Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Energy Fuels 2010, 24, 2795–2808 : DOI:10.

1021/ef100173j
Published on Web 04/23/2010

Chemistry and Association of Vanadium Compounds in Heavy Oil and Bitumen,


and Implications for Their Selective Removal†
Greg P. Dechaine* and Murray R. Gray
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB Canada T6G 2 V4

Received February 11, 2010. Revised Manuscript Received April 1, 2010

Most crude oils contain traces of vanadium, which cause significant detrimental impact during catalytic
conversion and combustion. The concentrations in bitumen and vacuum residue are generally much
higher, which poses a problem for the economical upgrading of these feedstocks. These problems are
relevant since the world reserves of conventional light oils are dwindling and being replaced by an
increasing amount of heavier feedstocks. In this article, the current understanding of the distribution and
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF MARYLAND COLG PARK on March 9, 2021 at 17:38:01 (UTC).

form of vanadium compounds within vacuum residue is critically discussed. The implications of this
chemistry on prospects for new separation methods, other than deasphalting, are considered. Although
only a small fraction of the vanadium is contained within the polar resin fraction (i.e., maltenes), this
fraction has been most often characterized to determine the chemical form of the vanadium compounds.
Various spectroscopic techniques have been used to determine that these resin soluble vanadium
compounds exist as metalloporphyrins characterized by their intense absorption of UV/visible radiation.
In the case of the asphaltene bound metals, this UV/visible absorbance is not observed and has historically
led to their distinction as “non-porphyrins”. However, more recent results using X-ray spectroscopies
(EXAFS and XANES), as well as a more in depth analysis of the UV/vis response of metalloporphyrins,
indicates that, although these asphaltene-bound vanadium compounds do not exhibit the characteristic
UV/visible absorption, they are indeed still bound in a porphyrinic structure. The fact that the majority of
the vanadium is contained within the highly aromatic, highly polar asphaltene fraction also poses
additional roadblocks to their selective removal. This fraction has been shown to associate/aggregate
significantly in most (if not all) solvents. Recent work on the nature and possible mechanisms of the
molecular association of asphaltenes in solution can be extended to help elucidate the molecular
interactions occurring between asphaltenes and metalloporphyrins, and hence the nature of the inclusion
of metalloporphyrins within the asphaltene fraction. Many different solvent systems including aromatics,
chloro-carbons, alcohols, ketones, as well as other polar solvents have been used to extract the
metalloporphyrins from the asphaltene fraction with limited success. In most cases, the effect of asphaltene
solubility and aggregation in the given solvent were not considered. The selective separation of
metalloporphyrins is clearly hampered by gaps in the basic understanding of the metalloporphyrin
properties and behavior in solution.

1. Introduction combustion, which poses a toxicity concern2-4 if emitted


directly to the environment from a stack, as well as a corrosion
Most crude oils contain traces of metal complexes of
concern for turbines when used in power generation.5 These
various forms. The most abundant and troublesome metal
problems are magnified by the fact that the world reserves of
complexes present in the organic portions of fossil fuel
conventional light oils are dwindling and being replaced by an
deposits are vanadium and nickel. Although these metals
ever increasing amount of heavier feedstocks. Therefore, the
are only present in trace quantities, they lead to deactivation
selective removal of these metal contaminants from heavy oils
of both desulphurization and cracking catalysts.1 For heavy
and bitumen is highly desirable. Because these components
oils and bitumen, the concentrations of these metals are
occur predominantly in the vacuum residue fraction, regard-
generally much higher, which poses a problem for the eco-
less of the properties of the original crude oil or bitumen, we
nomical upgrading of these feedstocks into saleable products.
will focus discussion on this fraction.
As well, the presence of vanadium compounds in product
The current industrial technologies that are applied on a
coke leads to the formation of vanadium pentoxide during
large scale for the removal of vanadium fall into one of three
categories: coking, deasphalting, and catalytic hydrodemeta-
*To whom correspondence should be addressed. Tel.: (780) 492-1107.
Fax: (780) 492-2881. E-mail: gpd@ualberta.ca.
lation. The most widely used technology is coking, which

A preliminary version of this paper was published in Greg Dechaine’s almost quantitatively captures the vanadium in the coke
Ph.D. Thesis (University of Alberta, Jan 26, 2010). byproduct.6 However, as indicated above, the presence of
(1) Branthaver, J. F. ACS Sym. Ser. 1987, 344, 188–204. vanadium in this product poses problems during combustion.
(2) Zychlinski, L.; Byczkowski, J. Z.; Kulkarni, A. P. Arch. Environ.
Con. Tox. 1991, 20, 295–298.
(3) Cooper, R. G. Indian J. Occup. Environ. Med. 2007, 11, 97–102. (5) Drbal, L. F.; Boston, P. G.; Westra, K. L. Power Plant Engineering
(4) Occupational Safety & Health Administration, U. S. Occupa- by Black & Veatch: Springer: New York, 1996.
tional Safety and Health Guideline for Vanadium Pentoxide. www.osha. (6) Siskin, M.; Kelemen, S. R.; Eppig, C. P.; Brown, L. D.; Afeworki,
gov (accessed Jan 7, 2010). M. Energy Fuels 2006, 20, 1227–1234.

r 2010 American Chemical Society 2795 pubs.acs.org/EF


Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

In cases where the byproduct coke is stockpiled, this is not and nickel compounds across acid-base-neutral (ABN) frac-
a concern, although stockpiling of coke represents a signifi- tions of Wilmington and Mayan crudes both before and after
cant loss of product and hence a loss of a potential revenue hydrotreating. They found that, for the feed materials, vana-
stream. The precipitation of some or all of the asphaltene dium was present in all of the ABN fractions.
fraction from the feed also results in significant removal Altgelt and Boduszynski13 measured the concentration of
of vanadium7 since the vanadium partitions preferentially vanadium as a function of atmospheric equivalent boiling
within this highly polar, highly aromatic fraction.8 However, point (AEBP) for several heavy oil samples and showed that
this precipitation process is nonselective and results in a the concentration of vanadium increases with increasing
significant loss of product in the form of a contaminated AEBP. This corroborates the previous findings that the
asphaltene stream. Finally, catalytic hydroprocessing selec- vanadium tends to concentrate in the residue fraction of
tively removes the vanadium as a vanadium sulfide deposit petroleum. Filby and Strong14 observed that vanadium is
on the catalyst in ebullated bed reactors.9 Although this distributed throughout all fractions of Athabasca bitumen,
process is the only selective process of the three, this selecti- although the majority (76.1%) is present with the asphaltene
vity comes at a substantial cost in catalyst as well as energy in fractions while the other 23.1% was present with the pentane-
the form of elevated temperatures, pressures, and hydrogen soluble maltenes. Pena et al.15 measured the vanadium dis-
consumption. tribution in Mexican offshore heavy crude and found as other
The fact that the majority of the vanadium is contained investigators did that the majority of the vanadium was
within the highly aromatic, highly polar asphaltene fraction present in the asphaltene fraction of the crude.
poses some significant difficulties. The asphaltene fraction is Finally, Yang et al.16 measured the distribution of vana-
defined as the portion of a petroleum feed that is toluene- dium compounds in various subfractions of Athabasca as-
soluble and n-alkane-insoluble (e.g., n-heptane). This opera- phaltenes. Each subfraction was obtained by precipitation
tional definition means that asphaltenes represent a solubility using an increasing amount of n-heptane precipitant. They
class and as such are a heterogeneous mixture of molecules. As found that the vanadium was distributed across all of the
is discussed later, this fraction of petroleum has been shown to subfractions, with a slight decrease in vanadium content as the
associate/aggregate significantly in most (if not all) solvents. ratio of n-heptane to bitumen increased. The fact that the
This aggregation poses some major difficulties in removing majority of the vanadium is contained within the highly
the vanadium compounds. aromatic, highly polar asphaltene fraction does pose some
This review examines the literature on the chemistry of the significant difficulties. As is discussed later, this fraction of
vanadium compounds in heavy oil and bitumen, in the petroleum has been shown to associate/aggregate significantly
vacuum residue fraction. To borrow the terminology of in most (if not all) solvents. This aggregation poses some
protein chemistry, we consider both the primary chemical major difficulties in removing the metal components.
coordination of the vanadyl species and the secondary struc-
ture of the molecules containing the vanadium. In this ana- 3. To Be or Not to Be (a Porphyrin);That Is the Question.
logy, the tertiary structure is the nature of the molecular The Characterization of Vanadium Compounds
aggregates within the crude oil that contain vanadium and
Although heavy oils and bitumen contain significantly
nonvanadium compounds. The implications of these struc-
higher concentrations of both vanadium and nickel than
tural features are then considered from the viewpoint of
conventional oils, the exact molecular form of these metals
selective removal of vanadium from vacuum residue, either
is still a point of contention among researchers within the field.
as inorganic metal compounds or as intact metal-organic
A wide variety of analytical methods have been employed to
complexes.
attempt to determine the exact molecular form of the vana-
dium compounds. Beyond dispute is the fact that a fraction of
2. Distribution of Vanadium Among Petroleum Fractions
the vanadium present in petroleum deposits is in the form of
Information on the distribution of vanadium is important vanadyl porphyrins.17 In this form, the vanadium is axially
in order to develop processes for the removal of vanadium. coordinated to an oxygen atom and also coordinated to the
Barwise and Whitehead10 measured the concentration of four nitrogen atoms of the porphyrin macrocycle, as illu-
vanadium for various boiling point fractions of a Boscan strated in Figure 1. These compounds are derived from
crude and found that very little of the vanadium is contained naturally occurring organic matter such as chlorophyll and
in the distillates (350-500 °C), while the majority of the heme (also shown in Figure 1). The most common forms of
vanadium is present with the residue fractions (>500 °C). vanadyl porphyrins identified in petroleum deposits are the
Reynolds8 separated several atmospheric residua using a Etio form (Figure 1b) and the DPEP form (Figure 1d). Other
modified ASTM 2007 separation (saturate, aromatic, resin, forms have been identified, such as the Rhodo or Benzo forms
asphaltene, or SARA) and determined that the majority (VOBenzo, Figure 1e), although not in the same abun-
(>90% in all cases, >95% for most) of the metals were dance. Recently, Qian et al.18 successfully identified vanadyl
contained in the polar fractions (both resin and asphaltene),
with a further majority concentrated in the asphaltenes. (13) Altgelt, K. H.; Boduszynski, M. M. Composition and Analysis of
Pearson and Green11,12 studied the distribution of vanadium Heavy Petroleum Fractions; Marcel Dekker, Inc.: New York, 1994; Vol. 54
(14) Filby, R. H.; Strong, D. AIChE Symp. Ser. 1991, 87, 1–9.
(15) Pena, M. E.; Manjarrez, A.; Campero, A. Fuel Process. Technol.
(7) Brons, G.; Yu, J. M. Energy Fuels 1995, 9, 641–647. 1996, 46, 171–182.
(8) Reynolds, J. G. Liq. Fuels Technol. 1985, 3, 73–105. (16) Yang, X.; Hamza, H.; Czarnecki, J. Energy Fuels 2004, 18, 770–
(9) Gray, M. R. Upgrading Petroleum Residues and Heavy Oils; Marcel 777.
Dekker, Inc.: New York, 1994. (17) Yen, T. F. Chemical Aspects of Metals in Native Petroleum. In
(10) Barwise, A. J. G.; Whitehead, E. V. Prepr.;Am. Chem. Soc., The Role of Trace Metals in Petroleum; Yen, T. F., Ed.; Ann Arbor Science
Div. Pet. Chem. 1980, 25, 268–279. Publichers Inc.: Ann Arbor, MI, 1975; pp 1-30.
(11) Pearson, C. D.; Green, J. B. Fuel 1989, 68, 456–464. (18) Qian, K.; Mennito, A. S.; Edwards, K. E.; Ferrughelli, D. T.
(12) Pearson, C. D.; Green, J. B. Fuel 1989, 68, 465–474. Rapid Commun. Mass Spectrom. 2008, 22, 2153–2160.
2796
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

