You are on page 1of 10

Article

Cite This: Energy Fuels 2019, 33, 9671−9680 pubs.acs.org/EF

Effect of the NiO/SiO2 Nanoparticles-Assisted Ultrasound Cavitation


Process on the Rheological Properties of Heavy Crude Oil: Steady
State Rheometry and Oscillatory Tests
Daniel Montes,† Esteban A. Taborda,† Mario Minale,*,‡ Farid B. Cortés,† and Camilo A. Franco*,†

Grupo de Investigación en Fenómenos de Superficie−Michael Polanyi, Facultad de Minas, Universidad Nacional de Colombia,
Sede Medellín 4309000, Colombia

Department of Engineering, University of Campania Luigi Vanvitelli, via Roma 29, 81031 Aversa, Caserta, Campania, Italy
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF GOTHENBURG on October 21, 2019 at 20:48:24 (UTC).

ABSTRACT: This manuscript has the primary objective of demonstrating the changes in the rheological behavior of heavy
crude oils (HO) in response to the application of ultrasound cavitation assisted with NiO-functionalized SiO2 nanoparticles
(SiNi1). A HO with an asphaltene mass fraction of 17.0% was used for the tests which were carried out at 25 °C and 1 atm with
a fixed ultrasound frequency and power of 37 kHz and 400 W, respectively. The viscosity measurements were performed on
four different samples: the HO in the absence of nanoparticles and ultrasound irradiation (sample A), the HO alongside
ultrasound irradiation (sample B), the HO with the addition of nanoparticles (sample C), and, finally, the HO in the presence
of the mentioned nanomaterial and ultrasound irradiation (sample D). It was observed that a single treatment, whatever it is,
only slightly changed the original HO rheology, while the cooperative action of ultrasound cavitation and nanoparticles addition
induced several measurable differences with respect to the HO: The viscosity was reduced up to 50−60%, depending on the
applied shear rate; the power per unit of volume dissipated during a hysteresis cycle was decreased of about the 70%; the sample
elasticity was measurably reduced, and accordingly, the relaxation time measurable for the original HO was not detectable
anymore. These findings can be explained by hypothesizing that the original HO viscoelastic microstructure is broken down by
the proposed combined treatments as the asphaltenes may stably adsorb on the nanoparticles and the subsequent size reduction
process of the asphaltene aggregates is enhanced by the ultrasound irradiation.

1. INTRODUCTION applied to a sample that has been previously at rest and the
Currently, heavy oil (HO) production and transportation are subsequent recovery of viscosity in time when the flow is
significantly increased to compensate for the depletion of light discontinued”.13 The transient behavior of a fluid can be
oil reservoirs.1,2 However, the physicochemical properties of investigated to study thixotropy, and in particular, the
these fluids, such as high density, high asphaltene content, and hysteresis technique, introduced by Green and Weltman,14
high viscosity, lead to low-efficiency operations with non-cost- can be used. It consists in increasing and decreasing the
effective results.3 applied shear rate from zero to a maximum value. When the
In heavy oils, the high content of asphaltenes and resins stress is plotted vs the shear rate, typically a hysteresis loop is
leads to the formation of complex hierarchical microstructures4 detected. The loop dimension will depend on the rate at which
that may also arrange to form a viscoelastic network,5 which the shear rate is varied and on the resting time.15,16 This effect
results in extra-high oil viscosity, that is an essential parameter may be also related to an irreversible microstructure
for production and transportation designing.6 Hence, several modification that induces a plastic deformation, usual in
investigations are focused on the crude oil dilution for solids,17 which is associated with an energy loss. The study of
adjusting its viscosity to pipeline standards.6−9 Nonetheless, the HO thixotropy is essential during the designing of
addition of solvents tends to worsen transportation or transport operations to avoid overestimations of the amounts
production operations. Indeed, the solvent, depending on its of diluents and of the temperature necessary to reach pipeline
chemical nature, may induce an undesired modification of the transportation standards or to avoid possible underestimations
crude oil microstructure that can drive even to the that could affect equipment integrity.18
precipitation of asphaltenes.4 Mortazavi-Manesh et al.18 studied the thixotropic behavior
Moreover, the complex heavy crude oil matrix also affects of Maya crude oil using hysteresis loop tests concluding that
other HO rheological properties, and often, both thixotropy10 thixotropy is reduced upon temperature increase. With
and viscoelasticity11,12 are observed. Viscoelasticity induces to dynamic experiments, Wardhaugh et al.12 characterized the
a fluid a behavior in between an elastic solid, able to recover an viscoelasticity of waxy crude oils to investigate how it affects
imposed strain when the applied stress is removed, and a
purely viscous fluid, able to dissipate all the energy Received: July 12, 2019
accumulated during a deformation. Thixotropy is defined as Revised: August 27, 2019
“the continuous decrease of viscosity with time when flow is Published: September 6, 2019

© 2019 American Chemical Society 9671 DOI: 10.1021/acs.energyfuels.9b02288


Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

Table 1. Investigated Samples


A B C D
heavy oil heavy oil + 2000 mg·L−1 SiNi1 heavy oil + sonication for 90 min heavy oil + 2000 mg·L−1 SiNi1 nanoparticles + sonication for 90
nanoparticles 400 W/37 kHz min 400 W/37 kHz

