You are on page 1of 37

Page 1 of 37 Industrial & Engineering Chemistry Research

1
2
3 Development of a Population Balance Model to Describe the Influence of Shear and
4
5
6 Nanoparticles on the Aggregation and Fragmentation of Asphaltene Aggregates
7
8
9 Nashaat N. Nassar,†,* Stefania Betancur, ‡ Sócrates Acevedo,§ Camilo A. Franco,‡
10
11 Farid B. Cortés‡,*
12
13 †
14
Department of Chemical and Petroleum Engineering, University of Calgary, 2500
15 University Drive NW, Calgary, Alberta, Canada.
16

17 Grupo de Investigación en Yacimientos de Hidrocarburos, Facultad de Minas,
18 Universidad Nacional de Colombia Sede Medellín, Kra 80 No. 65–223, Medellín,
19 Colombia.
20
21 §
22 Escuela de Química, Facultad de Ciencias, Universidad Central de Venezuela, Caracas
23 1041-A, Venezuela.
24
25 *nassar@ucalgary.ca
26
27 *fbcortes@unal.edu.co
28
29
30
31 ABSTRACT
32
33 The precipitation and deposition of asphaltenes is a primary problem related to the
34 processing, transportation and production of oil. Flocculation of asphaltene aggregates is
35 likely to occur during the production and processing of crude oil. Recently, it has been
36 shown that nanotechnology in the form of nanoparticles is useful for the inhibition or
37
38
prevention of asphaltene formation damage. Although it is well known that the adsorption
39 of asphaltenes on the nanoparticle surface would reduce the capacity of these asphaltic
40 compounds to interact with each other, limited studies have been performed regarding the
41 processes and the mechanisms associated with the effect of nanoparticles on the inhibition
42 of the formation damage due to asphaltenes. To better understand this phenomenon from a
43
mathematical approach, a population balance model (PBM) is proposed to describe the
44
45 kinetics of asphaltene flocculation-fragmentation in the presence of nanoparticles. The
46 model assumes that asphaltenes in the presence of a shear rate are related to the aggregation
47 and fragmentation phenomena and includes a term related to the asphaltene adsorption on
48 nanoparticles. An adsorption kinetic term was introduced into the model using the double
49 exponential model. Experimental data of the kinetics of asphaltene aggregation were
50
51
obtained by Dynamic Light Scattering (DLS) measurements at a fixed initial asphaltene
52 concentration of 1000 mg/L and with different Heptol mixtures. In this study, commercial
53 silica, γ-alumina and magnetite nanoparticles were used as adsorbents to study the effect of
54 the chemical nature of the nanoparticles on the inhibition of the asphaltene growth and for
55 model validation. Additionally, to demonstrate the versatility of the proposed model, the
56
effect of asphaltene was also evaluated. The obtained results from the proposed population
57
58 balance model agree well with the experimental data, within an RSME% < 9%.
59
60
1
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 2 of 37

1
2
3 1. INTRODUCTION
4
5
6 The continuous decline in the recovery of crude oil and the increasing oil demand require
7
8
9 that the current exploitation technologies be improved to enhance oil recovery from
10
11 conventional and unconventional resources. The production of conventional and
12
13
14
unconventional crude oil can be affected by several sources of damage under subsurface
15
16 and surface conditions that can affect oil production.1 Formation damage may occur during
17
18 different stages of oil and/or gas production such as production, drilling, hydraulic
19
20
21 fracturing, and workover operations.2 According to Bennion,1 formation damage is
22
23 technically defined as “any process that causes a reduction in the natural inherent
24
25 productivity of an oil or gas producing formation, or a reduction in the injectivity of a
26
27
28 water or gas injection well”. Heavy compounds in crude oil, such as asphaltenes, are a
29
30 source of damage in light,3 medium and heavy oils.4 Although the source of the damage is
31
32 the same, the type of damage is dependent on the crude oil quality and origin. In light crude
33
34
35 oils with low contents of asphaltenes, the damage is induced by the precipitation,
36
37 deposition and adsorption of asphaltenes on the porous media and the entire production
38
39
40
system.3 On the other hand, in crude oils with high asphaltene content, the damage is
41
42 associated with high viscosities that reduce the oil mobility and lead to fingering problems,
43
44 5, 6
resulting in the production of large amounts of water that negatively affect the
45
46
47 environment. Asphaltenes, the heaviest, most aromatic and surface-active fraction of crude
48
49 oil, are defined as the solubility class obtained from crude oil by fractionation using
50
51 solvents such as light paraffins.3, 7-9 According to the Yen-Mullins model, 7, 10
the island
52
53
54 architecture is the predominant structure of asphaltenes with an average molecular weight
55
56 close to 750 Da and an average size close to 1.5 nm. In addition, nanoaggregates, formed at
57
58
59
60
2
ACS Paragon Plus Environment
Page 3 of 37 Industrial & Engineering Chemistry Research

1
2
3 low concentrations tend to self-assemble at higher concentrations leading to large
4
5
6 aggregates that subsequently increase the heavy oil (HO) viscosity.4, 11 The asphaltene self-
7
8 association occurs as their concentration in the system increases, hence promoting the
9
10
formation of nanoaggregates (2 nm) and clusters (5 nm) with aggregation numbers lower
11
12
13 than 10.7, 10 Due to these characteristics of asphaltenes and the consequent problems in the
14
15 production and transportation of crude oil, several mathematical models have been
16
17
18
developed and reported in the literature to understand the phenomena associated with the
19
20 aggregation of asphaltenes. 12-24
21
22 18
Boek et.al. used a slot capillary experiment and the dynamics of stochastic rotation
23
24
25 (SRD) to study the potential interaction of asphaltenes under flow conditions and its effect
26
27 on the asphaltene deposition process. The authors indicated that at a certain flow rate (5
28
29 µL/min), the pressure drop across the capillary increased slowly. However, by increasing
30
31
32 the flow rate (10 µL/min) the aggregation and subsequent precipitation/deposition occurred
33
34 earlier; however, the asphaltene deposits were later dragged by the flow. The simulation
35
36
showed that for slow and fast flow rates, permanent and transient blocking occurred,
37
38
39 respectively. The colloidal interaction potential was identified as a key parameter in the
40
41 flocculation and deposition of asphaltenes under flow conditions and could be modified by
42
43
44
including co-precipitated elements.18
45
46
47 In this regards, a number of population balance equations have been proposed for
48
49 describing the asphaltenes growth.15, 16, 21-25
Khoshandam and Alamdari21 used mass and
50
51 population balance equations for modeling the growth of asphaltenes particles in Heptol
52
53
54 solutions. The proposed model employed four fitting parameters including the coefficient
55
56 of growth rate and the agglomeration coefficient; supersaturation as driving force was
57
58
59
60
3
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 4 of 37

1
2
3 quantitatively related to the agglomeration and growth mechanisms. These authors reported
4
5
6 a good fitting of the model towards the experimental observations made by dynamic light
7
8 scattering with a relative deviation of 10%.21 Maqbool et al.23 developed a geometric
9
10
population balance model that includes the Smoluchowski kernel for the simulation of the
11
12
13 asphaltenes aggregates growth in presence of a precipitant such as n-heptane. The model
14
15 fitted well the experimental observations regarding the asphaltenes aggregates size
16
17
18
distribution and is useful in the prediction of the precipitation onset time of asphaltenes at
19
20 low precipitant concentrations. Additionally, the authors found that the velocity at which
21
22 asphaltenes aggregate in larger particles increase as the concentration of precipitant
23
24
25 increases due to higher driving forces for asphaltenes aggregation. Rahimi and Solaimany-
26
27 Nazar22, 24
build up a fractal population balance model for the prediction of fractal
28
29 aggregates size distribution of Iranian crude oil asphaltene solutions in Heptol mixtures in
30
31
32 presence of shear. They reported that the fractal dimension of the aggregates is dependent
33
34 on the shear and a time-dependent function can be used for it’s description. Additionally,
35
36
the number average diameter curves showed a maximum and further decreases until a
37
38
39 steady state was reached. That trend was attributed to that after the maximum average size
40
41 time, the breakage rate overcome that of aggregation, resulting in smaller aggregates.22
42
43
44
Eskin et al.25 simulated the particle size evolution of asphaltenes in pipeline under turbulent
45
46 flows employing a population balance method. They estimated the collision efficiency,
47
48 particle critical size and particle-wall sticking by fitting with the experimental observations
49
50
51 obtained with a Couette device. Rahmani et al.15, 16 studied asphaltene size distribution in
52
53 organic solvents under a shear rate. Additionally, Rahmani et al. employed a mathematical
54
55 model based on a population balance to analyze the experimental data pertaining to the
56
57
58 aggregate size distribution of asphaltenes under certain shear rates, solvents and asphaltene
59
60
4
ACS Paragon Plus Environment
Page 5 of 37 Industrial & Engineering Chemistry Research

1
2
3 loadings. The model and experimental data described the asphaltene aggregation as two-
4
5
6 process phenomena involving the aggregation and fragmentation of asphaltenes induced by
7
8 the presence of a shear rate until reaching a steady-state aggregate size. 15, 16 Generally, by
9
10
increasing the shear rate, the rate of breakage of the asphaltene aggregates increases,
11
12
13 leading to the formation of smaller aggregate sizes. In addition, as the asphaltene loading or
14
15 the amount of precipitant increases, the rate of asphaltene aggregation growth is increased.
16
17 15, 16
18
19
20
21 Recently, the applications of nanoparticle technology to enhance upgrading and recover of
22
23 crude oil are well recognized in the oil industry due to their high stability, superior
24
25 adsorption and catalytic behavior, as well as their excellent dispersion ability.26, 27
26
27
28 Nanoparticles have been the subject of study in different fields of the Oil & Gas industry
29
30 such as in situ upgrading of heavy and extra heavy oils,26, 28-38 improve the efficiency of
31
32 the drilling fluids39-44 and drilling bit,45, 46
control/inhibition of formation damage due to
33
34
35 fines migration,47-50 organic51, 52 and inorganic scales,43, 53-55 oil removal from production
36
37 water,56, 57 and others.58, 59 Additionally, nanoparticles have become an interesting field of
38
39
40
study regarding the inhibition or prevention of asphaltene damage at the laboratory60-63 and
41
42 oil field scales.64 Therefore, our research group recently developed two models to describe
43
44 adsorption isotherms that accounts for the asphaltene self-association on the solid surface at
45
46
47 reservoir20 and surface19 conditions. These solid−liquid equilibrium (SLE) and solid−liquid
48
49 equilibrium at reservoir conditions (SLE−RC) models were based on the chemical theory
50
51 and provide valuable information about the adsorption affinity of certain adsorbents
52
53
54 towards asphaltenes and the method by which the asphaltenes self-associate over the solid
55
56 surface at equilibrium conditions.19, 20
The effects of the types of adsorbent, asphaltene
57
58
59
60
5
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 6 of 37

