Sunteți pe pagina 1din 16

SPE 165545

Surfactant-Steam Process: An Innovative Enhanced Heavy Oil Recovery


Method for Thermal Applications
K. Zeidani, SPE, S. C. Gupta, SPE, Cenovus Energy Inc

Copyright 2013, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Heavy Oil Conference Canada held in Calgary, Alberta, Canada, 1113 June 2013.
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
Surfactant-steam process (SSP) is a novel and potentially cost-effective process that utilizes a small amount of surfactant coinjected with steam to enhance the oil recovery of steam assisted gravity drainage (SAGD) well pairs. The mechanism of this
process involves interfacial tension (IFT) reduction, reservoir rock wettability alteration, oil relative permeability
enhancement, and in-situ emulsification.
SSP is expected to result in oil rate acceleration, steam-to-oil ratio (SOR) reduction and enhanced ultimate oil recovery
factor. Analogous enhancement is expected if the SSP is combined with other steam-based or steam-solvent processes.
This paper provides an introduction to this concept and presents a unique protocol that has been developed for screening
surfactants for co-injection with steam in SAGD process. In particular, the paper presents a scientific approach to surfactant
selection for SSP applications, describes the conditions in which the surfactants needs to be deployed within the reservoir,
and also predicts the potential synergies if use of different classes of surfactants is made. Novel experimental design on
different aspects of surfactant-steam phase behavior indicates the optimum surfactant concentrations for field trial
applications. Lab testing of selected surfactants on typical Canadian oilsands sand packs shows an improved incremental oil
recovery factor (RF) in the range of 6 to 16% (for different tested surfactants) compared to a SAGD base case.
SSP simulations were conducted for one of the surfactants that were tested in the lab. The simulation results indicate that this
particular surfactant on average accelerated the oil rate by 15% in the first 30 months of SSP operation, increased the ultimate
oil RF by 10%, and reduced the cumulative steam-to-oil ratio (CSOR) by almost 11% relative to a SAGD base case. In
addition, a sensitivity analysis was conducted to investigate the effect of surfactant concentration co-injected with steam. The
simulation results suggest that there is an optimum concentration for a given surfactant that needs to be explored through lab
investigations and field trials.
It is evident that once the SSP is successfully developed, the use of surfactant promises to improve environmental
performance and project economics of the in-situ oilsands recovery.
Introduction
Surfactants are compounds that lower interfacial tension (IFT) between two liquids (or liquid-solid), allowing easier
spreading. They may act as detergents, wetting agents, emulsifiers, foaming agents and dispersant. The use of surfactant in
oil industry and in particular in conventional resources is widely accepted as an effective enhanced oil recovery method. The
application of surfactant in Western Canadian oil deposits dates back to early 1900. When it comes to the use of surfactants
in oil sand deposits it has been utilized in surface mining for several decades. Recently, however, quite a few field
applications have been reported (Suncor energy 2011; Connacher Oil and Gas 2011), each aiming to co-inject chemicals with
steam in SAGD process. The physical properties of these chemicals are often not revealed in public domain. Often no or very
little information is provided on the criteria that the selection of these chemicals was based on. In few instances it appeared
that a chemical was selected based on a single feature, not necessarily ensuring that other compatibilities with reservoir and
operating conditions are in place. In light of this increased activity in the area of chemical co-injection with steam in SAGD,
the authors believe that the discussion on proper selection criteria for these chemicals (surfactants) is timely. This paper aims
to identify some of the critical elements of the surfactant-steam process (SSP) and to suggest a road map on the methodology
that could be followed at the lab and with simulations in order to asses the suitability of the selected surfactants for SSP.

SPE 165545

Surfactants Suitability for SSP


Selection of a surfactant for thermal application is a complex process mostly due to the nature of the process and the physico
chemical aspects of the reservoir rocks. This mainly relates to the condition in which the SSP is being carried out. In-situ
SAGD operations involve steam injection at elevated temperature of up to 320C. Extended exposure of surfactant to sever
temperate may degrade the surfactants if the right conditions are not met. The nature of reservoir rock chemistry and the
presence of many cations and anions are additional factors that may affect the SSP performance. In general terms a surfactant
can be considered as a suitable candidate for SSP application if it has the following characteristics: reduces IFT significantly;
vaporizes at downhole SAGD operating conditions; thermally stable at high temperatures; retain its properties and remain
effective at high temperatures; preferentially result in oil-in-water emulsion; enhances the reservoir wettability to water, and
be compatible with formation water. Surfactant availability and pricing obviously plays an essential role in the selection
process. It is also important that if a surfactant results in production of stable emulsion at downhole reservoir conditions then
it should be easy to break it down on surface. In the following sections each of above characteristics will be evaluated
separately and a detailed lab procedure will be given to measure certain aspect of the selected surfactants.
Surfactants Selection Criteria
There are varieties of surfactants available in the market that at the first glance may seem to be suitable for SSP application.
However, closer examination may reveals that most of these surfactants either can not be vaporized at SSP downhole
operating conditions or they may undergo thermal degradation. Another limiting factor is that some of these surfactants may
results in producing water-in-oil (W/O) emulsion which can be detrimental for the SSP recovery enhancement performance
owing to the fact that a W/O emulsion has a higher viscosity than the oil itself at a given operating conditions. It appears that
before investigating the type of surfactant that may work well for the SSP application it is more relevant to seek the type of
functionality that is expected from different type of surfactants and the estimated gain in process efficiency associated with
each class of surfactants. Surfactants can be classified based on the charge of their functional groups at their head. Anionic
surfactants are negatively charged while the cationic are the opposite (Rosen 2004). Zwitterionic surfactants have both
cationic and anionic centers attached to the same molecule. Nonionic surfactants do not have charge in their predominant
working range of pH. The first three class of surfactants appeared to be not suitable for SSP application either because it is
hard to vaporize them or because they may alter the reservoir rock wettability toward oil-wetness. However, the non-ionic
surfactants seem to be a good candidate for SSP application. They interact strongly with high energy structures such as
interfacial regions, phase boundaries and surfaces and can cause dramatic changes in crucial properties of molecular mixtures
including solubility, dispersion uniformity, viscosity, miscibility, and phase equilibrium.
Non-ionic Surfactants
Non-ionic surfactants are used extensively in the chemical industry in such areas as detergents, health and personal care,
coatings and polymers. The non-ionic surfactants include many subgroups but only phenol ethoxylates, alcohol ethoxylates,
and tertiary acetylenic diol including a tertiary acetylenic diol ethoxylates are covered in this paper. Zeidani and Gupta (2011
and 2012) provide an extensive list of other non-ionic surfactants that might be suitable for SSP application. The phenol
ethoxylate may have the formula:

