You are on page 1of 11

1676 Energy & Fuels 2004, 18, 1676-1686

Morphology of Aggregated Asphaltene Structural Models


Ä lvarez-Ramı́rez, and J. M. Martı́nez-Magadán
J. H. Pacheco-Sánchez,* F. A
Programa de Ingenierı́a Molecular, Instituto Mexicano del Petróleo,
Eje Central Lázaro Cárdenas 152, México D.F. 07730, México

Received April 12, 2004. Revised Manuscript Received July 8, 2004

Aggregated asphaltene structural models have been generated through a molecular simulation
geometry optimization process, using periodic boundary conditions. This methodology has been
validated by first applying it to a pure aromatic system. Initially, a random distribution of 35
molecules was chosen and a geometry optimization process was performed, allowing the cell
dimensions to vary without restrictions. The structure factor (S(k)) of an optimized final cell was
obtained and compared with experimental results, and the agreement between theoretical and
experimental S(k) profiles was satisfactory. This methodology was next used in the analysis of
the morphology of 32 asphaltene model molecules and their aromatic cores; asphaltene model
molecules were taken from literature. It is remarkable that face-to-face stacking of asphaltene
aggregates was observed, as well as π-offset and T-shaped stacking geometries. Finally, the effect
of aliphatic chains on the aggregates was also analyzed.

1. Introduction whereas the RDF is related to the observed structure


Aggregated asphaltene structural models would be factor S(k) by a Fourier transformation.7-12 In a pio-
useful in developing further understanding of important neering work, Yen et al.3 performed X-ray diffraction
problems in the petroleum industry, such as the ag- (XRD) measurements for asphaltenes in powdered solid
gregation, flocculation, and precipitation of asphaltenes. samples coming from Kuwait visbreaker tar oil. They
An operative definition of asphaltenes is that they are have compared their experimental S(k) profile for the
insoluble in low-molecular-weight n-alkanes but soluble aromatic component of their asphaltenes to that com-
in aromatic solvents. It is well-known that asphaltenes puted for a blend of five polynuclear aromatic com-
self-associate in stacked aggregates, which form clus- pounds of known structures.13 The S(k) profile and the
ters.1,2 However, the morphology of these clusters is not peak positions agree quite reasonably with that ob-
yet clearly known. The most common morphology tained by Diamond.13 Based on these results, they have
considered until now is the face-to-face type. In the proposed a model constituted by aromatic sheets that
literature, these interactions are usually only referenced are associated in a stacked cluster for an asphaltene
as aromatic-aromatic interactions, without going into solid phase. A very similar morphology that has been
the fundamental molecular details. Some efforts in extended to aggregates of asphaltene particles was
determining the morphologies of asphaltene-aggregated experimentally obtained by Dickie and Yen.2 Asphalt-
systems were performed by Yen et al.,3 Dickie and Yen,2 enes appear as unitary stacking sheets that are com-
Wiehe and Liang,4 and Evdokimov et al.5,6 posed of a highly condensed polynuclear system of
The morphology of these clusters is amorphous in aromatic rings bearing alkyl side chains. They proposed
nature. Therefore, a possible way to characterize them that asphaltene association occurs via a stacking of 3-6
geometrically is either through the radial distribution unitary sheets through π-π interactions. They define
function (RDF) or through the structure factor S(k); this these entities as unitary cells or particles, indicating
is also called isotropic scattering. The RDF is defined that the associations of such particles form micelles.
as the number of atoms lying at distances between r The morphology of polyaromatic compounds and
and r + dr from the center of an arbitrary origin atom,7 asphaltene aggregates have been studied theoretically
as induced aggregation in a vacuum,14,15 as spontaneous
* Author to whom correspondence should be addressed. E-mail
address: hpacheco@imp.mx. (8) McQuarrie, D. A. Statistical Mechanics; Harper and Row: New
(1) Sheu, E. Y. In Asphaltenes: Fundamentals and Applications; York, 1973.
Sheu, E. Y., Mullins, O. C., Eds.; Plenum Press: New York, 1995. (9) Hansen, J. P.; McDonald, I. R. Theory of Simple Liquids;
(2) Dickie, J. P.; Yen, T. F. Anal. Chem. 1967, 39, 1847-1852. Academic Press: New York, 1986.
(3) Yen, T. F.; Erdman, J. G.; Pollack, S. S. Anal. Chem. 1961, 33, (10) Tildesley, M. P.; Allen, D. J. Computer Simulation of Liquids;
1587-1594. Clarendon Press: Oxford, U.K., 1987.
(4) Wiehe, I. A.; Liang, K. S. Fluid Phase Equilib. 1996, 117, 201- (11) Lee, L. L. Molecular Thermodynamics of Nonideal Fluids;
210. Butterworth: Boston, 1988.
(5) Evdokimov, I. N.; Eliseev, N. Y.; Akhmetov, B. R. J. Pet. Sci. (12) Evans, D. F.; Wennerström, H. The Colloidal Domain; VCH
Eng. 2003, 37, 135-143. Publishers: New York, 1994.
(6) Evdokimov, I. N.; Eliseev, N. Y.; Akhmetov, B. R. J. Pet. Sci. (13) Diamond, R. Acta Crystallogr. 1957, 10, 359-364.
Eng. 2003, 37, 143-152. (14) Brandt, H. C. A.; Hendriks, E. M.; Michels, M. A. J.; Visser, F.
(7) ) Elliott, S. R. Physics of Amorphous Materials; Longman J. Phys. Chem. 1995, 99, 10430-10432.
Scientific and Technical: New York, 1990; Chapter 3. (15) Rogel, E. Colloids Surf. 1995, 104, 85-93.

