You are on page 1of 10

SPE 95404

A Study of Branched Alcohol Propoxylate Sulfate Surfactants for Improved Oil


Recovery
Y. Wu, SPE, P. Shuler, SPE, M. Blanco, Y. Tang, and W.A. Goddard, California Inst. of Technology

Copyright 2005, Society of Petroleum Engineers Inc.


This paper was prepared for presentation at the 2005 SPE Annual Technical Conference and
Exhibition held in Dallas, Texas, U.S.A., 9 12 October 2005.
This paper was selected for presentation by an SPE Program Committee following review of
information contained in a proposal submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
for commercial purposes without the written consent of the Society of Petroleum Engineers is
prohibited. Permission to reproduce in print is restricted to a proposal of not more than 300
words; illustrations may not be copied. The proposal must contain conspicuous
acknowledgment of where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

Abstract
This investigation considers one class of anionic surfactant, a
series of branched alcohol propoxylate sulfate surfactants, as
candidates for chemical EOR applications. This experimental
results show that these surfactants may be preferred candidates
for EOR as they can be effective at creating low interfacial
tension (IFT) at dilute concentrations, and without requiring
an alkaline agent or cosurfactant. In addition, some of the
formulations exhibit a low IFT at several percent sodium
chloride concentrations, and hence may be suitable for use in
more saline reservoirs.
Adsorption tests onto kaolinite clay indicate that the loss of
these surfactants can be comparable to or greater than other
types of anionic surfactants.
Surfactant performance was evaluated in oil recovery core
flooding tests. Selected formulations could displace most of
the waterflood residual oil in place even with dilute, 0.2 wt%
surfactant solutions from Berea sandstone cores.
Introduction
Surfactant enhanced oil recovery (EOR) has been investigated
for many years, especially starting in the 1970s and 1980s
when the technology was put on a sound scientific basis.
Unfortunately, the economic reality of the process
performance as experienced in early field trials largely
precluded widespread deployment of this technology.
However, the recent surge in crude oil process has provided
new impetus to consider employing chemical EOR.
The basic physics behind the surfactant flooding EOR
process is that the residual oil dispersed as micron-size ganglia
is trapped by high capillary forces within the porous media.
Increasing the fluid flow viscous forces or decreasing the
capillary forces holding the oil in place are required before the
oil can be pushed through the pore throats and sent on to a

production well. The rule of thumb for a successful surfactant


flood is that the interfacial tension between the crude oil and
the aqueous phase needs to be reduced to ultra-low values,
(target 0.001 mN/m), several orders of magnitude below that
of a typical reservoir brine-oil system.
Besides the requirement to achieve a low in-situ IFT,
another major factor that determines the technical and
economic success of a surfactant flood project is to minimize
the depletion of the injected surfactant, with the major sink
usually from solid adsorption onto clays in the reservoir.
A wide variety of surfactant has been investigated for their
potential efficacy for chemical EOR applications. With this
renewed interest in surfactant EOR, there is now the
opportunity to investigate surfactants not available or not
previously investigated during this earlier development of
chemical EOR technology. In part, for this reason, branched
alcohol propoylated sulfates were selected for this study.
Branched alcohol propoylated sulfates have emerged as an
effective type of surfactant for the removal of nonaqueous
phase liquids (NAPLs) from near surface, aquifer
contaminated sites1,2. This application to remediate shallow
subsurface aquifers by injecting a surfactant solution is a
relatively recent technology. This investigation considers
these same surfactants as EOR agents for oilfield applications.
Propoxylated sulfate surfactants have been shown to create
middle-phase micoemulsions versus crude oils, and
presumably achieve low interfacial tensions3. Another study
demonstrated surfactants as mixtures of ethoxylated and
propoxylated products could be formulated to provide
optimum performance for different oils and process
conditions4. Another motivation for this research is the need
for high performance surfactants and that surfactants with
branched-chain alkyl groups are shown in a recent study to
have lower IFT than those with straight-chain alkyl groups5.
The specific surfactants selected for this study are 18 different
branched alcohol propoxy sulfates, of the Alfoterra mn (m=1,
2, 3, 4, 5, 6; n=3, 5, 8) series supplied by Sasol Corporation.
Experimental
Materials.
The anionic surfactants investigated in this research are
branched alcohol propoxylate sulfate, sodium salt,
manufactured by Sasol North America Inc. The trade name
of these commercial anionic surfactants are Alforterra mn. In
the nomenclature Alfoterra mn, the second digit n (e.g., 5 in
Alfoterra 15) indicates the average number of propoxy

