You are on page 1of 10

Fuel 311 (2022) 122559

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Influence of asphaltene structural parameters on solubility


Ronald Nguele a, *, Albert Brandt Mbouopda Poupi b, Ghislain A. Mouthe Anombogo c,
Olalekan S. Alade d, Hakim Saibi b, *
a
Resource Production & Safety Engineering Laboratory, Kyushu University, Fukuoka 819- 0395 Japan
b
Department of Geosciences, College of Science, United Arab Emirates University, Al-Ain, 15551 Al Ain, United Arab Emirates
c
National Advanced School of Engineering of Maroua, University of Maroua, 46 Maroua, Cameroon
d
Center for Integrative Petroleum Research, College of Petroleum and Geosciences, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

A R T I C L E I N F O A B S T R A C T

Keywords: This paper investigates the influence of asphaltene’s structural parameters on its solubility, including aromatic
Asphaltenes sheet diameter and aromaticity. Aromatic sheet diameter (La) was determined by integrating the D1 and G bands
Structural parameters of Raman spectra. Aromaticity was determined by the ratio of the integrated Fourier Infrared peaks of aliphatic
Asphaltene solubility
carbon relative to aromatic carbon. Asphaltenes were extracted from five dead crude oils. Two types of solubility
Asphaltene precipitation
tests were conducted, with asphaltenes dissolved in both nonpolar (aromatic hydrocarbons) and polar (cyclic
ethers and aromatic amines) solvents and re-precipitated using pentane; in addition to a scenario where
asphaltenes were dissolved in toluene and re-precipitated using nonpolar (alkanes) and polar titrants (esters).
The largest La was estimated to be 7.34 nm, while the other asphaltenes have a La within the range of 1.52–2.35
nm. La increases monotonically with aromaticity, which was inversely proportional to the hydrogen bonding
capability. In terms of dispersant polarity, π-π interactions dominated, evidenced by a better asphaltene stability,
which increased with aromatic core size. A reverse trend was recorded with nonpolar precipitants due to
hydrogen bonding. This study argues that to inhibit the aggregation of asphaltene with large La, polar inhibitors
should be used.

1. Introduction either by the island or archipelago model [5]. The former model de­
scribes asphaltene molecule as a central, single aromatic with unpre­
The elemental chemistry of a typical crude oil consists of a mixture of dictable alkyl chains. The archipelago approach models asphaltene as a
carbon (C, up to 87%) and hydrogen (H, up to 14%), potentially in two fused aromatic and heteroaromatic rings, which interconnect by
addition to oxygen (O), nitrogen (N), sulfur (S) and even metals (M, up alkyl bridges [6]. Fig. 1 shows a typical molecular structure of
to 2%) [1]. These elements are found within the crude oil in the form of asphaltene.
paraffins (CnH2n+2), naphthalenes (CnH2n), and aromatics (C4r+2H2r+4), Because of the complex chemistry and nature of asphaltenes, a
where n and r represent the number of carbons and rings, respectively simpler definition describes asphaltenes as the fraction of crude oil
[2]. Their concentration within the crude oil dictates its nomenclature; soluble in aromatic solvents (e.g., benzene) and insoluble in normal
for example, a crude oil rich in straight-chain or branched paraffin alkanes (e.g., pentane) [8]. The presence of asphaltenes in crude oil is
would be described as a paraffinic crude oil [3]; likewise, an aromatic often detrimental to petroleum operations as their precipitation may
crude oil is predominantly rich in aromatic compounds. Alternatively, if reduce oil-bearing matrix permeability [9,10], cause low oil production
the oil’s composition is dominated by asphaltenes, it is termed an [6], or lead to pipeline clogging[11].
asphaltic crude. The aforementioned issues are due to an imbalance between the
Asphaltenes are dark brown to black friable solids with high mo­ favorable components of the oils (i.e., asphaltene and resins) and the
lecular weight that have no melting point and are composed of unfavorable components (i.e. asphaltenes and saturates) [12,13]. The
condensed aromatic components with variable amounts of heteroatoms mechanisms of asphaltene precipitation, which is a step-wise process,
[4]. From a structural point of view, asphaltenes could be described are triggered by the destabilization of asphaltene monomers in

* Corresponding authors.
E-mail addresses: nguele@mine.kyushu-u.ac.jp (R. Nguele), hakim.saibi@uaeu.ac.ae (H. Saibi).

https://doi.org/10.1016/j.fuel.2021.122559
Received 21 August 2021; Received in revised form 26 October 2021; Accepted 8 November 2021
Available online 14 November 2021
0016-2361/© 2021 Elsevier Ltd. All rights reserved.
R. Nguele et al. Fuel 311 (2022) 122559

