Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Journal of Petroleum Science and Engineering 139 (2016) 198–204

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Influence of the wettability of particles on the morphology and


stability of crude oil–particle aggregates in synthetic produced water
Bartłomiej Gaweł, Meysam Nourani, Thomas Tichelkamp, Gisle Øye n
Ugelstad Laboratory, Department of Chemical Engineering, Norwegian University of Science and Technology (NTNU), N-7491 Trondheim, Norway

art ic l e i nf o a b s t r a c t

Article history: Handling and treatment of produced water is becoming an increasing part of oil and gas production. A
Received 28 August 2015 major challenge in this respect is removal of oil drops stabilised by particles. Understanding the influence
Received in revised form of interfacial and wetting properties of crude oil–particle þ systems on the formation of particle stabi-
12 November 2015
lised emulsions is therefore essential. Elucidating these relationships was the aim of this investigation.
Accepted 16 December 2015
Available online 21 December 2015
Synthetic produced water containing different particles and crude oil were studied. The morphology of
aggregates formed by crude oil drops and particles was strongly related to the contact angle (θwo) at the
Keywords: water–oil–particle interfaces. High θwo resulted in formation of particle stabilised emulsions, while less
Produced water defined aggregates were formed at lower θwo Furthermore, coalescence of drops was related to the
Particle stabilised emulsions
coverage of particles at the interface. Stable drops covered by particles were only seen when the amount
Wettability
of particles was significant. Otherwise drops tended to coalesce, likely due to particle bridging.
o/w Emulsion stability
Separation & 2015 Elsevier B.V. All rights reserved.

1. Introduction Effective coalescence of oil drops is beneficial for ensuring good


quality of the produced water. It is well known that indigenous
Vast amounts of water are produced during oil and gas re- interfacially active components in the crude oil, like asphaltene
covery (Lee et al., 2011). The produced water contains a complex class, can cause formation of mechanically strong, viscoelastic
mixture of dissolved and dispersed components. The latter in- films at the interface (Horváth-Szabó et al., 2005; Pawar et al.,
cludes crude oil drops as well as various inorganic particles ori- 2011; Poteau et al., 2005; Spiecker et al., 2003). High elasticity of
ginating from the reservoir or appearing as scale and corrosion such films can slow down coalescence of oil drops (Gaweł et al.,
products. Removal of dispersed oil is considered to be crucial in 2014b). Furthermore, emulsions containing inorganic particles can
order to reduce the environmental impact of produced water. be even more stable than those stabilized by asphaltenes only
Therefore current regulations in the North Sea require that that (Hannisdal et al., 2006; Nianxi and Masliyah, 1996; Sullivan and
the amount of oil in water must be less than 30 ppm prior to Kilpatrick, 2002; Yan and Masliyah, 1994, 1997; Yan et al., 1999).
discharge to sea (OSPAR Commission, 2010). The ability of particles to stabilise emulsions is attributed to the
Pressure drops and shear forces during transport from the re- large free energy of adsorption of solids which are partially wetted
servoirs to production facilities can lead to extensive mixing of oil by both the oil and water phases (Aveyard et al., 2003; Binks,
drops and particles. The properties of both the oil phase and water 2002). This means that particle layers adsorbed at drop interfaces
phase can alter the surface properties, wettability and stability of can result in extremely stable emulsions (Binks and Lumsdon,
the particles (Dudášová et al., 2014, 2009a, 2009b, 2008; Marc- 1999; Binks and Whitby, 2004; Dong et al., 2014). The wettability
zewski and Szymula, 2002; Sullivan and Kilpatrick, 2002). Fur- is a key factor affecting the properties of the particle layers and
thermore, if particles become trapped at oil–water interfaces, subsequently the emulsion behaviour (Gu et al., 2003; Yan et al.,
particle stabilized emulsions can form. Oil drops stabilised by 2001). The contact angle for oil–water systems is related to in-
particles can be very difficult to remove by typical water treatment terfacial tensions (γ) according to the Young equation:
processes such as hydrocyclones and compact flotation units due
γSO = γSW + γOW cos θ OW (1)
to hindered coalescence and reduced Stokes velocity (Finborud,
2010; Markoff, 2009; Yan and Masliyah, 1997). where the subscripts s, o, and w represent solid, oil and water,
respectively. The oil–water contact angle (θow) is measured
n
Corresponding author. through the water phase. The γow is related to the chemical
E-mail address: gisle.oye@chemeng.ntnu.no (G. Øye). composition of the water phase and crude oil, while γso is related