10.1021/ef049911a CCC: $27.50 © 2004 American Chemical Society


Published on Web 08/19/2004
Aggregated Asphaltene Structural Models Energy & Fuels, Vol. 18, No. 6, 2004 1677

aggregation in a vacuum,16 and in the presence of


solvents15,17,18 by classical molecular dynamics (MD).19
An important result is that the structure of aggregates
could not be formed by considering only stacking
through π-π interactions. Two additional orientations,
such as those mentioned by Hunter and Saunders20 and
Leach,21 for aromatic-aromatic interactions seem to be
necessary. These authors summarized the results of
their investigations as follows: main orientations are
face-to-face geometry (π-π interactions), edge-on or
T-shaped geometry (π-σ interaction), and offset π-stacked
geometry (σ-σ interactions). Therefore, it was consid-
ered that these orientations can lead to different forms
of the asphaltene aggregates.2,16,22 These findings agree
with the results reported by Sheu,17 based on molecular
simulations for 64 asphaltene molecules, the structures
of which range in size from 3 aromatic rings to 11
aromatic rings, for simulations conducted in toluene. He
also found that asphaltene aggregates are formed not
only through face-to-face stacking but also through other
types of asphaltene clustering; such clustering is much
looser and rather irregular in appearance.
Generally, it is widely accepted that every designed
asphaltene molecule includes an aromatic component,
its own quantity of heteroatoms (such as N, S, and O),
and one or more aliphatic chains linked to the aromatic
region.23-26 Some asphaltene molecular models that
have the latter description have been proposed in the
literature. Four asphaltene structure model molecules
were selected from the following references: Groenzin
and Mullins,27 Speight and Moschopedis,28,29 Zajac et
al.,23 and Murgich et al.18 These molecules are repre-
Figure 1. Depiction of two asphaltene molecule structures:
sented in Figures 1 and 2. The corresponding asphalt- one designed by Groenzin and Mullins27 from Californian crude
enes of these references were extracted from Califor- oil, and the other was designed by Speight and Moschopedis28
nian, Venezuelan, Mayan, and Venezuelan crude oils, for Venezuelan crude oil.
respectively. Some physicochemical properties of these
molecules are given in Table 1. It is well-known that
petroleum fluids are comprised of asphaltene polydis-
perse systems. One of the main limits on the asphaltene
theoretical analysis lies in the diversity of structures
in which asphaltenes can exist. The difficulty associated
with the construction of molecular models of asphalt-

(16) Pacheco-Sánchez, J. H.; Zaragoza, P. I.; Martı́nez-Magadán, J.


M. Energy Fuels 2003, 17, 1346-1355.
(17) Sheu, E. Y. In Structures and Dynamics of Asphaltenes; Mullins,
O. C., Sheu, E. Y., Eds.; Plenum Press: New York, 1998; Chapter IV.
(18) Murgich, J.; Rodrı́guez, J.; Aray, Y. Energy Fuels 1996, 10, 68-
76.
(19) Cerius2 software, Molecular Simulations, Inc. (MSI) (now
Accelrys), San Diego, CA.
(20) Hunter, C. A.; Saunders, J. K. M. J. Am. Chem. Soc. 1990, 112,
2008-2010.
(21) Leach, A. R. Molecular Modeling; Addison-Wesley Longman,
Ltd.: Singapore, 1996.
(22) Takanohashi, T.; Sato, S.; Tanaka, R. Pet. Sci. Technol. 2003,
21, 491-505.
(23) Zajac, G. W.; Sethi, N. K.; Joseph, J. T. Scanning Microsc. 1994,
8, 463-470.
(24) Speight, J. G. The Chemistry and Technology of Petroleum;
Marcel Dekker: New York, 1999; Chapter X.
(25) Cimino, R.; Correra, S.; Del Bianco, A. In Asphaltenes: Fun-
damentals and Applications; Sheu, E. Y., Mullins, O. C., Eds.; Plenum
Press: New York, 1995; Chapter III.
(26) Pfeiffer, J. Ph.; Saal, R. N. J. Presented at the Sixteenth Colloid
Symposium, Stanford University, Palo Alto, CA, July 6-8, 1939; pp
139-165.
(27) Groenzin, H. G.; Mullins, O. C. Energy Fuels 2000, 14, 677-
684.
(28) Speight, J. G.; Moschopedis, S. E. Prepr.sAm. Chem. Soc., Div.
Pet. Chem. 1979, 24, 910-923. Figure 2. Depiction of two asphaltene molecule structures:
(29) Speight, J. G. The Chemistry and Technology of Petroleum; one designed by Zajac et al.23 from Mayan crude oil, and the
Marcel Dekker: New York, 1991. other exhibited by Murgich et al.18 for Venezuelan crude oil.
1678 Energy & Fuels, Vol. 18, No. 6, 2004 Pacheco-Sánchez et al.

Table 1. Physicochemical Properties of Each of the Asphaltene Molecules Used in This Work
value
property Mullins Speight Zajac Murgich
formula H98C72S1 H79C80N2O1S2 H63C57N1S1 H159C138N3O2S2
molecular weight (amu) 995.64 1149.67 794.2 1955.95
molecular volume (Å3) 794.6 892.7 637.3 1574.6
elemental analysis (%)
C 86.86 83.58 86.2 84.74
H 9.92 7.01 8 8.19
N 2.44 1.76 2.15
O 1.39 1.64
S 3.22 5.58 4.04 3.28
number of fused aromatic rings 7 14 9 24
CA 30 47 29 70
CS 42 33 28 68
HA 11 9 6 19
HS 87 70 57 140

