Sie sind auf Seite 1von 12

Advances in Colloid and Interface Science 217 (2015) 112

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science


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

Historical perspective

Thermodynamic modelling of asphaltene precipitation and


related phenomena
Esther Forte, Spencer E. Taylor
BP Centre for Petroleum and Surface Chemistry (BP-CPSC), Department of Chemistry, University of Surrey, Guildford, Surrey GU2 7XH, United Kingdom

a r t i c l e

i n f o

a b s t r a c t
Asphaltenes are considered to be the heaviest and most polar fractions of crude oils and are frequently implicated
in problems encountered during production and rening as a result of phase separation. In recent years,
considerable effort has been given to understanding the phase behaviour of these structurally heterogeneous
materials from both experimental and computational perspectives. Various experimental studies have conrmed
the long-advanced colloidal behaviour of asphaltenes in organic media, and this has inspired a number of
modelling strategies. The present review is specically concerned with advances in modelling asphaltene
phase behaviour with emphasis on the use of the statistical associating uid theory (SAFT), which it attempts
to place into the wider context of thermodynamic treatments.
2014 Elsevier B.V. All rights reserved.

Available online 15 December 2014


Keywords:
Asphaltene
Phase behaviour
Modelling
Equations of state
SAFT

Contents
1.
2.
3.

Introduction . . . . . . . . . . . . .
Lattice uid theories . . . . . . . . .
Equations of state. . . . . . . . . . .
3.1.
Cubic equations of state. . . . .
3.2.
Cubic plus association . . . . .
3.3.
Statistical associating uid theory
4.
Other approaches . . . . . . . . . . .
5.
Conclusions and future trends . . . . .
Acknowledgements . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

1. Introduction
Asphaltenes have attracted much research curiosity for many years.
Interest has ranged from their chemical characterisation in order to
dene their molecular structure, to understanding their solution and
interfacial behaviour, including their aggregation and precipitation
tendencies. However, even today, many of these aspects are still open
questions where further research is needed. The main reasons for this
stem from the limited or ambiguous knowledge concerning their
characterisation.
Crude oils are complex mixtures of thousands of components [1],
which are difcult to analyse and therefore characterise [2]. The vast
amount of components existing in crude oil has led to comparison
Corresponding author.
E-mail address: s.taylor@surrey.ac.uk (S.E. Taylor).

http://dx.doi.org/10.1016/j.cis.2014.12.002
0001-8686/ 2014 Elsevier B.V. All rights reserved.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

1
3
4
5
5
6
8
9
9
9

with the amount of genes in the genome, giving rise to the use of
the words petroleome and petroleomics. The difculties in
characterising the crude oil are also evident in the case of
asphaltenes, the heaviest nondistillable fraction in crude oil, where
the characteristics of this fraction will also depend on the source
of the crude oil [3]. Generically, they are classied solely on their
solubility properties, typically as the fraction of oil that is soluble
in toluene and insoluble in n-heptane.
In the oil industry upstream and downstream operations could each
benet from a comprehensive understanding of asphaltene-related
phenomena. Over the years, many studies have shown asphaltenes to
be active at water-oil interfaces, resulting in the stabilization of oil/
water emulsions, and solid-oil interfaces, causing modications to
rock wettability (refs. [4,5] and references therein). On the other hand,
a number of problems also result from their occulation and sedimentation caused by changes in solvency [6,7].

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

The phase behaviour of reservoir uids is intricate. Regions of up to


four phases in equilibrium (solid-liquid-liquid-vapour) have been
identied in asymmetric oil mixtures containing asphaltenes, resins
and light hydrocarbons as constituents [8]. Asphaltenes and other
petroleum fractions also exhibit complex path-dependent phase behaviour, which implies an added difculty in the prediction of asphaltene
precipitation. In particular precipitated asphaltenes have been observed
to undergo a complex transition from solid to liquid in a temperature
range of ~(300500) K [912] and additionally to present amphotropic
[13] liquid-crystal domains [14].
From a physicochemical viewpoint, asphaltenes have presented a
very interesting research topic, not least because of their colloidal nature in crude oils [1518] with the earliest theories based on interactions with structurally-similar resins. Asphaltenes are known to
contain a high proportion of polycyclic aromatic groups and aliphatic
chains. Two different molecular architectures have been proposed
which are consistent with established atomic compositions and supported by a range of analytical techniques. The rst is an archipelago
structure wherein differentiated aromatic groups are linked together
through alkyl bridges; the second is an island or continental structure wherein the core of the molecule is based on a single polycyclic aromatic hydrocarbon ring that contains peripheral alkane substituents.
Different authors have suggested models with one or the other being
predominant for different asphaltene samples (e.g., refs. [1925] favour
the archipelago architecture and [2631] favour the continental) or
asphaltene fractions [32], notwithstanding that the reliability of some
common experimental methods may be a matter of concern [3335]. A
predominant molecular structure for asphaltenes supporting the continental archetype has been identied in the Yen-Mullins model [3638]
(cf. Fig. 1); these structures can further assemble in the form of
nanoaggregates at low concentrations (~ 100 mg L1) and clusters at
higher concentration (~3 g L1). In the past, self-aggregation leading
to the formation of aggregates has led to much disagreement regarding
asphaltene molecular weights; over the years values have ranged from
a few hundred to several million (see refs. [26,3941] and references
therein). In turn, from a modelling perspective this has led to different
representations of the basic structural unit in the model (i.e., ranging
from molecular to aggregated).
However, the lack of a comprehensive understanding of the nature
of asphaltenes as well as their interactions with other components of
crude oil has motivated the development of a variety of different
approaches to model their behaviour in solution and mechanisms

involved in their aggregation and subsequent precipitation. Central to


the modelling discussions is the question of the intermolecular forces
involving the asphaltenic constituents and other components in crude
oil, which affect their solubility and aggregation properties. Perhaps
one of the most systematic studies is that of Wiehe [42], who studied
the dominant interactions in asphaltenic fractions through a solubility
analysis in a wide range of solvents arranged on the basis of their
complexing and eld force solubility parameters. These parameters
provide indications of whether directional (hydrogen-bonding or electrostatic) or nondirectional (van der Waals and other polar) interactions
dominate their solubility. Those with high eld force but low
complexing solubility parameters were found to be the best solvents,
supporting the idea that non-polar van der Waals interactions play a
major role. The importance of van der Waals dispersion interactions in
aggregates of asphaltenes and resins has also been supported by several
other authors [4345], although electrostatic interactions have been additionally suggested [44,46]. However, the contribution from hydrogenbonding interactions should not be disregarded in the case of
asphaltenes with a high proportion of polar functionality or heteroatom
content, as these are likely to promote their propensity to stabilise
water-oil emulsions [47].
Of particular interest from a modelling perspective is the nature of
the instability responsible for the precipitation of asphaltene-rich
phases for which both solid-liquid [48,49] and liquid-liquid [50] immiscibilities have been suggested, which has given rise to some disagreement in the approach to model these systems.
Furthermore, two different conceptual descriptions of the mechanism by which asphaltenes are considered as being stable in crude
and residual oils have provided the basis for modelling asphaltene stability over the years. As alluded to earlier, the rst supports the notion
that the phase behaviour of these systems is governed by their colloidal
nature, in which it was originally considered that asphaltenes are
dispersed in the oil matrix in the form of aggregates that are stabilised
by structurally similar resins which possess a slightly greater afnity
for the bulk oil. In this case, asphaltene precipitation is assumed to be
the result of a loss in the stabilising effects of the resin molecules. The
resulting colloidal picture led Nellensteyn [51] to introduce the concept
of a micelle for the rst time by in petroleum science. Later, Pfeiffer
and Saal [52] further developed the resin-stabilised asphaltene micelle
model. Within this rst perspective, most attention is paid to
asphaltene-asphaltene and asphaltene-resin association based on pi-pi
stacking of the aromatic units At that time, no specic details were

Molecule

Nanoaggregate

Cluster

(~1.5 nm)

(~2 nm)

(~5 nm)

Fig. 1. Conceptual view of the modied Yen-Mullins model. The proposed molecular structure is seen in the left image. The polycyclic aromatic hydrocarbon of the molecules is represented
by a at oval in the middle image; these can form nanoaggregates with aggregation numbers of about six which contain a single aromatic disordered stack in the interior (cf. middle
image). The latter can form clusters with aggregation number of about eight; this is represented in the right image, where the spheres depict the aromatic disordered stack in a
nanoaggregate. The lines are the (mainly) alkane substituents. Adapted from ref. [36].

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

given regarding the formation of micellar aggregates, although now,


based on the known solvophobic character of nonpolar solvents, intermolecular interactions are known to be much weaker than the more familiar hydrogen bonds in water, for example. These differences are
reected in measurable physical properties such as cohesive density
and surface tension; comparing water and nonpolar solvents, for example, indicates that the free energy benets of micellisation would be less
in a nonpolar solvent than in water. Therefore, for the formation of
inverse micelles in hydrocarbons, the solvophobic effect suggests
that the critical micelle (or aggregation) concentration will be less pronounced and aggregation numbers will be much smaller, as previously
elucidated by Ruckenstein and Nagarajan [53] thirty-ve years ago,
and supported by very recent experimental results [54].
The second description is based on a more general (molecular)
picture in which asphaltenes are considered to be dissolved in the
oil, rather than being suspended as colloids. Here, asphaltene precipitation is assumed to be the result of a phase-separation process, and
therefore no attention is paid to the structural characteristics of the
system. Within this second conceptual description, focus is placed
on mimicking the essential molecular characteristics of a set of representative components in the oil mixture. The role of resins is not clearly
distinguished from that of any other oil components and as a result
these are not always modelled as a separate molecular entity. One of
the rst thermodynamic solubility models (also called the liquid
model by the authors) is that of Hirschberg et al. [55]. In their work,
two pseudocomponent [asphalt (i.e., asphaltenes + resins) and solvent]
and three pseudocomponent (asphaltenes, resins and solvent) models
are considered, depending upon the conditions at which resins are associated with asphaltenes or are separate.
These models have been used by different authors with slight variations. Leontaritis and Mansoori [56] considered asphaltenes as solid-like
particles that are considered to be suspended in the crude oil in the form
of colloids stabilised by repulsive interactions brought about by resin
molecules adsorbed on the surface. The oil phase is assumed to be
asphaltene-free and it is only the chemical potential of the resins in
the solid and liquid phases that are used to determine the phase separation [56].1 Firoozabadi and coworkers [57,58] use a thermodynamic
model of micellisation based on the earlier works of Nagarajan and
Ruckenstein [59], Blankschtein et al. [60] and Puvvada and Blankschtein
[61], in which asphaltenes are assumed to aggregate in the core of micellar structures peptized by bipolar resins present on the surface. Equilibrium between the asphaltenic monomers in solution and those
forming micelles of a given number of aggregates is considered.2 A
colloidal-based description has also been suggested within a liquidliquid equilibrium interpretation [6264], wherein equilibrium is
based on both, asphaltene and resin, types of molecules.
The use of colloidal models is not without some controversy, however. Recent studies refuting the colloidal model are discussed by
Punnapala and Vargas [65], while micellar models of asphaltene aggregation have been suggested as being inadequate from a lack of evidence
for a critical micelle concentration, or concentration at which the
asphaltenes start to associate, using isothermal titration calorimetry
[66]. However, whilst this approach failed to detect nanoaggregate formation at low asphaltene concentrations, the nding is consistent with
the dominant entropic contribution to the process [67], and a number of
techniques have indeed provided evidence for nanoaggregate

