Cluster of Asphaltene Nanoaggregates by DC Conductivity and

Jul 22, 2014 - Schlumberger-Doll Research, Cambridge, Massachusetts 02139, United States. ABSTRACT: A model of the dominant molecular and stable ...
1 downloads 0 Views 4MB Size
Article pubs.acs.org/EF

Cluster of Asphaltene Nanoaggregates by DC Conductivity and Centrifugation Lamia Goual,† Mohammad Sedghi,† Farshid Mostowfi,‡ Richard McFarlane,§ Andrew E. Pomerantz,∥ Soheil Saraji,† and Oliver C. Mullins*,∥ †

Department of Chemical & Petroleum Engineering, University of Wyoming, Laramie, Wyoming 82071, United States Schlumberger−DBR Technology Center, Edmonton, Alberta T6N 1M9, Canada § Alberta Innovates - Technology Futures, Edmonton, Alberta T6N 1E4, Canada ∥ Schlumberger-Doll Research, Cambridge, Massachusetts 02139, United States ‡

ABSTRACT: A model of the dominant molecular and stable colloidal structures of asphaltenes has been proposed, the Yen− Mullins model. The formation of clusters of asphaltene nanoaggregates in toluene was reported elsewhere to occur at a concentration of a few grams per liter with a cluster aggregation number of approximately 8 (Mullins, O. C. Energy Fuels 2010, 24, 2179−2207). Here, we measure the DC-conductivity signal of toluene as a function of asphaltene concentration obtaining support for the critical clustering concentration (CCC) of a roughly 1.7 g/L in toluene. In addition, the small change in the Stokes drag at the CCC indicates that the cluster aggregation number is small, less than 10. The temperature variation of the CCC is measured to be small and within error, suggesting that cluster formation is entropically driven. Centrifugation experiments were also performed on asphaltene−toluene solutions at different concentrations. These experiments confirmed that a significant change in asphaltene aggregation occurs in the concentration range of the CCC. The CCC values from centrifugation and DC-conductivity measurements are roughly the same. The centrifugation experiments confirm a cluster size of ∼5 nm corroborating the small aggregation number found in the DC-conductivity experiments. These results add to the growing body of literature validating the Yen−Mullins model.



INTRODUCTION Asphaltenes are the solid, colloidal phase of crude oil.2 Their importance has long been understood within a flow assurance perspective.3 In addition, asphaltene gradients in oilfield reservoirs can be used to address a myriad of reservoir complexities such as reservoir connectivity.4−7 The utilization of asphaltene gradients to understand reservoir complexities is enormously enhanced by use of the industry’s first predictive equation of state of asphaltene gradients, the Flory−Huggins− Zuo EoS (FHZ EoS).8−10 This is the only formalism that has been shown to work for condensates through mobile heavy oils in field studies5−11 although there are other new formulations to treat asphaltene gradients as well.12 The FHZ EoS formalism depends explicitly on the size of the asphaltene molecular or colloidal species present in the corresponding reservoir crude oil. The molecular and stable nanocolloidal species of asphaltenes in crude oils and in laboratory solvents have been codified in the Yen−Mullins model (cf. Figure 1).1,13,14 Asphaltene Molecules. The mean asphaltene molecular weight is measured to be approximately 750 g/mol by many different mass spectral methods15−20 and molecular diffusion methods.21−25 The number of fused rings in the asphaltene polycyclic aromatic hydrocarbon (PAH) is approximately seven as obtained by NMR,26−28 Raman spectroscopy,29 direct molecular imaging,30,31 mass spectrometry18 and optical spectroscopy studies32,33 when combined with the importance of the aromatic sextets of the Clar representation from carbon X-ray Raman spectroscopy.34 Time-resolved fluorescence depolarization studies21,22 indicate that the predominant © XXXX American Chemical Society

Figure 1. Yen−Mullins model of asphaltenes showing the predominant molecular structure, nanoaggregate (with six molecules) and cluster (with eight nanoaggregates).1,13,14 In heavy oils, the clusters dominate.

molecular architecture consists of one or perhaps two PAHs per asphaltene molecule. Laser desorption, laser ionization mass spectrometry (L2MS) studies strongly support the predominance of one PAH per molecule, the so-called island model.35 Similar conclusions were corroborated by other mass spectral studies.36 Moreover, recent L2MS and surface assisted laser desorption ionization studies have shown that (1) laser mass spectrometry is sensitive to all asphaltenes including those that aggregate37 and (2) certain laser mass spectrometry studies have relatively uniform cross sections for asphaltene molecular types and additionally can lead to minimal fragmentation.38 Received: May 11, 2014 Revised: July 22, 2014

A

dx.doi.org/10.1021/ef5010682 | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

