Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

SPE 115386

Recent Advances in Surfactant EOR


George J. Hirasaki, SPE, Clarence A. Miller, SPE, and Maura Puerto, Rice University

Copyright 2008, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Denver, Colorado, USA, 21-24 September 2008.

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
Recent advances in surfactant EOR are reviewed. The addition of alkali to surfactant
flooding in the 1980s reduced the amount of surfactant required and the process became
known as alkaline surfactant polymer flooding (ASP). It was recently found that the adsorption
of anionic surfactants on calcite and dolomite can also be significantly reduced with sodium
carbonate as the alkali, thus making the process applicable for carbonate formations. The
same chemicals are also capable of altering the wettability of carbonate formations from
strongly oil-wet to preferentially water-wet. This wettability alteration in combination with ultra-
low interfacial tension (IFT) makes it possible to displace oil from preferentially oil-wet
carbonate matrix to fractures by oil-water gravity drainage.
The alkaline surfactant process co-injects alkali and synthetic surfactant. The alkali
generates soap in situ by reaction between the alkali and naphthenic acids in the crude oil. It
was recently recognized that the local ratio of soap/surfactant determines the local optimal
salinity for minimum IFT. Recognition of this dependence makes it possible to design a
strategy to maximize oil recovery with the least amount of surfactant and inject polymer with
the surfactant without phase separation. An additional benefit of the presence of the soap
component is that it generates an oil-rich colloidal dispersion which produces ultra-low IFT
over a much wider range of salinity than in its absence.
It was once thought that a co-solvent such as alcohol was necessary to make a
microemulsion without gel-like phases or a polymer-rich phase separating from the surfactant
solution. An example of an alternative to the use of alcohol is to blend two dissimilar
surfactants, a branched alkoxylated sulfate and a double-tailed, internal olefin sulfonate. The
single phase region with NaCl or CaCl2 is greater for the blend than for either surfactant alone.
It is also possible to incorporate polymer into such aqueous surfactant solutions without phase
separation under some conditions. The injected surfactant solution has under-optimum phase
behavior with the crude oil. It becomes optimum only as it mixes with the in situ generated
soap, which is generally more hydrophobic than the injected surfactant. However, some crude
oils do not have a sufficiently high acid number for this approach to work.
Foam can be used for mobility control by alternating slugs of gas with slugs of
surfactant solution. Besides effective oil displacement in a homogeneous sandpack, it
demonstrated greatly improved sweep in a layered sandpack.
2 [SPE 115386]

Introduction
It is generally considered that only about one third of the petroleum present in known
reservoirs is economically recoverable with established technology, i.e., primary recovery
methods utilizing gas pressure and other natural forces in the reservoir and secondary
recovery by waterflooding. It has long been an objective of the industry to develop improved
processes to increase overall recovery. However, the low oil prices which prevailed from the
mid-1980’s until recently provided little incentive for research on enhanced oil recovery (EOR),
especially surfactant processes with substantial initial cost for chemicals. With the current
higher prices and accompanying revival of interest it seems appropriate to review
understanding of and prospects for surfactant EOR.
Adding surfactant to injected water to reduce oil-water interfacial tension (IFT) and/or
alter wettability and thereby increase recovery is not a new idea (see, for instance, Uren and
Fahmy, 1927). Indeed, a few of early field trials where small amounts of surfactant were
injected did produce small increases in oil recovery. The increases were probably caused
mainly by wettability changes although the data were inconclusive for assessing mechanisms.
The results were not sufficiently promising to stimulate use of surfactants on a larger scale. A
related long-held concept for improving recovery is to generate surfactant in situ by injecting an
alkaline solution (Atkinson, 1927), which is less expensive than synthetic surfactants and
converts naphthenic acids in the crude oil to soaps. Early results were not encouraging, and
the relative importance of likely process mechanisms was not understood (Johnson, 1976).
Other references to early work on surfactants are given by Hill et al. (1973).
Two different approaches stimulated significant advances in surfactant EOR processes
in the 1960’s. The surfactants were made either by direct sulfonation of aromatic groups in
refinery streams or crude oils or by organic synthesis of alkyl/aryl sulfonates, which allowed for
the surfactant to be tailored to the reservoir of interest. The advantages of these surfactants
are their low cost, wide range of properties, and availability of raw materials in somewhat large
quantities.
Miscible flooding was an active area of research, but the solvents being considered
such as enriched gas and LPG exhibited poor reservoir sweep because the adverse mobility
ratio promoted viscous fingering and the low solvent density led to gravity override. Seeking a
solvent miscible with oil but having a higher viscosity and density, Gogarty and coworkers at
Marathon proposed using a slug of an oil-continuous microemulsion made of hydrocarbon, a
petroleum sulfonate surfactant, an alcohol, and water or brine (see review by Gogarty 1977).
Holm and coworkers at Union Oil advocated a similar process using a “soluble oil,” which was
also an oil-continuous microemulsion made mainly of crude oil, some mineral oil, petroleum
sulfonate, a cosolvent such as ethylene glycol monobutyl ether and water, as summarized by
Holm (1977). Slugs of these materials displaced oil miscibly and with better sweep than
previous solvents. However, it was not initially recognized that process success also
depended on maintaining ultralow IFT at the rear of the slug, where it was displaced by an
aqueous polymer solution and became a Winsor I microemulsion (Hirasaki, 1981).
The other approach involved injection of a surfactant formulation made of a petroleum
sulfonate and alcohol in an aqueous electrolyte solution. Key to success of this approach were
systematic studies of oil displacement leading to recognition that a dimensionless capillary
number ( μ v / σ ) controlled the amount of residual oil remaining after flooding an oil-containing
core at interstitial velocity, v with an aqueous solution having a viscosity, µ and IFT, σ with the
oil (Taber, 1969; Stegemeier, 1977; Melrose and Brandner, 1974; Foster, 1973). This work
revealed that at typical reservoir velocities, IFT had to be reduced from crude oil-brine values
[SPE 115386] 3

of 20-30 mN/m to values in the range of 0.001 to 0.01 mN/m to achieve low values of residual
oil saturation (<0.05).
Several research groups found that ultralow interfacial tensions in the required range
could be achieved using petroleum sulfonate/alcohol mixtures (Hill et al., 1973; Foster, 1973;
Cayias et al., 1977). They also found systematic variations of IFT when changing such
variables as salinity, oil composition, and temperature. An important contribution was the work
of Healy, et al (1976; see also Reed and Healy, 1977), who demonstrated a relationship
between IFT and microemulsion phase behavior and developed correlations between
microemulsion IFT’s with oil and brine and the volumes of oil and water solubilized per unit
volume of surfactant in the microemulsion. In particular they showed for anionic surfactants
that an optimal salinity exists where microemulsions form which solubilize equal volumes of oil
and water and which exhibit nearly equal IFT’s with oil and brine. Core tests using continuous
surfactant injection at the optimal salinity also yielded the highest recovery of waterflood
residual oil. Their studies used mixtures of an alcohol co-solvent with synthetic alkyl/aryl
sulfonates, in particular C9, C12, and C15 orthoxylene sulfonates, which can be made from
oligomers of propylene with more reproducible compositions than petroleum sulfonates.

Surfactant Requirements and Structures


In a successful displacement process the injected surfactant slug must first achieve
ultralow IFT to mobilize residual oil and create an oil bank where both oil and water flow as
continuous phases (Bourrel and Schechter, 1988). Second, it must maintain ultralow IFT at
the moving displacement front to prevent mobilized oil from being trapped by capillary forces.
Because of the way they are prepared, commercial surfactants are invariably mixtures of
multiple species, which raises questions as to whether chromatographic separation, i.e.,
preferential adsorption on pore surfaces or preferential partitioning into the oil phase of some
species, can cause IFT variations with possible adverse effects on oil recovery. When alcohol
is used in the formulation, it partitions among the bulk oil and brine phases and the surfactant
films in a manner different from the surfactant. The alcohol must then be careful selected and
tested to ensure there is no deleterious effect of chromatographic separation (Dwarakanath et
al., 2008). In the surfactant films, alcohol serves as a cosolvent, making them less rigid and
thereby increasing the rate of equilibration and preventing formation of undesirable viscous
phases and emulsions instead of the desired low-viscosity microemulsions. Alcohol can also
serve as a cosurfactant, altering, for instance, the optimal salinity required to achieve ultralow
IFT. Alcohols with short chains such as propanol increase optimal salinity for sulfonate
surfactants, while longer-chain alcohols such as pentanol and hexanol decrease optimal
salinity. For petroleum sulfonates and synthetic alkyl/aryl sulfonates with light crude oils, it has
been found that 2-butanol acts as a cosolvent but has less affect on optimal salinity than other
alcohols.
A disadvantage of using alcohol is that it decreases solubilization of oil and water in
microemulsions and hence increases the minimum value of IFT achievable with a given
surfactant (Salter, 1977). Also, it destabilizes foam that may be desired for mobility control
with the slug and in the drive. For temperatures below approximately 60°C, the need for
alcohol’s cosolvent effect can be reduced or eliminated by some combination of the following
strategies: (1) using surfactants with branched hydrocarbon chains, (2) adding ethylene oxide
(EO) and/or less hydrophilic propylene oxide (PO) groups to the surfactant, and (3) using
mixtures of surfactants with different hydrocarbon chain lengths or structures. Such measures
counter the tendency of long, straight hydrocarbon chains of nearly equal length to form
4 [SPE 115386]

condensed surfactant films and the lamellar liquid crystalline phase. At high temperatures,
increased thermal motion promotes more flexible surfactant films and disruption of ordered
structures, but does not always eliminate viscous phases and emulsions. Further studies are
needed to investigate usefulness of the above or other strategies for high temperatures.
The use of branched hydrocarbon chains to minimize or eliminate alcohol requirements
was discussed by Wade, Schechter and coworkers (Abe et al., 1986). An isotridecyl
hydrophobe was used in Exxon’s pilot test in the Loudon field. The hydrophilic part of the
surfactant was a chain consisting of short PO and EO segments and a sulfate group (Maerker
and Gale 1992). With this combination of the first two strategies, no alcohol was required.
The Neodol 67 hydrophobe developed and manufactured by Shell Chemical has an average of
1.5 methyl groups added randomly along a straight C16-C17 chain that provides another type of
branching. A propoxylated sulfate with this hydrophobe has been blended with an internal
olefin sulfonate, which is also branched, and used to displace West Texas crude oils in low-
salinity, ambient-temperature laboratory tests, both with (Levitt et al., 2006) and without (Liu et
al., 2008) alcohol. In this case all three strategies were combined. Internal olefin sulfonates
have twin tailed structure that has been shown to also be beneficial for the same reasons as
branching (Zhao et al. 2008).
Long-term surfactant stability at reservoir conditions is another surfactant requirement.
Provided that pH is maintained slightly alkaline amd calcium concentration is not too high,
hydrolysis of sulfate surfactants is limited for temperatures up to 50-60°C (Talley, 1988).
Surfactants with other head groups, most likely sulfonates, will be needed for reservoirs at
higher temperatures. Because the sulfonate group is added at different points along the
hydrocarbon chain during synthesis, internal olefin sulfonate (IOS) surfactants consist of
species with different branching. Results of laboratory studies of IOS phase behavior at high
temperatures have recently been presented (Barnes et al., 2008; Zhao et al., 2008).
Hydrocarbon chain lengths ranged from C15-C18 to C24-C28. However, the effect of dissolved
calcium and magnesium ions, which most likely cause surfactant precipitation, was not
investigated in these studies (see Fig. 2 of Liu, et al., 2008a). Moreover, the current availability
of internal olefins of long chain length, as would be required for low or moderate reservoir
salinities, is limited. Alpha olefin sulfonate (AOS) surfactants have the same potential
disadvantages and are not as highly branched as IOS surfactants.
Sulfonate groups cannot be added directly to alcohols including those with EO and/or
PO chains. One approach is to prepare sulfonates via glycidyl chloride or epichlorohydrin,
where a three-carbon chain is added between the EO or PO chain and the sulfonate group.
Wellington and Richardson (1997) described some results with such surfactants. Further
information on their synthesis and initial phase behavior results for propoxy glycidyl sulfonates
with the Neodol 67 hydrophobe are given by Barnes et al. (2008).
Until recent years, nearly all work was directed toward EOR in sandstone reservoirs
owing to concerns that in the high-divalent-ion environment of carbonate reservoirs, petroleum
or synthetic sulfonates would adsorb excessevly and/or form calcium and magnesium salts
that would either precipitate or partition into the oil phase. The exception was the work of
Adams and Schievelbein (1984), who conducted laboratory experiments and two field tests
showing that oil could be displaced in a carbonate reservoir using a mixture of petroleum
sulfonates and ethoxylated sulfate surfactants. The ethoxy groups add tolerance to divalent
ions. Recent work with carbonate reservoirs used ethoxylated or propoxylated sulfates, as
discussed below in the section on wettability.
[SPE 115386] 5