http://dx.doi.org/10.1016/j.petrol.2015.12.019
0920-4105/& 2015 Elsevier B.V. All rights reserved.
B. Gaweł et al. / Journal of Petroleum Science and Engineering 139 (2016) 198–204 199

to the composition of the crude oil and the adsorption of com- Table 2
ponents at the solid surface (Abdallah et al., 2007). Physical properties of the particles.
The free energy of adsorption of particles is dependent on the
Particle Shape Specific Density (g/cm3) Particle Average ag-
contact angle by the following expression: surface size(from gregate size
area (BET) suppliers) (μm)*
ΔGd = πr 2 (1 − cos (θ ) )2 (2) (m2/g)

where r is the diameter of the particles. This expression states that Kaolin Plate like 19.7 7 3 2.65 0.1–4 μm 2.0
adsorption of particles is strongest when θow is 90°. It is clear from CaCO3 Rhombic 19.9 7 2 2.71 70 nm 1.3
this that the wettability of particles is a very important factor for BaSO4 Prismatic 6.5 7 1 4.5 80–450 nm 1.1
emulsions stability. Another central parameter for particle stabi- Fe3O4 Spherical 45.2 7 3 4.8–5.1 20–30 nm 1.7
FeS Spherical 6.9 7 1 4.6–4.8 2 μm 2.1
lization emulsions is the surface coverage of the particles (Leal-
Calderon and Schmitt, 2008; Pawar et al., 2011). However, there *
Measured by Coulter Counter in 3.5 NaCl solutions.
are extensive variations in the literature with respect to how many
particles are required at the interface in order to constrain coa- densities and particle size/size distributions were provided by the
lescence (Binks and Lumsdon, 1999; Dong et al., 2014; Fre- suppliers, while surface area and average aggregate sizes were
lichowska et al., 2010; Pawar et al., 2011; Tambe and Sharma, 1993; measured previously (Dudášová et al., 2009a, 2009b).
Vignati et al., 2003). Many studies related to particle stabilisation
of emulsions have been focused on relatively simple model sys- 2.3. Brine composition
tems. During crude oil production, however, the fluid stream
consists of crude oil, water and particles which become mixed The brine solution was made by dissolving appropriate
during transport and can form various types of agglomerates like amounts of NaCl (99.5%, Merck, Germany), CaCl2 2H2O (99.5%,
particles stabilised emulsions, schmoo etc. In this work the pro- Merck, Germany), MgCl2  6H2O (99.5%, Merck, Germany), NaHCO3
duced water stream was mimicked by mixing emulsions and (99.5%, Merck, Germany) and Na2SO4 (98.5%, Acros Chemicals) in
particles under mild conditions. The main objective was to in- MQ water from a Millipore Simplicity System. The ionic compo-
vestigate the relations between wetting properties of particles in sition is listed in Table 3
crude oil – synthetic produced water systems and the stability of
the corresponding aggregates. Five different types of particles,
2.4. Contact angles measurements
typically present in produced water streams were used. The wet-
ting properties of the particles, morphology of particle–oil ag-
The particles were pressed into flat pellets (13 mm in diameter)
gregates and the corresponding stability in synthetic produced
under high pressure using a hydraulic press (Compac, Denmark).
water were followed by contact angle, microscopy and Turbiscan
The pellets were put in a custom made holder and immersed in
measurements.
the brine solution before an oil drop was placed on the pellet
surface by a syringe. Images were taken of the oil drop im-
mediately after attachment using an Optical Contact Angle Metre
2. Materials and methods
equipped with a computer-controlled high-speed camera
(CAM200, KSV Instruments). Subsequently, the contact angles
2.1. Crude oil
were determined by fitting the Young Laplace equation to the drop
profiles.
The crude oil used in this study has previously been char-
acterized with respect to its physicochemical and interfacial
properties. Details about the results and experimental procedures 2.5. Microscopy imaging
can be found elsewhere (Gaweł et al., 2014a, 2014b). Some of the
properties are presented in Table 1 for convenience. Crude oil was weighted into 50 ml Schott bottles and the brine
was added to give oil concentration of 10,000 ppm. The samples
were emulsified with an Ultra Turrax mixer (IKA, S25N-10G with
2.2. Particles
10 and 7.5 mm stator and rotor diameter respectively) at
20,000 rpm for 2 min. Subsequently, particles were added to the o/
The inorganic particles were chosen to represent minerals ty-
w emulsion to give a concentration of 5000 ppm. The system was
pically present in produced water (i.e. clay, scale and corrosion
dispersed using an automatic shaker (HS 501 digital IKA) at
products): kaolin (Aldrich, USA/Germany); CaCO3 (98.2%) (Speci-
300 rpm for 1 h. The shaking was used to mimic mixing under
alty Minerals Inc., USA); FeS (99.7%) (DLFTZ, Chang Hing, China);
mild shearing conditions.
BaSO4 (99%) and Fe3O4 (98 þ %) (Nanoamor, USA). Physical prop-
Images of the dispersions were taken by a Nikon LV 100D mi-
erties of the particles are listed in Table 2. The particle shape,
croscope. To avoid any influence of glass slides on the oil drops, the
Table 1
emulsions were poured into rectangular glass capillaries
Physicochemical properties of the crude oil (Gaweł
et al., 2014a). Table 3
Ionic composition of the brine solution.
Property Value
Ions Concentration (ppm)
Density at 20 °C (g/cm3) 0.81
Viscosity at 20 °C (cP) 3 Cl  62,810
Saturates (wt%) 80.0 Na þ 35,393
Aromatics (wt%) 18.0 Ca2 þ 3253
Resins (wt%) 1.9 Mg2 þ 909
Asphaltenes (wt%) 0.1 HCO3  218
TAN (mg/g) 0.4 SO42  49
TBN (mg/g) 0.6
200 B. Gaweł et al. / Journal of Petroleum Science and Engineering 139 (2016) 198–204