suspension in the crude oil. heptane-precipitated asphaltene by computing the ratio of the inte­
The precipitation commences when the asphaltene monomers stack grated peaks of aromatic C– – C at 1600 cm− 1 to those of aliphatic C–H
up to yield nanoaggregates, whose size varies from 2 to 5 nm due to the appearing at a wavelength of 1450 cm− 1 [16]. Rogel et al. (2001, 2015)
co-facial stacking induced by π- π bonding between the poly-condensed showed that the concentrations of functional groups can be estimated
cores [5,14,15]. Because of the propensity of asphaltene to self- from FTIR, assuming the extinction coefficient remains unchanged
aggregate, the nanoaggregates cluster further to larger sizes (5–12 [26,27]. In addition, x-ray spectroscopy (XPS) analyses the chemical
nm). The alteration of physico-chemical conditions of the environment composition of asphaltene, which is expressed as a function of the
in which asphaltenes are dispersed (e.g. injecting gas in the solution) chemical states of the element. From the available literature, petroleum
would prompt the formation of large flocs (size > 500 nm), which lead and coal-derived asphaltenes consist mainly of C1s, N1s, O1s, and S2p3/2
invariably to precipitation [16]. [6]; however, differentiating aromatic carbon from aliphatic carbon is of
Wiehe et al. (2012) suggested that the asphaltene precipitation oc­ considerable importance, and forms an essential part of the analytical
curs due to a decrease in asphaltene solubility, which is related to the process [28].
kinetics of asphaltene in a dispersed medium [17]. Consistent with the To help understand the contributions made by the different struc­
previous studies, Adams (2014) suggested that the precipitation tural parameters of asphaltene to its solubility, in this study, we focus on
occurred when there is an imbalance of the cooperative forces keeping the influence of aromatic sheet diameter and asphaltene aromaticity. We
the asphaltene monomers in suspension, including hydrogen bonding were particularly interested in understanding the behavior of asphaltene
(H-bonding), Brønsted acid-base interactions, metal coordination com­ with large aromatic sheet diameter (i.e., aromatic sheet diameter>2.1
plexes, aromatic π- π stacking, and hydrophobic pockets [18]. In sum­ nm) in different dispersants and titrants. For this purpose, five asphal­
mary, previous literature indicates that both intermolecular (e.g. H- tenes were precipitated from dead crude oils sampled across different
bonding) and intramolecular forces (e.g. π- π stacking) contribute to geological formations including Central and West Africa (one sample
asphaltene precipitation and should be accounted for when considering from each geographical location), the Middle East, Asia, and North
the thermodynamic conditions of the system [19]. America. Subsequently, their respective structural properties were
Bestougeff and Byramjee (1994) showed that asphaltene solubility estimated using Raman and FTIR spectroscopy, and XPS spectrometry
could be partially understood if the polymeric nature of the asphaltene is was used to determine the elemental composition of the asphaltenes.
well-constrained [20]. Daaou et al. (2009) reported that, based on ob­ Asphaltene solubility was also evaluated in both polar and nonpolar
servations from the Hassi Messaould Oilfield (Algeria), asphaltene pre­ media using gravimetric methods.
cipitation was proportional to the size of the poly-condensed aromatic
cores of the asphaltenes [21]. Riedeman et al. (2016) showed from the 2. Materials and methods
structural analyses of asphaltene that the larger the aromatic diameter
sheet of the asphaltenes, the higher their likelihood of precipitating 2.1. Materials
[22]. Their findings are consistent with the conclusions of Ibrahim et al.
(2004), who showed that precipitation likelihood is higher when a 2.1.1. Asphaltene samples
longer alkyl chain is attached to the poly-condensed aromatic core [23]. The investigated asphaltenes were extracted from five different
Thus, evaluating the contribution of the structural parameters of the crude oils sampled from the surface facilities in North America, Central
asphaltene is important to constrain the mechanisms of asphaltene and West Africa, Northern Japan, and the Middle East. The properties of
precipitation and, moreover, propose a potential mitigation approach. the parent oils are summarized in Table 1.
Typically, structural parameters are determined using spectral methods
[6,24]. In this work, we applied Raman, Infrared, and X-ray spectros­ 2.1.2. Chemicals
copy. Raman spectroscopy exploits the carbon structures of asphaltenes Table 2 outlines the type, class, name, chemical structure, supplier,
which are similar to those of microcrystalline graphite-like substances. and hydrogen-bonding capability of all solvents used in the present
Quantitatively, the position and integrated intensity of the D1 and G study.
bands allow the aromatic sheet diameter to be determined [22,25].
Fourier infrared (FTIR) is a complementary tool to Raman spectroscopy,
which uses the absorption of the functional groups present in the 2.2. Methods
asphaltenes to determine the aromaticity of the asphaltene.
Alhreez and Dongsheng (2019) determined the aromaticity of Scheme 1 shows the experimental flowchart followed in this work.
The experiments were broken down into two major groups starting

Fig. 1. A typical molecular structure of petroleum asphaltenes modified from Yen-Mullin model proposed by Martín-Martínez et al. [7]

2
R. Nguele et al. Fuel 311 (2022) 122559

Table 1 with the extraction and the characterization of the asphaltenes from the
a
Properties of dead oil samples used to extract asphaltenes. parent oils followed by a series of solubility tests. Note that all the ex­
Code A B C D E periments were conducted at room temperature and that the solvents
were used as received.
Origin Middle Central Asia North West
East Africa America Africa
Type of crude oil Light Heavy Heavy Extra- Bitumen 2.2.1. Asphaltene sources
heavy A modified standard procedure (ASTM D3279) was followed to
API gravity (o) 33.7 17.37 16.6 7.9 7.703 extract asphaltenes from their respective parent oils [31]. One gram of
Density (g/cm3) 0.853 0.947 0.955 1.014 1.013
Total acid number 0.774 8.536 3.187 1.853 6.587
crude oil was accurately weighed and transferred to a 100 ml Erlen­
(mg-KOH/g-oil) meyer flask. The crude oil was then diluted with toluene at a volume
Total base number 0.874 0.818 1.182 0.96 0.435 ratio of 1:10 oil:toluene. The mixture was stirred at 1500 rpm using a
(mg-HCl/g-oil) magnetic stirrer (D Lab, MS H280-pro) for 2 hr and the resulting solution
a
Both total acid number and total base number were measured following was then filtered through a 0.45 µm filter paper. One volume of the
standard procedures [29] filtrate was added to 20 volumes of heptane, followed by storage

