Você está na página 1de 10

SPE 143987

Emulsion Characteristics and Novel Demulsifiers for Treating Chemical


Induced Emulsions
Duy Nguyen, Nicholas Sadeghi, and Christopher Houston - Nalco Company

Copyright 2011, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Enhanced Oil Recovery Conference held in Kuala Lumpur, Malaysia, 1921 July 2011.
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
In this paper, the stability of oil-in-water emulsions produced from chemical enhanced oil recovery processes was investigated
over a wide range of parameters. These parameters are surfactant concentration, polymer concentration, mixing speed,
asphaltene concentration, salinity concentration, water cut, temperature, and alkaline concentration. Emulsion stability
decreased with an increase in temperature, salinity content, or water cut. Increasing surfactant concentration, polymer
concentration, or shear rate enhanced emulsion stability. One of the main contributions for the tight emulsion from alkaline
surfactant polymer (ASP) flood was the addition of alkaline. The surfactant, alkaline, and polymer decreased the size of oil
droplets, increased the surface charge of oil droplets, and increased the film elasticity, thereby making oil-water separation
difficult. Selected cationic surfactants (patents pending) proved much more effective than conventional nonionic resins and
polymeric cationic flocculants in separating oil-in-water emulsions. We also studied the effect of alkyl chain length (C8 C18) of benzyl and methyl quats on demulsifying efficiency and compared the performances of monoalkyl quat with dialkyl
quat. As the surfactant concentration in the brine decreased, the concentration of the cationic demulsifier required to separate
the emulsion decreased and the optimum chain length of the cationic demulsifier also changed. Particle video microscope and
focused beam reflectance measurement probes showed significant increase of the size of oil droplets and reduction in the
number of oil droplets in the presence of a cationic surfactant. This is in agreement of a decrease of the anionic charge on the
surface of the oil droplets and a reduction of the film elasticity in the cationic system. Application of this novel demulsifier
resulted in a much more effective oil/water separations process with the production of dry oil and clean water at a pilot ASP
flood that was experiencing very stable emulsions.
Introduction
Primary and secondary recovery techniques together are able to recover only about 35-50% of the oil in place. Since there is a
significant amount of oil remaining in the reservoir after the primary and secondary processes have been utilized, chemical
flooding (using surfactants, polymers and sometimes alkali) has been one of the technologies that can be used to recover up to
an additional 35%. It is estimated that several hundred chemical EOR field trials have been conducted over the past 50 years
with many occurring during the 1970s and 80s [1].
Chemical flooding today has been re-invigorated by the introduction of more cost-effective surfactants and polymers
coupled with improved reservoir modeling. Globally it is estimated that there are over 20 chemical EOR projects in field trial
or commercial stage in 2011. Additionally, the current (April, 2011) $100/bbl crude oil price environment continues to
stimulate interest in new projects by super majors, national oil companies and independents. Emulsions created by chemical
flooding have been extremely difficult to break due to the high concentration of alkali, surfactant and polymer tightly bound
with the oil and water. Traditional non-ionic demulsifiers have limited effect on the chemical emulsions. Some early 1980s
chemical floods were known to require several thousand parts per million (ppm) demulsifier in order to break the emulsion.
Many compounds have been proposed or used as demulsifiers. One type of widely used demulsifier capable of displacing
significant portions of the asphaltene layers in non-chemical EOR emulsions and promoting coalescence consists of
ethoxylated and/or propoxylated alkylphenol formaldehyde resins (APF) with molecular weights of a few thousand Daltons
[2]. As shown in Figure 1, addition of conventional demulsifiers of the type described above produced some separation, but it
was incomplete and the water content of the oil phase was unacceptable (11% BS&W). Adding alum or conventional cationic
polymeric flocculants to the emulsion was also unsuccessful at resolving the emulsion and caused viscous sediments at the

SPE 143987

bottom. This paper will discuss a new series of patent pending cationic demulsifiers which has been created that addresses the
unique nature of chemical EOR emulsions.