Figure 1. General structure and nomenclature of vanadyl porphyrins22.

porphyrins in unfractionated asphaltenes for the first time, of the intensity and sensitivity of the electronic absorption of
identifying VODPEP and VOBenzo as the dominant types. In UV/visible radiation by metalloporphyrins, electronic absorp-
addition, they observed cycloalkane-substituted and sulfur- tion (UV/visible) spectroscopy has been widely used in their
containing porphyrins. Some common vanadyl porphyrins identification and quantification in petroleum samples. How-
used as model compounds include vanadyl octaethylporphyrin ever, a significant portion of the vanadium present, particu-
(VOOEP, Figure 1c) and vanadyl meso-tetraphenylporphyrin larly the vanadium associated with the asphaltene fraction,
(VOTPP, Figure 1f). The nomenclature for the various vana- does not display this characteristic absorption. If we assume
dyl porphyrins shown in Figure 1 will be used throughout the that the extinction coefficients of the petroporphyrins are
remainder of this review. comparable to those of model compounds such as VOOEP,
One of the primary characteristics of the porphyrin macro- then the measured UV/visible absorbance is much too small to
cycle is intense absorption of UV/visible radiation.19 Because account for the total vanadium content of crude oils.14,20,21

(20) Senglet, N.; Williams, C.; Faure, D.; Des Courieres, T.; Guilard,
(19) Smith, K. M. General Features of the Structure and Chemistry of R. Fuel 1990, 69, 72–77.
Porphyrin Compounds. In Porphyrins and Metalloporphyrins; Smith, K. (21) Sugihara, J. M.; Bean, R. M. J. Chem. Eng. Data 1962, 7, 269–
M., Ed.; Elsevier Scientific Publishing Company: New York, 1975; pp 1-28. 271.
2797
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray
26
This discrepancy led to a distinction between the porphyrin Berthe et al. applied X-ray photoelectron spectroscopy
fraction of the metals, which absorbed as expected, and the (XPS or ESCA) to the analysis of vanadium compounds in
“non-porphyrin” fraction, which did not. heavy oils. They compared the binding energy of vanadium
3.1. X-Ray Absorption Spectroscopy. A very powerful in both Cabimas and Boscan asphaltenes to the binding
method available for characterizing the form of the vana- energy of vanadium in several model vanadyl compounds
dium in petroleum is the family of X-ray absorption spectro- with different coordination environments. The binding en-
scopies. In particular, extended X-ray absorption fine ergy of the vanadium in asphaltenes very closely matched
structure (EXAFS) spectroscopy and X-ray absorption that for the model compounds with a 4-N coordination
near-edge structure (XANES) spectroscopy can both be used environment. The binding energy observed with other co-
to obtain information regarding the bonding structure sur- ordination spheres (i.e., 4-O; 4-S; and various combinations
rounding the vanadium atoms within the petroleum and of O, S, and N) showed binding energies significantly
asphaltene matrix. different than that of the asphaltene bound vanadium.
Goulon et al.23 measured the EXAFS/XANES spectra for Loos et al.27 compared the EXAFS spectra of a series of
Boscan asphaltenes and compared them to those for oxova- vanadyl porphyrins with varying types of substitution at the
nadyl (VO2þ) OEP and thiovanadyl (VS2þ) OEP. As well, periphery of the porphyrin macrocycle: VODPEP, VOEtio,
the asphaltene spectra were compared to the spectra of VOOEP, VOtetrabenzyl, VOtetrapyridine, etc. The effect
authentic petroporphyrins extracted from the original crude of varying the size of the substituents but not location
sample using dichloromethane þ n-hexane mixtures fol- (e.g., Etio vs OEP) was low for the VdO and V-N bonds
lowed by chromatographic separations on aluminum oxide and became significant at distances farther from the vana-
and silica gel columns. Although UV/visible Soret absor- dyl center. Adding meso-tetrasubstitution (e.g., meso-
bance (see section 3.2) of the asphaltenes accounted for only tetraphenylporphyrin) did have a slight impact on the mag-
13-15% of the total vanadium present in the asphaltene nitude of the VdO and V-N signals. These slight variations
fractions, the EXAFS and XANES spectra of the whole at the center of the porphyrin core arise due to slight
asphaltenes were almost completely superimposed on the distortions induced in the porphyrin macrocycle by the
spectra for pure VOOEP and the authentic petroporphyrins different substituents at the periphery.
in toluene. This result indicated that, although the asphal- Zhang and Boduszynski28 compared the EXAFS and
tenes show very high levels of “non-porphyrinic” vanadium XANES spectra of asphaltenes to the spectra for VOTPP
(i.e., vanadium with no UV/visible absorbance), the vana- and several other species with vanadyl coordinated to other
dium within this fraction is of the oxovanadyl type coordi- types of ligands (vanadyl acetylacetonate and vanadyl
nated to four nitrogen atoms, as in the porphyrin macro- sulfate). They found that the vanadium is almost exclusively
cycle. The spectra for thiovanadyl OEP was also recorded coordinated to five ligands and not six ligands. They also
and showed a significantly different spectrum than all of the found that the XANES and EXAFS spectra of the asphal-
samples tested, indicating that the majority of the vanadium tenes very closely resembled the spectra for VOTPP, showing
in the asphaltene fraction is in the oxovandyl form. Goulon the characteristics of the VdO bond and the four V-N
et al.24 extended this analysis to Cabimas and Aramco bonds, thus indicating that the vanadium present in these
asphaltenes and obtained similar results. asphaltenes is bound in a porphyrinic structure.
Poncet et al.25 synthesized a VOEtio compound with the Miller et al.29 separated Mayan n-heptane asphaltenes into
four nitrogen atoms replaced by sulfur atoms. The EXAFS two fractions by Soxhlet extraction, yielding approximately
spectrum of this tetra-sulfur porphyrin ligand was dramati- 25% of the asphaltenes as soluble in n-heptane. Charac-
cally different than the traditional VOEtio spectra. When the terization of these two fractions using size exclusion
spectra of VOEtio were compared to those for authentic chromatography (SEC), vapor-pressure osmometry (VPO),
petroporphyrins,23 the spectra were nearly identical. There- and small angle neutron scattering (SANS) determined that
fore, the vanadium compounds present in asphaltenes are the soluble fraction did not associate in aromatic solvents,
unlikely to be coordinated to four sulfur atoms. while the insoluble fraction did form aggregated structures in
aromatic solvents. Therefore, the soluble fraction was named
(22) Czernuszewicz, R. S.; Maes, E. M.; Rankin, J. G. Resonance
“non-colloidal” asphaltenes and the insoluble fraction, “col-
Raman Spectroscopy of Petroporphyrins. In The Porphyrin Handbook; loidal” asphaltenes. Miller et al.30 then used UV/visible and
Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Press: San Diego, EXAFS spectroscopy to determine the form of the vanadium
CA, 2000; Vol. 7, pp 293-335. contained in the different fractions of this Mayan asphaltene
(23) Goulon, J.; Retournard, A.; Friant, P.; Goulonginet, C.; Berthe,
C.; Muller, J. F.; Poncet, J. L.; Guilard, R.; Escalier, J. C.; Neff, B. both before and after hydrocracking. The presence of UV/
J. Chem. Soc. Dalton 1984, 1095–1103. visible absorbance peaks in the electronic spectra of the
(24) Goulon, J.; Esselin, C.; Friant, P.; Berthe, C.; Muller, J. F.; Pncet, noncolloidal asphaltenes indicates that the metals are pre-
J. L.; Guilard, R.; Escalier, J. C.; Neff, B. In Structural Characterization
by X-Ray Absorption Spectroscopy (EXAFS/XANES) of the Vana- sent as metalloporphyrins, while the colloidal asphaltenes
dium Chemical Environment in Various Asphaltenes, Characterization
of heavy crude oils and petroleum residues. International Symposium on
Characterization of Heavy Crude Oils and Petroleum Residues, Lyon, (27) Loos, M.; Ascone, I.; Friant, P.; Ruiz-Lopez, M. F.; Goulon, J.;
France, June 25-27, 1984; Editions Technip: Paris, pp 158-163. Barbe, J. M.; Senglet, N.; Guilard, R.; Faure, D.; Des Courieres, T.
(25) Poncet, J. L.; Guilard, R.; Friant, P.; Goulonginet, C.; Goulon, J. Catal. Today 1990, 7, 497–513.
New J. Chem. 1984, 8, 583–590. (28) Zhang, G.; Boduszynski, M. M. In The Binding Structure of
(26) Berthe, C.; Muller, J.-F.; Cagniant, D.; Grimblot, J.; Bonnelle, Vanadium in Arabian Heavy Petroleum: An X-Ray Absorption Investiga-
J.-P. In Characterisation des Fractions Lourdes du Petrole par Spectro- tion, International Conference on Petroleum Refining and Petrochemical
scopie IR, UV-Visible, Spectroscopie Photoelectronique ESCA et Spectro- Processing, Beijing, China, September 11-15, 1990; Xianglin, H., Ed.;
metrie LAMMA en Relation avec L’environnement du Vanadium, International Academic Publishers: Beijing, China, 1991; pp 821-826.
Characterization of heavy crude oils and petroleum residues: Inter- (29) Miller, J. T.; Fisher, R. B.; Thiyagarajan, P.; Winans, R. E.;
national Symposium on Characterization of Heavy Crude Oils and Hunt, J. E. Energy Fuels 1998, 12, 1290–1298.
Petroleum Residues, Lyon, France, June 25-27, 1984; Editions Technip: (30) Miller, J. T.; Fisher, R. B.; van der Eerden, A. M.; Koningsberger,
Paris, pp 164-168. D. C. Energy Fuels 1999, 13, 719–727.
2798
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

