to have poor hydrolytic stability at temperatures greater
than 65 C (Talley 1988). However, recent research has shown that enhanced stability can be achieved in the pH range of 10 to 11
(Adkins et al. 2010). We have studied the synergistic interaction of alkoxy sulfates with IOS surfactants and have found that these inter-
actions have a positive effect on the stability of the alkoxy sulfates. The stability of PO-sulfates and PO-EO-sulfates in the presence of

December 2018 SPE Journal 2213

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2214 Total Pages: 16

1 wt% Na2CO3 was studied by measuring the surfactant activity by titration with cationic hyamine surfactant. These experimental
results show that PO-EO-sulfates in combination with a sulfonate such as an IOS retained more than 90% of their anionic activity at
100 C for 250 days (Fig. 18), whereas the PO-sulfates mixed with IOS retained only 70% anionic activity at 100 C after 200 days. At
80 C, the PO-sulfates mixed with the IOS retained more than 90% activity for 200 days (Fig. 19). These results indicate that there is a
synergistic interaction between the surfactants that increases their stability.

60
Oil, G1 Oil, G2
Water, G1 Water, G2

Solubilization Ratio (cm3/cm3)


50

40

30

20

10

0
5 15 25 35 45 55
Salinity (×1,000 ppm TDS)

Fig. 16—Phase behavior of Formulation G1 (green diamonds and dashed lines) consisting of 0.5 wt% C13-7PO-sulfate, 0.5 wt% C20–24
IOS, and 1 wt% DIPA-5EO, and Gemini surfactant Formulation G2 (orange circles and solid lines) consisting of 0.5 wt% C18-2EO-C18-
10PO-disulfate, 0.5 wt% C13-7PO-sulfate, and 1 wt% DIPA-5EO with Oil G at 608C.

0.3 PV
Polymer Drive
ASP Slug
100% 1.0
Oil Recovery (%) Oil Cut (%) So (%)

Cumulative oil
Oil cut
So Total surfactant retention = 0.10 mg/g
80% Gemini surfactant Gemini surfactant retention = 0.04 mg/g 0.8
C12-13-7PO-sulfate retention = 0.06 mg/g
TDA-7PO-sulfate

60% 0.6

Cp /Ci
40% 0.4

20% 0.2

0% 0.0
0.0 0.5 1.0 1.5
PV

Fig. 17— Oil-recovery results and effluent surfactant concentrations (Cp) relative to the injected concentrations (Ci) for Bentheimer
sandstone Coreflood 4 with Oil G at 608C.

1
Normalized Activity to Day 0

0.8

0.6

2-EHS + IOS
0.4
TDA-7PO-SO4 + IOS

0.2 TDA-35PO-30EO-SO4 + IOS

C18-35PO-30EO-SO4 + IOS
0
0 50 100 150 200 250 300
Days

Fig. 18—Anionic activity of alkoxy sulfate surfactants mixed with C15–18 IOS at 1008C with 1 wt% Na2CO3.

2214 December 2018 SPE Journal

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2215 Total Pages: 16

Normalized Activity to Day 0


0.8

2-EHS + IOS
0.6
TDA-7PO-SO4 + IOS

0.4

0.2

0
0 50 100 150 200 250
Days

Fig. 19—Anionic activity of 2-EHS and C12–13-7PO-sulfates mixed with C15–18 IOS at 808C with 1 wt% Na2CO3.

