You are on page 1of 8

Article

pubs.acs.org/EF

Role of Physicochemical and Interfacial Properties on the Binary


Coalescence of Crude Oil Drops in Synthetic Produced Water
Bartłomiej Gaweł, Caterina Lesaint, Sulalit Bandyopadhyay, and Gisle Øye*
Ugelstad Laboratory, Department of Chemical Engineering, Norwegian University of Science and Technology (NTNU), N-7491
Trondheim, Norway
*
S Supporting Information

ABSTRACT: An important aspect of efficient treatment of produced water is to promote the coalescence of dispersed oil drops.
Therefore, a fundamental understanding of the relations between interfacial and physicochemical properties of crude oils and the
coalescence process is essential, and elucidating these relationships was the aim of this investigation. Nine crude oils and fractions
of two crude oils, where acids, bases, and asphaltenes were selectively extracted, were included in the studies. The dynamic
interfacial tension and interfacial rheology of the oils were followed in synthetic produced water. The binary coalescence of oil
drops was followed by a micropipette setup, and characteristic times for three phenomena occurring during the coalescence
process were identified: the time of film thinning, the time for rupture of the thin film, and the apparent time of merger of the
two drops. Decreasing the density of the crude oils and increasing the strength of the interfacial layer were determined to
influence the coalescence process most significantly, by slowing it down. The elasticity and strength of the interfacial layers were
primarily associated with the presence of asphaltenes at the interface. The strength of the interfacial layer decreased as the
amount of aromatics and acidic components in the crude oil increased. In both cases, this was attributed to improved stability of
the asphaltenes in the bulk of the crude oil.

1. INTRODUCTION by the Bond Number,12 which is proportional to the density


Handling and treatment of produced water generated during the difference between the continuous and dispersed phases.
production of oil and gas is of great concern for the petroleum Furthermore, Gibbs interfacial elasticity affects the film thinning
industry. Produced water can be a mixture of formation water process, where high elasticity indicates formation of immobile
and injected seawater that is back-produced. It contains interfaces and reduced coalescence.13,14
dispersed oil and solids as well as dissolved organics, solids, Several studies on the coalescence between two isolated drops
and production chemicals. The total volume of produced water using well-defined model systems have been reported.15−18 A
on the Norwegian Continental Shelf (NCS) was 130 million m3 much more unexplored area is the coalescence of crude oil drops
in 2012.1 Even though increasing amounts are reinjected into the in complex systems, such as produced water. Understanding the
reservoirs, ∼85% of the water is still discharged to the ocean.2 interfacial phenomena in such systems is considered to be
The volumes of produced water increase as the fields mature, and important in view of increasing the separation efficiency of crude
there is considerable concern over its disposal, because of the oil from the produced water, which will be valuable for both the
presence of hazardous components. Current regulations at the environment and the oil industry. Previously, we have reported
NCS require that the oil-in-water concentration must be <30 on the coalescence between crude oil drops and bubbles.19 The
ppm if the water is discharged to sea.3 However, zero harmful aim of the current work was to investigate how the interfacial and
discharge is the objective. physicochemical properties of crude oils influence the binary
Technologies for treating produced water include methods coalescence crude oil drops. The dynamic interfacial tension and
such as gas flotation, extraction, filtration, and hydrocyclones.3−7 dynamic interfacial rheology were studied for nine crude oils in
Almost 90% of produced water treatment facilities include synthetic produced water. Furthermore, asphaltenes, as well as
hydrocyclones, because of high efficiency and low installation acidic and basic components, were selectively extracted from two
size.8 The removal efficiency of the oil drops is dependent on the of the crude oils in order to elucidate their influence on the
size of the drops and the density difference between oil and interfacial properties. Binary drop coalescence for the crude oils
continuous phase.9 However, the size of the drops is strongly was measured using a drop bubble micromanipulator.
related to the coalescence process of the drops.
Coalescence of oil drops involves close approach of the drops,
2. EXPERIMENTAL SECTION
formation and drainage of the thin aqueous film between the
drops, and finally rupture of the film.10 The time of film thinning 2.1. Characterization of Crude oils. Nine crude oils (denoted A−
(i.e., drainage time) is dependent on whether the drops are I) were characterized with respect to density, viscosity, SARA fractions,
deformable or not.10,11 For deformable drops, the thinning of the
film is within the Reynolds regime, where the drainage time Received: August 18, 2014
increases significantly with the size of the drops.10 The Revised: January 14, 2015
deformation of drops in the presence of gravity can be described