show no such absorbances (consistent with previous in- UV/visible radiation by metalloporphyrins, electronic ab-
vestigations). This trend continued for the hydrocracked sorption (UV/visible) spectroscopy has been widely used in
residuum as well. The EXAFS spectra for the untreated their identification and quantitation in petroleum samples.
noncolloidal and colloidal asphaltenes were qualitatively 3.2.1. Effect of Solvent. The solvent system can have
similar to the spectra for pure VOTPP. The investigators an impact on measured UV/visible spectra.31-33 Freeman
fit an assumed square planar porphyrin structural model to et al.34 used UV/visible spectroscopy to study the absorbance
the EXAFS data and calculated bond distances from the of VOOEP and VOEtio-I in methylene chloride, chloroform,
vanadium atom. Their values all agreed within experimental 1-2 dichloroethane, ethyl acetate, and toluene. The peak
error with the bond distances calculated for VOTPP (both height of the R band (570 nm) was used to quantify the
from the EXAFS spectra and XRD data for the model concentration of the metalloporphyrin. Overall, their analy-
VOTPP compound30). In the case of the hydrocracked sis was very thorough and represents a good benchmark for
residuum, Miller et al. found that the form of the vanadium applying UV/visible spectroscopy to the quantitation of
in the unconverted asphaltenes was still that of the vanadyl metalloporphyrins in solution. These investigators were
ion coordinated to four nitrogen atoms in a square planar capable of detecting concentrations as low as 0.1 μg/mL
arrangement. However, the form of the vanadium within the (∼0.1 ppm), despite the reduced intensity of the R band
resulting chlorobenzene insoluble solids could no longer be (ε = 28 000-32 000 L/mol 3 cm) versus the Soret band (ε >
fit by the standard porphyrinic model. Rather, a distorted 350 000 L/mol 3 cm35). They also showed that baseline correc-
octahedron including four N, an O, and another O or N tion of the peak height results in significant improvements in
closely approximates the measured EXAFS spectra. The reproducibility for metalloporphyrins. The two main results
presence of this additional ligand is similar to the axial obtained in this work were to show that the solvent has an
ligation observed between Lewis bases and vanadyl porphy- impact on the extinction coefficient of metalloporphyrins
rins in solution (see section 3.2.3). During processing at high (as much as 18% variation from one solvent to the next) and
temperatures, heteroatoms within the asphaltene matrix may to determine the solubility of these metalloporphyrins as a
chemically bind to the vacant ligand site of the vanadium function of the solubility parameter of the solvent.
center to form a new chemical bond. Ferrer and Baran36 measured the electronic absorption of
The evidence obtained by the X-ray methods is very VOTPP in various different solvents of different polarity and
consistent in its assertion that all of the vanadium present dielectric strength. The solvent had a significant impact on
in petroleum and asphaltene samples is present as a vanadyl the shape of the spectra when it was capable of axial
ion coordinated to four nitrogen ligands. The V-N bond interaction with the vanadyl atom, such as in the case of
distances observed are indicative of a porphyrin-like macro- DMSO, DMF, methanol, and pyridine. No indication was
cycle with only slight variations observed. However, as given as to the impact of the different solvents on the
indicated by Loos et al.,27 slight variations in the V-N bond intensity of the different peaks.
distances can be induced by different peripheral substituents. 3.2.2. Effect of Peripheral Substitution. Freeman et al.37
Any EXAFS signal obtained for a heterogeneous sample identified the peak locations for various vanadyl porphyrins
such as asphaltenes will undoubtedly include vanadyl por- in dichloromethane solvent using third derivative UV/visible
phyrins with a variety of peripheral substitutions and there- spectroscopy, and some of the results are shown in Table 1.
fore will represent an “average” spectrum. Such an average Foster et al.35 obtained the extinction coefficients and
spectrum would have some variation in appearance from wavelengths for the Soret bands of various vanadyl porphyr-
model compounds. However, the EXAFS spectra obtained ins in toluene, as shown in Table 2. Cant u et al.38 determined
by numerous different investigators on several different the location of the three peaks for VODPEP in dichloro-
asphaltene and petroleum samples all exhibited spectra methane as 410.5 nm (Soret peak, ε = 128 000), 533.6 nm
which were strikingly similar to those for model vanadyl (β peak, ε = 7060), and 574.5 nm (R peak, ε = 8720). The
compounds in the vicinity of the vanadyl core (VdO and results in Tables 1 and 2 and those of Cant u et al. indicate
V-N bonds), indicating that all of the vanadium is coordi- that the peripheral substitution of the porphyrin macrocycle
nated in such an environment. The subsequent variations of can have an impact on the location of the UV/visible peaks.
the sample spectra from the model compound spectra at When the peripheral substituents are in the same location
greater distances from the vanadyl core are indicative of (i.e., OEP, Etio, DMEP), the locations of the Soret and
variation in peripheral substitution, which is anticipated for visible peaks remain relatively fixed. However, when the
heterogeneous samples such as asphaltenes. peripheral substituents occupy different locations (such as
3.2. Electronic Absorption (UV/visible) Spectroscopy. As the meso positions in the case of TPP), the Soret peak is
mentioned, one of the primary characteristics of the por-
phyrin macrocycle is intense absorption of UV/visible radia- (31) Reichardt, C. Solvents and Solvent Effects in Organic Chemistry,
tion.19 The most intense absorption occurs in the vicinity of 3rd ed.; Wiley-VCH: Weinheim, Germany, 2003.
(32) Jauquet, M.; Laszlo, P. Influence of Solvents on Spectroscopy. In
400 nm (near-UV/violet) and is termed the Soret band. Solutions and Solubilities; Dack, M. R. J., Ed.; John Wiley & Sons: New
Simple free-base porphyrins also have four characteristic York, 1975; Vol. VIII, pp 195-258.
bands in the visible region, the location of which is dependent (33) Smith, B. C. Quantitative Spectroscopy: Theory and Practice;
Academic Press: Amsterdam, 2002.
on the peripheral substitution of the macrocycle. In the case (34) Freeman, D. H.; Swahn, I. D.; Hambright, P. Energy Fuels 1990,
of the metalloporphyrins, these four visible bands are 4, 699–704.
reduced to two bands referred to as R and β bands. Generally (35) Foster, N. S.; Day, J. W.; Filby, R. H.; Alford, A.; Rogers, D.
Org. Geochem. 2002, 33, 907–919.
speaking, the Soret band is much more sensitive than the two (36) Ferrer, E. G.; Baran, E. J. J. Electron Spectrosc. 1991, 57, 189–
visible bands and is the band of choice for quantitative 197.
analytical work,19 although as discussed later there are situa- (37) Freeman, D. H.; Saint Martin, D. C.; Boreham, C. J. Energy
Fuels 1993, 7, 194–199.
tions where the visible bands are a better choice. Because of (38) Cantu, R.; Stencel, J. R.; Czernuszewicz, R. S.; Jaffe, P. R.; Lash,
the intensity and sensitivity of the electronic absorption of T. D. Environ. Sci. Technol. 2000, 34, 192–198.
2799
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

Table 1. Location of Absorption Maxima for Various Vanadyl 3.2.3. Effect of Coordination and Association. The occur-
Porphyrins in Dichloromethane37 rence of coordination or dimerization/aggregation can alter
wavelength @ maximum (nm) the UV/visible spectra of metalloporphyrins.44 Walker
et al.45 monitored the UV/visible spectra of substituted
vanadyl porphyrin soret band β band R band VOTPP in toluene in the presence of small amounts of
no substitution 399.4 523.8 559.4 piperidine. The addition of piperidine resulted in the forma-
Etio I 406.6 532.8 570.7 tion of a porphyrin-piperidine adduct, causing a red shift of
octaethyl 407.3 533.2 570.9
12 nm for the Soret band with the extinction coefficient
DPEP 410.5 533.3 573.0
benzo 414.0 544.7 578.7 remaining unchanged.
Bonnett et al.46 used UV/visible spectroscopy to monitor
Table 2. Location and Extinction Coefficient for the Soret Band of the interaction of various basic solvents with VOOEP. They
Various Vanadyl Porphyrins in Toluene35 noted a change in the Soret and visible bands when strongly
soret band extinction coefficient coordinating solvents, in particular primary amines such as
vanadyl porphyrin wavelength (nm) (mM-1 cm-1) n-butylamine, were used. They propose that the primary
Etio I 407 400 amines can more readily coordinate to the vanadyl species
DMEP 407 390 because of reduced steric hindrance at the final remaining
OEP 407 364 vanadyl coordination site. The importance of steric hin-
TPP 424 528 drance on additional ligation of the vanadium atom arises
shifted to a longer wavelength (i.e., a red shift). Also, when because the vanadium atom lies above the plane of the four
the substituents begin to exhibit more cyclic (DPEP) and nitrogen atoms by 0.5 Å.47 The prospect ligand must be
aromatic (Benzo) characteristics, red shifts also occur. capable of penetrating the microwell created by the elevated
Trofimenko et al.39 observed minor shifts in the locations of vanadium atom in order to create the coordinating bond.
the absorption peaks for the different isomers of Etiopor- Shelnutt et al.48 observed similar tendencies and found that
phyrin (free base form) in benzene and primary alcohols, pyrrolidine coordinated much more readily with VOOEP
although the qualitative shape of the spectra remained un- than did pyridine, and this coordination caused a red shift in
changed. Therefore, even small changes in peripheral sub- the Soret band from 407 to 425 nm.
stitution (such as isomerization) can impact the electronic Bencosme et al.49 found similar steric effects when study-
absorption characteristics of the porphyrin. ing the axial ligation of various different Lewis bases with
Also worthy of note is the effect of substitution on the VOTPP in dichloromethane. They found that the ability of
magnitude of the absorption peaks. The extinction coeffi- the different Lewis bases to coordinate with the vanadyl
cient measured by Cant u et al.38 for VODPEP is ∼4 times structure followed the order nitrogenated > oxygenated g
lower than that obtained by Foster et al.35 for VOEtio. sulfonated. They also found that the ability of a Lewis base to
Therefore, if the extinction coefficient for VOEtio is used form an axial ligand with the out of plane vanadyl ion was
for quantitative studies of a mixture of VOEtio and VOD- related to steric factors (i.e., n-butylamine > tert-butyl
PEP, the total concentration of vanadyl porphyrins will be amine>diethylamine.triethylamine). Ozawa and Hanaki50
underestimated. compared the ligation tendency of three different oxo-tetra-
Peripheral substitution of the porphyrin macrocycle af- phenyl porphyrins: oxo-vanadium, oxo-chromium, and oxo-
fects the location of the absorption peaks of the electronic titanium. They found that the tendency of these metals to
spectra since they will impact the electronic structure of the coordinate with Lewis bases was directly related to the
porphyrin macrocycle.40 The electronic absorption of UV/ distance of the metal atom to the nitrogen basal plane. Oxo-
visible radiation by porphyrins is attributed to π-π* tran- titanium, which has the largest distance, showed little or
sitions,40-42 and therefore anything that affects the electron no tendency to coordinate with Lewis bases, while oxo-
structure will have an affect on the location and intensity of chromium, with the shortest distance, showed the highest
the absorption. The possibility of n-π* transitions are not tendency to coordinate. The fact that the distance to the
considered likely because of the symmetry of the n orbitals porphyrin basal plane dictates the tendency to coordinate
and because of the antisymmetry of the π* orbitals relative to would indicate that the coordination occurs axially with the
the plane of the porphyrin macrocycle.41 A more extensive metal center. Ferrer and Baran36 observed a similar pheno-
discussion on the effect of peripheral substituents on the menon where solvents capable of axial interactions (DMSO,
electronic absorption of the porphyrin macrocycle using DMF, methanol, and pyridine) produced a significant impact
the so-called “Mutual Atomic Effect” was presented by on the shape of the spectra.
Berezin.43 At this point, it is concluded that differences in This tendency of the metal porphyrins to coordinate with
peripheral substitution of the porphyrin macrocycle will various Lewis base heteroatoms could explain their tendency
have an impact on both the location and magnitude of
electronic absorption peaks and must be accounted for when (44) White, W. I. Aggregation of Porphyrins and Metalloporphyrins.
applying UV/visible spectroscopy for quantitation. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New York, 1978; Vol.
5, pp 303-339.
(45) Walker, F. A.; Hui, E.; Walker, J. M. J. Am. Chem. Soc. 1975, 97,
(39) Trofimenko, G. M.; Semeikin, A. S.; Berezin, M. B.; Berezin, 2390–2397.
B. D. Russ. J. Coord. Chem. 1996, 22, 476–480. (46) Bonnett, R.; Brewer, P.; Noro, K.; Noro, T. Tetrahedron 1978,
(40) Berezin, D. B.; Andrianov, V. G.; Semeikin, A. S. Opt. Spectrosc. 34, 379–385.
1996, 80, 554–561. (47) Drew, M. G. B.; Mitchell, P. C. H.; Scott, C. E. Inorg. Chim. Acta
(41) Berezin, B. D. Coordination Compounds of Porphyrins and 1984, 82, 63–68.
Phthalocyanines; John Wiley & Sons: Chichester, U. K., 1981. (48) Shelnutt, J. A.; Findsen, E. W.; Ondrias, M. R.; Alston, K. ACS
(42) Gouterman, M. Optical Spectra and Electronic Structure of Symp. Ser. 1987, 344, 368–383.
Porphyrins and Related Rings. In The Porphyrins; Dolphin, D., Ed.; (49) Bencosme, C. S.; Romero, C.; Simoni, S. Inorg. Chem. 1985, 24,
Academic Press: New York, 1978; Vol. III - Physical Chemi, pp 1-165. 1603–1604.
(43) Berezin, B. D. J. Appl. Spectrosc. 1999, 66, 521–527. (50) Ozawa, T.; Hanaki, A. Inorg. Chim. Acta 1988, 141, 49–51.
2800
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