Conclusions
Novel phenol and IBA ethoxylate and propoxy-ethoxylate cosolvents were synthesized for EOR applications using a commercially
viable synthesis process. These new phenol alkoxylate and IBA alkoxylate cosolvents were tested in both ASP and ACP formulations.
The new cosolvents significantly improved microemulsion-phase behavior compared with classical cosolvents. The addition of one PO
group to the ethoxylated alcohol led to a remarkable improvement in the performance of the cosolvents and reduced the required
cosolvent concentration by a factor of two. Very favorable performance attributes such as ultralow IFT, good aqueous stability, and low
microemulsion viscosity with Newtonian behavior have been obtained using these phenol alkoxylate and IBA alkoxylate cosolvents.
Classical cosolvents such as simple alcohols increase the IFT more than the new cosolvents phenol-PO-xEO and IBA-PO-xEO. The
positive effect of adding one PO in the cosolvent molecule has now been clearly established. The favorable effect of the PO is caused
by the branched nature of the PO coupled with more-favorable interaction with some crude oils. The highly branched DIPA ethoxylate
has also been found to be a good cosolvent. A special feature of this molecule is its alkaline nature because of its amine functionality,
which means it could be used to increase the pH of formulations with the goal of reducing surfactant retention.
New surfactants including 2-ethylhexanol-7PO-sulfate, C12–13 alkoxy sulfate, oleyl alkoxy carboxylate, and Gemini propoxy disul-
fate were tested in this study. When incorporated in optimized formulations, these new surfactants produced ultralow IFT, clear aqueous
solutions at and above optimal salinity, and low microemulsion viscosity with both active and inactive oils over a wide range of temper-
ature. The use of Na2CO3 in the formulations both reduces the surfactant retention and increases the thermal stability of the sulfate sur-
factants. All these formulations showed excellent oil recovery and low surfactant retention in sandstone cores. Formulations containing
the ultrashort hydrophobe surfactant 2-ethylhexanol-7PO-sulfate showed excellent behavior without any cosolvent. Increasing the
number of PO units in short hydrophobe C12–13-PO-sulfate surfactants and balancing the extra hydrophobicity with EO units also
produces good EOR surfactants, such as C12–13-45PO-10EO-sulfate.
Formulations consisting of mixtures of surfactants with different structures consistently show the best performance and good thermal
stability. Furthermore, no chromatographic separation has been observed when using these surfactant mixtures in corefloods. Surfactant
formulations must be tailored to the specific crude oil and temperature. Adjusting the number of POs and EOs in the primary surfactant
is a very practical way to optimize surfactant formulations. The cost of such surfactants decreases as the EO content increases.
Optimized formulations developed using these new chemicals have shown excellent performance with low surfactant retention in
cores. Low retention means less chemicals can be used to recover each barrel of oil from the reservoir. These chemicals can be made
commercially at low costs. The current cost of producing the new surfactants on a commercial scale is estimated to be approximately
USD 1.50/lbm of active surfactant. For Formulation F2 without any cosolvent, the surfactant retention was 0.11 mg/g rock in the sand-
stone with 12 wt% clay. Jang et al. (2016) measured surfactant retentions as low as 0.025 mg/g rock in similar cores by adding cosol-
vent to the polymer drive, so a value of 0.11 mg/g rock is not the lowest value that can be obtained with current technology. For a
reservoir with a porosity of 0.2, only 0.15 PV of 0.75 wt% surfactant solution is needed to satisfy a surfactant retention of 0.11 mg/g
rock. Assuming 24% of the initial oil in place is recovered, which is typical of optimized ASP floods in sandstones, the cost of the sur-
factant would be approximately USD 2.5/bbl of incremental oil. This example is intended to emphasize the remarkable decrease in the
surfactant cost that has resulted from a combination of improved surfactant technology and lower manufacturing cost. There is no impli-
cation that the surfactant cost will always be so low, nor does the surfactant cost represent the total cost to produce the oil.

Nomenclature
S* ¼ optimum salinity
Sorc ¼ residual oil saturation to chemical
c ¼ interfacial tension
r ¼ solubilization ratio
lME,S* ¼ viscosity of microemulsion at optimal salinity

Acknowledgments
The authors would like to thank the industrial affiliates of the Chemical Enhanced Oil Recovery research project at the University of
Texas at Austin for financial support of this research. Special thanks go to Harcros Chemicals, Shell Chemical Company, Stepan Com-
pany, Taminco Chemicals, and SNF Floerger for supplying surfactants, cosolvents, and polymers used in this study. We would like to
thank Patrick Lim, Tanya Xu, Brien Allen, Austin Lim, Sumudu Herath, Arnob Bhuyan, Mohsen Tagavifar, Dharmika Lansakara,
Leonard Chang, Jun Lu, Geegana Wickramasiri, Willie Martin, and Travis Pitcher for helping with the laboratory experiments.