the yield point. They found that the 3D internal microstructure 2. MATERIALS AND METHODS
of the investigated oils induces a solid-like behavior with 2.1. Materials. Nickel oxide nanocrystals functionalized over 7 nm
conventional elastic, creep, and fracture properties. Ghannam fumed silica nanoparticles (SiNi1) were synthesized following the
et al.16 evaluated the HO rheological properties with the procedure described in a previous study.33 More information about
addition of light hydrocarbons, concluding that the samples the characterization of the nanoparticles is reported in previous
were shear thinning and showed time dependency related to papers.29−31 The HO used has a density corresponding to 13°API and
approximately 40% of emulsified water. SARA measurement was
thixotropy, which decreases with the addition of light carried out through a microdeasphalting technique coupled with thin
hydrocarbons. Moreover, Taborda et al.19 studied the HO layer chromatography following the IP 469 standard35 and using a
internal structure alteration caused by nanoparticles addition TLC-FID/FPD Iatroscan MK6 (Iatron Laboratories Inc., Tokyo,
through steady state and dynamic rheological tests. They found Japan). The characterization results showed that mass fractions of
that the addition of nanoparticles significantly reduces the oil saturates, aromatics, resins, and asphaltenes are 18.2%, 23.1%, 41.7%,
and 17.0%, respectively.
viscosity, yield stress, and elasticity.
Four different samples are investigated: (A) HO, (B) HO with
In recent years, cavitation has been exploited as a technique nanoparticles, (C) sonicated HO, and (D) HO with nanoparticles
capable of reducing a fluid viscosity.20,21 Rahimi et al.22 studied subsequently sonicated (Table 1).
the combined effect of ultrasonic irradiation and heating on the 2.2. Methods. 2.2.1. Asphaltenes Characterization and
HO viscosity, showing that the viscosity reduction induced by Aggregates Size Measurements. The asphaltenes were characterized
cavitation increases with temperature. Furthermore, also the by their chemical composition, and aggregation kinetics were also
measured for confirming the existence of the viscoelastic network.
addition of nanoparticles to an HO improves the effect of First, the asphaltenes were extracted from the crude oil under
cavitation on the viscosity reduction thanks to their high investigation following the method described in previous works.24,32
dispersibility, high surface-area-to-volume-ratios, high asphal- Concisely, a certain HO volume was diluted with n-heptane in a 40:1
tenes selectivity, and adsorptive capacity.23−26 Askarian et al.27 crude oil/diluent ratio. The mixture was sonicated for 2 h and
employed metal nanoparticles for this purpose, obtaining a subsequently stirred for 20 h at 300 rpm; then it was filtered and
diluted again in a 4:1 ratio with n-heptane. The suspension was
final viscosity reduction of 20% due to the enhancement of the centrifuged for 30 min at 4500 rpm for obtaining the asphaltenes,
viscoelastic network disruption. Recently, hybrid nanoparticles which were washed with n-heptane several times until no impurities
of supported hygroscopic salts (SHS) showed excellent were observed in the residue.
asphaltene adsorptive capacities,28−32 making them suitable The asphaltenes chemical composition was characterized through
for enhancing the ultrasound-induced viscosity reduction. X-ray photoelectron spectroscopy (XPS) with a PHOIBOS 150 1D-
According to the authors, in the literature, the coupled effect DLD (SPECS GmbH, Berlin, Germany) photoelectronic X-ray
spectrometer (NAP-XPS) analyzer using monochromatic light of Al
of cavitation and nanoparticles addition on both viscosity and Kα (1486.7 Ev, 13 kV, 100 W) with 100 and 30 eV energy steps for
elasticity of HO has never been investigated. To the best of our general and high-resolution spectrum; the steps for general and high-
knowledge, only Montes et al.33 showed some preliminary resolution spectra were 1 and 0.1 eV respectively. Moreover, the
results on the sole heavy oil viscosity reduction induced by the asphaltenes mean aggregates size was measured through the dynamic
use of nanoparticles (NiO-functionalized SiO2, SiNi1) coupled light scattering technique (DLS) using a NanoPlus-3 (Micromeritics,
with ultrasound cavitation. The authors investigated the effect GA, USA). Briefly, n-heptane/toluene (Heptol) model solutions were
prepared with 50 vol % of toluene and 1000 mg·L−1 of asphaltenes,
of both the cavitation time and the nanoparticles concentration the mixture was stirred at 300 rpm for the entire duration of the
concluding that after 90 min of exposure with 2000 mg·L−1 of experiments, and aliquots were taken for carrying out the measure-
SiNi1 nanoparticles the viscosity of the HO at a shear rate of ments. The same procedure was applied consequently for Heptol
10 s−1 was reduced 44%. This result is achieved due to the HO model solutions at the same conditions of sample D: first, the SiNi1
internal microstructure modification induced by cavitation nanoparticles were added to the solutions, and then the obtained
suspensions were sonicated for 90 min before measuring asphaltenes
coupled to the asphaltenes adsorption on the nanoparticles. aggregation.
The aim of this paper is assessing the preliminary results of 2.2.2. Samples Preparation. The colloidal suspensions made of
Montes et al.33 by evaluating the effect of the proposed HO and 2000 mg·L−1 of SiNi1 nanoparticles were prepared by hand
combined treatments on the rheological properties of the stirring for 15 min at 30 °C in order to facilitate mixing homogeneity
heavy oil. The steady state behavior and the dynamic response due to the sample high viscosity, which made unfeasible the use of
of the material are investigated before and after the other mechanical methods of mixing.
Ultrasound cavitation was applied for 90 min to the pure HO
functionalized nanoparticles-assisted ultrasound cavitation. (sample C) and to the colloidal suspension discussed above (sample
Also, the HO thixotropy is investigated through hysteresis D). An Elmasonic E60H (Elma Schmidbauer GmbH, Singen,
loop tests. Since rheology is a very sensitive tool to investigate Germany) sonicator with a power of 400 W and frequency of 37
the microstructure of complex fluids, like the heavy oils,34 it is kHz was used, and the samples were monitored to avoid the
expected to understand the microstructure modification temperature increase induced by ultrasound exposure.
All of the samples were left at rest for 10 min before carrying out
induced by cavitation combined with nanotechnology treat-
any further tests.
ments so that this study might open a broader landscape about 2.2.3. Rheological Measurements. Prior to any rheological test,
the impacts of nanotechnology in the production and before loading, the samples were preconditioned by hand mixing for 5
transportation of heavy oils. min to start from a microstructure which resembles the one

9672 DOI: 10.1021/acs.energyfuels.9b02288


Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

encountered in real field applications more than that obtained at the roughness.46 For gaps below 0.5 mm, the viscosity will be
end of a preconditioning steady simple shear flow, as typically done in consequently underestimated and the data might be postprocessed
rheometry. to correct the experimental values.46,47 However, the residual wall slip
First, a flow curve is obtained by running a shear rate ramp soon will affect all of the investigated samples in the same way,46 and
after having loaded the sample in the rheometer. The shear rate was consequently, the viscosity will be identically underestimated. Since
increased from ∼3 to ∼100 s−1 in 2 min to limit oil light tense the goal of the paper is only to compare the behavior of different
evaporation and to obtain a flow curve characterization in a very short samples, as long as the data are taken at the same gap, the comparison
time, as typically required in control operations in real field consistency will not be affected by the wall slip. Thus, for the sake of
applications. simplicity, we will report in the following the data as they are
To compare the different samples, the degree of viscosity reduction measured without a postprocessing.
(DVR) is estimated. It is defined as All tests were run at 25 °C and 1 atm. All experiments were carried
ij η − η yz
DVR = jjjj HO treat z
zz × 100
out by triplicate, so to ensure that the initial conditions obtained with