enes is well-recognized, and the difficulty is even more modeling, using molecular simulations for a set of one
severe for a polydisperse system of a specific crude oil. type of asphaltene molecule.
However, a reasonable starting guess is that the amor-
phous solid asphaltene phase is mainly comprised of an 2. Methodology
average size molecule, as evidenced from experimental
The methodology used is based on force-field concepts; it is
work, which varies for each type of crude oil. Zajac et
an analytical function that is mainly composed of two types
al.23 proposed three asphaltene model molecules for of terms. The first terms are associated with bond-interaction
Mayan crude oil; they are built by rearrangement of energies, such as torsion, bending, and stretching. The second
some aromatic rings or some aliphatic chains for the terms are associated with nonbonded interactions, such as
asphaltene structures they proposed. As a consequence, Coulombic and van der Waals forces.31-34 Because of the force-
a polydisperse system can be constructed by a set of field features, this methodology is dependent on the molecular
monomer asphaltene molecular models in a vacuum, characterization of the crude oil under study. One objective
which, as a first approach, should be a good model to behind MD is to find a stable molecular configuration where
simulate asphaltene aggregates, as well as to investi- that configuration is located in a local minimum of the
gate how they agglomerate in an amorphous solid phase. potential energy surface around the initial configuration.
However, as a consequence of the high molecular weight, the
Self-association of covalent asphaltene model mol-
asphaltene motion in a MD simulation is known to be very
ecules was recently simulated through molecular simu- slow. For this reason, finding the local minimum energy
lations by Pacheco et al.16,30 The stacked asphaltene requires hundreds of picoseconds.
molecules were observed in the form of dimers, trimers, First, a model system constituted by a blend of 35 molecules
tetramers, and pentamers. The morphologies of those composed of equal quantities of five different polynuclear
aggregates were proposed as a stacking of asphaltenes aromatic molecules was set up, as proposed by Diamond.13
not only assembled face to face but also in T-shaped and These 35 molecules were randomly distributed in a simulation
offset orientations, in agreement with the results of cell using the Amorphous Cell program.35 The cell dimensions
Hunter and Saunders20 and Leach.21 Furthermore, (a, b, and c) are all equal to 63.25 Å, and the initial density in
using the same Groenzin and Mullins model utilized in the cell is equal to 0.1 g/cm3. Second, a complete cell geometry
optimization process was conducted, allowing free movement
this work, the interaction energy between two asphalt-
of molecules, including internal lengths and angles, bringing
ene models was calculated by minimizing the energy at as a consequence, the effect of a comprised cell with a
different distances, which, indeed, provided the energy representative density. The final density for the relaxed
as a function of distance. Similar morphologies can be structure was 1.39 g/cm3. This value is larger than the usual
observed in other systems,12 where it is also possible to experimental values.36 This behavior can be explained by the
observe face-to-face (FF), edge-to-face (EF), and edge- lack of aliphatic chains pendant to the aromatic cores. The
to-edge (EE) clay-particle associations. COMPASS98_02 force field35 was chosen because it was
In this work, a method to generate stable asphaltene designed for organic and inorganic molecules, and it has been
aggregates is presented; this methodology uses a peri- extensively applied to these types of systems with successful
results. An internal stress of 1 × 10-4 GPa was selected as
odic cell ensemble of 32 asphaltene molecules. To find
the criterion of convergence.
a stable structure, an optimization process was per- Figure 3 shows the same final structure from two different
formed, using a force-field method. Our objective is points of view, highlighting different molecules in each one.
mainly (i) to find morphologies of asphaltene aggregates The relaxed structure displays the presence of stacking, as
that properly describe the experimentally reported S(k) expected. Stacked polydisperse domain sheets were formed in
profile, (ii) to analyze the geometries through which the the cell; they can be seen in Figure 3a and b, where the
aggregates are formed, and (iii) to study the effect of
the aliphatic chains, on the aggregated structure, (31) Sun, H.; Rigby, D. Spectrochim. Acta A 1997, 53, 1301-1323.
(32) Rigby, D.; Sun, H.; Eichinger, B. E. Polym. Int. 1997, 44, 311-
through the evolution of the structure factor. To this 330.
end, a comparison between predicted and experimental (33) Sun, H.; Ren, P.; Fried, J. R. Comput. Theor. Polym. Sci. 1998,
structure factors is examined for asphaltene aggregation 8 (1-2), 229-246.
(34) Sun, H. J. Phys. Chem. B 1998, 102, 7338-7364.
(35) Force Field-Based Simulations; MSI, Inc.: San Diego, CA, 1997;
(30) Pacheco-Sánchez, J. H.; Alvarez-Ramı́rez, F.; Martı́nez-Ma- pp 29, 265-268.
gadán, J. M. Prepr.sAm. Chem. Soc., Div. Pet. Chem. 2003, 48, 71- (36) Rogacheva, O. V.; Rimaev, R. N.; Gubaidullin, V. Z.; Khakimov,
73. D. K. Colloid J. USSR 1980, 490-493.
Aggregated Asphaltene Structural Models Energy & Fuels, Vol. 18, No. 6, 2004 1679

Figure 4. Comparison between the structure factor S(k)


obtained in our simulation and those S(k) profiles reported by
Yen et al.3
asphaltene models, and the results will be compared against
at least one of those curves of the S(k) profile for several
asphaltenes reported by Yen et al.3