December 2018 SPE Journal 2215

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2216 Total Pages: 16

References
Abe, M., Schechter, D., Schechter, R. S. et al. 1986. Microemulsion Formation With Branched Tail Polyoxyethelene Sulfonate Surfactants. J. Colloid
Interfac. Sci. 114 (2): 342–356. https://doi.org/10.1016/0021-9797(86)90420-0.
Acosta, E. J., Harwell, J. H., and Sabatini, D. A. 2004. Self-Assembly in Linker-Modified Microemulsions. J. Colloid Interfac. Sci. 274 (2): 652–664.
https://doi.org/10.1016/j.jcis.2004.03.037.
Adkins, S., Arachchilage, G. P., Solairaj, S. et al. 2012. Development of Thermally and Chemically Stable Large-Hydrophobe Alkoxy Carboxylate Sur-
factants. Presented at the SPE Improved Oil Recovery Symposium, Tulsa, 14–18 April. SPE-154256-MS. https://doi.org/10.2118/154256-MS.
Adkins, S., Liyanage, P. J., Arachchilage, G. W. P. P. et al. 2010. A New Process for Manufacturing and Stabilizing High-Performance EOR Surfactants
at Low Cost for High-Temperature, High-Salinity Oil Reservoirs. Presented at the SPE Improved Oil Recovery Symposium, Tulsa, 24–28 April.
SPE-129923-MS. https://doi.org/10.2118/129923-MS.
Aoudia, M., Wade, W. H., and Weerasooriya, V. 1995. Optimal Microemulsions Formulated With Propoxylated Guerbet Alcohol and Propoxylated Tri-
decyl Alcohol Sodium Sulfates J. Disper. Sci. Technol. 16 (2): 115–135. https://doi.org/10.1080/01932699508943664.
Arf, T. G., LaBelle, G., Klaus, E. E. et al. 1987. EOR With Penn State Surfactants. SPE J. 2 (2): 166–176. SPE-12308-PA. https://doi.org/10.2118/12308-PA.
Barnes, J. R., Dirkzwager, H., Smit, J. R. et al. 2010. Application of Internal Olefin Sulfonates and Other Surfactants to EOR. Part I: Structure–
Performance Relationships for Selection at Different Reservoir Conditions. Presented at the SPE Improved Oil Recovery Symposium, Tulsa, 24–28
April . SPE-129766-MS. https://doi.org/10.2118/129766-MS.
Barnes, J. R., Groen, K., On, A. et al. 2012. Controlled Hydrophobe Branching to Match Surfactant to Crude Oil Composition for Chemical EOR. Pre-
sented at the SPE Improved Oil Recovery Symposium, Tulsa, 14–18 April. SPE-154084-MS. https://doi.org/10.2118/154084-MS.
Chang, L. Y., Lansakara-P., D. S. P., Jang, S. H. et al. 2016. Co-Solvent Partitioning and Retention. Presented at the SPE Improved Oil Recovery Confer-
ence in Tulsa, 11–13 April. SPE-179676-MS. https://doi.org/10.2118/179676-MS.
Flaaten, A. K., Nguyen, Q. P., Pope, G. A. et al. 2009. A Systematic Laboratory Approach to Low-Cost, High-Performance Chemical Flooding. SPE Res
Eval & Eng 12 (5): 713–723. SPE-113469-PA. https://doi.org/10.2118/113469-PA.
Fortenberry, R., Kim, D. H., Nizamidin, N. et al. 2015. Use of Cosolvents to Improve Alkaline/Polymer Flooding. SPE J. 20 (2): 255–266. SPE-166478-
PA. https://doi.org/10.2118/166478-PA.
Frank, C., Frielinghaus, H., and Allgaier, J. 2007. Nonionic Surfactants With Linear and Branched Hydrocarbon Tails: Compositional Analysis, Phase
Behavior, and Film Properties in Bicontinuous Microemulsions. Langmuir 23 (12): 6526–6535. https://doi.org/10.1021/la0637115.
Gale, W. W., Puerto, M. C., Ashcraft, T. L. et al. 1978. Propoxylated Ethoxylated Surfactants and Method of Recovering Oil Therewith. US Patent No.
US4293428.
Gao, B. and Sharma, M. M. 2013. A New Family of Anionic Surfactants for Enhanced-Oil-Recovery Applications. SPE J. 18 (5): 829–840. SPE-
159700-PA. https://doi.org/10.2118/159700-PA.
Ge, J. and Wang, Y. 2015. Surfactant Enhanced Oil Recovery in a High Temperature and High Salinity Carbonate Reservoir. J. Surfactants Deterg.
18 (6): 1043–1050. https://doi.org/10.1007/s11743-015-1735-1.
Graciaa, A., Lachaise, A., Cucuphat, C. et al. 1993a. Improving Solubilization in Microemulsions With Additives. 1. The Lipophilic Linker Role. Lang-
muir 9 (3): 669–672. https://doi.org/10.1021/la00027a010.
Graciaa, A., Lachaise, A., Cucuphat, C. et al. 1993b. Improving Solubilization in Microemulsions With Additives. 2. Long Chain Alcohols as Lipophilic
Linkers. Langmuir 9 (12): 3371–3374. https://doi.org/10.1021/la00036a008.