j zz
the manual mixing followed by the sample loading are actually
k {
η reproducible; the uncertainties are presented as error bars, and it is
HO (1)
worth mentioning that a maximum 3% error is recorded among the
where ηHO and ηtreat are the viscosities of the heavy oil (sample A) and different replicates in all of the rheological tests that we run. This
of the sample after a specific treatment (samples B−D), small error demonstrates the accuracy of the sample preparation and
respectively.36,37 The viscosities in eq 1 are measured at three of the measurement procedure as well as it guarantees the tests
characteristic shear rates during the increasing shear rate ramp. repeatability.
Hysteresis loop tests were performed to investigate the material
thixotropy on freshly loaded samples. The shear rate was first 3. RESULTS AND DISCUSSION
increased from ∼3 to ∼100 s−1 in 2 min, as in the flow curve The results reported in this study are mainly divided into four
determination, and then it was decreased from ∼100 to ∼3 s−1 always
in 2 min. In the increasing shear rate ramp, the sample viscosity sections: (I) asphaltenes chemical composition character-
change can be associated with the internal microstructure breakdown, ization and aggregation tests, (II) evaluation of the viscosity
while in the decreasing shear rate ramp, a microstructure buildup can reduction due to addition of nanoparticles, ultrasound
occur.15,38−40 Each whole experiment lasted about only 4 min. cavitation, and combination of the two, (III) preliminary
To characterize the sample hysteresis the area enclosed by the loop investigation of the thixotropy of the samples with a single
(WL) in a stress vs shear rate plot was evaluated.13 WL is the power hysteresis loop experiment, and (IV) assessment of the
per unit of volume dissipated by the sample in the loop test treatment effects by dynamical mechanical experiments.
γmin
̇ 3.1. Asphaltenes Chemical Composition Character-
WL = ∫γ ̇ (σBD − σBU)dγ ̇
(1a)
ization and Aggregation Tests. It has been mentioned
min
throughout this manuscript that the HO microstructure is
where σBD and σBU are the shear stresses in the increasing and formed by the crude oil heavy fractions such as asphaltenes and
decreasing shear rate ramp, respectively. γ̇min and γ̇max are the resins.48 In particular, the asphaltenes by means of their
minimum and maximum shear rate in the hysteresis loop test. The chemical structure including heteroatoms (oxygen, nitrogen
thixotropy index is also estimated. It is defined as the normalized area
sulfur)49,50 have a self-association trend fomented by different
enclosed by the loop in a viscosity vs shear rate plot
i BD − BU yz
thixotrophy index = jjj
mechanisms such as van der Waals forces, H-bonding, and
zz × 100
π−π stacking interactions,51 leading to aggregation of these
k BD { molecules, which causes formation of clusters.52
γ̇ It is well known that in light crude oils this behavior prompts
∫γ ̇ min (ηBD − ηBU)dγ ̇ the asphaltenes precipitation/deposition at certain pressure
min
= × 100
γ̇
∫γ ̇ min ηBD dγ ̇ conditions. However, in the HO cases, its microstructure is
min (2) usually configured in a viscoelastic network.49 The viscoelastic
where BD and BU represent the areas under the breakdown and network existence is driven by two particularly important
build-up curves, respectively, and ηBD and ηBU the viscosities during factors, a high amount of resins which stabilize the asphaltenes
the increasing and decreasing shear rate ramp, respectively.41 in the oil matrix, and the asphaltenes content that must be
Finally, the viscoelasticity of the samples was characterized by higher than a mass fraction of 5%.53,54 It can be said from the
dynamic experiments run on fresh samples. A frequency sweep test, asphaltenes content in the evaluated HO that the existence of a
from 0.63 to 250 rad/s, at a constant strain of 3% was performed. viscoelastic network is more than possible. Nevertheless,
Amplitude sweep tests allowed one to verify that with the chosen detailed analyses were carried out for obtaining some more
applied strain (3%) the samples are in the linear regime,33 and the insights over this matter.
torque of the rheometer is significantly higher than the lower limit of
the instrument.42,43 In Table S1 of the Supporting Information the chemical
All of the rheological measurements were carried out using a composition of the asphaltenes characterized through the XPS
Kinexus Pro+ (Malvern, UK) rheometer equipped with a Peltier cell technique can be observed. The presence of the mentioned
to control the temperature with a precision of 1 ×10−2 °C. With heteroatoms could be an indicator of asphaltenes−asphaltenes
multiphase fluids wall slip44 may occur, and to prevent or limit it a interaction forces.51 Hence, the asphaltenes clusters formation
serrated parallel plate−plate geometry with a diameter of 20 mm is was studied through aggregation kinetics and are shown in
used. It is equipped with a solvent trap to limit light tense Figure S1 of the Supporting Information, where an initial
evaporation, and a fixed gap of 0.3 mm is used for all of the aggregates size growth and a reduction until reaching a
experiments. Such a small gap was chosen to have the possibility to stabilized system after 230 min can be observed. This
exploit high shear rates. We checked the instrument alignment by
measuring the viscosity at different gaps with the serrated plates.45 We phenomenon is explained by the effect of the mixture shearing
found that the shear rate is independent of the gap for gaps larger than in the asphaltenes interactions and clusters configuration,
0.5 mm, while it decreases for lower gaps. The observed behavior is which is a similar behavior of that observed in the HO
due to an unavoidable intrinsic misalignment of the instrument and to microstructure breakdown after subjecting the fluid to
a residual wall slip induced by the flow through the geometry shearing.
9673 DOI: 10.1021/acs.energyfuels.9b02288
Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

Figure 1. Viscosity vs shear rate of the four samples of Table 1 as measured during the increasing shear rate ramp (BD). (A) Heavy oil (HO), (B)
HO with nanoparticles, (C) sonicated HO, and (D) HO with the addition nanoparticles and sonicated. Error bars are ± the standard deviation of
three sets of data. Lines are the predictions of Cross equation, eq 3, obtained with the best-fit parameters of Table 2. Gray regions highlight the data
taken to evaluate the degree of viscosity reduction (DVR).

Table 2. Best-Fit Parameters of the Cross Equation, Eq 3a


R2 η0 [Pa·s] η∞ [Pa·s] Α [s] m
sample A 0.999986 131.5 ± 5.5 26.4 ± 2.2 0.061 ± 0.004 1.29 ± 0.12
sample B 0.999996 113.5 ± 1.9 32.1 ± 0.6 0.067 ± 0.002 1.57 ± 0.06
sample C 0.999997 102.9 ± 1.4 30.7 ± 0.4 0.063 ± 0.002 1.67 ± 0.06
sample D 0.999931 70.4 ± 3.9 0 0.054 ± 0.007 0.96 ± 0.04
a
Uncertainties are the standard error of the estimated parameter, and R2 is the regression coefficient of determination.