Alcohol-free Surfactant Slugs for Injection


The surfactant slug to be injected should be a single-phase micellar solution. Especially
when polymer is added to increase slug viscosity, it is essential to prevent separation into
polymer-rich and surfactant-rich phases, which yields highly viscous phases unsuitable for
either injection or propagation through the formation (Trushenski, 1977). At low temperatures,
oil-free mixtures of petroleum sulfonate/alcohol or synthetic sulfonate/alcohol mixtures with
brine are often translucent micellar solutions at salinities well below optimal but contain
lamellar liquid crystal and exhibit birefringence near optimal salinity where ultralow IFT is found
on mixing with crude oil (Miller et al., 1986). In the absence of polymer, the lamellar phase is
often dispersed in brine as particles having maximum dimensions of at least several
micrometers. When polymer is added to such a turbid dispersion of the lamellar phase,
separation into a polymer-rich aqueous solution and a more concentrated dispersion occurs
(Qutubuddin, 1985). This undesirable behavior can sometimes be avoided by adding sufficient
alcohol. However, use of alcohol has disadvantages, as indicated above.
The lamellar phase was observed in surfactant/brine mixtures in the absence of oil even
for Exxon’s Loudon formulation mentioned above (Ghosh, 1985), where, as indicated
previously, branching and addition of EO/PO groups allowed low-viscosity microemulsions to
be formed and ultralow IFT to be achieved with the crude oil without the need to add alcohol.
Exxon avoided phase separation when polymer was added to the injected slug by including a
paraffinic white oil of high molecular weight in the formulation. That is, they injected a white oil-
in-water microemulsion (Winsor I), which became a bicontinuous microemulsion when mixed
with substantial volumes of crude oil in the reservoir (Maerker and Gale, 1992).
As temperature increases, the lamellar liquid-crystal phase may melt. However, the
surfactant/brine mixture is still unsuitable for injection if separation into two or more liquid
phases occurs near optimal salinity (Benton and Miller, 1983). Even if bulk phase separation
does not occur, turbid solutions are sometimes observed. These solutions usually have large,
anisotropic micelles and separate into surfactant-rich and polymer-rich phases on addition of
polymer. These phases can separate and/or plug the porous media into which it is injected.
Adding alcohol can reduce micelle size and prevent phase separation in some cases. As
indicated previously, addition of an oil which yields an oil-in-water microemulsion with nearly
spherical drops, is another approach for preventing phase separation when polymer is present.
Within limits, the higher the molecular weight of the oil added to produce an oil-in-water
microemulsion, the less oil is needed to formulate single phases with polymer for mobility
control.
Another approach to formulate single phase injection compositions would be to find
surfactants or surfactant blends that neither exhibit phase separation nor form turbid solutions
or liquid crystalline dispersions at conditions of interest. Blends of the branched surfactant
Neodol 67 propoxylated sulfate (N67-7POS) having an average of 7 PO groups with the twin-
tailed surfactant IOS 15/18, an internal olefin sulfonate made from a feedstock containing
mainly C15–C18 chains, are interesting in this respect. Fig. 1 shows phase behavior at ambient
temperature of 3 wt% aqueous solutions of such surfactant blends containing 1 wt% Na2CO3
and varying NaCl concentration but no alcohol or polymer. IOS 15/18 alone precipitates above
4 wt% NaCl in such solutions. In contrast, solutions of the propoxylated sulfate alone do not
precipitate but instead become cloudy above the same salinity as droplets of a second liquid
phase form and scatter light. Addition of IOS 15/18 to the propoxylated sulfate makes the
mixture more hydrophilic, thereby raising the salinity at which phase separation occurs to a
value higher than for either surfactant alone. For instance, the 4:1 blend (hereafter NI blend),
i.e., 80% N67-7POS and 20% IOS 15/18, exhibits phase separation at approximately 6 wt%
6 [SPE 115386]

NaCl although slight cloudiness occurs above approximately 3.5 wt% NaCl. Addition of 0.5
wt% of partially hydrolyzed polyacrylamide to a 0.5 wt% solution of this blend in a solution
containing 4 wt% NaCl and 1 wt% Na2CO3 produces phase separation although similar
addition of polymer to a solution containing only 2 wt% NaCl does not (Liu et al., 2008a).
Phase behavior studies show that the optimal salinity of this blend with a West Texas crude oil
is approximately 5 wt% NaCl (with 1 wt% Na2CO3) when the amount of surfactant present is
much greater than soap formed from the naphthenic acids in the crude oil. However, in
alkaline surfactant processes, it is best to inject at lower salinities, as discussed further below,
because the surfactant encounters conditions during the process with greater ratios of soap-to-
surfactant and correspondingly lower optimal salinities. Indeed, excellent recovery of the West
Texas crude oil was observed in sand pack experiments when a single-phase mixture of the
4:1 blend and polymer was injected at 2 wt% NaCl with 1 wt% Na2CO3 (Liu et al., 2008a). At
2% NaCl, the surfactant micelles are not highly anisotropic, and polymer and surfactant can
coexist in the same phase.
A similar approach was used by Falls et al. (1994) in an alkaline surfactant field test.
They added a small amount of the nonionic surfactant Neodol 25-12 to the main injected
surfactant, a blend of internal olefin sulfonates, to make the formulation sufficiently hydrophilic
to form a single micellar solution during storage at ambient temperature. Because this
surfactant becomes less hydrophilic at higher temperatures, it did not adversely affect process
performance at the reservoir temperature of approximately 57°C.
Thus, two approaches have been discussed to achieve alcohol-free single-phase slugs
for injection even when polymer is added to provide mobility control. One is that of the Exxon
Loudon project where a small amount of paraffinic oil having a higher optimal salinity than the
crude oil is added. Even at optimal salinity for the surfactant formulation with reservoir crude,
an oil-in-water microemulsion having nearly spherical drops is formed with the added oil. Its
high molar volume reduces its solubilization and hence microemulsion drop size, thereby
preventing phase separation. The second approach is to use an alkaline surfactant process,
where high oil recovery can be achieved if injection salinity is below the optimal salinity of the
injected synthetic surfactant with the crude oil.

Alkaline Surfactant Processes: Role of Alkali


Nelson et al. (1984) proposed injection of a solution containing both surfactant and
alkali for EOR. Such processes have attracted and continue to attract considerable interest.
They have been labelled by different names but will be collectively described here as alkaline
surfactant processes (Nelson, et al., 1984; Peru and Lorenz, 1990; Surkalo, 1990; Baviere, et
al., 1995).
The primary role of the alkali in an alkaline surfactant process is to reduce adsorption of
the surfactant during displacement through the formation and sequester divalent ions. An
additional benefit of alkali is that the soap is formed in situ from the naphthenic acid in the
crude oil (Johnson, 1976). As indicated previously, the presence of soap allows the surfactant
to be injected at lower salinities than if used alone, which further reduces adsorption and
facilitates incorporation of polymer in the surfactant slug. Also, alkali can alter formation
wettability to either more water-wet or more oil-wet state. In fractured oil-wet reservoirs, the
combined effect of alkali and surfactant in making the matrix prefentially water-wet is essential
for an effective process. These benefits of alkali will occur only where alkali is present. Thus it
is important to determine “alkali consumption”, which controls the rate of propagation of alkali
through the formation.
[SPE 115386] 7

Reduced surfactant adsorption. The discussion here will be limited to anionic


surfactants (Wesson and Harwell, 2000). The primary mechanism for the adsorption of anionic
surfactants on sandstone and carbonate formation material is the ionic attraction between
positively-charged mineral sites and the negative surfactant anion (Zhang and Somasundaran,
2006). Thus the role of the alkali is to be a “potential-determining-ion” to reverse the charge on
positively-charged mineral sites. The potential determining ions for oxide minerals are the
hydronium and hydroxide ions. The pH at which the charge reverses is the “isoelectric point” if
measured by electrophoresis (zeta potential) and the “point-of-zero-charge” if determined by
titration. The values are tabulated for most common minerals (Lyklema, 1995). Silica is
negatively charged at reservoir conditions and exhibits negligible adsorption of anionic
surfactants. Clays (at neutral pH) have negative charge at the faces and positive charge at the
edges. The clay edges are alumina-like and thus are expected to reverse their charge at a pH
of about 9. Carbonate formations and sandstone cementing material can be calcite or
dolomite. These latter minerals also have an isoelectric point of about pH 9 but carbonate ion
as well as the calcium and magnesium ions are more significant potential determining ions.
The zeta potential of calcite is negative even at neutral pH in the presence of 0.1 N
carbonate/bicarbonate ions (Hirasaki and Zhang, 2004). If a formation contains iron minerals,
the oxidation/reduction conditions influence whether the surface iron sites are Fe3+ or Fe2+.
Adsorption of anionic surfactant for one sandstone was found to be lower by more than a
factor of two for reducing rather than for oxidizing conditions (Wang, 1993). Surfactant
adsorption is only one component of surfactant retention. Phase trapping of surfactant can be
more significant and will be discussed later.
Alkaline preflush had been advocated to both sequester divalent ions and reduce
sulfonate adsorption (Holm and Robertson, 1981). In subsequent work, alkali has been
injected with the surfactant. Adsorption of anionic surfactants on Berea sandstone was
reduced several fold with addition of sodium carbonate for petroleum sulfonate (Bae and
Petrick, 1977) or with addition of sodium silicate or hydroxide for alcohol ethoxysulfate (Nelson,
et al., 1984). The reduction of adsorption on Berea sandstone with sodium bicarbonate was
68% in a dynamic experiment (Peru and Lorenz, 1990). Static and dynamic adsorption of
anionic surfactants on calcite and dolomite was decreased by an order of magnitude with
addition of sodium carbonate but insignificantly with sodium hydroxide, see Figs. 2, 3, and 4
(Hirasaki and Zhang, 2004; Seethepalli, et al., 2004; Zhang, et al., 2006; Liu, et al., 2008a;).
Divalent ion sequestration. The phase behavior of anionic surfactant systems is
much more sensitive to a change in divalent ions (e.g., Ca2+ and Mg2+) compared to
monovalent ions (e.g., Na+), especially at low surfactant concentrations (Nelson, 1981). This is
problematic in sandstones because of ion exchange between the clay, brine and surfactant
micelles (Hill, et al., 1977; Hirasaki, 1982). This exchange can result in the phase behavior
becoming over-optimum with resulting large surfactant retention (Glover, et al., 1979, Gupta,
1981). Alkali anions (e.g., carbonate, silicate, and phosphate) that have low solubility product
with divalent cations will sequester divalent cations to low concentrations (Holm and
Robertson, 1981). Hydroxide is not as effective for sequestration of calcium because the
solubility product of calcium hydroxide is not very low. Sodium metaborate has recently been
introduced as an alkali that may sequester divalent ions by complex formation rather than
precipitation (Flaaten, et al., 2008; ; Zhang et al. 2008).
Generation of soap. The original concept of alkali flooding was the reduction of oil-
water IFT by in situ generation of soap, which is an anionic surfactant, sodium naphthenate
(Jennings, 1975). Ultra-low IFT usually required injection of relatively fresh water with a low
concentration of alkali because optimal salinity (total electrolyte concentration) of the in situ
8 [SPE 115386]

