You are on page 1of 11

CIS-01585; No of Pages 11

Advances in Colloid and Interface Science xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science

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

Historical Perspective

Interfacial behavior of asphaltenes


Dominique Langevin a,⁎, Jean-François Argillier b
a
Laboratoire de Physique des Solides, Université Paris-Sud, CNRS, Université Paris Saclay, Bâtiment 510, 91405 Orsay Cedex, France
b
IFP Énergies nouvelles, 1-4 Avenue de Bois Préau, 92852 Rueil-Malmaison Cedex, France

a r t i c l e i n f o a b s t r a c t

Available online xxxx We review the existing literature on asphaltenes at various types of interfaces: oil–water, air–water, gas–oil and
solid–liquid, with more emphasis on the oil–water interfaces. We address the role of asphaltene aggregation,
Keywords: recently clarified for asphaltenes in bulk by the Yen–Mullins model. We discuss the questions of adsorption re-
Asphaltenes versibility and interfacial rheology, especially in connection with emulsion stability.
Resins © 2015 Elsevier B.V. All rights reserved.
Emulsions
Interfacial tension
Interfacial rheology

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Asphaltenes in bulk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2.1. Molecular asphaltenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2.2. Nano-aggregates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2.3. Clusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2.4. Precipitation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2.5. Influence of asphaltene extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Asphaltenes at oil–water interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.1. Asphaltenes in good solvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.2. Asphaltenes in poor solvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.3. Influence of salinity and pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.4. Behavior with added surfactants. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.5. Water in oil emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3.6. Demulsification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Asphaltenes at other type of interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.1. Asphaltenes at air–water surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.2. Asphaltenes at oil-gas surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.3. Asphaltenes at solid surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction The amount of asphaltenes in crude oil varies considerably, from


negligible amounts in volatile oils, small amounts in condensates, a
The name crude oil is used for naturally occurring and unprocessed few weight percent in fluid black oils and tens of weight percent in
petroleum. Crude oils are mixtures of many components, hydrocarbons heavy crude oils and bitumen. Their amount essentially determines
(mostly alkanes, cycloalkanes and aromatics) and other organic com- the oil viscosity which impacts oil recovery, transportation and treat-
pounds, containing nitrogen, oxygen, sulfur and traces of metals. ment in refineries. Their solubility in non-polar hydrocarbons is limited;
Asphaltenes are the components of crude oils with the highest molecular they can precipitate when temperature is lowered and be a serious
weight [1]. They are also the most polar and are made of polycyclic aro- problem during exploitation, storage and processing. Asphaltenes
matic hydrocarbon rings with peripheral alkane chains (Fig. 1). adsorb on solid surfaces and can alter the wettability of oil reservoirs,
hence also affect in this way the oil recovery process. They also adsorb
⁎ Corresponding author. at the oil–air interfaces and oil–water interfaces forming very stable

http://dx.doi.org/10.1016/j.cis.2015.10.005
0001-8686/© 2015 Elsevier B.V. All rights reserved.

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
2 D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx

Fig. 1. Typical asphaltene molecule (left) and common heteroatoms (right).

foams and emulsions, which are difficult to break afterwards. (about 1 wt.%) in carboxylic groups and phenols. Metals can be also
Asphaltenes are usually removed from the oil during upgrading pro- present in smaller amounts.
cesses and eventually hydrogenated. However, bitumen or Non-aqueous potentiometric titrations of asphaltenes suggest the
asphaltene-rich residues from refining processes are used in paving presence of acidic and basic functionalities [3]. When contacted with
and coating applications, taking advantage of the large temperature de- water, some asphaltene molecules can ionize and transfer into water.
pendence of their viscosity. Such applications were in use since the an- The transfer is small at low pH, but becomes important at high pH.
cient times. Asphaltenes are also found in coal and tar sands, which are This corresponds to partition coefficients between oil and water of 105
processed using mining methods. for acidic asphaltenes and 107 for basic ones [4].
Asphaltenes are usually defined as the oil fraction precipitated by It is now admitted that asphaltenes are small molecules, with an av-
addition of paraffinic solvents, usually pentane or heptane. Being a erage molecular weight of 750 Da. They contain mainly one polycyclic
solubility class, they have a wide range of chemical formulas. Other aromatic group made of about 7 rings in average, to which linear or
crude oil constituents of smaller polarity and molecular weight, resins, branched hydrocarbon chains are attached. Atomic force and scanning
are also surface active. They are obtained first by evaporating the lighter tunneling microscopy recently provided images of asphaltene mole-
oil components from the oil fraction not precipitated by addition of hep- cules with atomic resolution that confirmed the extreme variety of
tane. This fraction, called maltene, contains resins and other nonvolatile molecular structures (Fig. 2) [5].
constituents. The resins are extracted by liquid chromatography. The larger molecular weights currently reported in the literature
Asphaltenes also self-associate into different type of aggregates and arise from self-aggregation. It is to be noted that aggregation studies
their exact structure remained unclear for many years. Less than were made mostly in toluene, which is a good solvent for asphaltenes.
20 years ago, the reported measured molecular weights ranged from The state of aggregation in less good solvents remains less well
less than 1000 up to 109 Da. It was only very recently that a consensus known. The following discussion will therefore concern mostly
was reached with the Yen–Mullins model and that a better description asphaltene in toluene solutions.
of asphaltene molecules became available [2]. There are however still
many debates on how asphaltenes adsorb at surfaces. In this review, 2.2. Nano-aggregates
we will summarize the recent advances on petroleum asphaltenes.
We will mostly discuss oil–water interfaces, the topic which is the Asphaltenes self-assemble at very small concentrations. Above a
most active, but also review briefly the work on air–water, gas–oil and critical concentration of order 100 mg/L, called CNAC by Mullins, the
solid–oil interfaces. In many studies, asphaltenes are extracted and aggregation number remains small (b 10) and constant. This behavior
resolubilized in model solvents. The results are frequently compared is typical of aggregation of surfactant molecules in organic solvents
to measurements performed with the original crude oil or its dilutions which is progressive, starting with the formation of dimers, then trimers
when the viscosity is too high. In the following, we will mainly describe and following up with multimers. The formation of surfactant micelles
studies made with model asphaltenes solutions. in water occurs sharply because the free energy gained by a molecule
during incorporation in a micelle is typically 15 times the thermal ener-
2. Asphaltenes in bulk gy kBT (kB Boltzmann constant, T absolute temperature) while for
micelles in oil the free energy gain is only a few kBT [6]. The asphaltene
Let us first briefly summarize the description of asphaltenes nano-aggregates are formed due to the attractive interactions between
proposed in the Yen–Mullins model. the pi electrons of the rings. Because there is no X-ray signature of reg-
ular stacking of the polycyclic groups, the interior of the aggregate is
2.1. Molecular asphaltenes likely disordered (Fig. 3A).
The peripheral chains produce steric repulsion and once a small
The carbon of petroleum asphaltenes is about 50% aromatic. number of molecules are associated, it becomes difficult for a new one
Asphaltenes are the oil component with the lowest H/C ratio, due to to penetrate the aromatic core of the aggregate. Simulations confirmed
the presence of polycyclic aromatic groups (Fig. 1). They also contain this assumption, predicting small aggregation numbers as observed in
heteroatoms: sulfur, typically 5–8 wt.%, essentially present in the rings experiments.
in thiophenic groups; there are smaller amounts of nitrogen (about Combined X-ray and neutron scattering experiments provided a
1.5 wt.%) in the rings as pyrrolic and pyridinic groups and oxygen more precise picture: the nanoaggregates were shown to be disks of

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx 3

Fig. 2. Example of molecular structure of an asphaltene molecule according to Ref. [5].