formation in asphaltene systems, including high-Q ultrasonics [68],


NMR spectroscopy [69], electrical conductivity [67], small-angle neutron and x-ray scattering [70], mass spectrometry [70], centrifugation
[67], and the analysis of gravity gradients in oilelds [7173].
The reversibility of the process of asphaltene precipitation is also a
controversial topic. The effect of pressure on asphaltene precipitation
is generally thought to be reversible [40,74,75], whereas most disagreement is found regarding the inuence of temperature and composition.
The re-dissolution of precipitated asphaltene at low temperatures is
kinetically slow and long periods of time may be required to verify the
reversibility of the process. Some authors have regarded asphaltene precipitation only as partially reversible [76], especially when asphaltenes
are destabilised well beyond the onset conditions [77,78]. Apparent irreversibilities have also been associated with inappropriate reversibility
criteria [79] and to the existence of an energy barrier to asphaltene dissociation [80]. In spite of the difculties, the number of studies supporting
reversibility is remarkable [74,75,7983].
An interesting question is relevant to the energetic versus entropic
balance, specically in relation to the level of detail needed to model
these systems. In the models used to describe asphaltene precipitation,
there has been support coming from two different directions. One is an
effort to describe in detail the type of interactions playing a role, which
is considered as the dominant mechanism. The second perspective
considers asymmetry in sizes between the different components in
the mixture to be the dominant cause producing the instabilities and
segregation into two liquid phases. The latter has been shown to be an
important mechanism leading to uid-uid phase separation in
polymer solutions, or colloid-polymer systems, even in the case of
athermal or purely repulsive systems, examples of which are given
in refs. [8489].
In the next sections we will review the main approaches used to describe the phase behaviour of asphaltenes. This present review attempts
to complement earlier reviews on the subject [6,45,78,9092]. The most
recent are from Greeneld [92], where emphasis is on the application of
molecular simulation techniques, and Wiehe [45], which focuses on
asphaltene solubility and the use of solubility parameters and the
Flory-Huggins theory. In the present review we offer a detailed discussion on the main uid theories applied in this eld.
2. Lattice uid theories
A traditional approach to study asphaltene precipitation has been
the use of polymer models, in particular lattice theory [93]. The combination of the Flory-Huggins [94,95] polymer-solution theory with the
regular-solution theory of Scatchard-Hildebrand [96,97] has been the
most common methodology to study the precipitation of asphaltenes
in the literature. The conventional application of the theory is based
on a pseudo-binary approximation that assumes the mixture consists
of two pseudocomponents; these are the solvent (i.e., the oil) and the
polymer (i.e., the asphaltenes). According to this model, the system is
characterised based on the solubility parameter and the molar volume
of its components, where an expression to determine the solubility of
asphaltene in oil is obtained from evaluation of the Gibbs free energy
of mixing as3
h
i
v
2
Gmix RT ns ln s na lna ns a s a s
RT

This approach considers the process as being irreversible and not based on solving
equilibria for the asphaltenic monomers in solution and in the colloids. No distinction is
made regarding the number of asphaltenic units in the colloids and the equilibrium is
solved only based on the resins. In the absence of resins, asphaltene molecules occulate
and precipitate in this model.
2
Micelles of different size, i.e., different aggregation numbers, are regarded as different
species. The average state of an asphaltene molecule in a micellar core is to be assumed to
be similar to its state in a pure solid asphaltene phase. In the absence of resins, asphaltenic
molecules are assumed to precipitate immediately due to the low solubility of the
asphaltene monomeric molecules in the bulk.

where R is the universal gas constant, T is the absolute temperature, n is


the number of moles, is the volume fraction, is the solubility parameter and the subscripts a and s refer to asphaltenes and oil, respectively.
Because the effect of pressure is not taken into account within lattice
3
The rst terms are due to the entropy of mixing; the last one is due to the enthalpy of
mixing given by the regular solution model.

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

theories, combinations with cubic equations of state have been used to


solve for vapour-liquid equilibrium.
One of the rst applications of this Flory-Huggins-Hildebrand theory
was used by Hirschberg et al. [55] to calculate the amount of asphalt
precipitated from a liquid phase whose composition had been previously calculated using the Soave-Redlich-Kwong equation of state [98]. The
model of Hirschberg and coworkers was later used by other authors
[99101], but it is limited to a monodisperse asphaltenic fraction that
precipitates as a pure component. Modications have been suggested
to deal with the assumption that the precipitated phase contains not
only pure asphaltene but consists of a phase concentrated in
asphaltenes containing also a fraction of solvent. This is the work of
Cimino et al. [102], which, however, assumes a pure solvent phase and
cannot therefore be applied to estimate the amount of asphaltene precipitated. Other authors have avoided this type of simplifying assumption. This is the case of Wang and Buckley [103], where a correlation
with the refractive index of non-polar species is used to estimate the
solubility parameter of the oil [103], reducing the uncertainty of the
model. In the work of Yarranton et al. [104106], the bitumen or
heavy oil is considered to consist of the SARA (saturates, aromatics,
resins and asphaltenes) pseudocomponents, the solubility parameters
of which are obtained from a correlation with enthalpy of vaporisation
and molar volume data. Oil phases consisting of both asphaltene and
non-asphaltene components are also assumed in the work of
Mohammadi and Richon [107]. Further modications of the FloryHuggins theory have been presented and successfully applied to describe asphaltene precipitation. One example of such modications is
based on including a binary interaction parameter [108] that
characterises the interaction between unlike molecules, as applied by
Pazuki and Nikookar [109] and Modi and Edalat [110]; this interaction
parameter is determined as a function of molecular weight [109,110].
Mohammadi et al. [111] have combined the approach with chemical
theory to account for asphaltene self-association resulting in n-mer aggregates. The theory has been also extended to predict asphaltene concentration gradients in oil reservoirs by incorporation of a gravitational
term [112]. Recently, an application of the Flory-Huggins theory combined with the Non Random Two Liquid (NRTL) model [113] has been
presented to describe asphaltene precipitation in different solvent ratios
[114].
Other polymer theories have been used to model asphaltene phase
behaviour. In particular, a number of authors [104106,115119] have
developed thermodynamic models based on the Scott-Magat polymer
theory [120,121], where the heterogeneous structure and polydispersity in the asphaltene molecular weight is taken into account in the
model. Although there is recent practical evidence from sulphur analysis and speciation of reservoir crude oils that there are circumstances
where polydispersity can be ignored at the molecular [73] and
nanoaggregate/cluster [72] levels, polydisperse models have been
developed to account for size and molecular weight distributions of aggregates due to asphaltene self-aggregation [115,119,122,123]. In these,
the polydispersity is specied through a continuous molecular-weight
distribution function, which may be useful to account for a size distribution of aggregates due to asphaltene self-association. The rst application of this type of model to study precipitation of an asphaltene
fraction characterised by a molecular-weight distribution is that of
Kawanaka et al. [122] in which a gamma distribution is arbitrarily
chosen [122]; the variance of the distribution is treated as an adjustable
parameter. Although gamma distribution functions have been a common choice [105,106,115,124], Schultz-Zimm distribution functions
[104] and fractal distribution functions [119,123] have also been suggested. A comparison between these different distribution functions is
made in the study of Manshad and Edalat [117], where the use of a fractal distribution function leads to best agreement with experimental
data. A typical element in these studies is to model asphaltenes as the
only self-associating fraction. In the work of Yarranton et al. [124] a
single molecular weight distribution for a mixture of asphaltenes and

resins (or a combined pseudocomponent) is seen to best characterise


the self-association and precipitation of asphaltenes and resins that is
observed when low carbon number alkanes are used as precipitant.
The simplicity of lattice uid theories has motivated its extensive
application in the eld, but these approaches present important limitations. Until recently the effect of density or compressibility on the phase
behaviour was not taken into account to allow for the dependence on
pressure. However, this has been accomplished in a very practical way
by calculating pressure variations associated with asphaltene gradients
in oil columns, through the introduction of a gravity term into the equation of state [112,125].
Accounting for the effect of temperature is also not straightforward,
and so is usually described empirically. The temperature dependence of
the solubility parameters, densities and mass distribution is obtained
through tting to empirical correlations; however, some authors have
attempted the extension to different conditions [105,106,126]. An example is seen in Fig. 2, where predictions are calculated after tting to
asphaltene yields from n-heptane and then adjusting the temperature
dependence with another oil. A different drawback of regular solution
theory is the fact that the values for the solubility parameters of
asphaltenes and other crude oil components are not well known.
Although these are typically determined through correlation of
experimental data, solubility parameters are also obtained using
equations of state [93,109,123,127133]. Other authors have also
highlighted the need for a more detailed description of physicochemical phenomena than that underlying regular solution theory,
commenting on some of the limitations and misapplication of the
theory to these mixtures [134].

3. Equations of state
One of the advantages in the use of equations of state is the versatility in their application, which is not restricted to a range of conditions.
Equations of state are also not limited to modelling precipitation boundaries, but can be used to describe the entire phase equilibrium diagram
of the mixture. In this section we will review the application of cubic
equations of state, cubic plus association and the statistical associating
uid theory in the eld of asphaltene precipitation.

0.25
T=273 K
T=296 K

0.20

T=273 K
T=296 K
T=323
T=323 K

0.15

0.10

0.05

0.00
0.4

0.5

0.6

0.7

0.8

0.9

Fig. 2. Fractional yield (mass ratio of precipitate after washing and drying per mass of
heavy oil) versus solvent mass fraction s for a Lloydminster heavy oil diluted with either
n-pentane or n-heptane at various temperatures. Solid curves are results using the
regular- solution model of [106] for polydisperse asphaltenes; the continuous curves
correspond to dilution with n-pentane and the discontinuous curves to dilution
with n-heptane. Symbols are experimental data at temperatures of: red squares,
273 K; green circles, 296 K; blue triangles, 273 K; purple diamonds, 296 K and orange
inverted triangles, 323 K. Curves have the same colour code at each temperature.
Adapted from ref. [106].

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

3.1. Cubic equations of state


Cubic equations of state have a long history and tradition as widely
applied tools in industry. Since the rst cubic equation of van der
Waals in 1894 [135], a large number of cubic equations of state have
been proposed (excellent reviews can be found in refs. [136140]).
The most popular are perhaps those of Peng-Robinson (PR) [141] and
Soave-Redlich-Kwong (SRK) [98]. A number of authors have used
cubic equations to model asphaltene phase behaviour [58,142150]. In
the model of Kohse et al. [144], the Peng-Robinson equation of state is
used. In this work [144] a crude oil containing asphaltene is lumped
into twelve pseudocomponents; the heaviest pseudocomponent is
split into a non-precipitating component and a precipitating component
(essentially asphaltene, which is modelled as a pure solid). The parameters of the solid model are adjusted to match precipitation data for the
asphaltene of study. In the application of Sabbagh et al. [147], the PengRobinson equation of state is used to model asphaltene precipitation
from n-alkane-diluted bitumens. In their study, asphaltenes are considered to be macromolecular aggregates with a broad mass distribution
and represented by a large number of pseudocomponents (N 30); the
parameters for these fractions are obtained from those for asphaltene
monomers [147]. Peng-Robinson is also the equation of state of choice
in the work of Behar et al. [146], in which a large number of components
(29) and pseudocomponents (6) is again used to characterise the crude.
In their work, the thermodynamic model is integrated into an industrial
reservoir simulator with the aim of determining the asphaltene deposited at different stages of a reservoir [136]. A thermodynamic approach
using a cubic equation of state is also implemented into a reservoir simulator in the work of Fazelipour et al. [148]. Following the model of
Nghiem et al. [142], the heaviest fraction in the oil is assumed to consist
of a non-precipitating and a precipitating component. The latter is considered to be asphaltene, which is treated as a component that precipitates in a pure solid phase. Oil and gas phases are modelled with the PR
equation of state [148]. This solid model for asphaltene is also used in
the work of Tavakkoli et al. [151] and in the work of Jamaluddin et al.
[145] who use a seventeen component (and pseudocomponent) representation of the live oil of study. Another example using this equation
of state is that of Castellanos Daz et al. [150], in which this is applied
to bitumen, propane and carbon dioxide mixtures. At least sixteen
pseudocomponents are used to represent accurately the normal boiling
point curve of the bitumen of study. After optimising the interaction
parameters to t experimental saturation pressures of mixtures of
each solvent with the bitumen, the model is used to predict asphaltene
precipitation from n-heptane dilutions of the bitumen. However, the
prediction of asphaltene yields is seen to fail at high dilution ratios
[150].
A number of authors [57,58,143,149,152154] use the PengRobinson equation of state to calculate the fugacity coefcients of the
monomeric species within a thermodynamic micellisation model. In
this model, asphaltenes are assumed to associate to form micelles
with resins on their surface; monomeric asphaltenes, monomeric resins
and micelles are considered solutes in the solvent. The precipitate is in
most studies considered a mixture of asphaltenes and resins which is
solid at ambient conditions [58,149,154] and liquid at high temperature
[143,153].
The predictive capabilities of cubic equations of state have been
questioned, as for example in ref. [155], where results using the SRK
equation of state are compared to calculations with a more sophisticated approach based on the Statistical Associating Fluid Theory discussed
in Section 3.3. Parameters regressed with experimental onset pressure
data for a given percentage of injected gas are obtained and then used
to predict the temperature dependence of the onset pressure for
different amounts of injected gas. Considerable loss in predictive performance is observed for the cubic equation of state.
The success of cubic equations of state within the petroleum industry has been motivated by their simplicity and current state of maturity.