Figure 2. Measured asphaltene concentration gradients in a large oilfield (green ellipse in figure) in Saudi Arabia.11 An enormous concentration gradient of asphaltenes is observed in the heavy oil rim of this four-way sealing anticline. Left: the fit of eq 1 to the data in local sections of this giant field is nearly perfect. The one adjustable parameter is the cluster size; fitting this data obtains 5.2 nm cluster size whereas Figure 1 shows a 5.0 nm nominal cluster diameter. Right: the gradient over the 100 km heavy oil rim matches eq 2; fitting to the data (red squares) again gives 5.2 nm cluster size.

asphaltene nanoaggregates do not go to the oil−water interface, only asphaltene molecules load onto the interface.39,40 This is quite logical given the structure of the asphaltene nanoaggregates with their hydrophobic alkyl substituents facing outward (cf. Figure 1). Laser mass spectral studies have recently confirmed that the aggregation number of asphaltene nanoaggregates is about six.50 NMR studies,51 AC-conductivity studies,52 DC-conductivity studies53−55 and centrifugation of asphaltene−toluene solutions55,56 and live crude oil centrifugation57 all gave results consistent with small (100 °C versus room temperature laboratory results supports only small or secondary temperature effects. For a fixed temperature, all factors in Stokes drag (such as viscosity) cancel except for the radius before and after cluster formation. The ratio of slopes is equal to the ratio of radii. The cube then gives the volume difference, thus the aggregation number of the cluster. If one truly only had monomers of nanoaggregates prior to CCC, then the cluster aggregation number would be predicted to be 2, a dimer. However, some nanoaggregate dimerization, trimerization, etc. is expected prior to CCC. We interpret the cube of the ratio to mean the cluster aggregation number is very small; we believe less than 10. To

Figure 9. DC-conductivity vs asphaltene concentration in toluene for LC asphaltene acquired at room temperature. The CCC is seen to be ∼1300 mg/L. The red diamonds were used for linear fitting below the CCC (but above the CNAC) and the black triangles were used for linear fitting above the CCC.

Figure 8 has a room temperature CCC of ∼1.8 g/L. Thus, we find that there is some variability of the CCC, a rough value for virgin crude oil asphaltenes in 10−3 mass fraction, thus 10 times larger than the CNAC. Centrifugation. A robust means of seeing variation of aggregation is through centrifugation. In this context, centrifugation is not a precise measurement for CCC for several reasons. For example, there is variable collection efficiency of asphaltene species in different volume elements in the centrifuge tube. Variations in collection efficiency exist both vertically and laterally in the centrifugation tube. Nevertheless, if aggregation changes substantially, centrifugation is an excellent means to confirm this; there is no equivocation of the results. Figure 10 shows the results of centrifugation of asphaltene solutions of varying concentration. Only a short spin of ∼3 days was used, thereby providing a “cutoff size” of 4 nm for the asphaltene particles. This represents a lower limit. If the particles are growing to 5 nm particle size, then the centrifugation experiments performed here (with longer spin times for smaller particles) would underestimate the CCC. H

dx.doi.org/10.1021/ef5010682 | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

However, Figure 11 shows that there is an enormous increase in the fractional accumulation of HO asphaltenes over the range of concentration including the CCC. This is a robust confirmation that there is a large change in the nature of asphaltene aggregation at the CCC. The accumulated asphaltene fraction increases to nearly 100% meaning that these newly created asphaltene clusters are ≥4 nm per design of the experiment. These experiments, designed with a 4 nm cutoff diameter, are consistent with the finding of 5 nm asphaltene clusters from field studies.11 A different set of centrifugation experiments was performed in order to obtain size limits of asphaltene clusters. A solution of fixed, large concentration of both asphaltenes (3 g/L) was centrifuged for variable lengths of time. In these asphaltene sedimentation experiments, centrifugation was carried out at fixed rotational frequency and temperature, 18000 rpm and 25 °C, respectively. The (x,y) point for each curve (100%,100%) includes any collected asphaltene solid at the bottom of the tube. Figure 12 shows that for 0.47 h spin times, there was no

Figure 10. Distribution of HO asphaltenes (in solution and separated as a solid) in the centrifugation tube for different concentrations of the asphaltene solution (in mg/L). At high concentrations, there is substantial accumulation of asphaltenes at the bottom of the tube.

Figure 10 shows that at high concentrations there is a substantial accumulation of HO asphaltenes toward the base of the centrifuge tube consistent with the formation of clusters. Figure 11 shows the % asphaltene that accumulates toward or

Figure 11. Fractional accumulation of HO asphaltenes at the base of the centrifuge tube versus asphaltene concentration. There is a huge increase over the concentration range of the CCC going to nearly 100% of asphaltenes in clusters.