(VITROTUBES, 0.4  4 mm2) and sealed. Several images from dif- considered to be true contact angles, but rather used to evaluate
ferent parts of the samples were taken using conventional light the trends.
microscopy and fluorescence microscopy techniques. All images The contact angles when placing a drop of water onto pellets of
with particles were taken in UV light with fluorescence filter. This the same particles were reported previously (Dudášová et al.,
allowed visualisation of fluorescent crude oil species. 2009a, 2009b). Also in this case all the particles were partially
water wet, but with lower contact angles (r 25°). Furthermore,
2.6. Dispersion stability measurements the particles with highest contact angles were CaCO3 and FeS with
angles of 25° and 16°, respectively. These are the particles that
Lower concentration of particles and crude oil were used for gave the lowest contact angles in the oil–water-solid system pre-
the Turbiscan experiments in order to reduce turbidity of the sented here.
samples. Here, crude oil was weighted into glass vials and the An interesting point is that most studies related to contact
brine was added to give oil concentration of 1000 ppm. The angles report water–air contact angles, even though a lot of ap-
samples were emulsified in the same way as above. Subsequently, plications involve oil–water-solid phases. However, a few attempts
particles were added to the o/w emulsion to give a concentration have been made to investigate if there is any relationship between
of 1000 ppm and the system was dispersed by shaking the tubes the water–air and oil–water contact angles. For a range of silanized
by hand for 5 s. surfaces a linear relationship between the two contact angles were
The dispersions were transferred to a Turbiscan LAbExpert in- found, and the oil–water contact angles were 1.4–1.5 times larger
strument (Formulaction, France) immediately after preparation. than the air–water contact angle (Grate et al., 2012). No such re-
The stability was followed by recording the transmission (0° from lationship was found for the samples studied here, as the oil–
incident beam) and backscattering (135° from the incident beam) water contact angles were 1.9–6.4 times higher than the corre-
signals of a pulsed near-infra-red light source (λ ¼ 850 nm) that sponding air–water contact angles, Table 3. This might be due to
moved vertically along the full height of the sample. Complete adsorption of interfacially active components in the crude oil (i.e.
transmission and backscattering scans consisted of signals col-
asphaltenes and resins) onto the particles, resulting in lowering of
lected every 40 μm along the sample height (about 40 mm) and
the solid-oil interfacial tension (γso). This implies that the wetting
were collected every minute for the first hour, every 10 min for the
properties also could be influenced by the crude oil composition.
next 2 h and every 30 min for the last 10 h. The measurements
Here, asphaltenes likely adsorbed onto the solid-oil interface and
were carried out at 50 °C.
thus lowered the interfacial tension.
Since the samples studied here were dilute enough to allow
considerable transmission, the detected backscattering signals also
contained internal reflections from the measurement cell (Men- 3.2. Dispersion properties
gual et al., 1999a, 1999b). Consequently, only the transmission
signals were considered in the data analysis Fig. 1 shows the dispersions containing crude oil and particles
just after shaking at 300 rpm for 1 h. The separation started im-
mediately after the shaking was terminated and the o/w emulsion
3. Results and discussion (without any particles present), shown as a reference on the left,
quickly got a thin oil layer at the top, indicating that not all the oil
3.1. Wetting behaviour became dispersed. The samples with the particles generally