However, it should be recalled that cubic equations were established


following a simple description of molecules based on a molecular
model of a single sphere with a hard core and attractions mediated
through long-range dispersion forces. Unfortunately, such models may
not be appropriate for systems which exhibit hydrogen bonding or for
highly asymmetric molecules.
3.2. Cubic plus association
The effort in extending the range of application of cubic equations of
state to systems that associate as a result of directional interactions, e.g.,
hydrogen bonding as in water, motivated the development of hybrid
equations combining cubic equations of state with association theories.
This gave birth to the Cubic Plus Association (CPA) equation of state
[156], which in its initial formulation combined the SRK equation of
state with the thermodynamic perturbation theory of Wertheim
[157160]. The latter constitutes the association term in Statistical
Associating Fluid Theory introduced in the next section.
A number of authors have used the CPA equation of state to model
asphaltene phase behaviour [161165]. In the work of Li and
Firoozabadi [161,162], the PR equation of state is used within the CPA
framework. The rst study focuses on asphaltene precipitation in
model solutions as well as in heavy oil and bitumen, whereas the latter
two are considered to be a mixture of three pseudocomponents: saturates, aromatics/resins and asphaltenes [161]. A representative result
of their study can be seen in Fig. 3, in which the effect of temperature
is analysed in comparison with experimental data. In the second
study, asphaltene precipitation from a live oil is modelled, where the
live oil is considered to be a mixture of a considerable number of
components and pseudocomponents, together with a hydrocarbon
residue [162]. The latter is divided into a heavy pseudocomponent
and asphaltene. Self-association between asphaltene molecules as
well as between these and heavy pseudocomponent molecules is considered. Although requiring the use of a temperature-dependent
cross-association energy between asphaltene and heavy component
molecules, the approach successfully reproduces the bubble curve,
asphaltene precipitation boundary and gas-oil-asphaltene phase behaviour. The live oil is split in a similar manner in a mixture of pure components and pseudocomponents, resins and asphaltenes in the work of
Shirani et al. [164]. Different from the previous study, association of
asphaltenes with other components is neglected in favour of considering asphaltene self-association to be predominant. In the work of
Zhang et al. [163] a black-oil reservoir uid is characterised by eighteen
components and pseudocomponents and both asphaltene selfassociation and asphaltene-resin association are considered.
1.0

0.8
273K

0.6

296K
323K

0.4

0.2

0.0
0.4

0.5

0.6

0.7

0.8

0.9

1.0

Fig. 3. Fractional precipitation a (mass ratio of precipitate per mass dissolved) of


Athabasca asphaltenes in solutions of n-heptane and toluene versus the volume fraction
of n-heptane C7 at the temperatures shown. Continuous curves are results using the
CPA equation of state, whereas symbols are experimental data [147]. Adapted from ref.
[161].

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

Other association models, such as chemical theories incorporated in


cubic equations of state, have also been applied to the study of
asphaltene phase behaviour [166168]. In the work of Vafaie-Sefti
et al. [166] a model consisting of at least fteen components (and
pseudocomponents) is used with the PR equation of state combined
with a chemical contribution [169,170] to describe different oils; both
asphaltenes and resins are considered to associate. The PR equation of
state is also combined with a chemical approach in the study of Du
and Zhang [167]. While lighter components (~10 in number) are treated explicitly, components heavier than hexane are lumped into twelve
pseudofractions; the latter are also split into parafnic, naphthenic
and aromatic subfractions, where asphaltenes are considered to be the
heaviest aromatic fraction. Resins are not specied as a different fraction
and only asphaltene self-association is considered [167]. Shirani et al.
[168] compare the use of both PR and SRK coupled with a chemical
term to study asphaltene precipitation. In their work a model based
on about fteen components and pseudocomponents is used and both
asphaltene self-association and asphaltene-resin association are considered. Their results show a better performance of the PR equation of state
in comparison with experimental data [168]. In these chemical theories,
association is considered to be the result of a chemical reaction between
the monomers, giving rise to new distinct species. A limitation of these
approaches is that the corresponding temperature-dependent chemical
equilibrium constants appear in the thermodynamic relations and need
to be evaluated.4
3.3. Statistical associating uid theory
An alternative to previous approaches is the use of perturbation theories that are rooted in statistical mechanics. The Statistical Associating
Fluid Theory (SAFT) [173,174] is one such approach. The SAFT formalism stems from the thermodynamic perturbation theory of Wertheim
[157160,175,176] explicitly to take into account the possible nonspherical nature of molecules as well as association through directional
interactions characteristic of hydrogen bonding. The main assumption
is that of considering the thermodynamic properties of chain-like and
associating uids can be obtained from knowledge of the properties of
a monomeric uid of reference (consisting of single spherical segments). The equation is typically written in terms of the Helmholtz
free energy as a sum of contributions
AA

ideal

mono

chain

assoc

where Aideal is the ideal free energy, Amono is the contribution to the free
energy due to interactions between monomers or individual segments,
Achain accounts for the possibility of chain formation, and Aassoc is the
hydrogen-bonding intermolecular association. The equation is very versatile and a testament to this is the number of many different versions of
the theory that have been developed over the years. Although differences may be subtle, the main attribute leading to the various versions
is the choice of uid of reference.
In the original SAFT formulation [173,174], molecules are modelled
as associating chains of Lennard-Jones (LJ) segments, while in a simplied treatment [177,178], associating chain molecules of hard-sphere
(HS) segments are considered, SAFT-HS, and the dispersion forces are
treated at the van der Waals mean-eld level (i.e., in the absence of
correlations between particles). The SAFT-HS approach has proved to
be successful for modelling strongly hydrogen-bonded associating systems (such as water + alkanes [179]), although is not adequate for systems where dispersion interactions are dominant. Thus, the introduction
of mixing rules for the dispersion terms [180,181] led to the SAFT-VR approach [182,183], in which the uid is described in terms of associating
4

Refs. [171,172] offer a detailed discussion about these types of theories in comparison
to physical association schemes such as those that constitute the SAFT (and CPA)
formalisms.

chains of segments interacting through an attractive potential of variable


range (VR), typically a square-well, although the extension to Sutherland
and Yukawa potentials was also presented in the original manuscript. In
this version of the theory, the properties of the monomeric segments are
obtained through a Barker and Henderson high-temperature perturbation expansion [184186], truncated at second-order, from a reference
hard-sphere uid. Recently, a new formulation of the SAFT-VR theory
has been developed for Mie potentials (SAFT-VR Mie) [187], which improves the predictive capability of the equation in the description of
second-derivative properties. A third-order perturbation expansion is
further shown to improve the description of the vapour-liquid equilibrium of uids in the critical region [187]. The so-called soft-SAFT [188]
uses a Lennard-Jones reference uid as in the rst SAFT equation, but
with different expressions for the reference uid based on the equation
of Johnson et al. [189], a very accurate equation of state for mixtures of
associating Lennard-Jones chains. A version that has become very popular for chain uids is the perturbed-chain SAFT (PC-SAFT) [190,191] in
which the usual monomer reference system is replaced by a hardsphere-chain reference uid. A high-temperature expansion truncated
at second-order is used to obtain the attractive-chain perturbation contribution, but the formal expressions that would be obtained based on
a Barker and Henderson expansion were simplied using power series
in density with coefcients adjusted to t pure-component properties
of n-alkanes. More recently, other heteronuclear versions of the SAFT
equation have been developed such as the SAFT- approach [192194],
where the equation is coupled with a group contribution formalism.
The SAFT equation of state has been very successfully applied to
study the thermodynamic properties and phase behaviour of complex
uids and mixtures, and the interested reader is referred to the existing
reviews specically dedicated to this theory [171,172,195197].
In the area of present interest, a number of SAFT equations of state
have been applied to model the phase equilibria of asphaltenecontaining systems. In general, these applications consider that
asphaltene precipitation is based on a liquid-liquid equilibrium process,
although both colloidal and molecular models (cf., Section 1) have been
assumed. Following a molecular description, asphaltene clustering and
aggregation is considered a result of the molecular environment, and
not necessarily an initial assumption of the model; models that consider
asphaltene clusters (or small aggregates) have, however, also been used
in a number of works [64,65,155,198208] away from an explicit colloidal depiction. Similarities with the phase behaviour of oligo- or polymer
systems, have directed the focus of attention to molecular asymmetry
and van der Waals interactions, which are assumed to be the features
that dominate their phase behaviour. An advantage of the physicallysound pure-component parameters used in SAFT is that they correlate
well with molecular weight for families of compounds within a homologous series. This allows the estimation of intermolecular
model parameters for asphaltenes based on molecular weight. As
will be mentioned, this approach has been used in numerous studies
[65,198201,203,204,209].
In the work of Wu et al. [63,210], a formulation of the SAFT equation
of state at the level of SAFT-HS combined with colloid theory was used
to describe the precipitation of asphaltene from crude oil. The system is
assumed to consist of asphaltenes and resins (the solutes) interacting in
a structureless medium which has a screening effect over the dispersion
forces between asphaltenes and resins. The medium (oil) is treated at
the level of the McMillan-Mayer theory of solutions [48]5 and its effect
on the dispersion interaction between the components (asphaltenes
and resins) is determined through the use of a potential of mean-force
[212] expressed through the corresponding Hamaker constant [211]
in the oil. The Hamaker constant of the relevant interactions depends
5
Characterised by the use of a potential of mean-force in place of the intermolecular potential between solute molecules, as it is common treatment for colloidal systems [211],
and the pressure is considered to be the osmotic pressure due to the explicit treatment
of the effect of the solutes only.

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

These are the standard parameters used in SAFT frameworks, where the dispersioninteraction parameters are dened based on the Hamaker constants of the pure
components.
7
For example, average molecular weights and densities for pure asphaltene and resins
are used, from which, by considering a packing fraction in the solid range, the diameter of
asphaltene molecules can be derived. This is similar for resins, but with a packing fraction
in the liquid range.
8
As in previous work [63,210], the asphaltene model is considered to be based on a single hard-core sphere of large diameter (17 ), which may more appropriately mimic the
size of a small aggregate formed by three or four asphaltene monomeric units, instead of
a single one.