CH2CH2O

H
m ,
wherein R3 is hydrogen, or a linear or branched alkyl group, and m is greater than 1, such as a linear or branched alkyl group
having more than 2 carbon atoms.
Alcohol ethoxylates are the most important family of non-ionic surfactants. Alcohol ethoxylates are manufactured by reacting
an alcohol with ethylene oxide (EO) in the presence of a catalyst. The EO chain, frequently called the polyoxyethylene chain,
imparts water solubility to the alcohol ethoxylate. They may be a primary, secondary, or tertiary alcohol ethoxylate. The
alcohol ethoxylate may have the formula
R3

H
m ,
wherein R1 is a linear or branched alkyl group having more than 5 carbon atoms, and m is greater than 1.

R1

CH2CH2O

The tertiary acetylenic diol may have the formula:


CH3
HC
R4

OR5

OR5
(CH2)p C
CH3

C
CH3

CH3
(CH2)p

CH

R4 ,
wherein R4 is hydrogen or methyl; R5 is hydrogen or hydroxyethyl; and p is 13 when R5 is hydroxyethyl, or less than 3 when
R5 is hydrogen. The tertiary acetylenic diol may also be a tertiary acetylenic diol ethoxylate having the formula

SPE 165545

CH3
HC
R6

O
(CH2)q

CH2CH2O H CH CH O H
2
2
zO
y CH3

C
CH3

(CH2)q

CH

CH3

R6 ,
wherein R6 is hydrogen, or a linear or branched alkyl group, q is greater than 1, and at least one of y and z is greater than or
equal to 1.
Relevant to non-ionic surfactant there are two terms that needs to be defined, HLB and CMC. The HLB (HydrophileLipophile Balance) is an empirical expression for the relationship of the hydrophilic ("water-loving") and hydrophobic
("water-hating") groups of a surfactant. The relationship is defined on scale of 1 to 20. The higher the HLB value, the more
water-soluble the surfactant is. The HLB system is particularly useful to identify surfactants for oil and water emulsification.
Water-in-oil emulsions (W/O) require low HLB surfactants. Oil-in-water (O/W) emulsions often require higher HLB
surfactants.
The critical micelle concentration (CMC) is defined as the concentration of surfactants above which micelles form and all
additional surfactants added to the system go to micelles. The CMC is an important characteristic of a surfactant. Before
reaching the CMC, the surface tension changes strongly with the concentration of the surfactant. After reaching the CMC, the
surface tension remains relatively constant or changes with a lower slope. The value of the CMC for a given dispersant in a
given medium depends on temperature, pressure, and (sometimes strongly) on the presence and concentration of other
surface active substances. Micelles only form above critical micelle temperature. Surfactants with lower CMC provide
greater benefits for the project economics since lower dosage of surfactants is required. Going forward the content of this
paper will be focused mostly on the use of non-ionic surfactants for SSP applications.
SSP Related Laboratory Measurements
Experiments were conducted to determine the suitability of a surfactant for the SSP application. Examples of the surfactants
tested include phenol ethoxylates such as TRITON X-100 (TX-100), alcohol ethoxylates such as NOVELFROTH 190
(E-190), NOVELFROTH 234 (E-234), TERGITOL 15-S-9 (T-15-S-9), CARBOWETTM 76 (C-76), tertiary acetylenic
diols such as SURFYNOLTM 82 (S-82) and SURFYNOLTM 104PA (S-104PA). Certain properties of selected surfactants are
shown in Table 1. Within the context of this paper all references to surfactant concentrations (in ppm) refer to volume
concentrations or ratios of the surfactant to steam on a cold water equivalent basis, as measured at room temperature, which
was 22C unless otherwise specified.
Thermal Stability of Surfactants. Thermal stability of neat TX-100 (i.e., no detectable amount of water), as measured by its
Molecular Weight (MW), was studied under various combinations of temperature and pressure. The MW was measured
using freezing point depression. Initial measurement of the MW of TX-100 was found to be about 764.1 g/mol. TX-100 was
subsequently placed in an anaerobic environment at about 250C for 48 hours and the MW post heating was about 749.5
g/mol, indicating that TX-100 was thermally stable in an anaerobic environment at 250C for at least 48 hours. The thermal
stabilities of TX-100, E-190 and E-234 (all neat) in an anaerobic environment at other temperatures up to about 325C over
24 hours were similarly measured and the results are shown in Fig. 1.
These results suggest that in an anaerobic environment at 325C, all three surfactants could undergo partial thermal
decomposition. Given the degree of decomposition at this temperature, the surfactants would not likely be suitable for
extremely high temperature operations. It is worth mentioning that the tested surfactants had some impurities due to their
manufacturing grade and also the testing environment lacked the presence of steam which could have prevented surfactants
dehydration at higher temperatures. Accordingly, surfactants such as TX-100, E-190 and E-234 could perhaps be used with
an acceptable degree of decomposition at a temperature higher than 250C tested in this program. Suitability of other
surfactants from thermal stability viewpoint may be readily determined at various temperatures using the process described
herein.
Interfacial Tension (IFT) Measurements. IFT of oil-water mixture with various selected surfactants at different
concentrations was measured using standard methods. Specifically, a calibration curve of surfactant IFT at different
surfactant concentrations was obtained by measuring the IFT of the surfactant in a mixture of a refined non-polar mineral oil
of constant properties and distilled water. The calibration curve afforded determination of an effective amount of the
surfactant in such an oil-water mixture based on a measured surfactant IFT at room temperature. Fig. 2 illustrates IFT
measurements for TX-100, E-190 and E-234 surfactants in the oil-water mixture (mineral oil). IFT reducing properties of
TX-100, E-190, E-234, T-15-S-9, S-82, S-104PA and C-76 for Athabasca bitumen and water mixtures were also evaluated.
The experiments were conducted at about 60C and the IFT measurements are shown in Fig. 3. The measurements for E-190
and E-234 show an abnormal behavior up to a concentration of 100 ppm. This is possibly an attribute of some contamination
or procedural error in the experiment. Except this abnormality, it can be seen the E-190 and E-234 appeared to have relatively