1
2
3 chemical nature and the temperature were successfully described by the SLE and SLE-RC
4
5
6 models. In addition, the SLE model provides accurate information about the
7
8 thermodynamic properties of adsorption.19, 34
9
10
11 To the best of our knowledge, no mathematical model in the specialized literature has
12
13
14
included or referred to the effects of the adsorption process of asphaltenes onto
15
16 nanoparticles in the inhibition of asphaltene aggregation, consequently preventing the
17
18 asphaltene formation damage. Therefore, this study aims to present for the first time a
19
20
21 mathematical model that describes the behavior of asphaltene aggregation upon considering
22
23 the adsorption of asphaltenes onto different metal oxide nanoparticles. The mathematics
24
25 and other details of the model are described below.
26
27
28
29
30
31
2. EXPERIMENTAL
32
33
34
35
2.1.Reagents and adsorbents
36
37
38 n-Heptane (99%, Sigma-Aldrich, St. Louis, MO) was used as received for asphaltene
39
40 isolation and for the preparation of heavy oil model solutions. Toluene (99.5%, Merck
41
42
KGaA, Germany) was also used for heavy oil solutions preparation. Fumed silica and γ-
43
44
45 alumina nanoparticles were purchased from Sigma-Aldrich (St. Louis, MO). Magnetite
46
47 nanoparticles were obtained from Nanostructured & Amorphous Materials (Houston, TX).
48
49
50
Nanoparticle samples were characterized by N2 physisorption at 77 K, X-ray diffraction
51
52 (XRD, X Pert PRO MPD, PANalytical, Almelo, the Netherlands) and field emission
53
54 scanning electron microscopy (FESEM, JSM-6701F, JEOL, Japan) for surface area (SBET)
55
56
57 and particle size (dp). Details of the experimental apparatus and procedure used for
58
59
60
6
ACS Paragon Plus Environment
Page 7 of 37 Industrial & Engineering Chemistry Research

1
2
3 nanoparticle characterization can be found in a previous study.38 Panels a-f from Figure 1
4
5
6 show the FESEM images and the corresponding particle size probability distribution for a,
7
8 b) magnetite, c, d) γ-alumina and e, f) silica nanoparticles. Table 1 shows the estimated
9
10
values of the Brunauer, Emmett and Teller (BET) surface areas and the particle size of the
11
12
13 selected nanoparticles. As observed, silica nanoparticles have the largest surface area
14
15 followed by γ-alumina nanoparticles. Magnetite nanoparticles showed the smallest surface
16
17
18
area among the selected nanoparticles. It is also observed that the particle size decreases in
19
20 the order of magnetite > γ-alumina > silica. When comparing the mean crystallite size from
21
22 XRD and FESEM, the values from FESEM are greater than those from XRD, indicating
23
24
25 that a portion of the size distribution has a polycrystalline structure. Accordingly, the d48,
26
27 d45, and d46 values from the probability distributions for silica, γ-alumina and magnetite
28
29 nanoparticles would correspond to single-crystallite structures.
30
31
32
33 a
34
35
36
37
38
39
40
41
42
43
44
45
46
47 c
48
49
50
51
52
53
54
55
56
57
58
59
60
7
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 8 of 37

1
2
3
4 e f 0.02
5

Probability
6 0.015
7
8 0.01
9
10 0.005
11
12
13 0
14 0 3 6 9 12 15 18
15
16 Size (nm)
17
18
19
20 Figure 1. FESEM images and the corresponding particle size probability
21 distribution for a) and b) magnetite, c) and d) γ-alumina and e) and f) silica
22 nanoparticles.
23
24
25
26 Table 1. Estimated BET surface area, mean crystallite and particle size of selected
27 nanoparticles
28
29
30
dp (nm)
31 BET surface area
Material
32 (m2/g) XRD d50 SEM
33 mean crystallite particle diameter
34
35 Magnetite 102.2 70 97
36
37 γ-Alumina 223.2 49 52
38
39 Silica 389.1 7 9
40
41
42
43
44 2.2. n-C7 Asphaltenes
45
46
47 Two Colombian heavy and extra heavy oil (Yang, #229) samples were used as source of
48
49 asphaltenes. The AK18 EHO sample was produced in a field located in the Meta
50
51
52
department in the center region of Colombia. The Cedral HO sample was obtained from an
53
54 oil field located in the Putumayo department, southwest of Colombia. Table 2 shows some
55
56 properties of the selected oil samples. Solid n-C7 asphaltenes were extracted from the crude
57
58
59
60
8
ACS Paragon Plus Environment
Page 9 of 37 Industrial & Engineering Chemistry Research

1
2
3 oil by adding an excess amount of n-heptane following a standard procedure described in a
4
5
6 previous study.9, 65
7
8
9
10
11 Table 2. Properties of selected AK18 and Cedral crude oil samples and their extracted n-C7
12 asphaltenes
13
14 Crude oil
15 Property
16 AK18 Cedral
17
18 API gravity at 289 K (degrees) 7.2 17.3
19
Viscosity at 289 K (Cross, #1370) 2.5 × 106 4.0 × 105
20
21 n-C7 asphaltenes (wt %) 11.3 10.6
22
23
24
25
26 2.3. Asphaltene aggregation kinetics
27
28 It is well known that when a flocculation solvent, such as n-heptane, is added to an
29
30
31 asphaltene solution under shearing, a steady state will be reached after the particle
32
33 aggregation and fragmentation processes are equilibrated.16 Therefore, a desired amount of
34
35 n-C7 asphaltenes is added to toluene and magnetically stirred for 1 h at 300 rpm. Then, n-
36
37
38 heptane in the proportions of 20 v/v% (Heptol 20) and 40 v/v% (Heptol 40) is added to the
39
40 heavy oil model solutions to complete a fixed concentration of 1000 mg/L n-C7
41
42
asphaltenes. It is expected that the asphaltene size will increase as the amount of n-heptane
43
44
45 increases.66 Immediately, the solutions are magnetically stirred at a fixed shear rate of 300
46
47 rpm, and aliquots are taken to obtain the mean asphaltene aggregate size at a desired time.
48
49
50
DLS measurements for the experiments on asphaltene aggregation kinetics were performed
51
52 using a nanoplus-3 from Micromeritics (Norcross, ATL) set at a desired temperature and
53
54 equipped with a 0.9 mL glass cell. 66-68 It is noteworthy that no asphaltene precipitation was
55
56
57 observed for the employed Heptol solutions, as verified by the Oliensis Spot Test Number
58
59
60
9
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 10 of 37

1
2
3 69, 70
and Polarized Light Microscopy.37, 61 The asphaltene mean particle size was measured
4
5
6 until a steady state was reached. The scattering angle varies according to the solvent used to
7
8 optimize the intensity of the flocculation of scattering light. The mean aggregate diameter
9
10
11
of asphaltenes d asp (hydrodynamic diameter) is obtained from the Stokes-Einstein equation,
12
13 as follows: 71
14
15
16 kBT
17 d asp = (1)
18 3πη Da
19
20
21 where k B (1.38 x 10-23 m2kg·s-2K-1) is the Boltzmann constant , T (K) is temperature, η
22
23
24
(Cross, #1370) is the viscosity of the medium, and Da (m2·s-1) is the diffusion coefficient
25
26 of the particles.
27
28
29 In addition, to evaluate the inhibition of the asphaltene aggregation using nanoparticles, the
30
31
32
procedure was repeated in the presence of the selected nanomaterials with dose of 10 g/L of
33
34 solution. The nanoparticle dose was selected to allow the supernatant decantation for
35
36 accurate measurements of asphaltene aggregation by minimizing the risk of noise due to
37
38
39 suspended nanoparticles.
40
41
42 2.4.Adsorption kinetics experiments
43
44
45 Batch-mode adsorption experiments were performed at a ratio of 1:10 (L:g) model heavy
46
47 oil solution/mass of the nanoparticles following a procedure described in previous works.9,
48
49 65
In brief, the adsorption of asphaltene in the experiments was monitored for 100 min at a
50
51
52 fixed asphaltene initial concentration of 1000 mg/L, similar to the concentration used for
53
54 the aggregation kinetics tests, and the solution temperature was 293 K. Samples were
55
56
57
selected at predetermined time intervals and analyzed for asphaltene concentration using a
58
59
60
10
ACS Paragon Plus Environment
Page 11 of 37 Industrial & Engineering Chemistry Research

1
2
3 Genesys 10S UV-vis spectrophotometer (Thermo Scientific, Waltham, MA). The
4
5
6 calibration curves of asphaltene concentrations against absorbance were constructed at
7
8 fixed wavelengths of 290 and 285 nm for AK18 and Cedral n-C7 asphaltenes, respectively,
9
10
using lab-made standard model solutions with known concentrations. The adsorbed amount
11
12
13 of n-C7 asphaltenes relative to the mass of nanoparticles, q (mg/g), was estimated
14
15
16
following the mass balance qt = (Ci − Ct ) / M , where qt (mg/g) and Ct (mg/L) are the
17
18 concentration of asphaltenes in the solution at a desired time t (min), and M (g/L) is the
19
20
21 ratio of dry mass nanoparticles to solution volume.
22
23
24
25
3. AGGREGATION AND INHIBITION MODEL OF ASPHALTENE
26
27 GROWTH
28
29
30
3.1. General population balance equation
31
32 To model the kinetics of aggregation and inhibition, a PBM that accounts for the adsorption
33
34 of asphaltene moieties onto nanoparticles is proposed. The model assumes that asphaltene
35
36 aggregation in the presence of a shear rate is related to the aggregation and fragmentation
37
38
39 phenomena. Rahmani et al.16 proposed a PBM that describes the kinetics of aggregation
40
41 and fragmentation of asphaltene particles following the expression developed by Austin
42
43
44
197172 and Friedlander 197773 for the change rate of the particle size of a given size due to
45
46 the processes of aggregation and fragmentation. This model is based on the theory that at
47
48 the initial stages of the aggregation/fragmentation process, the asphaltene aggregate size
49
50
51 grows rapidly, i.e., aggregation is promoted. The fragmentation phenomena gains
52
53 importance as the aggregation increases and as the hydrodynamic stresses equalize to cause
54
55 stress of the aggregates.72, 74 Therefore, the asphaltene aggregate size is dependent on the
56
57
58 dominant phenomena in a specific moment. In this study, a modification of the model
59
60
11
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 12 of 37