groups in the molecule. The first number (m) is associated


with mainly with the size of the branched alkyl chain.
Alfoterra 1n, 2n, 3n, 4n, 5n, and 6n series have an alkyl chain
of approximately C13, C12,C14, C12-C13, C14-C15, and
C12-C13, respectively. An example molecular structure is
shown in Figure 1.
Methods.
Interfacial Tension (IFT) Measurement.
A matrix of test tube samples were prepared that encompassed
Alfoterra surfactant concentrations from 0.1 to 2 wt%, and in
NaCl brines from 1 to 20 wt%. In order to study equilibrium
phase behavior at oil and aqueous solution interface, the
model oil (n-octane) and surfactant aqueous solution were
mixed in a test tube in 1:1 volume ratio. The test tubes were
shaken at room temperature and left standing for at least two
weeks to achieve phase equilibrium.
The IFT between the top oil layer and bottom water layer
was measured obtained by a spinning drop tensiometer6,
Model 510 from Temco. An oleic phase drop (2 l) was
placed into a glass tube containing the aqueous phase, and
spun at high speed. Rotation continued until reaching an
equilibrium condition (typically in less than 2 hours), as
indicated by no drop shape change for 30 minutes at the test
temperature of 30 oC.
Adsorption measurement onto kaolinite.
Adsorption of Alfoterra surfactants onto kaolinite clay was
measured with this procedure: (1) kaolinite clay was heated in
an oven at 120 C for 2 hours in order to remove water and
any other adsorbed materials, (2) surfactant solutions were
prepared in 2.0 wt.% NaCl aqueous solution at three different
concentrations: 0.5, 1.0 and 2.0 wt.%, (active), (3) kaolinite
clay samples and surfactant solutions were mixed in 1:20 mass
ratio in a test tube, and then shaken for 12 hours at room
temperature to reach adsorption equilibrium. (4) the test tube
next was centrifuged to separate the solution and clay. (5) the
equilibrium surfactant concentration was analyzed in order to
determine the amount of chemical adsorbed.
Oil displacement coreflood experiments.
The tertiary oil recovery by surfactant solutions was evaluated
by core flooding tests using 1x 12 Berea sandstone cores
(brine permeability of approximately 300 md).
The
procedures are as follow: (1) saturate a dry core with brine and
determine the porosity of the core from the weight increase.
(2) flood the core with n-octane (used as model oil), and
collect the effluent until there is no more noticeable brine
production. (3) waterflood the core with the brine until no
more n-octane is produced. Based on the injected/produced
volumes in Steps 2 and 3, calculate the waterflood residual oil
saturation, the target for tertiary recovery.
The fluid
saturations calculated via changes in core weight and fluid
densities were consistent with the those based on the produced
fluid volumes. (4) inject test chemical solution into the core.
Collect the produced fluids and measure the cumulative noctane volume versus time until no more n-octane is produced.

SPE 95404

RESULTS AND DISCUSSION


Interfacial Tension Measurements; Initial Screening
Nine Alfoterra surfactants (2 wt%, active basis) were
selected for IFT screening testing at different NaCl
concentrations. Figures 2 - 10 present IFT data for these
Alfoterra surfactants over a range of aqueous phase salinities
(0 wt% - 6 wt.% NaCl) versus n-octane as the oil phase. The
IFT was measured with formulations that contained surfactant
only and also samples containing 1 wt% isopropanol as a cosurfactant. In all of these cases n-octane is the oil phase.
Figure 2 shows IFT results for Alforterra 23. The IFT of
this surfactant decreases dramatically with an increase of
salinity; at 6 wt.% NaCl, its IFT is as low as 0.003 mN/m. The
IFT at even higher salinity will be described later. At 3 wt%
NaCl and higher, IFT values of Alforterra 23 with and
without a co-surfactant are almost the same.
IFT results of
Alforterra 25 and Alforterra 28 surfactants are shown in
Figures 3 and 4, respectively. For both of these two
surfactants, the IFT values indicates an optimal salinity of 3
wt.% NaCl. The lowest IFT is about 0.01 mN/m. Again, the
effect of co-surfactant on IFT if any, is very small.
IFT results of Alforterra 33 and Alforterra 35 surfactants
are shown in Figures 5 and 6, respectively. At low salinity (at
or below 1 wt.% NaCl), addition of iso-propanol has no effect
on IFT, but this effect is significant above that salinity. For
Alforterra 33, addition of iso-propanol shows its IFT
minimum (as low as 0.005 mN/m) at 3 wt.% salinity, however
for Alforterra 35, iso-propanol increases the IFT.
Figure 7 shows the IFT of Alforterra 38 has a fairly broad
minimum over a range of salinities. This could be an
advantage because this behavior makes Alforterra 38 a
candidate for reservoirs with a wide range of salinity. The IFT
of this surfactant is as low as 0.008 mN/m, and the effect of
iso-propanol on IFT is very small.
IFT results for surfactants Alforterra 43, 45 and 48 are
shown in Figures 8, 9 and 10. In Figure 8 the IFT for
Alforterra 43 against n-octane is close to 0.1 mN/m, which is
not as low as that found with other Alforterra surfactants.
With addition of co-surfactant, iso-propanol, the IFT is even
higher. In Figure 9, the IFT for Alforterra 45 decreases to
less than 0.01 mN/m with an increase of salinity to 6 wt.%
NaCl. With addition of co-surfactant, its IFT decreases to as
low as 0.002 mN/m. IFT results of an Alforterra 48 aqueous
solution against n-octane with and without addition of cosurfactant, iso-propanol, are shown in Figure 10. The IFT
without co-surfactant remains almost a constant at about 0.2
mN/m; whereas with addition of iso-propanol, the IFT is
lowest at 3.0 wt.% (approximately 0.01 mN/m).
Interfacial Tension Measurement; High Salinity
In order to explore effect of salinity on IFT over a wider
range, 3 of these branched alcohol propoxylate sulfate anionic
surfactants, Alforterra 23, Alforterra 33 and Alforterra 43,
were selected to measure IFT of their aqueous solution against
n-octane at a salinity up to 20 wt.% NaCl. The previous IFT
data with these 3 surfactants (each with an average of 3 PO
groups) indicate the optimal salinity could be even above the 6
wt% NaCl salinity investigated before.