Table 2
Dispersants and titrants used in this study.
Solvent type Class Name Formula Supplier Hydrogen-bonding capabilitya

Dispersants Aromatic hydrocarbon Benzene C6H6 Junsei Poor


Toluene C6H5CH3 Junsei Poor
Xylene C6H4 (CH3)2 Junsei Poor
Ethylbenzene C6H5C2H5 Junsei Poor
Cyclic ethers Tetrahydrofuran (CH2)4O Junsei Moderate
Furan (CH)4O Sigma Aldrich Moderate
Furfural (CH)4OCO Sigma Aldrich Moderate
Aromatic amine Aniline C6H5-NH2 Sigma Aldrich Strong
Titrants Alkanes Pentane C5H12 Junsei Poor
Hexane C6H14 Junsei Poor
Heptane C7H16 Junsei Poor
Decane C10H22 Junsei Poor
Esters Methyl Acetate CH3COCH3 Junsei Moderate
Ethyl acetate C2H5COCH3 Junsei Moderate
Ethyl propionate C2H5COC2H5 Junsei Moderate
a
Hydrogen-bonding capability was extracted from Ref. [30].

Scheme 1. Experimental workflow followed in this study.

3
R. Nguele et al. Fuel 311 (2022) 122559

Table 3 2.2.3. Asphaltene solubility in different solvents


Elemental compositions of extracted asphaltenes. A gravimetric approach was employed to evaluate the interactions
Elemental composition (wt.%) between the extracted asphaltenes and different solvents following the
workflow illustrated in Scheme 2.
C1s O1s Si2p S2p Fe2p Zn2p
Two set of experiments were conducted in this study. The first set
A 51.4 42.8 3.90 1.92 – – evaluates the contribution of the dispersing medium. An asphaltene
B 57.2 39.3 2.27 1.26 – –
C 71.7 27.3 – – – 1.04
solution of 1 wt% was prepared by dissolving one gram of extracted
D 82.9 14.5 – 2.51 – – asphaltene into 100 g of either a polar (cyclic ethers) or nonpolar
E 83.0 16.7 – – 0.34 – dispersant (aromatic solvents). The asphaltenes were re-precipitated by
the addition of an excess of pentane (1:20 vol ratio) followed by filtra­
tion through a pre-weighed 0.22 µm filter paper.
overnight at atmospheric conditions. The precipitates were separated
The second set of experiments aimed to evaluate the contribution of
from the supernatant fluid by vacuum filtration using a pre-weighed
the titrant (i.e., solvent), which promotes the precipitation of asphal­
filter paper (0.45 µm). The filtered solids (asphaltenes) were then
tenes. In this series of experiments, a separate asphaltene solution
dried at ambient conditions; after this step, no further purification of the
(10,000 ppm) was prepared by dissolving the asphaltenes obtained from
precipitated asphaltenes was performed.
the parent oil into toluene. For each investigated titrant (Table 2), one
volume of the toluene solution was then added to 20 volumes of the
2.2.2. Spectral analyses
titrant. The removal of the precipitated asphaltenes was then performed
The spectral analyses used in this study included Raman, FTIR, and
per the procedure described above.
XPS spectroscopy.
The mass balance of asphaltene in the bulk solution could be written
Raman spectra were acquired using a laser Raman spectroscopy in­
as,
strument (ARAMIS, Aramis Horiba, Japan) which uses a 532 nm Lexel
Ar + Laser. The spectra were obtained in a backscattering configuration mi = mna + mcl + mpr (1)
with a 50 × long work distance objective used to focus the laser.
Fourier Transform Infrared Spectra (FT/IR) were acquired using a where mi is initial mass of the asphaltene (in g), mna is the mass of
Jasco spectrophotometer (Model 4100, Japan). All the absorption asphaltene nanoaggregates formed upon dissolving into the dispersant
spectra were acquired from 400 to 4000 cm− 1 with 48 scans and a res­
olution of 1 cm− 1; in addition, a baseline correction was performed prior
to the analysis.
X-ray photoelectron spectroscopy (XPS) was used to determine the
elemental composition of the extracted asphaltenes. The spectra were
obtained using an AXIS-ULTRA DLD instrument (Shimadzu). XPS results
revealed the photoemissions of Carbon (C1s) and oxygen (O1s) occurred
at binding energy values of 284 eV and 535 eV, respectively (Figure S1,
Supplementary file). Although less prominent, photoemission of sulfur
(S2p) at ~ 166 eV was also recorded. Additionally, transition metals
(iron, Fe2p) and metalloids (silicon, Si2p) were also photo-emitted.
Table 3 outlines the elemental composition of extracted asphaltenes.
The highest carbon content was obtained for the asphaltene extrac­
ted from bituminous crude (83%), while the lowest corresponds to the
asphaltene extracted from the Middle-Eastern heavy oil (51.7%). The
main takeaway from XPS analysis is that the total metal concentration
was too low (i.e., <3%) for their contribution to be significant to the
asphaltene solubility. A more appropriate technique i.e., ICP-Emission
Atomic Spectroscopy (AES) would provide more accurate results on
the influence of metal content to the asphaltene solubility. This point,
however, was beyond the scope of this work.
Fig. 2. Raman spectra of investigated asphaltenes.

Scheme 2. Experimental workflow used in the present study to compute the average solubility; the values 2–5 nm, 5–12 nm, and > 500 nm were extracted from
Ref. [16] and represent respectively the size of nanoaggregates, asphaltene clusters and aggregates.