generated soap is usually low (e.g., <1% electrolyte). If the alkali concentration is too low,
alkali consumption may result in a large retardation of the alkali displacement front. The
concept of alkaline surfactant flooding is to inject a surfactant with the alkaline solution such
that mixture of the in situ generated soap and injected surfactant has an optimal salinity that is
tailored to the reservoir fluids (Nelson, et al., 1984; Surkalo, 1990).
The common method used to determine the amount of naphthenic acid in crude oil is
the total acid number (TAN) determined by non-aqueous titration with a base (Fan and
Buckley, 2007). If sodium naphthenate is to act as a surfactant, it should partition into the
aqueous phase at low electrolyte concentrations and be measurable by hyamine titration for
anionic surfactants. It was found (see later) that the sodium naphthenate determined by
extraction into the aqueous phase and measured by hyamine titration is about one-half of the
TAN value (Liu, et al., 2008b). It is hypothesized that the TAN includes components that are
too lipophilic to be extracted to the aqueous phase and/or too hydrophilic to be detected by
hyamine titration.
Alkali consumption. The ASP process should be designed such that displacement
fronts of the alkali, surfactant, and polymer travel together. The mechanisms responsible for
the retardation of the alkali front include silica dissolution, clay dissolution with zeolite
precipitation, anhydrite or gypsum dissolution with calcite (or calcium hydroxide or silicate)
precipitation, dolomite dissolution with calcium and magnesium silicate precipitation, hydrogen
ion exchange, divalent ion exchange with precipitation, and mixing with divalent ions in
formation water with precipitation (Ehrlich and Wygal, 1977; Holm and Robertson, 1981;
Southwick, 1985; Cheng, 1986; Novosad and Novosad, 1982; Jensen and Radke, 1988;
Mohammadi, et al., 2008). Naphthenic acids in crude oil also react with alkali and thus
contribute to consumption but the amount is usually small compared to the mentioned
inorganic mineral reactions. Silica dissolution can be controlled by using a buffered system
such as sodium carbonate or silicate rather than hydroxide (Southwick, 1985). Clay
dissolution is strongly dependent on the pH, type of clay and is kinetically limited (Sydansk,
1981; Zhao, et al., 2004). Thus, acidic clay such as kaolinite and high temperature will
increase the importance of this mechanism.
A limitation of the application of sodium carbonate in carbonate formations is that if
anhydrite or gypsum is present, it will dissolve and precipitate as calcite. This is detrimental for
dolomite formations as they may have had their origin with evaporite deposits where gypsum is
usually present. An alternative alkali is sodium metaborate (Zhang, et al., 2008; Flaaten, et al.,
2008), which may complex calcium without precipitation.

Alkaline Surfactant Processes: Wettability Alteration


Wettability is the next most important factor in waterflood recovery after geology
(Morrow, 1990). The recovery efficiency of a flooding process is a function of the displacement
efficiency and sweep efficiency. These efficiencies are a function of the residual oil saturation
(waterflood and chemical flood) and mobility ratio, respectively. The residual oil saturation to
waterflooding is a function of wettability with the lowest value at intermediate wettability
(Jadhunandan and Morrow, 1995). The mobility ratio is a function of the ratio of water and oil
relative permeabilities at their respective endpoints or at a specific saturation. The mobility
ratio or relative permeability ratio becomes progressively larger as the wettability changes from
water-wet to oil-wet. (Anderson, 1987b). When a formation is strongly oil-wet, it can have both
a high waterflood residual oil saturation and unfavorable mobility ratio. In addition, an oil-wet
formation will have capillary resistance to imbibition of water (Anderson, 1987a). Formation
wettability can be altered by pH (Wagner and Leach, 1959; Ehrlich, et al., 1974; Takamura and
[SPE 115386] 9

Chow, 1985; Buckley, et al., 1989, Dubey and Doe, 1993), surfactants that adsorb on the
minerals (Somasundaran and Zhang, 2006) or remove adsorbed naphthenic acids (Standnes
and Austad, 2000), and acids or bases (Cuiec, 1977). These processes are now incorporated
into chemical flood simulators (Anderson, et al., 2006; Delshad, et al, 2006; Adibhatia and
Mohanty, 2007, 2008).
Sandstone formations. Wettability alteration to more water-wet or more oil-wet
conditions was proposed as one of the mechanisms of caustic flooding (Wagner and Leach,
1959; Ehrlich, et al., 1974; Johnson, 1976). Our current understanding of microemulsion
phase behavior and wettability is that the system wettability is likely to be preferentially water-
wet when the salinity is below the optimal salinity (Winsor I) and is likely to be preferentially oil-
wet when the salinity is above the optimal salinity (Winsor II), even in the absence of alkali.
The optimal salinity for a conventional alkali flooding system is dependent on the in situ
generated sodium naphthenate soap and is usually below about 1% electrolyte strength. As
salinity of reservoir brine typically exceeds this value, an alkali flood often generates over-
optimum and oil-wet conditions. We show later in this review that this behavior can be avoided
by injecting the alkali and surfactant in the Winsor I region. After mixing with the fluids in the
reservoir of interest, it will pass through the Winsor III, low IFT region. Even a high-salinity,
sandstone formation that is initially oil wet may be altered to preferentially water-wet by
injecting alkali with a hydrophilic surfactant in the Winsor I region.
Carbonate formations. Wettability alteration has received more attention recently for
carbonate formations compared to sandstones because carbonate formations are much more
likely to be preferentially oil-wet (Treiber, et al., 1972). Also, carbonate formations are more
likely to be fractured and will depend on spontaneous imbibition or buoyancy for displacement
of oil from the matrix to the fracture.
Wettability alteration tests on plates of calcite, marble, limestone, and dolomite with
different surfactants and sodium carbonate have been used to identify many systems that are
altered to preferentially water-wet with low anionic surfactant concentrations (Hirasaki and
Zhang, 2004; Seethepalli, et al., 2004; Zhang, et al., 2006; Adibhatta and Mohanty, 2008).
Sodium carbonate has an important role since the carbonate ion is a potential determining ion
for calcite and dolomite (Hirasaki and Zhang, 2004).
Spontaneous Imbibition. Spontaneous imbibition is the process by which a wetting
fluid is drawn into a porous medium by capillary action (Morrow and Mason, 2001). The
presence of surfactant in some cases lowers the interfacial tension and thus the capillary
pressure to negligible values. Spontaneous displacement can still occur in this case by
buoyancy or gravity drainage (Schechter, et al., 1994).
The research group of Austad has investigated spontaneous imbibition into chalk
formation material with enhancement by cationic and nonionic surfactants and/or sulfate ions
present in seawater (Standnes and Austad, 2000; Milter and Austad, 1996a, 1996b;
Høgnesen, et al., 2004, 2006). Spontaneous imbibition was most effective with dodecyl
trimethylammonium or amine surfactants. The mechanism is thought to be removal of
adsorbed naphthenic acids through ion-pairing with the cationic surfactant. Capillarity may
dominate during earlier time with gravity dominating later.
Laboratory and field testing of surfactant-enhanced imbibition was investigated for the
dolomite formation of the Yates field (Chen, et al., 2000; Yang and Wadleigh, 2000)). Both
mass balance and CT-scans showed increased oil recovery with 0.35% nonionic and
ethoxylated sulfate surfactants compared to reservoir brine. A single-well injection, soak and
production test showed increased oil recovery with decrease in WOR. The CT-scans showed
10 [SPE 115386]

that with formation brine, the oil recovery was counter-current imbibition but with surfactant the
displacement was dominated by gravity after early time.
An investigation compared the spontaneous recovery of the Yates system with 0.35%
ethoxylated alcohol and dodecyl trimethylammonium bromide (C12TAB, Standnes, et al.,
2002). The superior recovery of C12TAB was interpreted to be due to advancing contact
angle of 32° for C12TAB compared to 107° for the ethoxylated alcohol and 133° for brine.
Laboratory measurements of spontaneous oil recovery were made on the Yates system
or outcrop limestone with alkaline surfactant solutions with 0.05% anionic surfactant and
sodium carbonate (Hirasaki and Zhang, 2004; Seethepalli, 2004; Zhang et al, 2006; Adibhatla
and Mohanty, 2008). There was no recovery (Yates cores) or only little recovery (outcrop
cores) with brine as the imbibing fluid. However, the alkaline surfactant solutions recovered as
much as 60% of the oil. For a given system, temperature is an important factor for wettability
alteration and rate of imbibition oil recovery (Gupta and Mohanty, 2007).
Surfactant aided wettability alteration and spontaneous oil recovery may not have a
significant contribution from capillary pressure if the IFT is reduced to ultra-low values.
However, in these conditions gravity becomes an important contribution (Babadagli, 2001;
Hirasaki and Zhang, 2004, Zhang, et al., 2006; Adibhatia and Mohanty, 20007; Gupta and
Mohanty, 2007, 2008). Fig. 5 illustrates the simulated velocity field during spontaneous
displacement for a cylindrical core immersed in a 0.05% alkaline surfactant solution. The
aqueous phase enters from the sides and oil flows out from the top of the core. The
significance of gravity dominated displacement is that time scales proportionally with the
characteristic length of the matrix, whereas it would scale with the square of the length if the
process is capillary dominated (Hirasaki, and Zhang, 2004).
Najafabadi, et al., (2008) simulated a case of a fractured oil-wet system where negative
capillary pressure inhibited oil displacement from the matrix. Wettability alteration and low IFT
reduced the unfavorable capillary pressure so the pressure gradient from flow in the fractures
could displace oil from the matrix by viscous forces.