Similar results were obtained previously,37 corroborating the plateau, moreover, in previous studies involving nanoparticles-
hypothesis that the HO microstructure is formed by heavy-oil systems it nicely interpolated the experimental data37
asphaltenes clusters configured into a viscoelastic network. η − η∞
On the other hand, in Figure S1 the asphaltenes aggregation η = η∞ + 0
kinetics in the presence of nanoparticles and ultrasound 1 + (αγ )̇ m (3)
irradiation is also seen. It is observed that both mechanisms where η is the viscosity, α a characteristic relaxation time above
have an important effect on the formation of the clusters as which the system shows a shear thinning behavior, γ̇ the shear
these have a lower size and a faster stabilization in comparison rate, m a constant, η0 the zero shear viscosity, and η∞ the
with the system without treatment. viscosity plateau at high shear rates. The predictions of the
From these results it can be deduced that application of Cross equation, eq 3, are also shown in Figure 1 as solid gray
ultrasound alongside nanoparticles would have an important lines passing through the data. The best-fit parameters are
effect over HO microstructure. In this sense, this phenomenon reported in Table 2 for the four samples, together with the
was studied by rheological tests for identifying changes in the coefficient of determination, R2. Notice that for sample D we
fluid rheological properties. imposed η∞ = 0, since the data do not show even only the
3.2. Viscosity Reduction Evaluation. Figure 1 shows the trend toward a faraway plateau at high shear rates.
viscosity of the four investigated samples (Table 1) vs the Data in Figure 1 clearly show that the sole addition of
shear rate as measured during the breakdown (BD) ramp, i.e., nanoparticles and the sole sonication are able to similarly
while increasing the shear rate. All the samples show a shear decrease the HO viscosity without however qualitatively
thinning behavior, where the viscosity decreases with the shear changing the rheological behavior of the treated samples
rate. This is probably due to a shear-induced microstructure with respect to the native HO. Indeed, a tendency toward both
reorganization, more than to a purely viscoelastic effect. The the zero-shear and the high-shear plateau can be guessed from
viscosities of the four samples seem to tend to a zero-shear the data of samples A, B, and C, and their shear thinning
plateau as the shear rate is decreased, though the plateau value behavior is very similar, with the coefficient m (Table 2) of the
cannot be clearly estimated from the data at hand. A high shear Cross equation being almost the same. Conversely, when both
plateau seems to be approached by samples A, B, and C but functionalized nanoparticles addition and ultrasound cavitation
not by sample D. Thus, from a practical and engineering point treatment are applied to the HO a significant rheological
of view, the data can be interpolated with the Cross model,35 difference is observed: First, the viscosity is more than halved
eq 3, which accounts for both a zero-shear and a high-shear at each shear rate; the tendency toward a high-shear plateau is
9674 DOI: 10.1021/acs.energyfuels.9b02288
Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

Figure 2. Viscosity (histogram) and DVR (line−symbol plot) calculated at three different shear rates of the four samples of Table 1: (A) heavy oil
(HO), (B) HO with nanoparticles, (C) sonicated HO, and (D) HO with addition of nanoparticles and sonicated. Error bars are the standard
deviation of three sets of data.

not observable anymore; last but not least, the shear thinning can be mainly due to the ultrasound-induced weakening of the
behavior is changed and a terminal behavior with power law asphaltenes−asphaltenes interactions,22 causing a size reduc-
index n = 0.04 (eq 4) can be identified tion of their aggregates.64
For the employed nanoparticles concentration and ultra-
η = kγ ṅ − 1 (4) sound irradiation time conditions, it is observed that the sole
where k is the so-called consistency index. Such a pronounced ultrasound cavitation seems to be more effective than the sole
shear thinning resembles very much that of suspensions.55 addition of nanoparticles, which is in accordance with the
Notice that by comparing eq 4, first proposed by the Ostwald− results obtained in a previous study.33 Nevertheless, it is worth
de Waele56−58 and valid for power law fluids,59,60 with eq 3 mentioning that in these preceding results it was obtained that
with η∞ = 0 it is readily obtained that n = 1 − m and k = η0α−m. a dosage of 1000 mg·L−1 of nanoparticles had a better
To better quantify the viscosity reduction induced by each performance in viscosity reduction than the sole ultrasound
treatment and by the combination of the two, the viscosity irradiation, while by using 2000 mg·L−1 of SiNi1 with 90 min
reduction index, DVR, eq 1, is calculated at the three shear of ultrasound irradiation the process is enhanced due to a
rates highlighted with the gray shadows in Figure 1. The lowest better synergy between both mechanisms.
value, 3.7 s−1, is in the region where the trend toward the zero- Remarkably, the viscosity reduction measured at the highest
shear plateau can be observed, the highest value, 70.3 s−1, is in shear rate in both cases, the sole nanoparticles addition and the
the region where samples A, B, and C are approaching the sole ultrasound irradiation, is significantly smaller than those
high-shear plateau, and finally the intermediate value, 19.7 s−1, measured at smaller shear rates. This is due to the fact that the
is very close to the value of 20 s−1 at which the viscosity is high-shear rate plateau of the HO seems to be practically
conventionally measured to characterize the fluid rheology in unaffected by the single treatment, see Figure 1 and Table 2.
pipeline transportation.8 Figure 2 shows the viscosities and The dispersing action exerted by the flow results thus
DVR of the four samples. comparable to those induced either by the nanoparticles or
It can be seen that by adding the functionalized nano- by the ultrasound cavitation. The two treatments actually
particles to the HO (sample B) a viscosity reduction of about promote the asphaltene macroclusters dispersion or the
11% is achieved at the two lowest shear rates and a significantly
asphaltene network disruption, obtained mechanically by the
smaller reduction of 2.7% at the highest shear rate. This can be
flow, to lower shear rates.
attributed to the high asphaltenes selectivity and adsorption
Finally, the combination of the two treatments, i.e., the
capability of the used nanoparticles23,24 that were both
enhanced by assembling active sites of nickel oxide over the application of ultrasound cavitation in the presence of the
support surface.32,61,62 According to several studies the functionalized nanoparticles, significantly enhances the vis-
adsorption phenomenon is identified as a potential first cosity reduction at any shear rate obtaining a DVR of 50% at
mechanism of HO viscosity reduction,19,37,63 as it can modify 3.7 s−1, 53% at 19.7 s−1, and 62.5% at 20.3 s−1. As discussed
the internal fluid microstructure,19 disrupting the asphaltene above, the rheology is significantly modified as the high-shear
viscoelastic network or reducing the dimensions of the plateau is not observable anymore, and this leads to the very
asphaltene macroclusters,4 which are the primary causes of significant viscosity reduction gained at the highest shear rate.
the high viscosity of heavy oils. This can be explained considering that a much finer
Similarly, the application of ultrasound cavitation to the microstructure reorganization is induced by the coupled
heavy oil (sample C) induces a viscosity reduction of about treatments and that the obtained fine microstructure is more
18% at γ̇ = 3.7 s−l, 16% at 19.7 s−1, and 8.3% at 70.3 s−1. This stable than that of samples B and C.
9675 DOI: 10.1021/acs.energyfuels.9b02288
Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

Figure 3. Hysteresis loop tests: (a) viscosity vs shear rate; (b) shear stress vs shear rate. Sample A: Heavy oil (HO). Sample D: HO with the
addition of nanoparticles and sonication. Full symbols refer to the increasing shear rate ramps or breakdown (BD) curves and hollow symbols to
the decreasing shear rate ones or build-up (BU) curves. BD viscosity curves are the same in Figure 1. Error bars are ± the standard deviations of
three sets of data.