APF (a)
Figure 1:

Cationic flocculants (b)

Patent pending cationic demulsifier (c)

Conventional demulsifiers (a) and (b) vs. patent pending cationic demulsifer (c) in surfactant and polymer flood.

Experimental Methods
Materials: Pure cationic surfactants of the types, alkyltrimethylammonium halide (C6-C18) and
alkyldimethylbenzylammonium halide (C10-C18), were obtained from Sigma- Aldrich. The crude oils were from the USA
and their physical and chemical properties are listed in Table 1. The polymer (partially hydrolyzed polyacrylamide, HPAM)
from SNF Company has a molecular weight of 8x106 and 30% hydrolysis. Alcohol propoxylate sulfate and alkyl sulfonate
surfactants were obtained from Stepan Company. Brine solutions were prepared using various inorganic salts obtained from
Sigma-Aldrich. Diethylene glycol monobutyl ether (DGBE) and iso-butanol co-solvents were purchased from Aldrich
Chemical.
Emulsion preparation: The emulsion (100 ml) was produced in the lab by mixing the produced water (i.e., brine solution)
containing polymer and surfactants with the oil by shaking the six-ounce prescription bottle mechanically for 10 minutes using
a heavy-duty, two-speed Eberbach shaker. The demulsifier was added to the above emulsion and the bottle was again shaken
for another 3 minutes. The amounts (ppm) of demulsifiers added were based on the volume of the emulsion. The ratios of
produced water to oil used were varied from 1:1 to 9:1.
Particle Video Microscope (PVM): The PVM is from Mettler-Toledo Lasentec and consists of six-near IR lasers which
illuminate a small area in front of the probe face. The probe records high quality images even in dark and concentrated
suspensions or emulsions in real-time. Droplets between 2 microns and 1000 microns can be detected.
Interfacial Tension: The equilibrium interfacial tensions of brine and crude oil after phase separation were measured at 22oC,
using a University of Texas spinning drop tensiometer. Aliquots of oil and water were withdrawn with a 25-mL syringe
equipped with a 2-in. needle. Each aqueous phase was brought to temperature in the instrument as the continuous phase.
Then, 1.0 microliter drop of oil phase was introduced as the drop phase and the system was spun at variable speeds.
Demulsification tests: Demulsification tests were conducted in graduated six ounce prescription bottles to allow for rapid
water drop readings. All bottles used 100 ml of emulsion. After pouring the emulsion followed by chemical addition, the
bottles were allowed to reach the separator desired temperature via a water bath. Upon reaching the desired temperature, the
samples were shaken via a mechanical shaker and then returned to the water bath. Water drop readings were recorded in
milliliters as a function of time. Water drop values were also used to gauge emulsion stability, where a faster water drop
indicated lower emulsion stability. Water samples were taken from the bottoms of the bottles using a syringe and the oil
concentrations in water were deduced from the turbidity measurements. Following the oil drop readings, the resolved or
partially resolved oil from each bottle was analyzed for water content. Using a syringe with a needle, a small portion of the oil
(about 6 ml) was withdrawn. The tip of the syringe was set to 10-15 ml above the theoretical oil-water interface as determined
by the slug grindout value. This aliquot of oil was added to a graduated API centrifuge tube containing an equal volume of an
aromatic solvent and the contents were shaken by hand. The centrifuge tubes were then centrifuged on high speed for three
minutes. Following centrifugation, the percent residual emulsion, typically referred to as basic sediment (BS), and the percent
water (W) were noted for each bottle. After recording BS&W values, alkyl sulfonate surfactant (a chemical known to resolve
the remaining emulsion) was added to the centrifuge tube. Such chemicals are generally called slugging or knockout
chemicals and are typically low molecular weight sulfonate-based materials. After slugging, the tube was again shaken and
centrifuged as previously described. The BS was therefore completely eliminated and only water remained in the bottom part
of the tube. The slug grindout number is reported as a percentage. Smaller values of BS&W and slug indicate drier oil. The
time (in minutes) elapsed for the total volume of water to separate is taken as a measurement of the emulsion stability.