4
R. Nguele et al. Fuel 311 (2022) 122559

(in g), mcl is the mass of asphaltene clusters formed (in g), and mpr is the intensity of the G band (dimensionless), and ID1 is the integrated in­
precipitated asphaltene from the bulk solution (g). tensity of the D1 band (dimensionless).
Within the scope of the present work, mpr was obtained from vacuum To apply Eq. 3, Raman spectra were first baseline-corrected (Fig. 3)
the filtration experiments. Thus, the asphaltene apparent solubility (Sap) and the D1 and G bands were fitted using a Gaussian function (Fig. 4).
was defined as the concentration in asphaltenes that remains in sus­ The area under each respective band was then estimated and La was
pension in the bulk solution either nanoaggregates or asphaltene clus­ computed using Eq. 3. The results are summarized in Table 4.
ters or even both. Sap was obtained from Eq. 2, Based on literature, the expected theoretical range of La estimated
( ) from Raman spectroscopy is between 1.1 and 2.1 nm [22,25,34]. Our
Sap = 1 − mpr /mi (2)
results showed that only three of the five samples fall within this range,
A value of S close to 1 would express an apparent stable asphaltene namely, asphaltenes (A), (B), and (D) extracted respectively from
solution, whereas a value close to 0 would imply most of the asphaltenes Middle-Eastern, Central African, and North American crude. Compared
had been precipitated. It is worth mentioning the asphaltene solubility, to the aromatic sheet values from crude oils sampled in the same
measured using gravimetric approach, could be bias due to the poly­ geological regions [25], the La values obtained in this study were 11 and
dispersity of the asphaltenes. As such, the apparent solubility (Eq. 2) 15 % larger respectively for asphaltenes (A) and (D).
expresses only a part of the dissolved sampled and it is not representa­ A plausible explanation would be the maturation process of the
tive of the average concentration in asphaltenes that would precipitate. kerogen from which the oils were sampled [35]. The largest La was
obtained from the West-African bitumen, which was threefold higher
3. Results and discussion than any other asphaltene samples in our study. Unfortunately, and to
the best of our knowledge, there are no available works at present that
3.1. Structural parameters estimation report the aromatic sheet diameter of asphaltenes extracted from Afri­
can crudes. Nevertheless, these findings suggest that under the same
3.1.1. Raman studies and aromatic sheet diameter estimation conditions, asphaltene (E) would be less stable than all other asphaltenes
Fig. 2 shows the Raman spectra for all investigated asphaltenes. due to rapid flocculation of the clusters formed from nanoaggregates
Qualitatively, all acquired spectra were consistent with two shoul­ [19].
ders appearing around Raman shift values of 1350 cm− 1 and 1580 cm− 1,
termed respectively elsewhere as D1 and G bands [32]. The D1 band is 3.1.2. FTIR studies and asphaltene aromaticity
due to vibration of sp2 C–C in aromatic fused rings changed by in-plane Fig. 5 shows the infrared spectra of the extracted asphaltenes.
defects or heteroatoms. The G band, however, is attributed to stretching Characteristic absorptions were observed in the IR band at ~ 1450
of the sp2 C–C bond in ordered aromatic fused rings [33]. For most of cm− 1, ~1600 cm− 1, ~1700 cm- 1, ~2900 cm- 1, and ~ 3100 cm− 1.
the extracted asphaltenes, a slight shift was observed, which could be These peaks are attributed respectively to: (1) the bending of sp3 C–H
due to the carbon orientation within the asphaltene [6]. bonds; (2) the stretching of aromatic sp2 C; (3) a vibration stretch of
For all the asphaltene samples, except asphaltene (B), the G band is C–– O bonds from a functional group (e.g., acid); (4) sp3 C–H bonds
sharply defined whereas the D1 band appears broad. This suggests a from aliphatic chains; and (5) vibrational stretching of O–H from an
poorly ordered carbonaceous asphaltene or low short-range order of the alcohol [26,36]. Additionally, the IR spectra show a strong shoulder in
aromatic sheet [25]. Additionally, previous literature reports that a the IR band at a wavelength of ~ 2400 cm− 1 (gray shaded circles) for
stronger peak intensity suggests the presence of ring strain and thus a asphaltenes (C), (D), and (E). This IR band is often associated with the
large aromatic sheet diameter. Quantitatively, the D1 and G bands can absorption of gaseous carbon dioxide (CO2), which would suggest that
be used to estimate the size of the asphaltene structure. This is achieved some CO2 molecules might have been adsorbed to the sample surface
by integrating the intensity of peaks at or near 1350 cm− 1 and 1580 during drying, which was conducted at ambient conditions.
cm− 1 and applying Eq. 3: In order to estimate the aromaticity of the asphaltenes, the ratio of
the integrated peaks corresponding to the absorption of sp2 C– – C stretch
La = 4.4IG /ID1 (3) at 1600 cm− 1 to the absorption of sp3 C–H at 2900 cm- 1 should be used.
If the bending mode is considered, the aromaticity is computed from the
where La is the aromatic diameter sheet (in nm), IG is the integrated
ratio of the absorption of sp2 C–
– C stretch at 600–800 cm− 1 to the ab­
sorption of sp C H at 1450 cm- 1. In this work, the structural param­
3 –