Alkaline Surfactant Processes: Wide Region of Ultra-Low IFT


Surfactant processes for enhancing oil recovery are based on achieving ultra-low
interfacial tension (IFT) (e.g., <10-2 mN/m) to either raise the capillary number in forced
displacement or to raise the Bond number for gravity-driven displacement from matrix that is
surrounded by fractures or vugs bearing surfactant solution. For most anionic surfactants,
especially those sensitive to divalent ions, ultra-low IFT occurs only over a limited range of
conditions (Healy et al., 1976; IFT vs. salinity plot shows a narrow region of ultra-low IFT). It
has always been a process-design challenge to maintain or pass through these narrow
conditions of Winsor III or middle-phase microemulsion to have ultra-low IFT during the oil-
recovery displacement process. Recently it was found that blending alkyloxylated sulfate and
IOS surfactants in an alkaline environment produced ultra-low IFT for a wide range of
conditions even when the system has Winsor I phase behavior. This is an important discovery
because surfactant retention is much less for Winsor I compared to Winsor III phase behavior.
These results are reviewed here.
Conventional phase behavior for ultra-low IFT. The understanding of ultra-low IFT in
oil recovery processes was advanced when Healy et al., (1976) explained how the Winsor
definition of equilibrium microemulsion phase behavior (I, II, and III or lower-phase, upper-
phase, and middle-phase microemulsion) described the changes of phase behavior,
solubilization of oil and water, and IFT as a function of salinity for anionic surfactants. The
[SPE 115386] 11

surfactant is able to solubilize an increasing amount of oil and decreasing amount of water as
salinity is increased. The “optimal salinity” determined from phase behavior is the salinity at
which the microemulsion solubilizes equal amounts of oil and water. The optimal salinity at
which equilibrium IFTs between the microemulsion phase and excess oil or excess water
phase become equal (and thus the sum is a minimum) is close to the optimal salinity from
phase behavior. There are correlations between the “solubilization parameters” (ratio of
oil/surfactant, Vo/Vs or water/surfactant, Vw/Vs, by volume) and the IFTs of the microemulsion
with the respective excess phases (Huh, 1979). Thus one can estimate the value of the
equilibrium IFT at the optimal salinity from the value of the solubilization parameters at the
optimal salinity (where they are equal).
Nelson and Pope (1978) recognized that the appearance of a middle-phase
microemulsion (Winsor III) is dependent on the amounts of water, oil, and surfactant present.
Thus they defined the Type III phase environment as the range of salinity at which a middle-
phase microemulsion may exist if one was to scan the water-oil-surfactant ternary diagram.
This distinction is important at very high or very low surfactant concentrations because the
volume of the middle-phase microemulsion is proportional to the surfactant concentration. At
high surfactant concentrations, more of the excess phases is solubilized and thus the excess
phases have smaller volume or is not present. If the surfactant concentration is high enough,
the ‘middle-phase’ microemulsion phase may appear as a single phase at, or near optimal
conditions. On the other hand, at low surfactant concentrations but above the critical micelle
concentration, the volume of the middle-phase microemulsion is minute and its presence may
not be visually detected or sampled for IFT measurements.
The nano-structure of the microemulsion should be recognized to distinguish it from
macro-emulsions or liquid crystal dispersions or phases. Macroemulsions are nonequilibrium
dispersions that change with time or may be in a metasable condition. Liquid crystal phases
are condensed, ordered phases that usually are birefringent (rotate polarized light), viscous,
and tend to inhibit emulsion coalescence (Healey and Reed, 1974). Microemulsions are
equilibrium isotropic phases that may have a bicontinuous structure with near zero mean
curvature at or near optimal conditions (Scriven, 1976). (Microemulsions are oil-swollen
micelles in water at underoptium conditions and reversed micelles in oil at overoptimum
conditions.) It was once thought that it is necessary to have a co-solvent (alcohol) to have a
microemulsion with an anionic surfactant. However, it is now recognized that it is possible to
have microemulsions without alcohol at room temperature by using branched surfactants (Abe,
et al., 1986).
Salinity scan tests are used routinely to screen phase behavior of surfactant
formulations before conducting more time-consuming coreflood tests (Levitt et al., 2006;
Flaaten, et al., 2008). The minimum IFT is correlated with the solubilization parameters at the
optimal salinity. The presence of viscous, structured or birefringent phases and/or stable
macroemulsions is easily observed.
Middle-layer of lower-phase microemulsion. When salinity scan test is conducted at
low surfactant concentrations (e.g., 0.05%), the equilibrium phase behavior appears to go from
a lower-phase microemulsion to an upper-phase microemulsion over a narrow salinity range
(Zhang, et al., 2006: Liu, et al., 2008a). Middle-phase microemulsion is observed only at 3.2%
NaCl (with 1% Na2CO3) in this scan with 0.2% NI blend, a West Texas crude oil, but no
alcohol, Fig. 6. Thus IFT measurements for salinity scans for low surfactant concentration are
usually between the upper and lower phases observed in the sample tubes. The phases may
be (1) lower-phase microemulsion and excess oil (Winsor I), (2) excess brine and upper-phase
microemulsion (Winsor II), or (3) excess brine and excess oil (Winsor III). The value of the
12 [SPE 115386]

IFTs for these systems, measured between the upper and lower phases was not reproducible
until a protocol was developed to add a small volume of intermediate density material to the oil
drop for the spinning drop measurement. In Winsor III, this material would include the middle
phase microemulsion. In Winsor I (lower-phase microemulsion or under-optimum salinity) this
intermediate density material appeared to be a colloidal dispersion in the lower-phase
microemulsion, Fig. 7. This colloidal dispersion formed a middle layer between the excess oil
and lower-phase microemulsion of alkaline surfactant systems with crude oil with TAN of 0.3
mg KOH/g or greater. This middle layer was more opaque than the rest of the lower-phase
microemulsion and can be interpreted to be more oil-rich from its intermediate density. When
the oil/water ratio is increased, the volume of this layer also increased. Because this layer is
not observed in the absence of alkali, it is hypothesized that the dispersed colloidal material is
an oil-and-sodium naphthenate rich microemulsion that is in equilibrium with the remainder of
the lower-phase microemulsion. As shown later, the soap is more lipophilic than the added
surfactant. This layer is not just a macroemulsion dispersion that has not yet coalesced
because the presence of this material affects the value of the IFT.
IFT with and without alkali. The IFT of the system with and without alkali was
measured to test the hypothesis that the soap generated by the alkali is responsible for the
ultra-low IFT in the presence of the colloidal dispersion material. Fig. 8 compares the
measured IFT with and without alkali. In the absence of alkali, the lower-phase microemulsion
was homogeneous and the ultra-low IFT occurred only near the optimal salinity, as expected
for conventional surfactant EOR systems. In the presence of alkali and using the protocol to
assure that a small volume of colloidal dispersion was present, a wider, ultra-low IFT region
was observed, especially for underoptimum conditions. If the colloidal dispersion is not
present as a result of creaming or centrifugation, IFT behavior is similar to that in the absence
of alkali (Liu, et al., 2008a, Fig. 12). Apparently this colloidal dispersion contains surface-
active species responsible for lowering IFT between the lower-phase microemulsion and the
excess-oil phase in a manner similar to behavior of the middle-phase microemulsion between
the excess-brine and excess-oil phases.
Consistent with solubilization parameter. The wide range of ultra-low IFT was
verified by comparison with the Huh correlation (Huh, 1979). The solubilization ratios of the
alkaline NI blend with Yates crude oil are shown in Fig. 9. The IFTs calculated from the Huh
correlation and measured by spinning drop are shown in Fig. 10. Thus the measured IFT is
consistent with the volume of oil that is solubilized into the microemulsion phase(s)/volume of
surfactant. The generality on the wide ultra-low IFT was tested with another crude oil with a
TAN of 4.79 mg KOH/g. The lower-phase microemulsion of this crude oil was too dark to
observe the spinning drop for IFT measurements. The IFT estimated from the Huh correlation
shows this crude oil to also have a wider, ultra-low IFT region of salinities, especially in the
underoptimum region. (Liu, et al., 2008b).

Alkaline Surfactant Processes: Phase Behavior of Soap/Surfactant


Alkali saponifies the naphthenic acid in crude oil in situ to generate sodium
naphthenate, a soap that helps to generate low IFT during the displacement process. Thus an
alkaline surfactant system should be considered as a pseudo two-surfactant system, the
injected surfactant and the soap. The two surfactants will likely have different optimal salinity.
Thus a mixing rule is needed to model how the optimal salinity changes with surfactant and
soap concentrations.
WOR and surfactant concentration. Optimal salinity was observed to be a function of
surfactant concentration and WOR for an alkaline surfactant system, Fig. 11. However, all of
[SPE 115386] 13

these curves can be reduced to a single curve if plotted as a function of the soap/surfactant
ratio, Fig. 12. The latter figure compares the curves of optimal salinity for the TC blend and NI
blend surfactant formulations and the same crude oil.
Mixing rule. Modeling of alkaline surfactant flooding will benefit from a mixing rule for
the optimal salinity. When the TAN was used for the soap content, the experimental data
deviated significantly from the mixing rule of Salager, et al., (1979). An alternative approach to
determine the soap content of crude oil is to extract soap from the crude oil into alkaline,
alcoholic water and titrate for anionic surfactant content by hyamine titration. Fig. 13 shows
that the soap content by such aqueous extraction is about one-half of the TAN, which is
measured by nonaqueous titration. Apparently about one-half of the TAN content is too
hydrophobic to be extracted into the aqueous phase (e.g., asphaltene) and/or is too hydrophilic
to be detected by hyamine titration (Liu, et al., 2008b).
The Salager, et al. (1979) mixing rule was found to be followed reasonably well when
the aqueous titration method was used to quantify the soap content of the crude oil, Fig. 14,
right panel. The expression for the mixing rule is:
log ( Optmix ) = X soap log ( Optsoap ) + (1 − X soap ) log ( Optsurfactant )
where (1)
Soap
X soap =
Soap + Surfactant
The optimal salinity mixing rule is used in UTCHEM for simulation of alkaline surfactant
processes (Mohammadi, et al., 2008). In addition, they found that the optimum solubilization
ratio follows a linear mixing rule:
σ M* = X soap σ soap
*
+ X surfactant σ surfactant
*
(2)

where σ M* , σ soap
*
, and σ surfactant
*
are the optimal solubilization ratio of the mixture, soap, and
surfactant, respectively.
Winsor type phase behavior. The microemulsions with soap and synthetic surfactant
nominally have Winsor type I, III, II phase behavior with increasing electrolyte concentration
(Liu, et al., 2008a). The qualifier ‘nominally’ is used because variations from classic Winsor
behavior exist, e.g., the lower-phase microemulsion in the Winsor I environment may not be a
single homogeneous phase, as discussed earlier. The volume of the middle-phase
microemulsion at optimal conditions is proportional to the surfactant concentration and the
solubilization ratio at optimal conditions. With low surfactant concentrations (< 0.5%) normally
used for ASP processes, middle-phase microemulsion in a salinity scan may be observed over
only a narrow range of salinity. However, if a system has a large surfactant concentration (>
1%) and/or a large solubilization ratio, middle-phase microemulsions may be observed over a
wide range of salinity (Hirasaki, et al., 2005, 2006; Mohammadi, et al., 2008).
IFT measurements. The dependence of the optimal salinity on the soap/surfactant
ratio can be used to explain the difference of minimum equilibrium IFT that may be observed
with different surfactant concentrations and WOR. This also explains why the minimum IFT of
small oil drops on a calcite plate occurred at the optimal salinity of the TC blend surfactant with
zero soap fraction, i.e.,10%-12% NaCl (Zhang, et al., 2006; Fig. 4 and Fig. 13). The water/oil
ratio of the small drops was very high. Equilibrium IFT measurements had lower optimal
salinity because the soap/surfactant ratios were larger due to lower water/oil ratios.
14 [SPE 115386]

The dependence of optimal salinity on the soap/surfactant ratio also explains the
transient minimum IFT observations in spinning drop measurement of a fresh oil drop in fresh
surfactant solution (Liu, 2007). The soap concentration of the oil drop is changing as soap is
being extracted from the small oil drop into the much larger volume of surfactant solution.
Thus the soap/surfactant concentration ratio of the oil drop changes from a large value to near
zero with time and the minimum IFT (in time) occurs when the soap/surfactant ratio of the oil
drop corresponds to the ratio that is optimal for the salinity of the surfactant solution.