SPE 143987

Determination of zeta potential of oil droplets: Zeta potential measurements were conducted on a Horiba DT-1200,
Acoustic and Electroacoustic Spectrometer, by means of Colloid Vibration Current (CVI). Basically, an ultrasonic wave is
introduced and disturbs the double layer. The displacement of the ionic cloud creates a dipole moment. The sum of these
dipole moments over many particles creates an electrical field which causes the CVI to flow. The instrument measures the
CVI from which the zeta potential is determined from an empirical equation. Details about the technique can be found
elsewhere [3]. One of the main advantages of electroacoustics over the traditional electrokinetic methods is the elimination of
dilution as the instrument can measure up to 50% volume of dispersed liquids or solids. When diluted, droplet-droplet
interactions become less of importance which is not true for the systems we are dealing with. An emulsion at 70% water cut
was made as described previously and sat for two hours without agitation. The emulsion was then stirred gently with a stirring
bar and the zeta potential was measured.
Results and Discussions:
Effect of Formulation on Emulsion Stability: Emulsion stability depends both on the nature of the surfactant and the oil. In
1949, the hydrophilic-lipophilic balance (HLB) concept was introduced by Griffin [4] to describe the relative affinity of a
surfactant for water and oil. The simplicity of the HLB concept was its main advantage but it had several limitations such as it
did not take into account the effects of temperature, salinity, co-surfactant concentration, and the nature of the hydrophilic
group, and also was inaccurate when comparing different surfactant families.
In 1954, Winsor [5] introduced the R ratio that describes the molecular interaction energies between the surfactant
adsorbed at the interface and the water and oil phases surrounding it. Several authors [6,7] have used the Winsors R
parameter and showed that a maximum in demulsification performance corresponds to a physicochemical condition for which
R = 1. At such a condition, Salager et al. [8,9] also observed minimum stability occurs with emulsions made with oilsurfactant-water systems. However, the R ratio remains qualitative and is limited because the molecular interaction energies
cannot be determined experimentally.
In 1964, Shinoda [10] developed a method based on the determination of the phase inversion temperature (PIT) of a
surfactant-oil-water mixture heated under stirring. This method is based on the cloud point phenomenon associated with
nonionic surfactants and is more reliable than the above methods since it takes into account various variables such as
surfactant, oil, salinity, co-surfactant which affect the PIT. However, the PIT method is applicable only to nonionic surfactants
because ionic surfactants are much less sensitive to temperature.
In the 1970s, the drive of enhanced oil recovery stimulated researchers to improve upon Shinodas phase inversion
temperature and quantify Winsors R parameter. Salager et al. [11] introduced the hydrophilic lipophilic deviation (HLD)
concept, a dimensionless numerical expression which is expressed as a linear relationship including all formulation variables
such as the nature of the surfactant, oil nature, salinity nature and concentration, and the presence of alcohol cosurfactant, as
well as temperature. The HLD has been used as a quantitative design tool in the crude oil dehydration process to determine
regions where macroemulsions are likely to break easily [12]. The emulsion stability is very low when the surfactant has
exactly the same affinity for both oil and water phases. Such formulation for the emulsion quickest breaking has been called
optimum formulation and has been observed for a variety of surfactant-water-oil systems [6,12,13,14]. HLD is a
dimensionless number expressed, for anionic surfactants, by the following relation [15]:
HLD = + ln S kACN -t (T-25C) +aA
Eq (1)
where ACN is the carbon number of the n-alkane which is replaced by equivalent alkane number (EACN) for non-alkane oils,
A and S are alcohol and salt concentrations, is a characteristic parameter of the anionic surfactant which increases linearly
with the length of the lipophilic tail. k, t, and a are positive constants which depend on the surfactant head group and
electrolyte. As can be seen, the HLD value takes into account not only the HLB of the surfactant, but also the temperature, the
nature and the concentration of cosurfactant (e.g., alcohol), the electrolyte and the nature of the oil phase. When the HLD
value is zero or very close to zero, an optimum formulation is attained, indicating that the emulsion is easy to break or very
unstable. When HLD is positive or negative, the emulsion type is O/W or W/O, respectively. To obtain an optimal
formulation with a system which already contains an anionic surfactant(s) in which is negative (chemical EOR where HLD
<0), a demulsifier must be added to ensure the characteristic parameter of the mixture of the anionic surfactant and demulsifier
results in HLD=0 or close to zero, corresponding to minimum emulsion stability. This can be done by adding a cationic
surfactant to form ion pairs with an anionic surfactant, resulting in a positive value of [16]. These authors showed that the
ion pairs increased the detergency efficiency significantly. Interestingly, by applying the HLD concept, we have discovered
that selected cationic surfactants (patents pending) accelerated separation of oil-in-water emulsions representative of produced
emulsions expected during chemical EOR processes and yielded oil and water phases with greatly improved quality compared
to the initial emulsion. Furthermore, the concentration of demulsifier when combined with the anionic surfactant must
produce HLD = 0. This is in agreement with the observation that dehydration conditions are specific and that a small change
in the demulsifier concentration (above or below the optimum), or in temperature, may cause the emulsion shifting from being
unstable to stable. Also, because this optimum formulation corresponds to the proper mixture of demulsifier and anionic
surfactant at the interface, it depends on the anionic surfactant concentration, as will be discussed in this paper.

