Next Article in Journal
Aqueous Two-Phase Extraction, Antioxidant and Renal Protective Effects of Polysaccharides from Spores of Cordyceps cicadae
Previous Article in Journal
Tuning of Model Predictive Controllers Based on Hybrid Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Decomposition of n-C7 Asphaltenes Using Tungsten Oxides–Functionalized SiO2 Nanoparticles in Steam/Air Atmospheres

1
Grupo de Investigación en Fenómenos de Superficie–Michael Polanyi, Facultad de Minas, Universidad Nacional de Colombia Sede-Medellín, Cra. 80 No. 65-223, Medellín 050034, Colombia
2
Laboratorio de Investigación en Catálisis y Procesos, Departamento de Química, Universidad del Valle, Cali 760001, Colombia
3
Química de Recursos Energéticos y Medio Ambiente, Instituto de Química, Universidad de Antioquia UdeA, Calle 70 No. 52-21, Medellín 050010, Colombia
4
Departamento de Ciencia de los Materiales, Ingeniería Metalúrgica y Química Inorgánica, Facultad de Ciencias, Universidad de Cádiz, Campus Río San Pedro, 11510 Puerto Real, Spain
*
Authors to whom correspondence should be addressed.
Processes 2022, 10(2), 349; https://doi.org/10.3390/pr10020349
Submission received: 19 January 2022 / Revised: 4 February 2022 / Accepted: 5 February 2022 / Published: 11 February 2022
(This article belongs to the Section Energy Systems)

Abstract

:
A wide range of technologies are being developed to increase oil recovery, reserves, and perform in situ upgrading of heavy crude oils. In this study, supported tungsten oxide nanoparticles were synthesized, characterized, and evaluated for adsorption and catalytic performance during wet in situ combustion (6% of steam in the air, in volumetric fraction) of n-C7 asphaltenes. Silica nanoparticles of 30 nm in diameter were synthesized using a sol–gel methodology and functionalized with tungsten oxides, using three different concentrations and calcination temperatures: 1%, 3%, 5% (mass fraction), and 350 °C, 450 °C, and 650 °C, respectively. Equilibrium batch adsorption experiments were carried out at 25 ℃ with model solutions of n-C7 asphaltenes diluted in toluene at concentrations from 100 mg·L−1 to 2000 mg·L−1, and catalytic wet in situ combustion of adsorbed heavy fractions was carried out by thermogravimetric analysis coupled to FT-IR. The results showed improvements of asphaltenes decomposition by the action of the tungsten oxide nanoparticles due to the reduction in the decomposition temperature of the asphaltenes up to 120 °C in comparison with the system in the absence of WOX nanoparticles. Those synthesis parameters, such as temperature and impregnation dosage, play an important role in the adsorptive and catalytic activity of the materials, due to the different WOX–support interactions as were found through XPS. The mixture released during the catalyzed asphaltene decomposition in the wet air atmosphere reveals an increase in light hydrocarbons, methane, and hydrogen content. Hydrogen production was prioritized between 300 and 400 °C where, similarly, the reduction of CO, CH4, and the increase in CO2 content, associated with water–gas shift, and methane reforming reactions occur, respectively. The results show that these catalysts can be used either for in situ upgrading of crude oil, or any application where heavy fractions must be transformed.