3. Asphaltene Aggregation Structures


Using the same methodology as before for aromatic
molecules proposed by Yen et al,3 four periodic cells with
32 asphaltene model molecules were built. The cells
Figure 3. Final structure of 35 aromatic molecules obtained were built using the four asphaltene molecular models
after geometry optimization process. Panels a and b show two
angles of different stacking structures of the same final relaxed
mentioned previously; for simplicity, let us call them
cell; the macrostructure of asphaltic material representing just Mullins, Speight, Zajac, and Murgich. The minimization
the aromatic portion is shown in panel b. Panel c shows a of the energy was performed by allowing the relaxation
planar representation of Figure 3b, which is a typical Dickie of the cell dimensions to obtain spontaneous self-
and Yen2 model for aromatic components without considering aggregation of the asphaltene molecules. Asphaltene
aliphatic chains. self-aggregation in crude oil is believed to occur as a
spontaneous aggregation, which can be investigated
domains are themselves constituted by different types of using a geometry optimization process, such as that in
molecules, which give the polydisperse character to the formed this work. An initial density of ∼0.1 g/cm3 was chosen
aggregate. The generated aggregates are formed of dimer,
for all the cells. During the optimization process, the
trimer, and tetramer domains. This description is consistent
with the results found by Dickie and Yen.2 One very important equilibrium configuration of the system gradually de-
detail, in the present study, is that the face-to-face stacking creases the cell length, leading to a squeezed cell. The
geometrical orientation, as well as the offset π-stacked geom- cells that have been built in this way have the following
etry, was observed. final densities: 0.98 g/cm3 for the Mullins model mol-
The positions of every molecule and atom can be determined ecules, 1.04 g/cm3 for the Speight model molecules, 1.02
in the model, and then it is possible to get structural properties g/cm3 for the Zajac model molecules, and 0.69 g/cm3 for
as the spherically averaged distribution of interparticle vector the Murgich model molecules.
lengths (the radial distribution function, RDF) and its inverse 3.1. Mullins Model Case. The final structure of the
Fourier transformation (the structural factor, S(k)) within the cell that has been built using the Mullins asphaltene
model. The structure factor was obtained for this model model shows a very small number of stacking forma-
system, and good agreement was observed between the
tions, compared to that of aromatic molecules. This is
theoretical and experimental S(k) profile (Figure 4). In this
graph, we rescaled the abscissa axis (sin θ)/λ of Yen et al.3 by attributed to the presence of aliphatic chains, which
a value of 4π, such that this axis now is denoted as k, where hinder a close interaction between the aromatic cores
k ) 4π(sin θ)/λ. The dotted curve in Figure 4 represents the of the asphaltenes, as shown in Figure 5a.
experimental curve of the amorphous blend of five aromatic In particular, both the aliphatic and aromatic regions
compounds of known structure, according to those results images were isolated. Figure 5b shows the aliphatic
computed by Diamond.13 He computed the intensity of X-rays region image, which displays a homogeneous distribu-
diffracted from randomly oriented, perfect aromatic molecules tion over the cell. The aromatic region image, in Figure
of various sizes, using the Debye RDF. The dashed curve in 5c, shows two highlighted stacking formations: one in
Figure 4 represents approximately the experimental XRD an offset π-stacked geometry, and the other in a face-
pattern of the aromatic clusters in the asphaltene computed
to-face geometry. The stacking behavior can be ex-
by Yen et al.3 The solid curve in Figure 4 represents our own
simulation of the same amorphous blend of five aromatic plained by the repulsive steric effect of the aliphatic
compounds of known structure. Therefore, the structural chains, which decreases as the length of these chains
model in Figure 3 can be considered to be a good approximation diminishes. Asphaltene face-to-face stacking has a more-
of the real system. Because of these agreements, this meth- probable existence when short aliphatic chains overhang
odology will be extrapolated to different structure cases of the from the aromatic rings.
1680 Energy & Fuels, Vol. 18, No. 6, 2004 Pacheco-Sánchez et al.

Figure 6. Comparison between the asphaltene S(k) profile


of our relaxed cell and two aromatic S(k) profiles reported by
Yen et al.3

However, this structure shows an S(k) profile that is


in agreement with some experiments reported in the
literature. In particular, a direct comparison between
the S(k) profile associated with our structure and that
of the Ragusa crude oil asphaltene fraction experimen-
Figure 5. (a) Cell showing the final asphaltene structure of tally reported by Yen et al.3 is presented (Figure 7),
the Mullins model. (b) Isolated aliphatic region of the complete which corresponds to a more accurate comparison than
final structure. (c) Isolated aromatic region, where some the other S(k) profile of the experimentally obtained
stacking were highlighted. asphaltenes.
For an optimized configuration, the following two The fact that the model is able to reproduce the three
parameters of the simulation cell allow us to discuss experimental S(k) peaks justifies that the methodology
the diminishing of the stacking. The free-volume per- developed here produces an adequate asphaltene ag-
centage in Figure 5a is 34.68%, which means that gregate structure. Therefore, based on the morphology
asphaltene monomers are highly condensed in the cell. of the model, it is possible to infer the existence of more
The ratio of surface areas between the aromatic region than one type of relative orientation between the
and the aliphatic region is ∼0.41, which means that the asphaltene molecules. These orientations are not re-
aromatic portion is bigger than the aliphatic portion. stricted to FF; some other orientations also are included.
The small amount of stacking, in this case, is due to To analyze the distance between different aromatic
the aliphatic chain size and number, with respect to the regions in this aggregate, the RDF has been calculated,
aromatic core size, which does not allow a close interac- using the same atom in each one of the asphaltene
tion between the aromatic components of the asphalt- molecules of this system, where this atom is located in
ene. These chains form walls between the asphaltene the aromatic region; the first peak on the RDF corre-
nearest neighbors, which hinder the direct π-π interac- sponds to a minimum distance of 4.15 Å, which was
tion. The increased separation generated by the pres- measured between the centers of the aromatic region.
ence of the chains is reflected directly in the S(k) profile. The main reason this distance is deviated almost 15%
This phenomenon can be observed when the S(k) from the distance corresponding to the optimum binding
profiles for the aromatic and asphaltene cases are energy (3.55-3.7 Å) is because the distance between
compared. The presence of these chains produces a centers of aromatic regions is not the shortest one in
discrepancy between the aromatic and asphaltene S(k) each case, which reflects the effect of the aliphatic
profiles, which is reflected in the number of peaks and chains that is due to the use of a complete asphaltene
their positions, for k < 2.5 Å-1. It was observed that model.
the first peak (the nearest to zero) for the aromatic case It is important to mention herein that Yen et al.3 also
is still present in the asphaltene S(k) profile; however, measured a linear size distance of 16-20 Å for the
it is shifted to a k value of <2 Å-1. A new peak, at k ≈ stacking of four-unit asphaltene sheets. From this
0.5 Å-1, is observed; this peak is not present in the result, the following facts can be elucidated:
aromatic case, as shown in Figure 6. What we are (1) The distance 3.55-3.7 Å is underestimated, in the
testing in Figure 6 generally is that the S(k) profile of sense that corresponds only with the binding energy of
a complete asphaltene model has an important differ- asphaltene molecules; however, the size of the well
ence when is compared with the S(k) profile of aromatic potential includes a wider interaction range than this.16
cores. We then must compare simulated S(k) profiles of (2) The distance 16-20 Å corresponds to asphaltenes
whole asphaltene aggregates against experimental S(k) that are strongly influenced by aliphatic chains. Fur-
profiles of whole asphaltene aggregates. The other thermore, it is remarkable that Espinat et al.37 mea-
comparison must be between theoretical and experi-
(37) Espinat, D.; Rosenberg, E.; Scarsella, M.; Barre, L.; Fenistein,
mental S(k) profiles of aromatic cores of the asphaltene D.; Broseta, D. In Structures and Dynamics of Asphaltenes; Mullins,
aggregates. O. C., Sheu, E. Y., Eds.; Plenum Press: New York, 1998; Chapter V.
Aggregated Asphaltene Structural Models Energy & Fuels, Vol. 18, No. 6, 2004 1681