Composition Gradients
Displacement of residual oil by surfactant flooding requires reducing the IFT to ultra-low
values such that disconnected oil droplets can be mobilized. The ultra-low IFT generally exists
only in a narrow salinity range near the optimal salinity. During the 1970s and 1980s two
schools of thought developed on how ultra-low IFT could be achieved in the displacement
process (Gupta and Trushenski, 1979). One approach is to either preflush the formation to
reduce the formation salinity to a value near optimal or design the surfactant formulation such
that the optimal salinity is equal to the formation salinity, with the surfactant slug and drive
injected at the formation salinity (Maerker and Gale, 1992). In the former case success was
limited because the more viscous surfactant slug contacted portions of the reservoir which the
preflush bypassed. In the latter case, this problem is avoided because there is no change in
salinity due to dispersive mixing or cross-flow. The other approach is to have a salinity
gradient such that the system has over-optimum salinity ahead and under-optimum salinity
behind the active region. In this case, the salinity profile is certain to pass through the optimal
salinity somewhere in the displacement front region (Nelson, 1981).
Whether the salinity is constant or a salinity gradient is used, the electrolyte composition
is further challenged by divalent ions in the formation brine and ion-exchange from the clays to
the flowing phases (Hill, et al., 1977; Pope, et al., 1978; Glover, et al., 1979, Gupta, 1981). It
was discovered that the surfactant micelles or microemulsion droplets have an affinity for
divalent ions similar to that of the clays and thus act as a flowing ion-exchange medium
(Hirasaki, 1982; Hirasaki and Lawson, 1986). The problem of divalent ions is avoided by use
of an alkali such as sodium carbonate or sodium silicate (Holm and Robertson, 1981).
Salinity gradient. It was demonstrated that with a salinity gradient: (1) ahead of the
active region, the system is over-optimum, surfactant is retarded by partitioning into the oil-
phase, (2) the system passes through the active region of ultra-low IFT (Winsor III) where
residual oil displacement takes place, (3) behind the active region, the system is under-
optimum with lower-phase microemulsion and the surfactant propagates with the water velocity
(Glover et al., 1979; Pope et al., 1979; Hirasaki et al., 1983). Thus the salinity gradient tended
to focus the surfactant near the advancing displacement front where salinity is optimal and the
phase behavior is Winsor III, Fig. 15. Also, the salinity gradient helps to maintain polymer
flowing in the same phase with the surfactant for the Winsor I conditions behind the active
region. The polymer is in the excess-brine phase in the Winsor II & III phase environments
(Gupta, 1981, Tham et al., 1983). The example in Fig. 15 was injected over-optimum only for
illustration of surfactant transport with respect to salinity environment. Over-optimum salinity
environment (Winsor II) can have viscous, high-internal-phase, water-in-oil emulsions (Hirasaki
et al., 1983, Fig. 14) that may be bypassed by the subsequent lower salinity fluids. In practice,
the surfactant slug is injected in near-optimal to under-optimum salinity environment.
Soap/surfactant gradient. It was mentioned earlier that the optimal salinity changes
as the soap/surfactant ratio changes. Thus an alkaline surfactant flood will have a gradient of
optimal salinity due to a gradient in the soap/surfactant ratio unless the soap content is
[SPE 115386] 15

negligible or the surfactant and soap have identical optimal salinity (although surfactant would
likely not be used if the soap had a suitable optimal salinity). A gradient in the soap/surfactant
ratio exists because soap is generated in situ by interaction between the alkali and the
naphthenic acids in the crude oil while the synthetic surfactant is introduced with the injected
fluid.
The role of the soap/surfactant gradient in the ASP process was evaluated with a 1-D
finite difference simulator (Liu et al., 2008a, 2008b). Example composition and IFT profiles,
Fig. 16, show the IFT dropping to ultra-low values in a narrow region of the profile as the
optimal salinity passes across the system salinity, which was constant in this example. There
is only a short distance for the oil saturation to be reduced to a low value before the IFT again
increases and traps any oil that has not been displaced. The oil saturation that will be trapped
is approximately the saturation where the slope of the ultra-low IFT oil-water fractional flow
curve becomes less than the dimensionless velocity of the displacement front (Pope, 1980;
Hirasaki, 1981; Ramakrishnan and Wasan, 1988, 1989). Thus mobility control is important for
displacement efficiency in addition to sweep efficiency for ASP. Finite difference simulation
showed that recovery decreased from 95% to 86% as the aqueous viscosity decreased from
40 cp to 24 cp for an oil with viscosity of 19 cp (Liu, et al., 2008a). This is consistent with a
pair of experiments that differed only in polymer concentration.
The effects of salinity, surfactant concentration, acid number, slug size, and dispersion
on oil recovery are illustrated for a 0.2 PV slug and laboratory scale dispersion (Pe = 500) in
Fig. 17 (Liu, et al., 2008b). The system is that discussed in the IFT and phase behavior
sections, and the black dot in Fig. 17 represents conditions of the successful sand pack
experiment mentioned previously. The effects of surfactant concentration and acid number
(soap content of the crude oil) are combined in a single parameter, the soap fraction at the
waterflood residual oil saturation. The range of salinity for greater than 90% oil recovery is a
function of the soap fraction. The salinity for maximum oil recovery decreases from optimal
salinity of the surfactant to that of soap as the soap fraction increases (straight line in the
figure). The range of salinities for potentially high oil recovery is substantial, especially in the
under optimum region below the optimal line.
If the dispersion is increased to a representative field scale value (Pe = 50)(Lake, 1989)
with constant salinity and 0.2 PV slug, the region of greater than 90% recovery all but
disappears. Dilution by mixing at the front and back of the surfactant slug lowers the
surfactant concentration more than the soap concentration and the propagation velocity of the
soap/surfactant ratio for optimal salinity is greatly retarded. However, if the system is operated
with salinity gradient, high oil recovery is again possible (Liu, et al., 2008b). The lower salinity
of the drive compensates for the lower surfactant concentration such that region of optimal
salinity again propagates with a near unit velocity. In addition, injection of the surfactant slug
and polymer drive with a salinity that is less that the optimal salinity of the surfactant alone
makes it possible to inject the surfactant slug with polymer without separation of the surfactant
and polymer into separate phases (Gupta, 1981; Liu, et al., 2008a). Also, the salinity gradient
avoids the large surfactant retention from microemulsion trapping by the polymer drive (Glover,
et al., 1979; Hirasaki, et al., 1983).

Foam mobility control


Foam is usually considered as a means of mobility control for gas injection processes
such as steam foam or CO2 foam. Foam mobility control for surfactant flooding is a natural
progression since the system already has surfactant present (Lawson and Reisberg, 1980).
Moreover, at high temperatures foam may be favored because polymer degradation is a
16 [SPE 115386]

concern. In fact foam was used for mobility control for alkaline surfactant flooding in China
(Zhang, et al., 2000; Wang, et al., 2001). It has also been used as mobility control for
surfactant aquifer remediation (Hirasaki, et al., 1997, 2000).
ASP foam. The reduction of surfactant adsorption with alkali may result in the polymer
being the most expensive chemical in the ASP process. Experiments in 1-D sandpacks have
shown that an ASP process with the polymer drive replaced by a foam drive is equally efficient.
Fig. 18 is an experiment in which the ASP slug is alternated with equal size slugs of gas. The
foam drive consists of slugs of the better foaming surfactant component (without polymer)
alternated with equal size slugs of gas (Li, et al., 2008). Practically all of the 19 cp oil was
recovered after 1.2 TPV injected but with only 0.6 PV of liquid injected. Experiments with
different sands indicated that foam reduced mobility more in higher permeability media, making
it particularly attractive in layered systems.
ASP foam was used to recover a 266 cp, 4.8 mg KOH/g TAN crude oil, Fig. 19. What
was remarkable is that the apparent viscosity of the displacement process was only 80 cp or
less. Apparently, the viscous oil was being transported as an oil-in-water emulsion with much
less resistance that that of the crude oil.
Sweep of layered sands. Fig. 20 compares sweep in two cases of a 19:1 permeability
contrast layered sandpack initially filled with water dyed green. The sandpack is nearly
completely swept with 1.0 TPV of surfactant alternating with gas (SAG) while the low
permability layer is only one quarter swept with water only (Li, et al., 2008). The sweep
efficiency is compared in Fig. 21 as a function of the pore volumes of liquid injected for SAG,
WAG, and waterflooding.

Conclusions
The technology of surfactant flooding has advanced to overcome many of the past
causes for failures and to reduce the amount of surfactant required. These are summarized as
follows:
1. Surfactant adsorption can be significantly reduced in sandstone and carbonate
formations by injection of an alkali such as sodium carbonate. The alkali also
sequesters divalent ions. The reduced adsorption permits lower surfactant
concentrations. If anhydrite is present, sodium metaborate may offer an alternative
alkali.
2. A wide selection of surfactant structures is now available to meet requirements for
specific applications.
a. Branched alcohol alkoxylate sulfates and sulfonates are tolerant of divalent ions.
Ethoxylation increases optimal salinity. Propoxylation decreases optimal salinity.
In both cases, EO or PO, the optimal salinity decreases with increasing
temperature.
b. Alkyloxylated glycidyl sulfonate is more expensive than sulfate but is stable at
elevated temperatures.
c. Internal olefin sulfonates (IOS) are low-cost, double tailed surfactants.
3. Aqueous solutions of a blend of N67-7PO sulfate and IOS1518 with alkali have a larger
single-phase region extending to higher salinities and calcium ion concentrations than
either alone. This blend, without alcohol, can form a single phase for injection with
polymer but yet form microemulsions with crude oil without forming a gel.
[SPE 115386] 17

4. Soap generated in situ by the alkali is a co-surfactant that can change the phase
behavior of the injected surfactant solution from lower- to middle- to upper-phase
microemulsions. It is lower-phase when injected, middle-phase at the displacement
front, and upper-phase ahead of the displacement front.
5. Injection of the surfactant and polymer at salinity that is under-optimum with respect to
the injected surfactant avoids surfactant/polymer phase separation and microemulsion
trapping.
6. The soap generated in situ by the alkali causes a ‘middle layer’ to form and coexist with
the lower-phase microemulsion, which results in ultra-low IFT over a wide range of
salinity.
7. Anionic surfactants and sodium carbonate can alter wettability for either sandstone or
carbonate formations. Spontaneous oil displacement can occur by gravity drainage.
8. Foam can be used as the drive of the ASP process, in place of the polymer drive.
9. Foam can efficiently sweep layered systems.