1
2
3 employed by Rahmani et al. 16
is proposed to account for the effect of nanoparticles in
4
5
6 solution. We believe that the presence of nanoparticles will affect the kinetics of asphaltene
7
8 aggregation/fragmentation; hence, a new term representing aggregate inhibition provided
9
10
by nanoparticles through the adsorption process is included in the new version of the PBM
11
12
13 model, as described in equation 2:
14
15
16 dni dq
17 = B−D− (2)
dt dt
18
19
20 where,
21
22
23 dni
24 • is the rate number concentration (asphaltene particles per volume with size i,
dt
25
26
27
28 • B and D are general terms regarding the appearance and disappearance of asphaltene
29
30 particles of size i, respectively, and
31
32
33 dq
34 • is the term related to the rate of adsorption of asphaltene particles on
35 dt
36
37 nanoparticles.
38
39
40
A detailed expression of the inhibition model of asphaltene aggregations is shown in
41
42
43 equation 3,16, 73, 75 after considering the adsorption on nanoparticles
44
45
nmax nmax
46 dni 1 dq
47 = ∑ αβ jk (Vj , Vk )n j nk − ni ∑αβki (Vk ,Vi )nk − Bi ni + ∑ γ i, j B j n j − (3)
48 dt 2 j +k =i k =1 j =i +1 dt
49
50
where, ni is the number concentration of aggregates with size i that contains i primary
51
52
53 particles, and nmax is the maximum size of the particles that will form fragments of size i
54
55
56 before fragmentation. In equation 3 from right to left, the first term defines the adsorption
57
58
59
60
12
ACS Paragon Plus Environment
Page 13 of 37 Industrial & Engineering Chemistry Research

1
2
3 kinetic on nanoparticle, the second and fifth terms define the formation of i-sized particles,
4
5
6 and the third and fourth terms account for the disappearance of particles of size i. β jk is the
7
8
9 collision frequency between particles of volume V j and Vk due to Brownian motion and
10
11 fluid shear, as expressed in equation 4: 16, 76, 77
12
13
14
2 K BT d i − d j 2 G
15 B jk = ( ) + ( d i + d j )3 (4)
16 3 µ di d j 6
17
18
19 where, µ (Pa·s) is the solution viscosity, and G (1/s) is the shear rate. The volume of a
20
21 particle of size i is expressed as a function of the volume in the previous step, as follows:
22
23
24
fVi −1
25 Vi = (5)
26 1− ε
27
28 where, f is the sectional spacing, which is equal to 2, 16, 74, 78, 79
and ε is the aggregate
29
30
31 porosity that is assumed constant for aggregate formation from primary particles and is
32
33 used as a fitting parameter.
34
35
36 The collision efficiency, α , i.e., the fraction of collisions that results in aggregations, is
37
38
39 assumed equal to 1.0 because large aggregates have fractal and porous structures.78-80
40
41 Nevertheless, the collision efficiency is a function of the surface properties of the particles,
42
43 the structure of the aggregates, the diameter of the aggregates,16 the prevailing colloidal
44
45
46 force and hydrodynamic effects.81, 82
Additionally, γ i , j represents the volume fraction of
47
48
49 aggregates of size i originating from aggregates of size j expressed as γ i , j = V j Vi , and Bi is
50
51 the fragmentation or breakup rate of aggregates of size i , given as a function of particle
52
53
54 volume, as follows: 83-85
55
56
57
58
59
60
13
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 14 of 37

1
2
3 Bi = bVi 0.33 (6)
4
5
6 where, b is the breakup rate coefficient for shear-induced fragmentation.16 In this work, b
7
8
is assumed as a fitting parameter to correlate the experimental and simulation results.
9
10
11
12
3.2. Accounting for the kinetics of asphaltene adsorption onto nanoparticles
13
14
15
The inhibition term of asphaltene aggregation was introduced into the model using
16
17 adsorption kinetics of asphaltenes onto the nanoparticles. Asphaltenes are first found in the
18
19 bulk solution as large aggregates, and as nanoparticles gradually interact with the medium,
20
21
22 the adsorption forces become greater than the aggregation forces, and disaggregation of
23
24 asphaltene begins. Then, asphaltenes migrate from the bulk solution to the nanoparticle
25
26 surface depending on the adsorption potential induced by the adsorbent. In a determined
27
28
29 time, the nanoparticles adsorb a certain mass of asphaltenes that would not be able to
30
31 interact in the process of asphaltene aggregation; therefore, the volume of asphaltenes
32
33
34
available for growth decreases.
35
36
37
The double exponential model86, 87
is used to determine the adsorption rate. Equation 7
38
39 describes the double exponential model used to fit the kinetic adsorption experimental data,
40
41 respectively: 86, 87
42
43
44
Df Ds
45 qt = qm − exp(−k f t ) − exp(−kst ) (7)
46 M M
47
48 The double exponential model describes in two steps the kinetics of adsorption, namely, a
49
50
51 fast and a slow step.88 Therefore, k f (min-1) and ks (min-1), as well as Df (mg/L) and Ds
52
53
54 (mg/L) are the rate constants and adsorption coefficients of the fast and slow steps of the
55
56
57
58
59
60
14
ACS Paragon Plus Environment
Page 15 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
double exponential model, respectively, and qm (mg/g) is the maximum amount of
5
6 adsorbed asphaltenes.
7
8
9 Thus, the formation of large asphaltene aggregates in solution is inhibited in the presence of
10
11
12 nanoparticles, as the adsorption on the nanoparticle hinders or changes the growth of
13
14 asphaltene aggregate. Therefore, considering the adsorption kinetics of the nanoparticles,
15
16 the free asphaltene primary particles are re-calculated each time, and this population is used
17
18
19 in the model to find the aggregate size at a determined time. As described earlier, the free
20
21 asphaltene primary particle is defined as the smallest particle distribution detected by
22
23 means of the observation method employed (see next section). Additional information on
24
25
26 the primary particles re-calculation can be found in the Supporting Information.
27
28
29 3.3. Simulation characteristics
30
31
32 3.3.1 Size and number concentration of primary asphaltene particles: The initial size
33
34 distribution and the initial number concentration of the asphaltene particles were
35
36
37 determined using experimental data. The volume of the primary particles was calculated
38
39 according to the particle diameter obtained from DLS experiments assuming asphaltenes
40
41 aggregates as spherical particles.16 Then, at the initial state, the suspension is assumed
42
43
44 homogeneous, and the distribution of the primary particles is calculated using the mass and
45
46 density of asphaltenes assumed as 1200 kg/m3, as reported in the literature.15, 16, 89
47
48
49 3.3.2 Simulation conditions: The developed model of aggregation and inhibition of
50
51
asphaltenes considered in this study has the following characteristics and restrictions:
52
53
54
55
56
57
58
59
60
15
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 16 of 37

1
2
3
• It is a grey-box model in which the structure is given by the equation of the
4
5
6 population balance of asphaltenes, and the parameters were determined by
7
8 constitutive equations.
9
10
11
• It is on the macroscopic scale and discrete because the growth of aggregates occurs
12
13
14 over a wide size range; thus, a continuous model of flocculation may require
15
16 excessive computation time.
17
18
19
20
• Although the asphaltene aggregation is typically described by probability
21
22 distributions, the developed model has deterministic characteristics to obtain a
23
24 practical model for the inhibition process in the presence of nanoparticles.
25
26
27
28
• It has dynamic characteristics because the size of the asphaltene aggregates change
29
30 over time.
31
32
33 • It has distributed parameters because each “family” of asphaltenes has their own
34
35
36
values.
37
38
39 • Considering these characteristics, the model was solved numerically by a finite
40
41 difference scheme using the Quasi-Newton Method.90, 91
42
43
44
45 The model parameters are obtained by minimizing the root mean square error (RMSE%,
46
47 equation 8) between the theoretical and experimental data points using an implemented
48
49
50
routine in the program code, as follows:
51
52
53
1  p  Exp − Cal  
2

∑ 
54 RSME % = 100   (8)
55 p  1  Cal  
56
57
58
59
60
16
ACS Paragon Plus Environment
Page 17 of 37 Industrial & Engineering Chemistry Research

1
2
3
with p as the number of points, and Exp and Cal as the experimental and model
4
5
6 calculated values, respectively.
7
8
9 4. RESULTS AND DISCUSSIONS
10
11
12 The PBM proposed in this work describes the kinetics of asphaltene aggregation-
13
14
fragmentation in systems with different conditions in the absence and presence of
15
16
17 nanoparticles. The results and analyses are divided into two main sections: (i) kinetics
18
19 of asphaltene adsorption on the selected nanoparticles and (ii) kinetics of asphaltene
20
21
22
aggregation-fragmentation in the absence and presence of nanoparticles. The effects of
23
24 the amount of precipitant in the solution, the asphaltene source, the system temperature
25
26 and the nanoparticles’ chemical nature and dosage were evaluated to demonstrate the
27
28
29 PBM versatility.
30
31
32 4.1. Kinetics of n-C7 asphaltene adsorption on the selected nanoparticles
33
34 4.1.1. Effect of the nanoparticles’ chemical nature
35
36
Panels a and b from Figure 2 show the adsorption kinetics of asphaltenes on the selected
37
38
39 nanoparticles together with the double exponential adsorption kinetic model fit at
40
41 different n-heptane concentrations of a) Heptol 20 and b) Heptol 40 at a fixed
42
43
44
temperature of 293 K and a particles dosage of 10 g/L. For all of the cases, it was
45
46 observed that asphaltene adsorption onto nanoparticles is a fast process requiring times
47
48 less than 70 min to reach saturation. Note that at the beginning of the adsorption process
49
50
51 in the magnetite nanoparticles case and for Heptol 40, 93% of the total amount adsorbed
52
53 is reached in 10 min and is likely to be due to the non-porous characteristics of the
54
55 nanoparticles that favor a rapid first step of asphaltene adsorption just over the
56
57
58 nanoparticles surface.9, 31, 65
From Figure 2, it is also observed that for the three
59
60
17
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 18 of 37