1. Introduction

The growth of the population, as well as the economic and technological developments, imply a continuous increase in the global demand for energy [1]. Consequently, unconventional crude oils, such as heavy crude oil (HO) and extra heavy crude oil (EHO), among others, have gained significant importance [2,3,4]. These types of crude are characterized by high carbon–hydrogen ratios, and a lot of asphaltenes (2% to 20%, in mass fraction), high viscosity, and low American Petroleum Institute gravity (°API) [5].
Asphaltenes correspond to the fraction of the oil-soluble in aromatic hydrocarbons but are insoluble in alkanes [6]. Likewise, the asphaltenes contain plane polyaromatic hydrocarbons (PAH) with polar functional groups (mainly containing S, N, and O), metalloporphyrins of Ni and V, as well as aliphatic side chains [7,8,9]. The polycyclic nuclei are polarizable and are partially charged due mainly to the presence of heteroatoms, generating interactions of the types dipole–dipole, induced dipole–induced dipole, and π stacking [10]. In contrast, the aliphatic (peripheral) components avoid the proximity of the molecules because they generate steric hindrance [8,11,12,13]. The equilibrium between forces of attraction and repulsion gives rise to the tendency of aggregation in the asphaltenes, reaching to produce asphaltene aggregates of 5 nm–20 nm [14]. The pitfalls related to asphaltenes presence translate into higher economic expense and environmental impact [15]. Removing asphaltenes from crude oil would help to improve the crude oil quality and consequently reduce the costs of crude oil production and transportation, also generating a lower environmental impact [16]. Accordingly, several in situ technologies have been used for improving HO and EHO production to reduce the viscosity of the oil [17]. Procedures such as solvent injection, steam, and gas injection, as well as mechanical or thermal methods, are commonly used [14,18,19,20,21,22]. Among the thermal methods, wet in situ combustion is of interest as water is co-injected or alternately injected with the air to improve the heat consumption in the reservoir [23,24,25,26]. Wet in situ combustion is also called in situ steam generation or a combination of in situ combustion and waterflooding, where the heat is transferred to the oil bank front from the water injected [27]. During this process, different reactions occur simultaneously, generating physicochemical changes in the heavy oil, even leading to upgrading through viscosity reduction and API gravity increase. Nevertheless, the conversion of heavy and refractory fractions such as asphaltenes into lighter compounds struggles to occur behind the combustion zone, as the decomposition temperature could be higher than 400 °C [28,29,30,31].
The development of nanotechnology has allowed the oil and gas industry to be included among the broad applications of this phenomenon [32,33]. Adsorption of asphaltenes on different nanoparticles and the subsequent decomposition is considered as a viable and environmentally friendly solution for the development and recovery of HO and EHO [16,28,34,35,36,37,38,39]. Transition element oxides (TEO) provide a determinate characteristic that can be exploited to obtain selective adsorption of asphaltenes besides a better performance as catalysts for decomposition of them [39] and, at the same time, reducing problems associated with mobility, transport, production, and recovery [40,41]. Recently, supported and unsupported transition element oxide nanoparticles have been employed for asphaltene adsorption, and subsequently gasification [16,42,43,44,45], oxidation [28,36,46,47,48,49,50], and pyrolysis [35,51] as a new approach for the in situ upgrading of HO and EHO, providing useful information about effects of TEO and supported nanoparticles [16,35,36,46,52]. It is necessary to keep in mind that solid acid catalysts based on nickel or tungsten oxides clusters have shown to be active for dehydrogenation, isomerization, reforming, cracking, alcoholic dehydration, and olefin oligomerization reactions [52], in addition to the removal of sulfur and nitrogen atoms [53] correlated to the cracking reactions. On the other hand, the catalytic properties of a material are strongly related to their composition and the surface structure, which, in turn, can be determined by the conditions of synthesis [54,55], obtaining variations in the surface area, dispersion of the active phase, the resistance of acid sites, and effectiveness and selectivity in oxidative catalysis [55,56].
However, to the best of our knowledge, there are no studies related to the use of nanocatalysts in a wet in situ combustion atmosphere. Hence, this study employed, for the first time, nanomaterials based on tungsten oxides supported over silica nanoparticles and evaluated the effect of variations on synthesis method, such as the concentration of active phase precursor and calcination temperature, over the performance of the nanoparticles for adsorption and catalytic decomposition of asphaltenes in wet in situ combustion atmosphere. In this sense, the synthesis was evaluated using incipient wetness technique, using three different concentrations and calcination temperatures: 1%, 3%, and 5% (mass fraction), and 350 °C, 450 °C, and 650 °C, respectively. The nanomaterials were physical, and chemically characterized by surface area (SBET), dynamic light scattering technique (DLS), X-ray photoelectron spectroscopy (XPS), and H2 pulse chemisorption. The adsorption was conducted in batch–mode experiments at 25 °C and the wet combustion decomposition was carried out by thermogravimetric analysis coupled to a mass spectrometer.

2. Experimental Section

2.1. Materials

Silica nanoparticles used in this study were synthesized using HCl, hexadecyltrimethylammonium bromide (98.0%, Panreac, Spain), deionized water, and sodium silicate (Protokimica, Colombia) as a precursor. Na2WO4∙2H2O and (NH4)10H2(WO7)6 (Sigma-Aldrich, St. Louis, MO, USA) were used for the functionalization of nanoparticles. N-heptane (99%, Sigma-Aldrich, St. Louis, MO, USA) and toluene (99.5%, Merck GaG, Germany) were used for asphaltene isolation from crude oil and as a solvent on heavy oil model solutions, respectively.

2.2. Methods

2.2.1. Asphaltene Isolation and Characterization

Asphaltenes were isolated from a heavy crude oil (6.4° API) with an approximately contain asphaltenes of 13% (mass fraction), obtained from a reservoir located in southeast Colombia. Asphaltenes were extracted from the heavy crude oil sample by mixing each crude oil in a 40:1 ratio with n-heptane following the ASTM D2892 and ASTM D5236 [57,58,59]. An excess of n-heptane was added to the crude oil in a volume ratio of 1:40 (g·mL−1). The mixture was then sonicated for 2 h at 25 °C and further stirred at 300 rpm for 20 h; then it was centrifuged for 30 min at 4500 rpm. The precipitated asphaltenes were filtered and washed with n-heptane until obtaining a shiny black tone. Details of the chemical and physical properties of the asphaltenes can be found in previous work [29].