SPE 95404

Figure 11 shows IFT curves of Alforterra 23 aqueous


solution against n-octane versus salinity. The sharp minimum
in the IFT indicates the optimal salinity for Alforterra 23
surfactant is near 6.0 wt.% NaCl. It can be seen in this figure
that there is little effect from the presence of co-surfactant.
Simialr trends were found for Alforterra 33 surfactant, as
shown in Figure 12. The lowest IFT for Alforterra 33 is
0.005 mN/m at salinity of 6.0 wt.% NaCl without addition of
co-surfactant, and 0.002 mN/m at 3.0 wt.% with addition of
iso-propanol. The IFT results of Alforterra 43 are shown in
Figure 13. The optimal salinity for Alforterra 43 without
addition of co-surfactant is also 6.0 wt.%, however, its
minimum IFT of 0.1 mN/m is not as low as that for
surfactants Alforterra 23 and Alforterra 33. Above the
optimal salinity, IFT values increases rapidly for thiese
surfactant systems.
The strong sensitivity of IFT to salinity is expected , given
that these surfactants are anionic. These results suggest that
the so-called optimal salinity (salinity of minimum IFT) can
be as high as several percent NaCl concentration. Hence these
surfactants may be included in an EOR design for relatively
saline reservoirs.
In addition, the surfactants with more
propoxylated groups (the higher second number of the product
code indicates more PO groups) have a lower optimal salinity.
This is consistent with the notion that more PO groups make
the surfactant more hydrophobic.
Another trend is the addition of iso-propanol as a
cosurfactant has little effect. Possible exceptions are higher
IFT values with Alforterra 35, and IFT further reduced with
cosurfactant for Alforterra 33 and 38.
Interfacial Tension Measurement; Low Concentration
The IFT was also measured at much lower surfactant
concentrations, 0.2, or even 0.1 wt% (active) at the expected
optimal salinity. Again, the measurement used n-octane as the
oil phase and were performed at ambient temperature. The
results were listed in Table 1 and Table 2. Low IFTs at these
low surfactant concentrations would be attractive for further
study towards a field application.
Some follow-up IFT studies were carried out for 6 other
surfactants from this series at lower concentration. The IFT
results for the surfactants Alfoterra 13, 15, 18, 63, 65, and 68
at low surfactant concentration (0.20 wt.%) are shown in
Figures 14 and 15.
Samples Alfoterra 13, 15, and 18 were included later in
this testing program because they showed relatively low
adsorption in kaolinite clay adsorption test.
Samples
Alfoterra 63, 65, and 68 were added later because the
supplier indicated that these 3 surfactants might be the least
expensive to manufacture in large volumes. These IFT tests
screened the performance of these 6 products at 0.2 wt%
concentration (active). Because previous results already
demonstrated some of the other Alfoterra products could
achieve quite low IFT at low surfactant concentration, we
focused immediately on more severe test conditions.
As with the 2.0 wt% Alfoterra products screened
previously, the optimal salinity in general increases with an
increase in the number of propoxy groups (increase of
hydrophobicity). The IFT results indicate the optimal salinity
for some of these 6 Alfoterra surfactants in Figures 14 and 15