Figure 7. S(k) profile for the Mullins aggregate model, compared with two experimental S(k) profiles, and the S(k) profile obtained
from a relaxed aggregation model built only with the aromatic cores (AC) of Mullins asphaltenes.

sured a linear size distance of four unit sheets of 14-


28 Å for the stacking of asphaltenes.
(3) Knowing that the distance between monomers in
a dimer is 3.55-3.7 Å, the linear size of a tetramer (16-
20 Å or 14-28 Å) probably indicates two dimers
interacting between themselves, with aliphatic chains
in the middle of them providing an interaction distance
of 8.9-12.9 Å or 6.9-20.9 Å, respectively.
In addition, Figure 7 shows a comparison between the
S(k) profile of the complete asphaltene aggregate model
and that obtained from an aggregate model that was
built using only the aromatic Mullins asphaltene core,
which will be analyzed in the next section.
3.2. Speight Model Case. The relaxed aggregate
structure of the cell built with the Speight model shows
an abundance of stacking. However, the largest amount
of stacking is in an offset π-stacked geometry. In this
particular geometry, at least one structure that con-
tains five members has been observed (Figure 8). It has
also highlighted two more stacking geometries: one is
T-shaped, and the other one is a face-to-face geometry.
It can be attributed to the increment of stacking in the
Speight aggregate, compared with the Mullins model
case, to the highest absolute area of the aromatic
section. This gives greater contact area than in the
Mullins model case. However, this contact is not so Figure 8. (a) Cell showing the final asphaltene structure of
great, because the aliphatic chains do not allow a close the Speight model. (b) Isolated aliphatic region of the complete
interaction. A ratio of 0.64 between the aromatic surface final structure. (c) Isolated aromatic region, where one offset
area and the aliphatic surface area was observed. The π-stacking with five members, one dimer in T-shaped stacking,
and one dimer in face-to-face stacking were highlighted.
molecules are less flexible, because of the rigidity of the
aromatic sheets and the fact that, in this case, they are molecules can self-aggregate: face to face, offset
larger than those in the Mullins model case. This effect π-stacked, and T-shaped. The irregular stacking is
produces an increase in the free-volume percentage, attributed to the concavity obtained by the aromatic
which, in this case, is 37.80%. For this aggregate, the cores at the end of the energy optimization of the
minimum distance between the aromatic cores is 3.65 molecular simulation.
Å and, it is the first peak on the RDF. The ratio between the aromatic and aliphatic surface
The comparison between the S(k) profile for the areas is 0.65, which is almost equal to that observed in
Speight aggregate and the experimental S(k) profile the Speight case. However, the absolute size of the
shows good agreement in the position of the peaks. The aromatic area is smaller for the Zajac molecule than the
best agreement is found with the Ragusa asphaltene Speight molecule and is larger than that for the Mullins
crude oil (Figure 9). molecule, bringing, as a consequence, a higher stacking
3.3. Zajac Model Case. The Zajac molecule is a formation than that observed in the Mullins model case
singular one, because it has only one long aliphatic but smaller than that in the Speight model case. The
chain. The relaxed final aggregate structure exhibits free volume of 36.96% indicates the role of the only one
molecules with a great curvature in the aromatic region aliphatic chain: that, despite the aromatic region, in
that generally move away from these regions, stopping this case, the free volume is bigger than that in the
a higher stacking formation. The aggregate shows an Mullins model case. Note that their free volumes have
irregular stacking structure. However, Figure 10 il- almost the same magnitude. The minimum distance
lustrates at least three geometries in which these between the aromatic cores, for this case, is 3.25 Å and,
1682 Energy & Fuels, Vol. 18, No. 6, 2004 Pacheco-Sánchez et al.

Figure 9. S(k) profile for the Speight aggregate model, compared with two experimental S(k) profiles, and the S(k) profile obtained
from a relaxed aggregation model built only with the aromatic cores (AC) of Speight asphaltenes.