45
40
35

Precipitation on set

p / MPa

30
25
20
15

Bubble curve
10
5
0
340

360

380

400

420

T/K
Fig. 4. Precipitation onset and bubble point data of a crude oil. The symbols correspond to
experimental data [64], where the blue diamonds are asphaltene precipitation onset
points and the red circles are bubble points. The continuous curve corresponds to calculations of the bubble curve using a volume-shifted Peng-Robinson, whereas the discontinuous curve corresponds to calculations of the precipitation boundary using SAFT-VR in the
context of the McMillan-Mayer theory, following the work of Wu et al. [63,210]. Adapted
from ref. [64].

the nature of the interactions between the major components in


heavy oil. In this work the asymmetry in size between the different
components in the oil is considered to be the dominant mechanism in
the liquid-liquid phase separation. Vargas et al. [204] had already
pointed out existing similarities in the phase behaviour of polymer or
oligomer solutions with that of petroleum systems. Based on these observations, therefore, Artola et al. [209] built a representation of
asphaltenic molecules using polystyrene as a prototype molecule
combining both aromatic rings and alkyl chains. Asphaltene molecules
are then modelled as styrene oligomers or polymers using what the
authors call a polystyrene-asphaltene mapping approach. A simpler
description of the interactions between the components of the mixture
is used, where association interactions are not considered. Crude oil
systems are considered to be described by mixtures of asphaltene (or
polystyrene of a certain molecular weight) and alkane-like components.
The molecular model for the latter is based on what could be understood as a simplied homonuclear group-contribution approach
where alkane molecules (excepting methane) are built using the same
SW-potential parameters. The general features of the phase behaviour
of the asphaltene-containing oil of previous work [64] are described
with a basic lumping scheme where only four alkane-like pseudocomponents plus asphaltene are assumed to represent the crude.
These results are reproduced in Fig. 5. The asphaltene here is considered
100
90
80

Single-phase

70

p / MPa

on that of the oil, which is obtained as a function of the density and composition. A simple model is used for the asphaltenes, considered as attractive hard-core spheres; the resins are modelled as attractive hardcore chains. Association sites are added (two in asphaltene and one in
resin molecules) to mimic aggregation through directional short-range
association. Only asphaltene-asphaltene and asphaltene-resin associations are allowed. The molecular parameters for the asphaltene and
resin components6 are estimated from average properties7 to describe,
mainly qualitatively, experimental observations regarding the effect of
temperature and composition [210]. Because the medium is treated as
a continuum, it is not possible to calculate bubble pressures. In a latter
study, a volume-shifted PR equation of state is used to evaluate
vapour-liquid equilibria and obtain the composition of oil in equilibrium
with a gas phase [63]. Such a treatment is applied to several reservoir
uids based on experimental compositional data, where heavy fractions
are divided into a set of pseudocomponents represented by a molecular
weight and liquid density. Once the composition and density of the medium (liquid phase) are determined, asphaltene precipitation (understood as liquid-liquid equilibrium) is then calculated using the SAFT
molecular model presented earlier [210]. In this study, the estimated
values for the molecular parameters are kept constant with the exception of those relating to the association interactions (i.e. the number of
association sites in the asphaltene molecules and the asphalteneasphaltene and asphaltene-resin association energies), which are adjusted to t experimental onset pressure data for each particular crude oil
considered.
The SAFT-VR equation of state has also been applied to the study of
these systems. One of these applications is the work of BuenrostroGonzalez et al. [64]. In this study, the equation is used in the framework
of the McMillan-Mayer theory following the work of Wu et al. [63,210].
Analogous molecular models for the solutes are used (i.e., asphaltenes
and resins are respectively considered as attractive hard spheres8 and
hard chains). The dispersion interactions considered are based on Sutherland potentials, where the screening effects of the medium over the
energetic interaction of the solutes are, as before, estimated based on
the corresponding Hamaker constant. As in the studies of Wu et al.
[63,210] asphaltene and resins are also considered to associate through
short-ranged anisotropic interactions, which are thought to be dominant over dispersion interactions. The parameters of the model are
based on similar estimations, whereby a good prediction of asphaltene
precipitation with different n-alkanes is obtained through parameter
adjustment to a single titration curve. The same set of parameters is
then able to predict asphaltene precipitation at high pressure, with
only little adjustment of the association parameters to a few data points.
A representative result of these predictions is shown in Fig. 4. The order
of magnitude of the estimated energetic parameters is seen to agree
well with expected values for this type of interactions reported in the
literature [44,93,211,213]. As before, a volume-shifted Peng-Robinson
equation of state is also used here to explicitly model the nature of the
oil based on a considerable number of components and pseudocomponents, and determine the properties of the medium and
bubble-point pressures at conditions of vapour-liquid equilibria.
Another application of the SAFT-VR equation to model these systems
have been presented more recently by Artola et al. [209]. The SAFT
models described up to this point have focused on describing in detail

LLE

60
Precipitation
boundary

50
40
30
20
10
0

VLE

Bubble
curve

VLLE
0

200

400

600

800

T/K
Fig. 5. Constant composition p T uid phase diagram of a crude oil (symbols correspond
to experimental data [64]) modelled with a simple lumping scheme using SAFT-VR
(continuous curves). Adapted from ref. [219].

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

to be represented by a lower molecular weight than in the previous


study [64], assumed to be that of an individual molecule rather than
an aggregate. Comparisons with both bubble points and asphaltene precipitation data are depicted in the phase diagram. Instead of accurately
correlating experimental data, a key difference with other studies is that
of presenting an approach using minimal tting that is able to explain
the position of the bubble curve and the asphaltene-precipitation
boundary within the phase diagram. The bubble curve data is seen to
fall on the boundary of a three-phase vapour-liquid-liquid equilibrium
(VLLE) region, whereas the asphaltene-precipitation data is shown to
be related to the liquid-liquid (LLE) boundary of the diagram. The
PC-SAFT equation of state has also shown to be successful to model
the phase behaviour of asphaltene mixtures, as demonstrated by
the numerous works of Chapman and coworkers [155,198208]
and others [65,163,214216]. In most of these studies, the effects of
size asymmetry and van der Waals interactions are assumed to dominate
the phase behaviour, and association interactions are therefore not considered. Among the various applications, the equation of state has been
used to describe a recombined oil (stock-tank oil with its separator
gas) represented by six [198,200] or seven [199] pseudo-components.9
Molecular parameters are, where possible, derived from correlations
for n-alkanes and polynuclear aromatics with molecular weight.
Parameters for asphaltene, represented by pre-aggregate models of
~1700 g mol1, are obtained from ts to titration data for the refractive
index at the onset of asphaltene precipitation. Binary interaction parameters are adjusted based on vapour-liquid equilibrium properties
of mixtures of representative compounds of each pseudo-component
considered [198,200]. In the works of Gonzalez et al. [199] and Vargas
et al. [204], experimental data for the effect of gas injection over the
onset of asphaltene precipitation and bubble point pressure was
reproduced for a similar recombined oil sample. Another interesting
analysis presented with the PC-SAFT equation of state has been to examine the effect of the addition of carbon dioxide; when CO2 is used
as precipitant, different from other gases, a crossover temperature is
predicted with the equation, below which the addition of CO2 stabilises
the crude oil mixture [203] (see Fig. 6). Such predictions provide insight
to experimental observations [99]. An analysis of the effects of oil-based
drilling mud contamination on the onsets of asphaltene precipitation
and bubble point pressures has also been carried out [201,204], demonstrating that the equation is able to reproduce the experimental trends
that feature a pressure drop with increasing percentages of contamination. Although oil-based mud is known to be a precipitant for
asphaltenes, it dilutes the gaseous components of the oil, reducing
the gas-to-oil ratio and, consequently, the asphaltene instability
onset pressure [204]. The effect of asphaltene polydispersity on its
stability has also been examined [200,201,204,216]. In these studies
the polydisperse asphaltenes are subdivided into a number of pseudocomponents, which are characterised based on their solubilities in
dened fractions of n-alkanes, and where the lighter sub-fraction is
equivalent to what are generically known as resins. Even though mainly
qualitative results are shown, the approach demonstrates that the addition of resins stabilizes the asphaltenes via non-polar interactions only,
increasing the stable region of the phase diagram upon the addition of
precipitant [200]. Parameters for each asphaltene subfraction are
characterised by correlations of benzene derivatives in the study of
Gonzalez et al. [201]. Recently, Punnapala and Vargas [65] have
revisited the characterization procedure used in PC-SAFT so as to reduce
the number of adjustable parameters in the asphaltene model. In
their work, the model parameters for the asphaltene pseudo-fraction
are obtained from correlations for n-alkanes and polyaromatic compounds using the average molecular weight and the aromaticity as adjustable parameters regressed against onset pressures. Generally ten
pseudocomponents are used within this uid characterisation. Studies

Aromatics and resins are grouped into one pseudo-component.

65

Original
reservoir fluid

55

p / MPa

45
+ 20 % CO2

35
+ 10% CO2

25

15
320

370

420

470

520

T/K
Fig. 6. Constant composition p T uid phase diagram of a reservoir uid (symbols correspond to experimental data [220]) modelled using PC-SAFT (blue curves). The red
curves represent the PC-SAFT calculations upon the addition of a 10% mole fraction of
CO2 and the orange curves the corresponding to the addition of a 20% mole fraction of
the same component. In all cases the continuous curves correspond to bubble curve calculations and the discontinuous to the precipitation boundary. Adapted from ref. [203].

combining thermodynamic with reservoir and/or pipeline multiphase


ow modelling have also been carried out. One such study is the work
of Gonzalez et al. [202], in which a commercial software package including transport phenomena is combined with asphaltene and wax thermodynamic modelling using PC-SAFT and a solid-solution model
[217], respectively. In this way, estimations of solid precipitation for
the different conditions of pressure, temperature and composition
along the production lines and during the lifetime of the process can
be achieved. Further integration of molecular thermodynamics within
models to represent the mechanism of asphaltene transport in macroscopic systems has been shown by Vargas et al. [206], in which an approach to model a multi-step process of precipitation, aggregation,
advection and deposition of micro-aggregated asphaltenes in a
wellbore is proposed. This is based on reaction kinetic models to describe the rate of aggregation, precipitation and deposition. The transport of the micro-aggregates through the wellbore is then described
based on a material balance incorporating these phenomena [206].
Very recently, however, Sedghi and Goual [221] combined PC-SAFT
with Gibbs energies of asphaltene association parameters derived
from Molecular Dynamics (MD) simulations of asphaltene molecules
which were based on the average aggregation number of asphaltene
nanoaggregates (as appear in Fig. 1). This combination enabled these
authors to predict the onset of asphaltene precipitation in terms of pressure, temperature and system composition. In the same study, PC-SAFT
was shown to model reasonably well the inhibition of n-heptaneinduced asphaltene precipitation by n-octylphenol using crossassociation parameters calculated using mixing rules developed by
Sandler and coworkers [137,222].
In general, however, molecular parameters are in most cases xed or
estimated based on assumed approximations, the strong foundations of
the theory and the associated physical meaning of the parameters make
assigning values more reliable. At the moment, however, robust models
for the typical components found in crude oils have not been identied
to allow for prediction of thermophysical properties in different reservoir uids. With the limited amount of experimental data currently
available, increasing the complexity of the model may not be always
justied.

4. Other approaches
Although outside the scope of this review, efforts in this eld have
also been presented that use integral equation theories to study
structural properties of asphaltene-containing systems [218,219].