1
2
3 nanoparticles evaluated, the amount adsorbed increases as the amount of n-heptane in
4
5
6 the solutions increases. This reflects the increase of asphaltene aggregates in solution
7
8 and the subsequent decrease in asphaltene solubility. The results are in good agreement
9
10
with the trend observed by Franco et al.,37 who evaluated the adsorption of n-C7
11
12
13 asphaltenes from different Heptol solutions over fumed silica and functionalized fumed
14
15 silica nanoparticles and obtained higher asphaltene uptake as the amount of n-heptane
16
17
18
increased.
19
20
21
22
23
24 a 100 b 100
25
26 80 80
27
28
q (mg/g)
q (mg/g)

29 60 60
30
31
32 40 40 Silica
33 Silica
34 ɣ-alumina ɣ-alumina
35 20 20 Magnetite
Magnetite
36
37 Double exponential model Double exponential model
38
0 0
39 0 20 40 60 80 100 0 20 40 60 80 100
40
41 Time (min) Time (min)
42
43 Figure 2. Adsorption kinetics for AK18 n-C7 asphaltenes onto the selected
44 nanoparticles from solutions of a) Heptol 20 and b) Heptol 40. Temperature: 293 K;
45 adsorbent dose, 10 g/L. Continuous curves represent double exponential model
46
47 fitting (eq. 7).
48
49 The double exponential model was used to describe the adsorption kinetics of AK18 n-C7
50
51 asphaltenes onto the three employed nanoparticles. Table 3 shows the estimated parameters
52
53 for the model used, the respective correlation coefficient and the root square mean error. As
54
55
56 shown in Table 3, the double exponential model showed an excellent correlation towards
57
58
59
60
18
ACS Paragon Plus Environment
Page 19 of 37 Industrial & Engineering Chemistry Research

1
2
3
the experimental data with RSME % < 4.10%. This means that due to van der Waals and
4
5
6 electrostatic attraction forces, a first and rapid initial kinetic mechanism occurred where the
7
8 adsorbate diffuses toward the adsorbent.56, 57, 88 Values higher than 0.0 for the parameters ks
9
10
11 and Ds indicate that the fast step is followed by a gradual or slow adsorption of the
12
13
14 asphaltenes over the nanoparticle surface by complexation.56, 57, 92
These findings are in
15
16 good agreement with the studies reported by Nassar 9 and Franco et al.31, 65 Regarding the
17
18
19 adsorption efficiency in the fast step ( k f ), for the Heptol 20 solutions, it can be observed
20
21
that they tend to decrease in the order of magnetite > γ-alumina > silica, indicating that as
22
23
24 the surface area increases, longer times are required to reach the maximum capacity of
25
26 adsorption. However, the results of the double exponential model indicate that there is not a
27
28
29
second step for magnetite nanoparticles, and the rate of adsorption is higher for γ-alumina
30
31 than for silica nanoparticles. On the other hand, for Heptol 40, the trend followed by both
32
33 k f and ks is γ-alumina > silica > magnetite. In the case of magnetite nanoparticles in
34
35
36 Heptol 40, the dosage of nanoparticles was duplicated to 20 g/L, resulting in a decreasing in
37
38
39
the amount adsorbed in units of mg/g. However, when amount adsorbed is not normalized
40
41 by dry mass of nanoparticles, an increase in the n-C7 asphaltenes is observed when
42
43 comparing with the results at 10 g/L and could be due to a larger amount of active sites
44
45
46 available for adsorption. More details of the effect of the nanoparticles dosage can be found
47
48 in the Supporting Information.
49
50
51 4.1.2. Effect of asphaltene source
52
53
54 Adsorption of n-C7 asphaltenes depends on many factors and relevant to this work are
55
56
among others, structure and composition, capacity of sample to pack at the interface,
57
58
59
60
19
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 20 of 37

1
2
3 concentration, aggregation in solution and time.93 Thus any attempt to account for the
4
5
6 differences shown in Figure 3 would be speculative. Hence, we speculate that such a
7
8 difference could in part be due to differences in aggregation trends. As described earlier,93
9
10
aggregation in solution hinder the adsorption and this would be consistent with a higher
11
12
13 aggregation trend for the Cedral n-C7 asphaltenes. These arguments are consistent with the
14
15 higher hydrodynamic radii found for Cedral n-C7 asphaltenes when compared with the
16
17
18
AK18 n-C7 asphaltenes.
19
20
21 Table 3. Estimated parameters for the double exponential model for n-C7 asphaltene
22 obtained from Ak18
23
24 qm ,exp
25 Heptol qm Df kf Ds ks ×
26 Ratio
Sample 10-3 R2 RSME %
(mg/g) (mg/g) (mg/L) (min-1) (mg/L)
27 (min-1)
28
29 Magnetite 70.46 106.60 420.30 0.09 194.14 1.29 0.99 3.39
30 20 ɣ-alumina 77.80 104.58 325.04 0.12 323.70 2.09 0.98 4.10
31
Silica 88.49 70.17 401.18 0.20 0.00 0.00 0.99 1.30
32
33
34 Magnetite 78.73 119.69 211.95 0.14 287.66 1.97 0.99 1.59
35
40 ɣ-alumina 85.93 113.71 194.78 0.16 350.32 2.32 0.99 0.91
36
37 Silica 95.14 93.72 235.76 0.12 168.59 1.19 0.99 1.32
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
20
ACS Paragon Plus Environment
Page 21 of 37 Industrial & Engineering Chemistry Research

1
2
3
4 90
5 80
6
7 70
8 60
q (mg/g)
9
10 50
11 40
12
13 30 AK18 n-C7 asphaltenes
14 20 Cedral n-C7 asphaltenes
15
16 10 Double exponential model
17 0
18
19 0 20 40 60 80 100
20
21
Time (min)
22
23 Figure 3. Adsorption kinetics for AK18 and Cedral n-C7 asphaltenes from Heptol 40 on
24 magnetite nanoparticles. Temperature: 293 K; adsorbent dose, 10 g/L. Continuous curves
25 represent double exponential model fitting (eq. 7).
26
27
28
29 Again, the double exponential model showed excellent fit to the experimental data with a
30
31
32
R 2 of 0.99 and a RSME % of 0.43, respectively. In addition, the results indicated the
33
34 absence of the slow step by obtaining values of ks and Ds equal to 0.00, indicating that the
35
36
37 Cedral n-C7 asphaltene adsorption on magnetite nanoparticles is primarily a single-step
38
39 process. Values of qm , k f and Df were 47.86 mg/g, 0.65 (min-1) and 440.42 mg/L,
40
41
42 respectively, suggesting that the first adsorption step is more efficient for Cedral n-C7
43
44
45
asphaltenes than for AK18 n-C7 asphaltenes. However, higher qm value was obtained for
46
47 the latter than the former.
48
49
50
51
52 4.2. Asphaltene aggregation-fragmentation kinetics
53
54
55
The proposed PBM model was validated through series of DLS measurements for the
56
57 mean aggregate particle size of asphaltenes under different conditions of asphaltene
58
59
60
21
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 22 of 37

1
2
3 origin, Heptol solution and nanoparticle chemical nature. Different adsorption kinetic
4
5
6 models were used to account for the asphaltene adsorption on the selected nanoparticles
7
8 and their effect on the inhibition of the asphaltene growth. For each system, the kinetic
9
10
adsorption double exponential model was used for describing the n-C7 asphaltenes
11
12
13 adsorption over the nanoparticles.
14
15
16
17
18
4.2.1. Effect of the nanoparticles’ chemical nature on n-C7 asphaltene
19
20 aggregation-fragmentation
21
22
23 It is expected that the presence of nanoparticles in a solution of asphaltene aggregates
24
25 would create an adsorption potential that would promote the n-C7 asphaltenes adsorption
26
27
28 over the nanoparticles. Once the potential is induced, the interaction nanoparticle-n-C7
29
30 asphaltenes overcome that of n-C7 asphaltene- n-C7 asphaltenes and hence, the asphaltene
31
32 aggregate tends to disaggregate itself to migrate as molecules or smaller aggregates from
33
34
35 the bulk solution to the adsorbent surface. This means that as n-C7 asphaltenes adsorb on
36
37 the nanoparticles surface, the asphaltene particles available in the aggregation system
38
39
40
decrease. As the amount of asphaltenes adsorbed increases, the adsorption potential
41
42 decreases, and gradually, the nanoparticle- n-C7 asphaltenes interactions decrease until the
43
44 maximum amount adsorbed is reached. Then, the asphaltene aggregates are re-organized
45
46
47 but with a smaller aggregate size than prior to introducing the nanoparticles.
48
49
50 Figure 4 shows the evolution of the virgin AK18 n-C7 asphaltene aggregate size as a
51
52 function of time for asphaltenes in Heptol 20 and Heptol 40 along with the PBM fit. As
53
54 observed, when the amount of n-heptane increases, the average size of the asphaltene
55
56
57 aggregates at any time also increases. For example, in the Heptol 40 system, the asphaltene
58
59
60
22
ACS Paragon Plus Environment
Page 23 of 37 Industrial & Engineering Chemistry Research

1
2
3 aggregates reach a maximum size of 1086 nm, while in the Heptol 20 system, the
4
5
6 maximum size observed was 842 nm, which are found at 190 and 236 min, respectively.
7
8 This is in agreement with the results reported by Maqbool et al.,23 who reported that as the
9
10
amount of n-heptane increased, the speed at which asphaltenes self-associate to form larger
11
12
13 aggregates increased due to the to higher driving forces. This, of course, is expected in
14
15 view of the much lower solubility parameter of n-heptane when compared to toluene.94, 95
16
17
18
In addition, the ratio of polar/nonpolar moieties in the asphaltene structure defines the
19
20 polarity 95, 96 and leads to either a less or more favorable self-association phenomena.19
21
22
23 It is also observed that for both of the cases, the asphaltene size increases until a maximum
24
25 and then decreases. This is due to the kinetics of asphaltene growth in the presence of a
26
27
28 shear that involves both aggregation and fragmentation due to the asphaltene aggregates
29
30 collision.16 Table 4 shows the obtained parameters for the PBM fitting for kinetic
31
32 aggregation-fragmentation of asphaltenes in the absence and presence of nanoparticles. As
33
34
35 shown in Figure 4 and according to the obtained values of R2 and RSME % of 0.97 and
36
37 5.47, respectively, the model has a good fit to the experimental results. The parameters ε
38
39
40 and b varied depending on the system evaluated. For virgin asphaltenes, ε has values of
41
42 0.31 and 0.46 for Heptol 20 and Heptol 40 solutions, respectively, indicating a dependence
43
44
45
on solvent composition (Heptol). This is because by increasing the amount of n-heptane in
46
47 the Heptol solutions, the solubility of the asphaltenes decreases, leading to the formation of
48
49 larger aggregates, hence increasing the porosity. The values of the porosity of the
50
51
52 aggregates are in agreement with the findings by Rahmani and Masliyah.16 For the breakup
53
54 coefficient, b, the values had a weak dependence on the solvent composition for
55
56
57
58
59
60
23
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 24 of 37