to associate with the asphaltenes in petroleum samples. The occurs at a wavelength other than that being scanned for
asphaltenes are known to be enriched in heteroatoms (N, O, in the visible region. During hydrotreating, these large
and S) relative to the native petroleum. The metalloporphyr- peripheral groups are removed from the porphyrin
ins contained within the asphaltene fraction are likely co- backbone, causing a shift in the absorbance wavelength
ordinated with heteroatoms within the asphaltene molecules. to the region being monitored, resulting in an apparent
As indicated above, however, steric hindrance plays a major increase in porphyrin concentration
role in this additional association at the vanadyl center, and 2. Large polar asphaltene molecules, which retain metallo-
therefore only those heteroatoms which are present at a porphyrin molecules via molecular association, are con-
pendant position within the asphaltene molecule will be verted during hydrotreating, thus releasing metallopor-
accessible to such a coordination bond. This type of ligand phyrins. This type of molecular association is capable of
formation would also help to explain the absence of the Soret changing the electronic absorption wavelength of a
and visible bands at the usual locations for vanadyl porphy- species and hence would shift the wavelength outside
rins. As was observed for model systems, the formation of an of the region being monitored.Further work by these
additional ligand bond with the vanadium center causes a investigators61 with other vacuum residue fractions
significant shift in the location of the absorption bands, failed to reproduce this same appearance of porphyrins.
which may or may not cause them to pass undetected.
Freeman and O’Haver62 applied derivative UV/visible
The steric hindrance discussed above also reduces the like-
spectroscopy to quantify the concentration of metallopor-
lihood of dimerization of vanadyl porphyrins.51 Symmetrical
phyrins in deasphalted bitumen samples. These investigators
porphyrins such as VOTPP do not form significant amounts of
found significant absorbance by nonporphyrinic com-
a dimeric species,51 while unsymmetrical porphyrins show a
pounds in the region of the Soret band and therefore had
higher tendency to aggregate due to the ability to form head-to-
to resort to using the R band. In order to counteract the
toe species. Although the tendency of vanadyl porphyrins to
reduced sensitivity of this peak, data smoothing and second
self-dimerize is well documented,52-55 the conditions generally
derivative algorithms were applied to the spectrum. These
present in vacuum residue and asphaltene solutions should not
data analysis algorithms serve to improve the signal-to-noise
result in any significant dimerization.44,52
ratio, which is a major difficulty when dealing with complex
3.2.4. UV/visible Spectroscopy of Authentic Petroporphyrins.
A number of authors have attempted to quantify metallo- mixtures where significant background absorbance is pre-
porphyrins in crude oil by using the extinction coefficients of sent. In the end, they concluded that the optimal algorithm
isolated vanadyl compounds or model compounds. Groenn- was a second derivative, three-point sliding average algo-
ings56 treated crude oil with inorganic acids and organic rithm. Freeman et al.37 extended this analysis further by
solvents to isolate a fraction rich in demetalated porphyrins, applying third derivative UV/visible spectroscopy for the
which were then used to calibrate the analysis of the qualitative identification of metalloporphyrins. The use of
whole crude oil. Similar work by Sugihara and Bean,21 Biggs the third derivative of the absorbance allows for a much
et al., 57 Reynolds et al.,58 Reynolds et al.59 and Filby more precise identification of the exact wavelength (to within
and Strong14 found that the absorbance of the Soret band (0.1 nm) of an absorbance maximum since the third deri-
was 40-50% of the expected value based on model com- vative is characterized by a steep zero crossing at an absor-
pounds. In each case, this calculation was based on the bance peak. This method allowed the investigators to
extinction coefficient of isolated porphyrin compounds in differentiate a number of different metalloporphyrins on
solution. the basis of the UV/visible spectra alone. Unfortunately,
Van Berkel and Filby,60 and later Pearson and Green,11,12 this method requires that the petroporphyrins be separa-
studied the effect of thermal processing and hydrotreating, ted and/or purified prior to analysis. The complex mixtu-
respectively, on the distribution of vanadium and nickel res in vacuum residue cannot be analyzed directly to identify
compounds. They found that hydrotreating the crude oils the different porphyrins because of significant spectral
resulted in an increase in the porphyrinic vanadium content interferences.
for certain fractions. The increase is attributed to one of two Unfortunately, the majority of the UV/visible methods
possible causes: employed for the quantitation of metalloporphyrins in sam-
ples rely on chemical extraction of these species. As discussed
1. Hydrotreating leads to a change in the peripheral sub-
in section 4.3, it is unlikely that chemical extraction methods
stitution of highly substituted porphyrins. For porphyr-
would be capable of completely extracting all of the metallo-
ins with peripheral substitution other than octaethyl,
porphyrins present in a sample of vacuum residue or asphal-
etio, or tetraphenyl, it is possible that the absorbance
tene. This is further compounded by the fact that asphal-
tenes are suspected to form aggregated colloidal structures
(51) Boyd, P. D. W.; Smith, T. D.; Price, J. H.; Pilbrow, J. R. J. Chem. in most organic solvents, which could further deter the
Phys. 1972, 56, 1253–1263.
(52) Banci, L. Inorg. Chem. 1985, 24, 782–786. extraction of the metalloporphyrins from the asphaltene
(53) Maiya, G. B.; Krishnan, V. Inorg. Chem. 1985, 24, 3253–3257. phase.
(54) Lemtur, A.; Chakravorty, B. K.; Dhar, T. K.; Subramanian, J. Another major barrier to effective quantitation of vanadyl
J. Phys. Chem. 1984, 88, 5603–5608.
(55) Thanabal, V.; Krishnan, V. Inorg. Chem. 1982, 21, 3606–3613. porphyrins by UV/visible spectroscopy is the significant
(56) Groennings, S. Anal. Chem. 1953, 25, 938–941. absorption of UV/visible radiation by polycyclic aromatic
(57) Biggs, W. R.; Fetzer, J. C.; Brown, R. J.; Reynolds, J. G. Liq. hydrocarbons (PAHs). As the number of aromatic centers
Fuels Technol. 1985, 3, 397–421.
(58) Reynolds, J. G.; Biggs, W. R.; Bezman, S. A. ACS Symp. Ser. increases and as the degree of conjugation increases, the
1987, 344, 205–219. absorbance peak shifts to longer wavelengths due to a higher
(59) Reynolds, J. G.; Jones, E. L.; Bennett, J. A.; Biggs, W. R. Fuel
Sci. Techn. Int. 1989, 7, 625–642.
(60) Van Berkel, G. J.; Filby, R. H. ACS Symp. Ser. 1987, 344, 110– (61) Pearson, C. D.; Green, J. B. Energy Fuels 1993, 7, 338–346.
134. (62) Freeman, D. H.; O’Haver, T. C. Energy Fuels 1990, 4, 688–694.
2801
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray
63 64
electron density. Halasinski et al. recorded the UV/visible similarity of the curve to that for pure VOEtio indicate that the
spectra of a homologous series of PAHs, perylene, terrylene, vanadium species in all of the petroleum samples studied were
and quaterrylene, and observed absorption peaks in the region in the þ4 valence state (i.e., V4þ as in VO2þ).
of the vanadyl porphyrin bands (both Soret and visible) for all Tynan and Yen69 compared the type and shape of EPR
three molecules. Since asphaltenes are large PAHs with sizes spectra obtained for both authentic vanadium compounds in
resembling these homologues, it is likely that asphaltene mole- asphaltenes as well as for model vanadium compounds
cules will absorb radiation in the vicinity of the vanadyl added to a vanadium free asphaltene in various solvents.
porphyrin peaks and could easily mask their presence. Yokota For nonaggregated vanadium complexes (referred to as
et al.65 used UV/visible spectroscopy to monitor the absorption free), the EPR spectrum is said to be “isotropic” and displays
spectra of Athabasca asphaltenes as well as various fractions of a characteristic derivative spectrum, while an aggregated
said asphaltenes separated by SEC. The whole asphaltenes and vanadium compound displays a characteristic “anisotropic”
subsequent subfractions exhibited large structureless absorp- spectrum. When asphaltenes were added to a model vanadyl
tion bands covering the range 270-450 nm, which encompasses phthalocyanine complex, the EPR spectra shifted from an
the Soret band of the vanadyl porphyrins. This significant isotropic to an anisotropic form indicating association of the
interference from the asphaltenes poses significant problems vanadium complex with the asphaltenes. As the concentra-
for the quantitation of vanadyl porphyrins since the Soret peak tion of asphaltenes increased, so did the degree of aniso-
in most cases is virtually invisible. tropy. A similar effect was observed with the authentic
Antipenko and Zemtseva66 studied the impact of various vanadium compounds present in an asphaltene sample.
chemical species on the electronic spectra of vanadyl porphyrins. When the temperature of the solution was varied at a fixed
A native Russian oil was separated into numerous different asphaltene concentration, the degree of anisotropy de-
representative fractions including the usual SARA fractions creased with increasing temperature. Using the variation of
(saturates þ aromatics, asphaltenes, and resins), nitrogen base EPR derivative peak height with temperature, the investiga-
fractions, sulfur fractions, naphthenic acids, and other func- tors calculated a vanadium-asphaltene energy of associa-
tional groupings. In the case of VOEtio, significant changes in tion of 59.8 kJ/mol, regardless of the solvent used. When
the magnitude of the Soret band were observed in the presence of compared to approximate values for general van der Waals
various pure compounds and oil fractions. In particular, the forces70 (0.1-10 kJ/mol) and hydrogen bonds70 (10-40 kJ/
nitrogen bases and the saturates þ aromatics extracted from the mol), this value indicates relatively strong interactions, likely
native oil sample decreased the intensity of the Soret band by consisting of more than one noncovalent bond. Of the
over 30%, while some pure PAH compounds, in particular solvents studied (diphenylmethane, benzyl n-butyl ether,
phenanthrene and anthracene, also decreased the intensity of the 1-ethylnaphthalene, benzene, nitrobenzene, pyridine, and
Soret band by 30-40%. Carbazole, a model aromatic nitrogen tetrahydrofuran), tetrahydrofuran produced the most iso-
compound, also decreased the intensity of the Soret band by over tropic (free) vanadium. In a similar study, Selyutin et al.71
40%. Therefore, the lack of a vanadyl porphyrin Soret band in used ESR spectroscopy to monitor the dimerization of pure
asphaltenes could also be attributed to chemical interactions VOOEP. They found that, although at low temperatures
between the asphaltene molecules and the vanadyl porphyrins. (77 K) VOOEP showed dimerization tendencies, at 20 °C in
3.3. Other Analytical Techniques. 3.3.1. Electron Para- solution the dimeric species disappeared.
magnetic Resonance (EPR) Spectroscopy. Electron paramag- Many attempts have been made in the past to use EPR
netic resonance spectroscopy (EPR), also known as electron spectra to determine the bonding structure of the vanadium
spin resonance spectroscopy (ESR), measures the resonant species present in heavy oil and asphaltene samples. Yen
absorption of microwave radiation in the presence of a static et al.72 combined EPR and UV/visible spectroscopy and
magnetic field. The resonant absorption of the radiation in concluded that some of the vanadium was present as non-
the presence of the magnetic field occurs when there is a porphyrins. Dickson et al.73 and Dickson and Petrakis74
magnetic dipole created by the net spin or net orbital angular compared the spin Hamiltonian parameters of native vana-
momentum of an unpaired electron.67 Such a paramagnetic dium complexes and pure vanadium complexes73 obtained
resonance arises in a quadrivalent vanadium atom (V4þ) as a by fitting EPR spectra and concluded that a large portion of
result of an unpaired electron present in a 3d orbital. the vanadium compounds in the oil samples were present
Saraceno et al.68 compared the derivative EPR spectra of with ligand structures other than the four-nitrogen coordi-
pure VOEtio to that for a residue fraction with a similar nation sphere. Reynolds et al.75,76 also concluded on the
concentration of vanadium and found a peak at the same
characteristic location for vanadium. The height of this vana- (69) Tynan, E. C.; Yen, T. F. Fuel 1969, 48, 191–208.
dium peak varied linearly with total concentration with a slope (70) Ebbing, D. D.; Gammon, S. D. General Chemistry, 6th ed.;
closely matching the slope of a calibration line obtained using Houghton Mifflin Company: Boston, MA, 1999.
pure VOEtio dissolved in a vanadium free heavy oil distillate. (71) Selyutin, G. E.; Shklyaev, A. A.; Eletskii, N. P.; Anufrienko, V.
F.; Titov, V. I. Sov. J. Coord. Chem. 1979, 5, 1041–1046.
The linearity of the peak height vs concentration plots and the (72) Yen, T. F.; Boucher, L. J.; Dickie, J. P.; Tynan, E. C.; Vaughan,
G. B. J. Inst. Petrol 1969, 55, 87–99.
(73) Dickson, F. E.; Kunesh, C. J.; Mcginnis, E. L.; Petrakis, L. Anal.
(63) Rao, C. N. R. Ultra-Violet and Visible Spectroscopy: Chemical Chem. 1972, 44, 978–981.
Applications, 3rd ed.; Butterworths: London, 1975. (74) Dickson, F. E.; Petrakis, L. Anal. Chem. 1974, 46, 1129–1130.
(64) Halasinski, T. M.; Weisman, J. L.; Ruiterkamp, R.; Lee, T. J.; (75) Reynolds, J. G.; Biggs, W. R.; Fetzer, J. C. Liq. Fuels Technol.
Salama, F.; Head-Gordon, M. J. Phys. Chem. A 2003, 107, 3660–3669. 1985, 3, 423–448.
(65) Yokota, T.; Scriven, F.; Montgomery, D. S.; Strausz, O. P. Fuel (76) Reynolds, J. G.; Biggs, W. R.; Fetzer, J. C.; Gallegos, E. J.; Fish,
1986, 65, 1142–1149. R. H.; Komlenic, J. J.; Wines, B. K. In Molecular Characterization of
(66) Antipenko, V. R.; Zemtseva, L. I. Petrol. Chem. 1996, 36, 31–42. Vanadyl and Nickel Non-Porphyrin Compounds in Heavy Crude
(67) Wertz, J. E.; Bolton, J. R. Electron Spin Resonance: Elementary Petroleums and Residua, Characterization of heavy crude oils and
Theory and Practical Applications; McGraw Hill: New York, 1972. petroleum residues: International Symposium on Characterization of
(68) Saraceno, A. J.; Coggeshall, N. D.; Fanale, D. T. Anal. Chem. Heavy Crude Oils and Petroleum Residues, Lyon, France, June 25-27,
1961, 33, 500–505. 1984; Editions Technip: Paris, pp 153-157.
2802
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