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

5. Conclusions and future trends

Acknowledgements

As a consequence of asphaltenes being linked with various problems


in the petroleum industry, both upstream and downstream, a substantial amount of work in the literature has been concerned with the
prediction of asphaltene precipitation and related phenomena. One of
the main objectives of this article was therefore to delineate the state
of the art with respect to asphaltene stability modelling.
Since the nature of asphaltene self-interactions and those with other
oil components is not well understood, both experimental and theoretical research to gain an insight in this direction will continue to be invaluable. Results from experimental asphaltene science have advanced
knowledge considerably within the past decade, and it is now accepted
that asphaltenes lead an enigmatic colloidal existence based on different
levels of association, as encompassed in the Yen-Mullins model. Interestingly, recent evidence seems to indicate that this model is also consistent
with asphaltene adsorption at water-oil interfaces, in which adsorption
of the monomer units dominates; nanoaggregates and clusters appear
not to adsorb (at least not as strongly, or within the same timescale),
and the overall adsorption behaviour can be treated using a simple
Langmuir equation of state [223]. The signicance of this lies in the role
of asphaltenes in stabilising water-in-crude oil emulsions, for example,
where the build-up of viscous interfacial layers (skins) occurs [224],
which, additionally, may be amenable to theoretical treatments based
on ideas discussed herein relating to solution behaviour. It would
therefore appear that the development of ever-more reliable theoretical
approaches to asphaltene stability should follow from the improved understanding being achieved using an increasingly sophisticated range of
experimental methodologies to study asphaltenes in solution and at
interfaces.
In this respect, uid theories incorporating different levels of detail have shown to be sufcient for reproducing experimental observables found in crude oil systems. The use of lattice uid
theories combined with regular solution theory has been one of the
most frequently applied tools, perhaps motivated by its simplicity,
in that it may be linked to a higher level of empiricism and restricted
versatility. The application of cubic equations to describe asphaltene
precipitation has in general involved the use of a large number of components and pseudocomponents. It would be worth assessing the need
for such detailed lumping schemes in future advances.
Thermodynamic tools can help to ll the gaps in our understanding of
asphaltene precipitation phenomena. In general, molecular-based
equations of state can help in this respect, as their rm foundations
give condence in capturing the correct physical requirements.
However, additional experimental efforts are still essential to exploit
fully the potential of these approaches. Further work in the experimental
characterisation of asphaltenes and their interactions with other oil constituents may justify the computational effort needed to add more detail
into the molecular models. Increasing the currently sparse phase behaviour data available for these systems, covering wider ranges of experimental conditions, will be important to provide a more consistent and
robust development of lumping schemes and adjustment of intermolecular model parameters. Such efforts can certainly contribute to ensuring
that these tools are not used as mere correlations of experimental data
but ultimately to maximize the predictive power of these approaches.
Finally, molecular-based equations with well-dened potential energy functions open up an interesting eld of application through coupling with computer simulations such as molecular dynamics and
Monte Carlo [225]. This would allow the estimated intermolecular
model parameters from an equation of state to be used in molecular
simulations to study structural, interfacial and dynamic properties; the
recent SAFT Mie [187,194] is one such equation and it has been shown
to provide an accurate and reliable direct link with molecular simulations [225228]. Thus, the use of coupled approaches in the eld of
asphaltene stability could be a very promising approach to aid our
understanding of these complex systems.

The authors are grateful for nancial support and permission to publish from BP America. EF thanks A. J. Haslam and E. A. Mller for their
useful comments on this work. We also acknowledge the editor and
the anonymous referees for their valuable suggested improvements.
References
[1] Hughey CA, Rodgers RP, Marshall AG. Resolution of 11,000 compositionally distinct
components in a single electrospray ionization Fourier transform Ion cyclotron
resonance mass spectrum of crude oil. Anal Chem 2002;74(16):41459.
[2] Marshall AG, Rodgers RP. Petroleomics: The next grand challenge for chemical
analysis. Acc Chem Res 2004;37(1):539.
[3] Mullins OC. Review of the molecular structure and aggregation of asphaltenes and
petroleomics. SPE J 2008;13(1):4857.
[4] Maitland GC. Oil and gas production. Curr Opin Colloid Interface Sci 2000;5(56):
30111.
[5] Adams JJ. Asphaltene adsorption, a literature review. Energy Fuel 2014;28(5):
283156.
[6] Leontaritis KJ. Asphaltene deposition: A comprehensive description of problem
manifestations and modelling approaches. Oklahoma City, Oklahoma: SPE Production
Operations Symposium; 1989.
[7] Speight JG. Petroleum asphaltenes Part 1: Asphaltenes, resins and the structure of
petroleum. OGST Rev. IFP, 59(5); 2004 46777.
[8] Shaw JM, Zou XY. Challenges inherent in the development of predictive deposition
tools for asphaltene containing hydrocarbon uids. Pet Sci Technol 2004;22(78):
77386.
[9] Zhang Y, Takanohashi T, Sato S, Saito I, Tanaka R. Observation of glass transition in
asphaltenes. Energy Fuel 2004;18(1):2834.
[10] Maham Y, Chodakowski MG, Zhang XH, Shaw JM. Asphaltene phase behavior: prediction at a crossroads. Fluid Phase Equilib 2005;228-229(1):216.
[11] Sirota EB, Lin MY. Physical behavior of asphaltenes. Energy Fuel 2007;21(5):
280915.
[12] Lastovka V, Fulem M, Becerra M, Shaw JM. A similarity variable for estimating the
heat capacity of solid organic compounds: Part II. Application: Heat capacity
calculation for ill-dened organic solids. Fluid Phase Equilib 2008;268(12):
13441.
[13] Tschierske C. Amphotropic liquid crystals. Curr Opin Colloid Interface Sci 2002;
7(56):35570.
[14] Bagheri SR, Bazyleva A, Gray MR, McCaffrey WC, Shaw JM. Observation of liquid
crystals in heavy petroleum fractions. Energy Fuel 2010;24(8):432732.
[15] Ravey JC, Ducouret G, Espinat D. Asphaltene macrostructure by small angle neutron
scattering. Fuel 1988;67(11):15607.
[16] Carnahan NF, Quintero L, Pfund DM, Fulton JL, Smith RD, Capel M, et al. A small
angle x-ray scattering study of the effect of pressure on the aggregation of
asphaltene fractions in petroleum uids under near-critical solvent conditions.
Langmuir 1993;9(8):203544.
[17] da Silva Ramos AC, Haraguchi L, Notrispe FR, Loh W, Mohamed RS. Interfacial and
colloidal behavior of asphaltenes obtained from Brazilian crude oils. J Petrol Sci Eng
2001;32(24):20116.
[18] Mullins OC, Betancourt SS, Cribbs ME, Dubost FX, Creek JL, Andrews AB, et al. The
colloidal structure of crude oil and the structure of oil reservoirs. Energy Fuel
2007;21(5):278594.
[19] Strausz OP, Mojelsky TW, Lown EM, Kowalewski I, Behar F. Structural features of
Boscan and Duri asphaltenes. Energy Fuel 1999;13(2):22847.
[20] Strausz OP, Mojelsky TW, Faraji F, Lown EM, Peng P. Additional structural details on
Athabasca asphaltene and their ramications. Energy Fuel 1999;13(2):20727.
[21] Strausz OP, Peng P, Murgich J. About the colloidal nature of asphaltenes and the
mw of covalent monomeric units. Energy Fuel 2002;16(4):80922.
[22] Gawrys KL, Matthew Spiecker P, Kilpatrick PK. The role of asphaltene solubility and
chemical composition on asphaltene aggregation. Pet Sci Technol 2003;21(3-4):
46189.
[23] Rakotondradany F, Fenniri H, Rahimi P, Gawrys KL, Kilpatrick PK, Gray MR.
Hexabenzocoronene model compounds for asphaltene fractions: Synthesis and
characterization. Energy Fuel 2006;20(6):243947.
[24] Acevedo S, Castro A, Negrin JG, Fernandez A, Escobar G, Piscitelli V. Relations
between asphaltene structures and their physical and chemical properties: The
rosary-type structure. Energy Fuel 2007;21(4):216575.
[25] Alshareef AH, Scherer A, Tan X, Azyat K, Stryker JM, Tykwinski RR, et al. Formation
of archipelago structures during thermal cracking implicates a chemical mechanism for the formation of petroleum asphaltenes. Energy Fuel 2011;25(5):21306.
[26] Groenzin H, Mullins OC. Molecular size and structure of asphaltenes from various
sources. Energy Fuel 2000;14(3):67784.
[27] Andrews AB, Guerra RE, Mullins OC, Sen PN. Diffusivity of asphaltene molecules by
uorescence correlation spectroscopy. J Phys Chem A 2006;110(26):80937.
[28] Badre S, Goncalves CC, Norinaga K, Gustavson G, Mullins OC. Molecular size and
weight of asphaltene and asphaltene solubility fractions from coals, crude oils
and bitumen. Fuel 2006;85(1):111.
[29] Ruiz-Morales Y, Mullins OC. Polycyclic aromatic hydrocarbons of asphaltenes
analyzed by molecular orbital calculations with optical spectroscopy. Energy Fuel
2007;21(1):25665.
[30] Sabbah H, Morrow AL, Pomerantz AE, Zare RN. Evidence for island structures as the
dominant architecture of asphaltenes. Energy Fuel 2011;25(4):1597604.