1
2
3 experiments with virgin AK18 n-C7 asphaltenes with a value of 4.26 ± 0.01 for the Heptol
4
5
6 20 and Heptol 40 systems.
7
8
9 1200
10
11
12 1000
13
14 800
dasp (nm)

15
16 600
17
18
19 400 Heptol 40
20 Heptol 20
21 200 PBM
22
23 0
24
25 0 100 200 300
26
27 Time (min)
28
29 Figure 4. Kinetics of virgin AK18 n-C7 asphaltenes aggregation-fragmentation at a
30 temperature of 293 K in Heptol 20 and Heptol 40 solutions. The symbols are experimental
31
32 data, and the solid lines are from the PBM model (eq 3).
33
34
35
36
37
Figure 5 a and b shows the AK18 n-C7 asphaltene growth kinetics in the presence of the
38
39 selected nanoparticles for solutions of a) Heptol 20 and b) Heptol 40. Using DLS
40
41 experiments, the kinetic of asphaltene aggregation-fragmentation in the presence of silica
42
43
44 nanoparticles in Heptol 20 solutions could not be determined due to the low signal/noise
45
46 ratio. Probably this is due to the size of particles being lower that the detection limit of the
47
48 DLS technique used.97 This is coherent with the results observed in Figure 5b for n-C7
49
50
51 asphaltenes growth in the presence of silica nanoparticles in Heptol 40 that showed the
52
53 lowest asphaltene aggregates among the evaluated nanoparticles. For the other materials,
54
55 smaller aggregates were found in the order of γ-alumina < magnetite < virgin AK18 n-C7
56
57
58 asphaltenes, which was true for both Heptol 20 and Heptol 40 solutions. Although the
59
60
24
ACS Paragon Plus Environment
Page 25 of 37 Industrial & Engineering Chemistry Research

1
2
3 effect of nanoparticles in terms of affinity cannot be compared due to the differences in the
4
5
6 chemical nature of the materials, the adsorption process can be a key factor in
7
8 understanding the effect of the nanoparticles in the inhibition of the asphaltene self-
9
10
association in the oil matrix. The observed trend corresponds to the results obtained for the
11
12
13 adsorption kinetics that revealed a higher adsorption capacity for γ-alumina than for
14
15 magnetite nanoparticles. Due to the magnetic characteristics of magnetite nanoparticles,
16
17
18
they tend to aggregate due to anisotropic dipolar attraction and form clusters,98 hence
19
20 reducing the surface area available for asphaltene adsorption. Thus, higher adsorption
21
22 capacities will result in higher asphaltene adsorption and then will result in a reduction of
23
24
25 the asphaltenes available in the solution for the aggregation-fragmentation process.
26
27
28 The size of nanoparticles is another factor to be considered (see Table 1); in fact,
29
30 nanoparticles size and capacity to reduced aggregate size are in line with each other. We
31
32 suggest this is related to the easy diffusion of the nanoparticles into the solution. This
33
34
35 allows for a more effective scavenging of the dissolved n-C7 asphaltene material.
36
37
38 The proposed PBM also showed a good fit to the experimental results for n-C7 asphaltenes
39
40 in the presence of nanoparticles and was confirmed by values of R 2 > 0.94 and RMSE% <
41
42
43 8.7. The ε parameter decreased in the order of virgin AK18 n-C7 asphaltenes > magnetite
44
45 > γ-alumina > silica, indicating that smaller aggregates result from the addition of
46
47
48
nanoparticles to the system, which is in agreement with the experimental results in Figure
49
50 5. On the other hand, the b parameter did not follow a clear trend, according to the kinetics
51
52 of asphaltenes aggregation-fragmentation; this may be because the breakup coefficient
53
54
55 would change as a function of the resultant asphaltene concentration and could also be
56
57 affected by the presence of nanoparticles, their particle size and morphology. For example,
58
59
60
25
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 26 of 37

1
2
3 magnetite nanoparticles that have a more squared shape than alumina nanoparticles would
4
5
6 generate a different field of adsorption, thus directly affecting the rate of breakup of
7
8 asphaltene aggregates. In addition, this may be because the breakup rate coefficient
9
10
depends on not only the flocculants type and concentration but also the properties that are
11
12
13 strongly influenced by the presence of the nanoparticles, such as the primary particles, flock
14
15 structure and suspension of the medium.
16
17
18 Fragmentation of large aggregates by shear means that corresponding fragmentation energy
19
20
21 is small and comparable to the available thermal energy; thus, relative small changes in
22
23 temperature leads to significant changes in the aggregate size distribution. Also, when
24
25 applying a higher dosage of magnetite nanoparticles of 20 g/L in Heptol 40 solutions, it
26
27
28 was found that the asphaltene aggregates show a greater decrease as the amount of
29
30 nanoparticles may significantly alter the asphaltene aggregate size in solution because more
31
32 active sites will be available for adsorption. More information of nanoparticles dosage and
33
34
35 temperature dependence can be found in the Supporting Information.
36
37
38
39
40
41 Table 4. Estimated parameters of the population balance model (PBM).
42
43
44
Heptol b ×105 R2 RMSE%
45 Sample ε
46 Ratio
47
48
49
Virgin AK18 n-C7 asphaltenes 0.31 4.27 0.97 5.47
50 20 Magnetite 0.24 3.59 0.94 8.43
51 ɣ-alumina 0.20 5.57 0.94 8.68
52 Silica - - - -
53 Virgin AK18 n-C7 asphaltenes 0.46 4.25 0.98 4.73
54
55 Magnetite 0.41 6.03 0.98 4.28
40
56 ɣ-alumina 0.37 4.69 0.95 4.58
57 Silica 0.22 3.96 0.97 8.40
58
59
60
26
ACS Paragon Plus Environment
Page 27 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5 a 1000
6
7
8 800 Virgin AK18 n-
9 C7 asphaltenes
dasp (nm)

10
11
600 Magnetite
12
13 400
14 ɣ-alumina
15
16 200 PBM
17
18
19 0
20
21 0 100 200 300
22 Time (min)
23
24 1200
25 b
26 1000
27
28 Virgin AK18 n-C7
29 800 asphaltenes
30 Magnetite
dasp (nm)

31
32
600
33 ɣ alumina
34 400
35 Silica
36
37 200
38 PBM
39 0
40
41 0 100 200 300
42
43
Time (min)
44
45
46 Figure 5. Kinetics of asphaltenes aggregation-fragmentation in the presence of
47 nanoparticles for a) Heptol 20 and b) Heptol 40 at 293 K. The symbols are experimental
48 data, and the solid lines are from the PBM model (eq. 3).
49
50
51
52
53
54
55
56
57
58
59
60
27
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 28 of 37

1
2
3 4.2.2. Effect of n-C7 asphaltenes origin on their aggregation-fragmentation
4
5
6 Heptol 40 solutions and magnetite nanoparticles were selected to study the effect of n-C7
7
8
9 asphaltene origin on the processes of aggregation-fragmentation in the presence and
10
11 absence of nanoparticles. Figure 6 shows the kinetics of asphaltene growth for a) AK18 n-
12
13
14
C7 asphaltenes and b) Cedral n-C7 asphaltenes. From Figure 6, it can be observed that
15
16 Cedral n-C7 asphaltenes are more prone to aggregate in a Heptol 40 solution than AK18 n-
17
18 C7 asphaltenes, with maximum sizes of 1086 and 1655 nm, respectively. The PBM also
19
20
21 agrees with the experimental results with R 2 values of 0.95 and 0.98 and RMSE% values
22
23 of 6.99 and 4.86 for Cedral n-C7 asphaltenes in the absence and presence of magnetite
24
25
26 nanoparticles, respectively. The b parameter takes values of 3.96 × 10-5 and 4.02 × 10-5,
27
28 indicating that the breakup rate in the absence and presence of magnetite nanoparticles does
29
30 not vary significantly. This is also supported in Figure 6b where it can be seen that the
31
32
33 slope for the points between 100 and 300 min is similar for both of the systems with and
34
35 without nanoparticles. On the other hand, the ε parameter with a value of 0.47 is higher
36
37
38
than for virgin AK18 n-C7 asphaltenes and agrees with the experimental results as larger
39
40 aggregates were observed for the asphaltenes extracted from the Cedral crude oil. In
41
42 addition, for Cedral n-C7 asphaltenes in the presence of magnetite nanoparticles, the value
43
44
45 obtained for the porosity was 0.29, which was lower than the value obtained for virgin
46
47 Cedral n-C7 asphaltenes.
48
49
50
51
52
53
54
55
56
57
58
59
60
28
ACS Paragon Plus Environment
Page 29 of 37 Industrial & Engineering Chemistry Research

1
2
3
4 a 1200 b 2000
5
6 1000
7 1600
8 800

dasp (nm)
1200
dasp (nm)

9
Virgin Cedral n-C7 asphaltenes
10 600 Magnetite
11 800 PBM
12 400 Virgin AK18 n-C7 asphaltenes
13 400
14 200 Magnetite
15 PBM
16 0 0
17 0 100 200 300 0 100 200 300
18
19
Time (min) Time (min)
20
21
22 Figure 6. Kinetics of aggregation-fragmentation of a) AK18 n-C7 asphaltenes and b)
23 Cedral n-C7 in the absence and presence of magnetite nanoparticles at 293 K and for Heptol
24 40 solutions. The symbols are experimental data, and the solid lines are from the PBM
25 model (eq. 3).
26
27
28
29 5. CONCLUSIONS
30
31
A model to describe the asphaltene aggregation-fragmentation in the absence and
32
33
34 presence of different metal oxide nanoparticles that accounts for the adsorption process
35
36 was successfully developed. The model validity was sustained in population balance
37
38
39
equations, and the adsorption term was introduced by employing the double kinetic
40
41 exponential model. Under the different conditions examined all nanoparticles in
42
43 different degrees reduce the hydrodynamic radii of large aggregates in solution as a
44
45
46 result of adsorption. The effect of the nanoparticles’ chemical nature, asphaltene origin,
47
48 Heptol solution and temperature was successfully evaluated through DLS
49
50 measurements. In general, the population balance model had a good fit towards the
51
52
53 experimental results with RSME % < 9. For all of the systems evaluated, the double
54
55 exponential model showed a high affinity; therefore, it was incorporated into the PBM.
56
57
58
59
60
29
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 30 of 37