© XXXX American Chemical Society A DOI: 10.1021/ef501847q


Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

total acid number (TAN), and total base number (TBN). Descriptions ⎛A ⎞
of the detailed procedures can be found elsewhere.20 E″ = Δγ ⎜ 0 ⎟ sin(φ)
⎝ ΔA ⎠ (3)
2.2. Extraction of Acids and Bases. The extraction of acidic and
basic components from the crude oil, followed a modified version of the where Δγ is the amplitude of the periodic variation in surface tension, A0
solid phase extraction method developed by Simon et al.21,22 Detailed the area of the drop before oscillation, ΔA the amplitude of the
procedures for the extractions are described in the Supporting interfacial deformation, and φ the phase angle between the oscillating
Information. area and surface tension curves.
2.3. Aqueous Solution. The synthetic produced water was made by The time-dependent evolution of the elastic and viscous moduli was
dissolving appropriate amounts of NaCl (99.5%, Merck, Germany), followed by immersing a drop of crude oil into the appropriate aqueous
CaCl2·2H2O (99.5%, Merck, Germany), MgCl2·6H2O (99.5%, Merck, solution. The drop volume was sinusoidally varied by 5% at a frequency
Germany), NaHCO3 (99.5%, Merck, Germany), and Na2SO4 (98.5%, of 0.1 Hz, while images of the drop were recorded for 1 h using a drop
Acros Chemicals) in Milli-Q (MQ) water (Millipore Simplicity tensiometer (PAT-1, Sinterface Technologies, Germany). The corre-
System). The ionic concentration of the brine is listed in Table 1. sponding deformation of the interfacial area (ΔA/A0) was ∼4% and
within the range of linear deformation. The oscillating interfacial tension
was measured with a resolution of 20 points per second and was used to
Table 1. Ionic Composition of the Brine Solutions
calculate the time-dependent moduli, according to eqs 2 and 3 using a
ion concentration [ppm] range of 100 s. The measurements were carried out at 50 °C and were
reproducible within ±2 mN/m. It was not possible to conduct
Cl− 62810
measurements for crude oils C, D, F, and I, because of the high
Na+ 35393
viscosity of the oil or unstable drops during oscillations.
Ca2+ 3253 2.6. Strength of Interfacial Layers. Oscillation frequency scans
Mg2+ 909 were also carried out directly after the interfacial tension measurements
HCO3 − 218 described above had reached equilibrium. The oscillation frequency of
SO42− 49 the drop varied from 0.01 Hz to 0.1 Hz. Oscillations were carried out for
1000 s at 0.01 and 0.02 Hz, 500 s at 0.025 Hz, and 200 s at 0.1 Hz. The
deformation of the drop was ∼4%.
According to Lucassen and Van den Tempel (LVDT) theory, the
2.4. Dynamic Interfacial Tension Measurements. A pendant formation of elastic interface for systems with low concentration of
drop tensiometer (PAT-1, Sinterface Technologies, Germany) was used interfacially active components (e.g., asphaltenes) is related to
to measure the interfacial tensions between crude oil and synthetic diffusional exchange between the interface and bulk solution.23 In this
brine. A drop of crude oil was immersed into the aqueous solution by a case, the characteristic c modulus follows a power-law dependency over
curved needle, and images of the drop were recorded over time. frequency and the characteristic strength of the interface (S) can be
Subsequently, the contours of the drops were fitted to the Laplace− calculated according to the following equation:23
Young equation by a numerical procedure. Drop images were captured
every second for 4000 s, and the measurements were carried out at 50 |E| ∝ Sf n (4)
°C.
where
2.5. Dynamic Interfacial Rheology Measurements. The
interfacial rheology was described in terms of the Gibbs interfacial ⎛2⎞
dilatational modulus and is defined as the change in interfacial tension n = φ⎜ ⎟
⎝π ⎠ (5)
due to small variations in the area of a surface element:

E=
dγ |E| = E′2 + E″2 (6)
d ln A (1)
φ is the phase angle of the Gibb’s dilatational modulus E, f is the
where γ is the interfacial tension and A is the area of the surface element. oscillation frequency, and n is a constant (0 < n < 1). S and n were
The dilatational modulus can be expressed as E = E′ + iE″ when the extracted by fitting experimental parameters to the linear function of log
surface area of the drop is oscillated periodically. The real part, E′, is the E versus log f, after ensuring good correspondence between measured
elastic (storage) modulus, which accounts for the recoverable energy and calculated phase angles (see Figure S1 in the Supporting
stored in the interface, while the imaginary part, E″, is the viscous (loss) Information).
modulus, which accounts for the dissipation of energy through 2.7. Binary Drop Coalescence Measurements. The coalescence
relaxation processes. The two moduli are defined as follows for of two droplets was followed by attaching a drop and bubble
sinusoidal oscillations of the surface area at a given frequency: micromanipulator unit (DBMM-1, Sinterface Technologies, Germany)
⎛A ⎞ onto the PAT-1 drop tensiometer. The DBMM unit is essentially a
E′ = Δγ ⎜ 0 ⎟ cos(φ) capillary-pressure tensiometer that consists of two stainless-steel cells
⎝ ΔA ⎠ (2) containing piezo-excitation actuators and the pressure sensors in

Table 2. Characterisation Data for the Crude Oils


A B C D E F G H I
density @ 50 °C (g/cm3) 0.78 0.90 0.91 0.91 0.81 0.82 0.92 0.85 0.9
viscosity @ 50 °C (cP) 1.5 17.8 70.4 50.8 4.2 4.7 59.6 6.1 32.8
saturates (wt %) 80 44.3 25.6 58.1 67.7 62.2 53.5 60 52
aromatics (wt %) 18 38.8 49.6 30.7 27.1 31.2 29.9 36 32
resins (wt %) 1.9 16.4 10.9 9.5 4.4 5.9 14.4 6.7 13.9
asphaltenes (wt %) 0.1 0.5 13.9 1.7 0.8 0.7 2.5 0.3 2.1
resin/asphaltenes ratio 19 33 0.8 6 6 8 6 22 7
TAN (mg/g)a 0.4 (0.2) 2.3 (1.8) 0.5 0.3 0.3 0.3 2.2 2.7 1.4
TBN (mg/g)a 0.6 (0.6) 4.4 (3.6) 1.3 1.5 1.1 1.5 2.9 0.8 3.0

a
Values shown in parentheses represent values measured for asphaltenes-extracted fractions.

B DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

opposite positions. A capillary is connected to each of the cells so that


the tips are facing each other and the distance between them can be
manually controlled by an xyz-axis microtranslation stage. Further
details about the setup can be found elsewhere.24
The cells and capillaries were purged with water prior to each
experiment. The glass capillaries (internal diameter of ∼0.5 mm) then
were filled with crude oil by suction from a vial. Each capillary contained
∼10 μL of oil, which was usually sufficient for 5−10 repeated
measurements. The capillaries were immersed in the brine solution
and crude oil drops of similar size (diameter of about 0.5 mm) were
formed at each tip. One drop was approached toward the other
immediately after formation. All experiments were carried out at 50 °C.

3. RESULTS AND DISCUSSION


3.1. Physicochemical Properties of the Crude Oils. The
physicochemical properties of the crude oils are gathered in
Table 2. The densities ranged from 0.78 g/cm3 to 0.92 g/cm3,
where the three lightest oils (A, E, and F) had densities less than Figure 1. Dynamic interfacial tension of crude oil A and its fractions in
0.82 g/cm3 and the remaining oils had densities above 0.85 g/ brine.
cm3. The lightest crude oils contained the lowest amounts of
resins and highest amounts of saturates, which likely accounted
for the low densities and viscosities for these oils. The asphaltene
content of crude oil C (13.9 wt %) was considerably higher than components contributed to reducing the interfacial tension of
for the other oils, for which it ranged from 0.1 to 2.5 wt %. Crude the crude oils, which is in agreement with previous
oil C was also the only oil where the amount of asphaltenes was observations.25 Furthermore, the slower kinetics for these
higher than the amount of resins. The TBN values ranged from fractions suggests that relatively small interfacially active
0.6 mg/g to 4.4 mg/g. Five of the crude oils (A, C, D, E, and F) molecules were removed during the extraction of the acidic
had low TAN values, ranging from 0.3 mg/g to 0.5 mg/g, while and basic components and substitution by larger components
the remaining oils had higher values (1.4−2.7 mg/g). The only slows the kinetics. The asphaltene-extracted fraction, on the
acidic oil was crude oil H. other hand, had lower interfacial tension than the crude oil. It is
3.2. Interfacial Tension. The dynamic interfacial tension inferred from this that asphaltenes were also present at the crude
decayed uniformly, at different rates, toward equilibrium values oil/brine interface. The cross-sectional area that asphaltenes
(see Figure S2 in the Supporting Information). The kinetic occupy at the interface is relatively large26 and when they are
variation can be related to the abundance, molecular structure, extracted from the crude oil, smaller interfacially active
aggregation state, and mobility of the interfacially active components can access the available space. The higher fraction
species.25−31 The equilibrium interfacial tension of the crude of molecules at the interface can then account for the lower
oils in synthetic brine ranged from 4.2 mN/m to 17.4 mN/m, as interfacial tension observed for this fraction. Overall, the dynamic
shown in Table 3. interfacial tension data suggested that the extracted constituents
of one oil fraction become replaced at the interface by the
Table 3. Interfacial Properties of the Crude Oils constituents of the remaining fractions.
3.3. Dynamic Interfacial Rheology. The elastic moduli
crude oil IFTeq (mN/m) interfacial strength, S n
were larger than the viscous moduli to different extents for all the
A 8.9 27 0.33 investigated oils, indicating various degrees of elasticity of the
B 9.5 9 0.24
interfaces. Three of the crude oils (A, E, and H) showed an
C 4.2
increase of the elastic modulus over time, while it was
D 12.6
independent of time for the remaining oils (B and G) (see
E 17.4 27 0.30
Figure 2).
F 5.3 16 0.18
G 16.8 16 0.53
H 16.9 9 0.03
I 14.5 10 0.30