radius 3.2 nm and height of 0.67 nm; the core is a dense aromatic struc- unlikely. Molecular simulations could confirm if the casein type of clus-
ture, contrary to the shell, which is aliphatic [7]. tering is valid for asphaltenes (or not).
When the solvent quality is decreased, for instance by adding an al-
2.3. Clusters kane to toluene [14] or lowering the temperature [15], similar clusters
are formed, their size being only slightly larger: for instance their
At higher concentrations of asphaltenes, several g/L, larger aggre- mean radius is 14 nm in 45% heptane in toluene instead of 7 nm in
gates start forming. These aggregates remain however small pure toluene (the precipitation occurs above 56% heptane). It was
(radius ≲ 10 nm). Their dimensions are well defined and rather inde- shown that the clusters are swollen preferentially by toluene,
pendent of the type of asphaltenes. The binding energy is smaller explaining why their size is comparable to those observed in pure tolu-
than for nanoaggregates, and they are more strongly impacted by ene [16]. Asphaltene clusters in toluene and in crude oil are similar. This
external parameters such as temperature: this is likely why the vis- excludes the formerly well accepted idea of asphaltene aggregates steri-
cosity of asphaltenes solutions depends so strongly on temperature. cally stabilized by resins, other less crude oil components. The increased
The shape of these clusters has been the object of much debate: ves- solubility of asphaltenes in the presence of resins is therefore a mere sol-
icles [8], disks [9], and fractal assemblies of nanoaggregates [10] vent effect [8].
(Fig. 3B). The last hypothesis appears to be sounder, since it is able
to account for viscosity, X-ray and neutron scattering experiments 2.4. Precipitation
together. The clusters contain about 20 solvated nanoaggregates. A
question, yet unanswered, then arises: what mechanism limits the Asphaltenes become insoluble in less aromatic solvents. In this case,
cluster size? larger aggregates (flocs) rapidly appear and grow until precipitation is
Although asphaltenes are small molecules, their behavior at liquid completed. Asphaltene precipitates exhibit a multilayer structure, with
interfaces has striking similarities with that of proteins, which are mac- the polycyclic groups stacked rather regularly: at the difference of the
romolecules usually folded into globular structures [11]. Proteins such nanoaggregates, a quasi-Bragg peak is seen in X-ray experiments [8].
as casein form aggregates improperly called micelles, present in a dis- It is known that for a given asphaltene, precipitation occurs when the
persed state in milk, which also have a rather constant size (radius refractive index of the solvent falls below a given limit [17], suggesting
~100 nm) [12]. It is well admitted now that the more hydrophobic ca- that precipitation is due to non-specific van der Waals attraction.
seins locate in the micelle center, while the more hydrophilic ones are Because precipitation is most often reversible, this attraction is weak.
closer to the periphery. It is possible that a similar organization leads The reversibility also excludes the idea of asphaltene aggregates steri-
to the well-defined size in asphaltene large aggregates, with the more cally stabilized by resins. Indeed, precipitation would be irreversible in
polar asphaltene nano-aggregates in the center and the less polar ones this case [8].
at the periphery. Another option is electrostatic repulsion: when
charged entities cluster, the total charge increases and an optimal size 2.5. Influence of asphaltene extraction
is found [13]. However, the asphaltene charge in oil is very small (one
molecule charged in 105 in toluene (2)), and the average charge of the Asphaltenes being a solubility class, their molecular composition ap-
clusters is much less than one. The charge mechanism is therefore preciably depends on the extraction method [18]. Fig. 4 shows the visual
aspect of asphaltenes extracted using the most commonly used
solvents: pentane and heptane. One can see on the figure that the tex-
ture and even the color of these asphaltenes are different. The
asphaltenes extracted with pentane (C5) contain in addition molecules
that are soluble in heptane, such as resins.
To our opinion, this is the reason why C5 asphaltenes better
mimic heavy crude oils which are good solvents for asphaltenes. C5
asphaltenes contain more resins than the asphaltenes extracted
with heptane (C7) and precipitate less easily. Conversely, the C7
asphaltenes better mimic the so-called asphaltenic crude oils,
which are poor solvents for asphaltenes (also present in smaller
quantities).
Successive fractionations showed that the asphaltenes less soluble in
alkanes were more polar, their molecular weight was higher and they
contained more aromatic groups [18]. Neutron studies showed that
they formed larger clusters [20]. Similar results were found comparing
asphaltenes precipitated with different alkanes [10] or separated by
Fig. 3. Asphaltene aggregates (after Ref. [2]). centrifugation [21].

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
4 D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx

Fig. 4. Asphaltenes precipitated from a crude oil by using: A) n-pentane and B) n-heptane. Source: Ref. [19].

Not only asphaltenes depend on the crude oil from which they are exponential variation is expected for a interfacial reorganization with
extracted, but significant differences are found if different extraction a single characteristic time τ, as in the Maxwell model. The interfacial
methods are used. Care should therefore be taken when comparing tension varies more as a superposition of exponential decays, in partic-
results from different studies. ular with times much longer than τ. In the absence of an established re-
laxation model, the use of a single time can however allow viewing the
3. Asphaltenes at oil–water interfaces main trends.
It was argued that the long times arise from the fact that not all the
Because they are the more polar components of crude oil, asphaltenes adsorb at the interface [25]. Dabros and coworkers estimat-
asphaltenes adsorb at oil–water interfaces. The adsorption process is ed that no more than 0.4% of the asphaltenes could be held responsible
extremely slow, even at high bulk concentrations. for Athabasca bitumen emulsions stabilization ability [26]. Experiments
were done later with emulsions made with asphaltene solutions in mix-
3.1. Asphaltenes in good solvents tures of toluene and heptane. These emulsion experiments showed that
all the soluble asphaltenes extracted from this bitumen were surface ac-
Only few experiments were done below the CNAC with solutions of tive (the asphaltene concentration in the bulk solvent is nearly zero,
asphaltenes in toluene. Freer and Radke showed using pendent drop ex- provided the insoluble asphaltenes have been removed prior to emulsi-
periments and rinsing procedures that only a small fraction of fication) [27]. We will see in Section 3.3 that precipitated asphaltenes do
asphaltenes was adsorbed reversibly [22]. In addition, they measured not contribute to emulsion stabilization: the state of solubilization of
the interfacial compression elastic and loss moduli (also sometimes asphaltene in bitumen could therefore account for the low quantity of
called dilatational or dilational moduli). They showed that they could
fit the data with a superposition of the response of asphaltenes irrevers- 3000
ibly adsorbed (Maxwell model for the internal reorganization in the in-
terfacial layer) and reversibly adsorbed (Lucassen–van den Tempel
(LvdT) model). As reported early by Strassner [23] and later by many 2500
other groups, Freer and Radke mentioned the formation of wrinkles
upon compression (frequently called skin) that disappear after about a
characteristic time (sec)

few minutes. 2000 toluene


Adsorption times are still very long for more concentrated solutions, heptol
heptol+maltenes
of the order of 1000 s for 1 wt.% asphaltenes in toluene [24]. For a diffu-
sion controlled adsorption process, the interfacial concentration Γ 1500
initially varies with time t as:
rffiffiffiffi
D pffiffi 1000
Γ¼2 C t ð1Þ
π

C being the bulk concentration and D the diffusion coefficient of 500


asphaltene molecules. This equation can be used to estimate a charac-
teristic adsorption time τ. Using D = 3 10−10 m2/s and Γ = 4 mg/m2
leads to τ = 1 ms, much shorter than the experimental characteristic 0
time of 1000 ms. 0 5 10 15 20
Furthermore, the relaxation times in the interfacial layer remain Asphaltene weight percentage (%)
long as bulk concentration increases despite that diffusion times
becomes much shorter. This is illustrated in Fig. 5, where characteristic Fig. 5. Characteristic adsorption times for asphaltenes at the oil–water interface. Heptol is
a mixture of 50% toluene and 50% heptane. In the solvent heptol + maltenes, heptol per-
adsorption times are plotted versus asphaltene concentration. centage is 90% and maltenes + asphaltenes percentage is 10%. Redrawn from Ref. [24] (for
The relaxation times in Fig. 5 have been obtained by a simple expo- interpretation of the references to color in this figure legend, the reader is referred to the
nential fit of the time variation of the interfacial tension [24]. An web version of the article).