Although the synergistic effects of the two treatments are shows a thixotropy behavior more pronounced than that of the
now proven, further investigation of the rheological changes same oil once treated with both the addition of nanoparticles
induced by the treatments to assess the mobility and the flow and ultrasound cavitation. Moreover, it can be noticed that the
capacity improvements of the crude oil and to understand the first points of the BU curve of sample A (at the largest shear
mechanisms underneath these improvements is needed. In the rates) are slightly larger than the last points of the BD curve,
following, for the sake of brevity, we will focus only on samples because once at rest the microstructure starts to rebuildup, as it
A and D, which show the most relevant rheological differences. always happens in a thixotropic material. The same is not
3.3. Hysteresis Loop Tests. Crude oils can often show a observed for sample D. Besides this, it is worthwhile noticing
thixotropic behavior, and a fast way to investigate it is to run that the BU viscosity curve of sample A shows essentially a
hysteresis loop tests.10,18 The hysteresis loop may highlight the Newtonian behavior. This can be explained assuming that the
difference between the dynamics of the material internal flow breaks down the oil microstructure that is not
microstructure breakdown and the build-up in response to an reconstituted in the BU curve while rapidly decreasing, in
imposed increasing shear rate and a decreasing one, only 2 min, the shear rate from 100 to 3 s−1. If the
respectively.15,38 microstructure remains unaltered during the BU curve,
Results are shown in Figure 3, where the hysteresis loop accordingly the viscosity does not vary with the shear rate.
areas are shadowed. In Figure 3a the viscosity of the BD and Conversely, in sample D the HO original microstructure was
BU curves are shown for both samples A and D, while in already broken down by the synergistic action of asphaltene
Figure 3b the stress vs shear rate is plotted. In the last case, the adsorption on the nanoparticles and ultrasound cavitation, and
shadowed area represents WL (eq 1a), i.e., the power per unit thus, the initial configuration is characterized by a more
of volume dissipated by the sample in the loop test. It is clear dispersed system. It is probable that the shear thinning
that WL of HO is much larger than that of sample D. To behavior, in this case, is not due to a further breakdown of the
quantify the observed difference, WL is calculated by first microstructure but by a geometrical reorganization of it, which
smoothing the data with a cubic spline and then by numerically is reverted when the shear rate is decreased in the BU curve.
evaluating the integrals. The results are summarized in Table 3, 3.4. Viscoelastic Moduli. To further investigate if the
different rheological behavior of samples A and D is due to a
Table 3. Hysteresis Loop Parameters of the Untreated change of microstructure, the dynamic moduli of the two
Heavy Oil (Sample A) and of the Heavy Oil with the samples are measured. The elastic modulus G′, also known as
Addition of Nanoparticles and after Sonication (Sample D)a the storage modulus, is the modulus of the stress component in
phase with the strain, while the viscous modulus G′′, also
WL [Pa s−1] thixotropy index [%]
known as loss modulus, is the modulus of the stress
sample A 2.9 × 105 28.3 component in phase with the shear rate. Hence, G′ accounts
sample D 8.3 × 104 15.8 for the elastic response of the material and thus for the elastic
a
WL is the dissipated power per unit of volume; thixotropy index is energy stored during the deformation, while G′′ accounts for
defined in eq 2. the viscous, plastic, response of the fluid and thus for the
energy dissipated during the deformation.
and it is found that the dissipated power per unit volume by In Figure 4 G′ and G′′ are plotted vs the imposed frequency.
sample D amounts to about only the 29% of that dissipated by A clear fluid behavior, where G′′ is significantly larger than G′,
sample A. Also, the thixotropy index, eq 2, is calculated by is exhibited by both samples, though the low-frequency limit of
numerically evaluating the viscosity integrals after having linear viscoelasticity, where G′′ is proportional to the
smoothed the data with a cubic spline, and the results are frequency and G′ to the frequency to the power of two, is
shown in Table 3. not reached in the investigated frequencies. The loss moduli of
The large difference between both WL and the thixotropy the two samples almost coincide, while small differences are
index of the two samples suggests that the original heavy oil observable in G′ moduli where that of sample D is slightly
9676 DOI: 10.1021/acs.energyfuels.9b02288
Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

correspond to low shear rates, in the region approaching the


zero-shear plateau. In this region sample D appears to be less
viscous than sample A, in agreement with the viscosity data
discussed in section 3.1.
The phase angles of the two samples are measurably
different, and that of sample A is always smaller than that of
sample D, confirming that the latter sample is less elastic than
the former one. The phase angles are always larger than 45°
because G′′ is always larger than G′ and, as discussed above,
the materials exhibit a fluid behavior.
3.5. Relaxation Spectra. A viscoelastic fluid has a
behavior in between that of a purely viscous fluid and a purely
elastic material; this implies that this material can dissipate part
of the energy stored during the deformation with different
relaxation mechanisms, each of them with a characteristic
Figure 4. G′ (full symbols) and G′′ (hollow symbols) vs frequency. relaxation time. Thus, to better analyze the different elasticity
Circles are for sample A: Heavy oil (HO). Diamonds for Sample D: of the two samples, their relaxation spectra are here calculated.
HO with addition of nanoparticles and sonicated. The relaxation spectrum allows identifying the characteristic
relaxation times of the sample if the last ones lie within the
smaller than that of sample A at any frequency. This suggests investigated frequencies.
that the differences between the two samples are mainly related The relaxation spectra are calculated from the elastic
to their different elasticity, more than to their different viscous modulus, G′, of Figure 4 with the approximated equation, eq
behavior. 5, of second order proposed by Tschoegl66 and also from the
To better highlight the differences between the two samples, loss modulus G′′ with the approximated equation, eq 6, of the
in Figure 5 both the complex viscosity (Figure 5a) and the second order of Schwarzl and Staverman66
1 2
H(t ) = dG′/d ln ω − d G′/d(lnω)2 |1/ ω= t / 2
2 (5)

H(t ) = (2/π )[G″ − d2G″/d(ln ω)2 ]1/ ω= t (6)

where H(t) is the relaxation spectrum, ω is the frequency (rad


s−1), and t is the time (s). To calculate the spectra from the
moduli, data of Figure 4 are first smoothed with a third-order
Bezier polynomial, so to get rid of false results that may come
from the scatter of the data, and eqs 5 and 6 are then applied.
The results are shown in Figure 6, where solid lines refer to
sample A and dashed ones to sample D. Moreover, black lines
are for the spectra calculated from G′, eq 5, and red lines are

Figure 5. (A) Complex viscosity vs frequency and (B) phase angle vs


frequency of sample A (full symbols) and sample D (hollow symbols).

phase angle (Figure 5b) are plotted vs the frequency. These


data can be calculated from the moduli: In materials like
polymer melts or polymer solution, the complex viscosity can
be directly compared to the viscosity while the phase angle Figure 6. Relaxation spectra vs time estimated from the elastic
highlights the relative contribution of the elastic behavior with modulus G′ (black lines) and from the loss modulus G′′ (red lines) of
respect to the viscous one, being 90° for purely viscous Figure 4. Spectra of the untreated HO, sample A, are plotted as solid
materials and 0° for purely elastic ones.65 lines, while those of the HO in the presence of 2000 mg·L−1 of SiNi1
The complex viscosity of the two samples is comparable, and after 90 min of ultrasound cavitation, sample D, are plotted as dashed
measurable differences are observed at low frequencies, which lines.