Figure 12. Cumulative collection of asphaltenes (y-axis) vs the cumulative solution (x-axis) as a function of centrifugation time for a fixed concentration of 3g/L. Short spin times yield no accumulation of asphaltenes toward the bottom (slope = 1), longer spin times yield almost total accumulation at the bottom of the centrifugation tube (slope ≪ 1), providing limits for the size of asphaltene clusters. Results for two different asphaltenes, LC and HO, are shown.

at the base of the centrifuge tube as a function of the solution concentration. The CCC, the concentration where growth in the cluster size ceases, is approximately 1.5 g/L. The lower limit of 4 nm designed for the centrifugation experiments might account for the difference in the comparison between the values of CCC from centrifugation versus DC-conductivity (which is ∼1.8 g/L). At low concentrations, there is some accumulation of HO asphaltenes at the base of the centrifuge tube. Some of this accumulation is due to different sensitivity of collection efficiency of asphaltene particles versus position in the tube. Smaller particles near the far wall of the tube and lower in the tube will have similar collection efficiency as larger particles further from the wall and higher up in the tube. Consequently, centrifugation is not precise in terms of particle sizing.

accumulation of asphaltenes toward the bottom of the centrifuge tube. Table 2 shows that 50 nm particles would travel the length of the centrifuge tube in this time; the asphaltene clusters are much small than 50 nm. This same conclusion was obtained from noting that such large particles would accumulate at the base of crude oil columns in reservoirs, thus would not be present in normal crude oils. Even for the 11.47 h spin, the accumulation is only partial, indicating that the clusters are less than 10 nm. Note that in this spin time, some accumulation of asphaltenes of 5 nm is expected. We have not tried to quantify spin time versus size except to zeroth order (Table 2) due to the variable g-forces in the centrifuge tube and the variable velocities of asphaltenes before and after they hit the sloped outer wall of the tilted centrifuge tube. At 1.96 days spin time, both asphaltenes show a high degree of accumulation at the base of the centrifuge tube. This I

dx.doi.org/10.1021/ef5010682 | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

experiments provide robust determination of a large change in asphaltene aggregation at the CCC. In addition, the centrifugation and DC-conductivity experiments show that the cluster diameter is roughly 5 nm. Here, the temperature dependence of the critical cluster concentration (CCC) was found to be relatively small. This implies that, similar to nanoaggregate formation, cluster formation is predominantly mediated by entropy increase. The magnitude of the entropy change is ∼11 cal/K·mol, which is considerably smaller than that associated with nanoaggregate formation. Any high energy sites in asphaltene molecules would be (noncovalently) bonded in the nanoaggregate, thus only weak bonding remains for cluster formation. Analysis of the ratio of Stokes radii of the nanoaggregate and cluster (below and above the CCC) indicates that the cluster aggregation number is small and similar to many other findings such as in NMR and SANS and SAXS studies. The determinations found here for the number of nanoaggregates in a cluster and for the CCC are similar to those values in several other literature reports. The tight link between asphaltene nanoscience and oilfield reservoirs that are bigger by at least 13 orders of magnitude in linear dimension is a welcome development enabling a much clearer picture of reservoir fluid dynamics in geologic time. Prior to understanding asphaltene nanostructures, understanding these reservoir dynamics was simply precluded. In turn, reservoir studies are now constraining models of asphaltene nanostructures. This new feedback loop links fundamental science with major economic concerns, which is beneficial to all involved.

corresponds to a nominal asphaltene cluster size of 5 nm. A much longer spin did not cause accumulation of all asphaltene. A low concentration of asphaltene remained in the supernatant fluid after long spin times. This is consistent with the phase equilibrium model of asphaltene aggregation, which was shown to apply to high-Q ultrasonics data.44,45 That is, at low concentrations, asphaltene clusters are not expected, only asphaltene nanoaggregates or even molecules at lower concentrations. In other words, at high concentrations, most of the asphaltene population is in clusters, but there remains a small concentration of asphaltene nanoaggregates. The greater sedimentation of the HO asphaltene versus the LC asphaltene indicates that the HO clusters are somewhat larger than the LC clusters. One factor that is unknown is whether these asphaltenes slide down the test tube, tilted at 24°, at the same rate once the asphaltenes hit the far wall. Assuming this and other factors cancel out, the comparison of the data for both asphaltenes for the 11.74 h shows that the HO asphaltenes went roughly twice as far in the x-direction in Figure 12 from the x=y curve. The sedimentation velocity is proportional to the square of the particle size (cf. eq2). Thus, one obtains the estimate that the HO clusters are bigger by a factor of 1.4 than the LC clusters. If the LC clusters are 5 nm, then this gives HO clusters of 7 nm. However, comparison of the DC conductivity does not indicate that the HO clusters are much large than the LC clusters. It may be that more precise methods are needed to resolve subtle differences in cluster size. Current oilfield reservoir studies are incorporating the influence of asphaltene clusters in the dynamics of reservoir fluids.85 It appears that clusters play a critical role in establishing convective gravity currents that give rise to tar mat formation. These tar mats are often regional and can preclude aquifer support of production, leading to expensive alterations of field development planning.85 In other cases, where asphaltene cluster diffusion is overwhelmed by reservoir fluid processes that cause rapid asphaltene destabilization, upstructure mobile and permeable bitumen deposition has been encountered.85 The link between asphaltene nanoscience and large scale oil reservoir processes that occur in geologic time is enabling a first-principles approach to assessing oil production risks; this is beneficial for asset managers and oilfield scientists alike.