SPE 165545

comparable IFT reduction effects. The results also indicate that the rest of surfactants were effective at reducing oil-water IFT
at the conditions studied, with oil-water IFT values at high surfactant concentrations being comparable to the relatively low
values observed with E-190 and E-234. TX-100, T-15-S-9, S-82, S-104PA and C-76 appeared to more effectively reduce IFT
relative to E-190 and E-234, especially at low surfactant concentrations in the range of 10500 ppm. Also, the hydrocarbon
samples used for testing T-15-S-9, TX-100, S-82, S-104PA and C-76 appeared to have a different baseline IFT (about 2829
mN/m) compared to E-190 and E-234 (about 36 mN/m). This difference may relate to trace contamination or compositional
changes between two batches of hydrocarbons sampled at different times. The results also indicate that the change in IFT was
minimal for TX-100, T-15-S-9, and C-76 beyond the 300 to 500 ppm concentration which this range can be viewed as their
approximate CMC value. Other surfactants (i.e. E-190, E-234, S-82, and S-104 PA) seemed to have a CMC value in the
vicinity of 1000 ppm concentration.
Vapor Pressures of Surfactants. Closed system vapor pressure (absolute pressure in units of kPaa) of TX-100, E-190, E234, T-15-S-9, S-82, S-104PA and C-76 (all neat) at various temperatures ranging from about 20C to about 325C was
measured using a reactor cylinder placed inside an oven .The results are summarized in Fig. 4 and compared with the steam
vapor pressures. The results in Fig. 4 indicate that S-82 and S-104PA has a very similar vapor pressure to steam. E-190, E234, TX-100, C-76 and T-15-S-9 had relatively low vapor pressures relative to steam up to about 275C but their vapor
pressure increased sharply beyond this temperature. Sharp rise of the vapor pressure beyond 275C can also happen on
account of some thermal degradation but this was not ascertained. Normally, a pour pressure curve similar to the dashed line
in Fig. 4 is expected for E-190. Similar trends are expected for other surfactants beyond the 275C. Since the vapor pressure
of the steam is known then the fraction of surfactant vapor in the gaseous phase at a given temperature can be determined for
given surfactant dissolved concentrations.
Surfactants Vaporization and Effectiveness at Elevated Temperatures. The vapor pressure data presented in earlier
section provides the knowledge to estimate the extent of vaporization for a given surfactant at a given operating temperature.
However, it does not tell the reader if a surfactant was thermally degraded during the vaporization process. Also, it lacks the
knowledge on whether or not the vaporized surfactant retained its properties and the vapor fraction still capable or reducing
the IFT once it condenses to a liquid phase. For this purpose a new set of experiments were designed to investigate the level
of vaporization and the effectiveness of those surfactants that had relatively low vapor pressure. Fig. 5 shows the schematic
of the experimental setup. Distilled water was pumped into a test cylinder, which was placed in an oven heated to 275C.
Surfactant E-190 of known concentration (2000 ppm) was also injected into the test cylinder at a pressure above the vapor
pressure of the distilled water at a specified temperature. The pressure in the test cylinder was then dropped to the water
saturation pressure. A mixture of steam and surfactant was captured and was subsequently cooled to room temperature at
about 20C in a condenser. Samples were collected at defined time increments and subjected to the IFT measurements as
described in above (i.e IFT between refined oil and the condensate of steam plus surfactant mixture). The objective was to
measure the reduction in IFT from the baseline, and hence, the effective concentration of the surfactant that was vaporized in
the steam phase (i.e., the level of vaporization). E-190 study was conducted at the following conditions:
Water Injection Rate = 100 mL/h for the entire test run
Surfactant Injection Rate = 0.2 mL/h for the entire test run
Injection Pump Pressure = 5500 kPag
Mixing Vessel Pressure = 5494 kPag
Temperature of Mixing Vessel = 275C
Results for vaporization of E-190 in the steam phase (at about 275C) are shown in Table 2. The results suggest measured
IFT between the oil sample and the condensed steam plus surfactant of about 16.4 mN/m. This equates to an effective E-190
concentration of about 2000 ppm. This indicates that almost 100% of E-190 was vaporized into the steam phase at 275C and
the vaporized surfactant retained its properties to reduce IFT. Similar measurements were performed for few other surfactants
but due to some experimental issues they need to be repeated and they are not reported here.
Surfactants Static Adsorption Testing. Static adsorption tests were conducted to determine the extent of surfactant losses
due to adsorption to sand grains at high temperature and high pressure. The surfactants studied included T-15-S-9, E-190, E234, S-82, S-104PA and C-76. In a typical experiment, two cylinders were each charged with about 300 grams of cleaned
sand from homogenized oil sands core sample from a McMurray formation. Each cylinder was evacuated for about 2 hours
to remove gas from the system. The cylinders were then placed in a high temperature oven and shaken vigorously once every
8 hours. The cylinders were then heated to about 225C. About 300 mL of distilled water with a surfactant concentration of
about 2000 mg/L was injected into each cylinder. The pressure in each cylinder was expected to be close to the steam
pressure at 225C. Immediately after injection of water and the surfactant, the cylinders were rotated to mix water and the
surfactant with the oil sands (e.g., about every 6 hours for a 24 hours period). The cylinders were cooled to room temperature
and free water was removed from the cylinders without removing any sands. The IFT was measured between the removed
water and refined mineral oil in accordance with above IFT measurement techniques. The surfactant content in the removed