Table 4 shows the contact angles for a drop of the crude oil
placed on the different particles surrounded by water. All the
particles are partially water wetting with contact angles less than
90°. The lowest contact angles were seen for CaCO3 and FeS. Ac-
cording to Eq. (2), this corresponds to low free energy for de-
tachment of particles from oil–water interfaces. The contact angles
were somewhat higher for BaSO4 and Fe3O4 and highest for
kaolinite.
It is obvious that surface roughness will influence the contact
angle measurements by the current method. Particle size, shape
and specific surface area are parameters that are expected to have
relatively large influence on the surface roughness. Considering
CaCO3 and FeS, however, these resulted in similar contact angles,
while the particle shape, particle size and specific surface were
markedly different (Table 2). It was therefore considered that the
chemical properties had larger influence on the observed contact
angle than the differences in surface roughness of the pellets from
the various powders. Furthermore, the measured values were not

Table 4
Oil–water (θow) and water–air (θow) contact angles for the various particles.

Particle θow(°) θow(°) θow/θwa

Kaolinite 707 10 12 72 5,8


CaCO3 48 7 10 25 72 1,9
BaSO4 587 10 9 72 6,4
Fe3O4 577 10 10 72 5,7 Fig. 1. The dispersions immediately after shaking. From left to right the following
FeS 477 10 16 72 2,9 particles were added to the o/w emulsions: Pure oil, kaolinite, CaCO3, BaSO4, Fe3O4
and FeS. Bottom image shows enlarged view of kaolinite, CaCO3 and BaSO4 sample.
B. Gaweł et al. / Journal of Petroleum Science and Engineering 139 (2016) 198–204 201

Fig. 2. Crude oil drops (1 wt%) covered by BaSO4 particles (0.5 wt%).

separated into three layers; a top layer that included mixture of


crude oil and particles, a middle layer with different degrees of
turbidity and a bottom layer which primarily contained particles,
but also some oil remains. Furthermore, it could be distinguished
between two groups of behaviour. The following observations
were made for the samples containing kaolinite, BaSO4 and Fe3O4:

1. Initially, drops covered with particles rose to the top of the


sample. The attachment of particles to the oil water interface
occurred during the shaking of the samples. This is illustrated
for the BaSO4 system in Fig. 2, where the drops have diameters
around 100 μm and the particles are about 1.1 μm. Particle
aggregates can also be seen.
2. Dense-packed emulsion layers formed at the top of the samples.
This is shown for the BaSO4 system in Fig. 3A. Coalescence
within the dense-packed layer resulted in a thin oil film at the
top and formation of particle-rich agglomerates, most likely
containing remains of trapped oil, which gradually sank to the
bottom of the sample. The sediment of the BaSO4 system is
shown in the image in Fig. 3B.