may exceed even 10 wt% salinity. None of the measured IFT


values are especially low, as compared to the best Alfoterra
systems reported earlier.
It is encouraging that for several cases, the IFT is quite low
at all three surfactant concentrations. The Alfoterra 23
perhaps has the best results, followed by the Alfoterra 45 at
6.0 wt% NaCl and Alfoterra 28 at 3.0 wt% NaCl and without
iso-propanol addition. In some cases, the IFT does not
increase monotonically with a decrease in the Alfoterra
surfactant concentration. This behavior could be associated
with experimental errors in the IFT measurement and/or
reflect that the optimal salinity can shift with a change in
surfactant concentration.
IFT Measurements; Crude Oil Systems
All the previous IFT results were measured by use of n-octane
as oil phase. While n-octane can be a reasonable proxy for
crude oil for phase behavior/IFT studies, measurement with
actual crude oil is more relevant. Two different crude oil
samples were selected for IFT measurements. One is Crude
Oil #4, which has a density of 0.857 g/cc, and the other one is
a GOM Oil, whose density is 0.812 g/cc at 45 C. We selected
a test temperature as 45 C and brine salinity of about 3 wt%
NaCl, representative of conditions at the Oil #4 reservoir. The
second oil, GOM (from the Gulf of Mexico), is a lighter,
somewhat waxy crude oil. The measured IFT results for
different Alfoterra surfactants and the two crude oils at 45 C
are listed in Tables 3 and 4. These results indicate the IFT
values are lower for the crude oils than the n-octane at these
same conditions. We speculate that the IFT values could be
even lower versus these crude oils if the brine salinity was
adjusted to be optimal. This is based on previous results of
this study where n-octane and Alfoterra surfactants at ambient
temperature show an optimal salinity of several percent NaCl.
Molecular Dynamic simulation of surfactant/water/oil
A full atomistic simulation of water/octant interface mediated
by an Alfoterra type of surfactant was carried out to gain
some insights into the phase behavior/IFT of these surfactants.
These particular simulations investigated the effect of the
number of propylene oxide units in the surfactant molecule on
the interfacial tension between water and oil, which is noctane in this study. Each simulated system has the same
composition with 600 water molecules, 60 n-octane (C8)
molecules and 16 surfactant molecules with a C16 alkyl chain.
The number of propylene oxide groups considered was
1(PPO_1), 3(PPO_3), 5(PPO_5) and 8(PPO_8).
The
methodologies to perform the Molecular Dynamic (MD)
calculations to arrive at an equilibrated structure and IFT value
are documented elsewhere7. A snapshot of the simulated
distribution of the molecules is shown in Figure 16.
Table 5 presents the the total densities of the equilibrated
systems (which are more or less the same), their different
dimensions, and IFT values. Compared to PPO_1 and PPO_8,
surfactants PPO_3 and PPO_5 induce smaller cross-sectional
area; the area occupied by individual surfactant molecules of
PPO_3 and PPO_5 is 38% smaller than PPO_1 and PPO_8.
This means surfactants PPO_3 and PPO_5 more effectively
reduce the interfacial area and create a smaller IFT value

compared with PPO_1 and PPO_8. The notion of having an


optimum number of PO groups for low IFT seems logical.
For simplicity, these simulations did not include any added
dissolved salt. Direct comparison to experimental data is
difficult, but note that the measured IFT values are 1 - 5 mN/m
in fresh water when measured for the Alfoterra 3n series, for
example (see Figures 5 7), are comparable to the calculated
values via these MD calculations.
Adsorption tests onto kaolinite
Adsorption results for this test series at different equilibrium
concentration are listed in Table 6, and show three trends:
(1) Adsorption amount increases with increase of
equilibrium concentration of surfactant. There are only three
exceptions with Alfoterra 15, 25 and 28. For Alfoterra 15
and 25, with an apparent adsorption maxima at a bulk
concentration of about 1 wt%. For Alfoterra 28, adsorption
decreases with increase of equilibrium concentration.
(2) The general trend for adsorption sorted by Alfoterra
surfactant
series
from
lowest
to
highest
is:
Alfoterra 1ns < 2ns < 6ns < 5ns < 4ns < 3ns (n=3, 5, 8).
(3) Generally adsorption decreases with an increase of the
number of propoxy (PPO) groups in the surfactant molecule.
There are exceptions, such as with surfactants Alfoterra 1n
and Alfoterra 3n at concentrations about 1 and 2 wt.%. For
Alfoterra 1n series, Alfoterra 15 shows highest adsorption at
about 1 wt% equilibrium concentration, but it shows lowest,
even a negative adsorption at 2 wt% concentration. For the
Alfoterra 3n series, Alfoterra 35 shows highest adsorption at
1 and 2 wt% concentrations.
The mechanisms of surfactant adsorption at the kaolinite
/liquid interface has been previously studied 8-12. It is
recognized that two steps, more or less separated, can describe
the formation of the adsorption layer. Adsorption onto
minerals at low equilibrium concentration is generally due to
the interaction between the polar head of the amphiphile
molecule and some specific sites of the surface, e.g., H-bonds,
electrostatic forces. For higher concentration, aggregates are
formed at the interface as a result of lateral interactions
between hydrophobic chains. This aggregation is due to the
same forces as those responsible for bulk micelle formation. In
Alfoterra surfactant molecules, more propoxy groups make
the surfactant more hydrophobic, the interaction between
hydrophobic chains, therefore, become stronger, which
weakens the interaction between the polar head of the
surfactant molecules and the specific sites on the kaolinite clay
surface. This might explain a decrease of adsorption of
Alfoterra surfactants with an increase of the number of
propoxy groups in their molecules.
On the other hand, adsorption of surfactant on mineral
surface also depends on many other factors.
Several
physicochemical processes can be expected to occur, such as
hydrolysis of surface species, ion-exchange, electrostatic
adsorption and dissolution of the clay constitute, and
adsorption or precipitation of resultant complexes.
Similar to some other studies of surfactant adsorption from
aqueous solutions onto kaolinite clay, an adsorption maxima
was found in some cases here. Recently, a mechanism
responsible for the adsorption maximum has been proposed 1315
. When aqueous surfactant solution contacts with kaolinite