SPE 165545

water after high temperature adsorption was then determined based on these measurements. The results are shown in Fig. 6.
The results indicate virtually zero adsorption losses with T-15-S-9 and S-82, indicating that these surfactants did not change
significantly in their IFT reduction property in the refined oil-water system and thus T-15-S-9 and S-82 had a minimal
adsorption to the cleaned sand grains. This makes them economically more suitable surfactants choices for SSP application.
E-234, and to a slightly lesser extent C-76 and E-190, had relatively significant adsorption losses after the high temperature
static testing, indicating that significantly higher concentrations of these surfactants would be required in a field application
to generate a low IFT condition.
Coreflood Tests. Five coreflood tests on identical homogenized cleaned oil sand packs of 3.81 cm length were conducted
with sample solutions containing different sample surfactants in water (2000 mg/L). In a typical experiment, a homogenized
oil sand pack was prepared with about 10% initial water saturation and 90% bitumen saturation. The pack was then flooded
with cleaned (water free) oil at both 80C and 225C to evaluate the initial permeability to bitumen, which was found to be
5400 mD (milli-Darcy) at 80C, and 4100 mD at 225C (the second value is slightly lower due to connate water thermal
expansion effects and thermal grain expansion effects). Representative oil sand pack and test parameters are listed below:
Sand Status = Cleaned/Homogenized
Oil Source = Typical Athabasca Bitumen
Pack Length = 40 cm
Pack Diameter = 3.81 cm
Pack Flow Area = 11.4 cm2
Pack Porosity = 0.38 frac
Pack Pore Volume = 173.28 cm3
Test Temperature = 225C
Test Pore Pressure = 3450 kPag
Test Overburden Pressure = 7000 kPag
The first test conducted was a baseline run with no surfactant added, and the second test used 2000 ppm T-15-S-9 in the fresh
water phase. The procedure for T-15-S-9 testing was repeated using each of 2000 ppm E-190, S-82, and C-76. Comparative
results from the tests for percent recovery of original oil in place (OOIP) as a function of cumulative volume of water/steam
injected are summarized in Fig. 7. It was observed that all tested surfactants (T-15-S-9, S-82, E-190 and C-76) improved oil
recovery rate. Additionally, these results suggest that T-15-S-9 appeared to have a significant effect on both the rate of oil
recovery and the reduction of final residual oil saturation. Test 2 (T-15-S-9) recovered about 10% additional OOIP on pure
water flood at 225C in contrast to the baseline test, and almost 16% overall incremental percent recovery of OOIP when the
steam flood phase was taken into consideration. Thus, T-15-S-9 appeared to improve thermal recovery performance in both
hot water and steam displacement modes. Similarly, C-76, E-190, and S-82 resulted in incremental RF of about 6, 11, and
12%, respectively, relative to the steam-alone base case. It is concluded that the enhanced performance was probably due to
reduced residual oil saturation, enhanced relative permeability, IFT reduction, and perhaps wettability alteration in the
presence of tested surfactants.
Discussion of Experimental Results
The type of experimental tests reported in above sections provides a road map for selecting surfactants for SSP application.
There are several reasons for the need of such experiments. The first and foremost feature of any surfactants is its ability to
reduce the IFT to a great extent at the process conditions. Essentially all surfactants are considered to be IFT reducer. If a
surfactant is considered to be only IFT reducer or wetting agent then the knowledge of IFT data perhaps indicates whether or
not it meets the first criterion. However, if a surfactant is considered to be an emulsifier then additional tests are required to
evaluate its suitability. In particular the concept of HLB becomes very important. As such the selection of surfactant for SSP
application should exclude any emulsifiers with low HLB in order to avoid harming the process recovery performance. It also
should be noted that a surfactant with HLB of less 10 can be considered for SSP application as long as it is not an emulsifier
(i.e. wetting agent or only IFT reducer). Aside from the HLB concept the oil and surfactant molecular compatibility play a
significant role in determining the main functionality of a surfactant in an oil-water mixture. The extent of IFT reduction
indicates that the composition of the oil can also play a significant role. Looking at the IFT results of E-190 and E-234 with
refined oil shows that the scale of IFT reduction was not as extensive as the IFT reduction with bitumen. It seems that these
two surfactants may have reacted with the bitumen (acids, resins, and ashphaltene particles present in the bitumen) and
created additional In-Situ surfactants. It is postulated that a particular surfactant may be chosen for treating hydrocarbons
from a particular source based on molecular compatibility in terms of carbon chain length and the composition of the
hydrocarbon source. For example, the concentration of organic acids present in the hydrocarbon source may affect the
performance of certain surfactants. Therefore, in order to determine the suitability of a surfactant for treating hydrocarbons
from a particular source, standard laboratory tests may be performed as described in above.
Regarding evaluating the thermal stability of a surfactant for SSP application the MW measurement is only one of the
suggested methods that can be utilized for this purpose. There exist other methods that might cost more but perhaps provide