SPE 143987

Characterization of oil: As shown in Table 1, the physical and chemical properties of the oils were measured and
characterized. The API gravity and viscosity were measured at ambient temperature. SARA analysis was performed to
identify the total saturates, aromatics, resins, and asphaltenes. Resins and asphaltenes are known to stabilize the emulsion.
Table 1
Crude Oil Physical Properties

API (at 20C)


Viscosity (at 20C, cP)
Saturates
Aromatics
Resins
Asphaltenes

SP Oil
30.03
13.10
15.4
56.42
4.49
1.61

ASP Oil
28.92
17.95
50
8.7
4.3
0.3

Characterization of produced water: As shown in Tables 2, 3, and 4, the concentrations of surfactants and polymer in the
produced water from SP and ASP floodings are above 1000 ppm. Surfactants and polymer were mainly responsible for the
stability of oil droplets. Polymer increased the interfacial film between water and oil, decreased the zeta potential (more
negative) and blocked aggregation and flocculation of oil droplets. Surfactants decreased the interfacial tension (IFT) and zeta
potential, making oil droplets difficult to approach and coalesce. Alkali such as sodium carbonate reacted with acidic
components of the crude oil and converted them into natural surfactants which can further stabilize the emulsion.
Table 2
Surfactant/Polymer Flooding Produced Water Formulations

NaCl, %
CaCl22H2O , %
Polymer, %
Surfactant (sulfate), %
Surfactant (sulfonate), %
Iso-butyl alcohol, %

Produced Water
1.0
0.18
0.12
0.15
0.05
0.4

Table 3
Alkali/Surfactant/Polymer Concentrations in Produced Water from ASP flooding

Na2CO3 , %
NaCl, %
CaCl22H2O, %
MgCl26H2O, %
NaHCO3 , %
Polymer, %
Surfactant (Sulfonate), %
Surfactant (Sulfate), %
Iso-butyl alcohol, %

Produced Water
1.0
1.0
0.015
0.139
0.043
0.15
0.113
0.038
0.75

Table 4
High Surfactant/Polymer Flooding Produced Water Formulations

NaCl, %
CaCl22H2O , %
Polymer, %
Surfactant (sulfate), %
Surfactant (sulfonate), %
Iso-butyl alcohol, %

Produced Water
1.0
0.18
0.12
0.3
0.1
0.4

SPE 143987

Effects of cationic demulsifier concentrations (patents pending) and salinity:


In our previous work [17],
dodecyltrimethylammonium bromide (C12TAB) proved very effective at breaking the ASP emulsions at 25oC. Figure 2
shows the emulsion stability variation, as a funcion of C12TAB concentration for 1% NaCl and 3% NaCl. Since the stablity
curve is so flattened near the minimum and the range of concentration associated with low stability is so wide (100 ppm to
1,000 ppm for 3% NaCl), the product is robust, thereby providing a large margin for error if overdosed. In other words, any
error in dosing the product close to the optimum formulation would not cause an increase in emulsion stability. However,
addition of C12TAB beyond their optimum (>1000 ppm) resulted in an increase of the emulsion stability (e.g., 3000 ppm),
perhaps due to the formation of an elastic and viscous film caused by the excess of demulsifier. We observed two phenomena
when the salinity is increased from 1% to 3% (Fig. 2). First, the minimum stability occurred at a broader range of demulsifer
concentration. This feature is very beneficial in the field because it indicates a robust condition. Secondly, below the
optimum dosage, the separation time (i.e., emulsion stability) decreased drastically.
Since we are dealing with oil-in-water emulsions, the HLD value is negative. To shift the HLD from negative to nearly zero
(i.e., minimum stability), one can increase the salinity (see Eq. 1). This is in agreement with our observation. Also, an
increase in salinity caused salting-out of the demulsifier (C12TAB) and particularly, the anionic surfactants(sulfate and
sulfonate), resulting in adsorption of these products at the oil-water interface as opposed to being present in the water phase.
This will shift the HLD from negative to nearly zero values. In other words, the increase in salinity enhanced the demulsifier
efficiency.

Figure 2: Stability of the emulsions at 20% oil cut in ASP flood for 1% and 3% NaCL as a function of C12TAB concentration

Effect of the nature of oil: Figure 3 shows the SP and ASP emulsions treated with alkyltrimethylammonium bromides with
different alkyl chain lengths (C8 to C18). The optimum chain length shifted from C8 (SP flood) to C12 (ASP flood). It has
been shown that cyclisation, in particular with double bonds, reduces the equivalent alkane carbon number (EACN) [18]. For
example, the EACN is 6 for hexane, 3.5 for cyclohexane, and -3 for benzene. As shown in Table 1, the SP crude oil contains
much more aromatics than saturates while the ASP crude oil contains more saturates than aromatics. Therefore, an increase in
EACN (from SP crude oil to ASP crude oil) required an increase in the hydrophobic group present in the
alkyltrimethylammonium bromide. For this reason, a transition HLD ~ 0 (SP flood) to HLD <0 (ASP flood) occurs for
C8TAB (octyltrimethylammonium bromide), whence emulsion stability increased. For the same reason, a transition HLD>0
(SP flood) to HLD~0 (ASP flood) takes place for C12TAB, whence emulsion stability decreases.

SPE 143987

Figure 3: SP and ASP bottle tests for 200 ppm alkyltrimethylammonium bromide demulsifiers at 25oC and 30% oil cut

Effect of the temperature: Increasing the temperature from 25oC to 60oC increased the emulsion (ASP) stability for the
treated and untreated samples (Figure 4). The following may explain for the observation. An increase in temperature leads to
an increase in the interactions between the surfactant and the water phase, due to the dissociation of the ionic surfactant. As a
result, the surfactant becomes a little more hydrophilic, causing a more negative HLD value.

Figure 4: Thermal tests for ASP emulsions

Effect of the anionic surfactant concentration: The tests compare the performance of methyl quat demulsifiers in the
normal surfactant SP fluid (2000 ppm anionic surfactant, Table 2) with those in the high surfactant SP fluid (4000 ppm anionic
surfactant, Table 4). C8-TAB was the best performing demulsifier for the normal surfactant SP fluid (2000 ppm anionic
surfactant), but C14-TAB had the best performance for the high surfactant SP fluid (4,000 ppm anionic surfactant) (Figure 5).
The alkyltrimethylammonium bromide demulsifiers were tested at a higher dosage, 300 ppm, in the high surfactant SP fluid
(Figure 6). The C8-TAB performance increased with the higher dosage; however, C12-TAB and C14-TAB still exhibited the
best performance at the 4000 ppm surfactant fluid. This suggests a directly proportional relationship between demulsifier
concentration and surfactant concentration of the SP fluid. It also appears that with the increase in surfactant concentration,
the optimum chain length of demulsifier also increases. Rondon et al. [19] reported that the optimum demulsifier
concentration is proportional to the asphaltene concentration for water-in-oil emulsions.