SPE 95404

clay, the concentration of Ca2+ and Mg2+ in the solution


increases due to the ion-exchange process. At low surfactant
concentration, surfactant molecules exist in the solution as
monomers. Meanwhile, the adsorption of surfactant increases
with increase in the concentration. At high concentration,
when the adsorption density of surfactant on kaolinite surface
is high enough to make the concentration product of the
adsorbed surfactant anion and Ca2+ and Mg2+ greater than their
solubility products, a precipitate will form on the kaolinite
surface. However, with a further increase in surfactant
concentration in aqueous solution, surfactant molecules
aggregate and micelles are formed in the bulk solution, which
enhance the solubility of the precipitate. Consequently, the
micelles compete with the surface of kaolinie for the adsorbed
molecules and dissolve some of the precipitate on the surface.
As a result, the adsorption decreases at high concentrations.
The adsorption levels for the Alfoterra surfactant series
onto kaolinite (roughly 5 50 mg/gm) are comparable or
greater than that reported for some other anionic surfactants
tested under similar conditions (moderate NaCl brine, near
ambient temperature). For example, one investigation reports
adsorption onto kaolinite of 5 -10 mg/gm for a series of linear
alkyl benzene sulfonates, and another study for a petroleum
sulfonate shows 5 50 mg/gm. 16,17
Core Flooding Oil Recovery Tests
Alfoterra 38:
Based on the above favorable IFT measurements, we selected
the Alfoterra 38 for a tertiary oil recovery core flooding test.
This surfactant displayed a broad region of low IFT versus
salinity, with the optimum salinity at about 3 wt% NaCl.
The run conditions for this test include:
Connate brine: 10 wt% NaCl
Surfactant slug: 2 wt% surfactant in 1 wt% NaCl, 0.25 PV
Polymer drive: Alcoflood 1235, 350 ppm in 1 wt% NaCl
Flow rate: 3 ft/day
Note that in this experiment the connate brine is quite high,
10 wt% NaCl, and the chemical slugs are formulated at a
much lower 1 wt% NaCl. This was designed to test two things
1) that an Alfoterra surfactant formulation could recover
significant tertiary oil when the in-situ brine is fairly high
salinity, and 2) that the so-called salinity gradient approach
is a reasonable strategy. That is, the in-situ brine is above the
optimal salinity (about 3 wt% NaCl) for the chemical
formulation, and the injected solutions are below this optimum
salinity. The tertiary oil recovery curve is shown in Figure 17.
After surfactant flooding, the tertiary oil recovery was found
to be more than 50% of the waterflood residual.
Alfoterra 23:
Based on previous favorable IFT measurements with this
Alfoterra 23 surfactant at 0.2 wt%, it was selected for a
coreflood oil displacement experiment. It appears to have a
broad salinity region of low IFT, with the optimum salinity at
approximately 6 wt% NaCl.
The run conditions for this test include:
Connate brine: 10 wt% NaCl
Surfactant slug: 0.2 wt% Alfoterra in 5 wt% NaCl, 0.8 PV
Polymer drive: Alcoflood 1235, 350 ppm in 1 wt% NaCl
Flow rate: 3 ft/day

SPE 95404

The tertiary oil recovery curve is shown in Figure 18. After


surfactant flooding, the tertiary oil recovery was found to be
close to 40% in volume.
The purpose of this core flooding test is to confirm that a
low concentration of Alfoterra 23 could recover significant
tertiary oil.
In particular, the design of the coreflood
experiment was intended to have the displacement occur at a
favorable, low IFT environment by having the large surfactant
slug at 5 wt% NaCl, just below the optimal salinity. Finally,
the chase, polymer slug, was made up in a low salinity (1 wt%
NaCl) brine so as to be well below optimum and complete an
overall salinity gradient design.
While the tertiary oil recovery is only mediocre, the
chemical efficiency of this coreflood experiment is relatively
good. The mass of Alfoterra 23 surfactant injected is less than
that used in the previous, Alfoterra 38 coreflood test, and the
percent tertiary recovery being just slightly less with the
Alfoterra 23. As with the previous oil recovery experiment,
significant residual oil recovery occurred even with the
connate brine being a relatively high, 10 wt% NaCl.
Conclusions
Key findings from this study include:
1. Branched alcohol propoxylate sulfate surfactants at
concentrations as low as 0.1 wt% can create an IFT of 0.01
mN/m, or less, between a brine and n-octane or a crude oil.
2. The optimal salinity (minimum IFT condition ) is several
percent NaCl for the products tested. The optimal salinity
decreases with an increase of propoxy groups.
3. Generally, adsorption of these surfactants on kaolinite clay
decreases with an increase of the number of propoxy
groups. The amount adsorbed is in a range of 3 - 25 mg
per gram kaolinite at equilibrium concentrations 0.4% wt
or lower; and 4 - 65 mg per gram clay at around 2% wt
concentration.
4. The tertiary oil recovery in core flooding tests was
approximately half of waterflood residual oil with
surfactant concentrations as low as 0.2 wt%.
Acknowledgements
This work was conducted at the California Institute of
Technology and financially supported by the Department of
Energy of Energy grant DE-FC26-01BC15362. The authors
thank Sasol for supplying surfactant samples, and Chevron
and Akzo Nobel for their collaboration on this project.
References
1. Jayanti, S., et. al.: Laboratory Evaluation of Custom-Designed
Surfactants to Remediate NAPL Source Zones, Environ. Sci.
Technol., 36, 5491-5497, 2002.
2. Coi, K.C., et.al.: Recent Advances in Surfactant Remediation
of Contaminated Soils, paper SPE 52732, presented at
SPE/EPA E&P Environmental Conference, Austin, TX, 1-3
March, 1999.
3. Aoudia, M., Wade, W., and Weerasooriya, Optimum
microemulsions formulated with propoxylated Guerbet alcohol
and propoxylated tridecyl alcohol sodium sulfates
.J. Dispersion Sci. Technol., 16, 115-135, 1996.
4. Minana-Peez, M., Graciaa, A., Lachaise, J., and Salager, J.,
Solubilization of polar oils with extended surfactants. J.
Colloids Surf. A, 100, 217-224, 1995.