eters were determined from IR bands showing the stretching vibration


mode as it appears prominent in all the IR spectra. In addition, the in­
tegrated peak ratio 3400 cm- 1/3100 cm− 1 would estimate the likelihood
of an asphaltene to aggregate due to hydrogen bonding [16]. Lastly, the
integrated peak ratio 1700 cm- 1/2900 cm− 1 estimates the concentration
of carboxylic acid to aliphatic carbon. It should be noted that deter­
mining the aforementioned parameters assumes that the extinction co­
efficients of the functional groups do not change among the fraction
[26,37]. For each spectrum, the peaks of interest were baseline-
corrected, deconvoluted using Gaussian fit, and the integrated areas
were then computed.
The plots of the deconvoluted spectra obtained are appended in the
Supporting file of this manuscript (Figure S3, S4). The results of the
estimated parameters as a function of aromatic sheet diameter are
summarized in Table 5.
The ratio 1600 cm- 1/2900 cm− 1, an indicator of the aromaticity,
somewhat increased with increasing aromatic core size; the larger the
La, the higher the aromaticity. Nonetheless, our results are consistent
with the findings reported by Rogel et al. [26,38]. On the other hand, the
Fig. 3. Baseline-corrected Raman spectra for extracted asphaltenes. Approxi­ hydrogen bonding capability (3400 cm- 1/3100 cm− 1) and carboxylic-
mate extent of D1 and G bands highlighted in gray. to-aliphatic carbon ratio (1700 cm− 1/2900 cm− 1) showed opposing

5
R. Nguele et al. Fuel 311 (2022) 122559

Fig. 4. Raman raw data (white points) and baseline-corrected spectra (in black and blue) fitted with the Gaussian function of extracted asphaltenes (in red) for
samples A–E; The black line denotes D1 band, while the G band is represented by the blue line. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

trends. The former decreases with increasing La while the latter shows a concentration in precipitated asphaltenes was respectively ± 0.05,
reverse trend. The observations discussed above appear true for all the ± 0.67, ± 1.47, and ± 0.65 mg/g respectively for asphaltene A, C, D, and
investigated samples except asphaltene D, which shows the largest E. A summary of the concentration in precipitated asphaltenes is pre­
values for either parameters. These results would suggest that asphal­ sented in Table 6.
tenes with large core sizes might hinder H-bonding and their aggrega­ On average, the precipitate values obtained from 1 g of dispersed
tion might be a result of polar interactions. asphaltene in aromatic hydrocarbons (i.e. non-polar solvents) were
24.9, 5.2, 2.4, and 3.4 mg/g for asphaltenes (D), (A), (C), and (E),
3.2. Structural parameters and asphaltene solubility respectively. The larger the size of the asphaltene core size, the lower the
obtained precipitate value. The concentration increased with the length
Herein, the structural parameters of extracted asphaltenes will be of the alkyl chain attached to the phenyl group, irrespective of the core
correlated with the apparent solubility. It is worth mentioning that the size of the asphaltenes (Table 5). The same trend was observed in polar
subsequent results were found by studying one parameter at a time. solvents including cyclic ethers and aromatic amine.
However, a multivariable analysis would be a better approach to analyze In terms of asphaltene apparent solubility, we found that (1)
the data than checking one parameter at a time. asphaltenes were apparently more stable in nonpolar solvents, and (2)
their solubility increases with the aromatic core size (Fig. 6).
3.2.1. Influence of dispersant type Fig. 6 further indicates that the self-aggregation of the asphaltenes is
To evaluate the contribution of dispersants, a total of eight solvents controlled by the nature of the dispersant. Gray et al. (2011) argued that
were used, including four non-polar and four polar solvents. Each set of the process of asphaltene self-assembly is a combination of both intra-
experiments was repeated three times. The repeatability error of the and intermolecular forces, including π-π stacking and H-bonding [39],

6
R. Nguele et al. Fuel 311 (2022) 122559

Table 4 Table 6
Raman spectra analysis and calculated La Values using the Gaussian Function for Summary of the concentration of precipitated asphaltenes in different disper­
extracted asphaltenes.a sants; all concentrations are expressed in mg/g.
Sample Origin Band Raman FWHM Area La Dispersant D (2.03 A (2.07 C (2.35 E (7.34
Shift (cm− 1) (nm) nm) nm) nm) nm)
Class Name
(cm− 1)
Aromatic Benzene 0.846 2.515 1.077 0.980
A Middle East D1 1374.6 260.5 61195.9 2.07
hydrocarbon Toluene 7.793 4.780 2.041 1.148
G 1565.9 147.3 28726.4
Xylene 85.81 – 5.659 4.366
B Central D1 1403.3 183.3 78683.0 1.52
Ethylbenzene 5.324 8.272 0.745 7.283
Africa G 1576.6 174.5 27186.8
Cyclic ethers Tetrahydrofuran 11.27 2.372 1.045 0.810
C Northern D1 1363.5 85.60 4224.3 2.35
Furan 192.6 – 1.076 0.766
Asia G 1570.4 70.77 2255.7
Furfural 80.99 – 1.053 1.503
D North D1 1395.8 167.7 93852.2 2.03
Aromatic Aniline 150.6 8.691 4.042 6.074
America G 1570.5 113.7 43218.6
amine
E West Africa D1 1400.0 74.32 33068.2 7.34
G 1607.6 81.38 55168.5
Ref. Saudi Arabia D1 1343 174.6 215848.3 1.88
[25] (Middle East) G 1584 74.2 92299.5
Canada D1 1360 198.81 259549.7 1.72
(North G 1576 111.69 101401.3
America)
a
FWHM: Full width at half maximum.

Fig. 6. Influence of aromatic core size on average asphaltene apparent solu­


bility; asphaltene samples were dissolved in different dispersants and precipi­
tated using pentane.