1
2
3
In addition, the estimated PBM parameters ε and b were in agreement with the
4
5
6 behavior of the experimental curves of asphaltene aggregation-fragmentation.
7
8
9
10
11 6. ACKNOWLEDGMENTS
12
13 The authors acknowledge The Universidad Nacional de Colombia for logistical and
14 financial support and COLCIENCIAS and ECOPETROL for the support provided in
15 Agreement 264 of 2013 and the Cooperation Agreement 04 of 2013, respectively. Thanks
16 are due to Mr. Juan D. Aristizabal for his fruitful support. The authors are also grateful to
17
18
the Natural Sciences and Engineering Research Council of Canada (NSERC), Nexen-
19 CNOOC Ltd, and Alberta Innovates-Energy and Environment Solutions (AIEES) for the
20 financial support provided through the NSERC/NEXEN/AIEES Industrial Research Chair
21 in Catalysis for Bitumen Upgrading. Also, the contribution of facilities from the Canada
22 Foundation for Innovation, the Institute for Sustainable Energy, Environment and
23 Economy, the Schulich School of Engineering and the Faculty of Science at the University
24
25 of Calgary are greatly appreciated.
26
27
28 7. Supporting Information Available
29
30 Supporting Information includes three main sections namely: Section S1. Re-calculating
31
32 free asphaltenes primary particles, Section S2. Effect of nanoparticle dosage and system
33 temperature on the n-C7 asphaltenes adsorption and Section S3. Effect of nanoparticle
34
dosage and system temperature on n-C7 asphaltene aggregation-fragmentation. This
35
36 material is available free of charge via the Internet at http://pubs.acs.org.
37
38
39
40 8. REFERENCES
41
42 (1) Bennion, D. B. An Overview of Formation Damage Mechanisms Causing a Reduction in the
43 Productivity and Injectivity of Oil and Gas Producing Formations. J. Can. Petrol. Technol. 2002, 41,
44 (11), 29.
45 (2) Civan, F. Reservoir Formation Damage. Gulf Professional Publishing, 2011.
46
(3) Adams, J. J. Asphaltene Adsorption, a Literature Review. Energy Fuels 2014, 28, (5), 2831.
47
48 (4) Leontaritis, K. J.; Amaefule, J. O.; Charles, R. E. A Systematic Approach for the Prevention
49 and Treatment of Formation Damage Caused by Asphaltene Deposition. SPE Prod. Facil. 1994, 9,
50 (03), 157.
51 (5) Slobod, R. L.; Thomas, R. A. Effect of Transverse Diffusion on Fingering in Miscible-Phase
52 Displacement. Soc. Petrol. Eng. J. 1963, 3, (01), 9.
53 (6) Homsy, G. M. Viscous Fingering in Porous Media. Annu. Rev. Fluid Mech. 1987, 19, (1), 271.
54
(7) Mullins, O. C. The Asphaltenes. Annu. Rev. Anal. Chem. 2011, 4, 393.
55
56
57
58
59
60
30
ACS Paragon Plus Environment
Page 31 of 37 Industrial & Engineering Chemistry Research

1
2
3 (8) Mullins, O. C.; Sabbah, H.; Eyssautier, J. l.; Pomerantz, A. E.; Barré, L.; Andrews, A. B.; Ruiz-
4
Morales, Y.; Mostowfi, F.; McFarlane, R.; Goual, L. Advances in Asphaltene Science and the Yen–
5
6 Mullins Model. Energy Fuels 2012, 26, (7), 3986.
7 (9) Nassar, N. N. Asphaltene Adsorption onto Alumina Nanoparticles: Kinetics and
8 Thermodynamic Studies. Energy Fuels 2010, 24, (8), 4116.
9 (10) Mullins, O. C. The Modified Yen Model†. Energy Fuels 2010, 24, (4), 2179.
10 (11) Ghanavati, M.; Shojaei, M.-J.; SA, A. R. Effects of Asphaltene Content and Temperature on
11 Viscosity of Iranian Heavy Crude Oil: Experimental and Modeling Study. Energy Fuels 2013, 27,
12
(12), 7217.
13
14 (12) Rogel, E. Studies on Asphaltene Aggregation Via Computational Chemistry. Colloids Surf. A
15 1995, 104, (1), 85.
16 (13) Rogel, E. Thermodynamic Modeling of Asphaltene Aggregation. Langmuir 2004, 20, (3),
17 1003.
18 (14) Murgich, J. Molecular Simulation and the Aggregation of the Heavy Fractions in Crude Oils.
19 Mol. Simul. 2003, 29, (6-7), 451.
20
21 (15) Rahmani, N. H.; Dabros, T.; Masliyah, J. H. Evolution of Asphaltene Floc Size Distribution in
22 Organic Solvents under Shear. Chem. Eng. Sci. 2004, 59, (3), 685.
23 (16) Rahmani, N. H.; Masliyah, J. H.; Dabros, T. Characterization of Asphaltenes Aggregation
24 and Fragmentation in a Shear Field. AIChE J. 2003, 49, (7), 1645.
25 (17) Pacheco-Sánchez, J.; Zaragoza, I.; Martínez-Magadán, J. Asphaltene Aggregation under
26 Vacuum at Different Temperatures by Molecular Dynamics. Energy Fuels 2003, 17, (5), 1346.
27
(18) Boek, E. S.; Ladva, H. K.; Crawshaw, J. P.; Padding, J. T. Deposition of Colloidal Asphaltene
28
29 in Capillary Flow: Experiments and Mesoscopic SimulaƟon†. Energy Fuels 2008, 22, (2), 805.
30 (19) Montoya, T.; Coral, D.; Franco, C. A.; Nassar, N. N.; Cortés, F. B. A Novel Solid–Liquid
31 Equilibrium Model for Describing the Adsorption of Associating Asphaltene Molecules onto Solid
32 Surfaces Based on the “Chemical Theory”. Energy Fuels 2014, 28, (8), 4963.
33 (20) Nassar, N. N.; Tatiana Montoya, T.; Franco, C. A.; Cortés, F. B.; Pereira-Almao, P. R. A New
34 Model for Describing the Adsorption of Asphaltenes on Porous Media at a High Pressure and
35
36
Temperature under Flow Conditions. Energy Fuels 2015 29, (7), 4210.
37 (21) Khoshandam, A.; Alamdari, A. Kinetics of Asphaltene Precipitation in a Heptane− Toluene
38 Mixture. Energy Fuels 2010, 24, (3), 1917.
39 (22) Rahimi, H.; Solaimany Nazar, A. R. Asphaltene Aggregates Fractal Restructuring Model, a
40 Population Balance Approach. Energy Fuels 2009, 24, (2), 1088.
41 (23) Maqbool, T.; Raha, S.; Hoepfner, M. P.; Fogler, H. S. Modeling the Aggregation of
42
Asphaltene Nanoaggregates in Crude Oil− Precipitant Systems. Energy Fuels 2011, 25, (4), 1585.
43
44 (24) Solaimany-Nazar, A. R.; Rahimi, H. Dynamic Determination of Asphaltene Aggregate Size
45 Distribution in Shear Induced Organic Solvents. Energy Fuels 2008, 22, (5), 3435.
46 (25) Eskin, D.; Ratulowski, J.; Akbarzadeh, K.; Pan, S. Modelling Asphaltene Deposition in
47 Turbulent Pipeline Flows. Can. J. Chem. Eng. 2011, 89, (3), 421.
48 (26) Hashemi, R.; Nassar, N. N.; Almao, P. P. Nanoparticle Technology for Heavy Oil in-Situ
49 Upgrading and Recovery Enhancement: Opportunities and Challenges. Appl. Energy 2014, 133,
50
374.
51
52 (27) Hashemi, R.; Nassar, N. N.; Pereira-Almao, P. Transport Behavior of Multimetallic
53 Ultradispersed Nanoparticles in an Oil-Sands-Packed Bed Column at a High Temperature and
54 Pressure. Energy Fuels 2012, 26, (3), 1645.
55 (28) Nassar, N. N.; Hassan, A.; Pereira-Almao, P. Application of Nanotechnology for Heavy Oil
56 Upgrading: Catalytic Steam Gasification/Cracking of Asphaltenes. Energy Fuels 2011, 25, (4), 1566.
57
58
59
60
31
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 32 of 37