AUTHOR INFORMATION

Corresponding Author

*O. C. Mullins. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Mullins, O. C. The Modified Yen Model. Energy Fuels 2010, 24, 2179−2207. (2) Asphaltenes, Heavy Oils and Petroleomics; Mullins, O. C., Sheu, E. Y., Hammami, A., Marshall, A. G., Eds.; Springer: New York, 2007. (3) Hammami, A.; Ratulowski, J. Precipitation and Deposition of Asphaltenes in Production Systems: A Flow Assurance Overview. In Asphaltenes, Heavy Oils and Petroleomics; Mullins, O. C., Sheu, E. Y., Hammami, A., Marshall, A. G., Eds.; Springer: New York, 2007; Chapter 23. (4) Mullins, O. C. The Physics of Reservoir Fluids; Discovery through Downhole Fluid Analysis; Schlumberger Press: Houston, TX, 2008. (5) Mullins, O. C.; Betancourt, S. S.; Cribbs, M. E.; Creek, J. L.; Andrews, A. B.; Dubost, F.; Venkataramanan, L. The Colloidal Structure of Crude Oil and the Structure of Reservoirs. Energy Fuels 2007, 21, 2785−2794. (6) Dong, C.; Petro, D.; Pomerantz, A. E.; Nelson, R. L.; Latifzai, A. S.; Nouvelle, X.; Zuo, J. Y.; Reddy, C. M.; Mullins, O. C. New Thermodynamic Modeling of Reservoir Crude Oil. Fuel 2014, 117, 839−850. (7) Mullins, O. C.; Pomerantz, A. E.; Zuo, J. Y.; Dong, C. Downhole Fluid Analysis and Asphaltene Science for Petroleum Reservoir Evaluation. Annu. Rev. Chem. Biomol. Eng. 2014, 5, 325−345. (8) Freed, D. E.; Mullins, O. C.; Zuo, J. Y. Asphaltene Gradients in the Presence of GOR Gradients. Energy Fuels 2010, 24, 3942−3949. (9) Zuo, J. Y.; Mullins, O. C.; Freed, D. E.; Dong, C.; Elshahawi, H.; Seifert, D. Advances of the Flory-Huggins-Zuo Equation of State for Asphaltene Gradients and Formation Evaluation. Energy Fuels 2013, 27, 1722−1735.



CONCLUSIONS Stable asphaltene clusters of nanoaggregates are an important component of the asphaltene nanostructure, as indicated in the Yen−Mullins model. Clusters play a crucial role in creating massive asphaltene and viscosity gradients in oilfield reservoirs. Prior to understanding the nature of asphaltene clusters, there was great uncertainty regarding the origin of heavy oil gradients in reservoirs. Now with the understanding of clusters and use of the Flory−Huggins−Zuo equation of state, large gradients of heavy oil are routinely interpreted to be either in thermodynamic equilibrium or in a state of disequilibrium due to oftenidentified reservoir fluid processes. In addition, asphaltene clusters are evidently important in reservoir processes that can result in formation of sealing tar mats that can have enormous impact on the economics of oilfields. It is essential to expand understanding of asphaltene clusters. The CCC for n-heptane asphaltenes in toluene determined by DC conductivity is at approximately 1.7 g/L, with comparable results found with centrifugation. The CCC in crude oils can be quite different and is often at a much higher concentration. The centrifugation J

dx.doi.org/10.1021/ef5010682 | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