2.2.2. Nanoparticles Preparation

The synthesis method used for obtaining SiO2 nanoparticles was developed in previous works [60,61]. Some modifications to the protocol were made with the purpose of obtaining particle sizes that adjust to the IUPAC definition of nanoparticle (1 nm–100 nm) [62]. The procedure consists of the preparation of a silica suspension from an aqueous solution of sodium silicate (24.5% volumetric fraction, silicate/water) by the addition of hydrochloric acid (1 M) through the precipitation method under constant stirring. A sodium silicate solution (SSS) was prepared and heated until adjusting the temperature to 55 °C; using a constant stirring speed of 300 rpm, cetyltrimethylammonium bromide was dissolved until obtaining a concentration of 10% (in volumetric fraction) with respect to the SSS, then, a HCl solution was slowly added. The use of the acid had two purposes: first, to initiate the formation of a gel (pH: 9–9.5); and the second one was to acidify the solution and bring it to a pH of 1–1.5. Between two steps, the solution was stirred for 10 min, with no further addition of HCl, to prevent the formation of a strong silica gel. Then, the gel was aged for 20 h at 55 °C, and the nanoparticles obtained were washed using ethanol. Finally, the silica nanoparticles were dried for 8 h at 120 °C and calcined at 550 °C.

2.2.3. Surface Functionalization

Silica nanoparticles were dried at 120 °C. Sodium tungstate dihydrate Na2WO4∙2H2O and ammonium tungstate (NH4)10H2(W2O7)6 were used. The precursors were dissolved in a volume of water equal to the pore volume of a determined amount of the catalyst support. To determine volume of liquid required, the selected solid silica is dried at room temperature for several weeks, weighed, and immersed in alcohol for 20 min The bubbling observed on the surface of the silica indicates that trapped air is being released. Once saturated, the silica is placed on an alcohol-saturated paper towel and immediately weighed at room temperature (20 ºC). This process is completed in less than 10 s. The internal volume of the silica is calculated by converting the difference in masses of the alcohol-saturated silica and the dry aggregate to volume using the density of the alcohol.
The impregnation procedure in the support was carried out using incipient wet impregnation, which was carried out by adding, drop by drop, the precursor solution to the support until saturation.
Then, the impregnated materials were then dried at 120 °C for six hours, followed by calcination [28,37]. The calcination temperature was varied at 350 °C, 450 °C, and 650 °C separately for 4 h, obtaining the tungsten oxide nanoparticles. The hybrid or supported hygroscopic salt (called SHS) in this study was named using the chemical symbol of the support (i.e., Si) followed by the symbol of tungsten (i.e., W). If the precursor salt was the sodium tungstate dihydrate, the Na symbol was added after W. Then, it was included in the mass fraction percentage of the nominal tungsten used [28]. The number after the space indicates the calcination temperature of the material. For example, the sample SiWNa1 350 corresponds to nanosilica doped with 1% (in mass fraction) of sodium tungstate dihydrate and calcined at 350 °C. Table 1 lists the characteristics of the tungsten functionalized silica nanoparticles (WSN) used in this study.

2.2.4. Nanoparticles Characterization

The silica nanoparticles were characterized using dynamic light scattering technique (DLS) for hydrodynamic diameter by a nanoplus-3 from Micromeritics, USA. Characterization of functional groups on the nanoparticles and WSN surface was performed by Fourier transform infrared spectroscopy (FTIR) with an IRAffinity-1 FTIR spectrophotometer from Shimadzu, Japan. Determination of tungsten dispersion and the average size of crystal on silica nanoparticles or mean particle size of active phase (MSAP) was performed by H2 pulse chemisorption with a Chembet 3000 (Quantachrome Instruments, Boynton Beach, FL, USA) with a high-sensitivity thermal conductivity detector. The procedure is detailed in other work [63]. Quantitative hydrogen pulses (10 μL) were injected until the surface was saturated, and W–H species were formed [64,65,66]. The surface area (SBET) of catalysts were obtained by N2 adsorption/desorption experiments in an Autosorb-1 from Quantachrome [67,68]. The SBET was estimated based on the Brunauer−Emmet−Teller (BET) method which is detailed elsewhere [69]. X-ray photoelectron spectroscopy tests were performed for selected catalysts on a Specs X-ray photoelectronic spectrometer (NAP-XPS) with a PHOIBOS 150 1D-DLD analyzer, using a monochromatic source of Al-Kα (1486.7 eV, 13 kV, 100 W). Details of the experimental setup are found in a previous work [44].