Acknowledgment
The authors acknowledge the financial support by DOE grant DE-FC26-03NT15406 and
the Rice University Consortium on Processes in Porous Media. The information and insight
we gained from our long collaboration with Gary Pope and Kishore Mohanty is acknowledged.

References
Abe, M., Schechter, D., Schechter, R.S., Wade, W.H., Weerasooriya, U., and Yiv, S. 1986.
Microemulsion Formation with Branched Tail Polyoxyethelene Sulfonate Surfactants, JCIS, Vol.
114, No. 2, 342-356.
Adams, W.T. and Schievelbein, V.H. 1984. Surfactant flooding carbonate reservoirs, SPE/DOE
12686, presented at Symp. on EOR, Tulsa, OK.
Adibhatia, B. and Mohanty, K.K. 2007. Simulation of Surfactant-Aided Gravity Drainage in Fractured
Carbonates, SPE 106161 presented at the SPE Simulation Symp., Houston, TX, 26-28 Feb.
Adibhatla, B. and Mohanty, K.K. 2008. Oil Recovery from Fractured Carbonates by Surfactant-Aided
Gravity Drainage: Laboratory Experiments and Mechanistic Simulations, SPERE&E, Feb., 119-
130.
Anderson, G.A., Delshad, M., King, C.B., Mohammadi, H, Pope, G.A. 2006. Optimization of Chemical
Flooding in a Mixed-Wet Dolomite Reservoir, SPE 100062 presented at the SPE/DOE IOR
Symp., Tulsa, OK, 22-26 April.
Anderson, W.G. 1987a. Wettability Literature Survey – Part 4: Effects of Wettability on Capillary
Pressure, JPT, Oct., 1283-1300.
Anderson, W.G. 1987b. Wettability Literature Survey – part 5: The Effect of Wettability on Relative
Permeability, JPT, Nov., 1453-1468.
Atkinson, H. 1927. U.S. Patent 1,651,311.
Austad, T., Matre, B., Milter, J., Saevareid, A., and Oyno, L. 1998. Chemical flooding of oil reservoirs 8.
Spontaneous oil expulsion from oil- and water-wet low permeability chalk material by imbibition
of aqueous surfactant solutions, Colloids and Surfaces A. Physicochemical and Engineering
Aspects, Vol. 137,117-129.
18 [SPE 115386]

Babadagli, T. 2001. Scaling of Cocurrent and Countercurrent Capillary Imbibition for Surfactant and
Polymer Injection in Naturally fractured reservoirs, SPEJ, Dec., 465-478.
Bae, J.J. and Petrick, C.B. 1977. Adsorption/Retention of Petroleum Sulfonate in Berea Cores, SPEJ,
Oct., 353-357.
Barnes, J., Smit, J.P., Smit, J.R., Shpakoff, P.G., Raney, K.H., and Puerto, M.C. 2008. Development of
surfactants for chemical flooding at difficult reservoir conditions, SPE 113313, presented at
SPE/DOE IOR Symp., Tulsa, OK, 19-23 April.
Baviere, M., Glenat, P., Plazanet, V., and Labrid, J. 1995. Improved EOR by Use of Chemicals in
Combination, SPERE, Aug., 187-193.
Benton, W.J. and Miller, C.A. 1983. Lyotropic liquid crystalline phases and dispersions in dilute anionic
surfactant-alcohol-brine systems. 1. Patterns of phase behavior, J. Phys. Chem., 87, 4981-
4991.
Bourrel, M. and Schechter, R.S. 1988. Microemulsions and related systems, Surface Science Series,
Vol. 30, Marcel Dekker, Inc., New York.
Buckley, J.S., Takamura, K., and Morrow, N.R. 1989. Influence of Electrical Surface Charges on the
Wetting Properties of Crude Oil, SPERE, Aug., 332-340.
Cayias, J.L., Schechter, R.S., and Wade, W.H. 1977. The Utilization of petroleum Sulfonates for
Producing low Interfacial Tensions between Hydrocarbons and Water, J. Colloid Interface Sci.,
59.1, 31-38.
Chen, H.L., Lucas, L.R., Nogaret, L.A.D., Yang, H.D., and Kenyon, D.E. 2000. Laboratory Monitoring of
Surfactant Imbibition Using Computerized Tomography, SPE 59006 paper prepared for
presentation at the 2000 SPE IPCE, Villahermosa, Mexico, 1-3 February.
Cheng, K.H. 1986. Chemical Consumption During Alkaline Flooding: A Comparative Evaluation,
SPE/DOE 14944 presented at the SPE/DOE Fifth Symposium on EOR, Tulsa, OK, 20-23 April.
Cuiec, L. 1977. Study of Problems Related to the Restoration of the Natural State of Core Samples. J.
of Canadian Pet. Tech., Oct.-Dec., 68-80.
Delshad, M., Najafabadi, N.F., Anderson, G.A., Pope, G.A., and Sepehmoori, K. 2006. Modeling
Wettability Alteration in Naturally Fractured Reservoirs, SPE 100081 presented at the SPE/DOE
Symposium on IOR, Tulsa, OK, 22-26 April.
Dubey, S.T. and Doe, P.H. 1993. Base Number and Wetting Properties of Crude Oil, SPERE, August,
195-200.
Dwarakanath, V., Chaturvedi, T., Jackson, A.C., Malik, T., Siregar, A., and Zhao, P. 2008. Using Co-
Solvents to Provide Gradients and Improve Oil Recovery during Chemical Flooding in a Light Oil
Reservoir, SPE 113965, paper presented at the SPE/DOE IOR Symp., 19-23 April.
Ehrlich, R., Hasiba, H.M., and Raimondi, P. 1974. Alkaline Waterflooding for Wettability Alteration-
Evaluating a Potential Field Application, JPT, Dec. 1335-1342.
Ehrlich, R. and Wygal, R.J. 1977. Interrelation of Crude Oil and Rock Properties with the Recovery of
Oil by Caustic Waterflooding, SPEJ, Aug., 263-270.
Falls, A.H., Thigpen, D.R., Nelson, R.C., Claston, J.W., Lawson, J.B., Good, P.A., Ueber, R.C., and
Shahin, G.T. 1994. Field Test of Cosurfactant-Enhanced Alkaline Flooding, SPERE, Aug. 217-
223.
Fan, T. and Buckley, J.S. 2007. Acid Number Measurements Revisited, SPEJ, Dec., 496-500.
Flaaten, A.K., Nguyen, Q.P., Pope, G.A., and Zhang, J. 2008. A Systematic laboratory Approach to
Low-Cost, high-Performance Chemical Flooding, SPE 113469 paper presented at the 2008
SPE IOR Symposium, Tulsa, OK, 19-23 April.
[SPE 115386] 19

Foster, W.R. 1973. A low-tension waterflooding process, JPT, 25, 205.


Ghosh, O. 1985. Liquid crystal to microemulsion transitions in anionic surfactant-oil-brine systems,
PhD thesis, Rice University.
Glover, C.J., Puerto, M.C., Maerker, J.M., and Sandvik, E.L. 1979. Surfactant Phase Behavior and
Retention in Porous Media, SPEJ, June, 183-193.
Gogarty, W.B. 1977. Oil recovery with surfactants: history and a current appraisal, in Improved Oil
Recovery by Surfactant and Polymer Flooding, D.O. Shah and R.S. Schechter (eds.), Academic
Press, New York, 27-54.
Gupta, R. and Mohanty, K.K. 2007. Temperature effects on Surfactant-Aided Imbibition Into fractured
Carbonates, SPE 110204 presented at the SPE ATC&E, Anaheim, CA, 11-14 November.
Gupta, R. and Mohanty, K.K. 2008. Wettability Alteration of Fractured Carbonate Reservoirs, SPE
113407 presented at the 2008 SPE/DOE IOR Symp., Tulsa, OK, 19-23 April.
Gupta, S.P. 1981. Dispersive Mixing Effects on the Sloss Field Micellar System, SPE/DOE 9782
presented at the SPE/DOE Second Joint Symp. on EOR, Tulsa, OK, 5-8 April.
Gupta, S.P. and Trushenski, S.P. 1979. Micellar Flooding – Compositional Effects on Oil Displacement,
SPEJ, April, 116-128; Trans., AIME, 267.
Healey, R.N. and Reed, R.L. 1974. Physicochemical Aspects of Microemulsion Flooding, SPEJ, Oct.
491-501; Trans. AIME 265.
Healy, R.N., Reed, R.L., and Stenmark, D.G. 1976. Multiphase Microemulsion Systems, SPEJ, Vol. 16,
147-160; Trans. AIME 261.
Hill, H.J., Helfferich, F.G., Lake, L.W., and Reisberg, J. 1977. Cation Exchange and Chemical Flooding,
JPT, Oct., 1336-1338.
Hill, H.J., Reisberg, J., and Stegemeier, G.L. 1973. Aqueous surfactant systems for oil recovery, J.
Pet. Techol., 25, 186-194.
Hirasaki, G.J., 1981. Application of the Theory of Multicomponent, Multiphase Displacement to Three-
Component, Two-Phase Surfactant Flooding, SPEJ, April, 191-204.
Hirasaki, G.J. 1982. Ion Exchange with Clays in the Presence of Surfactant, SPEJ, April, 181-192.
Hirasaki, G.J., Jackson, R.E., Jin, M., Lawson, J.B., Londergan, J., Meinardus, H., Miller, C.A., Pope,
G.A., Szafranski, R., and Tanzil, D. 2000. Field Demonstration of the Surfactant/Foam Process for
Remediation of a Heterogeneous Aquifer Contaminated with DNAPL. NAPL Removal: Surfactants,
Foams, and Microemulsions, S. Fiorenza, C.A. Miller, C.L. Oubre, and C.H. Ward, ed., Lewis
Publishers.
Hirasaki, G.J. and Lawson, J.B. 1986. An Electrostatic Approach to the Association of Sodium and
Calcium with Surfactant Micelles, SPERE, March, 119-130.
Hirasaki, G.J., Miller, C.A., Pope, G.A. 2005. Surfactant Based Enhanced Oil Recovery and Foam
Mobility Control; 2nd Annual Report to DOE, DE-FC26-03NT15406, July.
Hirasaki, G.J., Miller, C.A., and Pope, G.A., 2006. Surfactant Based Oil Recovery and Foam Mobility
Control, Final report to DOE, DE-FC26-03NT15406, June.
Hirasaki, G.J., van Domselaar, H.R., and Nelson, R.C., 1983. Evaluation of the Salinity Gradient
Concept in Surfactant Flooding, SPEJ, June, 486-500.
Hirasaki, G.J. and Zhang, D.L. 2004. Surface Chemistry of Oil Recovery from Fractured, Oil-Wet,
Carbonate Formations, SPEJ, June, 151-162.
Høgnesen, E.J., Standnes, D.C. and Austad, T., 2004: Scaling Spontaneous imbibition of Aqueous
Surfactant solution into Preferential Oil-Wet Carbonates. Energy & Fuels, 18, 1665-1675.
20 [SPE 115386]