1
2
3 (29) Nassar, N. N.; Hassan, A.; Pereira-Almao, P. Metal Oxide Nanoparticles for Asphaltene
4
Adsorption and Oxidation. Energy Fuels 2011, 25, (3), 1017.
5
6 (30) Nassar, N. N.; Hassan, A.; Carbognani, L.; Lopez-Linares, F.; Pereira-Almao, P. Iron Oxide
7 Nanoparticles for Rapid Adsorption and Enhanced Catalytic Oxidation of Thermally Cracked
8 Asphaltenes. Fuel 2012, 95, 257.
9 (31) Franco, C. A.; Montoya, T.; Nassar, N. N.; Pereira-Almao, P.; Cortés, F. B. Adsorption and
10 Subsequent Oxidation of Colombian Asphaltenes onto Nickel and/or Palladium Oxide Supported
11 on Fumed Silica Nanoparticles. Energy Fuels 2013, 27, (12), 7336.
12
(32) Hashemi, R.; Nassar, N. N.; Pereira Almao, P. In Situ Upgrading of Athabasca Bitumen Using
13
14 Multimetallic Ultradispersed Nanocatalysts in an Oil Sands Packed-Bed Column: Part 1. Produced
15 Liquid Quality Enhancement. Energy Fuels 2013, 28, (2), 1338.
16 (33) Hashemi, R.; Nassar, N. N.; Pereira Almao, P. Enhanced Heavy Oil Recovery by in Situ
17 Prepared Ultradispersed Multimetallic Nanoparticles: A Study of Hot Fluid Flooding for Athabasca
18 Bitumen Recovery. Energy Fuels 2013, 27, (4), 2194.
19 (34) Franco, C. A.; Nassar, N. N.; Montoya, T.; Cortés, F. B. NiO and PdO Supported on Fumed
20
21 Silica Nanoparticles for Adsorption and Catalytic Steam Gasification of Colombian C7 Asphaltenes.
22 In Handbook on Oil Production Research; Ambrosio, J., Ed. Nova Science Publishers: 2014; pp 101-
23 145.
24 (35) Nassar, N. N.; Hassan, A.; Vitale, G. Comparing Kinetics and Mechanism of Adsorption and
25 Thermo-Oxidative Decomposition of Athabasca Asphaltenes onto TiO2, ZrO2, and CeO2
26 Nanoparticles. Appl. Catal. A Gen. 2014, 484, 161.
27
(36) Hashemi, R.; Nassar, N. N.; Pereira Almao, P. In Situ Upgrading of Athabasca Bitumen Using
28
29 Multimetallic Ultradispersed Nanocatalysts in an Oil Sands Packed-Bed Column: Part 2. Solid
30 Analysis and Gaseous Product Distribution. Energy Fuels 2014, 28, (2), 1351.
31 (37) Franco, C. A.; Nassar, N. N.; Montoya, T.; Ruíz, M. A.; Cortés, F. B. Influence of Asphaltene
32 Aggregation on the Adsorption and Catalytic Behavior of Nanoparticles. Energy Fuels 2015, 29, (3),
33 1610.
34 (38) Nassar, N. N.; Franco, C. A.; Montoya, T.; Cortés, F. B.; Hassan, A. Effect of Oxide Support
35
36
on Ni–Pd Bimetallic Nanocatalysts for Steam Gasification of Nc 7 Asphaltenes. Fuel 2015, 156, 110.
37 (39) Jimenez, M. A.; Genolet, L. C.; Chavez, J. C.; Espin, D. Method for Treating Drilling Fluid
38 Using Nanoparticles. U.S. Patent 6,579,832, 2003.
39 (40) Sensoy, T.; Chenevert, M. E.; Sharma, M. M. Minimizing Water Invasion in Shales Using
40 Nanoparticles. Conference Paper SPE-124429-MS. Society of Petroleum Engineers, 2009. ISBN
41 978-1-55563-263-2.
42
(41) Zakaria, M.; Husein, M. M.; Harland, G. Novel Nanoparticle-Based Drilling Fluid with
43
44 Improved Characteristics. Conference Paper SPE-156992-MS. Society of Petroleum Engineers,
45 2012. ISBN 978-1-61399-206-7.
46 (42) Srivatsa, J. T.; Ziaja, M. B. An Experimental Investigation on Use of Nanoparticles as Fluid
47 Loss Additives in a Surfactant-Polymer Based Drilling Fluids. Conference Paper IPTC-14952-MS.
48 International Petroleum Technology Conference (Publisher), 2011. ISBN 978-1-61399-148-0.
49 (43) Shen, D.; Zhang, P.; Kan, A. T.; Fu, G.; Alsaiari, H. A.; Tomson, M. B. Control Placement of
50
Scale Inhibitors in the Formation with Stable Ca-Dtpmp Nanoparticle Suspension and Its Transport
51
52 in Porous Medium. Conference Paper SPE-114063-MS. Society of Petroleum Engineers, 2008. ISBN
53 978-1-55563-169-7.
54 (44) Betancur-Márquez, S.; Alzate-Espinosa, G. A.; Cortés, F. B. Optimization of Drilling Fluids
55 Using Functionalized Nanoparticles: Loss Filtration Reduction and Thickness Mudcake. Boletín de
56 Ciencias de la Tierra 2014, 35, 5.
57
58
59
60
32
ACS Paragon Plus Environment
Page 33 of 37 Industrial & Engineering Chemistry Research

1
2
3 (45) Sengupta, S.; Kumar, A. Nano-Ceramic Coatings-a Means of Enhancing Bit Life and
4
Reducing Drill String Trips. Conference Paper IPTC-16474-MS. International Petroleum Technology
5
6 Conference (Publisher), 2013. ISBN 978-1-61399-218-0.
7 (46) Chakraborty, S.; Agrawal, G.; DiGiovanni, A.; Scott, D. E. The Trick Is the Surface-
8 Functionalized Nanodiamond PDC Technology. Conference Paper SPE-157039-MS. Society of
9 Petroleum Engineers, 2012. ISBN 978-1-61399-206-7.
10 (47) Huang, T.; Crews, J. B.; Willingham, J. R. Using Nanoparticle Technology to Control Fine
11 Migration. Conference Paper SPE-115384-MS. Society of Petroleum Engineers, 2008. ISBN 978-1-
12
55563-147-5.
13
14 (48) Mera, C.A.; Franco, C.A.; Cortés, F. B. Uso De Nanopartículas De Sílice Para La
15 Estabilización De Finos En Lechos Empacados De Arena Ottawa. Revista Informador Técnico 2013,
16 77, (1), 27.
17 (49) Ogolo, N. C.; Olafuyi, O. A.; Onyekonwu, M. Effect of Nanoparticles on Migrating Fines in
18 Formations. Conference Paper SPE-155213-MS. Society of Petroleum Engineers, 2012. ISBN 978-1-
19 61399-206-7.
20
21 (50) Ogolo, N.; Olafuyi, O.; Onyekonwu, M. Impact of Hydrocarbon on the Performance of
22 Nanoparticles in Control of Fines Migration. Conference Paper SPE-167503-MS. Society of
23 Petroleum Engineers, 2013. ISBN 978-1-61399-301-9.
24 (51) Yang, F.; Paso, K.; Norrman, J.; Li, C.; Oschmann, H.; Sjöblom, J. Hydrophilic Nanoparticles
25 Facilitate Wax Inhibition. Energy Fuels 2015, 29, (3), 1368.
26 (52) Wang, F.; Zhang, D.; Ding, Y.; Zhang, L.; Yang, M.; Jiang, B.; Zhang, S.; Ai, M.; Liu, G.; Zhi, S.
27
The Effect of Nanohybrid Materials on the Pour-Point and Viscosity Depressing of Waxy Crude Oil.
28
29 Chin. Sci. Bull. 2011, 56, (1), 14.
30 (53) Tung, N. P.; Phong, N. T. P.; Long, B. Q. K.; Duy, N. H. Scale Inhibitors for Co-Deposited
31 Calcium Sulfate and Calcium Carbonate in Squeeze Process in White Tiger Oilfield. Conference
32 Paper SPE-87467-MS. Society of Petroleum Engineers, 2004. ISBN 978-1-55563-985-3.
33 (54) Safari, M.; Golsefatan, A.; Jamialahmadi, M. Inhibition of Scale Formation Using Silica
34 Nanoparticle. J. Disp. Sci. Technol. 2014, 35, (10), 1502.
35
36
(55) Zhang, P.; Shen, D.; Fan, C.; Kan, A.; Tomson, M. Surfactant-Assisted Synthesis of Metal-
37 Phosphonate Inhibitor Nanoparticles and Transport in Porous Media. Journal Paper SPE-121552-
38 PA. Society of Petroleum Engineers, 2010. ISSN 1086-055X.
39 (56) Franco, C. A.; Cortés, F. B.; Nassar, N. N. Adsorptive Removal of Oil Spill from Oil-in-Fresh
40 Water Emulsions by Hydrophobic Alumina Nanoparticles Functionalized with Petroleum Vacuum
41 Residue. J. Colloid Interf. Sci. 2014, 425, 168.
42
(57) Franco, C. A.; Nassar, N. N.; Cortés, F. B. Removal of Oil from Oil-in-Saltwater Emulsions by
43
44 Adsorption onto Nano-Alumina Functionalized with Petroleum Vacuum Residue. J. Colloid Interf.
45 Sci. 2014, 433, 58.
46 (58) Barati, R.; Johnson, S. J.; McCool, S.; Green, D. W.; Willhite, G. P.; Liang, J. T. Fracturing
47 Fluid Cleanup by Controlled Release of Enzymes from Polyelectrolyte Complex Nanoparticles. J.
48 App. Polym. Sci. 2011, 121, (3), 1292.
49 (59) Huang, T.; Crews, J. B. Nanotechnology Applications in Viscoelastic Surfactant Stimulation
50
Fluids. Journal Paper SPE-107728-PA. Society of Petroleum Engineers, 2008. ISSN 1930-1855.
51
52 (60) Franco, C. A.; Nassar, N. N.; Ruiz, M. A.; Pereira-Almao, P.; Cortés, F. B. Nanoparticles for
53 Inhibition of Asphaltenes Damage: Adsorption Study and Displacement Test on Porous Media.
54 Energy Fuels 2013, 27, (6), 2899.
55 (61) Mohammadi, M.; Akbari, M.; Fakhroueian, Z.; Bahramian, A.; Azin, R.; Arya, S. Inhibition of
56 Asphaltene Precipitation by Tio2, Sio2, and Zro2 Nanofluids. Energy Fuels 2011, 25, (7), 3150.
57
58
59
60
33
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 34 of 37