2.2.5. Equilibrium Adsorption Isotherms

A stock solution of 2000 mg·L−1 was prepared by dissolving the extracted asphaltenes in toluene; later, model solutions were made ranging from 100 mg·L−1 to 2000 mg·L−1. Batch adsorption experiments were carried out at 25 °C, adding the nanoparticles to the model solutions at a fixed nanoparticles-to-solution ratio of 100 mg to 10 mL [70]. The mixtures were stirred for 12 h at 300 rpm to ensure equilibrium; then, the WSN with the adsorbed asphaltenes were separated from the asphaltenes in solution by centrifugation. The absorbance of the supernatant was measured using a Genesys 10S UV–Vis spectrophotometer, taking the toluene as a blank. A calibration curve was previously constructed for a wavelength of 295 nm, and, in this way, determining the adsorbed amount of asphaltenes per nanoparticle area surface is determined by the mass balance of the analysis. The amount adsorbed q (mg·m−2) was determined using Equation (1):
q = C 0 C E V A
where C0 (mg·L−1) is the initial concentration of asphaltenes, CE (mg·L−1) is equilibrium concentrations in solution after adsorption, V (L) is volume of the solution, and A (m2∙g−1) is the dry surface area of the employed material. The adsorption isotherms for the asphaltenes onto the nanoparticles were modeled using the solid–liquid equilibrium (SLE) model [38] (see Equations (S1) to (S3) in the support information).

2.2.6. Thermogravimetric Analysis

To evaluate the catalytic behavior of the materials, the thermal stability of the asphaltenes (adsorbate) in the presence and absence of the adsorbent was studied. Thermogravimetric tests were performed on asphaltenes, and nanoparticles obtained after the batch adsorption tests (previously dried) on a TGA Q50 from TA Instruments, Inc, New Castle, DE. The samples were dried for 12 h, and 5 mg of each one was taken for thermogravimetric analysis kept low to circumvent the diffusion limitations [71]. The samples were analyzed on wet combustion atmosphere from 30 °C to 600 °C at three different rates (5 °C·min−1, 10 °C·min−1, and 20 °C·min−1) using a wet airflow (5%, volumetric fraction) rate constant 100 mL·min−1 throughout the experiment. Steam is generated in an evaporator (filled with deionized water) with a temperature controller, which fixes the vapor pressure based on the temperature set point. The wet airflow is obtained when the airflow stream is injected through the evaporator, reaching its saturation at the vapor pressure required [72]. For the adsorbents selected as the best catalysts, their capacity to produce in situ hydrogen and other calorific gases was analyzed via mass spectrometry using a mass spectrometer (Shimadzu GC-MS, Tokyo, Japan). Details of the setup are found in other works [28,73]. The components targeted for the analysis by single-ion monitoring in mass spectrometry were H2 (2 g·mol−1), CO (28 g·mol−1), CO2 (44 g·mol−1), and LHC (C2H4, C2H6, CH4). Reproducibility was ensured, realizing the experiments by triplicate. In addition, the effective activation energies were estimated by the method of Ozawa–Flynn–Wall (OFW). This methodology is extensively found in the literature and reported in previous reports [28,73], with good reproducibility (see Equations (S4)–(S7)).