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx 5

surface-active asphaltenes in the experiments of Ref. [25] and [26]. observed in insoluble monolayers of proteins [38] or of nanoparticles
More recent emulsion experiments showed that as much of 98% [39].
asphaltenes could be removed without affecting emulsion stability [28]. Interfacial shear rheological investigations were also performed [40].
Even assuming that only 2% of asphaltenes adsorb at the interface in The shear modulus is very small and non-measurable during the first
the estimation above, the diffusion time will only be 2.5 s instead of hours. It increases afterwards, becoming larger than the loss modulus,
1000 s. In order to match with the experimentally measured time, one indicating that the interfacial layer became solid. Interestingly, the ap-
should assume that only a fraction of 10− 3 of the total amount of pearance of the skin also requires aging and could be correlated to the
asphaltenes is surface active, which is not consistent with Ref. [28]. Fur- existence of a sufficiently large shear modulus. Indeed, in protein or
thermore, if adsorption was diffusion controlled, one could not explain nanoparticle monolayers, the skins are formed when the interfacial
why the adsorption time decreases only slowly with increasing layers are solid.
asphaltene concentration.
The thickness of the oil–water interfacial layer has been measured 3.2. Asphaltenes in poor solvents
with emulsions exploiting the contrast matching procedure offered by
neutron scattering techniques. These emulsions were prepared from Except for heavy oils, crude oils are not good solvents for asphaltenes,
3 g/L asphaltene in xylene solutions, and contained large clusters. Xy- due to the presence of appreciable amounts of alkanes. In order to mimic
lene is a good solvent for asphaltenes, very similar to toluene but safer the crude oil, many studies were performed using mixtures of toluene
in terms of health hazards, and increasingly used in experiments. It and heptane, called heptol. In the following heptol will be used for the
was found that the interfacial layer thickness was similar to the diame- mixture of equal volumes of each solvent, unless stated otherwise.
ter of the asphaltene clusters (7 nm, twice their gyration radius) [29,30]. Pauchard and coworkers used very dilute asphaltene solutions in
The measured thicknesses were similar to those measured by Masliyah mixtures of toluene and viscous polyolefin oils [41,42]. They showed
and coworkers for the thin films separating two water drops in toluene, that the initial time decay of the interfacial tension is diffusion con-
8.5 nm [31]. These results suggest that the clusters are adsorbed as a trolled. Adsorption slows down afterwards and can be modeled with a
whole and that the interface is no longer in a situation of monolayer random sequential adsorption [43]. Using fast compression–expansion
coverage. The adsorbed amounts are larger than for monolayer cover- cycles, they showed that at the difference of equilibration between in-
age, of order 5 mg/m2. They are slightly smaller than those measured terface and bulk, which is very slow, the equilibration in the interfacial
by gravimetric methods by Yarranton and coworkers, probably because layer is rapid. They fitted the measured variation of the interfacial ten-
they used a poorer solvent [32]. Interestingly, the size of the clusters re- sion γ with the surface coverage Γ using a Langmuir equation of state,
maining in the oil phase is slightly smaller than the initial size, meaning leading to a limit surface concentration Γ∞ = 3.2 mg/m2. Such a value
that the adsorbed clusters are the largest. This is as expected, as the is consistent with monolayer coverage and with the polycyclic group
more polar asphaltenes assemble in larger clusters [21]. sitting flat at the interface. Pauchard and coworkers also performed
Note that the picture of Fig. 3 suggests that the clusters are hydro- emulsion studies and showed that stability is achieved once emulsion
phobic, but this is a cartoon drawn for clusters in bulk oil. During clus- drops are covered by a compact asphaltene monolayer (Γ = Γ∞) [42].
ters adsorption, they likely reorganize, as water soluble proteins do Emulsions studies at low asphaltene concentration in heptol were
when they adsorb to an air–water or oil–water interface [33]. This reor- also performed by Yarranton and coworkers and surface coverages
ganization could also affect the interfacial tension and explain why the were estimated by gravimetric methods [27]. These authors found a
adsorption times are not diffusion controlled. Adsorption barriers similar surface coverage (around 3.5 mg/m2). They also showed that
could also be present and slow down adsorption as well. most of the soluble asphaltenes adsorb at the oil–water interface
Interfacial compression rheological determinations were made for
these interfaces [34,35]. It was shown that the compression elastic mod-
ulus E increases with time and reaches equilibrium values after times
much longer than the interfacial tension, typically one or two days.
The modulus also increases with asphaltene concentration up to a max-
imum in the concentration region where nano-aggregates are present
(~0.1–10 g/L) and decreases afterwards. This behavior is as predicted
by the interface-bulk exchanges controlled by diffusion (LvdT model),
but the concentration of the maximum is incompatible with calculated
values that are much smaller. We have seen earlier that adsorption
was essentially irreversible and it is not surprising to see that the LvdT
model cannot account for the maximum observed.
Yarranton and coworkers used water micro-droplets and measured
the interfacial tension changes upon compressing an oil drop. They ob-
served after surface aging over one hour a sharp decrease in compress-
ibility of the interfacial layer [36]. They proposed that this was due to a
structural transition in the adsorbed layer. Upon further compression,
crumpling of the layer was seen: the drop surface becomes rough be-
cause of an apparent buckling. This effect was attributed to the presence
of a rigid skin at the surface, already mentioned earlier. Crumpling oc-
curs above a certain reduction of area, called crumpling point, with
the relative area reduction at this point called crumpling ratio (Fig. 6).
Alvarez and coworkers reported measurements of the crumpling
ratio at long times [35]. This ratio correlates well with the compression
modulus as found by Yarranton and coworkers. At short aging times, the
crumpling ratios are lower, as also seen in their experiments. Yarranton
and coworkers propose that at low aging time, asphaltenes are revers-
ibly or nearly reversibly adsorbed but that at longer times, adsorption Fig. 6. Water drop in diluted bitumen: A) initial shape; B) after compression by aspiration
becomes progressively irreversible. Similar phenomena are currently in the pipette. From Ref. [37].

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
6 D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx

whereas precipitated asphaltenes do not adsorb at all. At larger As for experiments with toluene solutions, the compression elastic
asphaltene concentrations, the surface coverage increases up to modulus has been investigated with oscillating drop and Langmuir
10 mg/m2 at concentrations of 40 g/L, well above monolayer coverage trough experiments, as well as with droplets held on micropipettes.
[32]. The surface coverages are somewhat lower for better solvents Yarranton and coworkers showed that the micropipette technique
and depend slightly on the type of asphaltene. leads to comparable surface pressure curves than in Langmuir troughs
Langmuir trough studies were also performed by several groups. In and comparable compression elastic moduli than with oscillating drop
these experiments, the asphaltenes are spread onto the water interface, methods [36]. Upon aging and as in toluene, a high compressibility
generally using a concentrated solution of asphaltenes in toluene. After phase progressively transform into a low compressibility one. After
evaporation of the toluene, the water is covered by pure solvents, gen- one hour in heptol (instead of a few days in toluene), the surface pres-
erally heptol. This is equivalent in principle to a interface between sure curve exhibit a transition between a high compressibility phase
water and a very dilute asphaltene solution in heptol. However, in at large surface area and a low compressibility phase at small surface
view of the irreversibility of the adsorption, it is not obvious that the area. When the area is decreased further, crumpling is observed. This
clusters deposited at the interface transform into asphaltene mono- behavior is remarkably similar to that of colloidal particles at surfaces,
layers. Masliyah and coworkers conducted tests to confirm that for which it has been postulated that the intermediate transition was
asphaltene molecules remain at the interface and do not migrate in a jamming transition [46]
the bulk phases, at least within the time frame of the experiments (~1 Kilpatrick and coworkers showed that both compression and shear
hour) [44]. They also found very long interfacial tension evolutions elastic moduli reach equilibrium values after long times, typically
(~ 1 hour), consistent with internal reorganization processes in the 1 day [47,48]. The shear modulus is more variable with the type of
layer. The asphaltene layers were found more compressible at the asphaltene, the more polar asphaltenes form more rigid layers and the
heptol-water interface than at the air–water interface: the larger the corresponding emulsions are more stable. They also showed that
amount of toluene in heptol, the larger the compressibility. The absence asphaltenes are mostly irreversibly adsorbed using washing proce-
of dissolution of asphaltene in bulk heptol showed that asphaltene ad- dures: the shear modulus continues to increase with time many hours
sorption is irreversible, at least at concentrations of 2 g/L in toluene, after replacement of the asphaltene solution by pure solvent. The equi-
the spreading solvent. Note that proteins behave in a similar way: librium value of the modulus is however smaller is the washing proce-
they can also be spread at liquid surfaces, and although they are water dure is performed early. Adsorption is thus only partially irreversible
soluble, they do not migrate toward the water phase. at short times and becomes more irreversible as the layer ages, as also
The characteristic decay time τ of the interfacial tension between found by Yarranton and coworkers [36].
water and asphaltene solutions in heptol is shown in Fig. 5. The time τ The shear elastic modulus being larger than the shear loss modulus,
increases with bulk asphaltene concentration, while it decreases for these interfaces are solid-like. They were recently shown to behave as
toluene solutions. In heptol solutions, asphaltenes are close to the pre- soft glasses [49]. The shear elastic modulus is large, much larger than
cipitation limit and the surfaces should be denser, hence need more for asphaltenes in toluene [40,48]. At the difference of interfacial ten-
time to reorganize. Interestingly, adding 10% maltenes give times τ sim- sion, aging times are smaller than for toluene solutions. Crumpling is ob-
ilar to those observed in toluene, probably because the asphaltenes are served by compressing the surface at smaller compression ratio [36].
better solubilized. These results suggest that toluene might be a better Compression rheological experiments were done in cyclohexane so-
model for crude oil (that contains resins) than usually assumed. It lutions, a solvent intermediate between toluene and heptane for
should be recalled that τ describes only the initial evolution of the inter- asphaltenes [50]. These experiments cannot prove that the interface is
facial tension that can be much longer. Fig. 7 shows results obtained solid as shear rheological experiments do. They can however evidence
with small droplets held on micropipettes were droplets can be aged features compatible with solid state, for instance using the time fre-
in situ for very long periods of time [45]. One sees that the equilibrium quency superposition method. It was shown in this way that the layer
value of the interfacial tension does not depend ultimately on bitumen becomes glassy-like below a temperature of 30 °C [50].
concentration within experimental error.
The bitumen used in the experiments of Fig. 6 contains 17–20% 3.3. Influence of salinity and pH
asphaltenes, 0.001% bitumen would then correspond to 2 mg/L
asphaltenes, well below the CNAC. This example shows that the deter- The water present in oil reservoirs, or injected for recovery purposes,
mination of CNAC or other aggregation concentrations is not possible frequently contains salt. The influence of salt on interfacial properties is
using interfacial tension data with this type of dilution. therefore of interest. Significant variations of the interfacial tension are
observed only at high salt concentrations (~1 M) [51] and are likely due
to the screening of the charges of ionized groups at the interface.
The time dependence of the compression and shear moduli however
is strongly affected by salt [16]. Indeed the interfacial elasticity is more
affected by surface rearrangements and consolidation than the interfa-
cial tension, again, as in protein solutions. The salt effect has been attrib-
uted to the fact that asphaltenes at oil–water interfaces possess both
positive and negative charges resulting in attractive interactions pro-
moting rapid consolidation. Salt screens these attractive interactions
and slows down consolidation.
The influence of pH is more significant. In the case of toluene solu-
tions of asphaltenes, the time τ and the interfacial tension are maximum
at neutral pH, while the elastic compression modulus and the emulsion
stability are minimum (Fig. 8) [52].
This is correlated with a maximum in adsorbed amount and a mini-
mum in interfacial thickness [29]. This means that the more solvated the
clusters, the better the emulsion stability. Note that the opposite is seen
when replacing toluene by heptol at neutral pH: the clusters are less sol-
vated but the emulsion stability is higher. In fact the emulsion stability
Fig. 7. Interfacial tension between water and dilutions of bitumen in heptol. After Ref. [45]. seems more related to the interfacial shear modulus in this case. The

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx 7

pH dependence of the shear modulus has not been investigated to date. 12


5% Asphaltenes in Xylene
It is however likely that G′ will be zero at high pH, since skins do not 11 5% Asphaltenes in Xylene/0.1% SDBS
form in this case [23,52] Diluted crude oil
One sees that the interfacial tension becomes very low at high pH. 10 Diluted crude oil/0.1% SDBS
This is associated to the transfer into the water phase of oil constituents
9
able to ionize. This transfer has been quantified by measuring the pH of
the water phase after equilibration with an oil phase called pHf as a 8
function of the initial pH called pHi [53]. An example is shown in Fig. 9
7

pHf
[4].
One sees in Fig. 9 that the behavior of oil is similar to that of 6
asphaltenes solutions, meaning that the species transferred are likely
5
asphaltenes (here extracted with pentane). Fig. 9 shows that pHf N pHi
for acidic water, corresponding to a transfer of few basic oil species in 4
water [54]. The transfer is much more important at high pH, as con-
firmed by visual observation of water phases [55]. The species trans- 3
ferred at high pH could be carboxylic or naphtenic acids. The 2
important transfer suggests that asphaltenes (and crude oil) contain 2 3 4 5 6 7 8 9 10 11 12
larger quantities of acidic species than of basic ones. Because some of pHi
the basic species will remain at the interface, the interfacial tension is
lower at high pH (Fig. 7). Fig. 9. Final pH versus initial pH of the water phase after equilibration with diluted crude
Note that the salt effect on interfacial tension is significant at low and oil and model oil (5% asphaltenes in toluene), in the presence or absence of added surfac-
high pH, consistent with a pH dependent surface charge of adsorbed tant. After Ref. [4] (for interpretation of the references to color in this figure legend, the
reader is referred to the web version of the article).
asphaltenes [16].

3.4. Behavior with added surfactants were expelled from the interface, the equilibration times were very
long, much longer than without the asphaltenes [24]. It is therefore like-
Surfactants can be used to favor the formation of oil in water emul- ly that a duplex layer is formed, with a layer of surfactant on the top of a
sions for transportation of extra heavy oil purposes. Indeed, oil in water layer of asphaltenes.
emulsions are much less viscous than the water in oil emulsions when Other surfactants were studied, sodium dodecyl sulfate (SDS) and
the oil contains large quantities of asphaltenes [56]. Surfactants can sodium dodecyl benzene sulfonate (SDBS) [59]. SDS and Triton X-405
also be injected in oil reservoirs (chemical enhanced oil recovery) to im- gave similar results, but SDBS led to very small interfacial tensions at
prove oil recovery: they decrease the oil–water interfacial tension and acid and basic pH, a phenomenon associated with a more important
the capillary pressures, improving oil mobilization. transfer of oil species in water (Fig. 9) [54,55]. At alkaline pHs, the inter-
Early experiments by Strassner showed that synthetic surfactants facial tension first decreases with time and after reaching a minimum,
reduce the oil–water interfacial tension to a much greater extent that increases again before reaching an equilibrium value. This behavior is
the natural surfactants such as asphaltenes and resins at intermediate the consequence of the transfer of asphaltenes into water revealed by
pH values [23]. Synthetic surfactants can therefore displace asphaltenes Fig. 8. The existence of the interfacial tension minimum also evidence
and resins from the interface as they do for proteins [57]. However, al- the existence of a desorption barrier [60].
though synthetic surfactants significantly reduce or suppress skin for-
mation, the emulsion stability is much better in the presence of the 3.5. Water in oil emulsions
natural surfactants [58]. Experiments done with asphaltene solutions
in toluene and aqueous solutions of a typical surfactant, Triton X405, Asphaltenes, fine solids and/or waxes are the main stabilizers of
showed that the oil–water interface tension was lower than without water in petroleum emulsions [32]. Asphaltenes stabilize emulsions to
surfactant and equal to the interfacial tension between pure toluene the greatest extent when they are close or at the point of precipitation
and surfactant solution. Although this might suggest that asphaltenes after which stabilization decreases. The stability also decreases when