(10) Freed, D. E.; Mullins, O. C.; Zuo, J. Y. Heuristics for Equilibrium Distributions of Asphaltenes in the Presence of GOR Gradients. Energy Fuels 2014, in press. (11) Mullins, O. C.; Zuo, J. Y.; Seifert, D.; Zeybek, M. Clusters of Asphaltene Nanoaggregates Observed in Oil Reservoirs. Energy Fuels 2013, 27, 1752−1761. (12) Panuganti, S. R.; Vargas, F. M.; Chapman, W. G. Modeling of Reservoir Connectivity and Tar-Mat Using Gravity-Induced Asphaltene Compositional Grading. Energy Fuels 2012, 26, 2548−2557. (13) Mullins, O. C.; Sabbah, H.; Eyssautier, J.; Pomerantz, A. E.; Barré, L.; Andrews, A. B.; Ruiz-Morales, Y.; Mostowfi, F.; McFarlane, R.; Goual, L.; Lepkowicz, R.; Cooper, T.; Orbulescu, J.; Leblanc, J. M.; Edwards, J.; Zare, R. N. Advances in Asphaltene Science and the YenMullins Model. Energy Fuels 2012, 26, 3986−4003. (14) Mullins, O. C. The Asphaltenes. Annu. Rev. Anal. Chem. 2011, 4, 393−418. (15) Hortal, A. R.; Hurtado, P.; Martinez-Haya, B.; Mullins, O. C. Molecular Weight Distributions of Coal and Crude Oil Asphaltenes from Laser Desorption Ionization Experiments. Energy Fuels 2007, 21, 2863−2868. (16) Pomerantz, A. E.; Hammond, M. R.; Morrow, A. L.; Mullins, O. C.; Zare, R. N. Two Step Laser Mass Spectrometry of Asphaltenes. J. Am. Chem. Soc. 2008, 130, 7216−7217. (17) Pomerantz, A. E.; Hammond, M. R.; Morrow, A. L.; Mullins, O. C.; Zare, R. N. Asphaltene Molecular Weight Distribution Determined by Two-Step Laser Mass Spectrometry. Energy Fuels 2009, 23, 1162− 1168. (18) Hurt, M. R.; Borton, D. J.; Heewon Choi, J.; Kenttämaa, H. I. Comparison of the Structures of Molecules in Coal and Petroleum Asphaltenes by Using Mass Spectrometry. Energy Fuels 2013, 27, 3653−3658. (19) Hortal, A. R.; Martınez-Haya, B.; Lobato, M. D.; Pedrosa, J. M.; Lago, S. J. On the Determination of Molecular Weight Distributions of Asphaltenes and Their Aggregates in Laser Desorption Ionization Experiments. Mass Spectrom. 2006, 41, 960. (20) Qian, K.; Edwards, K. E.; Siskin, M.; Olmstead, W. N.; Mennito, A. S.; Dechert, G. J.; Hoosain, N. E. Desorption and Ionization of Heavy Petroleum Molecules and Measurement of Molecular Weight Distributions. Energy Fuels 2007, 21, 1042−1047. (21) Groenzin, H.; Mullins, O. C. Asphaltene Molecular Size and Structure. J. Phys. Chem. A 1999, 103, 11237−11245. (22) Groenzin, H.; Mullins, O. C. Molecular Sizes of Asphaltenes from Different Origin. Energy Fuels 2000, 14, 677. (23) Andrews, A. B.; Guerra, R.; Mullins, O. C.; Sen, P. N. Diffusivity of Asphaltene Molecules by Fluorescence Correlation Spectroscopy. J. Phys. Chem. A 2006, 110, 8095. (24) Andrews, A. B.; Shih, W.-C.; Mullins, O. C.; Norinaga, K. Molecular Size of Various Asphaltenes by Fluorescence Correlation Spectroscopy. Appl. Spectrosc. 2011, 65, 1348−1356. (25) Freed, D. E.; Lisitza, N. V.; Sen, P. N.; Song, Y. Q. Asphaltene Molecular Composition and Dynamics from Diffusion Measurements. In Asphaltenes, Heavy Oils and Petroleomics; Mullins, O. C., Sheu, E. Y., Hammami, A., Marshall, A. G., Eds.; Springer: New York, 2007; pp 279−296. (26) Andrews, A. B.; Edwards, J. C.; Mullins, O. C.; Pomerantz, A. E. A Comparison of Coal and Petroleum Asphaltenes by 13C Nuclear Magnetic Resonance and DEPT. Energy Fuels 2011, 25, 3068−3076. (27) Majumdar, R. D.; Gerken, M.; Mikula, R.; Hazendonk, P. Validation of the Yen−Mullins Model of Athabasca Oil-Sands Asphaltenes using Solution-State 1H NMR Relaxation and 2D HSQC Spectroscopy. Energy Fuels 2013, 27 (11), 6528−6537. (28) Fergoug, T.; Bouhadda, Y. Determination of Hassi Messaoud Asphaltene Aromatic Structure from 1H & 13C NMR Analysis. Fuel 2014, 115, 521−526. (29) Bouhadda, Y.; Bormann, D.; Sheu, E. Y.; Bendedouch, D.; Krallafa, A.; Daaou, M. Characterization of Algerian Hassi-Messaoud Asphaltene Structure Using Raman Spectrometry and X-ray Diffraction. Fuel 2007, 86, 1855−1864.