3. Results and Discussion

3.1. Nanoparticle Characterization

The results obtained for the hydrodynamic diameter of synthesized SiO2 nanoparticles are in the nanometer range with a mean value of 12.5 nm ± 2 nm. This size is suitable for application in oilfields because its size is smaller than pore throats of heavy oil reservoirs [74]. Table 1 shows the properties of the nanoparticles, BET surface area, tungsten dispersion, and average size of the tungsten nanoparticles in the supported catalyst. Each type of nanoparticle exhibited a different chemical nature.
Pulsed H2 chemisorption determines the exposed metallic atoms at the surface, using H2 as a molecule probe, which does not interact significantly with the support material (SiO2), but forms a strong chemical bond to the surface metallic atoms causing the H2 spillover and consequently its consumption. The W surface area and dispersion were calculated, assuming dissociative chemisorption of hydrogen with a stoichiometry of W–H of 1 [65,75]. Results of the tungsten mean particle size under the measurement of reduced particles for the studied samples are also shown in Table 1 and confirm the formation of tungsten nanoparticles (tungsten oxides) over the support surface. It can be observed that by increasing the calcination temperature, the obtained nanoparticles are bigger in size due to the thermally promoted particle growth. Meanwhile, the particle dispersion has an inverse trend: as the calcination temperature is increased, particle dispersion decreases. The tendency obtained for the variation of impregnation dosage used is similar; a greater quantity implies a crystal with larger size, and, in turn, lower dispersion. Hence, the formation of the tungsten oxide nanoparticles over the SiO2 surface is strongly influenced by synthesis conditions.
The obtained SBET values for the nanomaterials are also listed in Table 1. As expected, the sample prepared at 350 °C has the highest surface area of 122.3 m2·g−1. The values of the measured SBET followed the order 350 °C > 450 °C > 650 °C and 1% > 3% > 5%, according to calcination temperature and impregnation dosage, respectively. As the temperature of calcination is raised, the surface area decreases [76]. It is well known that the temperature favors the atomic movements and, therefore, an increase of the crystallite size. Thus, a bigger size of the crystalline domain implies an increase of the particle dimension and a reduction of its surface area. In porous materials, it is also reported that the increase of the temperature can induce a collapse of the porous structure and a densification of the material. The behavior described in the reduction of the surface area is in agreement with the reported in other works [28,33].
XPS was used to investigate the oxidation states and chemical environment of W species on the surface of the as-synthesized catalysts; the results are shown in Figure 1. In an XPS survey spectra (see Figure S1) of selected catalysts, binding energies of the O1s and Si2s levels for all the catalysts agreed with the values obtained for silica (O1s = 532.9 eV and Si1s = 155.3 eV) [77]. The catalysts present significant signals at 531.3 eV and 285 eV, possibly due to the presence of hydroxyl groups and carbonate species or C–C bonds in carbon impurities (adventitious carbon) [78,79].
The W4f HR-XPS core-level spectrum (Figure 1) reveals that the W4f doublet (5/2 and 7/2) presents at least four different signals corresponding to different W species at the surface. From approximately 33.8 eV to 36.2 eV and from 36.2 eV to 40 eV represent the spin orbit splitting W4f 7/2 and W4f 5/2, respectively. Then, each of these regions was decomposed into two different signals, corresponding to different oxidation states or chemical environments of the tungsten superficial atoms. In general, the signals are displaced to higher energy with the increasing calcination temperature, indicating that the tungsten is surrounded by a high electronegative environment or is more oxidized when the temperature increases. W4f electrons from WO3 have binding energies at around 35.9 eV and 38.0 eV for 7/2 and 5/2 orbits, respectively [80,81,82]. As is shown in Figure 1a, another signal displaced to less binding energy is found in all W4f HR-XPS. These signals vary between 34.8 eV to 35.5 eV and 36.6 eV to 37.3 eV for 7/2 and 5/2 orbits, respectively. Generally, the spectrum of supported WO3 is broader than massif WO3 due to the interaction between the tungsten oxides and the support [83]. On another side, the broadening of the spectrum toward the high and low binding energy flack suggests that the supported tungsten is not fully oxidized, and possibly the surface contains small amounts of W5+ on the supported catalysts. The analysis of the high-resolution XPS for W4f suggests that the surface of the tungsten nanoparticles contains a large quantity of defects which diminish with the increasing of the temperature. These defects cause a nonstoichiometry oxide by the apparition of oxygen vacancies V O 2   at the surface and in the bulk as well. A nonstoichiometry oxide can be written as shown in Equation (2), where the missing charges can be balanced by a reduction of the W cations. In the Supplementary Information is shown the whole process (in Kröger–Vink notation) of the vacancy formation and reduction of the cations inside of a tungsten oxide (Equation (S8)–(S10)). Thus, the changes in BE shown in W4f high-resolution XPS (Figure 1) are associated with the oxygen vacancy formation and the cation reduction from 6+ to 5+, in accord with the BE values.
W O 3 δ   W 1 2 δ 6 + W 2 δ 5 + O 3 δ 2 δ V O 2   ,
Different reports show that the tungsten oxides can produce oxygen vacancies, and that its quantity can be modulated though the heating temperature and the atmosphere as well [84,85,86,87,88]. These defects play an important role in the physicochemical properties, such as catalytic [89] and photocatalytic activity [85], quantum confinement [86], and ionic conductivity [85], among others.
The atomic ratio of surface elements is shown in Figure 1c; all the catalysts showed small changes of Si/O ratio whereas W/Si and W/O ratios were different between catalysts. Considering the ratio of W/O and (W)/Si, which indicated the surface dispersion and concentration of tungsten on a silica support, a direct relation is seen to calcination temperature, relatively greater than the impregnation dosage used on each one [77,89]