10

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

[31] da Silva Oliveira EC, Cunha Neto A, Lacerda Jr V, Ribeiro de Castro EV, Cabral de
Menezes SM. Study of Brazilian asphaltene aggregation by nuclear magnetic
resonance spectroscopy. Fuel 2014;117(1):14651.
[32] Acevedo S, Escobar O, Echevarria L, Gutirrez LB, Mndez B. Structural analysis of
soluble and insoluble fractions of asphaltenes isolated using the PNP method.
Relation between asphaltene structure and solubility. Energy Fuel 2004;
18(2):30511.
[33] Morgan TJ, Millan M, Behrouzi M, Herod AA, Kandiyoti R. On the limitations of UVuorescence spectroscopy in the detection of high-mass hydrocarbon molecules.
Energy Fuel 2005;19(1):1649.
[34] Strausz OP, Safarik I, Lown EM, Morales-Izquierdo A. A critique of asphaltene
uorescence decay and depolarization-based claims about molecular weight and
molecular architecture. Energy Fuel 2008;22(2):115666.
[35] Alvarez-Ramrez F, Ruiz-Morales Y. Island versus archipelago architecture for
asphaltenes: Polycyclic aromatic hydrocarbon dimer theoretical studies. Energy
Fuel 2013;27(4):1791808.
[36] Mullins OC. The modied Yen model. Energy Fuel 2010;24(4):2179207.
[37] Mullins OC. The asphaltenes. Annu Rev Anal Chem 2011;4:393418.
[38] Mullins OC, Sabbah H, Eyssautier J, Pomerantz AE, Barr L, Andrews AB, et al.
Advances in asphaltene science and the Yen-Mullins model. Energy Fuel
2012;26(7):39864003.
[39] Mansoori GA, Vazquez D, Shariaty-Niassar M. Polydispersity of heavy organics in
crude oils and their role in oil well fouling. J Pet Sci Eng 2007;58(34):37590.
[40] Sjblom J, Aske N, Auem IH, Brandal , Havre TE, Sther , et al. Our current understanding of water-in-crude oil emulsions: Recent characterization techniques
and high pressure performance. Adv Colloid Interface Sci 2003;100102(1):
399473.
[41] Groenzin H, Mullins OC. Asphaltene molecular size and structure. J Phys Chem A
1999;103(50):1123745.
[42] Wiehe IA. Two-dimensional solubility parameter mapping of heavy oils. Fuel Sci
Technol Int 1996;14(12):289312.
[43] Buckley JS, Hirasaki GJ, Liu Y, Von Drasek S, Wang JX, Gill BS. Asphaltene precipitation and solvent properties of crude oils. Pet Sci Technol 1998;16(34):25185.
[44] Murgich J. Intermolecular forces in aggregates of asphaltenes and resins. Pet Sci
Technol 2002;20(910):98397.
[45] Wiehe IA. Asphaltene solubility and uid compatibility. Energy Fuel 2012;26(7):
400416.
[46] Taylor SE. The electrodeposition of asphaltenes and implications for asphaltene
structure and stability in crude and residual oils. Fuel 1998;77(8):8218.
[47] Spiecker PM, Gawrys KL, Kilpatrick PK. Effects of petroleum resins on asphaltene
aggregation and water-in-oil emulsion formation. Colloids Surf A 2003;220(13):
927.
[48] McMillan Jr WG, Mayer JE. The statistical thermodynamics of multicomponent
systems. J Chem Phys 1945;13(7):276305.
[49] Andersen SI, Stenby EI. Thermodynamics of asphaltene precipitation and dissolution investigation of temperature and solvent effects. Fuel Sci Technol Int 1996;
14(12):26187.
[50] Godbole SP, Thele KJ, Reinbold EW. EOS modeling and experimental observations
of three-hydrocarbon-phase equilibria. SPE Reserv Eng 1995;10(2):1018.
[51] Nellensteyn FJ. Relation of the micelle to the medium in asphalt. Inst Pet Technol
1928;14:1348.
[52] Pfeiffer JP, Saal RNJ. Asphaltic bitumen as colloid system. J Phys Chem 1940;44(2):
13949.
[53] Ruckenstein E, Nagarajan R. Aggregation of amphiphiles in nonaqueous media. J
Phys Chem 1980;84(11):134958.
[54] Smith GN, Brown P, Rogers SE, Eastoe J. Evidence for a critical micelle concentration
of surfactants in hydrocarbon solvents. Langmuir 2013;29(10):32528.
[55] Hirschberg A, de Jong LNJ, Schipper BA, Meijer JG. Inuence of temperature and
pressure on asphaltene occulation. SPE J 1984;24(3):28393.
[56] Leontaritis KJ, Mansoori GA. Asphaltene occulation during oil production and processing: A thermodynamic colloidal model. San Antonio, Texas: SPE International
Symposium on Oileld Chemistry; 1987.
[57] Victorov AI, Firoozabadi A. Thermodynamic micellization model of asphalteneprecipitation from petroleum uids. AIChE J 1996;42(6):175364.
[58] Pan H, Firoozabadi A. A thermodynamic micellization model for asphaltene
precipitation: Part I: Micellar size and growth. SPE Prod Facil 1998;13(2):
11827.
[59] Nagarajan R, Ruckenstein E. Theory of surfactant self assembly: A predictive
molecular thermodynamic approach. Langmuir 1991;7(12):293469.
[60] Blankschtein D, Thurston GM, Benedek GB. Theory of phase separation in micellar
solutions. Phys Rev Lett 1985;54:9558.
[61] Puvvada S, Blankschtein D. Thermodynamic description of micellization, phase
behavior, and phase separation of aqueous solutions of surfactant mixtures. J
Phys Chem 1992;96(13):556779.
[62] Wu J, Prausnitz JM. Phase equilibria for systems containing hydrocarbons, water,
and salt: An extended Peng-Robinson equation of state. Ind Eng Chem Res 1998;
37(5):163443.
[63] Wu J, Prausnitz JM, Firoozabadi A. Molecular thermodynamics of asphaltene
precipitation in reservoir uids. AIChE J 2000;46(1):197209.
[64] Buenrostro-Gonzalez E, Lira-Galeana C, Gil-Villegas A, Wu J. Asphaltene precipitation in crude oils: Theory and experiments. AIChE J 2004;50(10):255270.
[65] Punnapala S, Vargas FM. Revisiting the PC-SAFT characterization procedure for an
improved asphaltene precipitation prediction. Fuel 2013;108(0):41729.
[66] Merino-Garcia D, Andersen SI. Calorimetric evidence about the application of the
concept of CMC to asphaltene self-association. J Dispers Sci Technol 2005;26(2):
21725.

[67] Goual L, Sedghi M, Zeng H, Mostow F, McFarlane R, Mullins OC. On the formation
and properties of asphaltene nanoaggregates and clusters by DC-conductivity and
centrifugation. Fuel 2011;90(7):248090.
[68] Andreatta G, Bostrom N, Mullins OC. High-Q ultrasonic determination of the critical
nanoaggregate concentration of asphaltenes and the critical micelle concentration
of standard surfactants. Langmuir 2005;21(7):272836.
[69] Majumdar RD, Gerken M, Mikula R, Hazendonk P. Validation of the YenMullins
model of Athabasca oil-sands asphaltenes using solution-state 1H NMR relaxation
and 2D HSQC spectroscopy. Energy Fuel 2013;27(11):652837.
[70] Wu Q, Seifert DJ, Pomerantz AE, Mullins OC, Zare RN. Constant asphaltene molecular and nanoaggregate mass in a gravitationally segregated reservoir. Energy Fuel
2014;28(1):47582.
[71] Eyssautier J, Levitz P, Espinat D, Jestin J, Gummel J, Grillo I, et al. Insight into
asphaltene nanoaggregate structure inferred by small angle neutron and X-ray
scattering. J Phys Chem B 2011;115(21):682737.
[72] Mullins OC, Zuo JY, Seifert D, Zeybek M. Clusters of asphaltene nanoaggregates
observed in oil reservoirs. Energy Fuel 2013;27(4):175261.
[73] Pomerantz AE, Seifert DJ, Bake KD, Craddock PR, Mullins OC, Kodalen BG, et al.
Sulfur chemistry of asphaltenes from a highly compositionally graded oil column.
Energy Fuel 2013;27(8):46048.
[74] Hammami A, Phelps CH, Monger-McClure T, Little TM. Asphaltene precipitation
from live oils: An experimental investigation of onset conditions and reversibility.
Energy Fuel 2000;14(1):148.
[75] Joshi NB, Mullins OC, Jamaluddin A, Creek J, McFadden J. Asphaltene precipitation
from live crude oil. Energy Fuel 2001;15(4):97986.
[76] Rassamdana H, Dabir B, Nematy M, Farhani M, Sahimi M. Asphalt occulation and
deposition: I. The onset of precipitation. AIChE J 1996;42(1):1022.
[77] Wang JX, Brower KR, Buckley JS. Observation of asphaltene destabilization at
elevated temperature and pressure. SPE J 2000;5(4):4205.
[78] Pina A, Mougin P, Behar E. Caractrisation des asphaltnes et modlisation de la
oculation - tat de lArt. OGST - Rev IFP, 61. 3; 2006 31943.
[79] Buckley JS. Predicting the onset of asphaltene precipitation from refractive index
measurements. Energy Fuel 1999;13(2):32832.
[80] Beck J, Svrcek WY, Yarranton HW. Hysteresis in asphaltene precipitation and
redissolution. Energy Fuel 2005;19(3):9447.
[81] Mohamed RS, Loh W, Ramos ACS, Delgado CC, Almeida VR. Reversibility and
inhibition of asphaltene precipitation in Brazilian crude oils. Pet Sci Technol
1999;17(78):87796.
[82] Peramanu S, Singh C, Agrawala M, Yarranton HW. Investigation on the reversibility
of asphaltene precipitation. Energy Fuel 2001;15(4):9107.
[83] Abedini A, Ashoori S, Torabi F, Saki Y, Dinarvand N. Mechanism of the reversibility
of asphaltene precipitation in crude oil. J Pet Sci Eng 2011;78(2):31620.
[84] Kerley GI. Equations of state and gasgas separation in soft-sphere mixtures. J
Chem Phys 1989;91(2):120410.
[85] Frenkel D, Louis AA. Phase separation in binary hard-core mixtures: An exact result.
Phys Rev Lett 1992;68:33635.
[86] Lekkerkerker HNW, Poon WCK, Pusey PN, Stroobants A, Warren PB. Phase behaviour
of colloid + polymer mixtures. Europhys Lett 1992;20(6):559.
[87] Dijkstra M, Frenkel D, Hansen JP. Phase separation in binary hard-core mixtures. J
Chem Phys 1994;101(4):317989.
[88] Bolhuis PG, Louis AA, Hansen JP. Inuence of polymer excluded volume on the
phase-behavior of colloid-polymer mixtures. Phys Rev Lett 2002;89:128302.
[89] Paricaud P, Varga S, Jackson G. Study of the demixing transition in model athermal
mixtures of colloids and exible self-excluding polymers using the thermodynamic
perturbation theory of Wertheim. J Chem Phys 2003;118(18):852536.
[90] Andersen SI, Speight JG. Thermodynamic models for asphaltene solubility and
precipitation. J Pet Sci Eng 1999;22(13):5366.
[91] Monteagudo JEP, Lage PLC, Rajagopal K. Towards a polydisperse molecular thermodynamic model for asphaltene precipitation in live-oil. Fluid Phase Equilib 2001;
187188:44371.
[92] Greeneld ML. Molecular modelling and simulation of asphaltenes and bituminous
materials. Int J Pavement Eng 2011;12(4):32541.
[93] Prausnitz JM, Lichtenthaler RN, de Azevedo EG. Molecular thermodynamics of
uid-phase equilibria. Prentice Hall; 1999.
[94] Flory PJ. Thermodynamics of high polymer solutions. J Chem Phys 1941;9(8):
6601.
[95] Huggins ML. Solutions of long chain compounds. J Chem Phys 1941;9(5):440-440.
[96] Hildebrand JH. Solubility. XII. Regular solutions. J Am Chem Soc 1929;51(1):6680.
[97] Hildebrand JH, Prausnitz JM, Scott RL. Regular and related solutions. New York: Van
Nostrand Reinhold Company; 1970.
[98] Soave G. Equilibrium constants from a modied Redlich-Kwong equation of state.
Chem Eng Sci 1972;27(6):1197203.
[99] Verdier S, Carrier H, Andersen SI, Daridon JL. Study of pressure and temperature effects on asphaltene stability in presence of CO2. Energy Fuel 2006;20(4):158490.
[100] Hamouda AA, Chukwudeme EA, Mirza D. Investigating the effect of CO2 ooding on
asphaltenic oil recovery and reservoir wettability. Energy Fuel 2009;23(2):111827.
[101] da Silva NAE, Oliveira VRD, Costa GMN. Modeling and simulation of asphaltene
precipitation by normal pressure depletion. J Pet Sci Eng 2013;109:12332.
[102] Cimino R, Correra S, Sacomani PA, Carniani C. Thermodynamic modelling for prediction of asphaltene deposition in live oils. San Antonio, Texas: SPE International
Symposium on Oileld Chemistry; 1995.
[103] Wang JX, Buckley JS. A two-component solubility model of the onset of asphaltene
occulation in crude oils. Energy Fuel 2001;15(5):100412.
[104] Alboudwarej H, Akbarzadeh K, Beck J, Svrcek WY, Yarranton HW. Regular solution
model for asphaltene precipitation from bitumens and solvents. AIChE J 2003;
49(11):294856.