basis of EPR spin Hamiltonian parameters that the vana- In order for the method to be valid, such interactions must
dium was present in nonporphyrin coordination, although be avoided or minimized.80,81
they were unable to pinpoint an exact coordination type. A number of studies have combined the SEC of asphal-
EPR spin Hamiltonian parameters obtained by Graham77 tenes with an analysis of fractions for metal content,
for Boscan and Circle Cliffs asphaltenes were similar to including Fish and Komlenic,82 Fish et al.,83 Biggs
VOOEP. Finally, Reynolds8 and Reynolds et al.78 deter- et al.,57,84 Reynolds et al.,58 Fish et al.,85 Reynolds and
mined EPR spin Hamiltonian parameters for resins and Biggs,86 Sundararaman et al.,87 and Reynolds et al.59
asphaltene fractions and once again concluded on the basis Unfortunately, the SEC separation of a polydisperse
of these parameters that there are nonporphyrin vanadium mixture like asphaltenes is not selective for molecular size
coordination environments. (Davison et al.88). The tendency for asphaltenes to associ-
The majority of the work done using spin Hamiltonian ate in solution changes their apparent size significantly. In
parameters (the g-factor, go, and the hyperfine splitting addition, asphaltene fractions have a tendency to be
constant, Ao) reports and differentiates these values to an adsorbed on the column (chemical interaction), and the
extremely high precision (five significant figures for go). Very elution order of polar fractions is significantly altered by
few analytical techniques are capable of this level of preci- the choice of solvent. Both of these effects can shift the
sion, particularly when nonlinear regression of a highly apparent size of the eluting material. As well, the response
complex model such as the spin Hamiltonian is used. Also, of asphalt samples in an SEC experiment deviates signifi-
asphaltene samples are heterogeneous, and therefore the cantly from the response of polystyrene standards to a
calculated values for go are average values for the entire similar experimental setup.88 The use of polystyrene stan-
sample. This level of precision is probably not justified. dards for calibrating the MW vs elution volume response
Even at a precision of 0.1%, which for most analytical of the experiment is not valid for asphaltene samples.
techniques would be deemed excellent, the isotropic go value These problems make the method invalid for drawing
would only be accurate to the third decimal place and conclusions regarding the size and ligand structure of the
would not reveal any significant differences. Malhotra and vanadium compounds present in the asphaltene fractions
Buckmaster79 used high precision 34 GHz EPR spectroscopy of crude oil samples.
along with statistical analysis of variance methods to con- 3.3.3. Interaction of Radioactive Tracers with Asphaltenes.
Nguyen and Filby89 used nickel complexes containing
clude that the spin Hamiltonian parameters derived from
radioactive nickel isotopes (63Ni) to monitor the interac-
EPR spectra are not capable of differentiating between
tion of model nickel compounds (both porphyrin and
different coordination structures around the vanadyl ion.
nonporphyrin) with Athabasca asphaltenes. When added
This implies that EPR spectroscopy is not a suitable tool for
to a solution of asphaltenes, the model nickel compounds
identifying the ligand structure of vanadium in vacuum
adsorbed/interacted with the asphaltenes. It was not clear
residue and asphaltene samples. This method has, however, whether this interaction was due to association (i.e.,
given us strong proof that the vast majority (if not all) of the chemisorption) or due to ligand interactions between
vanadium present in petroleum samples is in the form of nickel and functional groups within the asphaltene mole-
vanadyl ions. As well, EPR spectroscopy is of value in cules. Attempts to coprecipitate the model compounds
assessing the aggregation characteristics of the vanadium with the asphaltenes found that changing the precipita-
complexes, as shown in the work by Tynan and Yen69 and tion solvent had little or no impact on the amount of
Selyutin et al.71 model nickel compounds present in the precipitated as-
3.3.2. Size Exclusion Chromatography (SEC) with Element phaltenes. Once again, they were unable to determine
Specific Detection. One general class of methods used to whether the asphaltene-Ni interaction was by a π-π
study the form of vanadium compounds involves the use of interaction or by axial bonding of asphaltene functional
size exclusion chromatography (SEC, also known as gel groups and the Ni2þ ion. Unfortunately, this work did not
permeation chromatography or GPC). The basic premise investigate the asphaltene-nickel interaction in stronger
of SEC is that the difference in retention times of different asphaltene solvents where the asphaltene species are not
compounds within the column is brought about exclusively present as precipitates but rather as smaller colloidal
on the basis of molecular size. Small molecules penetrate the particles.
porous network of the column packing and are retained in 3.3.4. Mass Spectrometry. Mass spectrometry has been
the column for a longer period of time. Larger molecules do applied by numerous investigators to identify the molecular
not penetrate as extensively and are eluted from the column forms of vanadyl porphyrins, as well as the distribution of the
sooner. In this manner, different molecules are separated on
the basis of their size and hence the name size exclusion
chromatography. One of the primary assumptions in the (82) Fish, R. H.; Komlenic, J. J. Anal. Chem. 1984, 56, 510–517.
application of SEC is that there are no chemical interactions (83) Fish, R. H.; Komlenic, J. J.; Wines, B. K. Anal. Chem. 1984, 56,
2452–2460.
occurring between the sample and the column packing. (84) Biggs, W. R.; Brown, R. J.; Fetzer, J. C. Energy Fuels 1987, 1,
257–262.
(85) Fish, R. H.; Reynolds, J. G.; Gallegos, E. J. ACS Symp. Ser.
(77) Graham, W. R. M. ACS Symp. Ser. 1987, 344, 358–367. 1987, 344, 332–349.
(78) Reynolds, J. G.; Gallegos, E. J.; Fish, R. H.; Komlenic, J. J. (86) Reynolds, J. G.; Biggs, W. R. Prepr.;Am. Chem. Soc., Div. Pet.
Energy Fuels 1987, 1, 36–44. Chem. 1987, 32, 398–405.
(79) Malhotra, V. M.; Buckmaster, H. A. Fuel 1985, 64, 335–341. (87) Sundararaman, P.; Biggs, W. R.; Reynolds, J. G.; Fetzer, J. C.
(80) Malawer, E. G.; Senak, L. Introduction to Size Exclusion Geochim. Cosmochim. Acta 1988, 52, 2337–2341.
Chromatography. In Handbook of Size Exclusion Chromatography and (88) Davison, R. R.; Glover, C. J.; Burr, B. L.; Bullin, J. A. Size
Related Techniques, 2nd ed.; Wu, C.-s., Ed.; Marcel Dekker, Inc.: New York, Exclusion Chromatography of Asphalts. In Handbook of Size Exclusion
2004; Chapter 1. Chromatography and Related Techniques, 2nd ed.; Wu, C.-s., Ed.; Marcel
(81) Skoog, D. A.; Leary, J. J. Principles of Instrumental Analysis, 4th Dekker, Inc.: New York, 2004; Chapter 8.
ed.; Saunders College Publishing: Fort Worth, TX, 1992. (89) Nguyen, S. N.; Filby, R. H. ACS Symp. Ser. 1987, 344, 384–401.
2803
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray
90-96
various molecular forms in crude oil samples. Mass using atmospheric pressure photoionization (APPI) FT-
spectrometry is a key analytical method for qualitatively iden- ICR-MS. Not only did they identify homologues of several
tifying the different porphyrinic forms identified in Figure 1. of the expected forms of vanadyl porphyrins (VODPEP,
Beato et al.97 used electron ionization tandem mass spectro- VORhodo, and VOEtio), they also identified several sulfur-
metry (EIMS/MS) to study vanadium compounds in a containing vanadyl porphyrins. This report is the first evi-
New Albany bitumen and its pyrolysate using the same dence of sulfur species directly attached to vanadyl porphyrins.
separation method as Van Berkel and Filby.60 Their results This is also added proof of the complex nature of the
indicated that both the bitumen and pyrolysate contained molecular environment surrounding vanadium atoms within
vanadyl porphyrins of similar structure. They also concluded the asphaltene fraction.
that the appearance of porphyrins in pyrolysate is due to an 3.4. Final Assessment. Although a number of methods have
enhanced solubilization/desorption mechanism and not due been used to determine the chemical structures of the organo-
to C-C bond scission, indicating that metalloporphyrins are vanadium compounds present in crude oils, examination of
held in the asphaltene/kerogen matrix by association rather the primary chemical environment of the vanadium by X-ray
than chemically bound. spectroscopies (EXAFS, XANES, and XPS) gives a consis-
Grigsby and Green98 used low-eV high-resolution mass tent result; the nearest neighbor atoms surrounding the
spectrometry (HR/MS) to characterize the nonporphyrinic vanadium atom are a single oxygen atom (making a vanadyl
vanadium present in the >700 °C resid fraction of a Cerro- ion) and four nitrogen atoms in a square pyramidal structure
Negro crude. They identified several porphyrinic vanadyl corresponding to the porphyrin macrocycle. The agreement
structures within the non-UV absorbing fraction. They between the spectra for the model vanadyl porphyrins and the
explained the discrepancy by identifying several cycloalkyl nonabsorbing petroporphyrins is excellent.
and aromatic substituents on the periphery of the vanadyl The majority of studies that identified vanadyl porphyrins
porphyrin macrocycle, which would lead to much lower from crude oil relied on solvent extracts that did not include
response in the visible region, thus resulting in an under- all of the vanadium present in the original crude oil; there-
estimation of the porphyrin content. They were also able to fore, they provide at best a partial picture of the secondary
identify homologues of the Etio and DPEP forms containing structure of the vanadium species. By analyzing unfractio-
additional carbons attached as peripheral substituents. nated asphaltenes, Qian et al.18 were able to indentify
Rodgers et al.99 used electrospray ionization Fourier trans- elemental compositions consistent with homologous series
form ion cyclotron resonance mass spectrometry (ESI-FT-ICR- with one or more fused benzene rings, one or more fused
MS) to characterize the petroporphyrin fractions of a Cerro cycloalkyl rings, and vanadyl porphyrins with sulfur. This
Negro crude oil extracted using chromatography and solvent- result indicates a much more diverse range of chemical
based extraction methods. They identified homologues of the structures for vanadium than previous studies. The periph-
VOEtio and VODPEP forms containing additional carbons, eral substitution of the porphyrin ring, coupled with aggre-
indicating variation in peripheral substitutions. Their analysis gation with other asphaltenic species, would substantially
also indicated the presence of dimers of vanadyl porphyrins. change the intensity and position of the Soret band.
However, their analysis requires the addition of ionic species The evidence suggests, therefore, that the UV-visible studies
(Hþ and Naþ) to charge the analytes for analysis. The presence of porphyrin concentrations in crude oil suffered from overly
of ionic species, and in particular the intentional charging of the simplistic assumptions. By ignoring the effect of molecular
porphyrin, can have a significant impact on the dimerization association in solution and peripheral substitution on the
of the porphyrin monomers.53,55 Also, methanol was used to extinction coefficients at a given wavelength, they likely reached
produce the analyte solutions. Methanol has already been a false conclusion and attempted to define a “non-porphyrin”
identified as a poor solvent for vanadyl porphyrins with a fraction of vanadium. The evidence indicates that all of the
maximum solubility of 3 μg/mL.34 The solutions used by vanadium is in vanadyl porphyrins, although the substitution
Rodgers et al.99 are very close to this solubility limit which on the porphyrin ring may be highly variable and extensive.
could result in the formation of aggregates and/or crystals as a
result of exceeding saturation in the solvent. 4. Aggregation and Solubility Characteristics of Asphaltenes
Generally speaking, the samples analyzed with mass spec- and Metalloporphyrins
trometry are chemical extracts of the petroleum rather than
the whole sample or even an asphaltene sample. These Asphaltenes can adopt a colloidal character in solution,
methods cannot, therefore, help elucidate the form of the based on various analytical methods including small angle
vanadyl porphyrins contained with the asphaltene fraction. neutron scattering (SANS),100-102 small-angle X-ray scat-
However, Qian et al.18 analyzed a whole asphaltene sample tering (SAXS),103-106 vapor phase osmometry (VPO),107-111