SPE 165545

more reliable results. For example the MW measurement conducted in this study shows that selected alcohol ethoxylates (E190 and E-234) and phenol ethoxylates (TX-100) may undergo thermal decomposition to a certain degree at temperature
325C. Similar but perhaps more accurate results were obtained when Mitsuda et al. (1989) examined the TX-100 thermal
stability with differential scanning calorimetry (DSC) and IR spectrophotometry below 350C. They reported that in the
absence of oxygen (carrier gas like nitrogen or argon) TX-100 evaporated without decomposition. Similarly Evetts et al.
(1995) studied the thermal stability of neat alcohol ethoxylates and alcohol ethoxylates mixed with water (50:50 vol/vol) at
high temperatures. They utilized extensively aerated conditions that were modeled by thermogravimetry with mass
spectrometric detection and differential scanning calorimetry with vapor phase sampling and analysis. They concluded that
C14-15 alcohol with an average 7 moles of EO per mole of alcohol is thermally stable to 300C in inert atmosphere.
Interestingly, they reported that alcohol ethoxylates are significantly more stable under heat when diluted with steam.
The static adsorption tests are essential to evaluate the extent of the surfactants adsorption on porous media. If a surfactant
adsorb extensively on a porous media it means that greater concentration is required in the field in order to make sure the
surfactant can reach out the periphery of the steam chamber. This may harm project economics due to higher surfactant cost.
On the other hand surfactant adsorption can be viewed as positive phenomena since the greater the adsorption is the greater
of the amount oil that can be unlocked from the sand grains (i.e. lower residual oil saturation). Also, in the case of
surfactants with emulsifying characteristics it is preferred to have the surfactant left inside the reservoir in order to minimize
the risk to a SAGD facility due to producing the surfactant to surface and perhaps producing tough emulsions that can not be
broken easily.
Finally the coreflood tests were conducted at 2000 ppm of surfactant concentration in order to be able compare the results. In
reality each surfactant has its own optimum concentration which the test should be carried out.
SSP Simulation Study
The knowledge of vapor pressure of a surfactant is an important consideration to be evaluated for SSP application. Such
information allows the assessment of surfactant vaporization and the extent in which it can be presented in the gaseous phase
at a given operating condition. The vapor pressure can be utilized in a thermal simulator to predict the concentration of
surfactant in different region of the steam chamber and its partition within the vapor and liquid phases. Alternately, one can
utilize the vapor pressure data and predict the phase behavior of a surfactant using the Raoults law which gives:
sat
sat
Ptotal x steam Psteam
x surfac tan t Psurfac
tan t

y surfac tan t
x surfac tan t

sat
Psurfac
tan t

Ptotal

where Ptotal is the mixture's total vapor pressure and x is the mole fraction of steam and surfactant in the liquid mixture; y is
the mole fraction of a component (steam or surfactant) in the vapor phase.
Additionally, the vapor density of a surfactant can be approximated by using the ideal gas law assuming an ideal mixture of
gases exists within the SSP steam chamber.
Proper simulation of SSP process requires the knowledge of PVT data as discussed above along with other physical
properties of each surfactant on a broad scale of operating pressures and temperatures. Such information for the most part is
not available through the surfactants manufacturer. Also, measurement of such properties in the lab can be time consuming
and costly. Using the in-house built simulator, the authors aimed to overcome some of these difficulties through taking a
simplified approach yet somewhat representative of the actual reservoir process. The approach includes combining and
incorporating all enhancement functionality of the surfactant into the relative permeability function. As widely accepted in
the literature the relative permeability functional dependence can be approximated by:

k rw f ( S w )

krg f ( S g )

kro f ( S w , S g )

The kro function is rarely known and even when it is, the form is not convenient for use in numerical simulators. Practical
approaches are based on the estimation of Three-Phase relative permeability from two set of Two-Phase data. As such the
relative permeability in oil-water and oil-gas systems are modified as follow:

k row f ( S w )

k rog f ( S g )