Huh, C. 1979. Interfacial Tensions and Solubilizing Ability of Microemulsion Phase That Coexists With Oil and Brine. J. Colloid Interfac. Sci. 71 (2):
408–426. https://doi.org/10.1016/0021-9797(79)90249-2.
Iglauer, S., Wu, Y., Shuler, P. et al. 2009. Alkyl Polyglycoside Surfactant-Alcohol Cosolvent Formulations for Improved Oil Recovery. Colloid. Surface.
A 339 (1–3): 48–59. https://doi.org/10.1016/j.colsurfa.2009.01.015.
Jang, S. H., Liyanage, P. J., Tagavifar, M. et al. 2016. A Systematic Method for Reducing Surfactants to Extremely Low Levels. Presented at SPE
Improved Oil Recovery Symposium, Tulsa, 11–13 April. SPE-179685-MS. https://doi.org/10.2118/179685-MS.
Jürgenson, G. A., Bittner, C., Kurkal-Siebert, V. et al. 2015. Alkyl Ether Carboxylate Surfactants for Chemically Enhanced Oil Recovery in Harsh Field
Conditions. Presented at the SPE Enhanced Oil Recovery Symposium, Kuala Lumpur, 11–13 August. SPE-174589-MS. https://doi.org/10.2118/
174589-MS.
Kamal, M. S. 2016. A Review of Gemini Surfactants: Potential Application in Enhanced Oil Recovery. J. Surfactants Deterg. 19 (2): 223–236. https://
doi.org/10.1007/s11743-015-1776-5.
Levitt, D. A., Jackson, A. C., Heinson, C. et al. 2009. Identification and Evaluation of High-Performance EOR Surfactants. SPE Res Eval & Eng 12 (2):
243–253. SPE-100089-PA. https://doi.org/10.2118/100089-PA.
Liyanage, P. J., Lu, J., Arachchilage G. W. P. et al. 2015a. A Novel Class of Large-Hydrophobe Tristyrylphenol (TSP) Alkoxy Sulfate Surfactants for
Chemical Enhanced Oil Recovery. J. Pet. Sci. Eng. 128 (April): 73–85. https://doi.org/10.1016/j.petrol.2015.02.023.
Liyanage, P. J., Upamali, K. A. N., Weerasooriya, U. et al. 2015b. Latest Developments on Enhanced Oil Recovery Related to Oleochemicals. Oral pre-
sentation given at the Malaysian Palm Oil Board International Palm Oil Congress and Exhibition, Kuala Lumpur, 6–8 October.
Lu, J., Britton, C., Solairaj, S. et al. 2014. Novel Large Hydrophobe Carboxylate Surfactant for Enhanced Oil Recovery. SPE J. 19 (6). 1024–1034. SPE-
154261-PA. https://doi.org/10.2118/154261-PA.
Menger, F. M. and Littau, C. A. 1993. Gemini Surfactants: A New Class of Self-Assembling Molecules. J. Am. Chem. Soc. 115 (2): 10083–10090.
https://doi.org/10.1021/ja00075a025.
Miñana-Perez, M., Graciaa, A., Lachaise, J. et al. 1996. Systems Containing Mixtures of Extended Surfactants and Conventional Nonionics: Phase
Behavior and Solubilization in Microemulsion. Proc., 4th World Surfactant Congress, Barcelona, Spain, 3–7 June, Vol. 2, 226–234.
Pandey, A., Koduru, N., Stanley, M. et al. 2016. Results of ASP Pilot in Mangala Field: A Success Story. Presented at the SPE Improved Oil Recovery
Symposium, Tulsa, 11–13 April. SPE-179700-MS. https://doi.org/10.2118/179700-MS.
Pandey, A., Suresh Kumar, M., Jha, M. K. et al. 2012. Chemical EOR Pilot in Mangala Field: Results of Initial Polymer Flood Phase. Presented at the
SPE Improved Oil Recovery Symposium, Tulsa, 14–18 April. SPE-154159-MS. https://doi.org/10.2118/154159-MS.
Prasad, D., Pandey, A., Suresh Kumar, M. et al. 2014. Pilot to Full-Field Polymer Application in One of the Largest Onshore Field in India. Presented at
the SPE Improved Oil Recovery Symposium, Tulsa, 12–16 April. SPE-169146-MS. https://doi.org/10.2118/169146-MS.
Puerto, M., Hirasaki, G. J., Miller, C. A. et al. 2012. Surfactant Systems for EOR in High Temperature, High Salinity Environments. SPE J. 17 (1):
11–19. SPE-129675-PA. https://doi.org/10.2118/129675-PA.
Rajapaksha, S., Britton, C., McNeil, R. I. et al. 2014. Restoration of Reservoir Cores to Reservoir Condition Before Chemical Flooding Tests. Presented
at the SPE Improved Oil Recovery Symposium, Tulsa, 12–16 April. SPE-169887-MS. https://doi.org/10.2118/169887-MS.
Sahni, V., Dean, R. M., Britton, C. et al. 2010. The Role of Co-Solvents and Co-Surfactants in Making Chemical Floods Robust. Presented at SPE
Improved Oil Recovery Symposium, Tulsa, 24–28 April. SPE-130007-MS. https://doi.org/10.2118/130007-MS.