Table 7
Fig. 5. FTIR spectra of heptane-precipitated asphaltenes. Summary of the concentration of precipitated asphaltenes in different titrants.
All concentrations are expressed in mg/g.
Titrant B D A C E
Table 5
Estimated aromaticity, hydrogen bonding propensity and carboxyl group con­ Class Name (1.52 (2.03 (2.07 (2.35 (7.34
nm) nm) nm) nm) nm)
tent as a function of aromatic sheet diameter
Alkane Pentane 28.25 7.793 4.780 2.041 1.148
La Aromaticity Carboxylic-to- Aggregation
Hexane 7.869 – 9.567 44.16 130.0
(nm) (1600 cm− 1/ aliphatic carbon ratio propensity (3400
Heptane 10.49 0.510 9.751 1.255 4.775
2900 cm− 1) (1700 cm− 1/2900 cm− 1/3100 cm− 1)
Decane 20.27 0.583 8.413 21.30 109.7
cm− 1)
Esters Methyl 73.30 45.51 8.411 9.144 1.290
B 1.52 0.119 ± 0.001 Not detected on IR Not detected on IR Acetate
D 2.03 9.960 ± 3.064 5.459 ± 2.422 0.687 ± 0.161 Ethyl acetate 45.99 1.570 12.36 15.48 0.924
A 2.07 0.505 ± 0.120 0.505 ± 0.120 10.52 ± 2.767 Ethyl 10.89 – 21.16 7.107 21.78
C 2.35 0.426 ± 0.188 0.369 ± 0.001 2.257 ± 0.503 propionate
E 7.34 1.521 ± 0.802 0.998 ± 0.694 0.371 ± 0.089

3.2.2. Influence of titrants


whose formation propensity was estimated by FTIR spectroscopy To investigate the influence of titrants, asphaltenes were re-
(Fig. 5b). The solubility of asphaltene appears to be inversely propor­ precipitated from a bulk solution consisting of 1 g of asphaltene
tional to the H-bonding, which itself decreases with aromatic core size. dispersed in 100 g of toluene. Two types of titrants were considered,
This observation validates the concept that large aromatic cores are including alkanes and esters. Herein, the series of experiments were
sterically hindered due to their aliphatic side chains and alkyl chains repeated also three times and the repeatability error for asphaltene A, B,
[19]. These findings are also consistent with the results reported by C, D, and E was ± 0.05, ± 0.30, ± 0.68, ± 0.64, and ± 0.74 mg/g,
Shetty et al. (1996), who argued that aromatic solvents tend to decrease respectively. The results are summarized in Table 7.
π-π interactions [40]. The polarity of the solvents alters the stability of Fig. 7 shows the relationship between asphaltene solubility and the
asphaltene such that H-bonding is promoted over π-π interactions. aromatic core size.

7
R. Nguele et al. Fuel 311 (2022) 122559

Fig. 7. Effect of titrant polarity on asphaltene solubility of asphaltenes; asphaltene samples were diluted in toluene.

Fig. 8. Asphaltene precipitation and solubility in the presence of different inhibitors.

As opposed to the contribution of dispersants, we found that carbon dioxide (CO2) were performed. CO2 was selected because of its
asphaltenes with large aromatic core sizes were less soluble in nonpolar reported impact on asphaltene precipitation [1,41].
solvents (Fig. 7). The average precipitation in alkanes increased mark­
edly from 16.7 to 61.4 mg/g as core size increased from 1.52 to 7.34 nm
3.3. Solvent polarity and asphaltene precipitation
(Table 6). However, an opposing trend was observed when polar sol­
vents were used. The general tendency shows that large asphaltenes’
Two oxygen-containing polar chemicals were selected based on their
cores precipitate less when polar solvents are used, suggesting that large
polarity; these included polyvinyl alcohol (PVOH) and ethylene glycol
aromatic cores are stabilized by H-bonding.
(EG). The preparation of the inhibitors as well as the experimental
It is worth highlighting that the apparent solubility, reported in
procedure are appended in the Supporting file of this manuscript. The
Figs. 6 and 7, was relatively large (>0.80 g/g). It might suggest at first
results of the analysis are shown in Fig. 8.
sight that most of the asphaltenes remained in the bulk solution either as
The general trend shows that fewer asphaltenes precipitate when the
nanoaggregates, clusters or both. This would mean that the apparent
core size is large (Fig. 8a) and consequently more stable (Fig. 8b). This
solubility should be seen as the propensity of either the dispersant or the finding is consistent with the data shown in Fig. 7. For all cases, the
titrant to promote the formation of clusters, and to floc those clusters
bubbling of CO2 initiated the self-aggregation of asphaltenes, as evi­
into large aggregates (precipitates). denced by the decrease in asphaltene solubility. The addition of EG
Nevertheless, these findings from Figs. 6 and 7 evidence the role of
caused more asphaltenes to precipitate, plausibly due to the effect of
H-bonding on asphaltene precipitation, and, furthermore, the role of π-π strong H-bonding.
bonding. Extending these observations to engineering applications (the
As an example, the concentration of precipitated asphaltenes in the
petroleum industry, for example), one can theorize that asphaltene control experiment for asphaltene (A) was calculated to be 2.29 mg/g.
precipitation from a bulk solution can be prevented if the polarity of the
Bubbling CO2 in the blank solution increased the concentration fourfold
medium is controlled. To verify this hypothesis, a series of experiments (~9.15 mg/g). The same bubbling performed in blank and EG yielded a
in which asphaltene was precipitated from a model crude oil using
concentration of 22.9 mg/g, a value tenfold larger than the initial