Høgnesen, E.J., Olsen, M, and Austad, T. 2006. Capillary and Gravity Dominated Flow Regimes in
Displacement of Oil from an Oil-Wet Chalk Using Cationic Surfactant, Energy & Fuels, Vol. 20,
1118-1122.
Holm, L.W. 1977 Soluble oils for improved oil recovery, in Improved Oil Recovery by Surfactant and
Polymer Flooding, D.O. Shah and R.S. Schechter (eds.), Academic Press, New York, 453-485.
Holm, L.W. and Robertson, S.D. 1981. Improved Micellar/Polymer Flooding with High-pH Chemicals,
JPT, January, 161-171.
Huh, C. 1979. Interfacial Tensions and Solubilizing Ability of a Microemulsion Phase the Coexists with
Oil and Brine, JCIS, Vol. 71, No. 2, 408-426.
Jadhunandan, P.P. and Morrow, N.R. 1995. Effect of Wettability on Waterflood Recovery for Crude-
oil/Brine/Rock Systems, SPERE, Feb., 40-46.
Jennings, H.Y, Jr. 1975. A Study of Caustic Solution-Crude Oil Interfacial Tensions, SPEJ, June, 197-
202.
Jensen, J.A. and Radke, C.J. 1988. Chromatographic Transport of Alkaline Buffers Through Reservoir
Rock, SPERE, Aug., 849-856.
Johnson, C.E., 1976. Status of Caustic and Emulsion Methods, JPT, Jan., 85-92.
Lake, L.W., 1989. Enhanced Oil Recovery, Prentice Hall, Englewood Cliffs, NJ, pg. 166.
Lawson, J.B. and Reisberg, J. 1980. Alternate Slugs of Gas and Dilute Surfactant for Mobility Control
during Chemical Flooding, SPE 8839 presented at the SPE/DOE EOR Symposium, Tulsa, OK,
20-23 April.
Levitt, D.B., Jackson, A.C., Heinson, C., Britton, L.N., Malik, T., Varadarajan, D., and Pope, G.A. 2006.
Identification and evaluation of high-performance EOR surfactants, SPE 100089, presented at
SPE/DOE IOR Symp., Tulsa, OK.
Li, R.F., Yan, W. Liu, S., Hirasaki, G.J., and Miller, C.A. 2008. Foam Mobility Control for Surfactant
EOR, SPE 113910 presented at the 2008 SPE IOR Symp., Tulsa, OK, 19-23 April.
Liu, S. 2007. Alkaline Surfactant Polymer Enhanced Oil Recovery Processes, Ph.D. thesis, Rice
University.
Liu, S., Zhang, D.L., Yan, W., Puerto, M., Hirasaki, G.J., and Miller, C.A. 2008a. Favorable attributes of
alkaline-surfactant-polymer flooding, SPEJ, 13, 5-16.
Liu, S., Li, R.F., Miller, C.A., and Hirasaki, G.J. 2008b. ASP Processes: Wide Range of Conditions for
Good Recovery, SPE 113936 presented at the 2008 SPE IOR Symp., Tulsa, OK, 19-23 April.
Lyklema, J. 1995. Fundamentals of Interface and Colloid Science, Vol. II Interfaces, Academic Press
Inc., San Diego, A3.1-A3.6.
Maerker, J.M. and Gale, W.W. 1992. Surfactant flood process design for Loudon, SPERE, 7, 36-44.
Melrose, J.C. and Brandner, C.F. 1974. The role of capillary forces in determining microscopic
displacement efficiency for oil recovery by waterflooding, J. Canad. Petrol. Tech., 13, 54-62.
Miller, C.A., Ghosh, O., and Benton, W.J. 1986. Behavior of Dilute Lamellar Liquid Crystalline Phases,
Colloids and Surfaces, 19, 197-223.
Milter, J. and Austad, T. 1996a. Chemical flooding of oil reservoirs 6. Evaluation of the mechanisms for
oil expulsion by spontaneous imbibition of brine with and without surfactant in mixed-wet, low
permeability chalk material, Colloids and Surfaces A: Physicochemical and Engineering
Aspects, Vol. 117, 109-115.
Milter, J. and Austad, T. 1996b. Chemical flooding of oil reservoirs 7. Oil expulsion by spontaneous
imbibition of brine with and without surfactant in mixed-wet, low permeability chalk material,
Colloids and Surfaces A: Physicochemical and Engineering Aspects, Vol. 117, 109-115.
[SPE 115386] 21

Mohammadi, H., Delshad, M., and Pope, G.A. 2008. Mechanistic Modeling of
Alkaline/Surfactant/Polymer Floods, SPE 110212 presented at the SPE IOR Symp., Tulsa, OK,
19-23 April.
Morrow, N.R. 1990. Wettability and Its Effect on Oil Recovery, JPT, Dec., 1476-1484.
Morrow, N.R. and Mason, G. 2001. Recovery of oil by spontaneous imbibition, Current opinion in
Colloid & Interface Science, Vol. 6, 321-337.
Najafabadi, N.F., Delshad, M., Sepehrnoori, K., Nguyen, Q.P., and Zhang, J. 2008. Chemical Flooding
of Fracture Carbonates Using Wettability Modifiers, SPE 113369 presented at the SPE/DOE
IOR Symposium, Tulsa, OK, 19-23 April.
Nelson, R.C., 1981. Further Studies on Phase Relations in Chemical Flooding, in Surface Phenomena
in Enhanced Oil Recovery, Shah, D.O., ed., Plnum Press, New York, 73-104.
Nelson, R.C., Lawson, J.B., Thigpen, D.R., and Stegemeier, G.L. 1984. Cosurfactant-Enhanced
Alkaline Flooding, SPE/DOE 12672 presented at the SPE/DOE Fourth Symp. on EOR, Tulsa,
OK April 15-18.
Nelson, R.C. and Pope, G.A. 1978. Phase Relationships in Chemical Flooding, SPEJ, Vol. 18, 325-338.
Novosad, Z. and Novosad, J. 1982. The Effect of Hydrogen Ion Exchange on Alkalinity Loss in Alkaline
Flooding, SPE 10605 presented at the SPE 6th Int. Symp. on Oilfield and Geothermal
Chemistry, Dallas, TX, 25-27 January.
Pope, G.A., Lake, L.W., Helfferich. 1978. Cation Exchange in Chemical Flooding: Part 1 – Basic Theory
Without Dispersion, SPEJ, Vol. 18, No. 6, 418-434.
Pope, G.A., 1980. The Application of Fractional Flow Theory to Enhanced Oil Recovery, SPEJ, June,
191-205.
Pope, G.A., Wang, B. and Tsaur, K. 1979. A Sensitivity Study of Micellar/Polymer Flooding, SPEJ,
Dec., 357-368.
Qutubuddin, S., Miller, C.A., Benton, W.J., and Fort, T., Jr. 1985 Effects of polymers, electrolytes, and
pH on microemulsion phase behavior, in Macro- and Microemulsions, D.O. Shah (ed.), ACS
Symp. Ser. 272, ACS, Washington, D.C., 223-251.
Ramakrishnan, T.S. and Wasan, D.T. 1988. The Role of Adsorption in High pH Flooding, SPE 17408
presented at the SPE California Regional Meeting, Long Beach, CA, 23-25 March.
Ramakrishnan, T.S. and Wasan, D.T. 1989. Fractional-Flow Model for High-pH Flooding, SPERE, Feb.,
59-68.
Reed, R.L. and Healy, R.N. 1977. Some physico-chemical aspects of microemulsion flooding: a
review, in Improved Oil Recovery by Surfactant and Polymer Flooding, D.O. Shah and R.S.
Schechter (eds.), Academic Press, New York, 383-437.
Salter, S.J. 1977. The influence of type and amount of alcohol on surfactant-oil-brine phase behavior
and properties, SPE 6843, presented at SPE Annual Meeting, Denver, CO.
Salager, J.L., Bourrel, M., Schechter, R.S., and Wade, W.H. 1979. Mixing Rules for Optimum Phase-
Behavior Formulations of Surfactant/Oil/Water Systems, SPEJ, Oct., 271-278.
Schechter, D.S., Zhou, D., and Orr, F.M., Jr. 1994. Low IFT drainage and imbibition, J. Pet. Sci. & Eng.,
Vol. 11, 283-300.
Scriven, L.E. 1976. Equilibrium bicontinuous structure, Nature, Vol. 263, No. 5573, 123-125.
Seethepalli, A., Adibhatla, B., and Mohanty, K.K. 2004. Physicochemical Interactions During Surfactant
Flooding of Fractured Carbonate Reservoirs, SPEJ, Dec., 411-418.
Somasundaran, P. and Zhang, L. 2006. Adsorption of surfactant on minerals for wettability control in
improved oil recovery processes, J. Pet. Sci. & Eng., Vol. 52, 198-212.
22 [SPE 115386]

Southwick, J.G. 1985. Solubility of Silica in Alkaline Solutions: Implications for Alkaline Flooding, SPEJ,
Dec., 857-864.
Standnes, D.C. and Austad, T. 2000. Wettability alteration in chalk 2. Mechanism for wettability
alteration from oil-wet to water-wet using surfactants, J. Pet Sci. Eng., Vol. 28, 123-143.
Standnes, D.C., Nogaret, L.A.D., Chen, H.-C.,Austad, T. 2002. An Evaluation of Spontaneous
Imbibition of Water into Oil-Wet Carbonate Reservoir Cores Using a Nonionic and a Cationic
Surfactant, Energy &Fuels, Vol. 16, 1557-1564.
Stegemeier, G.L. 1977. Mechanisms of entrapment and mobilization of oil in porous media, in
Improved Oil Recovery by Surfactant and Polymer Flooding, D.O. Shah and R.S. Schechter
(eds.), Academic Press, New York, 55-91.
Surkalo, H. 1990. Enhanced Alkaline Flooding, JPT, Jan., 6-7.
Sydansk, R.D. 1981. Elevated Temperature Caustic-Sandstone Interactions – Implications for
Improving Oil Recovery, SPE/DOE 9810 presented at the 1981 SPE/DOE Second Joint Symp.
on EOR, Tulsa, OK, 5-8 April.
Taber, J.J. 1969. Dynamic and static forces required to remove a discontinuous oil phase from porous
media containing both oil and water, Soc. Pet. Eng. J., 9, 3-12.
Takamura, K. and Chow, R.S. 1985. The Electrical Properties of the Bitumen/Water Interface – Part II.
Application of the Ionizable Surface-Group Model, Colloids and Surfaces, Vol. 15, 35-48.
Talley, L.D. 1988. Hydrolytic Stability of Alkylethoxy Sulfates, SPERE, 3, 235-242.
Tham, M.J., Nelson, R.C., and Hirasaki, G.J. 1983. Study of the Oil Wedge Phenomena through the
Use of a Chemical Flood Simulator, SPEJ, Oct., 746-757.
Treiber, L.E., Archer, D.L., and Owens, W.W. 1972. A Laboratory Evaluation of the Wettability of Fifty
Oil-Producing Reservoirs, SPEJ, Dec., 531-540.
Trushenski, S.P. 1977. Micellar flooding: sulfonate-polymer interaction, in Improved Oil Recovery by
Surfactant and Polymer Flooding, D.O. Shah and R.S. Schechter (eds.), Academic Press, New
York, 555-575.
Uren, L.C. and Fahmy, E.H. 1927. Factors influencing the recovery of petroleum from unconsolidated
sands by waterflooding, Petrol. Dev. Technol., Petrol. Div. AIME, 318.
Wagner, O.R. and Leach, R.O. 1959. Improving Oil displacement Efficiency by Wettability Adjustment,
Pet. Trans. AIME, Vol. 216, 65-72.
Wang, D., Cheng, J., Yang, Z., Li, Q., Wu, W., Yu, H. 2001. Successful Field Test of the First Ultra-Low
Interfacial Tension Foam Flood. SPE 72147 presented at the SPE Asia Pacific Improved Oil
Recovery Conference, 6-9 October, Kuala Lumpur, Malaysia.
Wang, F.H.L. 1993. Effects of Reservoir Anaerobic, Reducing Conditions on Surfactant Retention in
Chemical Flooding, SPERE, May, 108-116.
Wellington, S.L. and Richardson, E.A. 1997. Low Concentration Enhanced Waterflooding, SPEJ, 2,
389.
Wessen, L.L. and Harwell, J.H. 2000. Surfactant Adsorption in Porous Media, in Surfactants:
Fundamentals and Applications in the Petroleum Industry, Schramm, L.L., ed., Cambridge
University Press, New York.
Yang, H.D. and Wadleigh, E.E. 2000. Dilute Surfactant IOR – Design Improvement for Massive
Fractured Carbonate Applications, SPE 59009, paper prepared for presentation at the 2000
SPE IPCE, Villahermosa, Mexico, 1-3 February.
Zhang, R. and Somasundaran, P. 2006. Advances in adsorption of surfactants and their mixtures at
solid/solution interfaces, Adv. Colloid Interface Sci., Vol. 123-126, 213-229.
[SPE 115386] 23