9677 DOI: 10.1021/acs.energyfuels.9b02288


Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

used for the spectra calculated from G′′, eq 6. The reconstitute itself very quickly. Furthermore, the dynamic data
characteristic relaxation times are identified as the abscissas in the linear regime show that sample D is less elastic than the
of the relative maxima of the spectra. Sample A shows a single HO, and this is corroborated by the estimates of their
relaxation time equal to 1 × 10−2 s if estimated from the relaxation spectra that highlight the existence of a relaxation
spectrum calculated from G′ (black solid line) and equal to 8 × time only for the HO, sample A. This last result implies that
10−3 s if estimated from the spectrum calculated from G′′ (red the microstructure of HO is either a viscoelastic network of
solid line). Conversely, both relaxation spectra of sample D are asphaltenes or it is made of large asphaltenes clusters that can
monotonous, and thus, no relaxation time is identifiable. This elastically deform; conversely, sample D is probably a
implies that if a relaxation time of sample D exists, it must lie suspension-like one.
outside the time range here exploited with the dynamic The combined treatments consisting of the addition of
moduli, and thus, it must be shorter than 4 × 10−3 s. nanoparticles with the subsequent ultrasound cavitation of the
The relaxation spectrum is a material function; thus, in resulting suspension is then able to break the original HO
principle, that calculated with eq 5 should coincide with that microstructure. We hypothesize that the asphaltenes stably
calculated with eq 6. This seldom happens because of two adsorb on the nanoparticles and that the size reduction of the
reasons; first, the two equations are approximated; second, the asphaltenes is clearly enhanced by the ultrasound irradiation.
data are affected by an unavoidable experimental error that Although further studies are needed to better investigate the
propagates into the calculated spectra. However, the results mechanisms involving the heavy oil microstructure modifica-
that we obtained are quite robust because the relaxation times tion, this paper not only proves the suitability of the
of sample A estimated with the two different procedures are in application of this technology to increase the HO flow
acceptable agreement, and both relaxation spectra calculated capacity but also demonstrates its enhanced performance by
for sample D are monotonous. possibly providing a better and more efficient HO production
The existence of a relaxation time is in accordance with the and transportation design.
existence of a 3D microstructure formed by the aggregates of
asphaltenes that causes the viscoelastic behavior of the HO,
which then shows a more pronounced elasticity and a larger

*
ASSOCIATED CONTENT
S Supporting Information
viscosity with respect to the treated sample D. Sample D, in The Supporting Information is available free of charge on the
agreement, shows the absence of a relaxation time, at least in ACS Publications website at DOI: 10.1021/acs.energy-
the time range corresponding to the frequencies at which the fuels.9b02288.
experiments were carried out. This suggests that the original Chemical composition characterized through the XPS
viscoelastic 3D network of the HO is disrupted by addition of technique for the asphaltenes obtained from the
the functionalized nanoparticles coupled to the ultrasound evaluated crude oil; mean aggregate size measured
cavitation. through the dynamic light scattering (DLS) technique at
These results agree with those reported by several authors in 298 K using a heptol 50 solution with the asphaltenes
the literature, e.g., Dickinson et al.,67 Soo and Woo,68 and obtained from the evaluated crude oil (PDF)
Behzadfar et al.42 found that bitumens and heavy oils have
characteristic relaxation times associated to the presence of a
3D internal microstructure. ■ AUTHOR INFORMATION
Corresponding Authors
4. CONCLUSIONS *E-mail: mario.minale@unicampania.it.
A novel process involving ultrasound cavitation assisted by *E-mail: caafrancoar@unal.edu.co.
nickel oxide nanoparticles functionalized over nanoparticulated ORCID
silica support was applied to increase the heavy oil mobility at Mario Minale: 0000-0002-7756-3536
reservoir conditions and/or to improve the heavy oil flow Farid B. Cortés: 0000-0003-1207-3859
capacity at surface and transportation conditions. The adopted Camilo A. Franco: 0000-0002-6886-8338
method led to a viscosity reduction ranging from 50% to more Notes
than 63% by increasing the applied shear rate. We proved that The authors declare no competing financial interest.


the sole addition of nanoparticles or ultrasound cavitation
irradiation is not capable of modifying the HO rheology ACKNOWLEDGMENTS
significantly. A small viscosity reduction is achieved in both
The authors acknowledge COLCIENCIAS and Agencia
cases, while the combination of the two treatments not only
Nacional de Hidrocarburos (ANH-Colombia) for their
reduces the viscosity much more than every single treatment
support provided in Agreement 272 of 2017. We also
but also modifies qualitatively the HO flow curve. Indeed, the
recognize the Universidad Nacional de Colombia for logistical
high shear plateau, which starts to be approached by the HO
and financial support.


for shear rates larger than 70 s−1, disappears for sample D,
where the nanoparticles are added to the HO, and the obtained REFERENCES
suspension is subsequently irradiated. Moreover, the response
of the two samples to a hysteresis loop test is also significantly (1) Alboudwarej, H.; Felix, J.; Taylor, S.; Badry, R.; Bremner, C.;
Brough, B.; Skeates, C.; Baker, A.; Palmer, D.; Pattison, K. La
different as the HO has a power per unit volume dissipated importancia del petróleo pesado. Oilfield Rev. 2006, 18, 38−59.
during the test 3.5 times larger than that of sample D. This last (2) Owen, N. A.; Inderwildi, O. R.; King, D. A. The status of
result indicates that the original HO microstructure once conventional world oil reservesHype or cause for concern? Energy
broken down by the flow needs a long time to rebuildup; Policy 2010, 38 (8), 4743−4749.
conversely, the microstructure of sample D is so dispersed that (3) Urquhart, R., Heavy oil transportation-present and future. J. Can.
either it cannot be broken down by the flow or it is able to Pet. Technol. 1986, 25, (02). DOI: 10.2118/86-02-05