(90) Baker, E. W.; Palmer, S. E. Geochemistry of Porphyrins. In The (100) Tanaka, R.; Hunt, J. E.; Winans, R. E.; Thiyagarajan, P.; Sato,
Porphyrins; Dolphin, D., Ed.; Academic Press: New York, 1978; Vol. I, pp S.; Takanohashi, T. Energy Fuels 2003, 17, 127–134.
486-552. (101) Spiecker, P. M.; Gawrys, K. L.; Kilpatrick, P. K. J. Colloid
(91) Blumer, M.; Rudrum, M. J. Inst. Petrol. 1970, 57, 99–106. Interface Sci. 2003, 267, 178–193.
(92) Baker, E. W.; Yen, T. F.; Dickie, J. P.; Rhodes, R. E.; Clark, L. F. (102) Thiyagarajan, P.; Hunt, J. E.; Winans, R. E.; Anderson, K. B.;
J. Am. Chem. Soc. 1967, 89, 3631–3639. Miller, J. T. Energy Fuels 1995, 9, 829–833.
(93) Blumer, T. Geochim. Cosmochim. Acta 1964, 28, 1147–1154. (103) Cosultchi, A.; Bosch, P.; Lara, V. H. Colloid Polym. Sci. 2003,
(94) Strong, D.; Filby, R. H. ACS Symp. Ser. 1987, 344, 154–172. 281, 325–330.
(95) Van Berkel, G. J.; Quino~nes, M. A.; Quirke, M. E. J. Energy Fuels (104) Savvidis, T. G.; Fenistein, D.; Barre, L.; Behar, E. AIChE J.
1993, 7, 411–419. 2001, 47, 206–211.
(96) Xu, H.; Que, G.; Yu, D.; Lu, J. R. Energy Fuels 2005, 19, 517–524. (105) Bardon, C.; Barre, L.; Espinat, D.; Guille, V.; Li, M. H.;
(97) Beato, B. D.; Yost, R. A.; Van Berkel, G. J.; Filby, R. H.; Quirke, Lambard, J.; Ravey, J. C.; Rosenberg, E.; Zemb, T. Fuel Sci. Technol.
J. M. E. Org. Geochem. 1991, 17, 93–105. Int. 1996, 14, 203–242.
(98) Grigsby, R. D.; Green, J. B. Energy Fuels 1997, 11, 602–609. (106) Kim, H.-g.; Long, R. B. Ind. Eng. Chem. Fund. 1979, 18, 60–63.
(99) Rodgers, R. P.; Hendrickson, C. L.; Emmett, M. R.; Marshall, A. (107) Strausz, O. P.; Peng, P.; Murgich, J. Energy Fuels 2002, 16, 809–
G.; Greaney, M.; Qian, K. Can. J. Chem. 2001, 79, 546–551. 822.
2804
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

and to a lesser extent electron paramagnetic resonance of an insoluble colloidal particle of asphaltene molecules,
spectroscopy (EPR).112 This tendency to aggregate and forming a steric barrier preventing further aggregation.121
precipitate in n-alkane solvents (primarily n-pentane or Upon the addition of a low polarity, low solubility-
n-heptane) provides the operational definition for asphal- parameter solvent, these resin molecules are dissolved, thus
tenes. The molecular aggregation of asphaltenes has been exposing the insoluble asphaltene core and leading to floc-
detected at surprisingly high temperatures, suggesting that culation and precipitation. Subsequent authors proposed
the aggregates may be remarkably stable in the heavy oil or smaller clusters of asphaltenes and suggested an analogy to
bitumen.100,113 As mentioned previously, the majority of surfactant micelles,122 eventually reaching the limit sug-
the vanadyl porphyrins are concentrated in this asphaltene gested by Speight123 that only a single asphaltene molecule
fraction. This partitioning of the vanadyl compounds in is surrounded by resins. In essence, the resins behave as a
the asphaltene fraction could be attributed to their low surfactant for the asphaltene molecules. Recent work by
solubility in most (if not all) solvents, or it could be because Andersen and co-workers124-129 using microcalorimetry
the porphyrins are molecularly associated to asphaltene showed that asphaltenes in toluene do not fit the behavior
molecules. of a micelle-like system (i.e., critical micelle concentration).
4.1. Solubility Modeling of Asphaltene Precipitation. The Rather, the asphaltene-toluene system fits the behavior of a
lack of solubility of porphyrins and their various derivatives stepwise association with aggregates of variable size similar
in organic solvents has been well documented.114-119 As to the dye Rhodamine 6G.128 Therefore, the analogy of
discussed in section 4.3, Freeman et al.34 determined the asphaltene aggregates to micelles is false and the use of the
solubility of VOOEP and VOEtio in various solvents and term “micelle” in conjunction with asphaltene behavior is
found them to be on the order of 10-5 mol/L (103 ppmw). not recommended. If this model is accurate, then the metallo-
The type and location of peripheral substituents does have an porphyrin molecules are not likely to associate with the
impact on the solubility of the porphyrin macrocycle. asphaltenes within the crude oil. Rather, the highly aromatic
Trofimenko et al.39 showed that the solubility of four vanadyl porphyrins are also bound to resins in a similar
isomers of Etioporphyrin (free base form) can vary by as manner. Upon the addition of the solvent and solvation of
much as an order of magnitude due to changes in the crystal the resins, the metalloporphyrins are released and precipitate
packing structure of the different isomers. Salcedo et al.120 with the asphaltenes.
applied DFT theory to investigate the effect of periphe- The second model adopts a thermodynamic approach to
ral substitution on solubility of vanadyl porphyrins and explain the stabilizing effect of the resins. This model attri-
found that the addition of peripheral substituents leads butes the destabilization of the asphaltene micelles to a
to different degrees of out of plane distortion of the porphyr- decrease in the solvent power of the medium toward the
in macrocycle, which in turn leads to variations in the asphaltene monomers. As the solvent power of the surround-
dipole moments of the metalloporphyrin. According to their ing medium decreases, the solubility of the asphaltene mono-
model, it is this porphyrin dipole moment, coupled with the mers decreases up to a point where these monomers are no
solvent dipole moment, that dictates the solubility. Many longer fully soluble. Because the size of the asphaltene
solvents including aromatics, ketones, alcohols, amines, etc. aggregates is dependent on temperature, concentration,
have been tested with little or no success for complete and solvent identity, the aggregation mechanism is thought
dissolution of porphyrins and metalloporphyrins. Therefore, to be reversible and hence governed by thermodynamic
the presence of the vanadyl porphyrins with the insolu- equilibrium. If this model is correct, then the high concen-
ble asphaltene fraction could easily be the result of a copreci- tration of metalloporphyrins in the asphaltene fraction is
pitation mechanism rather than a molecular association controlled by the solubility (or lack thereof) of the metallo-
mechanism porphyrins in the solvent medium and not by association of
Two models have been proposed to explain the role of the the metalloporphyrin molecules with the asphaltenes.
resins in stabilizing the asphaltene fraction within the oil. The Yarranton and co-workers have developed two types
first model proposes that the resins associate at the periphery of models, one for molecular association, to explain
apparent molecular weight measurements. Agrawala and
Yarranton110 proposed that the aggregation of the asphal-
(108) Andersen, S. I. Fuel Sci. Techn. Int. 1994, 12, 51–74. tene molecules occurs via a pseudopolymerization mecha-
(109) Yarranton, H. W.; Alboudwarej, H.; Jakher, R. Ind. Eng. Chem.
Res. 2000, 39, 2916–2924. nism whereby the aggregation of the asphaltene molecules
(110) Agrawala, M.; Yarranton, H. W. Ind. Eng. Chem. Res. 2001, 40, occurs via active heteroatom centers within the asphaltene
4664–4672.
(111) Yarranton, H. W.; Fox, W. A.; Svrcek, W. Y. Can. J. Chem.
Eng. 2007, 85, 635–642. (121) Pfeiffer, J. P.; Saal, R. N. J. J. Phys. Chem. 1940, 44, 139–149.
(112) Wong, G. K.; Yen, T. F. J. Petrol. Sci. Eng. 2000, 28, 55–64. (122) Yen, T. F.; Erdman, J. G.; Pollack, S. S. Anal. Chem. 1961, 33,
(113) Tanaka, R.; Sato, E.; Hunt, J. E.; Winans, R. E.; Sato, S.; 1587–1594.
Takanohashi, T. Energy Fuels 2004, 18, 1118–1125. (123) Speight, J. G. Asphaltenes and the Structure of Petroleum. In
(114) Mamardashvili, G. M.; Mamardashvili, N. Z.; Golubchicov, Petroleum Chemistry and Refining; Speight, J. G., Ed.; Taylor and Francis:
O. A.; Berezin, B. D. J. Mol. Liq. 2001, 91, 189–191. Washington, DC, 1998; pp 103-120.
(115) Mamardashvili, G. M.; Mamardashvili, N. Z.; Golubchikov, (124) Merino-Garcia, D.; Andersen, S. I. Langmuir 2004, 20, 4559–
O. A.; Berezin, B. D. Russ. J. Phys. Chem. 1999, 73, 894–898. 4565.
(116) Trofimenko, G. M.; Mamardashvili, N. Z.; Golubchikov, (125) Merino-Garcia, D.; Andersen, S. I. Langmuir 2004, 20, 1473–
O. A.; Berezin, B. D. Russ. J. Phys. Chem. 1997, 71, 240–243. 1480.
(117) Berezin, B. D.; Koifman, O. I.; Nikitina, G. E. Russ. J. Phys. (126) Merino-Garcia, D.; Murgich, J.; Andersen, S. I. Petrol. Sci.
Chem. 1980, 54, 1413–1415. Technol. 2004, 22, 735–758.
(118) Berezin, B. D.; Koifman, O. I.; Zelov, V. V.; Nikitina, G. E. (127) Juyal, P.; Merino-Garcia, D.; Andersen, S. I. Energy Fuels 2005,
Russ. J. Phys. Chem. 1978, 52, 1281–1282. 19, 1272–1281.
(119) Koifman, O. I.; Berezin, B. D.; Zelov, V. V.; Nikitina, G. E. (128) Merino-Garcia, D.; Andersen, S. I. J. Disper. Sci. Technol. 2005,
Russ. J. Phys. Chem. 1978, 52, 1032–1034. 26, 217–225.
(120) Salcedo, R.; Zaragoza, I. P.; Martinez-Magadan, J. M.; Garcia- (129) Verdier, S.; Plantier, F.; Bessieres, D.; Andersen, S. I.; Stenby,
Cruz, I. THEOCHEM 2003, 626, 195–201. E. H.; Carrier, H. Energy Fuels 2007, 21, 3583–3587.
2805
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