SPE 143987

Figure 5: Effect of anionic surfactant concentration in SP emulsions at 22 C and 200 ppm demulsifier

Figure 6: Performance of demulsifier at a higher dosage, 300 ppm, with 4000 ppm anionic surfactant

Effect of the water-oil ratio (WOR): The effect of WOR on emulsion stability is complicated and has been investigated by
several authors [20,21]. First, as the WOR is varied, the anionic surfactants (sulfate and sulfonate) and demulsifier (cationic
surfactant) mixture or the ion pairs that is adsorbed at the water-oil interface also changes. Secondly, the WOR also affect the
emulsion properties such as morphology, droplet size, and stability [22]. Table 5 shows the effect of WOR on emulsion
(surfactant-polymer flood) stability at a fixed demulsifier concentration (200 ppm C8TAB). Interestingly, at this dosage,
C8TAB was found to remain effective at various WORs. Conversely, the emulsion stability increased with a decrease in the
WOR for the untreated.
Table 5: Effect of oil cut on emulsion stability treated with C8TAB at 25C with SP emulsion
Dosage (ppm)
200
200
200
200
200
0
0
0

% Oil cut
50
40
30
20
10
50
30
10

Stability (min)
8
8
8
8
8
>240
>240
20

Comment
Good interface
Good interface
Good interface
Good interface
Good interface
Rag layer
Rag layer
Rag layer

Effect of re-emulsification: After 1 day of phase separation, the untreated and treated samples (100 ppm and 200 ppm
C12TAB) were re-shaken. Figure 6 shows that, for the treated samples, it took additional 3 minutes to achieve 100%
separation when compared to the fresh samples. The treated samples, however, produced a much faster phase separation and
better water and oil quality than the untreated.

SPE 143987

Figure 7: Effect of re-emulsification for ASP emulsion at 25oC and 20% oil cut

Effect of demulsifier on the size of oil droplets: Figures 7 and 8 show the particle video microscope images of the oil
droplets in the SP emulsion at 30% oil cut with and without the addition of the demulsifier (C8TAB) at 0 min and 60 min,
respectively. The size of oil droplets for the untreated was about 50 microns and did not change much after 15 minutes. On
the other hand, initially larger droplets were observed when the emulsion was treated with 500 ppm C8TAB and coalesced in a
short period of time (3 minutes). PVM helps visualize droplets in dark and opaque systems and monitor droplet
rearrangement and droplet morphology in real time.

Figure 8: Particle video microscope images for the untreated sample in the SP emulsion at 30% oil cut

Figure 9: Particle video microscope images for 500 ppm C8TAB in the SP emulsion at 30% oil cut

Interfacial tension measurements: Interfacial tension measurements have been investigated by several authors to give
insights into the mechanisms of demulsifier adsorption at the oil-water interface [23,24]. Figure 9 shows results from
equilibrium interfacial tension measurements on C6TAB, C12TAB, C16TAB, and the untreated samples in an ASP emulsions
at 22oC and 60oC. The appearance of the bottles of these samples after 1 day was also shown in Figure 9. It is observed that

SPE 143987

C12TAB and C16TAB exhibited the lowest interfacial tension. The interfacial tension measured for these samples (0.44
dyne/cm) were about 50% smaller than those measured from the untreated and C6TAB samples (0.93 dyne/cm and 0.87
dyne/cm, respectively). The agreement between the interfacial tension measurements and bottle tests is qualitative; i.e., a good
separation with good water quality and low BS&W was observed for the system that exhibited lowest equilibrium interfacial
tension. This finding is consistent with the interpretation of the optimum demulsification efficiency in which a minimum
stability is achieved when the demulsifier adsorbed at the interface exhibits the same affinity for the oil and water phases.
Typically, the interfacial tension values for microemulsion-water interfaces range from 0.01 0.0001 dyne/cm [25], which are
much lower than that measured for C16TAB (0.43dyne/cm). Perhaps due to the low concentration of demulsifiers used in
these experiments (200 ppm ), a middle (microemulsion) layer was not observed. Increasing the temperature from 22oC to
60oC produced a lower interfacial tension for the untreated and a higher interfacial tension for the C12TAB sample. This
finding is consistent with more stable emulsions.