No strong correlations were found between the IFT values and


the amount of any of the crude oil fractions. However, the
influence of crude oil composition on the interfacial tension was
further investigated by selectively extracting acidic, basic, and
asphaltenic components from two of the oils (A and B). As
presented in Figure 1, the dynamic interfacial tension decayed
uniformly, but at different rates, for the various fractions of crude
oil A.
The trends were similar for crude oil B (see Figure S3 in the
Supporting Information). Both the acid- and base-extracted
fractions had higher interfacial tensions than the crude oil.
Consequently, both acidic and basic interfacially active Figure 2. Dynamic interfacial elasticity of the crude oils in brine.

C DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

The increase in elasticity is attributed to adsorption and pounds for the interfacial elasticity of the crude oils were clearly
subsequent rearrangement of asphaltenes at the interface, and revealed. The low elastic modulus for the asphaltenes-extracted
this has also been reported by other researchers.23,32,33 fraction confirms the importance of the asphaltenes for the
Further investigations of the various fractions of crude oils A elasticity of the crude oil/water interface. Molecular mechanisms
and B suggested differences in the synergies between that could account for the fast formation of an elastic intarface for
components in the crude oils, which were decisive for their the acid-extracted fraction are (i) destabilization and enhanced
interfacial properties. The elastic and viscous moduli as a adsorption of asphaltenes at the interface in the absence of acidic
function of time for both crude oils and their fractions are shown compounds (i.e., the acidic components stabilize asphaltenes in
in Figure 3. the crude oil) and (ii) more densely packed asphaltenes at the
interface when the interfacially active acidic components were
removed. A combination of the above is also possible.
The elastic moduli of all fractions of crude oil B were
independent of time and were lower than the corresponding
fractions of crude oil A (see Figure 3B).
All the fractions had lower elasticity moduli, compared to the
crude oil, and it was lowest for the asphaltenes- extracted fraction.
This demonstrated that the asphaltenes influenced the interfacial
properties of this crude oil as well. However, no synergies
between the asphaltenes and acidic components were seen.
Notably, both the TAN and TBN values were somewhat lower
than the crude oil in this case. This might provide good
stabilization of the asphaltenes in the bulk of the crude oil.
Furthermore, lower average polarity of this fraction,20 in
combination with higher content of aromatics, might provide
better solubility conditions for the asphaltenes, thereby making
them less susceptible to form viscoelastic interfaces.
3.4. Interfacial Strength. The viscoelastic behavior can be
modeled based on Lucassen and Van den Tempel (LVDT)
theory. In this model, the viscoelastic behavior is mostly related
to diffusion of the molecules to and from the interface, rather
than formation of the physical network.23,26 The latter could be
related to non-Laplacian behavior of the drops, which was not
observed during our measurements. The interfacial strengths (S)
and n values of the crude oils, calculated from eq 4, are listed in
Table 3. Crude oil H had a considerably more elastic interface
(i.e., lower n value) than the other oils. This was the only acidic
oil and it is also known to contain tetrameric carboxylic acids.
These acids can form highly rigid interfaces, when calcium is
present in the water phase.34 Generally, the interfacial strength
decreased with increasing contents of aromatics (see Figure 4)
and increasing amounts of acidic compounds (TAN) (see Figure
Figure 3. Dynamic interfacial elasticity (closed symbols) and viscosity 5) in the crude oil.
(open symbols) of (A) crude oil A and its fractions and (B) crude oil B
and its fractions.