5. Rosen, M.J., Wang, H., Shen, P., and Zhu, Y.: Ultralow
Interfacial Tension for Enhanced Oil Recovery at Very Low
Surfactant Concentrations, Langmuir, 21, 3749 3756, 2005.
6. Cayias, J. L.; Schecher, R. S.; Wade, W. H.: The
Measurement of Low Interfacial Tension via the Spinning
Drop Technique, ACS Symposium Series, No 8, 234, 1971.
7. Jang, S.S., Lin, S.T., Maita, P.K., Blanco, M., Goddard III,
W.A., Shuler, P.J., and Tang, Y.: Molecular Dynamics
Study of a Surfactant-Mediated Decane-Water Interface:
Effect of Molecular Architecture of Alkyl Benzene
Sulfonate, J. Phys Chem. B, 108, 12130-12140, 2004.
8. Cases, J. M.; Villieras, F. Thermodynamic model of ionic and
nonionic surfactants adsorption-abstraction on heterogeneous
surfaces. Langmuir (1992), 8(5), 1251-64.
9. Tiberg, Fredrik; Joensson, Bengt; Tang, Ji-an; Lindman, Bjoern.
Ellipsometry Studies of the Self-Assembly of Nonionic
Surfactants at the Silica-Water Interface: Equilibrium Aspects.
Langmuir (1994), 10(7), 2294-300.
10. Luciani, L.; Denoyel, R. Adsorption of polydisperse surfactants
on solid surfaces: an ellipsometric study. Journal of Colloid
and Interface Science (1997), 188(1), 75-80.
11. Bohmer, Marcel R.; Koopal, Luuk K.; Janssen, Rob; Lee, Ellen
M.; Thomas, Robert K.; Rennie, Adrian R. Adsorption of
nonionic surfactants on hydrophilic surfaces. An experimental
and theoretical study on association in the adsorbed layer.
Langmuir (1992), 8(9), 2228-39.
12. Manne, S.; Schaeffer, T. E.; Huo, Q.; Hansma, P. K.; Morse, D.
E.; Stucky, G. D.; Aksay, I. A. Gemini Surfactants at SolidLiquid Interfaces: Control of Interfacial Aggregate Geometry.
Langmuir (1997), 13(24), 6382-6387.
13. Hanna, H. S.; Somasundaran, P.. Physico-chemical aspects of
adsorption at solid/liquid interfaces. II. Mahogany
sulfonate/berea sandstone, kaolinite. Sch. Mines, Columbia
Univ., New York, NY, USA. Editor(s): Shah, Dinesh O.;
Schechter, Robert S. Improved Oil Recovery Surfactant
Polym. Flooding, [Pap. AIChE Symp.] (1977), Meeting Date
1976, 253-74.
14. Yang, C.-Z.; Yan, H.-K.; Li, G.-Z.; Cui, G.-Z.; Yuan, H.
Fundamental and Advances in Combined Chemical Flooding
(in Chinese), p. 174, Yu, G.-Y. et al. Ed. China Petroleum Press,
Beijing, 2002.
15. Mukerjee, Pasupati; Anavil, Aroonsri. Adsorption of ionic
surfactants to porous glass. Exclusion of micelles and other
solutes from adsorbed layers and the problem of adsorption
maxima. ACS Symposium Series (1975), 8 (Adsorption
Interfaces, Pap. Symp., 1974), 107-28.
16. Barakat, Y.; El-Mergawy, S. A.; El-Zein, S. M.; Mead, A. I.
Adsorption of alkylbenzene sulfonates onto mineral surfaces.
Indian Journal of Chemical Technology (1995), 2(3), 162-6.
17. Baviere, M., Ruaux, E., and Defives, D.: Sulfonate
Retention by Kaolinite at High pH: Effect of Inorganic
Anions, paper SPE 21031, SPE International Chemitry
Symposium, Ananheim, CA, 20-22 February, 1991.

SPE 95404

H3C
H3C

CH4

CH2

8 ~ 10

(CH2)2
CH

CH3
O

CH2

CH2

CH

SO3 Na+

Figure 1. Molecular Structure of Branched Alcohol Propoxylate Sulfate, Sodium Salt.