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112
[105] Akbarzadeh K, Dhillon A, Svrcek WY, Yarranton HW. Methodology for the characterization and modeling of asphaltene precipitation from heavy oils diluted
with n-alkanes. Energy Fuel 2004;18(5):143441.
[106] Akbarzadeh K, Alboudwarej H, Svrcek WY, Yarranton HW. A generalized regular
solution model for asphaltene precipitation from n-alkane diluted heavy oils and
bitumens. Fluid Phase Equilib 2005;232(12):15970.
[107] Mohammadi AH, Richon D. A monodisperse thermodynamic model for estimating
asphaltene precipitation. AIChE J 2007;53(11):29407.
[108] Funk EW, Prausnitz JM. Thermodynamic properties of liquid mixtures: Aromaticsaturated hydrocarbon systems. Ind Eng Chem 1970;62(9):815.
[109] Pazuki GR, Nikookar M. A modied Flory-Huggins model for prediction of
asphaltenes precipitation in crude oil. Fuel 2006;85(78):10836.
[110] Modi AM, Edalat M. A simplied thermodynamic modelling procedure for
predicting asphaltene precipitation. Fuel 2006;85(1718):261621.
[111] Mohammadi AH, Eslamimanesh A, Richon D. Monodisperse thermodynamic model
based on chemical + Flory-Huggins polymer solution theories for predicting
asphaltene precipitation. Ind Eng Chem Res 2012;51(10):404155.
[112] Zuo JY, Mullins OC, Freed D, Elshahawi H, Dong C, Seifert DJ. Advances in the FloryHuggins-Zuo equation of state for asphaltene gradients and formation evaluation.
Energy Fuel 2013;27(4):172235.
[113] Chen CC. A segment-based local composition model for the Gibbs energy of
polymer solutions. Fluid Phase Equilib 1993;83(0):30112.
[114] Shahebrahimi Y, Zonnouri A. A new combinatorial thermodynamics model for
asphaltene precipitation. J Pet Sci Eng 2013;109:639.
[115] Branco VAM, Mansoori GA, Xavier LCD, Park SJ, Mana H. Asphaltene occulation
and collapse from petroleum uids. J Pet Sci Eng 2001;32(2-4):21730.
[116] Orangi HS, Modarress H, Fazlali A, Namazi MH. Phase behavior of binary mixture of
asphaltene + solvent and ternary mixture of asphaltene + solvent + precipitant.
Fluid Phase Equilib 2006;245(2):11724.
[117] Manshad AK, Edalat M. Application of continuous polydisperse molecular thermodynamics for modeling asphaltene precipitation in crude oil systems. Energy Fuel
2008;22(4):267886.
[118] Mohammadi AH, Richon D. The Scott-Magat polymer theory for determining onset
of precipitation of dissolved asphaltene in the solvent + precipitant solution. Open
Thermodyn J 2008;2:136.
[119] Nakhli H, Alizadeh A, Moqadam MS, Afshari S, Kharrat R, Ghazanfari MH. Monitoring
of asphaltene precipitation: Experimental and modeling study. J Pet Sci Eng 2011;
78(2):38495.
[120] Scott RL, Magat M. The thermodynamics of high-polymer solutions: I. The free
energy of mixing of solvents and polymers of heterogeneous distribution. J Chem
Phys 1945;13(5):1727.
[121] Scott RL. The thermodynamics of high-polymer solutions: II. The solubility and
fractionation of a polymer of heterogeneous distribution. J Chem Phys 1945;
13(5):17887.
[122] Kawanaka S, Park SJ, Mansoori GA. Organic deposition from reservoir uids: a
thermodynamic predictive technique. SPE Reserv Eng 1991;6(2):18592.
[123] Manshad AK, Manshad MK, Rostami H, Mohseni SM, Vaghe M. The association
thermodynamics modeling of asphaltene precipitation. Pet Sci Technol 2014;
32(1):5160.
[124] Yarranton HW, Fox WA, Svrcek WY. Effect of resins on asphaltene self-association
and solubility. Can J Chem Eng 2007;85(5):63542.
[125] Zuo JY, Mullins OC, Freed D, Zhang D. A simple relation between solubility parameters and densities for live reservoir uids. J Chem Eng Data 2010;55(9):29649.
[126] Tharanivasan AK, Yarranton HW, Taylor SD. Application of a regular solution-based
model to asphaltene precipitation from live oils. Energy Fuel 2011;25(2):52838.
[127] Akbarzadeh K, Ayatollahi S, Moshfeghian M, Alboudwarej H, Yarranton HW.
Estimation of SARA fraction properties with the SRK EOS. J Can Pet Technol
2004;43(9):319.
[128] Omidkhah MR, Nikookar M, Pazuki GR, Sahranavard L, Emadi MA. Calculation of
phase behaviour of asphaltene precipitation by using a new EOS. Abu Dhabi,
UAE: Abu Dhabi International Petroleum Exhibition and Conference; 2006.
[129] Pazuki GR, Nikookar M, Omidkhah MR. Application of a new cubic equation of state
to computation of phase behavior of uids and asphaltene precipitation in crude
oil. Fluid Phase Equilib 2007;254(12):428.
[130] Buckley JS, Wang J, Creek JL. Solubility of the least-soluble asphaltenes. In: Mullins
OC, Sheu EY, Hammami A, Marshall AG, editors. Asphaltenes, heavy oils and
petroleomics. New York: Springer; 2007.
[131] Nikookar M, Omidkhah MR, Pazuki GR. Prediction of density and solubility parameter of heavy oils and SARA fractions using Cubic equations of state. Pet Sci Technol
2008;26(16):190412.
[132] Nikookar M, Pazuki GR, Omidkhah MR, Sahranavard L. Modication of a thermodynamic model and an equation of state for accurate calculation of asphaltene
precipitation phase behavior. Fuel 2008;87(1):8591.
[133] Eslamimanesh A, Esmaeilzadeh F. Estimation of solubility parameter by the
modied {ER} equation of state. Fluid Phase Equilib 2010;291(2):14150.
[134] Nikooyeh K, Shaw JM. On the applicability of the regular solution theory to
asphaltene + diluent mixtures. Energy Fuel 2012;26(1):57685.
[135] van der Waals JD. The thermodynamic theory of capillarity under the hypothesis of
a continuous variation of density. Z Phys Chem 1894;13:657725.
[136] Assael MJ, Trusler JPM, Tsolakis TF. Thermophysical properties of uids: An introduction to their prediction. Imperial College Press; 1996.
[137] Orbey H, Sandler SI. Cubic equations of state and their mixing rules. Cambridge Ser.
Chem. Eng.; 1998.
[138] Anderko A. Chapter 4: Cubic and generalized van der waals equations. In: Sengers
JV, Kayser RF, Peters CJ, White HJ, editors. Equations of state for uids and uid
mixtures. Amsterdam: Elsevier; 2000.

11

[139] Goodwin ARH, Sengers JV, editors. Applied thermodynamics of uids. London:
Royal Soc. Chemistry; 2010.
[140] Kontogeorgis GM, Folas GK, editors. Thermodynamic models for industrial
applications. From classical and advanced mixing rules to association theories.
Chichester, West Sussex, UK: John Wiley and Sons Ltd.; 2010.
[141] Peng DY, Robinson DB. A new two-constant equation of state. Ind Eng Chem
Fundam 1976;15(1):5964.
[142] Nghiem LX, Hassam MS, Nutakki R, George AED. Efcient modelling of asphaltene
precipitation. Houston, Texas: SPE Annual Technical Conference and Exhibition;
1993.
[143] Pan H, Firoozabadi A. Thermodynamic micellization model for asphaltene precipitation from reservoir crudes at high pressures and temperatures. San Antonio,
Texas: Society of Petroleum Engineers Annual Technical Conference and Exhibition;
1997.
[144] Kohse BF, Nghiem LX, Maeda H, Ohno K. Modelling phase behaviour including the
effect of pressure and temperature on asphaltene precipitation. Brisbane, Australia:
SPE Asia Pacic Oil and Gas Conference and Exhibition; 2000.
[145] Jamaluddin AKM, Nighswander JN, Kohse BF, El Mahdi A, Binbrek MA, Hogg PF.
Experimental and theoretical assessment of the asphaltene precipitation characteristics of the Sahil eld under a proposed miscible gas injection scheme. Abu Dhabi, UAE:
Abu Dhabi International Petroleum Exhibition and Conference; 2000.
[146] Behar E, Mougin P, Pina A. Integration of asphaltenes occulation modeling into
Athos reservoir simulator. OGST - Rev. IFP, 58(6); 2003 63746.
[147] Sabbagh O, Akbarzadeh K, Badamchi-Zadeh A, Svrcek WY, Yarranton HW. Applying
the PR-EOS to asphaltene precipitation from n-alkane diluted heavy oils and
bitumens. Energy Fuel 2006;20(2):62534.
[148] Fazelipour W, Pope GA, Sepehrnoori K. Development of a fully implicit, parallel,
EOS compositional simulator to model asphaltene precipitation in petroleum
reservoirs. Denver, Colorado, USA: SPE Annual Technical Conference and Exhibition;
2008.
[149] Mahdavi S, Kharrat R. Asphaltene precipitation prediction using a micellization
model based on experimental data. Pet Sci Technol 2011;29(11):113346.
[150] Castellanos Daz O, Modaresghazani J, Satyro MA, Yarranton HW. Modeling the
phase behavior of heavy oil and solvent mixtures. Fluid Phase Equilib 2011;
304(12):7485.
[151] Tavakkoli M, Kharrat R, Masihi M, Ghazanfari MH. Prediction of asphaltene precipitation during pressure depletion and CO2 injection for heavy crude. Pet Sci Technol
2010;28(9):892902.
[152] Victorov AI, Smirnova NA. Description of asphaltene polydispersity and precipitation by means of thermodynamic model of self assembly. Fluid Phase Equilib
1999;158160(0):47180.
[153] Pan H, Firoozabadi A. Thermodynamic micellization model for asphaltene precipitation from reservoir crudes at high pressures and temperatures. SPE Prod Facil
2000;15(1):5865.
[154] Fahim MA, Al-Sahhaf TA, Elkilani AS. Prediction of asphaltene precipitation for
Kuwaiti crude using thermodynamic micellization model. Ind Eng Chem Res
2001;40(12):274856.
[155] Panuganti SR, Vargas FM, Chapman WG. Modeling reservoir connectivity and tar
mat using gravity-induced asphaltene compositional grading. Energy Fuel 2012;
26(5):254857.
[156] Kontogeorgis GM, Voutsas EC, Yakoumis IV, Tassios DP. An equation of state for
associating uids. Ind Eng Chem Res 1996;35(11):43108.
[157] Wertheim MS. Fluids with highly directional attractive forces. I. Statistical
themodynamics. J Stat Phys 1984;35(12):1934.
[158] Wertheim MS. Fluids with highly directional attractive forces. II. Thermodynamic
perturbation theory and integral equations. J Stat Phys 1984;35(12):3547.
[159] Wertheim MS. Fluids with highly directional attractive forces. III. Multiple attractive
sites. J Stat Phys 1986;42(34):45976.
[160] Wertheim MS. Fluids with highly directional attractive forces. IV. Equilibrium
polymerization. J Stat Phys 1986;42:477.
[161] Li Z, Firoozabadi A. Modeling asphaltene precipitation by n-alkanes from heavy oils
and bitumens using cubic-plus-association equation of state. Energy Fuel 2010;
24(2):110613.
[162] Li Z, Firoozabadi A. Cubic-plus-association equation of state for asphaltene precipitation in live oils. Energy Fuel 2010;24(5):295663.
[163] Zhang X, Pedrosa N, Moorwood T. Modeling asphaltene phase behavior: Comparison
of methods for ow assurance studies. Energy Fuel 2012;26(5):261120.
[164] Shirani B, Nikazar M, Mousavi-Dehghani SA. Prediction of asphaltene phase behavior
in live oil with CPA equation of state. Fuel 2012;97(0):8996.
[165] Nasrabadi H, Moortgat J, Firoozabadi A. A new three-phase multicomponent compositional model for asphaltene precipitation using CPA-EOS. Woodlands, TX, USA:
SPE Reservoir Simulation Symposium; 2013.
[166] Vafaie-Sefti M, Mousavi-Dehghani SA, Mohammad-Zadeh M. A simple model for
asphaltene deposition in petroleum mixtures. Fluid Phase Equilib 2003;206(12):
111.
[167] Du JL, Zhang D. A thermodynamic model for the prediction of asphaltene precipitation. Pet Sci Technol 2004;22(78):102333.
[168] Shirani B, Nikazar M, Naseri A, Mousavi-Dehghani SA. Modeling of asphaltene precipitation utilizing association equation of state. Fuel 2012;93(1):5966.
[169] Anderko A. A simple equation of state incorporating association. Fluid Phase Equilib
1989;45(1):3967.
[170] Anderko A, Malanowski S. Calculation of solid-liquid, liquid-liquid and vapor-liquid
equilibria by means of an equation of state incorporating association. Fluid Phase
Equilib 1989;48(1):22341.
[171] Mller EA, Gubbins KE. Molecular-based equations of state for associating uids: A
review of SAFT and related approaches. Ind Eng Chem Res 2001;40(10):2193211.