Figure10:EffectofchainlengthonequilibriuminterfacialtensionsofASPemulsionstreatedwithalkyltrimethylammoniumbromide

Effect of demulsifier on zeta potential: Figure 10 shows the effect of demulsifier dosage (octyltrimethylammonium
bromide) on the zeta potentials of oil droplets. It can be seen that zeta potentials of the oil droplets became less negative with
increasing demulsifier concentration, indicating that the ion-pairs of a cationic demulsifier and an anionic surfactant can
adsorb at the oil-water interface and replace the surfactants on the surface of oil droplets. For example, zeta potential
increased from -49 mV with no demulsifier to -28 mV with 400 ppm demulsifier. The decrease in electrostatic repulsion
between oil droplets promotes their coalescence. Surfactant and polymer in produced water can adsorb on the surface of oil
droplets and change the zeta potential. Since both the polymer and surfactants are anionic, the adsorption of these chemicals
increases the density of negative electric charge on the surface of oil droplets, thereby stabilizing the oil droplets via
electrostatic stabilization mechanism. The demulsifier decreased the anionic charge on the surface of oil droplets, lowered the
film elasticity, and decreased the interfacial tension reduction rate, making phase separation faster.

Figure 11: Effect of demulsifier dosage (C8TAB) on zeta potential

10

SPE 143987

Conclusions:
Ion-pairs formed between the cationic demulsifiers (patents pending) and anionic surfactants may explain for the accelerated
separation of oil-in-water emulsions with good oil and water quality during chemical EOR processes. At the optimum
demulsification efficiency, the interfacial tension is lowest , suggesting that the ion-pairs exhibit the same affinity for both
phases. Particle video microscope confirmed that 500 ppm C12TAB produced significant coalescence shortly after it was
added to the emulsions. The cationic surfactant decreased the anionic charge on the surface of the oil droplets, thereby
facilitating the approach and coalescence of oil droplets. This is in agreement with an increase of the oil droplet size in the
presence of the demulsifier. The WOR of the SP emulsion has a significant affect on the emulsion breaking rate for the
untreated sample; however, when treated with 200 ppm C8TAB, the emulsion breaking rate is quick and essentially is the
same for a variety of WORs. This suggests that C8TAB is quite robust. Increasing the temperature and decreasing the
salinity enhance the emulsion stability.
Acknowledgements
The authors would like to thank Nalco for permission to publish this work and express sincere gratitude to Professors George
Hirasaki and Clarence Miller at Rice University for valuable discussion.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.