10.00

10.00

No iso-propanol

1.00

1% iso-propanol

0.10

0.10

0.01

0.01

0.00
0.00

No iso-propanol

1% iso-propanol

IFT (mN/m)

IFT (mN/m)

1.00

1.00

2.00

3.00 4.00
NaCl%w t.

5.00

6.00

0.00
0.00

7.00

Figure 2. IFT for Alfoterra 23 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)
10.00

1.00

2.00

3.00 4.00
NaCl%w t.

5.00

6.00

7.00

Figure 3. IFT for Alfoterra 25 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)
10.00
No iso-propanol

No iso-propanol

1.00

1% iso-propanol

1.00
IFT (mN/m)

IFT (mN/m)

1% iso-propanol
0.10

0.01

0.01

0.00
0.00

0.10

1.00

2.00

3.00 4.00
NaCl%wt.

5.00

6.00

0.00
0.00

7.00

3.00

4.00

5.00

6.00

7.00

Figure 5. IFT for Alfoterra 33 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)
10.00

1.00

1.00
IFT (mN/m)

10.00

IFT (mN/m)

2.00

NaCl%w t.

Figure 4. IFT for Alfoterra 28 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

0.10

0.01

1.00

No iso-propanol
1% iso-propanol

0.10

0.01

No iso-propanol
1% iso-propanol

0.00
0.00

1.00

2.00

3.00

4.00

5.00

6.00

7.00

0.00
0.00

1.00

2.00

3.00

4.00

5.00

6.00

7.00

NaCl%wt.

NaCl%w t.

Figure 6. IFT for Alfoterra 35 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

Figure 7. IFT for Alfoterra 38 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

10.00

10.00

1.00

1.00

No iso-propanol

IFT (mN/m)

IFT (mN/m)

SPE 95404

0.10

0.01

1% iso-propanol

0.10

0.01

No Iso-propanol
1% iso-propanol

0.00
0.00

1.00

2.00

3.00

4.00

5.00

6.00

0.00
0.00

7.00

1.00

NaCl%w t.

3.00 4.00
NaCl%w t.

5.00

6.00

7.00

Figure 9. IFT for Alfoterra 45 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

Figure 8. IFT for Alfoterra 43 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

10.00

2.00

10.00
No iso-propanol

1.00

1% iso-propanol

IFT (mN/m)

IFT (mN/m)

1.00

0.10

0.01

0.00
0.00

0.10

No iso-propanol

0.01

1% iso-propanol

0.00
1.00

2.00

3.00

4.00

5.00

6.00

7.00

0.0

5.0

NaCl%w t.

1.00
IFT (mN/m)

1.00
IFT (mN/m)

10.00

No iso-propanol

25.0

10.00

0.01

20.0

Figure 11. IFT for Alfoterra 23 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

Figure 10. IFT for Alfoterra 48 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

0.10

10.0
15.0
NaCl%wt.

0.10

0.01

No iso-propanol

1% iso-propanol

1% iso-propanol
0.00

0.00
0.0

5.0

10.0

15.0

20.0

25.0

NaCl%w t.

0.0

5.0

10.0

15.0

Figure 12. IFT for Alfoterra 33 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

20.0

25.0

NaCl%w t.

Figure 13. IFT for Alfoterra 43 Surfactant


(2.0 wt.%, 25 C, Oil phase: n-Octane)

SPE 95404

Table 1. IFT Results of Alfoterra Surfactants


at 0.2 wt.% (Oil Phase: n-Octane)
Surfactant

Alfoterra 23
Alfoterra 23
Alfoterra 28
Alfoterra 28
Alfoterra 33
Alfoterra 33
Alfoterra 38
Alfoterra 38
Alfoterra 38
Alfoterra 45
Alfoterra 45
Alfoterra 48

IPA
(wt.%)
0
0.1
0
0.1
0.1
0
0
0.1
0.1
0
0.1
0.1

NaCl
(wt.%)
6.0
6.0
3.0
3.0
3.0
6.0
3.0
3.0
6.0
6.0
6.0
3.0

Table 2. IFT Results of Alfoterra Surfactants


at 0.1 wt.% (Oil Phase: n-Octane)

IFT
(mN/m)
0.009
0.006
0.040
0.019
0.006
0.111
0.081
0.121
0.249
0.012
0.011
0.014

Surfactant
Alfoterra 23
Alfoterra 23
Alfoterra 28
Alfoterra 28
Alfoterra 33
Alfoterra 33
Alfoterra 38
Alfoterra 38
Alfoterra 38
Alfoterra 45
Alfoterra 45
Alfoterra 48

IPA: iso-propanol .

IPA
(wt.%)
0
0.05
0
0.05
0.05
0
0
0.05
0.05
0
0.05
0.05

NaCl
(wt.%)
6.0
6.0
3.0
3.0
3.0
6.0
3.0
3.0
6.0
6.0
6.0
3.0

IFT
(mN/m)
0.016
0.011
0.050
0.024
0.011
ND
ND
ND
ND
0.018
0.034
0.048

IPA: iso-propanol. ND: Not determined.