molecule. Resins act as a terminator in the process since they and therefore is not entirely predictive. More advanced
contain little or no active heteroatomic centers. In the molecular association models (e.g., SAFT) have been pro-
presence of a significant amount of resins, little or no posed138,139 to model the asphaltene precipitation pheno-
association of asphaltenes occurs. This idea was extended menon with moderate success, although the complexity of
to modeling of precipitation. Yarranton et al.110,130-133 these models makes them somewhat less attractive.
successfully modeled the precipitation behavior of asphal- In light of the success of these Flory-Huggins regular
tenes in various solvents using a modification of the Flory- solution models for predicting the extent of asphaltene
Huggins regular solution model initially developed by precipitation, this same model could also successfully model
Cimino et al.134,135 The asphaltene fraction is assumed to the precipitation behavior of the vanadyl porphyrins in
consist of monomeric asphaltene species which aggregate to similar solutions. In fact, the work of Freeman at al.34
form larger particles similar to polymerization. The pure indicates that the solubility parameter of the solvent does
component properties of the asphaltenic fraction (i.e., solu- play a significant role in the solubility of vanadyl porphyrins.
bility parameter, density, molar mass, etc.) are then modeled 4.2. Proposed Mechanisms For Asphaltene Aggregation.
using various fitted distribution functions incorporating this Akbarzadeh et al.140 investigated the association behavior of
stepwise association model. The model parameters were model pyrene compounds in solution with different types of
tuned to fit the experimental density and precipitation data. polar substituents. They found that the nonpolar polyaro-
One of the main drawbacks of the regular solution model matic compounds pyrene and dipyrenyl decane did not
of Yarranton et al. is that the results are highly sensitive to associate in o-dichlorobenzene. When ketone and hydroxyl
the values of the solubility parameters of the various com- functional groups were added to the molecules, significant
ponents. Unfortunately, these values for asphaltenes and association took place. This result supported the contention
various other SARA fractions are not well-known. In an that the associations in asphaltene molecules occur via
effort to overcome this shortcoming, Akbarzadeh et al.136 heteroatomic sites rather than only simple π interactions.
applied the SRK EOS to predict the heat of vaporization and Rakotondradany et al.141 performed similar tests using
density, and in turn the solubility parameter, of SARA symmetric, nonpolar alkyl substituted hexabenzocoronene
fractions. Their method relied on correlations obtained with (HBC) and observed significant association in toluene.
model PAHs to predict the pure component properties of
Molecular simulations indicated that the most stable asso-
SARA fractions based on molecular weight. The model was
ciation configurations occurred as a result of π stacking of
tuned to experimental density and precipitation data. This
the polyaromatic core. Tan et al.142,143 demonstrated that
approach was taken one step further by Sabbagh et al.,137
multiple interactions, including π-bonding between PAH
who modeled asphaltene precipitation as a liquid-liquid
groups, polar nitrogen groups, and hydrogen bonds in a
equilibrium using the Peng-Robinson EOS. Their metho-
dology was similar to that of Akbarzadeh et al.136 and single molecule could combine to give association in solution
produced similar results. when each single interaction was too weak on its own.
The fact that relatively simple thermodynamic theories Vanadium compounds have been implicated in schemes
such as the regular solution model and cubic EOSs can for aggregation of asphaltenes since Yen et al.69,72 Vanadyl
successfully predict the solid-liquid equilibrium behavior porphyrins are capable of forming additional ligand struc-
of a complex mixture of asphaltenes lends credibility to the tures with heteroatoms, and vanadyl porphyrins have a flat,
claim that the aggregation of asphaltenes is indeed a thermo- aromatic core rich in π electrons, very similar to the HBC
dynamically controlled equilibrium process. The use of core. Yudin et al.144 observed that, as the vanadium content
property distributions rather than discrete properties to of asphaltenes increases, so too does the tendency of the
describe the entire asphaltene fraction is appropriate since asphaltenes to aggregate. The expectation of π-bonding of
this fraction is composed of a large number of species and vanadyl porphyrins to other aromatic species was tested
hence would exhibit a range of properties. However, all of systematically by Yin et al.145 They found no evidence of
these models require fitting of experimental fractionation measurable interactions of VOOEP or VOTPP in solution
(SARA) and precipitation data and therefore are tuned to with PAHs, alkyl PAHs such as HBC, or pyrene derivatives.
the specific system being modeled. This requires a greater They suggested that the porphyrin ring must be more sub-
amount of experimental data to ensure reasonable accuracy stituted in order for association with asphaltenic com-
pounds to occur in solution. This observation implies that
(130) Akbarzadeh, K.; Alboudwarej, H.; Svrcek, W. Y.; Yarranton, simple vanadyl porphyrins will be at most weakly bound to
H. W. Fluid Phase Equilib. 2005, 232, 159–170.
(131) Akbarzadeh, K.; Dhillon, A.; Svrcek, W. Y.; Yarranton, H. W.
Energy Fuels 2004, 18, 1434–1441. (138) Buenrostro-Gonzalez, E.; Lira-Galeana, C.; Gil-Villegas, A.;
(132) Alboudwarej, H.; Akbarzadeh, K.; Beck, J.; Svrcek, W. Y.; Wu, J. AIChE J. 2004, 50, 2552–2570.
Yarranton, H. W. AIChE J. 2003, 49, 2948–2956. (139) Wu, J.; Prausnitz, J. M.; Firoozabadi, A. AIChE J. 2000, 46,
(133) Yarranton, H. W.; Masliyah, J. H. AIChE J. 1996, 42, 3533– 197–209.
3543. (140) Akbarzadeh, K.; Bressler, D. C.; Wang, J.; Gawrys, K. L.;
(134) Cimino, R.; Correra, S.; Sacomani, P. A.; Carniani, C. In Gray, M. R.; Kilpatrick, P. K.; Yarranton, H. W. Energy Fuels 2005, 19,
Thermodynamic modelling for prediction of asphaltene deposition in 1268–1271.
live oils. International Symposium on Oilfield Chemistry, Feb 14-17, (141) Rakotondradany, F.; Fenniri, H.; Rahimi, P.; Gawrys, K. L.;
1995, San Antonio, TX; Society of Petroleum Engineers (SPE): Kilpatrick, P. K.; Gray, M. R. Energy Fuels 2006, 20, 2439–2447.
Richardson, TX, 1995; pp 499-512. (142) Tan, X.; Fenniri, H.; Gray, M. R. Energy Fuels 2008, 22, 715–
(135) Cimino, R.; Correra, S.; Del Bianco, A.; Lockhart, T. P. 720.
Solubility and Phase behaviour of Asphaltenes in Hydrocarbon Media. (143) Tan, X.; Fenniri, H.; Gray, M. R. Energy Fuels 2009, 23, 3687–
In Asphaltenes: Fundamentals and Applications; Sheu, E. Y., Mullins, 3693.
O. C., Eds.; Plenum Press: New York, 1995; pp 97-130. (144) Yudin, I. K.; Nikolaenko, G. L.; Gorodetskii, E. E.; Markhashov,
(136) Akbarzadeh, K.; Ayatollahi, S.; Moshfeghian, M.; Alboudwarej, E. L.; Frot, D.; Briolant, Y.; Agayan, V. A.; Anisimov, M. A. Petrol. Sci.
H.; Yarranton, H. W. J. Can. Petrol. Technol. 2004, 43, 31–39. Technol. 1998, 16, 395–414.
(137) Sabbagh, O.; Akbarzadeh, K.; Badamchi-Zadeh, A.; Svrcek, (145) Yin, C.-X.; Tan, X.; Mullen, K.; Stryker, J. M.; Gray, M. R.
W. Y.; Yarranton, H. W. Energy Fuels 2006, 20, 625–634. Energy Fuels 2008, 22, 2465–2469.
2806
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