(30) Zajac, G. W.; Sethi, N. K.; Joseph, J. T. Molecular imaging of petroleum asphaltenes by scanning tunneling microscopy: verification of structure from 13C and proton NMR data. Scanning Microsc. 1994, 8, 463−470. (31) Sharma, A.; Groenzin, H.; Tomita, A.; Mullins, O. C. Probing Order in Asphaltenes and Aromatic Ring Systems by HRTEM. Energy Fuels 2002, 16, 490−96. (32) Ruiz-Morales, Y.; Mullins, O. C. Polycyclic Aromatic Hyodrocarbons of Asphaltenes Analyzed by Molecular Orbital Calculations with Optical Spectroscopy. Energy Fuels 2007, 21, 256− 265. (33) Ruiz-Morales, Y.; Wu, X.; Mullins, O. C. Electronic Absorption Edge of Crude Oils and Asphaltenes Analyzed by Molecular Orbital Calculations with Optical Spectroscopy. Energy Fuels 2007, 21, 944. (34) Bergmann, U.; Mullins, O. C.; Cramer, S. P. X-ray Raman Spectroscopy of Carbon in Asphaltene: Light Element Characterization with Bulk Sensitivity. Anal. Chem. 2000, 72, 2609−2612. (35) Sabbah, H.; Morrow, A. L.; Pomerantz, A. E.; Zare, R. N. Evidence for Island Structures as the Dominant Architecture of Asphaltenes. Energy Fuels 2011, 25, 1597−1604. (36) Borton, D., II; Pinkston, D. S.; Hurt, M. R.; Tan, X.; Azyat, K.; Scherer, A.; Tykwinski, R.; Gray, M.; Qian, K.; Kenttama, H. I. Molecular Structures of Asphaltenes Based on the Dissociation Reactions of Their Ions in Mass Spectrometry. Energy Fuels 2010, 24, 5548−5559. (37) Pomerantz, A. E., et al. Asphaltene Molecular Weight and Architecture and Nanoaggregate Weight by Laser Mass Spectrometry, Manuscript in preparation. (38) Wu, Q.; Pomerantz, A. E.; Mullins, O. C.; Zare, R. N. Fragmentation and Aggregation in Laser Desorption Laser Ionization and Surface Assisted Laser Desorption Ionization Mass Spectrometry. J. Am. Soc. Mass Spec. 2013, 24, 1116−1122. (39) Rane, J. P.; Harbottle, D.; Pauchard, V.; Couzis, A.; Banerjee, S. Adsorption kinetics of asphaltenes at the oil−water interface and nanoaggregation in the bulk. Langmuir 2012, 28, 9986−9995. (40) Rane, J. P.; Pauchard, V.; Couzis, A.; Banerjee, S. Interfacial rheology of asphaltenes at oil-water interfaces and interpretation of the equation of state. Langmuir 2013, 29, 4750−4759. (41) Andrews, A. B.; McClelland, A.; Korkeila, O.; Krummel, A.; Mullins, O. C.; Demidov, A.; Chen, Z. Sum frequency generation studies of Langmuir films of complex surfactants and asphaltenes. Langmuir 2011, 27, 6049−6058. (42) Karimi, A.; Qian, K.; Olmstead, W. N.; Freund, H.; Yung, C.; Gray, M. R. Quantitative Evidence for Bridged Structures in Asphaltenes by Thin Film Pyrolysis. Energy Fuels 2011, 25, 3581− 3589. (43) Alshareef, A. H.; Scherer, A.; Tan, X.; Azyat, K.; M. Stryker, J. M.; Tykwinski, R. R.; Gray, M. R. Formation of Archipelago Structures during Thermal Cracking Implicates a Chemical Mechanism for the Formation of Petroleum Asphaltenes. Energy Fuels 2011, 25, 2130− 2136. (44) Andreatta, G.; Bostrom, N.; Mullins, O. C. High-Q Ultrasonic Determination of the Critical Nanoaggregate Concentration of Asphaltenes and the Critical Micelle Concentration of Standard Surfactants. Langmuir 2005, 21, 2728. (45) Andreatta, G.; Goncalves, C. C.; Buffin, G.; Bostrom, N.; Quintella, C. M.; Arteaga-Larios, F.; Perez, E.; Mullins, O. C. Nanoaggregates and Structure-Function Relations in Asphaltenes. Energy Fuels 2005, 19, 1282−1289. (46) Eyssautier, J.; Levitz, P.; Espinat, D.; Jestin, J.; Gummel, J.; Grillo, I.; Barre, L. Insight into Asphaltene Nanoaggregate Structure Inferred by Small Angle Neutron and X-ray Scattering. J. Phys. Chem. B 2011, 115, 6827−6837. (47) Eyssautier, J.; Henaut, I.; Levitz, P.; Espinat, D.; Barre, L. Organization of Asphaltenes in a Vacuum Residue: A Small-Angle Xray Scattering (SAXS)−Viscosity Approach at High Temperatures. Energy Fuels 2012, 26, 2696−2704. (48) Eyssautier, J.; Espinat, D.; Gummel, J.; Levitz, P.; Becerra, M.; Shaw, J.; Barre, L. Mesoscale Organization in a Physically Separated K

dx.doi.org/10.1021/ef5010682 | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