The elastic moduli were higher than the viscous moduli for all
the fractions, and this indicates some degree of elasticity for all
the interfaces. The elastic moduli for crude oil A and its base-
extracted fraction were comparable and increased slowly over
time (see Figure 3A). For the acid-extracted fraction, however,
the elastic modulus was significantly higher and increased
markedly over time. Finally, the asphaltene-extracted fraction
had the lowest elastic modulus with no time dependence. This
fraction had lower TAN value, but similar TBN values as the
crude oil (see Table 2). The absence of basic groups in the
asphaltenes could account for the low stabilization of this fraction
by acidic components in the resin fraction. The similarities
between the base-extracted fraction and crude oil showed that
the basic components did not have any major impact on the Figure 4. Dependency of the interfacial strength (S) on the amount of
viscoelastic properties of the crude oil/brine interface. On the aromatics in the crude oils. The solid line is the regression line for the
other hand, synergies between asphaltenes and acidic com- data.

D DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

point A). After stabilization, one drop was approached toward


the other until the separation distance reached zero (indicated by
point B). At this point, the drops were separated by a thin
aqueous film. After some time, a rapid decrease in the pressure
difference was observed between point C and point D. In some
cases, a break point was observed, and the decrease in pressure
difference slowed before reaching a stable value, as shown
between point D and point E in Figure 6. In other cases, this
break point was not observed. Images corresponding to the
various points are also shown in Figure 6, while videos with and
without the retarded decrease between point D and point E are
available in the Supporting Information. The time for the entire
process from point B to point E was referenced as the
coalescence time. This is divided into three characteristic times
for the subprocesses described above: drainage time, rupture
Figure 5. Dependency of the interfacial strength (S) on the total acid
number (TAN) of the crude oils. The solid line is the regression line for
time, and apparent merging time.
the data. Drainage Time (B−C). Drainage time is defined as the time
between the approach of the drops and rupture (rapid change of
the pressure); in this stage, the pressure is relatively constant,
The aromatic character of the asphaltenes makes them soluble since the drainage does not reduce the interfacial area of the
in the aromatic fraction of the oil, and the reduction in strength drops. The time corresponds to the thinning of the aqueous film
was attributed to less aggregation and reduced interfacial activity
between the drops.
of asphaltenes as the aromatics content increased. The
Rupture Time (C−D). The rupture time is characterized by a
weakening of interfacial strength with increasing TAN value
rapid change of the pressure over time, which is attributed to the
suggested synergetic effects between the acidic components and
asphaltenes, as outlined above. Notably, the interfacial strength breakage of the aqueous film between drops that results in the
also decreased with increasing resin contents (see Figure S4 in contact between the oil drops.
the Supporting Information). This suggested that most of the Apparent Merging Time (D-E). The apparent merging time is
acidic compounds were present in the resin fractions of the crude characterized by a slow change of the pressure over time and is
oils. attributed to the flow of crude oil and equilibration process.
3.5. Drop−Drop Coalescence. The differential pressures Notably, this process was too fast to be detected for samples with
across the oil/water interfaces and the distances between the low viscosity.
approaching drops were acquired as a function of time during the Points B, C, D and E were chosen “by eye”.
binary drop coalescence measurements. Table 4 lists the times for the various processes.
Figure 6 shows the differential pressure across an oil drop The coalescence time varied from ∼2 s to 9 s among the crude
(solid line) along with the distance between two drops (dashed oils. The drainage time (i.e., time of film thinning) constituted
line) in a typical experiment. The differential pressure stabilized most of this time, while the rupture of the film occurred faster
when the predefined size of the drop was reached (indicated by than 0.6 s. The apparent merging time of two drops was <2 s.

Figure 6. Differential pressure (solid line) and distance between drops (dashed line), as a function of time for a typical drop−drop coalescence
experiment: stable oil drop (point A), zero gap between the drops (point B), rupture of thin film starts (point C), merging of drops starts (point D), and
coalescence process completed (point E). The images on the right correspond to the various stages on the graph.

E DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

Table 4. Characteristic Times during the Coalescence Process down by elastic interfaces. Note that the strength and the drop
binary coalescence measurements were carried out at different
crude coalescence drainage rupture time apparent merging
oil time (s) time (s) (s) timea (s) time scales. Nevertheless, formation of the viscoelastic layer can
start quickly after drop formation (Figure 2) and it is anticipated
A 5.84 ± 1.55 5.5 ± 1.5 0.34 ± 0.05 ND
that the drainage time would become longer if longer aging was
B 5.65 ± 0.65 4.7 ± 0.4 0.35 ± 0.05 0.6 ± 0.2
possible in the coalescence experiments.
C 4.43 ± 1.31 1.9 ± 1.0 0.63 ± 0.11 1.9 ± 0.2
The rupture time was strongly related to the asphaltene
D
content of the crude oils (see Figure 9). Several mechanisms for
E 9.10 ± 1.30 8.8 ± 1.2 0.30 ± 0.10 ND
F 5.93 ± 2.02 5.6 ± 2.0 0.33 ± 0.02 ND
G 4.84 ± 1.23 3.4 ± 0.6 0.44 ± 0.03 1 ± 0.6
H 5.11 ± 0.91 4.8 ± 0.9 0.31 ± 0.01 ND
I 1.84 ± 0.51 1.0 ± 0.3 0.34 ± 0.01 0.5 ± 0.2
a
ND = ot detectable.

The drainage times tended to decrease with increased density


of the crude oil (see Figure 7). This was attributed to increasing

Figure 9. Dependency of the rupture time on the asphaltene content of


the crude oils. The solid line is the regression line for the data.

the rupture of thin liquid films have been proposed.35 In the so-
called “capillary wave mechanism”, thermally excited fluctuations
of the opposite interfaces come in contact and cause film rupture
at a critical film thickness. It is then reasonable that the rupture
time becomes longer when the asphaltenes content (and,
Figure 7. Dependency of the drainage time on the density of the crude consequently, also the viscosity (see Figure S5 in the Supporting
oils. The solid line is the regression line for the data. Information) increases, since this will lower the flexibility of the
interfaces. Detectable apparent merging times occurred only
when the viscosity of the crude oils was >17 mPa s and increased
Bond numbers when the density difference between the crude oil as the viscosity increased (see Figure 10). This demonstrated that
and aqueous phase increased. The Reynolds regime then became the viscous forces were the most important factor for the rupture
more dominating for the drainage process, and it has been shown and merging processes.
that this causes the film thinning to slow down for drop sizes
above 100 μm.10 It is also known that the interfacial elasticity 4. CONCLUSIONS
slows down the film thinning rate.10 The increased drainage time The relationship between interfacial properties and coalescence
with increasing strength was in agreement with this (Figure 8), of crude oil drops was directly investigated. Asphaltenes was
and this observation suggested that the film thinning was slowed found to be the main factor opposing the coalescence of crude oil

Figure 8. Dependency of the drainage time on the interfacial strength Figure 10. Dependency of the merging time on the viscosity of the crude
(S) of the crude oils. The solid line is the regression line for the data. oils.

F DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

drops. Coalescence was hindered by an elastic oil/water (10) Ivanov, I. B.; Kralchevsky, P. A. Stability of emulsions under
interface, which was associated with the interfacial activity of equilibrium and dynamic conditions. Colloids Surf., A 1997, 128, 155−
asphaltenes. This decreased if the asphaltenes solubility in crude 175.
oil was enhanced by high aromatics content in the oil. (11) Ivanov, I. B.; Danov, K. D.; Kralchevsky, P. A. Flocculation and
Furthermore, the interfacial activity of asphaltenes could be coalescence of micron-size emulsion droplets. Colloids Surf., A 1999,
152, 161−182.
hindered by acidic components from crude oil. These
(12) Kočaŕ ková, H.; Rouyer, F.; Pigeonneau, F. Film drainage of
components tended to stabilize the asphaltenes in the bulk, viscous liquid on top of bare bubble: Influence of the Bond number.
and thus reduced the elastic character of the interface.