heteroatomic associations, and π-π associations likely lead


to the partitioning of the vanadyl porphyrins with the
asphaltene fraction.
4.3. Solution-Based Metalloporphyrin Extraction Methods.
A number of researchers have used solvent precipitation and
extraction schemes to enrich vanadium compounds in a
single fraction. The work of Freeman et al.34 showed that
vanadyl porphyrins have very low solubilities in common
organic solvents (see Figure 2). Chlorinated solvents
(chloroform and dichloromethane) showed the highest so-
lubility, while all other solvents showed very little solvent
power for the vanadyl porphyrins.
Several patents have been issued for processes to extract
vanadyl porphyrins using polar solvents such as 2-pyrroli-
done147 and butyrolactone,148 although the data provided in
the patents are for low vanadium oils with little or no
asphaltenes present. Overfield149 used a specific range of
Hansen solubility parameters to identify optimal solvents
and identified ethylene carbonate, propylene carbonate,
dimethyl sulfone, and ethylene trithiocarbonate as optimum
solvents for vanadyl porphyrin extraction. However, once
again, the data presented in the patent were for low vana-
Figure 2. Solubility of vanadyl porphyrins at 23 ( 2 °C as a function dium oil, although in this case it did contain asphaltenes
of the solubility parameter of the solvent.34 (vacuum residue). These solvents were capable of extracting
the vanadium compounds selectively, although the process
aggregates of asphaltenes, while more complex species will be required a large number of equilibrium stages to be effective.
strongly held. Galimov et al.150 examined the efficacy of a large variety of
Sirota146 presents a much different view of the observed polar solvents at their boiling point for the extraction of
asphaltene phase behavior. Rather than using a colloidal vanadyl porphyrins from Russian asphaltene samples. None
model to explain the precipitation and morphology of of the solvents examined were capable of extracting more
asphaltenes, Sirota adopts a liquid-liquid phase separation than 65% of the free porphyrins initially present.
model to explain the observed phenomena. He argues that Yin et al.151 chemically tagged the vanadyl porphyrins
the fractal morphology observed in precipitated asphaltenes within asphaltene samples by reacting with oxalyl chloride
can be explained on the basis of liquid viscosity and surface followed by a primary amine to replace the oxygen atom with
tension restrictions which prevent the coalescence of nano- nitrogen coupled to an octadecyl or perfluoroctyl side chain.
scale liquid droplets to form larger liquid regions. If these These tagged species were then subjected to selective affinity
droplets are cooled below the glass transition point of the chromatography. As was observed with the solvent-based
asphaltene entity, then these minute droplets solidify in a extraction methods above, the highest removal of vanadyl
fractal-like structure, giving the appearance of colloids that porphyrins achieved with this method was 57%. The lack of
have aggregated to form a large precipitate. This theory was full recovery of the vanadyl porphyrins was attributed
illustrated by precipitating polystyrene in cyclohexane. If the mainly to the aggregation behavior of the asphaltenes in
polystyrene solution was slowly cooled across the glass solution.
transition point, the resulting morphology was that of Chemical extraction methods are, therefore, not likely to
smooth droplets of polystyrene in the cyclohexane solvent. be capable of completely extracting all of the metallopor-
However, if the polystyrene solution was cooled rapidly by phyrins present in a petroleum sample. This chemical limita-
injecting some of it into cold iso-octane, a fractal-like tion is further compounded by the fact that asphaltenes form
structure was formed similar to that observed with asphal- aggregated colloidal structures in most organic solvents,
tenes. The fact that a well-defined model compound (which which could further deter the extraction of the metallopor-
could be operationally defined as an asphaltene since it is phyrins from the asphaltene phase.
toluene-soluble but n-heptane-insoluble) behaved in vir-
tually the same manner as actual asphaltenes provides 5. Implications for Selective Removal of Vanadium
compelling evidence for the liquid-liquid glass transition Components from Vacuum Residues
theory. However, the length scales observed by Sirota were
on the order of micrometers and as such are more in line with The commercial processes for processing of vacuum residue
the precipitation behavior observed in n-heptane and other give distinctly different fates for the vanadium components.
asphaltene precipitants and as such are not necessarily
applicable to nanoaggregates in solvents such as toluene. (147) Lerner, B. J. Removing metals with a 2-pyrrolidone-alcohol
mixture. Patent 3052627, 09/04/1962.
This type of liquid-liquid formalism was used by Yarranton (148) Kimberlin, C. N. J.; Mattox, W. J. Butyrolactone solvent
et al.130-132 along with regular solution theory to model the extraction process for removal of metal contaminants. Patent
precipitation of asphaltenes in n-alkanes. The interactions 2913394, 11/17/1959.
(149) Overfield, R. E. Integrated method for extracting nickel and
between the vanadyl porphyrins and the asphaltenes are vanadium compounds from oils. Patent 4643821, 02/17/1987.
very complex, and a combination of solubility interactions, (150) Galimov, R. A.; Krivonozhkina, L. B.; Abushayeva, V. V.;
Romanov, G. V. Petrol. Chem. 1993, 33, 539–543.
(151) Yin, C.-X.; Stryker, J. M.; Gray, M. R. Energy Fuels 2009, 23,
(146) Sirota, E. B. Energy Fuels 2005, 19, 1290–1296. 2600–2605.
2807
Energy Fuels 2010, 24, 2795–2808 : DOI:10.1021/ef100173j Dechaine and Gray

The most widely used technology is coking, which almost adsorbent can efficiently access the metal centers when they
quantitatively captures the vanadium in the coke byproduct. are in such a cluster. Similarly, photochemical or electroche-
The primary chemical environment of the vanadium is un- mical processing will be ineffective for this fraction of the
changed through this process, although the secondary struc- metal. Small molecules could possibly penetrate such clusters
ture is significantly altered by thermal cracking, dehydration, and react with the vanadyl group, but the resulting products
and addition reactions.6 Similarly, residue fluid catalytic may be no more accessible to separation than the original
cracking would trap vanadium in coke deposits on the catalyst material.151
pellets, which would then be burned when the catalyst is What are the prospects for completely dispersing the as-
regenerated. The resulting vanadium on the FCC catalyst phaltenes to render the metals accessible for catalytic reaction,
destroys the catalyst structure.152,153 Deasphalting gives a adsorption, or derivatization for separation? While increasing
metal-rich and a metal-depleted fraction, but the separation temperature and solvent strength reduce the degree of aggre-
is nonselective in that complete removal is only achieved with gation, these conditions have not been demonstrated to
a very large yield of asphalt.7 Catalytic hydroprocessing completely disaggregate asphaltenes. Given that multiple
selectively removes the vanadium as vanadium sulfide, which interactions between polyfunctional molecules likely account
accumulates in the catalyst and eventually renders it inactive; for the strength of asphaltene aggregation, then a successful
therefore, high metal content feeds require high addition rates strategy must interrupt π-π interactions, hydrogen bonding,
of the catalyst in ebullated bed reactors.9 Only catalytic and polar interactions simultaneously. The Hansen solubility
demetalation can be considered a truly selective removal parameter approach166,167 suggests that a mixed solvent could
technique, but it comes at a substantial cost in catalyst. offer benefits that cannot be achieved by a single solvent
Several different reactive demetalation schemes have been that interrupts only one type of intermolecular interaction.
explored including oxidative demetalation,154-157 biological Yarranton’s observation that the apparent molecular weight
demetalation,158-160 electrolytic demetalation,161,162 ultraso- of asphaltenes can be suppressed by the presence of polar
nic irradiation þ adsorption,163 photochemical reaction þ aromatic compounds110 suggests that a high concentration of
liquid extraction,164 and absorption by Mo complexes.165 such material could effectively disperse the asphaltene ma-
These varied processes yield vanadium removals ranging terial, although the kinetics of such a disaggregation are not
between 20 and 78%, although in many cases the overall mass known. Any scheme that seeks to use a sorbent to remove the
balance is not considered, indicating that a portion of the metals selectively then needs to ensure that the metal species
vanadium removal is not selective and results in significant interact more with the surface than with the other species in
product loss. The inability of the aforementioned schemes to solution.
fully account for the vanadium is linked to the associative
behavior of the asphaltene fraction to which a large portion of 6. Closing Remarks
the vanadium is linked.
On the basis of the preceding discussion, the goal of
The work of Tanaka et al.113 has important implications for
selectively removing the vanadium compounds from vacuum
any effort to remove metals from vacuum residue or asphal-
residue or asphaltene samples would seem to be out of reach.
tene by chemical means. If the aggregate persists at tempera-
Agreement on the actual form of the metal complexes within
tures of over 350 °C, then a portion of the vanadyl compounds
these samples has not even been achieved yet. Nevertheless,
will always be segregated from the free solution. No catalyst or
the incentives for development of selective removal processes
(152) Speronello, B. K.; Reagan, W. J. Oil Gas J. 1984, 82, 139–143. warrant additional research toward this goal. In particular,
(153) Tsiatouras, V. A.; Evmiridis, N. P.; AF Tsiatouras, V. A.; future research should be directed at understanding the
Evmiridis, N. P. Ind. Eng. Chem. Res. 2008, 47, 9288–9296. physical form of the vanadium components (aggregated vs
(154) Kimberlin, C. N. J.; Ellert, H. G.; Adams, C. E.; Hamner, G. P.
Demetallization with Hydrofluoric Acid. Patent 3,203,892, 08/31/1965. free, adsorbed vs chemically bound in the aggregates) and to
(155) Sugihara, J. M.; Branthaver, J. F.; Willcox, K. W. Oxidative also understand under which conditions the vanadium com-
demetallation of oxovanadium(IV) porphyrins. In The Role of Trace pounds are susceptible to removal (i.e., free). This information
Metals in Petroleum; Yen, T. F., Ed.; Ann Arbor Science Publichers Inc.:
Ann Arbor, MI, 1975; pp 183-193. is essential in selecting the conditions for successful separation
(156) Gould, K. A. Fuel 1980, 59, 733–736. of these components. As well, it is clear that any successful
(157) Mann, D. P.; Kukes, S. G.; Coombs, D. M. Metals Removal vanadium separation scheme will need to address the associa-
with a Light Hydrocarbon and an Organophosphorous Compound.
Patent 4,518,484, 05/21/1985. tive behavior of the asphaltenes with which the majority of the
(158) Fedorak, P. M.; Semple, K. M.; Vazquez-Duhalt, R.; Westlake, vanadium is aggregated. Therefore, research efforts directed
D. W. S. Enzyme Microb. Tech. 1993, 15, 429–437. at understanding and eventually disrupting this association
(159) Xu, G.-W.; Mitchell, K. W.; Monticello, D. J. Fuel Product
Produced by Demetalizing a Fossil Fuel with an Enzyme. Patent will be critical to successfully separating the vanadium. This
5,726,056, 03/10/1998. will help point us in the right direction toward the ultimate
(160) Dedeles, G. R.; Abe, A.; Saito, K.; Asano, K.; Saito, K.; goal of selectively removing these troublesome contaminants
Yokota, A.; Tomita, F. J. Biosci. Bioeng. 2000, 90, 515–521.
(161) Greaney, M. A.; Kerby, M. C., Jr.; Olmstead, W. N.; Wiehe, I. from bitumen and crude oils.
A. Method for Demetallating Refinrey Feedstreams. Patent 5,529,684,
06/25/1996. Acknowledgment. The authors acknowledge the financial sup-
(162) Ovalles, C.; Rojas, I.; Acevedo, S.; Escobar, G.; Jorge, G.; port of the Imperial Oil-Alberta Ingenuity Center for Oilsands
Gutierrez, L. B.; Rincon, A.; Scharifker, B. Fuel Process. Technol. 1996, Innovation for funding this project and also the Natural Sciences
48, 159–172. and Engineering Research Council of Canada (NSERC) in the
(163) Sakanishi, K.; Yamashita, N.; Whitehurst, D. D.; Mochida, I.
Catal. Today 1998, 43, 241–247. form of a scholarship for G.P.D.
(164) Shiraishi, Y.; Hirai, T.; Komasawa, I. Ind. Eng. Chem. Res.
2000, 39, 1345–1355. (166) Hansen, C. M. Ind. Eng. Chem. Prod. Res. Dev. 1969, 8, 2–11.
(165) Sakanishi, K.; Saito, I.; Watanabe, I.; Mochida, I. Fuel 2004, 83, (167) Wiehe, I. A. Process Chemistry of Petroleum Macromolecules;
1889–1893. CRC Press: Boca Raton, FL, 2008.

2808

You might also like