3.2. Asphaltene Adsorption Isotherms

Effect of Synthesis Conditions

Figure 2a–f shows the obtained adsorption isotherms of asphaltenes by SiWNa and SiW samples at 25 °C with the SLE model. For every sample, the sorption isotherms revealed an increase in the adsorbed asphaltene content as the asphaltene concentration increased. It should be noted that SHS adsorbed more asphaltenes than SiO2 nanoparticles for all asphaltene concentrations. The difference was more noticeable at higher concentrations (CE > 200 mg·L−1), which agrees with previous results [28,36,46,47,48,49,50]. These results agree with literature reports for the adsorption of different asphaltene types onto different solid surfaces [36,37,42,43,73]. The adsorption isotherms follow a type Ib behavior according to the IUPAC classification [87], indicating that there is a strong affinity between the n-C7 asphaltenes and the synthesized materials [31,90].
On another side, considering the acidic character of tungsten oxide, this behavior suggests the basicity of the asphaltenes in addition the presence of heteroatoms such as nitrogen and oxygen in the asphaltene molecules, which favor the interaction with the nanoparticle surface [91,92]. It could be observed in Figure 2a–f that there is a trend on the adsorptive capacity highly influenced by synthesis conditions of WSN following the general orders SiWNa1 > SiWNa3 > SiWNa5 > SiO2 and SiWNa350 > SiWNa 450 > SiWNa 650 > SiO2. The SLE model parameters are enlisted in Table 2. Additionally, the adsorption affinity, represented by the reciprocal of H value, increases for lower tungsten impregnation dosage, except for the virgin support. The same trend can be observed by the degree of asphaltene self-association on active sites, which is represented by K. This parameter decreases with the increase in impregnation dosage, and the lowest value is obtained for the virgin support. Conversely, the inverse trend can be observed for the qm parameter. From values of qm and the H parameter, it can be observed that the functionalized nanoparticles show more affinity and adsorption capacity than the support. In addition, lower values of the K parameter suggest that the inclusion of tungsten oxide allows the inhibition of the adsorbate self-association.
Van der Waals forces, polar, and acid–base (Lewis and/or Brønsted) interactions seem to be the dominant forces that contribute to the adsorption of the asphaltenes onto tungsten oxide and/or salt surfaces [93]. Previous studies concluded that the adsorption of asphaltenes over silica is due to the chemisorption of asphaltenes on the silica silanol groups [94]. On the other hand, acid sites are responsible for adsorption of basic molecules that contain heteroatoms [91,92]. Due to the low acidity of silica, the presence of tungsten oxide generates the formation of Brønsted acid sites and the increase of Lewis acid sites [95], increasing interactions with the basic sites of the asphaltenes, or heteroatoms N, O, and S, which makes the interaction stronger and consequently increases the number of asphaltenes adsorbed.
Figure 2g shows the obtained adsorption isotherms of asphaltenes by SiW at 25 °C with the SLE model. The asphaltene uptake is enhanced after functionalizing the SiO2 with the trend SiW1 > SiW3 > SiW5 > Si and SiW 350 > SiW 450 > SiW 650 > SiO2. Figure S2 compares the obtained adsorption isotherms of asphaltenes by SiW1 350 and SiWNa1 350 at 25 °C with the SLE model. In this case, the amount adsorbed of asphaltenes is higher than SiWNa nanoparticles; in addition, SiW nanoparticles showed a similar adsorption affinity than SiWNa. These differences could be due to electronic effects and different interactions between the active phase and the support (as seen in XPS analysis) that influence the selectivity for different functional groups or heteroatoms in n-C7 asphaltene molecules [43].
Figure 3 presents the trends of the H and K parameters of the SLE model. Figure 3a shows the trends for functionalized nanoparticles from sodium tungstate at different dosage calcination temperatures. Here, in the parameter H, as the calcination temperature increases, the value increases, which suggests that the higher the calcination temperature, the lower the affinity. On the other hand, parameter K shows that as the calcination temperature increases for the same dosage, the self-association of the asphaltenes increases.
During XPS analysis, a direct relationship between calcination temperature and W/O and W/Si ratios was demonstrated, which indicated that the surface dispersion and concentration of tungsten on a silica support depends on the calcination temperature used. In addition, the tungsten nanoparticles contain a large quantity of defects that are diminishing with the increasing calcination temperature. All these properties affect the adsorptive behavior of nanoparticles and their ability to inhibit the self-association degree of asphaltenes on their surfaces.
Figure 4b shows that, likewise for nanoparticles functionalized from ammonium tungstate, the affinity value increases as the calcination temperature increases, which suggests a lower affinity. Regarding the parameter K, there is a slight increase in the value, which suggests a greater self-association of the asphaltenes as the calcination temperature increases.
Asphaltene adsorption over the surface of the nanoparticles depends on factors such as asphaltene concentration on the medium, affinity interactions, and availability of active sites [38,61]. However, at medium and high uptake [38,43], due to their self-associated tendency asphaltenes, molecules will form aggregates around the high-energy sites until saturation [61,96]. On another side, the asphaltenes adsorption is subject to the adsorbent surface selectivity to the aromatic part of the asphaltenes or heteroatoms present in their structure, generating different adsorption orientations (perpendicular, parallel, or flatways) [61]. Following this order of ideas, with the functionalization of the nanoparticles, selectivity is induced, and the conditions used during material synthesis could influence the amount of active sites available for asphaltene adsorption; at the same time, the obtained results suggest that these conditions affect the characteristics of supported metal oxide, such as dispersion and active phase size, influence the adsorptive capacity of support, and enhance its affinity for asphaltenes [97]. It is important to remark that the adsorption capacity seems to be enhanced after functionalization at a lower dosage.