1
2
3 (62) Hashemi, S. I.; Fazelabdolabadi, B.; Moradi, S.; Rashidi, A. M.; Shahrabadi, A.; Bagherzadeh,
4
H. On the Application of NiO Nanoparticles to Mitigate in Situ Asphaltene Deposition in Carbonate
5
6 Porous Matrix. Appl. Nanosci. 2015, 1.
7 (63) Kazemzadeh, Y.; Malayeri, M.; Riazi, M.; Parsaei, R. Impact of Fe3O4 Nanoparticles on
8 Asphaltene Precipitation During CO2 Injection. J. Nat. Gas Sci. Eng. 2015, 22, 227.
9 (64) Zabala, R.; Mora, E.; Botero, O.; Cespedes, C.; Guarin, L.; Franco, C.; Cortes, F.; Patino, J.;
10 Ospina, N. Nano-Technology for Asphaltenes Inhibition in Cupiagua South Wells. J. Petrol. Technol.
11 2014, 66, (2), 117.
12
(65) Franco, C.; Patiño, E.; Benjumea, P.; Ruiz, M. A.; Cortés, F. B. Kinetic and Thermodynamic
13
14 Equilibrium of Asphaltenes Sorption onto Nanoparticles of Nickel Oxide Supported on
15 Nanoparticulated Alumina. Fuel 2013, 105, 408.
16 (66) Ramalho, J. B. V.; Lechuga, F. C.; Lucas, E. F. Effect of the Structure of Commercial Poly
17 (Ethylene Oxide-B-Propylene Oxide) Demulsifier Bases on the Demulsification of Water-in-Crude
18 Oil Emulsions: Elucidation of the Demulsification Mechanism. Química Nova 2010, 33, (8), 1664.
19 (67) Yudin, I. K.; Anisimov, M. A. Dynamic Light Scattering Monitoring of Asphaltene
20
21 Aggregation in Crude Oils and Hydrocarbon Solutions; In Asphaltenes, Heavy Oils, and
22 Petroleomics; Mullins, O.C., Sheu, E.Y., Hammami, A., Marshall, A.G. , Eds.; Springer: 2007; pp 439.
23 (68) Rane, J. P.; Harbottle, D.; Pauchard, V.; Couzis, A.; Banerjee, S. Adsorption Kinetics of
24 Asphaltenes at the Oil–Water Interface and Nanoaggregation in the Bulk. Langmuir 2012, 28, (26),
25 9986.
26 (69) Asomaning, S. Test Methods for Determining Asphaltene Stability in Crude Oils. Petrol. Sci.
27
Technol. 2003, 21, (3-4), 581.
28
29 (70) Oliensis, G. L. THE OLIENSIS SPOT TEST--WHAT JUSTIFICATION IS THERE FOR ITS USE?
30 00932284, Association of Asphalt Paving Technologists, St. Paul, MN, USA, 1935, 88-93.
31 (71) Berne, B.; Pecora, R. Dynamic Light Scattering with Applications to Chemistry. Biology and
32 Physics; Wiley: New York, 1976.
33 (72) Austin, L. G. Introduction to the Mathematical Description of Grinding as a Rate Process.
34 Powder Technol. 1971, 5, (1), 1.
35
36
(73) Friedlander, S. K. Smoke, Dust and Haze: Fundamentals of Aerosol Behavior; Wiley-
37 Interscience: New York, 1977.
38 (74) Hounslow, M.; Ryall, R.; Marshall, V. A Discretized Population Balance for Nucleation,
39 Growth, and Aggregation. AIChE J. 1988, 34, (11), 1821.
40 (75) Jindal, P.; Austin, L. A Review Introduction to the Mathematical Description of Grinding as
41 a Rate Process. Powder Tech 1976, 5, 1.
42
(76) Elimelech, M.; Jia, X.; Gregory, J.; Williams, R. Particle Deposition & Aggregation:
43
44 Measurement, Modelling and Simulation; Butterworth-Heinemann: Oxford, 1998.
45 (77) Smoluchowski, M. Versuch Einer Mathematischen Theorie Der Koagulationskinetik
46 Kolloider Lösungen. Zeitschrift für physikalische Chemie 1917, 92, (2), 129.
47 (78) Kusters, K. A.; Pratsinis, S. E.; Thoma, S. G.; Smith, D. M. Ultrasonic Fragmentation of
48 Agglomerate Powders. Chem. Eng. Sci. 1993, 48, (24), 4119.
49 (79) Spicer, P. T.; Pratsinis, S. E. Coagulation and Fragmentation: Universal Steady-State
50
Particle-Size Distribution. AIChE J. 1996, 42, (6), 1612.
51
52 (80) Serra, T.; Casamitjana, X. Effect of the Shear and Volume Fraction on the Aggregation and
53 Breakup of Particles. AIChE J. 1998, 44, (8), 1724.
54 (81) Higashitani, K.; Iimura, K.; Sanda, H. Simulation of Deformation and Breakup of Large
55 Aggregates in Flows of Viscous Fluids. Chem. Eng. Sci. 2001, 56, (9), 2927.
56 (82) Burban, P. Y.; Lick, W.; Lick, J. The Flocculation of Fine-Grained Sediments in Estuarine
57 Waters. J. Geophys. Res. Oceans 1989, 94, (C6), 8323.
58
59
60
34
ACS Paragon Plus Environment
Page 35 of 37 Industrial & Engineering Chemistry Research

1
2
3 (83) Boadway, J. D. Dynamics of Growth and Breakage of Alum Floc in Presence of Fluid Shear.
4
J. Env. Eng. Div. 1978, 104, (5), 901.
5
6 (84) Kapur, P. Self-Preserving Size Spectra of Comminuted Particles. Chem. Eng. Sci. 1972, 27,
7 (2), 425.
8 (85) Pandya, J.; Spielman, L. Floc Breakage in Agitated Suspensions: Effect of Agitation Rate.
9 Chem. Eng. Sci. 1983, 38, (12).
10 (86) Wilczak, A.; Keinath, T. M. Kinetics of Sorption and Desorption of Copper (Ii) and Lead (Ii)
11 on Activated Carbon. Water Environ. Res. 1993, 65, (3), 238.
12
(87) Chiron, N.; Guilet, R.; Deydier, E. Adsorption of Cu (Ii) and Pb (Ii) onto a Grafted Silica:
13
14 Isotherms and Kinetic Models. Water Res. 2003, 37, (13), 3079.
15 (88) Shayan, N. N.; Mirzayi, B. Adsorption and Removal of Asphaltene Using Synthesized
16 Maghemite and Hematite Nanoparticles. Energy Fuels 2015, 29, (3), 1397.
17 (89) Yarranton, H.W. Asphaltene Solubility and Asphaltene Stabilized Water-in-Oil Emulsions.
18 Ph.D. Dissertation, University of Alberta, Edmonton, Alberta, Canada, 1994.
19 (90) Dennis, J., John E; Moré, J. J. Quasi-Newton Methods, Motivation and Theory. SIAM Rev.
20
21 1977, 19, (1), 46.
22 (91) Cameron, I. T.; Hangos, K. Process Modelling and Model Analysis; Academic Press, 2001.
23 (92) Nassar, N. N. Kinetics, Equilibrium and Thermodynamic Studies on the Adsorptive Removal
24 of Nickel, Cadmium and Cobalt from Wastewater by Superparamagnetic Iron Oxide
25 Nanoadsorbents. Can. J. Chem. Eng. 2012, 90, (5), 1231.
26 (93) Acevedo, S.; Castillo, J.; Fernández, A.; Goncalves, S.; Ranaudo, M. A. A Study of Multilayer
27
Adsorption of Asphaltenes on Glass Surfaces by Photothermal Surface Deformation. Relation of
28
29 This Adsorption to Aggregate Formation in Solution. Energy Fuels 1998, 12, (2), 386.
30 (94) Spiecker, P. M.; Gawrys, K. L.; Trail, C. B.; Kilpatrick, P. K. Effects of Petroleum Resins on
31 Asphaltene Aggregation and Water-in-Oil Emulsion Formation. Colloids Surf., A 2003, 220, (1), 9.
32 (95) Spiecker, P. M.; Gawrys, K. L.; Kilpatrick, P. K. Aggregation and Solubility Behavior of
33 Asphaltenes and Their Subfractions. J. Colloid Interface Sci. 2003, 267, (1), 178.
34 (96) Pernyeszi, T.; Patzko, A.; Berkesi, O.; Dékány, I. Asphaltene Adsorption on Clays and Crude
35
36
Oil Reservoir Rocks. Colloids Surf., A 1998, 137, (1), 373.
37 (97) Anisimov, M. A.; Ganeeva, Y. M.; Gorodetskii, E.; Deshabo, V.; Kosov, V.; Kuryakov, V.;
38 Yudin, D.; Yudin, I. Effects of Resins on Aggregation and Stability of Asphaltenes. Energy Fuels
39 2014, 28, (10), 6200.
40 (98) Lu, Y.; Yin, Y.; Mayers, B. T.; Xia, Y., Modifying the Surface Properties of Superparamagnetic
41 Iron Oxide Nanoparticles through a Sol-Gel Approach. Nano Lett. 2002, 2, (3), 183.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
35
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 36 of 37

1
2
3
4
5
6
7
8
9
10
11 List of figure captions.
12
13 Figure 1. FESEM images and the corresponding particle size probability distribution for a) and b)
14 magnetite, c) and d) γ-alumina and e) and f) silica nanoparticles. ...................................................... 8
15 Figure 2. Adsorption kinetics for AK18 n-C7 asphaltenes onto the selected nanoparticles from
16
17
solutions of a) Heptol 20 and b) Heptol 40. Temperature: 293 K; adsorbent dose, 10 g/L.
18 Continuous curves represent double exponential model fitting (eq. 7). ............................................ 18
19 Figure 3. Adsorption kinetics for AK18 and Cedral n-C7 asphaltenes from Heptol 40 on magnetite
20 nanoparticles. Temperature: 293 K; adsorbent dose, 10 g/L. Continuous curves represent double
21 exponential model fitting (eq. 7). ...................................................................................................... 21
22
23
Figure 4. Kinetics of virgin AK18 n-C7 asphaltenes aggregation-fragmentation at a temperature of
24 293 K in Heptol 20 and Heptol 40 solutions. The symbols are experimental data, and the solid lines
25 are from the PBM model (eq 3). ....................................................................................................... 24
26 Figure 5. Kinetics of asphaltenes aggregation-fragmentation in the presence of nanoparticles for a)
27 Heptol 20 and b) Heptol 40 at 293 K. The symbols are experimental data, and the solid lines are
28
from the PBM model (eq. 3). ............................................................................................................ 27
29
30 Figure 6. Kinetics of aggregation-fragmentation of a) AK18 n-C7 asphaltenes and b) Cedral n-C7 in
31 the absence and presence of magnetite nanoparticles at 293 K and for Heptol 40 solutions. The
32 symbols are experimental data, and the solid lines are from the PBM model (eq. 3). ...................... 29
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
36
ACS Paragon Plus Environment
Page 37 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5 TOC Graphic
6
7
8
9 1200
10
Mean asphaltene aggregate diameter (nm)

11
12
13 1000
14
15
16
17 800
18
19
20
21 600
22
23
24
25 400
26
27
28 Virgin n-C7 asphaltenes in Heptol 40
29 200 n-C7 asphaltenes with magnetite nanoparticles in Heptol 40
30 Virgin n-C7 asphaltenes in Heptol 20
31 n-C7 asphaltenes with magnetite nanoparticles in Heptol 20
32 PBM
33 0
34
35
0 100 200 300
36 Time (min)
37
38
39
40
41
42
43
44
45
46
37
ACS Paragon Plus Environment
47
48

You might also like