3.4. Murgich Model Case. The Murgich asphaltene


molecular model possesses the largest aromatic area
among all the asphaltene models analyzed here. The
rigidity of this area in the molecule does not allow it to
be folded to fill all the space. Therefore, the relaxed
aggregate structure obtained from the Murgich molecule
produces empty regions and jammed clusters of as-
phaltenes, instead of stacking configurations. The free
volume in this case is 57.30%, which is significantly
bigger than that observed in the other cases. Visualizing
carefully, however, these jammed clusters are formed
mainly by two of our three main stacking configura-
tions: offset π-stacked and edge-on or T-shaped con-
figurations. The most frequent configuration obtained
is the offset π-stacked orientation. The face-to-face
geometry configurations are not observed, which can be
attributed to a very extensive aromatic region in the
asphaltene model (Figure 12). A remarkable difference
with the other cases is the formation of two large
Figure 10. (a) Cell showing the final asphaltene structure of clusters in the simulation cell, meaning that the ali-
Zajac model. (b) Isolated aliphatic region of the complete final phatic regions are not uniformly distributed in the cell.
structure. (c) Isolated aromatic region exhibits highlighted The ratio of aromatic and aliphatic surface areas was
three geometries with two members at least: one offset
π-stacking, one edge-on or T-shaped, and one face to face. In
determined to be 0.55, which is smaller than that
this case, the AC finished having a clear concavity. observed for the Speight and Zajac cases but larger than
the Mullins case. Nevertheless, as in the Speight case,
it is the first peak on the RDF. This value is deviated the absolute size of the aromatic area has an important
almost 9% from the distance corresponding to the role in making the stacking formation more possible.
optimum binding energy (3.55 Å). The minimum distance between the aromatic cores is
The aggregate S(k) profile, as in previous cases, shows 3.65 Å, which agrees with the known distance of
all the peaks in the positions reported experimentally. optimum binding energy for two asphaltenes. However,
The exception is the position of the second peak, where the calculated peak between 1 Å-1 and 2 Å-1 is barely
there is good agreement for the Ragusa asphaltene described, considering that the first peak hides its own
aggregates (Figure 11). shoulder peak in this region (Figure 13).

Figure 11. S(k) profile for the Zajac aggregate model, compared to both the two experimental S(k) profiles, and the S(k) profile
obtained from a relaxed aggregation model built only with the aromatic cores (AC) of Zajac asphaltenes.
Aggregated Asphaltene Structural Models Energy & Fuels, Vol. 18, No. 6, 2004 1683

Table 2. Final Densities of Asphaltene Molecules after


Energy Optimization for Both the Whole Asphaltene
Model (WAM) and the Aromatic Asphaltene Core Cell
(AACC) Cases
density (g/cm3)
model WAM AACC
Mullins 0.98 1.26
Speight 1.04 1.36
Zajac 1.02 1.28
Murgich 0.69 1.24

Table 3. Free Volumes of the Final Optimized Cells for


Both the Whole Asphaltene Model (WAM) and the
Aromatic Asphaltene Core Cell (AACC) Cases
free volume (%)
model WAM AACC
Mullins 34.68% 28.13%
Speight 37.80% 31.91%
Zajac 36.96% 31.81%
Murgich 57.30% 36.04%

Figure 12. (a) Cell showing the final asphaltene structure of As shown in Figure 14, there is a higher stacking
Murgich model. (b) Isolated aliphatic region of the complete
final structure. (c) Isolated aromatic region, where one is offset tendency of the aromatic sheets; for example, in the
π-stacked, one is edge-on, and one is face-to-face stacking with Mullins case, one trimer and one pentamer are high-
two members was highlighted. lighted. However, the others are not selected, to pre-
serve clarity in the figure. As a comparison, only two
However, the comparison between the S(k) profile for dimers are highlighted in the isolated aromatic region
Murgich asphaltene aggregates and experimental re- shown in Figure 5, the Mullins asphaltene case. In the
sults shows the closest agreement, in regard to the Speight case, a trimer and a tetramer were chosen,
position of its shoulder peak and the first peak of the whereas, in the Zajac case, a dimer and a pentamer are
Gilsonite asphaltenes (see Figure 13). shown. Finally, in the Murgich aromatic model, just a
hexamer is highlighted. The geometric form chosen for
4. Aromatic Cores the Mullins and Murgich cases is offset π-stacked.
Correspondingly, the Speight and Zajac selected geo-
To explain the role of the aliphatic chains and the metric forms are offset π-stacked and T-shaped. A strict
aromatic core, all the aliphatic chains were cut from the face-to-face geometry can hardly be selected in any of
asphaltene models. After the aromatic cores were these cases. The free volumes of the AACC models are
obtained, our methodology was applied on all resulting smaller than those in the WAM, as shown in Table 3.
purely aromatic cores. The final densities of the aro- The difference between the free-volume percentages
matic asphaltene core cell (AACC) models without of the WAM and AACC models is a clear evidence of
aliphatic chains are higher than the whole asphaltene the role of aliphatic chains, because a great compression
model (WAM) (see Table 2). of the sheets results when these are removed from the
Except for the Murgich case, the order of magnitude systems. The aliphatic chains connected to the aromatic
of the WAM density values agree with those reported cores then stop the attraction between aromatic cores
by Speight24 for the specific gravity (density 60/60 °F38) of asphaltenes, which produces a large amount of empty
for various asphaltene samples obtained from different space in the aggregation of asphaltenes.
bitumens.39 In all cases, the AACC model shows its When the structure factors of the AACC and WAM
stacking presence, showing different geometries as a models are compared, a second peak shift to a smaller
mixture of the three geometries (face-to-face, edge-on value of k for the WAM model, with respect to the AACC
or T-shaped, offset π-stacked; see Figure 14). model, is observed, as in the aromatic blend molecules

Figure 13. S(k) profile for the Murgich aggregate model, compared to both the two experimental S(k) profiles available and the
S(k) profile obtained from a relaxed aggregation model built with the aromatic cores (AC) of Murgich asphaltenes.
1684 Energy & Fuels, Vol. 18, No. 6, 2004 Pacheco-Sánchez et al.