10.00

10.00
Alfoterra 13

Alfoterra 63

Alfoterra 15

Alfoterra 65

Alfoterra 18

Alfoterra 68

1.00

IFT (mN/m)

IFT (mN/m)

1.00

0.10

0.10

0.01
0.0

3.0

6.0

9.0

12.0

Salinity (wt% NaCl)

Figure 14. IFT Results of Alfoterra 13, 15 and 18 at Low


Conc. (0.2 wt.%) in Various Salinity (Oil
Phase: n-Octane, Room Temperature).

Table 3. IFT Results of Alfoterra Surfactants at


Crude Oil \ Brine Interface (T=45 C)

Alfoterra

IFT for
IFT for
Surfactant Conc. (wt%)
Crude #4
GOM Oil
23
0.2
0.94
ND
28
0.2
0.22
0.49
38
0.2
0.23
0.72
45
0.2
0.65
ND
48
0.2
0.31
ND
68
0.2
0.69
ND
Brine Composiiton NaCl 2.9 g/l CaCl2.2H2O 0.1 g/l

ND Not Determined

0.01
0.0

2.0

4.0

6.0

8.0

10.0

12.0

Salinity (wt.% NaCl)

Figure 15. IFT Results of Alfoterra 63, 65 and 68 at Low


Conc. (0.2 wt.%) in Various Salinity (Oil
Phase: n-Octane, Room Temperature).

Table 4. IFT Results of Alfoterra 28 in NaCl Brine


(Surfactant Conc. 0.20 wt.%, T=45 C)

n-Octane
GOM Crude Oil
NaCl (wt%) IFT(mN/m)
NaCl (wt%) IFT(mN/m)
1
0.14
1
0.11
3
0.007
3
0.003
4
0.22
4
0.078
Densities at 45 C
n-Octane 0.695 g/cc GOM Crude Oil 0.812 g/cc

SPE 95404

Oil Phase

Water Phase

Oil Phase

Figure 16. Atomistic Simulations of Water/Oil Interface Mediated by Branched


Alcohol Propylene Sulfate, Sodium Salt

Table 5. Density and Dimension of Whole System

PPO
number

Density

Dimensions of Cell

Cross Sectional
Area, Avg.

Approx.
IFT
(dyn/cm)

Avg.

Std. Dev.

a, Avg.

b, Avg.

c, Avg.

0.8977

0.0008

26.94

32.92

54.23

888.8648

30

0.8968

0.0015

20.70

25.88

96.02

535.7678

0.9027

0.0012

21.53

26.91

94.40

579.3339

0.9083

0.0011

26.99

33.74

65.29

910.5655

Table 6. Adsorption of Alfoterra Surfactants on Kaolinite Clay from 2 wt.% NaCl Aqueous Solution
(Solid to solution mass ratio: 1:20, Equilibrium at room temperature, 25 C, for 12 hours)
Surfactant

Alfoterra 13
Alfoterra 15
Alfoterra 18
Alfoterra 23
Alfoterra 25
Alfoterra 28
Alfoterra 33
Alfoterra 35
Alfoterra 38
Alfoterra 43
Alfoterra 45
Alfoterra 48
Alfoterra 53
Alfoterra 55
Alfoterra 58
Alfoterra 63
Alfoterra 65
Alfoterra 68

Ceq(wt%)

(mg/g)

Ceq(wt%)

(mg/g)

Ceq(wt%)

(mg/g)

0.45
0.45
0.48
0.42
0.48
0.50
0.39
0.36
0.44
0.39
0.47
0.50
0.38
0.43
0.44
0.39
0.45
0.47

8.9
9.2
3.0
14.5
3.2
0.4
20.8
26.1
11.9
20.0
12.0
6.2
22.5
13.8
10.3
20.0
9.1
6.2

0.96
0.93
0.97
0.89
0.96
1.01
0.80
0.76
0.89
0.84
0.94
0.91
0.78
0.86
0.90
0.77
0.90
0.93

6.7
14.3
6.1
20.6
7.9
-2.2
38.3
45.1
20.6
30.5
23.6
12.6
40.1
22.6
15.2
42.5
18.8
14.3

1.98
2.02
1.96
1.78
1.97
2.03
1.70
1.64
1.83
1.75
1.95
1.85
1.66
1.77
1.80
1.75
1.85
1.88

4.0
-3.4
7.2
42.5
5.4
-5.4
55.7
66.1
30.1
46.7
35.9
27.1
61.7
40.0
33.6
47.7
27.6
22.9

10

SPE 95404

60

50

40
Tertiary Oil Recovery(V%)

Tertiary Oil Recovery (V%)

50

40

30

20

10

30

20

10

0
0

0.5

1.5

2.5

0.5

1.5

2.5

Pore Volume(PV) Injected

Pore Volume(PV) Injected

Figure 17. Tertiary oil recovery by Alfoterra 38 (2%wt.,


0.25 PV) followed by polymer Alcoflood 1235
(350 ppm, 2 PV).

Figure 18.

Tertiary oil recovery by Alfoterra 23 (.2%wt.,


0.8 PV) followed by polymer Alcoflood 1235
(350 ppm, 1.2 PV).

You might also like