8
R. Nguele et al. Fuel 311 (2022) 122559

conditions. The same trend was also observed for larger aromatic cores. org/10.1016/j.fuel.2021.122559.
These results validate that asphaltene solubility is inversely proportional
to the size of aromatic cores in a polar medium. References
In non-polar solvents, or at least a less polar medium (i.e. PVOH in
this study), asphaltenes are likely to be more soluble due to the alter­ [1] Alimohammadi S, Zendehboudi S, James L. A comprehensive review of asphaltene
deposition in petroleum reservoirs: theory, challenges, and tips. Fuel 2019;252:
ation of H-bonding between the H-donor of the solvent and the alkyl 753–91.
chain of the asphaltene, which is consistent with the literature [42]. As [2] Speight JG, Moschopedis SE. On the Molecular Nature of Petroleum Asphaltenes,
discussed above, the inhibition of the asphaltene precipitation requires 1982, p. 1–15.
[3] Schiessler RW. The Science of Petroleum: Crude Oils, Chemical and Physical
the use of chemicals that primarily alter the H-bonding responsible for Properties, Vol. V, Part I. Benjamin T. Brooks and A. E. Dunstan, Eds. New York:
the asphaltenes’ self-aggregation. The concentration of the inhibitor Oxford Univ. Press, 1950. 200 pp. $11.00. Science (80-) 1951;113:454–454.
would also be likely to have an impact, however, this aspect is beyond [4] Speight JG. Petroleum asphaltenes - part 1: asphaltenes, resins and the structure of
petroleum. Oil Gas Sci Technol 2004;59(5):467–77.
the scope of the current study. [5] Mullins OC, Sabbah H, Eyssautier J, Pomerantz AE, Barré L, Andrews AB, et al.
Advances in asphaltene science and the Yen-Mullins Model. Energy Fuels 2012;26
4. Conclusions (7):3986–4003.
[6] Zuo P, Qu S, Shen W. Asphaltenes: separations, structural analysis and
applications. J Energy Chem 2019;34:186–207.
In this study, we investigate the contribution of structural parame­ [7] Martín-Martínez FJ, Fini EH, Buehler MJ. Molecular asphaltene models based on
ters of asphaltenes on their solubility; in particular, we explored the Clar sextet theory. RSC Adv 2015;5(1):753–9.
effects of aromaticity and aromatic sheet diameters. Asphaltenes were [8] Liao. What are asphaltenes in petroleum , oilsands , and heavy oil ? Energy & Fuels
2009;1:1–4.
extracted from five different crude oils originating from North America, [9] Hu C, Sabio JC, Yen A, Joshi N, Hartman RL. Role of water on the precipitation and
Asia, Central and West Africa, and the Middle East. Both Raman and deposition of asphaltenes in packed-bed microreactors. Ind Eng Chem Res 2015;54
Fourier Infrared spectroscopy were employed to estimate, respectively, (16):4103–12.
[10] Nguyen MT, Nguyen DLT, Xia C, Nguyen TB, Shokouhimehr M, Sana SS, et al.
the aromatic sheet diameter and the aromaticity of the extracted Recent advances in asphaltene transformation in heavy oil hydroprocessing:
asphaltenes. Asphaltene solubility was further evaluated by gravimetric Progress, challenges, and future perspectives. vol. 213.
analyses. [11] Yao B, Chen W, Li C, Yang F, Sun G, Wang G, et al. Polar asphaltenes facilitate the
flow improving performance of polyethylene-vinyl acetate. Fuel Process Technol
Our findings show that the aromaticity increases with the aromatic 2020;207:106481.
sheet diameter, which itself is inversely proportional to the hydrogen [12] Kaiser MJ. Flow assurance issues. Offshore Pipeline Constr Ind 2020:39–60.
capability of the asphaltenes. The solubility of asphaltenes depends on [13] Wang JX, Buckley JS, Burke NA, Creek JL. Anticipating Asphaltene Problems
Offshore-A Practical Approach. Proc Annu Offshore Technol Conf 2003;2003-May:
the polarity of the precipitant rather than that of the dispersed medium. 1686–95.
In terms of the dispersant, asphaltenes were more stable in non-polar [14] Dickie JP, Yen TF. Macrostructures of the asphaltic fractions by various
solvents, and their solubility increases with aromatic core size due to instrumental methods. Anal Chem 1967;39(14):1847–52.
[15] Hoepfner MP, Limsakoune V, Chuenmeechao V, Maqbool T, Fogler HS.
π-π interactions. However, asphaltenes precipitate less when polar pre­
A fundamental study of asphaltene deposition. Energy Fuels 2013;27(2):725–35.
cipitants are used due to the effects of hydrogen bonding. From this [16] Alhreez M, Wen D. Molecular structure characterization of asphaltene in the
study, we conclude that to inhibit the aggregation of large fused cores, presence of inhibitors with nanoemulsions. RSC Adv 2019;9(34):19560–70.
polar inhibitors should be used, whereas nonpolar chemicals could be [17] Wiehe IA. Asphaltene Solubility and Fluid Compatibility. Energy Fuels 2012;26(7):
4004–16.
effective for small aromatic cores. [18] Adams JJ. Asphaltene adsorption, a literature review. Energy Fuels 2014;28(5):
2831–56.
Funding [19] Sedghi M, Goual L, Welch W, Kubelka J. Effect of asphaltene structure on
association and aggregation using molecular dynamics. J Phys Chem B 2013;117
(18):5765–76.
This work was supported by United Arab Emirates University (Fund [20] Bestougeff MA, Byramjee RJ. Chapter 3 Chemical Constitution of Asphaltenes. Dev.
Number 12S016). Pet. Sci., vol. 40, Elsevier; 1994, p. 67–94.
[21] Daaou M, Bendedouch D, Bouhadda Y, Vernex-Loset L, Modaressi A, Rogalski M.
Explaining the flocculation of Hassi Messaoud asphaltenes in terms of structural
CRediT authorship contribution statement characteristics of monomers and aggregates. Energy Fuels 2009;23(11):5556–63.
[22] Riedeman JS, Kadasala NR, Wei A, Kenttämaa HI. Characterization of asphaltene
deposits by using mass spectrometry and Raman Spectroscopy. Energy Fuels 2016;
Ronald Nguele: Conceptualization, Methodology, Writing – original 30(2):805–9.
draft. Albert Brandt Mbouopda Poupi: Formal analysis, Investigation, [23] Ibrahim HH, Idem RO. Interrelationships between asphaltene precipitation
Writing – review & editing. Ghislain A. Mouthe Anombogo: Re­ inhibitor effectiveness, asphaltenes characteristics, and precipitation behavior
during n-heptane (light paraffin hydrocarbon)-induced asphaltene precipitation.
sources. Olalekan S. Alade: Resources. Hakim Saibi: Funding acqui­
Energy and Fuels.
sition, Project administration, Supervision, Writing – review & editing. [24] Wattana P, Fogler HS, Yen A, Carmen Garcìa MarìaD, Carbognani L.
Characterization of polarity-based asphaltene subfractions. Energy Fuels 2005;19
(1):101–10.
Declaration of Competing Interest
[25] Abdallah WA, Yang Y. Raman spectrum of asphaltene. Energy Fuels 2012;26(11):
6888–96.
The authors declare that they have no known competing financial [26] Rogel E, Miao T, Vien J, Roye M. Comparing asphaltenes: Deposit versus crude oil.
interests or personal relationships that could have appeared to influence Fuel 2015;147:155–60.
[27] Rogel E, León O, Espidel Y, González Y. Asphaltene stability in crude oils. SPE Prod
the work reported in this paper. Facil 2001;16:84–8.
[28] Guzmán HéctorJ, Isquierdo F, Carbognani L, Vitale G, Scott CE, Pereira-Almao P.
Acknowledgments X-ray photoelectron spectroscopy analysis of hydrotreated athabasca asphaltenes.
Energy Fuels 2017;31(10):10706–17.
[29] D664-18e2 A. Standard Test Method for Acid Number of Petroleum Products by
The authors extend their gratitude to Mitsubishi Chemicals for sup­ Potentiometric Titration. ASTM Int., West Conshohocken, PA: 2018.
plying the polymer used in this study. Japan Petroleum Exploration is [30] Barton AFM. CRC Handbook of Solubility Parameters and Other Cohesion
Parameters. Second Edition. Boca Raton, FL: CRC Press; 1991.
also acknowledged for supplying the dead oils from which asphaltenes [31] Nguele R, Sasaki K. Asphaltene behavior at the interface oil-nanofluids:
(C) and (D) were extracted. This research was funded by the Research implications to adsorption. Colloids Surf A Physicochem Eng Asp 2021;622:
Office of United Arab Emirates University (Fund 12S016). 126630.
[32] Tuinstra F, Koenig JL. Raman spectrum of graphite. J Chem Phys 1970;53(3):
1126–30.
Appendix A. Supplementary data [33] Larkin PJ. IR and Raman Spectroscopy. Elsevier; 2011.