9678 DOI: 10.1021/acs.energyfuels.9b02288


Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

(4) Minale, M.; Merola, M. C.; Carotenuto, C. Effect of solvents on (27) Askarian, M.; Vatani, A.; Edalat, M. Heavy oil upgrading in a
the microstructure aggregation of a heavy crude oil. Fuel Process. hydrodynamic cavitation system: CFD modelling, effect of the
Technol. 2018, 177, 299−308. presence of hydrogen donor and metal nanoparticles. Can. J. Chem.
(5) Pierre, C.; Barré, L.; Pina, A.; Moan, M. Composition and heavy Eng. 2017, 95, 670−679.
oil rheology. Oil Gas Sci. Technol. 2004, 59 (5), 489−501. (28) Cardona, L.; Arias-Madrid, D.; Cortés, F. B.; Lopera, S. H.;
(6) Martinez-Palou, R.; Mosqueira, M. d. L.; Zapata-Rendon, B.; Franco, C. A. Heavy Oil Upgrading and Enhanced Recovery in a
Mar-Juarez, E.; Bernal-Huicochea, C.; de la Cruz Clavel-Lopez, J.; Steam Injection Process Assisted by NiO-and PdO-Functionalized
Aburto, J. Transportation of heavy and extra-heavy crude oil by SiO2 Nanoparticulated Catalysts. Catalysts 2018, 8 (4), 132.
pipeline: A review. J. Pet. Sci. Eng. 2011, 75 (3-4), 274−282. (29) Franco, C. A.; Montoya, T.; Nassar, N. N.; Cortés, F. B. NiO
(7) Gateau, P.; Hénaut, I.; Barré, L.; Argillier, J. Heavy oil dilution. and PdO Supported on Fumed Silica Nanoparticles for Adsorption
Oil Gas Sci. Technol. 2004, 59 (5), 503−509. and Catalytic Steam Gasification of Colombian C7 Asphaltenes.
(8) Hart, A. A review of technologies for transporting heavy crude Handbook on Oil Production Research, 1st ed.; Ambrosio, J., Ed.;
oil and bitumen via pipelines. J. Pet. Explor. Prod. Technol. 2014, 4 (3), NOVA SCIENCE, 2014; Chapter 3, pp 101−146.
327−336. (30) Franco, C. A.; Montoya, T.; Nassar, N. N.; Pereira-Almao, P.;
(9) Hasan, M. D. A.; Shaw, J. M. Rheology of Reconstituted Crude Cortés, F. B. Adsorption and subsequent oxidation of colombian
Oils: Artifacts and Asphaltenes. Energy Fuels 2010, 24 (12), 6417− asphaltenes onto Nickel and/or Palladium oxide supported on fumed
6427. silica nanoparticles. Energy Fuels 2013, 27 (12), 7336−7347.
(10) Mortazavi-Manesh, S.; Shaw, J. M. Thixotropic rheological (31) Cortés, F. B.; Mejía, J. M.; Ruiz, M. A.; Benjumea, P.; Riffel, D.
behavior of Maya crude oil. Energy Fuels 2014, 28 (2), 972−979. B. Sorption of asphaltenes onto nanoparticles of nickel oxide
(11) Visintin, R. F.; Lapasin, R.; Vignati, E.; D’Antona, P.; Lockhart, supported on nanoparticulated silica gel. Energy Fuels 2012, 26 (3),
T. P. Rheological behavior and structural interpretation of waxy crude 1725−1730.
oil gels. Langmuir 2005, 21 (14), 6240−6249. (32) Franco, C. A.; Nassar, N. N.; Ruiz, M. A.; Pereira-Almao, P.;
(12) Wardhaugh, L.; Boger, D. The measurement and description of Cortés, F. B. Nanoparticles for inhibition of asphaltenes damage:
the yielding behavior of waxy crude oil. J. Rheol. 1991, 35 (6), 1121− adsorption study and displacement test on porous media. Energy Fuels
1156. 2013, 27 (6), 2899−2907.
(13) Mewis, J.; Wagner, N. J. Thixotropy. Adv. Colloid Interface Sci. (33) Pinzón, D. M. Rheological Demonstration of Heavy Oil
2009, 147−148, 214−227. Viscosity Reduction by NiO/SiO2 Nanoparticles-Assisted Ultrasound
(14) Green, H.; Weltmann, R. Analysis of Thixotropy of Pigment- Cavitation. SPE Annual Technical Conference and Exhibition; Society of
Vehicle Suspensions - Basic Principles of the Hysteresis Loop. Ind. Petroleum Engineers, 2018.
Eng. Chem., Anal. Ed. 1943, 15 (3), 201−206. (34) Merola, M. C.; Carotenuto, C.; Gargiulo, V.; Stanzione, F.;
(15) Cheng, D. C.; Evans, F. Phenomenological characterization of Ciajolo, A.; Minale, M. Chemical-physical analysis of rheologically
the rheological behaviour of inelastic reversible thixotropic and different samples of a heavy crude oil. Fuel Process. Technol. 2016, 148,
antithixotropic fluids. Br. J. Appl. Phys. 1965, 16 (11), 1599. 236−247.
(16) Ghannam, M. T.; Hasan, S. W.; Abu-Jdayil, B.; Esmail, N. (35) Institute, E. IP 469: Determination of saturated, aromatic and
Rheological properties of heavy & light crude oil mixtures for polar compounds in petroleum products by thin layer chromatography
improving flowability. J. Pet. Sci. Eng. 2012, 81, 122−128. and flame ionization detection; Energy Institute Publications, 2006.
(17) Meyers, M. A.; Chawla, K. K. Mechanical Behavior of Materials; (36) Aristizábal-Fontal, J. E.; Cortés, F. B.; Franco, C. A. Viscosity
Cambridge University Press: New York, 2008. reduction of extra heavy crude oil by magnetite nanoparticle-based
(18) Mortazavi-Manesh, S.; Shaw, J. M. Effect of diluents on the ferrofluids; Adsorption Science & Technology, 2017.
rheological properties of Maya crude oil. Energy Fuels 2016, 30 (2), (37) Taborda, E. A.; Franco, C. A.; Lopera, S. H.; Alvarado, V.;
766−772. Cortés, F. B. Effect of nanoparticles/nanofluids on the rheology of
(19) Taborda, E. A.; Alvarado, V.; Franco, C. A.; Cortés, F. B. heavy crude oil and its mobility on porous media at reservoir
Rheological demonstration of alteration in the heavy crude oil fluid conditions. Fuel 2016, 184, 222−232.
structure upon addition of nanoparticles. Fuel 2017, 189, 322−333. (38) Hahn, S. J.; Ree, T.; Eyring, H. Flow mechanism of thixotropic
(20) Kaushik, P.; Kumar, A.; Bhaskar, T.; Sharma, Y.; Tandon, D.; substances. Ind. Eng. Chem. 1959, 51 (7), 856−857.
Goyal, H. Ultrasound cavitation technique for up-gradation of vacuum (39) Moore, F. The rheology of ceramic slip and bodies. Trans. Br.
residue. Fuel Process. Technol. 2012, 93 (1), 73−77. Ceram. Soc. 1959, 58, 470−492.
(21) Plesset, M. S.; Prosperetti, A. Bubble dynamics and cavitation. (40) Peter, S. Zur Theorie der Rheopexie. Rheol. Acta 1964, 3 (3),
Annu. Rev. Fluid Mech. 1977, 9 (1), 145−185. 178−180.
(22) Rahimi, M. A.; Ramazani SA, A.; Alijani Alijanvand, H.; (41) Benchabane, A.; Bekkour, K. Rheological properties of
Ghazanfari, M. H.; Ghanavati, M. Effect of ultrasonic irradiation carboxymethyl cellulose (CMC) solutions. Colloid Polym. Sci. 2008,
treatment on rheological behaviour of extra heavy crude oil: A 286 (10), 1173.
solution method for transportation improvement. Can. J. Chem. Eng. (42) Behzadfar, E.; Hatzikiriakos, S. G. Viscoelastic properties and
2017, 95 (1), 83−91. constitutive modelling of bitumen. Fuel 2013, 108, 391−399.
(23) Betancur, S.; Carmona, J. C.; Nassar, N. N.; Franco, C. A.; (43) Dimitriou, C. J.; McKinley, G. H. A comprehensive constitutive
Cortés, F. B. Role of particle size and surface acidity of silica gel law for waxy crude oil: a thixotropic yield stress fluid. Soft Matter
nanoparticles in inhibition of formation damage by asphaltene in oil 2014, 10 (35), 6619−6644.
reservoirs. Ind. Eng. Chem. Res. 2016, 55 (21), 6122−6132. (44) Carotenuto, C.; Vananroye, A.; Vermant, J.; Minale, M.
(24) Franco, C. A.; Lozano, M. M.; Acevedo, S.; Nassar, N. N.; Predicting the apparent wall slip when using roughened geometries: A
Cortés, F. B. Effects of resin I on asphaltene adsorption onto porous medium approach. J. Rheol. 2015, 59 (5), 1131−1149.
nanoparticles: a novel method for obtaining asphaltenes/resin (45) Carotenuto, C.; Marinello, F.; Minale, M. A new experimental
isotherms. Energy Fuels 2016, 30 (1), 264−272. technique to study the flow in a porous layer via rheological tests. AIP
(25) Guzmán, J. D.; Betancur, S.; Carrasco-Marín, F.; Franco, C. A.; Conf. Proc. 2012, 1453 (1), 29−34.
Nassar, N. N.; Cortés, F. B. Importance of the adsorption method (46) Carotenuto, C.; Minale, M. On the use of rough geometries in
used for obtaining the nanoparticle dosage for asphaltene-related rheometry. J. Non-Newtonian Fluid Mech. 2013, 198, 39−47.
treatments. Energy Fuels 2016, 30 (3), 2052−2059. (47) Paduano, L. P.; Schweizer, T.; Carotenuto, C.; Vermant, J.;
(26) Franco, C. A.; Zabala, R.; Cortés, F. B. Nanotechnology applied Minale, M. Rough geometries with viscoelastic Boger fluids:
to the enhancement of oil and gas productivity and recovery of Predicting the apparent wall slip with a porous medium approach. J.
Colombian fields. J. Pet. Sci. Eng. 2017, 157, 39−55. Rheol. 2019, 63 (4), 569−582.