2216 December 2018 SPE Journal

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025


J179702 DOI: 10.2118/179702-PA Date: 17-November-18 Stage: Page: 2217 Total Pages: 16

Salter, S. J. 1977. The Influence of Type and Amount of Alcohol on Surfactant-Oil-Brine Phase Behavior and Properties. Presented at the SPE Annual
Fall Technical Conference and Exhibition, Denver, 9–12 October. SPE-6843-MS. https://doi.org/10.2118/6843-MS.
Sanz, C. A. and Pope, G. A. 1995. Alcohol-Free Chemical Flooding: From Surfactant Screening to Coreflood Design. Presented at the SPE International
Symposium on Oilfield Chemistry, San Antonio, Texas, 14–17 February. SPE-28956-MS. https://doi.org/10.2118/28956-MS.
Solairaj, S., Britton, C., Kim, D. H. et al. 2012. Measurement and Analysis of Surfactant Retention. Presented at the SPE Improved Oil Recovery Sympo-
sium, Tulsa, 14–18 April. SPE-154247-MS. https://doi.org/10.2118/154247-MS.
Song, B., Hu, X., Shui, X. et al. 2015. A New Type of Renewable Surfactants for Enhanced Oil Recovery: Dialkylpolyoxyethylene Ether Methyl Car-
boxyl Betaines. Colloid. Surface. A 489 (20 January): 433–440. https://doi.org/10.1016/j.colsurfa.2015.11.018.
Tagavifar, M., Herath, S., Weerasooriya, U. P. et al. 2017. Measurement of Microemulsion Viscosity and Its Implications for Chemical Enhanced Oil
Recovery. SPE J. 23 (1): 66–83. SPE-179672-PA. https://doi.org/10.2118/179672-PA.
Talley, L. D. 1988. Hydrolytic Stability of Alkylethoxy Sulfates. SPE Res Eng 3 (1): 235–242. SPE-14912-PA. https://doi.org/10.2118/14912-PA.
Walker, D., Britton, C., Kim, D. H. et al. 2012. The Impact of Microemulsion Viscosity on Oil Recovery. Presented at the SPE Improved Oil Recovery
Symposium, Tulsa, 14–18 April. SPE-154275-MS. https://doi.org/10.2118/154275-MS.
Yang, H. T., Britton, C., Liyanage, P. J. et al. 2010. Low-Cost High-Performance Chemicals for Enhanced Oil Recovery. Presented at the SPE Improved
Oil Recovery Symposium, Tulsa, 24–28 April. SPE-129978-MS. https://doi.org/10.2118/129978-MS.