Phys. Fluids 2013, 25, 022105.
(13) Maldonado-Valderrama, J.; Martin-Rodriguez, A.; Galvez-Ruiz,
ASSOCIATED CONTENT M. J.; Miller, R.; Langevin, D.; Cabrerizo-Vilchez, M. A. Foams and
*
S Supporting Information emulsions of β-casein examined by interfacial rheology. Colloids Surf., A
Depictions (Figures S1−S5), as well as detailed description of 2008, 323, 116−122.
extraction procedures and videos of coalescence experiments, as (14) Arla, D.; Sinquin, A.; Palermo, T.; Hurtevent, C.; Graciaa, A.;
noted in the text, are provided. This material is available free of Dicharry, C. Influence of pH and Water Content on the Type and
charge via the Internet at http://pubs.acs.org. Stability of Acidic Crude Oil Emulsions. Energy Fuels 2006, 21, 1337−


1342.
AUTHOR INFORMATION (15) Chen, J.-D.; Hahn, P. S.; Slattery, J. C. Coalescence time for a
small drop or bubble at a fluid−fluid interface. AIChE J. 1984, 30, 622−
Corresponding Author 630.
* Tel.: (+47) 73 59 41 35. Fax: (+47) 73 59 40 80. E-mail: gisle. (16) Krebs, T.; Schron, C. G. P. H.; Boom, R. M. Coalescence kinetics
oye@chemeng.ntnu.no. of oil-in-water emulsions studied with microfluidics. Fuel 2013, 106,
Author Contributions 327−334.
The manuscript was written through contributions of all authors. (17) Won, J. Y.; Krägel, J.; Gochev, G.; Ulaganathan, V.; Javadi, A.;
All authors have given approval to the final version of the Makievski, A. V.; Miller, R. Bubble-bubble interaction in aqueous β-
manuscript. lactoglobulin solutions. Food Hydrocolloids 2014, 34, 15−21.
(18) Krebs, T.; Ershov, D.; Schroen, C. G. P. H.; Boom, R. M.
Notes
Coalescence and compression in centrifuged emulsions studied with in
The authors declare no competing financial interest.


situ optical microscopy. Soft Matter 2013, 9, 4026−4035.
(19) Eftekhardadkhah, M.; Øye, G. Induction and coverage times for
ACKNOWLEDGMENTS crude oil droplets spreading on air bubbles. Environ. Sci. Technol. 2013,
The authors are grateful to the industrial sponsors (Con- 47, 14154−14160.
ocoPhillips Skandinavia, ENI Norge, Schlumberger Norge, (20) Gaweł, B.; Eftekhardadkhah, M.; Øye, G. Elemental Composition
PWMS, Statoil Petroleum, and Total E&P Norge) of the joint and Fourier Transform Infrared Spectroscopy Analysis of Crude Oils
industrial program “Produced Water Management: Fundamen- and Their Fractions. Energy Fuels 2014, 28, 997−1003.
tal Understanding of the Fluids” for financial support. Dr. Jens (21) Simon, S.; Nordgard, E.; Bruheim, P.; Sjöblom, J. Determination
Norrman is thanked for proofreading the manuscript. of C-80 tetra-acid content in calcium naphthenate deposits. J.