The samples containing CaCO3 and FeS behaved differently


from the above description in the way that the particles/ag-
gregates sank directly to the bottom of the sample, i.e. without
initially being creamed together with the oil. However, interac-
tions of CaCO3 particles with the water–air surface resulted in a Fig. 3. The top layer (A) and bottom layer (B) of the dispersion prepared with crude
top layer of particles in this case. oil (1 wt%) and BaSO4 particles (0.5 wt%).
Microscopy imaging also showed that the differences in dis-
persion behaviour were associated with different morphologies of
particle-crude oil agglomerates in the samples, Fig. 4. The top left required to provide stability towards coalescence vary widely be-
image shows oil drops in the emulsion before addition of particles tween various systems reported in the literature. Some show that
and the subsequent shaking. The average drop size was 5.5 μm. No even small amounts of interfacial particles will constrain coales-
spherical oil drops were observed in the samples containing CaCO3 cence (Binks and Lumsdon, 1999; Dong et al., 2014; Vignati et al.,
and FeS particles. Instead, irregularly shaped structures were seen, 2003), while others show that coalescence can take place until full
and particularly in the sample with CaCO3 particles extended coverage with particles is reached (Pawar et al., 2011). The results
network structures were apparent. Notably, these were the parti- here suggest that coalescence continued until particles were
cles with the lowest water–oil contact angles, and consequently densely packed at the oil drops (Fig. 2). The mechanism might be
the lowest affinity for the oil–water interface. particle bridging between neighbouring oil drops at low particle
Spherical drops were, on the other hand, observed in the coverage. This would result in increased contact time between the
samples containing BaSO4, Fe3O4 and kaolinite particles. These drops and promote coalescence (and flocculation) in the samples.
were the samples with the highest water–oil contact angles (θwo) Such bridging of fluid interfaces by particles has been reported by
and thereby highest affinity for the oil–water interface. The aver- others (Nixon et al., 1999).
age drop sizes for the samples containing BaSO4, Fe3O4 and kao-
linite particles were 73 μm, 52 μm and 42 μm, respectively. This 3.3. Stability measurements
was significantly larger than the average drop size of 5.5 μm seen
in the o/w emulsion prior to the addition of particles. This means Sedimentation, creaming, aggregation, flocculation and coa-
that considerable coalescence took place during shaking of the lescence are the kinetic phenomena that can destabilize the in-
samples. Flocculation of larger drops was also observed vestigated dispersions and ultimately give rise to phase separation.
(Figs. 2 and 4). Furthermore, it was clear that the particles influ- Information about these phenomena was obtained by following
enced the coalescence process. The smallest average drop size the changes in transmission signals over time. Typical transmis-
(42 μm) was observed with kaolinite present. The higher affinity sion data obtained by the Turbiscan LAb are shown in Fig. 5. Going
of kaolinite to the interface might provide a more efficient barrier from left to right along the x axis corresponds to moving from the
towards further coalescence than for the samples with BaSO4 and bottom to the top of the sample (i.e. the height of the dispersions
Fe3O4 present. However, the particle shape might also be a factor. was about 40 mm). The corresponding transmissions are shown
The extent of particle coverage at an oil–water interface along the y axis, while the time dependency is indicated by the
202 B. Gaweł et al. / Journal of Petroleum Science and Engineering 139 (2016) 198–204

Fig. 4. The o/w emulsion without added particles (oil drops were added colour to increase the contrast) and the dispersions were and the dispersions were various particles
(0.5 wt%) were added to the emulsion (1 wt%).