Above methodology was modified further for the SSP system in which the relative permeability is modified based on the
concentration of surfactant (Csurf) presented in the system.

krow f ( S w , Csurf )

krog f ( S g , Csurf )

SPE 165545

The relative permeability curves were developed based on the coreflood experimental results and were utilized in this
simulation study. As explained in previous section the sand packs steam (plus surfactant) flooding was performed at two
stages: first the sand packs were flooded with hot water and then it was exposed to saturated steam to represent the actual
process the way it happens downhole at reservoir condition. As it is expected the oil RF was improved from hot water
flooding to steam flooding and this improvement was greater with the use of surfactants, each yielded relatively higher RF
relative to the base case. Based on these data we modified the krow endpoint to match the residual (immobile) oil saturation at
the end of hot water flooding stage. The incremental reduction in residual saturation during the steam flooding stage was
matched through modifying the end point of krog curve. It was assumed that the krw and krg of steam plus surfactant case (SSP
cases) will remain similar to the steam-alone case (SAGD case). Such assumption seems to be reasonable for the available
experimental data but certainly it should be investigated through independent study and it provides room to improve the
assumptions and consequently the simulation results. Figures 8 and 9 show the relative permeability curves that were
developed for the above-mentioned five experiments (i.e. steam-only, E-190, S-82, C-76, and T-15-S-9 tests). Only two out
of the five experiments were simulated to be served as example for SSP simulation. The SAGD (steam-only) base case and
the SSP case (using 2000 ppm Surfynol, S-82) were simulated. Also, a sensitivity analysis of S-82 surfactant concentration
was performed aiming to identify the optimum concentration that needs to be tested and verified at the field scale. For this
purpose three SSP simulations having 50, 500 and 2000 ppm of S-82 were conducted. An additional simplifying assumption
was made that the residual (immobile) oil saturations for krog and krow curves are directly proportional to the surfactant
concentration in water, with these saturations falling on the base case values in absence of surfactant and their maximum
reduced values at the CMC of the surfactant. This assumption should be verified in future studies aiming to improve the
modeling accuracy of the SSP process.
Two-Dimensional models were used in the simulation studies. The simulation studies were based on dead-oil modeling
which had no solution gas in order to avoid any potential complexity with unknowns associated with the non-condensable gas
flow in steam chamber. Following reservoir conditions and properties were assumed:
Reservoir Dimensions:
Model Grid Block Dimension:
Original Oil in Place:
Horizontal Permeability:
Vertical Permeability:
Oil Gravity:
Initial reservoir Pressure:
Steam Injection Strategy:
Operation Mode:
Start of Surfactant Injection:
Vertical Well Spacing:

100m wide x 30m thick x 800m long


1m wide x 1m thick x 800m long
80% (by volume)
6 Darcy
2 Darcy
8.5 API
2600 kPaa
Controlled by downhole injection pressure
10 months gas lift operation mode (4600 kPaa) followed by ESP mode (2600 kPaa)
Right after circulation at 60 days
4m

The simulation results that are presented in Fig. 10 show the steam injection and oil production rates for all cases. It indicates
that surfactant S-82, at injected concentration of 2000 ppm, on average accelerated the oil rate by 15% for the first 30 months
of surfactant injection. Afterwards, the oil rate dropped gradually over time but remained at higher level relative to the SAGD
base case for the rest of SSP life. This figure also shows that lower concentrations of surfactant co-injected at 50 and 500
ppm did not improve the oil rate noticeably during gas lift (first 10 months) operation. As the chamber operating pressure
was dropped to 2600 kPaa (ESP operating mode) it appears that the 500 ppm case boosted the oil rate by almost 17% for the
next 4 years of operation. On the other hand, it seems that the 50 ppm case did not improve the oil rate in the first 30 months
of the operation but it increased the oil rate by almost 23% for the rest of SSP life.
Fig. 11 shows the CSOR for all cases. It shows that at 30 months of SSP operation the CSOR has dropped by 7.5 and 10%
for 500 and 2000 ppm cases, respectively, relative to the SAGD base case. The 50 ppm case had no impact on CSOR during
this period. At about 7 years of SSP operation the reduction in CSOR was about 11% for both 500 and 2000 ppm cases; and
at similar timing the 50 ppm case was effective to reduce the CSOR by almost 8% relative to the SAGD base case.
Finally, Fig. 12 compares the RF for all cases. This figure indicates an incremental oil RF of about 4 and 7.5% for the 500
and 2000 ppm cases, respectively, relative to the SAGD base case at the end of 3 years of SSP operation. It also indicates that
as the chamber growth continues and at the end of 7 years of operation the 50, 500 and 2000 ppm cases may produce
incremental oil RF at about 6, 8.4, and 9% (respectively) relative to the SAGD base case.
In general, it can be seen that in the case of S-82 surfactant higher concentrations (2000 ppm) are very effective in enhancing
the SSP performance during the entire life of operation. Low (50 ppm) and moderate (500 ppm) surfactant concentrations
seem to improve the process efficiency at the middle and late stages of the operation. As steam chamber hits the SAGD-top
boundary, which is normally followed by reduction in steam chamber operating pressure, the moderate surfactant
concentrations become more effective in improving the process efficiency. Lower surfactant concentrations appear to
improve the process efficiency at later stages of SSP operation. It appears that there exists an optimum surfactant