Thomas, S.; Chemical EOR: The past- Does it have a future? SPE distinguished lecture, 2005
Krawczy, M., Wsan, D., Shetty, C. Chemical demulsification of petroleum emulsions using oil-soluble demulsifiers. Ind. Eng. Chem. Res.,
v. 30, p. 367-375, 1991.
Dukhin, A., Goetz, P. Acoustic and electroacoustic spectroscopy for characterizing concentrated dispersions and emulsions. Adv. Colloid
Interface Sci., v. 92, p. 73, 2001.
Griffin, W.C. J. Soc. Cosmet. Chem., v.1, p. 311-326, 1949.
Winsor, P.A., Solvent Properties of Amphiphilic Compounds, Butterworth, London, 1954.
Goldszal, A; Bourrel, M. Demulsification of crude oil emulsions: correlation to microemulsion phase behavior. Ind. Eng. Chem. Res., v.39,
pg 2746, 2000.
Pena, A.A, Hirasaki, G., and Miller, C., Chemically Induced Destabilization of Water-in-Crude-Oil, Ind. Eng. Chem. Res., v.44, 1139-1149,
2005.
Salager, J.L.; Minana-Perez, M.; Perez-Sanchez, M.; Ramirez-Gouveia, M.; Rojas, C. Surfactant-oil-water systems near the affinity
inversion. 3. The two kinds of emulsion inversion. J. Dispersion Sci. Technol., v.4, p.313, 1983.
Salager, J. L. Emulsion properties and related know-how to attain them. In Pharmaceutical Emulsions and Suspensions; Nielloud, F., MartiMestres, G., Eds.; Marcel Dekker, New York, 1998.
Shinoda, K.; Arai, H. , J. Phys. Chem. 1964, v.68, p.3485, 1964.
Salager, J. L.; Marquez, N.; Graciaa; Lachaise, J. Langmuir, v.16, p.5534, 2000.
Salager, J. L. Int. Chem. Eng., v.30, p.103-116, 1990.
Rodon, M.; Bouriat, P.; Lachaise, J. Breaking of water-in-crude oil emulsion. 1. Physicochemical Phenomenology of demulsifier action,
Energy & Fuels, v.20, p.1600-1604, 2006.
Bourrel, M.; Graciaa, A.; Schechter, R.; Wade, W. H., The relation of emulsion stability to phase behavior and interfacial tension of
surfactant systems, J. Colloid Interface Sci., v.72, p.161-163, 1979.
Salager, J. L.; Morgan, J. C.; Schechter, R. S.; Wade, W. H.; Vasquez, E., Optimum formulation of surfactant/water/oil systems for minimum
interfacial tension or phase behavior. Soc. Petrol. Eng. J., v.19, p. 107-115, 1979.
Kiran, S.K; Acosta, E.J; Malhotra, V.K., Microemulsion modeling of anionic-cationic mixtures for use in detergency. Presented at the AOCS
Nguyen, D.; Sadeghi, N. The selection of the right demulsifier for chemical enhanced oil recovery. SPE # 140860. Presented at the SPE
International Symposium on Oilfield Chemistry, April 2011.
Salager, J. L.; Anton, R.; Forgiarini, A.; Marquez, L. Chapter 3: Formulation of Microemulsion in Microemulsions: background, new
concepts, applications, perspectives. Edited by Stubenrauch C., Blackwell Publishing Ltd., 2009.
Rondon, M.; Pereira, J.C.; Bouriat, P.; Graciaa, A.; Lachaise, J.; Salager, J.L. Breaking of water-in-crude oil emulsions. 2. Influence of
asphaltene concentration and diluent nature on demulsifier action. Energy & Fuels, v. 22, p.702-707, 2008.
Borges, B.; Rondon, M.; Sereno, O.; Asuaje, J. Breaking of water-in-crude oil emulsion. 3. Influence of salinity and water-oil ratio on
demulsifier action. Energy & Fuels, v.23, p.1568-1574, 2009.
Salager, J. The fundamental basis for the action of a chemical dehydrant. Influence of the physical and chemical formulation on the stability
of an emulsion. International Chemical Engineering, v.30, p.103-116, 1990.
Salager, J.L.; Minana-Perez, M.; Perez-Sanchez, M.; Ramirez-Gouveia, M.; Rojas, C. J. Dispersion Sci. Technol. v.4, p.313, 1983.
Goldszal, A.; Bourrel, M. Demulsification of crude oil emulsions: correlation to microemulsion phase behavior. Ind. Eng. Chem. Res., v. 39,
p.2746-2752, 2000.
Breen, P.J. Adsorption kinetics of demulsifiers to an expanded oil-water interface. ACS Symp. Ser., v. 615,
p. 268, 1995.
Salager, J.L. Guidelines for the formulation, composition and stirring to attain desired emulsion properties. In Surfactants in solution;
Chattopadhyay, A.K., Mittal, K.L., Eds.; Surfactant Science Series 64; Marcel Dekker; New York, 1996.

Você também pode gostar