of both AC asphaltenes of Mullins, Speight, and Zajac


structure models, as shown in Figures 7, 9, and 11. This
agrees with the interpretation of Wiehe and Liang,4
meaning that the highest shoulder of the second peak
of the S(k) profile in Baxterville asphaltene is aromatic
and the other is paraffinic. The same figures also exhibit
Ragusa asphaltene,3 as it represents the best match
against asphaltenes of these types. Ragusa asphaltene
includes the major portion of the second peak of both
AC and asphaltenes of the Murgich average structure
model; however, Gilsonite asphaltene3 is the best match
against asphaltenes of the Murgich average structure
model. Third, in Figure 15, the height value on the S(k)
profile of AC compounds3 and AC Wafra A13 asphalt-
enes is observed to be larger than all the second peaks
of our asphaltene models. In addition to this point, the
k value of these experimental models is larger than that
of three of our models; only the second peak of the S(k)
profile on the AC of Speight asphaltene almost matches
the k value of the AC compounds.

5. Discussion
Figure 14. Different relaxed aromatic-core aggregated struc-
tures: (a) Mullins, (b) Speight, (c) Zajac, and (d) Murgich. Our methodology for calculating the minimum dis-
tance between two asphaltenes in the stacking can give
studied by Yen et al.3 These shifts are called the γ-band us a reasonable prediction if the asphaltene molecule
and the (002)-band, respectively. Our results are due has a compacted aromatic region that is well-designed,
to the fact that the WAM model has a stronger presence which can be done using the methodology developed by
of aliphatic chains than does the AACC model. This Ruiz-Morales.40 This is due to very good results that
effect was also very well-identified by Wiehe and Liang.4 have been obtained for the distance of binding energy
They associated paraffins to the smallest 2θ value and between two asphaltene molecules on the Speight and
aromatics to the largest 2θ value on this peak of the Murgich cases; however, although the Zajac case is 9%
structure factor of asphaltenes that they obtained. This deviated from experimental values, the Mullins case has
shift behavior is independent of the asphaltene model a deviation of 15%. These deviations clearly indicate the
molecule used, as is shown in the sequence of Figures presence of aliphatic chains, which restrict both the
7, 9, 11, and 13. When the aliphatic chains are removed closest interaction distance between two asphaltene
from the aromatic core in the Murgich case, the S(k) molecules and its geometry of stacking. The following
profile recovers the characteristic peak at ∼1.2 Å-1; in comment of Ebert,41 that “only 36% of the aromatic
addition, the Murgich AACC density recovers typical carbon was in stacks of two (‘dimers’) and 64% of the
aromatic density values (see Table 2). aromatic carbon was not in a stack of any kind (‘amor-
Given all these facts, aliphatic chains have a role in phous’)” suggests agreement with our observations.
inhibiting the stacking formation. Depending of the Another point is related to the number of aromatic
asphaltene model, there are different inhibition mech- rings in the AC of one asphaltene molecule. Ruiz-
anisms, such as intercalation of aliphatic chains be- Morales40 suggests that asphaltenes present a polyaro-
tween aromatic sheets or pulling the molecule in one matic core size of 1-2 aromatic systems with 4-10
side and only allowing some degree of stacking in fused rings in each one; according to the Mullins
another place. An example of the presence of the effect experimental work.27 Rogacheva et al.36 reported as-
of aliphatic chains is the Murgich case, where the phaltenes with 4-10 fused aromatic rings. Rogel42 has
aliphatic chains cause a delay in the molecules forming drawn polyaromatic rings with 9-14 fused rings to
regions of empty space and clusters (see Figure 12). This represent the aromatic moieties of asphaltenes. We used
effect is not observed without aliphatic chains (see four different asphaltene molecules, with 7, 9, 14, and
Figure 14). 24 aromatic rings, following previous models in the
literature. We believe that the use of molecular simula-
We have made the following three observations in our
tions to calculate the structure factor S(k) is a good
graphs of the S(k) profile for k ) 0 to k > 0 in Figure
starting point to improve this design.
14. First, the first peak on the aromatic core (AC) of
By cutting aliphatic chains of asphaltenes, just ACs
Mullins asphaltenes is the smallest one, whereas the
were obtained. These ACs were also energetically
first peak on the AC of Murgich asphaltenes is the
optimized, and their structure factors S(k) were ob-
largest one, and its second peak is the smallest one
tained, as shown in Figure 15. Figures 8c and 14a show
among all of the AC asphaltenes. Second, the S(k) profile
that an asphaltene molecule designed by Mullins with
of Baxterville asphaltene3 includes the major component
seven fused aromatic rings exhibits at least one 5-aro-
(38) Relative density (60/60 °F), which was called density by Speight, matic-stack system. As we can also see in Figure 15,
is the ratio of the mass of a given volume of liquid at 60 °F to the
mass of an equal volume of pure water at the same temperature. (40) Ruiz-Morales, Y. J. Phys. Chem. A 2002, 106, 11238-11308.
(39) Speight, J. G. The Chemistry and Technology of Petroleum; (41) Ebert, L. B. Fuel Sci. Technol. Int. 1995, 13, 941-944.
Marcel Dekker: New York, 1999; p 315. (42) Rogel, E. Langmuir 2004, 20, 1003-1012.
Aggregated Asphaltene Structural Models Energy & Fuels, Vol. 18, No. 6, 2004 1685

Figure 15. Comparison between structure factors of aromatic cores (AC) calculated by us (continuous line) and those proposed
by Yen et al.3 (dotted and dashed lines).