Vacuum Residue: Comparison to Asphaltenes in a Simple Solvent. Energy Fuels 2012, 26, 2680−2687. (49) Sedghi, M.; Goual, L.; Welch, W.; Kubelka, J. Molecular Dynamics of Asphaltene Aggregation in Organic Solvents. J. Phys. Chem. B 2013, 117, 5765−5776. (50) Wu, Q.; Pomerantz, A. E.; Mullins, O. C.; Zare, R. N. Laserbased Mass Spectrometric Determination of Aggregation Numbers for Petroleum- and Coal-Derived Asphaltenes. Energy Fuels 2014, 28, 475−482. (51) Freed, D. E.; Lisitza, N. V.; Sen, P. N.; Song, Y. Q. A Study of Asphaltene Nanoaggregation by NMR. Energy Fuels 2009, 23, 1189− 93. (52) Sheu, E. Y.; Long, Y.; Hamza, H. Asphaltene Self-Association and Precipitation in Solvents; AC Conductivity Measurements. In Asphaltenes, Heavy Oils and Petroleomics; Mullins, O. C., Sheu, E. Y., Hammami, A., Marshall, A. G., Eds.; Springer: New York, 2007; Chapter 10. (53) Zeng, H.; Song, Y. Q.; Johnson, D. L.; Mullins, O. C. Critical Nanoaggregate Concentration of Asphaltenes by Low Frequency Conductivity. Energy Fuels 2009, 23, 1201−1208. (54) Sedghi, M.; Goual, L. Role of Resins on Asphaltene Stability. Energy Fuels 2010, 24, 2275−2280. (55) Goual, L.; Sedghi, M.; Zeng, H.; Mostowfi, F.; McFarlane, R.; Mullins, O. C. On the Formation and Properties of Asphaltene Nanoaggregates and Cluster by DC-Conductivity and Centrifugation. Fuel 2011, 90, 2480−2490. (56) Mostowfi, F.; Indo, K.; Mullins, O. C.; McFarlane, R. Asphaltene Nanoaggregates and the Critical Nanoaggregate Concentration from Centrifugation. Energy Fuels 2009, 23, 1194−1200. (57) Indo, K.; Ratulowski, J.; Dindoruk, B.; Gao, J.; Zuo, J. Y.; Mullins, O. C. Asphaltene Nanoaggregates Measured in a Live Crude Oil by Centrifugation. Energy Fuels 2009, 23, 4460−4469. (58) Betancourt, S. S.; Ventura, G. T.; Pomerantz, A. E.; Viloria, O.; Dubost, F. X.; Zuo, J. Y.; Monson, G.; Bustamante, D.; Purcell, J. M.; Nelson, R. K.; Rodgers, R. P.; Reddy, C. M.; Marshall, A. G.; Mullins, O. C. Nanoaggregates of Asphaltenes in a Reservoir Crude Oil. Energy Fuels 2009, 23, 1178−1188. (59) McKenna, A. M.; Donald, L. J.; Fitzsimmons, J. E.; Juyal, P.; Spicer, V.; Standing, K. G.; Marshall, A. G.; Rodgers, R. P. Heavy Petroleum Composition. 3. Asphaltene Aggregation. Energy Fuels 2013, 27 (3), 1246−1256. (60) Mullins, O. C.; Martinez-Haya, B.; Marshall, A. G. Contrasting Perspective on Asphaltene Molecular Weight. This Comment vs the Overview of A. A. Herod, K. D. Bartle, and R. Kandiyoti. Energy Fuels 2008, 22, 1765−1773. (61) Friberg, S. E. Micellization. In Asphaltenes, Heavy Oils and Petroleomics; Mullins, O. C., Sheu, E. Y., Hammami, A., Marshall, A. G., Eds.; Springer: New York, 2007; Chapter 7. (62) Goncalves, S.; Castillo, J.; Fernandez, A.; Hung, J. Absorbance and Fluorescence Spectroscopy on the Aggregation of Asphaltene− toluene Solutions. Fuel 2004, 83, 1823−1828. (63) Kurup, A.; Valori, A.; Bachman, H. N.; Korb, J. P.; Hurlimann, M.; Zielinski, L. Frequency Dependent Magnetic Resonance Response of Heavy Crude Oils: Methods and Applications, SPE 168070; SPE Saudi Arabia Section Technical Symposium and Exhibition, Khobar, Saudi Arabia, May 19−22, 2013. (64) Korb, J. P.; Louis-Joseph, A.; Benamsili, L. Probing Structure and Dynamics of Bulk and Confined Crude Oils by Multiscale NMR Spectroscopy, Diffusometry, and Relaxometry. J. Phys. Chem. B 2013, 2013, 7002−7014. (65) Jain, S.; Valeriy, V.; Ginzburg, V. V.; Jog, P.; Weinhold, J.; Srivastava, R.; Chapman, W. G. Modeling Polymer-Induced Interactions between Two Grafted Surfaces: Comparison between Interfacial Statistical Associating Fluid Theory and Self-Consistent Field Theory. J. Chem. Phys. 2009, 131, 044908. (66) George, G. N.; Gorbaty, M. L. Sulfur K-edge X-ray Absorption Spectroscopy of Petroleum Asphaltenes and Model Compounds. J. Am. Chem. Soc. 1989, 111, 3182−3186.