3.3. Catalytic Steam Decomposition of n-C7 Asphaltenes

3.3.1. Mass Loss Analysis

The tests were carried out in a wet air atmosphere for a specific asphaltene loading of 0.20 mg·m−2. Figure 4 shows the rate of mass loss for the virgin n-C7 asphaltenes in the absence and presence of SiWNa. Figure 4a–c compares the presence of nanoparticles using the same impregnation dosage, and Figure 4e–f compare the presence of nanoparticles using the same calcination temperature during nanoparticles synthesis. As seen for virgin n-C7 asphaltenes, the reaction has an onset approximately at 325 °C, with a maximum peak at 465 °C and an offset near to 574 °C. In this case, the presence of silica nanoparticles did not prove to be effective in reducing the decomposition temperature of asphaltenes. Figure 5 shows that, generally, the curve for asphaltenes rate of mass loss in the presence of WSN shifts to the left, confirming that decomposition occurs at a lower temperature. For the best scenarios evaluated with the WSN (SiWNa1 350, SiWNa450, and SiWNa650), the maximum peak of mass loss rate was reduced from 465 °C to 325 °C. In previous reports, the support (SiO2) was evaluated without any functionalization, showing that it reduces the temperature of transformation of asphaltenes a bit under different atmospheres, including inert or thermal cracking [73], steam gasification, and air (or combustion [28]). In this case, the presence and absence of water into the stream was also evaluated, showing that the decomposition profile remains unchanged. However, a change in the produced gases was observed and more gasification products were found (H2, CO, and CH4). These products will be analyzed in the following section.
The conversion of n-C7 asphaltenes to gaseous products can be associated with their partial oxidation, pyrolysis, and gasification [16]. Steam presence in the decomposition of asphaltenes as carbonaceous source implies different reactions, such as water–gas shift reaction and methanation [43]. The catalytic decomposition of asphaltenes in the presence of WSN can be divided into two reaction regions, LTR (low-temperature reaction, before 400 °C) and HTR (high-temperature reaction, between 400 °C and 600 °C). Initially, reactions on the breakdown of alkyl chains take place and the dissociation of weak bonds (S–C and N–C) [93,98], followed by the opening of polycyclic aromatic hydrocarbons and radical or addition reactions that can form heavier compounds (these compounds could be larger from the starting species turning into coke) [51,99]. With the increase in the temperature, there are still some reactions attributed to the gasification and oxidation of resultant compounds produced by free radicals that were not stabilized and due to addition reactions [42]. In the catalytic decomposition process, irrespective of the atmosphere (wet air), the first step includes the oxidation and low-temperature gasification of the asphaltenes. Hence, the volatile content is released, and the remaining condensed aromatic sheets subsequently react. Later, the oxygen present reacts with condensed aromatic sheets oxidizing first outer layers and, at relatively higher temperatures, the internal layers [93].
The inclusion of WSN reduces the temperature of asphaltene consumption. In all cases, two main peaks are observed, one on LTR and another on HTR. It is inferred that the first peak corresponds to the breaking of weakest bonds of asphaltenes and the second, to the decomposition of more stable structures, such as aromatic condensed rings. The presence of H2O can enhance the process of asphaltenes decomposition, due to the reaction of gaseous H2 or O2 previously adsorbed such as water over the nanoparticle matrix and active sites with adsorbed molecules that would be cracked, and free radicals formed which are highly reactive and can aid to decompose other molecules, i.e., raw asphaltenes. According to the results obtained, the surface properties and the catalytic activity of the nanoparticles affect how asphaltenes and water are adsorbed over the active sites, leading to a highly efficient process in energetic terms [43]. In the studied cases, the first peak temperature is considerably reduced by nanoparticles at approximately 100 °C and for the WSN that shows the best catalytic activity, SiWNa1 350. Changing the precursor does not show significant catalytic activity (Figure 4g), a factor that could be attributed to poor dispersion, acidity, affinity, or the big size of the active phase. It is considered that acid sites influence the asphaltenes conversion [100]. Isomerization reactions occur on the Brønsted acid sites of the WO3/SiO2 catalysts, and studies have concluded that by poisoning the surface acid sites with small amounts of alkali elements, such as Na, K, Rb, and Cs, branched metathesis products were decreased, that is, isomerization was reduced, obtaining a greater degree of asphaltenes consumption [101,102,103].
Regarding the SLE parameters, it can also be noticed a direct relation between the catalytic effect of the synthesized materials with the affinity for asphaltene adsorption and the degree of self-association. With lower values of H and K parameters of the SLE model, the catalytic behavior of the surface is higher. The asphaltene auto-association degree would impact the catalytic activity of the nanoparticle blocking this one, due to that the formation of large aggregates could block some active sites [61]. The highest tungsten oxide dispersion and the lowest active phase showed a higher reduction of decomposition temperature.