Zhang, D.L., Liu, S., Puerto, M., Miller, C.A., and Hirasaki, G.J. 2006. Wettability alteration and
spontaneous imbibition in oil-wet carbonate formations, J. Pet. Sci. Eng., Vol. 52, 213-226.
Zhang, J., Nguyen, Q.P., Flaaten, A.K., Pope, G.A. 2008. Mechanisms of Enhanced Natural Imbibition
with Novel Chemicals, SPE 113453 paper prepared for presentation at the 2008 SPE IOR
Symp., Tulsa, OK, 19-23 April.
Zhang, Y., Yue, X., Dong, J., Liu, Y. 2000. New and Effective Foam Flooding To Recover Oil in
Heterogeneous Reservoir. SPE 59367 presented at the SPE/DOE Improved Oil Recovery Symp.,
3-5 April, Tulsa, OK.
Zhao, P., Jackson, A.C., Britton, C., Kim, D.H., Britton, L.N., Levitt, D.B., and Pope, G.A. 2008.
Development of high-performance surfactants for difficult oils, SPE 113432, presented at
SPE/DOE IOR Symp., Tulsa, OK, 19-23 April.
24 [SPE 115386]

10
9 Multi-Phase
8 Region
Phase boundary
7 Clear solution
% NaCl 6 2 clear phases
5 Precipitation
4
1-Phase Cloudy solution
3 * * Cloudy after
Region
2 * 9 months.
1 *
0 *
1:1 4:1 9:1
IOS
N67
N67:IOS (w/w)
Fig. 1. Effect of added NaCl on phase behavior of 3 wt% solutions of
N67/IOS mixtures containing 1 wt% Na2CO3. (Liu, et al., 2008a)

2.5
Without Na2CO3
Adsorption Density

2.0
(10-3*mmol/m^2)

1.5

1.0
With Na2CO3(0.2M,0.3M,0.4M)

0.5

0.0
0.0 0.5 1.0 1.5 2.0 2.5
Residual Surfactant Concentration (mmol/L)
Fig. 2 Static adsorption of TC Blend surfactant on dolomite
sand. (Zhang, et al., 2006)
[SPE 115386] 25

Dimensionless Concentration
1.0 Expr. 1
v=1.2 feet/day
0.9 beta=0.34±0.03 Ex pr. 2
φ = 0.335 Cl -
0.8
Expr. 2
0.7 v=12 feet/day
beta=0.22±0.03 Ex pr. 1
0.6 φ = 0.338

0.5
Calculat ed
0.4 f r om dolom it e
isot her m
0.3
bet a= 0.40
0.2
0.1
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Injected Volume (PV)
Fig. 3 Dynamic adsorption of 0.2% TC Blend surfactant without
Na2CO3 on dolomite sand. (Zhang, et al., 2006)
Dimensionless Concentration

1.0 beta=0.07±0.04
0.9 v=1.2 feet/day
φ = 0.337
0.8
0.7
0.6
0.5
0.4
0.3 Experimental Date for NaCl

0.2 Experimental Date for Surfactant


Simulation Curve for Surfactant
0.1
Simulation Curve for NaCl
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Injected Volume (PV)
Fig. 4 Dynamic adsorption of 0.2% TC Blend/0.3M Na2CO3 on
dolomite sand. (Zhang, et al., 2006)
26 [SPE 115386]

Aqueous Phase Velocity 10.8852

0.09

Distance from Bottom(m) ---> 0.08

0.07

0.06

0.05

0.04

0.03

0.02

0.01

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018
Distance from Center(m) --->

(a)

Oil Phase Velocity 10.8852

0.09

0.08
Distance from Bottom(m) --->

0.07

0.06

0.05

0.04

0.03

0.02

0.01

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018
Distance from Center(m) --->

(b)
Fig. 5. Velocity field in a cylindrical core immersed in surfactant
solution at 10.89 days (a) Aqueous phase (b) Oil phase (from
Gupta and Mohanty, 2007)
[SPE 115386] 27

x= 0.2 0.8 1.4 2.0 2.6 3.2 3.6 4.0 4.5

Excess
oil

Colloidal
dispersion

Lower phase
microemulsion

Fig. 7 View of dispersion region near


Fig. 6 Salinity scan for 0.2% NI blend, 1% Na2CO3 with interface for 2% NaCl sample from
MY4 crude oil for WOR=3 after settling time of 28 days at salinity scan except after 23 days
25°C. x = wt.% NaCl. (from Liu, et al., 2008a) settling. (from Liu, et al., 2008a)

1.E+01
Without Na2CO3
With 1% Na2CO3
1.E+00
IFT(mN/m)

1.E-01

1.E-02

1.E-03

1.E-04
0 1 2 3 4 5 6
Salinity(% NaCl)
Fig. 8 Measured IFT of system with and without Na2CO3
(from Liu, et al., 2008a)
28 [SPE 115386]

1000

Vw /Vs
Solubilization ratio

100

10
Vo/Vs

1
2 2.5 3 3.5 4
NaCl, %
Fig. 9 Measured solubilization ratios of salinity scan
(from Liu, et al., 2008a).

1.E-01
Chun-huh Correlation

Spinning drop measurement


.

1.E-02
IFT, mN/m

1.E-03

1.E-04
2.0 2.5 3.0 3.5 4.0
NaCl, %

Fig. 10 Comparison of the IFT from the solubilization parameter and spinning
drop measurements (from Liu, et al., 2008a)
[SPE 115386] 29

14
Optimal NaCl Conc, % 12

10

4 WOR=10
2 WOR=3
WOR=1
0
0.01 0.1 1 10

Surfactant Concentration, %
Fig. 11 Optimal sodium chloride concentration of TC blend as a function of WOR and
surfactant concentration (settled for more than 6 months). (from Zhang, et al., 2006)

14
WOR=1 (TC Blend)
12
Optimal NaCl Conc., %

WOR=3 (TC Blend)


10 WOR=10 (TC Blend)
8 NI blend
NI Blend
6
TC Blend
4

2
0
1.E-02 1.E-01 1.E+00 1.E+01
Soap/Synthetic surfactant Mole Ratio

Fig. 12 Optimal salinity as a function of soap-surfactant ratio for NI and TC


surfactant blends with MY4 crude oil. (Liu, et al., 2008a)
30 [SPE 115386]

2.5

Acid Numberby Soap Titration (mg KOH/g)


Crude B
2

1.5

1 SWCQ y = 0.5003x - 0.0094


R2 = 0.9945
0.5 MY

CM
OMF
0
0 1 2 3 4 5
Total Acid Number by Nonaqueous Titration (mg KOH/g)

Fig. 13 Acid number by soap extraction as a function of total acid number


by nonaqueous titration (from Liu, et al, 2008b)

(a) Nonaqueous phase titration (b) Soap extraction by NaOH


Acid number= 0.75 mg KOH /g Acid number= 0.44 mg KOH/g
Opt vs Soap Fraction (theory) Opt vs Soap Fraction (exp)
Opt vs Soap Fraction (theory) Opt vs Soap Fraction (exp)
10 10
Optimal Salinity
Optiman Salinity

1 1

0.1 0.1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
Xsoap
Xsoap

Fig. 14. Relationship of optimal salinity and soap mole fraction by different acid number methods
for NI Blend and Yates oil. (from Liu, et al., 2008b)
[SPE 115386] 31

Fig. 15 Oil and surfactant production during experiment with finite, over-
optimum surfactant slug and salinity gradient (Hirasaki, et al, 1983).
32 [SPE 115386]

Fig. 16 Profiles for large slug (0.5 PV) with low dispersion near optimal salinity (2 % NaCl).
(from Liu, et al., 2008b)
[SPE 115386] 33

5.0
4.0
70% 50%
3.0 30%

2.0 90%

90%

1.0
Optimum
70%
Curve
50%
0.5
30%

Sor
X Soap
Acid No.=0.2mg/g, surfactant concentration=0.14%, salinity=4.0% NaCl (over-optimum)
Acid No.=0.2mg/g, surfactant concentration=0.14%, salinity=2.0% NaCl (near-optimum)

Acid No.=0.2mg/g, surfactant concentration=0.14%, salinity=1.0% NaCl (under-optimum)

Fig. 17 Recovery factor with small slug (0.2 PV) and low dispersion (Pe=500).
(from Liu, et al., 2008b)
34 [SPE 115386]

Fig. 18 Displacement profiles for the displacement of MY residual crude oil by ASPF in
40 darcy sandpack. (from Li, et al., 2008)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 2.0
Total PV
0 0.1 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
Liquid PV

NIP Air NIP Air NIP Air IOS Air IOS Air IOS Air IOS Air IOS Air IOS

1 ft/day 20 ft/day

Fig. 19 Profiles of the displacement of 266 cp Crude B with ASP and ASPF. (Li, et al., 2008)
[SPE 115386] 35

SAG, 6 psi, fg=1/3 Water only, 4 psi

0.0 TPV

0.2 TPV

0.4 TPV

0.6 TPV

0.8 TPV

1.0 TPV

Fig. 20 Comparison of SAG with waterflood in 19:1 permeability ratio sandpack.


(Li, et al., 2008)

SAG fg=2/3, 8psi


SAG fg=2/3, 6psi
SAG
SAG fg=4/5, 4psi
1.0
SAG fg=2/3, 4psi
WAG
Sweep Efficiency

0.8 SAG fg=3/4, 4psi


0.6 SAG fg=2/3, 2psi
Waterflood
SAG fg=1/3, 6psi
0.4
SAG fg=1/2, 4psi
0.2 WAG fg=4/5, 4psi
WAG fg=3/4, 4psi
0.0
0 0.5 1 1.5 2 2.5 3 WAG fg=2/3, 4psi
WAG fg=1/2, 4psi
PV's of Liquid Injected
Water fg=0, 4psi
Fig. 21 Sweep in 19:1 permeability contrast sandpack with SAG, WAG, and Waterflood.
(Li, et al., 2008)

You might also like