Karasinghe A. Nadeeka Upamali is a research associate at the Center for Petroleum and Geosystems Engineering at the Univer-
sity of Texas at Austin. Her expertise includes designing and developing new surfactant formulations for ASP, SP, and ACP in
CEOR with various types of crude oils and different conditions. Upamali also specializes in chemical synthesis, characterization,
and improvement of surfactants and methods development for analysis of EOR chemicals and fluids. She was also a part of the
development of surfactant formulations and methods for several field trials. Upamali holds a PhD degree in photochemistry, with
7 years of experience working with EOR chemicals and methods. She is a member of SPE.
Jith Liyanage is a research associate at the Center for Petroleum and Geosystems Engineering at the University of Texas at
Austin. He has worked on multiple CEOR projects from laboratory inception to pilot and field trials, and his focus has been devel-
oping surfactant and other chemistries for EOR projects. Liyanage is also interested in chemical analysis and designing experi-
ments and laboratory setups for EOR projects, including high-temperature/high-pressure corefloods. He holds a bachelor’s
degree in chemistry from Wichita State University. Liyanage is a member of SPE.
Sung Hyun Jang is a senior research scientist at Kemira. Before joining Kemira in January 2018, he worked at the Center for Petro-
leum and Geosystems Engineering at the University of Texas at Austin for 5 years as a post-doctoral-degree fellow and research
associate. Jang’s research interests are centered on CEOR technologies using surfactants and polymers. He holds bachelor’s
and PhD degrees in chemical engineering, both from Seoul National University, South Korea. Jang is a member of SPE.
Erin Shook is a student at the University of Texas at Austin. She is currently a candidate for a master’s degree in information stud-
ies. Previously, Shook was a technical staff member in the Petroleum and Geosystems Engineering Department at the University
of Texas. She holds a bachelor’s degree in anthropology from the University of Texas at Austin.
Upali Peter Weerasooriya is a retired faculty member of, and a current senior technical consultant to, the Center for Petroleum
and Geosystems Engineering at the University of Texas at Austin. In 2008, he retired from Harcross Chemicals as vice president
and chief technology officer and joined the CEOR program at the University of Texas at Austin. Weerisooriya is currently on the
Board of Directors of Harcros Global Manufacturing, Venus Etoxyethers (India), Esteem Industries (India), and UltimTE Enhanced
Oil Recovery Services. His expertise is in the design and application of surfactants and other chemicals in the EOR and detergent
industries. Weerasooriya holds or has authored applications for more than 50 US patents and has authored or coauthored more
than 50 scientific publications. He holds a PhD degree in chemistry from the University of Texas at Austin. In 1999, Weerasooriya
won the Best Surfactant Scientist Award (Rosen Award) from the American Oil Chemists Society. He is a member of SPE.
Gary A. Pope is a professor in the Hildebrand Department of Petroleum and Geosystems Engineering at the University of Texas at
Austin, where he holds the Texaco Centennial Chair in Petroleum Engineering. His teaching and research are in the areas of
EOR, reservoir engineering, natural-gas engineering, and reservoir simulation. Pope holds a bachelor’s degree from Oklahoma
State University and a PhD degree from Rice University, both in chemical engineering. He was elected to the National Academy
of Engineering in 1999. Pope is a Distinguished Member of SPE and holds SPE Honorary Member status and he has received the
AIME Environmental Conservation Award and AIME Distinguished Service Award, the Hocott Distinguished Centennial Engineer-
ing Research Award, the SPE Improved Oil Recovery Pioneer Award, the SPE/AIME Anthony F. Lucas Gold Medal, the SPE John
Franklin Carll Award, the SPE Distinguished Achievement Award, and the SPE Reservoir Engineering Award.

December 2018 SPE Journal 2217

ID: jaganm Time: 13:54 I Path: S:/J###/Vol00000/180025/Comp/APPFile/SA-J###180025

Вам также может понравиться