SPE 165545

concentration that needs to be investigated at the lab and through simulation for each surfactant. It is expected that if the
optimum concentration is identified it may improve the project economics significantly.
Conclusions
SSP appears to be an attractive process that produces improved results compared to SAGD. Based on preliminary
lab and simulation results it appears that the SSP may accelerate the oil rate, reduce cumulative steam-to-oil ratio
(CSOR), and enhance the ultimate oil recovery factor (RF).
A set of criteria have been developed for selecting and testing surfactants for SSP application.
Different classes of surfactants were studied and few surfactants were selected for lab testing. It appeared that most
of these selected surfactants are potentially suitable candidates for SSP application. It is evident that field trials and
economical evaluation are essentials for further development of this novel process.
SSP simulations were conducted based on reasonable assumptions and limited measured PVT data and other
physical properties that allowed further understanding of the SSP. It is believed that the current modeling approach
could serve as a basis for future enhancement of SSP simulation studies.
Acknowledgment
The authors are grateful to Cenovus Energy for permission to publish this paper. Also, they would like to acknowledge the
support from colleagues in both Christina Lake and Technology Development teams.
Nomenclature
C-76
CMC
CSOR
E-190
E-234
EO
ESP
g/mol
h
HLB
krg
krow
krog
krw
kPaa
kPag
mD
mg/L
mL/h
mN/m
MW
O/W
OOIP
Ptotal
ppm
PVT
RF
S-104PA
S-82
SAGD
SOR
SSP
T-15-S-9
TX-100
W/O
x
y

=
=
=
=
=
=
=
=
=
=
=

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

CARBOWETTM 76 surfactant
critical micelle concentration
cumulative steam-to-oil ratio
NOVELFROTH 190 surfactant
NOVELFROTH 234 surfactant
ethylene oxide
electrical submersible pumps
gram per mole
hour
hydrophile-lipophile balance
gas relative permeability
oil relative permeability in oil-water system
oil relative permeability in oil-gas system
water relative permeability
kilo Pascal absolute
kilo Pascal gauge
milli-Ddarcy
milli-gram per liter
milli-liter per hour
milli-Newton per meter
molecular weight
oil-in-waterOW emulsion
original oil in place
mixture's total vapor pressure
part per million
pressure-volume-temperature
oil recovery factor
SURFYNOLTM 104PA surfactant
SURFYNOLTM 82 surfactant
steam assisted gravity drainage
steam-to-oil ratio
surfactant-steam process
TERGITOL 15-S-9 surfactant
TRITON X-100 surfactant
water-in-oil emulsion
mole fraction of a component in the liquid phase
mole fraction of a component in the vapor phase

SPE 165545

References
1. Butler, R.M. 1994. Steam Assisted Gravity Drainage. Concept, Development, Performance, and Future. Journal of
Canadian. Petroleum Technology 33(2) 44-50.
2. Cenovus Energy Inc. 2012. Surfactant-Steam Process (SSP) pilot project application for experimental scheme
approval, Energy Resources Conservative Board application No. 1724747, Alberta, Canada April 11.
3. Connacher Oil and Gas Limited. 2011. Application to amend approval N0. 10587E to add surfactant to the steam
injected into well pairs 102W-04 & 102W-05 at Connachers Pod One SAGD facility, Energy Resources
Conservative Board application No. 1707322, Alberta, Canada November 15.
4. Engel, D.B., Alan, E., Goliaszewski, McDaniel, C.R. 2010. Separatory and emulsion breaking processes. US patent
No. 7,771,588B2 August 10.
5. Evetts, S., Cynthia K., Levin, M., and Stafford, M. 1995. High-Temperature stability of alcohol ethoxylates. JAOCS
Vol. 72, no. 7.
6. Handy, L.L., Amaefuel, J.O., Ziegler, V.M., and Ershaghi, I. 1979. Thermal stability of surfactants for reservoir
application. Paper SPE 7867, presented at SPE Fourth International Symposium on Oilfield and Chemistry, Huston,
USA January 22-24.
7. Hart, P.R., Stefan B. J., Srivastava, P., and Debord, J.D. 2009. Method for enhancing heavy hydrocarbon recovery.
US patent application No. 2009/0218099 September 3.
8. Mitsuda, K., Kimura H., and Murahashi, T.1989. Evaporation and decomposition of Triton X-100 under various
gases and temperatures, Journal of Materials Science 24 413419.
9. Ozum, B. 2010. Method for improving bitumen recovery from oil sands by production of surfactants from bitumen
asphaltenes. US patent application No. 2010/0147742A1 June 17.
10. Rosen, M.J. 2004. Surfactants and interfacial phenomena, third edition, Hoboken, New Jersey, USA John Wiley
and Sons.
11. Scamehorn, J.F., Sabatini, D.A., and Harwell, J.H. 2004. Surfactant, Part I: Fundamentals, In Encyclopedia of
Supramolecular Chemistry; Atwood, J., and Stead, J., eds., Marcel Dekker Inc., New York, 1458 1469.
12. Srivastava, P., Sadestsky, V., Debord, J., Stefan, B., and Orr, B. 2010. Development of a steam-additive technology
to enhance thermal recovery of heavy oil. Paper SPE 133465 presented at the SPE Annual Technical Conference
and Exhibition, Florence, Italy September 19-22.
13. Suncor Energy Inc. 2011. Application for Chemical (Alkali and/or surfactant) Pad 22 co-injection test, Suncor
Mackay River oil sands project, Energy Resources Conservative Board application No. 1690728, Alberta, Canada
June 16.
14. Zeidani, K., and Gupta, S.C. 2012. Hydrocarbon recovery from bituminous sands with injection of surfactant
vapour. Canadian patent application No. CA 2791492 September 28.
15. Zeidani, K., and Gupta, S.C. 2011. Hydrocarbon recovery from bituminous sands with injection of surfactant
vapour. US provisional patent application No. 61/541712 September 30.