different colours for each scan along the sample where the purple
and red scans represent the first and last measurement,
respectively.
The increase in the transmission at the bottom of the sample in
Fig. 5A was associated with creaming of drops in the o/w emulsion
for the initial measurements, while the creaming became less
dominant at longer times. Flocculation and coalescence of drops
was followed by changes in transmission in the region of the
sample where creaming could be neglected, i.e. between 20 and
30 mm above the bottom of the sample. The increase in trans-
mission caused by flocculation/coalescence is in agreement with
Mie theory, predicting larger scattering in the forward directions
by particles larger than the wavelength of the light source (Men-
gual et al., 1999a, 1999b).
Fig. 5B shows the transmission data for the suspended BaSO4
particles. The increase in transmission at the top of the sample
corresponds to sedimentation, due to reduced concentration of
particles. In this case the sedimentation could not be neglected
between 20 and 30 mm above the bottom of the sample (the se-
dimentation front dominates also in this region). Similar beha-
viour was observed for FeS and kaolinite particles. For CaCO3 and
Fe3O4 particles, however, the sedimentation was preceded by
flocculation (see Fig. S1 in Supplementary material) and the
change in transmission in this range was associated with ag-
gregation of particles.
The transmission data for the sample where BaSO4 particles
and crude oil drops were mixed are shown in Fig. 5C. Some se-
dimentation was observed for this sample. Similar observations
were made when kaolinite and FeS particles were mixed with the
emulsion, while creaming was seen for the samples mixed with
CaCO3 and Fe3O4 particles (See Fig. S1 in Supplementary material).
The sedimentation or creaming could be neglected between 20
and 30 mm above the bottom of the sample for all the dispersions.
The change in transmission in this range could then be due to
Fig. 5. Turbiscan transmission profiles for (A) the o/w emulsion (1000 ppm oil) flocculation and coalescence of drops, in agreement with micro-
without particles, (B) the suspension of BaSO4 particles (1000 ppm) and (C) the scopy observations, as well as aggregation of particles.
dispersion where BaSO4 particles (1000 ppm) were added to the emulsion. Each
Fig. 6 shows the transmission in the middle of the sample for
measurement is identified by a colour and the corresponding measurement times
are listed to the right. (For interpretation of the references to colour in this figure the suspensions with the various particles. For the suspensions
legend, the reader is referred to the web version of this article.) with BaSO4, FeS and kaolinite there was an induction time of about
B. Gaweł et al. / Journal of Petroleum Science and Engineering 139 (2016) 198–204 203

microscopy.

4. Conclusions

The behaviour and stability of crude oil drops dispersed to-


gether with various inorganic particles in synthetic produced
water were investigated. It was found that the three phase contact
angle (θwo) at the water–oil–particles interface was important for
the morphology of crude oil–particle aggregates. High θwo corre-
sponded to formation of particle stabilised emulsions, while low
θwo resulted in less well-defined types of aggregates. It was also
suggested that the extent of coalescence of oil drops was depen-
dent on the particle coverage at the oil–water interface.

Author contributions
Fig. 6. Time dependent increase of transmission in the middle (i.e. between 20 and
30 mm above the bottom) of the samples with the various suspended particles. The
The manuscript was written through contributions of all au-
concentration of the particles and the crude oil was 1000 ppm. thors. All authors have given approval to the final version of the
manuscript.

Acknowledgements

The authors are grateful to the industrial sponsors (Con-


ocoPhillips Skandinavia, ENI Norge, Schlumberger Norge PWMS,
Statoil Petroleum and Total E&P Norge) (Project number
40114800) of the joint industrial programme “Produced Water
Management: Fundamental Understanding of the Fluids” for fi-
nancial support.

Appendix A. Supplementary material

Supplementary data associated with this article can be found in


the online version at http://dx.doi.org/10.1016/j.petrol.2015.12.019.