12

E. Forte, S.E. Taylor / Advances in Colloid and Interface Science 217 (2015) 112

[172] Mller EA, Gubbins KE. Chapter 12: Associating uids and uid mixtures. In:
Sengers JV, Kayser RF, Peters CJ, White HJ, editors. Equations of state for uids
and uid mixtures. Amsterdam: Elsevier; 2000.
[173] Chapman WG, Gubbins KE, Jackson G, Radosz M. SAFT: Equation-of-state solution
model for associating uids. Fluid Phase Equilib 1989;52:318.
[174] Chapman WG, Gubbins KE, Jackson G, Radosz M. New reference equation of state
for associating liquids. Ind Eng Chem Res 1990;29(8):170921.
[175] Wertheim MS. Fluids of dimerizing hard spheres, and uid mixtures of hard
spheres and dispheres. J Chem Phys 1986;85(5):292936.
[176] Wertheim MS. Thermodynamic perturbation theory of polymerization. J Chem
Phys 1987;87(12):732331.
[177] Jackson G, Chapman WG, Gubbins KE. Phase equilibria of associating uids.
Spherical molecules with multiple bonding sites. Mol Phys 1988;65(1):131.
[178] Chapman WG, Jackson G, Gubbins KE. Phase equilibria of associating uids. Chain
molecules with multiple bonding sites. Mol Phys 1988;65(5):105779.
[179] Galindo A, Whitehead PJ, Jackson G, Burgess AN. Predicting the high-pressure
phase equilibria of water + n-alkanes using a simplied SAFT theory with transferable intermolecular interaction parameters. J Phys Chem 1996;100(16):678192.
[180] Huang SH, Radosz M. Equation of state for small, large, polydisperse, and associating
molecules. Ind Eng Chem Res 1990;29(11):228494.
[181] Huang SH, Radosz M. Equation of state for small, large, polydisperse, and associating
molecules: extension to uid mixtures. Ind Eng Chem Res 1991;30(8):19942005.
[182] Gil-Villegas A, Galindo A, Whitehead PJ, Mills SJ, Jackson G, Burgess AN. Statistical
associating uid theory for chain molecules with attractive potentials of variable
range. J Chem Phys 1997;106(10):416886.
[183] Galindo A, Davies LA, Gil-Villegas A, Jackson G. The thermodynamics of mixtures
and the corresponding mixing rules in the SAFT-VR approach for potentials of
variable range. Mol Phys 1998;93(2):24152.
[184] Barker JA, Henderson D. Perturbation theory and equation of state for uids: The
square-well potential. J Chem Phys 1967;47(8):285661.
[185] Barker JA, Henderson D. Perturbation theory and equation of state for uids. II. A
successful theory of liquids. J Chem Phys 1967;47(11):471421.
[186] Barker JA, Henderson D. What is liquid? Understanding the states of matter. Rev
Mod Phys 1976;48(4):587671.
[187] Latte T, Apostolakou A, Avendao C, Galindo A, Adjiman CS, Mller EA, et al.
Accurate statistical associating uid theory for chain molecules formed from Mie
segments. J Chem Phys 2013;139(15):154504.
[188] Blas FJ, Vega LF. Thermodynamic behaviour of homonuclear and heteronuclear
Lennard-Jones chains with association sites from simulation and theory. Mol
Phys 1997;92(1):13550.
[189] Johnson JK, Mller EA, Gubbins KE. Equation of state for Lennard-Jones chains. J
Phys Chem 1994;98(25):64139.
[190] Gross J, Sadowski G. Application of perturbation theory to a hard-chain reference
uid: an equation of state for square-well chains. Fluid Phase Equilib 2000;
168(2):18399.
[191] Gross J, Sadowski G. Perturbed-Chain SAFT: An equation of state based on a perturbation theory for chain molecules. Ind Eng Chem Res 2001;40(4):124460.
[192] Lymperiadis A, Adjiman CS, Galindo A, Jackson G. A group contribution method for
associating chain molecules based on the Statistical Associating Fluid Theory
(SAFT-gamma). J Chem Phys 2007;127(23):23490325.
[193] Lymperiadis A, Adjiman CS, Jackson G, Galindo A. A generalisation of the SAFTgamma group contribution method for groups comprising multiple spherical
segments. Fluid Phase Equilib 2008;274(12):85104.
[194] Papaioannou V, Latte T, Avendano CA, Adjiman CS, Jackson G, Mller EA, et al.
Group contribution methodology based on the statistical associating uid theory
for heteronuclear molecules formed from Mie segments. J Chem Phys 2014;140:
054107.
[195] Economou IG. Statistical associating uid theory: A successful model for the calculation of thermodynamic and phase equilibrium properties of complex uid
mixtures. Ind Eng Chem Res 2002;41(5):95362.
[196] Tan SP, Adidharma H, Radosz M. Recent advances and applications of statistical
associating uid theory. Ind Eng Chem Res 2008;47(21):806382.
[197] McCabe C, Galindo A. Chapter 8: SAFT associating uids and uids mixtures. In:
Goodwin ARH, Sengers JV, editors. Applied thermodynamics of uids. London:
Royal Society of Chemistry; 2010.
[198] Ting PD, Hirasaki GJ, Chapman WG. Modeling of asphaltene phase behavior with
the SAFT equation of state. Pet Sci Technol 2003;21(34):64761.
[199] Gonzalez DL, Ting PD, Hirasaki GJ, Chapman WG. Prediction of asphaltene instability
under gas injection with the PC-SAFT equation of state. Energy Fuel 2005;19(4):
12304.
[200] Ting PD, Gonzalez DL, Hirasaki GJ, Chapman WG. Application of the PC-SAFT equation of state to asphaltene phase behavior. In: Mullins OC, Sheu EY, Hammami A,
Marshall AG, editors. Asphaltenes, heavy oils and petroleomics. New York: Springer;
2007.

[201] Gonzalez DL, Hirasaki GJ, Creek J, Chapman WG. Modeling of asphaltene precipitation due to changes in composition using the perturbed chain statistical associating
uid theory equation of state. Energy Fuel 2007;21(3):123142.
[202] Gonzalez DL, Jamaluddin AKM, Solbakken T, Hirasaki GJ, Chapman WG. Impact of
ow assurance in the development of a deepwater prospect. Anaheim, California,
USA: SPE Annual Technical Conference and Exhibition; 2007.
[203] Gonzalez DL, Vargas FM, Hirasaki GJ, Chapman WG. Modeling study of CO2-induced
asphaltene precipitation. Energy Fuel 2008;22(2):75762.
[204] Vargas FM, Gonzalez DL, Hirasaki GJ, Chapman WG. Modeling asphaltene phase
behavior in crude oil systems using the perturbed chain form of the statistical
associating uid theory (PC-SAFT) equation of state. Energy Fuel 2009;23(3):
11406.
[205] Vargas FM, Gonzalez DL, Creek JL, Wang J, Buckley J, Hirasaki GJ, et al. Development
of a general method for modelling asphaltene stability. Energy Fuel 2009;23(3):
114754.
[206] Vargas FM, Creek JL, Chapman WG. On the development of an asphaltene deposition simulator. Energy Fuel 2010;24(4):22949.
[207] Kurup AS, Vargas FM, Wang J, Buckley J, Creek JL, Subramani HJ, et al. Development
and application of an asphaltene deposition tool (ADEPT) for well bores. Energy
Fuel 2011;25(10):450616.
[208] Kurup AS, Buckley J, Wang J, Subramani HJ, Creek JL, Chapman WG. Asphaltene
deposition tool: Field case application protocol. Houston, Texas, USA: Offshore
Technology Conference; 2012.
[209] Artola PA, Pereira FE, Adjiman CS, Galindo A, Mller EA, Jackson G, et al. Understanding the uid phase behaviour of crude oil: Asphaltene precipitation. Fluid
Phase Equilib 2011;306(1):12936.
[210] Wu J, Prausnitz JM, Firoozabadi A. Molecular-thermodynamic framework for
asphaltene-oil equilibria. AIChE J 1998;44(5):118899.
[211] Israelachvili JN. Intermolecular and Surface Forces. Academic Press; 1992.
[212] McQuarrie DA, editor. Statistical mechanics. New York: Harper- Collins Publishers;
1976.
[213] Ortega-Rodrguez A, Cruz SA, Gil-Villegas A, Guevara- Rodrguez F, Lira-Galeana C.
Molecular view of the asphaltene aggregation behavior in asphaltene-resin
mixtures. Energy Fuel 2003;17(4):11008.
[214] Pedersen KS, Srensen CH. PC-SAFT equation of state applied to petroleum
reservoir uids. Anaheim, California, USA: SPE Annual Technical Conference and
Exhibition; 2007.
[215] Tabatabaei-Nejad SA, Khodapanah E. Application of Chebyshev polynomials to predict phase behavior of uids containing asphaltene and associating components
using {SAFT} equation of state. Fuel 2010;89(9):251121.
[216] Tavakkoli M, Panuganti SR, Taghikhani V, Pishvaie MR, Chapman WG. Understanding the polydisperse behavior of asphaltenes during precipitation. Fuel 2014;
117(1):20617.
[217] Zuo JY, Zhang DD, Ng HJ. An improved thermodynamic model for wax precipitation
from petroleum uids. Chem Eng Sci 2001;56(24):69417.
[218] Ortega-Rodriguez A, Duda Y, Guevara-Rodriguez F, Lira- Galeana C. Stability and
aggregation of asphaltenes in asphaltene resin-solvent mixtures. Energy Fuel
2004;18(3):67481.
[219] Duda Y, Lira-Galeana C. Thermodynamics of asphaltene structure and aggregation.
Fluid Phase Equilib 2006;241(12):25767.
[220] Jamaluddin AKM, Joshi N, Iwere F, Gurpinar O. An investigation of asphalteneinstability
under nitrogen injection. Villahermosa, Mexico: SPE International Petroleum Conference and Exhibition; 2002.
[221] Sedghi M, Goual L. PC-SAFT modeling of asphaltene phase behavior in the presence
of nonionic dispersants. Fluid Phase Equilib 2014;369(1):8694.
[222] Wolbach JP, Sandler SI. Using molecular orbital calculations to describe the phase
behavior of cross-associating mixtures. Ind Eng Chem Res 1998;37(8):291728.
[223] Rane JP, Pauchard V, Couzis A, Banerjee S. Interfacial rheology of asphaltenes at oilwater interfaces and interpretation of the equation of state. Langmuir 2013;
29(15):47509.
[224] Taylor SE. Resolving crude oil emulsions. Chem Ind 1992;20:7703.
[225] Mller EA, Jackson G. Force eld parameters from the SAFT- equation of state for
use in coarse-grained molecular simulations. Annu Rev Chem Biomol Eng 2014;5:
40527.
[226] Avendao C, Latte T, Galindo A, Adjiman CS, Jackson G, Mller EA. SAFT- force
eld for the simulation of molecular uids. 1. A single-site coarse grained model
of carbon dioxide. J Phys Chem B 2011;115(38):1115469.
[227] Avendao C, Latte T, Adjiman CS, Galindo A, Mller EA, Jackson G. SAFT- force
eld for the simulation of molecular uids: 2. Coarse grained models of greenhouse
gases, refrigerants, and long alkanes. J Phys Chem B 2013;117(9):271733.
[228] Latte T, Avendao C, Papaioannou V, Galindo A, Adjiman CS, Jackson G, et al.
SAFT- force eld for the simulation of molecular uids. 3. Coarse-grained models
of benzene and hetero-group models of n-decylbenzene. J Phys Chem B 2012;
110(11-12):1189203.

Das könnte Ihnen auch gefallen