Table 1. Properties of selected surfactants for lab testing


MW
(g/mol)

CMC
(ppm)

HLB

TRITONTM X-100 (TX-100)


TERGITOL 15-S-9 (T-15-S-9)
NOVELFROTH 190 (E-190)
NOVELFROTH 234 (E-234)
CARBOWETTM 76 (C-76)
SURFYNOLTM 82 (S-82)

624
596
228
291
169

189
52
-

13.6
13.3
11.0
12.7
7.5
5.0

SURFYNOLTM 104PA (S-104PA)

226

4.0

Surfactant

10

SPE 165545

Table 2. E-190 vaporization result at 275C

Sample #
Baseline (steam
injection phase)
SSP
(steam and 2000
ppm surfactant coinjection at steady
state condition)

Total
Time
(h)

Total H2O
Injected
(mL)

Total
Surfactant
Injected
(mL)

Avg. IFT
(H2O-Oil)
(mN/m)

Avg. IFT
(SampleOil)
(mN/m)

Concentration of
E-190
(ppm)

22.50

2251.10

0.00

36.90

36.80

0.00

119.21

11941.30

9.80

36.90

16.40

2000.00

900
800
Initial MW
Molecular Mass, g/mol

700
Post 250C MW
600

Post 325C MW

500
400
300
200
100
0
TX-100

E-190
Surfactant Type

Fig. 1. Surfactants MW measurement pre- and post-heating

E-234

SPE 165545

11

40
35

TX-100

IFT, mN/m

30

E-190

25

E-234

20
15
10
5
0
0

500

1000

1500

2000

2500

Surfactant Concentration in Water, ppm

Fig. 2. IFT of selected surfactants with refined oil

40

35

TX-100
E-190

30
E-234
T-15-S-9

IFT, mN/m

25

C-76
20

S-82
S-104PA

15

10

0
0.01

0.1

10

100

1000

Surfactant Concentration in Water, ppm

Fig. 3. IFT of selected surfactants with typical Athabasca bitumen

10000

100000

12

SPE 165545

100,000
T-15-S-9
S-82
S-104PA

10,000

Vapour Pressure, kPaa

C-76
E-234
1,000

TX-100
Steam
E-190

100

Expected E-190

10

1
0

50

100

150
200
Temperature, C

Fig. 4. Surfactants vapor pressure measurement.

E-190 Inj Pump

Oven at 275 deg C

Fig. 5. Schematic of surfactants vaporization experimental setup.

250

300

350

SPE 165545

13

100
90

Adsorption Losses, %

80
70
60
50
40
30
20
10
0
T-15-S-9

E-190

S-104PA

E-234

C-76

S-82

Fig. 6. Surfactant adsorption losses to a typical Athabasca reservoir sand grains.

100
90

Bitumen recovery Factor, %

85
80
70

81
80
75

69

60
50
T-15-S-9
40

S-82

30

E-190

20

C-76
Base case - no surfactant

10
0
1

10

100
Cumulative Injection, mL

1000

10000

Fig. 7. The results of five sand pack steam flood tests (one with steam only and the rest utilized steam
plus 2000 ppm of different surfactants).

14

SPE 165545

E-190
T-15-S-9
krg

1.0

1.0

0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0
0.0

0.2

0.4

0.6

0.8

Krg

Krog

C-76
S-82
Base case - no surfactant

1.0

Sg

Fig. 8. Relative permeability for oil - water+ surfactant system.

S-82
C-76
Krw

1.0

1.0

0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Sw

Fig. 9. Relative permeability for oil steam + surfactant system.

0.7

0.8

0.9

1.0

Krw

Krow

T-15-S-9
E-190
Base case - no surfactant

SPE 165545

15

1000
SAGD oil

900

SAGD steam
SSP oil (50 ppm)

800
Steam Rate, Oil Rate, m3/d

SSP steam (50 ppm)


700

SSP oil (500 ppm)


SSP steam (500 ppm)

600

SSP oil (2000 ppm)


SSP steam (2000 ppm)

500
400
300
200
100
0
0.0

1.0

2.0

3.0
4.0
Times, Years

5.0

6.0

7.0

Fig. 10. Simulation results, steam and oil daily rates for SAGD and SSP with different S-82 surfactant
concentrations.

3.00

Cumulative Steam-Oil Ratio (CSOR)

2.80

SAGD
SSP 50 ppm
SSP 500 ppm
SSP 2000 ppm

2.60

2.40

2.20

2.00

1.80

1.60
0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

Time, Years

Fig. 11. Simulation results, CSOR for SAGD and SSP with different S-82 surfactant concentrations.

16

SPE 165545

90%

80%

Bitumen Recovery (RF), %

70%

60%

50%

SAGD
SSP 50 ppm
SSP 500 ppm
SSP 2000 ppm

40%

30%

20%

10%

0%
0.0

1.0

2.0

3.0

4.0

5.0

6.0

Time, Years

Fig. 12. Simulation results, bitumen RF for SAGD and SSP with different S-82 concentrations.

7.0

S-ar putea să vă placă și