the AC of Speight asphaltenes with 14 fused aromatic pendent on Euclidean geometric forms of the molecule,
rings gave the best agreement between its second S(k) because of the potential they use, as the hard-sphere
peak and the first S(k) peak of AC compounds; however, intermolecular potential of interaction. This is a gross
the AC of Mullins asphaltenes is deviated from match- limitation, because the center of mass is translated,
ing. Second, the S(k) peak of the AC of Zajac asphaltenes which produces deviations from real systems as as-
with nine aromatic fused rings has a very similar phaltenes are. They usually compare their results
behavior to that of the AC of Mullins asphaltenes. In against MD simulations, Monte Carlo (MC) calculations,
the case of Murgich asphaltene, we found that the AC and/or experiments.45 Yen et al.3 used a very clever
of 24 fused aromatic rings has a behavior more similar methodology to compare the XRD experimental pattern
to asphaltenes than to ACs of asphaltenes (see Figure for aromatic clusters in petroleum asphaltene with the
13). After this analysis, it can be concluded that the experimental pattern for a blend of five polynuclear
Murgich asphaltene case is only meaningful when aromatic compounds of known structure by computing
aliphatic chains are removed. the Debye RDF implemented by Diamond,13 which is
We must stress that minimization of the energy is precisely the structure factor S(k). We calculated a very
independent of the temperature, and a complete study good approximation of the S(k) profile by optimizing the
of the stability using entropy and thermodynamic energy of a 35-membered aromatic system, using the
potentials is not possible within the present methodol- same five aromatic compounds mentioned previously in
ogy. At this time, we just want to show that we found a groups with seven-membered rings.
very good matching between this theory and experiment Whether the approved SF profile is our guide to
stabilizing asphaltenes by molecular simulations of validate generated morphologies, the effects of polydis-
energy optimization. persity on the SF profile are well-known, as shown in
It must be stated that, knowing the interaction some studies developed by D’Aguanno and Klein.43 In
potential, the structure factor S(k) can be approximated that work, the effects of polydispersity in charged
from the solution of the Ornstein-Zernicke (O-Z) colloidal dispersions are exhibited in a manner in which
integral equation and an additional closure relation (PY, the SF peaks are displaced, modifying its height and
MSA, RMSA, HNC, RY, etc.).43,44 The methodology of width at different levels of polydispersity. Asphaltenes
Henderson-Barker-Abraham and of density functional are known to constitute a polydisperse system; however,
theory (DFT) can be used to obtain the solution of the in the literature, it is usual procedure to characterize
O-Z equation.9,45,46 However, these methods are de- crude oil systems by means of an asphaltene average
molecule.18,23,27,28 Because of this restriction on the
(43) D’Aguanno, B.; Klein, R. J. Chem. Soc., Faraday Trans. 1991, characterization of the entire polydispersity of the
87 (3), 379-390.
(44) Ortega-Rodrı́guez, A.; Cruz, S. A.; Gil-Villegas, A.; Guevara-
asphaltenic system, as a first approach, we have taken
Rodrı́guez, F.; Lira-Galeana, C. Energy Fuels 2003, 17, 1100-1108.
(45) von Grünberg H. H.; Klein R. J. Chem. Phys. 1999, 110 (11), (46) Pacheco-Sánchez, J. H.; Rodrı́guez, A. G. Rev. Inst. Mex. Pet.
5421-5431. 1993, 25 (2), 55-60.
1686 Energy & Fuels, Vol. 18, No. 6, 2004 Pacheco-Sánchez et al.

a set of average molecules reported in the literature as and 2 Å-1 is associated with the stacking. We observed
the starting point in this study. stacking clusters when aromatic molecules are used.
Finally, our results here can help to improve the Based on the similarity of the S(k) profiles, we can
experimental methodology for designing asphaltene conclude similarities in the asphaltene agglomeration
molecules, which can be useful, as many researchers of the Zajac, Mullins, and Speight cases with Ragusa-
have shown.14-18,22,30,42,47 type asphaltene, and of the Murgich case with Gilsonite-
type asphaltene.
6. Conclusions Generally, the positions of the main S(k) peaks were
correctly reproduced, showing a discrepancy on the
One possible way to determine the morphology of second peak region. Such a discrepancy can be at-
asphaltene aggregates is through modeling and simu- tributed to the tail presence in the aromatic sheets that
lating such aggregates. This is because (i) our simulated inhibits the π-π, π-σ, and σ-σ interactions, bringing,
structure factors (S(k)) of asphaltene aggregates agree as a consequence, an increase of the distance between
with those reported by Yen; (ii) the molecular models these sheets. The good correlation between the experi-
of asphaltenes are experimentally designed from ex- mental and predicted S(k) profiles would imply that the
tracted crude oils; and (iii) experimentally, it is possible asphaltene aggregate structure is an adequate model
to simply guess about the morphology of the aggregates. for studying the asphaltene aggregation phenomenon.
We have proposed a procedure for generating as- Therefore, based on the present simulations, we
phaltene aggregated structures, based on a minimiza- conclude that a nearest-neighbor stacking in face-to-face
tion of the energy of the system in periodic cells. These geometry is not the only possible orientation. The
periodic cells were constructed using different asphalt- observed asphaltene aggregate structure represents just
ene models, such as those devised in the research of one of the possible ensemble structures that could exist
Mullins, Speight, Zajac, and Murgich. Each cell was in the system. The final cells are good models as a first
built with one type of asphaltene model in an amor- approach for representing asphaltene aggregates as well
phous arrangement to optimize them. The developed as the way in which they agglomerate in an amorphous
methodology is capable of reproducing the position of solid phase. In addition, we applied our methodology to
the experimental S(k) profile. Important differences the different asphaltene models used but without their
were observed, with respect to the AC cases. The aliphatic chains, finding a similar stacking behavior for
presence of a shift in the peak located between 1 Å-1 each case.
(47) Rogel, E.; León, O. Energy Fuels 2001, 15, 1077-1086. EF049911A

You might also like