30 30
compress elastic modus (mN/m)

25
interfacial tension (mN/m)

25
1 g/L asphaltenes

20 1 g/L asphaltenes 20

15 15

10 100 mg/L asphaltenes 10

100 mg/L asphaltenes


5 5
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
pH pH

Fig. 8. Interfacial tension and interfacial compression modulus as a function of pH for water–toluene asphaltenes solutions containing two different concentrations of asphaltenes, 0.1 and
1 g/L after a waiting time of 5400 s. Redrawn from Ref. [52] (for interpretation of the references to color in this figure legend, the reader is referred to the web version of the article).

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
8 D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx

resins are added [61]. The emulsion stability is therefore governed pri- 3.6. Demulsification
marily by the state of solubility of the asphaltenes in the crude oil.
Emulsion stability can be evaluated from the amount of resolved Stable water in oil emulsions are frequently obtained during produc-
water after centrifugation. Fig. 10 shows the aspect of a crude oil emul- tion and refining, especially when the oil contains appreciable amounts
sion after centrifugation at 4000 rpm during 30 min. The water phase at of asphaltenes. The crude oil refining processes requires that the water
the bottom originates from coalescence of water drops. The oil phase at content is reduced typically to less than 0.5%. Several demulsification
the top originates from sedimentation of oil drops enhanced by the cen- methods have been used to separate water from oil: centrifugation,
trifugal field. ultrasound, high voltage electric fields, increase of temperature, addi-
The percentage of resolved water is currently used to characterize tion of organic diluents to increase the difference of density between
emulsion stability, other methods including the application of electric oil and water. The efficiency of these methods to reach the target of
fields [63]. More microscopic methods involve the visualization of coa- 0.5% water is generally not sufficient, and demulsifiers are generally
lescence of two individual drops [64]. added [44].
When the solvent quality decreases, the emulsion stability in- The action of demulsifiers has been explained by Kim and Wasan
creases [61], likely because of the presence of a thicker and glassy [69]. During thinning of the liquid films between emulsion drops, the
interface layer. When the asphaltenes start to precipitate, they no species adsorbed at the film surfaces are entrained toward the film
longer adsorb at the oil–water interface and are unable to stabilize boundary, an interfacial tension gradient develops and counteracts the
emulsions [27]. motion of these species. Kim and Wasan proposed that the demulsifier
In general, the degree of coalescence correlates well with the elastic molecules should be soluble both in oil and water and adsorb fast at the
compression modulus. Exceptions were observed: for instance Dicharry surface. The demulsifier can then heal the interfacial tension gradients
and coworkers found that emulsions made with oils deprived of while the stabilizing species are expelled from the film center. If the
asphaltenes are much less stable than those stabilized by asphaltenes, demulsifier molecules not only replace the in-situ stabilizing species,
even when the interface compression modulus of the asphaltene-free but form surface layers with small compression elastic modulus, the
interfaces is the largest [65]. In this work, the shear modulus has not emulsion can be easily destabilized.
been measured, but the glassy behavior of the layers containing A number of Langmuir trough studies supported these findings [70,
asphaltenes has been evidenced by temperature studies as in Ref. [50]. 71]. Brewster angle microscopy showed that demulsifier molecules
The emulsion stability could then be better correlated to the shear mod- were able to disrupt the asphaltene film and that they dominate the in-
ulus than to the compression one. terface even during compression [72].
Both compression and shear interface elastic moduli increase with It was also shown that the phase separation rate is correlated with
time, consistent with the fact that aged emulsions are currently more the phase diagram oil–water–demulsifier, as in microemulsion systems:
stable than freshly prepared ones. the kinetics is faster if a three-phase equilibrium is observed (Winsor III
The influence of asphaltene concentration on emulsion stability system) [73]. This is in line with Kim and Wasan results, which showed
has been studied by Sjöblom and coworkers [66]. The stability be- that the demulsifier should be partitioned between oil and water.
comes very good above 2 g/L (in heptol with 60% toluene), consistent
with the independence of the shear modulus on asphaltene concen- 4. Asphaltenes at other type of interfaces
tration in this range, while the compression modulus decreases.
Other evidences of the importance of the shear properties on emul- 4.1. Asphaltenes at air–water surfaces
sion stability is the fact that coalescence of two oil drops is more
difficult under compression than under shear [67]. This is as for pro- A number of studies of asphaltenes spread at the air–water surface
tein emulsions that are very stable upon compression, but unstable were made with Langmuir troughs. A comprehensive review of the re-
upon shear that produces cracks above the yield stress in the surface cent work is presented in Ref. [74]. Although this situation is far from
layers [68]. practical conditions in which asphaltenes are usually encountered,
these studies benefited from the use of complementary techniques,
such as imaging (using Brewster angle microscopy, BAM) and spectros-
copy (visible and infrared). The surface layers can be deposited onto a
solid substrate using Langmuir–Blodgett methods, and then be imaged
at the nanoscale with Atomic Force Microscopy (AFM), which also pro-
vide the measurement of surface layer thickness. In general, AFM imag-
ing is performed on dried surfaces, and the formation of large and thick
asphaltene aggregates is frequently reported. These large aggregates
could be produced during the drying process, and care should be
taken while interpreting the pictures.
When toluene is used as spreading solvent, the curves surface pres-
sure versus area A per mass deposited at the surface (isotherms) are
very similar, irrespective of the source of asphaltenes: the pressure
starts to increase below 0.35 m2/mg and a kink is seen around A =
0.2 m2/mg announcing collapse. These asphaltene monolayers are
more expanded and less compressible than those at the air-heptol sur-
face [44]. Lobato and coworkers [75] showed that the isotherms are
more condensed (the pressure rises at smaller area) and are steeper,
when the concentration of the spreading solution is above 2 g/L, i.e.
when this solutions contains clusters. The aspect of the surface domains
changes as well, the reflectivity increases suggesting that as seen at the
oil–water surface, the clusters adsorb as a whole. At the difference of
surfactant molecules, domains are observed even when the surface
pressure is zero. This is reminiscent of situations observed with hydro-
Fig. 10. Crude oil emulsion after centrifugation. Picture from Ref. [62]. phobic polymers [76] or nanoparticles [39]: these entities are too