(67) Waldo, G. S.; Mullins, O. C.; Penner-Hahn, J. E.; Cramer, S. P. Determination of the Chemical Environment of Sulfur in Petroleum Asphaltenes by X-ray Absorption Spectroscopy. Fuel 1992, 71, 53−57. (68) Mitra-Kirtley, S.; Mullins, O. C.; Ralston, C. Y.; Sellis, D.; Pareis, C. Sulfur Speciation by XANES in Maltenes Resins and Asphaltenes and Whole Crude Oils. Appl. Spectrosc. 1998, 1998, 1522. (69) Mitra-Kirtley, S.; Mullins, O. C.; Chen, J.; Van Elp, J.; George, S. J.; Cramer, S. P. Determination of the Nitrogen Chemical Structures in Petroleum Asphaltenes Using XANES Spectroscopy. J. Am. Chem. Soc. 1993, 115, 252−258. (70) Mitra-Kirtley, S.; Mullins, O. C.; Van Elp, J.; Cramer, S. P. Nitrogen Chemical Structure in Coal and Asphaltenes. Fuel 1993, 72, 133. (71) Pomerantz, A. E.; Bake, K. D.; Craddock, P. R.; Qureshi, A.; Zeybek, M.; Mullins, O. C.; Kodalen, B. G.; Mitra-Kirtley, S.; Bolin, T. B.; Seifert, D. J. Sulfur Speciation in Asphaltenes from a Highly Compositionally Graded Oil Column. Energy Fuels 2013, 27, 4604− 4608. (72) Wu, Q.; Seifert, D. J.; Pomerantz, A. E.; Mullins, O. C.; Zare, R. N. Constant Asphaltene Molecular and Nanoaggregate Mass in a Gravitationally Segregated Reservoir. Energy Fuels 2014, in press. (73) Anisimov, M. A.; Yudin, I. K.; Nikitin, V.; Nikolaenko, G.; Chernoutsan, A.; Toulhoat, H.; Frot, D.; Briolant, Y. Asphaltene Aggregation in Hydrocarbon Solutions Studied by Photon Correlation Spectroscopy. J. Phys. Chem. 1995, 99, 9576−9580. (74) Yudin, I. K.; Anisimov, M. A. Dynamic light scattering monitoring of asphaltene aggregation in crude oils and hydrocarbon solutions. In Asphaltenes, Heavy Oils and Petroleomics; Mullins, O. C., Sheu, E. Y., Hammami, A., Marshall, A. G., Eds.; Springer: New York, 2007; Ch. 17 in Ref 1. (75) Maqbool, T.; Balgoa, A. T.; Fogler, H. S. Revisiting Asphaltene Precipitation from Crude Oils: A Case of Neglected Kinetic Effects. Energy Fuels 2009, 23, 3681−3686. (76) Maqbool, T.; Raha, S.; Hoepfner, M. P.; Fogler, H. S. Modeling the Aggregation of Asphaltene Nanoaggregates in Crude Oil− Precipitant Systems. Energy Fuels 2011, 25, 1585−1596. (77) Hoepfner, M. P.; Limsakoune, V.; Chuenmeechao, V.; Maqbool, T.; Fogler, H. S. A Fundamental Study of Asphaltene Deposition. Energy Fuels 2013, 27, 725−735. (78) Hoepfner, M. P.; Fogler, H. S. Multiscale Scattering Investigations of Asphaltene Cluster Breakup, Nanoaggregate Dissociation, and Molecular Ordering. Langmuir 2013, 29, 15423− 15432. (79) Goual, L.; Sedghi, M.; Wang, X.; Zhu, Z. Asphaltene Aggregation and Impact of Alkylphenols. Langmuir 2014, 30, 5394− 403. (80) Goual, L. Impedance Spectroscopy of Petroleum Fluids at Low Frequency. Energy Fuels 2009, 23, 2090−4. (81) Joshi, N. B.; Mullins, O. C.; Jamaluddin, A.; Creek, J. L; McFadden, J. Asphaltene Precipitation from Live Crude Oils. Energy Fuels 2001, 15, 979. (82) Santos, F. J. V.; de Castro, C. A. N.; Dymond, J. H.; Dalaouti, N. K.; Assael, M. J.; Nagashima, A. Standard Reference Data for the Viscosity of Toluene. J. Phys. Chem. Ref. Data 2006, 35, 1−8. (83) Professor Walter Chapman, Rice University, private communication. (84) Rane, J., et al., private communication. (85) Mullins, O. C.; Zuo, J. Y.; Wang, K.; Hammond, P. S.; De Santo, I.; Dumont, H.; Mishra, V. K.; Chen, L.; Pomerantz, A. E.; Dong, C.; Elshahawi, H.; Seifert D. J. The Dynamics of Reservoir Fluids and Their Systematic Variations. 55th Annual SPWLA Symposium, Abu Dhabi, UAE, May 18−22, 2014.

L

dx.doi.org/10.1021/ef5010682 | Energy Fuels XXXX, XXX, XXX−XXX