Fig. 7. Time dependent increase of transmission in the middle (i.e. between 20 and References
30 mm above the bottom) of the o/w emulsion and the dispersions where the
various particles were added to the emulsion. The concentration of crude oil and
Abdallah, W., Buckley, J.S., Carnegie, A., Edwards, J., Herold, B., Fordham, E., Graue,
particles was 1000 ppm.
A., Habashy, T., Seleznev, N., Signer, C., Hussain, H., Montaron, B., Ziauddin, M.,
2007. Fundamentals of wettability. Oilfield Rev. 19, 44–61.
Aveyard, R., Binks, B.P., Clint, J.H., 2003. Emulsions stabilised solely by colloidal
10 min before any detectable change in the transmission. This
particles. Adv. Colloid Interface Sci. 100–102, 503–546.
suggested that sedimentation was the more dominant process in Binks, B.P., 2002. Particles as surfactants-similarities and differences. Curr. Opin.
these samples, which also was in agreement with the overall Colloid Interface Sci. 7, 21–41.
Binks, B.P., Lumsdon, S.O., 1999. Stability of oil-in-water emulsions stabilised by
transmission data. Furthermore, the rapid increase in transmission
silica particles. Phys. Chem. Chem. Phys. 1, 3007–3016.
for the BaSO4 sample was due to an overlap with the sedimenta- Binks, B.P., Whitby, C.P., 2004. Silica particle-stabilized emulsions of silicone oil and
tion front of the sample. In the CaCO3 and Fe3O4 suspensions on water: aspects of emulsification. Langmuir 20, 1130–1137.
the other hand, aggregation of particles were initially the domi- Dong, J., Worthen, A.J., Foster, L.M., Chen, Y., Cornell, K.A., Bryant, S.L., Truskett, T.M.,
Bielawski, C.W., Johnston, K.P., 2014. Modified montmorillonite clay micro-
nant process. The size of the calcium carbonate and iron oxide particles for stable oil-in-seawater emulsions. ACS Appl. Mater. Interfaces 6,
particles were smaller compared to the other types of particles 11502–11513.
(Dudášová et al., 2009a, 2009b), and the aggregates of a certain Dudášová, D., Flåten, G.R., Sjöblom, J., Øye, G., 2009a. Study of asphaltenes ad-
sorption onto different minerals and clays. Part 2. Particle characterization and
size was required before sedimentation became significant. suspension stability. Colloids Surf. A 335, 62–72.
The transmission profiles in the middle of the samples where Dudášová, D., Rune Flåten, G., Sjöblom, J., Øye, G., 2009b. Stability of binary and
particles were mixed with the o/w emulsion, are shown in Fig. 7. ternary model oil-field particle suspensions: a multivariate analysis approach. J.
Colloid Interface Sci. 337, 464–471.
Overall, the transmission increased faster when the emulsion was Dudášová, D., Simon, S., Hemmingsen, P.V., Sjöblom, J., 2008. Study of asphaltenes
mixed with particles. The profile of the sample containing CaCO3 adsorption onto different minerals and clays. Part 1. Experimental adsorption
was most similar to that of the o/w emulsions. These particles had with UV depletion detection. Colloids Surf. A 317, 1–9.
Dudášová, D., Sjöblom, J., Øye, G., 2014. Characterization and suspension stability of
low affinity for the oil–water interface and irregular aggregates
particles recovered from offshore produced water. Ind. Eng. Chem. Res. 53,
were observed in the microscope. Fast increase and overall high 1431–1436.
transmission were seen for the samples containing BaSO4 and Finborud, A., 2010. Design for optimal operation of produced water treatment.
kaolinite. For these particles, particle stabilised drops were ob- TEKNA Produced Water Management Conference, Stavanger.
Frelichowska, J., Bolzinger, M.-A., Chevalier, Y., 2010. Effects of solid particle content
served in the microscope. The increased transmission was also on properties of o/w Pickering emulsions. J. Colloid Interface Sci. 351, 348–356.
consistent with the coalescence and flocculation observed by Gaweł, B., Eftekhardadkhah, M., Øye, G., 2014a. Elemental composition and fourier
204 B. Gaweł et al. / Journal of Petroleum Science and Engineering 139 (2016) 198–204