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx 9

hydrophobic to fully spread on water and prefer to gather together into agreement with the measured zeta potential. Increasing the solution
isolated cohesive surface domains. pH or the temperature increases significantly the repulsive force bar-
A pronounced hysteresis is observed during a compression–expan- riers and decreases adhesion forces [84]. The increase of repulsion
sion cycle, in contrast with the behavior of whole petroleum. The hys- with increasing pH is likely associated to the ionization of asphaltene
teresis decreases if a second cycle is performed. This was interpreted acidic groups, which are more numerous than the basic groups as
assuming that the asphaltene molecules initially flat at the surface tilt seen in Section 2.1.
progressively and irreversibly upon compression, as confirmed by sur- Studies of asphaltenes at solid surfaces are also relevant to oil
face spectroscopy [75]. transport in porous reservoirs. When asphaltenic crude oils are in-
Langmuir–Blodgett films from asphaltene layers deposited with volved, the wettability is mixed, portions of the reservoir rock are
dilute solutions (nano-aggregate regime) were performed [77] The oil wet, and others are water-wet. When crude oil first invades a
thickness of the determined by AFM is about 1–2 nm, consistent with water-filled reservoir, the asphaltenes originally at oil–water surface
the size of the nano-aggregate. Spectroscopic investigations demon- may adsorb onto the solid surfaces. Radke and coworkers [85]
strated that the polycyclic groups were oriented parallel to the surface showed using atomic force microscopy, that adsorption occurs if
and the peripheral alkanes perpendicularly [78]. the water film between oil drops and the solid ruptures. Advancing
and receding contact angles were shown to increase in time over
4.2. Asphaltenes at oil-gas surfaces days. These contact angles with the asphaltene-coated solid surface
control pore-level drainage and imbibition events. Aging of the oil–
Oil–gas surfaces are encountered in oil foams, which can occur dur- water surface and the asphaltene-coated surface are therefore both
ing production or refining of crude oils. Such foams are especially stable important in the evolution of mixed wettability.
in crude oils containing large amounts of asphaltenes [79] and pose
problems in oil–gas separation plants. Few characterizations of these
surfaces can be found in the literature. 5. Conclusions
Kilpatrick and coworkers showed that as emulsions, foams were
more stable near the asphaltene solubility limit, likely for the same rea- Although the behavior of asphaltenes in bulk is now being better un-
sons [80]. They also found that there is a good correlation between the derstood, the behavior at surfaces is still debated. It seems that at low
foamability and foam lifetime with the viscosity of asphaltene solutions, concentration asphaltenes adsorb in the form of molecular layers, but
as it could be expected. Indeed, liquid drainage between foam bubbles is that at higher concentrations, asphaltene aggregates adsorb instead.
slowed down and film rupture is delayed. Bauget and coworkers The structure of the layers remains to be understood, in particular,
showed that the correlation holds only if the foam is formed rapidly how the molecules are oriented at the surface.
[81]. If the foam is formed more slowly, the foam lifetime increases Asphaltenes solutions model well the behavior of crude oil, provided
more rapidly with asphaltene concentration. This additional increase the type of asphaltenes and solvent are well chosen. C5 asphaltenes in
is likely due to the consolidation of the asphaltene layers adsorbed at toluene model well the heavy crude oils, while the lighter oils are better
the bubble surfaces. The oil surface tension is low, and asphaltenes pro- modeled by C7 asphaltenes in heptol.
duce only small variations, a few mN/m. The adsorption times are very There is a clear relation between emulsion stability and interfacial
long, still longer that at oil–water surfaces [24]. The elastic compression rheology in the case of toluene solutions, where the compression mod-
moduli become measurable only after a few hours. The foam films life- ulus is larger than the shear modulus. However, the shear modulus
times have also been measured, they increase exponentially with the could play a more important role for emulsion stability in the case of
elastic modulus as asphaltene concentration is increased. The shear heptol solutions. The existence of a non-zero shear modulus appears as-
moduli have not been measured and likely affect foam stability as sociated to the formation of skins at the oil–water interface. Further
they do for emulsion stability. work remains to be done in order to relate better emulsion stability
and interfacial rheology. This question is not particular to crude oil
4.3. Asphaltenes at solid surfaces emulsions, it is equally central in emulsions stabilized by proteins or
particles (Pickering emulsions), in which case solid-like layers are also
Asphaltenes also adsorb at solid–liquid surfaces. This is of impor- present at the oil–water interface.
tance for understanding oil reservoir wettability, and also the stability Fewer studies have been made to understand the role of pH, salt and
of water in oil emulsion produced in situ because oil frequently contains added surfactants. The very low interfacial tensions encountered at high
clay particles, hydrophobized by the oil. Asphaltenes adsorb at the par- pH are transient and associated to a slow transport of oil species into
ticles surface through ion binding of their acidic groups with clay cat- water. Surfactants are able to displace asphaltenes from oil–water inter-
ions. Particle stabilized emulsions are usually very stable, and the faces although not fully. The surfactant efficiency depends on its molec-
topic has received much attention. Masliyah and coworkers showed ular structure, for reasons still to be elucidated. The role of salt is minor
that bitumen emulsion stability was enhanced by the presence of clay at neutral pH, but becomes important at acidic or basic pHs.
particles, especially small ones, large particles showing no stability en- The behavior of asphaltenes at oil–water, air–water and solid–water
hancement [32]. Kilpatrick and coworkers showed that as for emulsions interfaces is remarkably similar (adsorbed amounts, layer thicknesses,
without solids, the extent of asphaltene aggregation is the dominant aging). Their behavior at gas–oil interfaces has been much less investi-
factor for the stability control [63]. Adsorption isotherms of asphaltenes gated, but the growing interest in crude oil foams will likely lead to
onto hydrophilic solid substrates were very similar to those at oil–water new studies in the near future.
surfaces, including for the amount adsorbed, a few mg/m2, which in-
creases with bulk asphaltene concentration [82]. At pH below 7, the
zeta potential of colloidal silica is smaller than that of the same silica Acknowledgements
covered by asphaltenes, probably because the asphaltene basic groups
ionize [83]. Atomic force microscopy has been used by Masliyah and co- We acknowledge IFPEN for continuing financial support of our
workers to investigate the force between silica surfaces coated with work on asphaltenes during many years. We thank IFPEN colleagues
asphaltenes in aqueous solutions. A repulsive force develops with time for many interesting discussions and our present and past coworkers
suggesting asphaltene rearrangements after contact with water. The re- on asphaltene studies. We also acknowledge the COST action
pulsion is suppressed by addition of salt, confirming the electrostatic na- MP1106. D.L. thanks Vincent Pauchard for very stimulating
ture of this force. The measured surface potential is in reasonable discussions.

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
10 D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx

References [38] Li JB, Zhang Y, Yan LL. Multilayer formation on a curved drop surface. Angew Chem
Int Ed 2001;40(5):891–4.
[1] Speight JG. Petroleum asphaltenes – Part 1 – Asphaltenes, resins and the structure of [39] Zang DY, Rio E, Langevin D, Wei B, Binks BP. Viscoelastic properties of silica nanopar-
petroleum. Oil Gas Sci Technol Rev IFP Energ Nouv 2004;59(5):467–77. ticle monolayers at the air–water interface. Eur Phys J E Soft Matter 2010;31(2):
[2] Mullins OC. The modified Yen model. Energy Fuel 2010;24:2179–207. 125–34.
[3] Dutta PK, Holland RJ. Acid-base characteristics of petroleum asphaltenes as studied [40] Harbottle D, Chen Q, Moorthy K, Wang L, Xu S, Liu Q, et al. Problematic stabilizing
by non-aqueous potentiometric titrations. Fuel 1984;63(2):197–201. films in petroleum emulsions: Shear rheological response of viscoelastic asphaltene
[4] Hutin A, Argillier J-F, Langevin D. Mass transfer of asphaltenes between oil and films and the effect on drop coalescence. Langmuir 2014;30(23):6730–8.
water. OGST; 2015(submitted for publication). [41] Rane JP, Harbottle D, Pauchard V, Couzis A, Banerjee S. Adsorption kinetics of
[5] Schuler B, Meyer G, Peña D, Mullins OC, Gross L. Unraveling the molecular structures asphaltenes at the oil–water interface and nanoaggregation in the bulk. Langmuir
of asphaltenes by atomic force microscopy. J Am Chem Soc 2015;137(31):9870–6. 2012;28(26):9986–95.
[6] Evans F, Wennerström W. ##The Colloidal Domain. second ed. Wiley; 1999. [42] Pauchard V, Roy T. Blockage of coalescence of water droplets in asphaltenes solu-
[7] Eyssautier J, Levitz P, Espinat D, Jestin J, Gummel J, Grillo I, et al. Insight into tions: A jamming perspective. Colloids Surf, A 2014;443:410–7.
asphaltene nanoaggregate structure inferred by small angle neutron and X-ray scat- [43] Pauchard V, Rane JP, Zarkar S, Couzis A, Banerjee S. Long-term adsorption kinetics of
tering. J Phys Chem B 2011;115(21):6827–37. asphaltenes at the oil–water interface: A random sequential adsorption perspective.
[8] Porte G, Zhou HG, Lazzeri V. Reversible description of asphaltene colloidal associa- Langmuir 2014;30(28):8381–90.
tion and precipitation. Langmuir 2003;19(1):40–7. [44] Zhang LY, Xu Z, Masliyah JH. Langmuir and Langmuir−Blodgett films of mixed
[9] Gawrys KL, Kilpatrick PK. Asphaltenic aggregates are polydisperse oblate cylinders. J asphaltene and a demulsifier. Langmuir 2003;19(23):9730–41.
Colloid Interface Sci 2005;288(2):325–34. [45] Czarnecki J, Moran K. On the stabilization mechanism of water-in-oil emulsions in
[10] Barre L, Jestin J, Morisset A, Palermo T, Simon S. Relation between nanoscale struc- petroleum systems. Energy Fuel 2005;19(5):2074–9.
ture of asphaltene aggregates and their macroscopic solution properties. Oil Gas [46] Maestro A, Rio E, Drenckhan W, Langevin D, Salonen A. Foams stabilised by mixtures
Sci Technol Rev IFP Energ Nouv 2009;64(5):617–28. of nanoparticles and oppositely charged surfactants: Relationship between bubble
[11] Freer EM, Yim KS, Fuller GG, Radke CJ. Interfacial rheology of globular and flexible shrinkage and foam coarsening. Soft Matter 2014;10(36):6975–83.
proteins at the hexadecane/water interface: Comparison of shear and dilatation de- [47] Yang X, Verruto VJ, Kilpatrick PK. Dynamic asphaltene-resin exchange at the oil/
formation. J Phys Chem B 2004;108(12):3835–44. water interface: Time-dependent W/O emulsion stability for asphaltene/resin
[12] Walstra P. On the stability of casein micelles. J Dairy Sci 1990;73(8):1965–79. model oils. Energy Fuel 2007;21(3):1343–9.
[13] Groenewold J, Kegel WK. Anomalously large equilibrium clusters of colloids. J Phys [48] Spiecker PM, Kilpatrick PK. Interfacial rheology of petroleum asphaltenes at the oil–
Chem B 2001;105(47):11702–9. water interface. Langmuir 2004;20(10):4022–32.
[14] Fenistein D, Barre L, Broseta D, Espinat D, Livet A, Roux JN, et al. Viscosimetric and [49] Samaniuk JR, Hermans E, Verwijlen T, Pauchard V, Vermant J. Soft-glassy rheology of
neutron scattering study of asphaltene aggregates in mixed toluene/heptane sol- asphaltenes at liquid interfaces. J Dispers Sci Technol 2015;36(10):1444–51.
vents. Langmuir 1998;14(5):1013–20. [50] Bouriat P, El Kerri N, Graciaa A, Lachaise J. Properties of a two-dimensional
[15] Roux JN, Broseta D, Deme B. SANS study of asphaltene aggregation: Concentration asphaltene network at the water — cyclohexane interface deduced from dynamic
and solvent quality effects. Langmuir 2001;17(16):5085–92. tensiometry. Langmuir 2004;20(18):7459–64.
[16] Verruto VJ, Le RK, Kilpatrick PK. Adsorption and molecular rearrangement of ampho- [51] Alves DR, Carneiro JSA, Oliveira IF, Facanha Jr F, Santos AF, Dariva C, et al. Influence of
teric species at oil–water interfaces. J Phys Chem B 2009;113(42):13788–99. the salinity on the interfacial properties of a Brazilian crude oil-brine systems. Fuel
[17] Buckley JS, Hirasaki GJ, Liu Y, Von Drasek S, Wang JX, Gil BS. Asphaltene precipitation 2014;118:21–6.
and solvent properties of crude oils. Pet Sci Technol 1998;16(3-4):251–85. [52] Poteau S, Argillier JF, Langevin D, Pincet F, Perez E. Influence of pH on stability and
[18] Fossen M, Kallevik H, Knudsen KD, Sjöblom J. Asphaltenes precipitated by a two-step dynamic properties of asphaltenes and other amphiphilic molecules at the oil–
precipitation procedure. 2. Physical and chemical characteristics. Energy Fuel 2011; water interface. Energy Fuel 2005;19(4):1337–41.
25(8):3552–67. [53] McLean JD, Kilpatrick PK. Effects of asphaltene solvency on stability of water-in-
[19] paraffindepositionandcontrol.wikispaces.com/file/view/Asphltene_Aggregate.jpg/ crude-oil emulsions. J Colloid Interface Sci 1997;189(2):242–53.
183212945/800x376/Asphltene_Aggregate.jpg. [54] Hutin A, Argillier J-F, Langevin D. Mass transfer between crude oil and water. Part 1:
[20] Fossen M, Kallevik H, Knudsen KD, Sjoblom J. Asphaltenes precipitated by a two-step Effect of oil components. Energy Fuel 2014;28(12):7331–6.
precipitation procedure. 1. Interfacial tension and solvent properties. Energy Fuel [55] Hutin A, Argillier J-F, Langevin D. Mass transfer between crude oil and water. Part 2:
2007;21(2):1030–7. Effect of sodium dodecyl benzenesulfonate for enhanced oil recovery. Energy Fuel
[21] Barre L, Simon S, Palermo T. Solution properties of asphaltenes. Langmuir 2008; 2014;28(12):7337–42.
24(8):3709–17. [56] Langevin D, Poteau S, Henaut I, Argillier JF. Crude oil emulsion properties and their
[22] Freer EM, Radke CJ. Relaxation of asphaltenes at the toluene/water interface: Diffu- application to heavy oil transportation. Oil Gas Sci Technol Rev IFP Energ Nouv
sion exchange and surface rearrangement. J Adhes 2004;80(6):481–96. 2004;59(5):511–21.
[23] Strassner JE. Effect of ph on interfacial films and stability of crude oil–water emul- [57] Mackie AR, Gunning AP, Wilde PJ, Morris VJ. Orogenic displacement of protein from
sions. J Petrol Tech 1968;20(3):303. the air/water interface by competitive adsorption. J Colloid Interface Sci 1999;
[24] Jeribi M, Almir-Assad B, Langevin D, Henaut I, Argillier JF. Adsorption kinetics of 210(1):157–66.
asphaltenes at liquid interfaces. J Colloid Interface Sci 2002;256(2):268–72. [58] Moran K, Czarnecki J. Competitive adsorption of sodium naphthenates and naturally
[25] Chaverot P, Cagna A, Glita S, Rondelez F. Interfacial tension of bitumen–water inter- occurring species at water-in-crude oil emulsion droplet surfaces. Colloids Surf, A
faces. Part 1: Influence of endogenous surfactants at acidic pH. Energy Fuel 2008; 2007;292(2-3):87–98.
22(2):790–8. [59] Trabelsi S, Argillier J-F, Dalmazzone C, Hutin A, Bazin B, Langevin D. Effect of added
[26] Xu Y, Dabros T, Hamza H, Shefantook W. Destabilization of water in bitumen emul- surfactants in an enhanced alkaline/heavy oil system. Energy Fuel 2011;25(4):
sion by washing with water. Pet Sci Technol 1999;17(9-10):1051–70. 1681–5.
[27] Yarranton HW, Hussein H, Masliyah JH. Water-in-hydrocarbon emulsions stabilized [60] Rubin E, Radke CJ. Dynamic interfacial-tension minima in finite systems. Chem Eng
by asphaltenes at low concentrations. J Colloid Interface Sci 2000;228(1):52–63. Sci 1980;35(5):1129–38.
[28] Yang F, Tchoukov P, Pensini E, Dabros T, Czarnecki J, Masliyah J, et al. Asphaltene [61] McLean JD, Kilpatrick PK. Effects of asphaltene aggregation in model heptane-
subfractions responsible for stabilizing water-in-crude oil emulsions. Part 1: Interfa- toluene mixtures on stability of water-in-oil emulsions. J Colloid Interface Sci
cial behaviors. Energy Fuel 2014;28(11):6897–904. 1997;196(1):23–34.
[29] Alvarez G, Jestin J, Argillier JF, Langevin D. Small-angle neutron scattering study of [62] Hoebler-Poteau S. Relation entre les propriétés interfaciales et la stabilité des émul-
crude oil emulsions: Structure of the oil–water interfaces. Langmuir 2009;25(7): sions de brut lourd. , 6Université Paris; 2006(thesis).
3985–90. [63] Sullivan AP, Kilpatrick PK. The effects of inorganic solid particles on water and crude
[30] Jestin J, Simon S, Zupancic L, Barre L. A small angle neutron scattering study of the oil emulsion stability. Ind Eng Chem Res 2002;41(14):3389–404.
adsorbed asphaltene layer in water-in-hydrocarbon emulsions: Structural descrip- [64] Yarranton HW, Urrutia P, Sztukowski DM. Effect of interfacial rheology on model
tion related to stability. Langmuir 2007;23(21):10471–8. emulsion coalescence — II. Emulsion coalescence. J Colloid Interface Sci 2007;
[31] Taylor SD, Czarnecki J, Masliyah J. Disjoining pressure isotherms of water-in- 310(1):253–9.
bitumen emulsion films. J Colloid Interface Sci 2002;252(1):149–60. [65] Dicharry C, Arla D, Sinquin A, Graciaa A, Bouriat P. Stability of water/crude oil emul-
[32] Gafonova OV, Yarranton HW. The stabilization of water-in-hydrocarbon emulsions sions based on interfacial dilatational rheology. J Colloid Interface Sci 2006;297(2):
by asphaltenes and resins. J Colloid Interface Sci 2001;241(2):469–78. 785–91.
[33] Dickinson E. Adsorbed protein layers at fluid interfaces: Interactions, structure and [66] Fan Y, Simon S, Sjoblom J. Interfacial shear rheology of asphaltenes at oil–water in-
surface rheology. Colloids Surf B Biointerfaces 1999;15(2):161–76. terface and its relation to emulsion stability: Influence of concentration, solvent aro-
[34] Sztukowski DM, Yarranton HW. Rheology of asphaltene−toluene/water interfaces. maticity and nonionic surfactant. Colloids Surf, A 2010;366(1-3):120–8.
Langmuir 2005;21(25):11651–8. [67] Yeung A, Moran K, Masliyah J, Czarnecki J. Shear-induced coalescence of emulsified
[35] Alvarez G, Poteau S, Argillier J-F, Langevin D, Salager J-L. Heavy oil–water interfacial oil drops. J Colloid Interface Sci 2003;265(2):439–43.
properties and emulsion stability: Influence of dilution. Energy Fuel 2009;23(1): [68] Hotrum NE, Stuart MAC, van Vliet T, van Aken GA. Flow and fracture phenomena in
294–9. adsorbed protein layers at the air/water interface in connection with spreading oil
[36] Yarranton HW, Sztukowski DM, Urrutia P. Effect of interfacial rheology on model droplets. Langmuir 2003;19(24):10210–6.
emulsion coalescence — I. Interfacial rheology. J Colloid Interface Sci 2007;310(1): [69] Kim YH, Wasan DT. Effect of demulsifier partitioning on the destabilization of water-
246–52. in-oil emulsions. Ind Eng Chem Res 1996;35(4):1141–9.
[37] Yeung A, Dabros T, Czarnecki J, Masliyah J. On the interfacial properties of [70] Ese MH, Galet L, Clausse D, Sjoblom J. Properties of Langmuir surface and interfacial
micrometre-sized water droplets in crude oil. Proc R Soc Lond A Math Phys Sci films built up by asphaltenes and resins: Influence of chemical demulsifiers. J Colloid
1999;455(1990):3709–23. Interface Sci 1999;220(2):293–301.