9679 DOI: 10.1021/acs.energyfuels.9b02288


Energy Fuels 2019, 33, 9671−9680
Energy & Fuels Article

(48) Mullins, O. C. The asphaltenes. Annu. Rev. Anal. Chem. 2011, 4, (68) Soo Cho, K.; Woo Park, G. Fixed-point iteration for relaxation
393−418. spectrum from dynamic mechanical data. J. Rheol. 2013, 57 (2), 647−
(49) Mullins, O. C.; Sabbah, H.; Eyssautier, J.; Pomerantz, A. E.; 678.
Barré, L.; Andrews, A. B.; Ruiz-Morales, Y.; Mostowfi, F.; McFarlane,
R.; Goual, L. Advances in asphaltene science and the Yen-Mullins
model. Energy Fuels 2012, 26 (7), 3986−4003.
(50) Yen, T.; Chilingarian, G. Asphaltenes and asphalts. Develop-
ments in Petroleum Science; Elsevier Science BV: The Netherlands,
1994.
(51) Duong, A.; Smith, K. J. A model of ceramic membrane fouling
during heavy oil ultrafiltration. Can. J. Chem. Eng. 1997, 75 (6),
1122−1129.
(52) Mullins, O. C.; Andrews, B.; Pomerantz, A.; Dong, C.; Zuo, J.
Y.; Pfeiffer, T.; Latifzai, A. S.; Elshahawi, H.; Barre, L. E.; Larter, S.
Impact of Asphaltene nanoscience on understanding oilfield
reservoirs. SPE Annual Technical Conference and Exhibition; Society
of Petroleum Engineers, 2011.
(53) Mullins, O. C.; Betancourt, S. S.; Cribbs, M. E.; Dubost, F. X.;
Creek, J. L.; Andrews, A. B.; Venkataramanan, L. The colloidal
structure of crude oil and the structure of oil reservoirs. Energy Fuels
2007, 21 (5), 2785−2794.
(54) 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; Springer, 2007;
pp 439−468.
(55) Carotenuto, C.; Merola, M. C.; Á lvarez-Romero, M.; Coppola,
E.; Minale, M. Rheology of natural slurries involved in a rapid
mudflow with different soil organic carbon content. Colloids Surf., A
2015, 466, 57−65.
(56) Barnes, H. A.; Hutton, J. F.; Walters, K. An introduction to
rheology; Elsevier, 1989.
(57) Mitschka, P. Nicht-Newtonsche Flüssigkeiten II. Drehströ-
mungen Ostwald-de Waelescher Nicht-Newtonscher Flüssigkeiten.
Collect. Czech. Chem. Commun. 1964, 29 (12), 2892−2905.
(58) Mooney, M.; Black, S. A generalized fluidity power law and
laws of extrusion. J. Colloid Sci. 1952, 7 (3), 204−217.
(59) Barnes, H. A. A handbook of elementary rheology; University of
Wales, Institute of Non-Newtonian Fluid Mechanics, 2000.
(60) Wan Nik, W.B.; Ani, F.N.; Masjuki, H.H.; Eng Giap, S.G.
Rheology of bio-edible oils according to several rheological models
and its potential as hydraulic fluid. Ind. Crops Prod. 2005, 22 (3),
249−255.
(61) Lozano, M. M.; Franco, C. A.; Acevedo, S. A.; Nassar, N. N.;
Cortés, F. B. Effects of resin I on the catalytic oxidation of n-C 7
asphaltenes in the presence of silica-based nanoparticles. RSC Adv.
2016, 6 (78), 74630−74642.
(62) Montoya, T.; Argel, B. L.; Nassar, N. N.; Franco, C. A.; Cortés,
F. B. Kinetics and mechanisms of the catalytic thermal cracking of
asphaltenes adsorbed on supported nanoparticles. Pet. Sci. 2016, 13
(3), 561−571.
(63) Taborda, E. A.; Franco, C. A.; Ruiz, M. A.; Alvarado, V.; Cortés,
F. B. Experimental and theoretical study of viscosity reduction in
heavy crude oils by addition of nanoparticles. Energy Fuels 2017, 31
(2), 1329−1338.
(64) Hemmati-Sarapardeh, A.; Dabir, B.; Ahmadi, M.; Mohammadi,
A. H.; Husein, M. M. Toward mechanistic understanding of
asphaltene aggregation behavior in toluene: The roles of asphaltene
structure, aging time, temperature, and ultrasonic radiation. J. Mol.
Liq. 2018, 264, 410−424.
(65) Rosa, M.; Grassia, L.; D’Amore, A.; Carotenuto, C.; Minale, M.
Rheology and mechanics of polyether(ether)ketone - Polyetherimide
blends for composites in aeronautics. AIP Conf. Proc. 2016, 1736 (1),
020177.
(66) Ferry, J. D. Viscoelastic properties of polymers, 3rd ed.; John
Wiley & Sons, 1980.
(67) Dickinson, E.; Witt, H. The dynamic shear modulus of paving
asphalts as a function of frequency. Trans. Soc. Rheol. 1974, 18 (4),
591−606.

9680 DOI: 10.1021/acs.energyfuels.9b02288


Energy Fuels 2019, 33, 9671−9680

You might also like