transform infrared spectroscopy analysis of crude oils and their fractions. En- Nianxi, Y., Masliyah, J.H., 1996. Demulsification of solids-stabilized oil-in-water
ergy Fuels 28, 997–1003. emulsions. Colloids Surf. A 117, 15–25.
Gaweł, B., Lesaint, C., Bandyopadhyay, S., Øye, G., 2014b. The role of physico- Nixon, A.J., Pacek, A.W., Nienow, A.W., Norton, I.T., 1999. Influence of fat crystals on
chemical and interfacial properties on the binary coalescence of crude oil drops the equilibrium drop size and coalescence rate in agitated sunflower oil/water
in synthetic produced water. Energy Fuels 29, 512–519. dispersions. Inst. Chem. Eng. Symp. Ser., 177–186.
Grate, J.W., Dehoff, K.J., Warner, M.G., Pittman, J.W., Wietsma, T.W., Zhang, C., OSPAR Commission, 2010. Discharges, Spills and Emissions from Offshore Oil and
Oostrom, M., 2012. Correlation of oil–water and air–water contact angles of Gas Installations, London.
diverse silanized surfaces and relationship to fluid interfacial tensions. Lang- Pawar, A.B., Caggioni, M., Ergun, R., Hartel, R.W., Spicer, P.T., 2011. Arrested coa-
muir 28, 7182–7188. lescence in Pickering emulsions. Soft Matter 7, 7710–7716.
Gu, G., Zhou, Z., Xu, Z., Masliyah, J.H., 2003. Role of fine kaolinite clay in toluene- Poteau, S., Argillier, J.-Fo, Langevin, D., Pincet d.r., F., Perez, E., 2005. Influence of pH
diluted bitumen/water emulsion. Colloids Surf. A 215, 141–153. on Stability and dynamic properties of asphaltenes and other amphiphilic
Hannisdal, A., Ese, M.-H., Hemmingsen, P.V., Sjöblom, J., 2006. Particle-stabilized molecules at the oil–water interface. Energy Fuels 19, 1337–1341.
emulsions: effect of heavy crude oil components pre-adsorbed onto stabilizing Spiecker, P.M., Gawrys, K.L., Trail, C.B., Kilpatrick, P.K., 2003. Effects of petroleum
solids. Colloids Surf. A 276, 45–58. resins on asphaltene aggregation and water-in-oil emulsion formation. Colloids
Horváth-Szabó, G., Masliyah, J.H., Elliott, J.A.W., Yarranton, H.W., Czarnecki, J., 2005. Surf. A 220, 9–27.
Adsorption isotherms of associating asphaltenes at oil/water interfaces based Sullivan, A.P., Kilpatrick, P.K., 2002. The effects of inorganic solid particles on water
on the dependence of interfacial tension on solvent activity. J. Colloid Interface and crude oil emulsion stability. Ind. Eng. Chem. Res. 41, 3389–3404.
Sci. 283, 5–17. Tambe, D.E., Sharma, M.M., 1993. Factors controlling the stability of colloid-stabi-
Leal-Calderon, F., Schmitt, V., 2008. Solid-stabilized emulsions. Curr. Opin. Colloid lized emulsions. I. An experimental investigation. J. Colloid Interface Sci. 157,
Interface Sci. 13, 217–227. 244–253.
Lee, K., Neff, J., DeBlois, E., 2011. Produced Water: Overview of Composition, Fates, Vignati, E., Piazza, R., Lockhart, T.P., 2003. Pickering emulsions: interfacial tension,
and Effects, Produced Water. Springer, New York, pp. 3–54. colloidal layer morphology, and trapped-particle motion. Langmuir 19,
Marczewski, A.W., Szymula, M., 2002. Adsorption of asphaltenes from toluene on 6650–6656.
mineral surface. Colloids Surf. A 208, 259–266. Yan, N., Gray, M.R., Masliyah, J.H., 2001. On water-in-oil emulsions stabilized by fine
Markoff, C., 2009. Evaluation of available technological solutions for valhall pro- solids. Colloids Surf. A 193, 97–107.
duced water. TEKNA Produced Water Management Conference, Stavanger. Yan, N., Masliyah, J.H., 1994. Adsorption and desorption of clay particles at the oil–
Mengual, O., Meunier, G., Cayre, I., Puech, K., Snabre, P., 1999a. Characterisation of water interface. J. Colloid Interface Sci. 168, 386–392.
instability of concentrated dispersions by a new optical analyser: The TUR- Yan, N., Masliyah, J.H., 1997. Creaming behavior of solids-stabilized oil-in-water
BISCAN MA 1000. Colloids Surf. A 152, 111–123. emulsions. Ind. Eng. Chem. Res. 36, 1122–1129.
Mengual, O., Meunier, G., Cayré, I., Puech, K., Snabre, P., 1999b. TURBISCAN MA Yan, Z., Elliott, J.A.W., Masliyah, J.H., 1999. Roles of various bitumen components in
2000: multiple light scattering measurement for concentrated emulsion and the stability of water-in- diluted-bitumen emulsions. J. Colloid Interface Sci.
suspension instability analysis. Talanta 50, 445–456. 220, 329–337.

You might also like