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005
D. Langevin, J.-F. Argillier / Advances in Colloid and Interface Science xxx (2015) xxx–xxx 11

[71] Le Follotec A, Pezron I, Noik C, Dalmazzone C, Metlas-Komunjer L. Triblock co- [79] Abivin P, Henaut I, Chaudemanche C, Argillier JF, Chinesta F, Moan M. Dispersed sys-
polymers as destabilizers of water-in-crude oil emulsions. Colloids Surf, A tems in heavy crude oils. Oil Gas Sci Technol Rev IFP Energ Nouv 2009;64(5):
2010;365(1-3):162–70. 557–70.
[72] Fan Y, Simon S, Sjoeblom J. Influence of nonionic surfactants on the surface and [80] Zaki NN, Poindexter MK, Kilpatrick PK. Factors contributing to petroleum foaming. 2.
interfacial film properties of asphaltenes investigated by langmuir balance and Synthetic crude oil systems. Energy Fuel 2002;16(3):711–7.
brewster angle microscopy. Langmuir 2010;26(13):10497–505. [81] Bauget F, Langevin D, Lenormand R. Dynamic surface properties of asphaltenes and
[73] Goldszal A, Bourrel M. Demulsification of crude oil emulsions: Correlation to resins at the oil–air interface. J Colloid Interface Sci 2001;239(2):501–8.
microemulsion phase behavior. Ind Eng Chem Res 2000;39(8):2746–51. [82] Ekholm P, Blomberg E, Claesson P, Auflem IH, Sjoblom J, Kornfeldt A. A quartz crystal
[74] Hua Y, Angle CW. Brewster angle microscopy of Langmuir films of athabasca bitu- microbalance study of the adsorption of asphaltenes and resins onto a hydrophilic
mens, n-C5 asphaltenes, and SAGD bitumen during pressure–area hysteresis. Lang- surface. J Colloid Interface Sci 2002;247(2):342–50.
muir 2012;29(1):244–63. [83] Hannisdal A, Ese MH, Hemmingsen PV, Sjoblom J. Particle-stabilized emulsions:
[75] Lobato MD, Pedrosa JM, Möbius D, Lago S. Optical characterization of asphaltenes at Effect of heavy crude oil components pre-adsorbed onto stabilizing solids. Colloids
the air−water interface. Langmuir 2009;25(3):1377–84. Surf, A 2006;276(1-3):45–58.
[76] Mann EK, Henon S, Langevin D, Meunier J. Molecular layers of a polymer at the free- [84] Abraham T, Christendat D, Karan K, Xu Z, Masliyah J. Asphaltene–silica interactions
water surface — microscopy at the brewster-angle. J Phys II 1992;2(9):1683–704. in aqueous solutions: Direct force measurements combined with electrokinetic
[77] Andrews AB, McClelland A, Korkeila O, Demidov A, Krummel A, Mullins OC, et al. studies. Ind Eng Chem Res 2002;41(9):2170–7.
Molecular orientation of asphaltenes and PAH model compounds in Langmuir– [85] Freer EM, Svitova T, Radke CJ. The role of interfacial rheology in reservoir mixed
Blodgett films using sum frequency generation spectroscopy. Langmuir 2011; wettability. J Pet Sci Eng 2003;39(1-2):137–58.
27(10):6049–58.
[78] Orbulescu J, Mullins OC, Leblanc RM. Surface chemistry and spectroscopy of UG8
asphaltene Langmuir film, Part 1. Langmuir 2010;26(19):15257–64.

Please cite this article as: Langevin D, Argillier J-F, Interfacial behavior of asphaltenes, Adv Colloid Interface Sci (2015), http://dx.doi.org/10.1016/
j.cis.2015.10.005

You might also like