Supplementary data to this article can be found online at https://doi.

9
R. Nguele et al. Fuel 311 (2022) 122559

[34] Li K, Vasiliu M, McAlpin CR, Yang Y, Dixon DA, Voorhees KJ, et al. Further insights [39] Gray MR, Tykwinski RR, Stryker JM, Tan X. Supramolecular assembly model for
into the structure and chemistry of the Gilsonite asphaltene from a combined aggregation of petroleum asphaltenes. Energy Fuels 2011;25(7):3125–34.
theoretical and experimental approach. Fuel 2015;157:16–20. [40] Shetty AS, Zhang J, Moore JS. Aromatic π-stacking in solution as revealed through
[35] Speight JG. Occurrence and Formation of Crude Oil and Natural Gas. Subsea Deep. the aggregation of phenylacetylene macrocycles. J Am Chem Soc 1996;118(5):
Oil Gas Sci. Technol., Gulf Professional Publishing; 2015, p. 1–43. 1019–27.
[36] Coelho RR, Hovell I, de Mello Monte MB, Middea A, Lopes de Souza A. [41] Fakher S, Imqam A. Asphaltene precipitation and deposition during CO2 injection
Characterisation of aliphatic chains in vacuum residues (VRs) of asphaltenes and in nano shale pore structure and its impact on oil recovery. Fuel 2019;237:
resins using molecular modelling and FTIR techniques. Fuel Process Technol 2006; 1029–39.
87(4):325–33. [42] Nguele R, Hirota H, Sugai Y, Sasaki K. Role of polymer-based nanofluids on
[37] Bellamy LJ. The Infra-red Spectra of Complex Molecules. vol. 3. Dordrecht: asphaltene adsorption during carbon dioxide (CO2) injection. Energy Fuels 2021;
Springer Netherlands; 1975. 35(18):14746–57. https://doi.org/10.1021/acs.energyfuels.1c02333.
[38] Rogel E, Moir M. Effect of precipitation time and solvent power on asphaltene
characteristics. Fuel 2017;208:271–80.

10

You might also like