Chromatogr., A 2008, 1200, 136−143.
REFERENCES (22) Nenningsland, A. L.; Simon, S.; Sjöblom, J. Surface Properties of
Basic Components Extracted from Petroleum Crude Oil. Energy Fuels
(1) Bakke, T.; Klungsøyr, J.; Sanni, S. Environmental impacts of 2010, 24, 6501−6505.
produced water and drilling waste discharges from the Norwegian (23) Rane, J. P.; Pauchard, V.; Couzis, A.; Banerjee, S. Interfacial
offshore petroleum industry. Mar. Environ. Res. 2013, 92, 154−169. Rheology of Asphaltenes at Oil-Water Interfaces and Interpretation of
(2) Lee, K.; Neff, J.; DeBlois, E. Produced Water; Springer: New York, the Equation of State. Langmuir 2013, 29, 4750−4759.
2011; pp 3−54. (24) Won, J. Y.; Krägel, J.; Makievski, A. V.; Javadi, A.; Gochev, G.;
(3) Fakhru’l-Razi, A.; Pendashteh, A.; Abdullah, L. C.; Biak, D. R. A.; Loglio, G.; Pandolfini, P.; Leser, M. E.; Gehin-Delval, C.; Miller, R. Drop
Madaeni, S. S.; Abidin, Z. Z. Review of technologies for oil and gas
and bubble micro manipulator (DBMM)A unique tool for mimicking
produced water treatment. J. Hazard. Mater. 2009, 170, 530−551.
processes in foams and emulsions. Colloids Surf., A 2014, 441, 807−814.
(4) Das, P. C. Selection of technology for produced water treatment.
(25) Farooq, U.; Simon, S.; Tweheyo, M. T.; Øye, G.; Sjoblom, J.
SPE J. 2012, 1122−1128.
Interfacial Tension Measurements Between Oil Fractions of a Crude Oil
(5) Puprasert, C.; Hebrard, G.; Lopez, L.; Aurelle, Y. Potential of using
Hydrocyclone and Hydrocyclone equipped with Grit pot as a pre- and Aqueous Solutions with Different Ionic Composition and pH. J.
treatment in run-off water treatment. Chem. Eng. Process. 2004, 43, 67− Dispersion Sci. Technol. 2013, 34, 701−708.
83. (26) Rane, J. P.; Harbottle, D.; Pauchard, V.; Couzis, A.; Banerjee, S.
(6) Knudsen, B. L.; Hjelsvold, M.; Frost, T. K.; Svarstad, M. B. E.; Adsorption Kinetics of Asphaltenes at the Oil−Water Interface and
Grini, P. G.; Willumsen, C. F.; Torvik, H. Meeting the zero discharge Nanoaggregation in the Bulk. Langmuir 2012, 28, 9986−9995.
challenge for produced water. In SPE International Conference on Health, (27) Poteau, S.; Argillier, J.-F.; Langevin, D.; Pincet, F.; Perez, E.
Safety and Environment in Oil and Gas Exploration and Production, Influence of pH on Stability and Dynamic Properties of Asphaltenes and
Society of Petroleum Engineers: Calgary, Alberta, Canada, March Other Amphiphilic Molecules at the Oil−Water Interface. Energy Fuels
29−31, 2004; pp 525−530. 2005, 19, 1337−1341.
(7) Veil, J. A. Produced Water; Springer: New York, 2011; pp 3−54. (28) Varadaraj, R.; Brons, C. Molecular origins of heavy crude oil
(8) Hemmingsen, P. V.; Silset, A.; Hannisdal, A.; Sjöblom, J. Emulsions interfacial activity part 2: Fundamental interfacial properties of model
of heavy crude oils. I: Influence of viscosity, temperature, and dilution. J. naphthenic acids and naphthenic acids separated from heavy crude oils.
Dispersion Sci. Technol. 2005, 26, 615−627. Energy Fuels 2007, 21, 199−204.
(9) Silset, A.; Flaten, G. R.; Helness, H.; Melin, E.; Øye, G.; Sjöblom, J. (29) Brandal, O.; Sjöblom, J.; Øye, G. Interfacial behavior of
A Multivariate Analysis on the Influence of Indigenous Crude Oil naphthenic acids and multivalent cations in systems with oil and
Components on the Quality of Produced Water. Comparison Between water. I. A pendant drop study of interactions between n-dodlecyl
Bench and Rig Scale Experiments. J. Dispersion Sci. Technol. 2010, 31, benzoic acid and divalent cations. J. Dispersion Sci. Technol. 2004, 25,
392−408. 367−374.

G DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

(30) Pauchard, V.; Sjöblom, J.; Kokal, S.; Bouriat, P.; Dicharry, C.;
Muller, H.; Al-Hajji, A. Role of Naphthenic Acids in Emulsion Tightness
for a Low-Total-Acid-Number (TAN)/High-Asphaltenes Oil. Energy
Fuels 2008, 23, 1269−1279.
(31) Magnusson, H. n.; Hanneseth, A. M. D.; Sjöblom, J. Character-
ization of C80 Naphthenic Acid and Its Calcium Naphthenate. J.
Dispersion Sci. Technol. 2008, 29, 464−473.
(32) Freer, E. M.; Radke, C. J. Relaxation of asphaltenes at the toluene/
water interface: diffusion exchange and surface rearrangement. J. Adhes.
2004, 80, 481−496.
(33) Mullins, O. C. The Modified Yen Model. Energy Fuels 2010, 24,
2179−2207.
(34) Nordgård, E. L.; Magnusson, H.; Hanneseth, A.-M. D.; Sjöblom, J.
Model compounds for C80 isoprenoid tetraacids: Part II. Interfacial
reactions, physicochemical properties and comparison with indigenous
tetraacids. Colloids Surf., A 2009, 340, 99−108.
(35) Ivan, B. I.; Krassimir, D. D.; Peter, A. K. Encyclopedic Handbook of
Emulsion Technology; CRC Press: Boca Raton, FL, 2001; pp 621−659.

H DOI: 10.1021/ef501847q
Energy Fuels XXXX, XXX, XXX−XXX

You might also like