3.3.2. Estimation of the Effective Activation Energies

Another way to validate the catalytic activity of the nanoparticles consists of the estimation of the effective activation energies through the isoconversional OFW methodology. For this case, the nanoparticle that demonstrated the highest efficiency in hydrogen production and reduction of asphaltene decomposition temperature was selected to analyze the effective activation energy behavior (that is, SiWNa1 350). For comparative purposes, the nanoparticle was selected with the same metal dosage, calcined at the same temperature, but with different precursors, that is, SiW1 350.
Figure 5 shows the estimated effective activation energies as a function of the degree of conversion. For virgin n-C7 asphaltenes, the value of Ea decreases as the conversion increases, which agrees well with previous reports under different reaction atmospheres [28,72,95]. Conversely, for the decomposition n-C7 asphaltenes in the presence of SiWNa1 350 and SiW1 350 nanoparticles, the opposite trend is observed. It is worth remarking that, in all cases for conversion < 80%, the effective activation energy is shown to be lower compared with the virgin n-C7 asphaltene; this again confirms the catalytic activity of the nanoparticles. For the catalytic asphaltene decomposition, the activation energy increases with the degree of conversion, suggesting that the catalytic decomposition consists of a complex mechanism that involves multiple steps [42,104]. SiWNa1 350 displayed the lowest activation energy throughout the process, suggesting that this sample is more catalytically active toward SiW1 350. The reduction of the effective activation energy is of primary importance to enhance the wet in-situ combustion, as a more efficient heat consumption can be ensured and could lead to higher oil recovery with upgraded characteristics.
As discussed above, the temperature of calcination affects the size of particles and the surface chemical environment. For lower temperatures, a small particle, high dispersion, and more quantity of defects are found. Though the XPS analysis, lower binding energies were found when the calcination temperature was set at 350 ℃, showing that the W cations were in the lowest oxidation state (on average). Therefore, for these catalysts (350 ℃), it is expected to find the largest quantity of surface defects, such as oxygen vacancies and reduced tungsten cations. These couples of punctual defects V O 2   W W x constitute a redox or basic–acid site that promotes the adsorption and activation of the different reactants, such as water, oxygen, and hydrocarbons, favoring the reaction kinetics (see Figure S3).

3.3.3. Analysis of the Gaseous Products Evolved during Catalytic Decomposition Process

During the asphaltenes decomposition though wet air gasification, different gases such as CO, CH4, H2, light hydrocarbons (LHC), and CO2 can be released. In this study, the releases gases during virgin asphaltene and catalytic steam gasification with SiWNa1 350 were analyzed using mass spectrometry. The SiWNa1 350 sample was selected based on catalytic results reported in previous sections. The experimental nonisothermal heating at 10 °C·min−1 was considered, and the results of the gaseous products are shown in Figure 6. As expected, the gaseous mixture was composed of CO2, CH4, CO, C2H2, C2H4, C6H6, and H2. The gases C2H2, C2H4, and C6H6 will be referred to as light hydrocarbons (LHC). Some traces of other gases were evidenced in different temperature intervals, but they were neglected (selective distribution < 1.0% vol).
Panel a shows the evolution of the different gases with temperature change for virgin asphaltene decomposition. The gas mixture was composed mainly for CO and CO2. The CO evolved increased with temperature because of the reverse Boudouard reaction. The CO2 is produced by different mechanisms, including the transformation of ketones and other oxygenated groups in the asphaltene chemical structure. The produced LHC was evidenced during all the temperature range evaluated. However, at lower temperatures (<200 °C), the higher release of them was noted because of the breaking and decomposition of aliphatic chains and low-molecular-weight hydrocarbons. Finally, H2 release was poor in the noncatalyzed sample because these heavy fractions follow other reaction mechanisms that prioritize the production of other gases, such as CO2 and CO [29].
The results of the catalyzed sample are shown in panel b. The results show an increase in LHC, CH4, and H2 during the wide range of temperature. At lower temperatures (<250 °C), the LHC production is slightly higher than the uncatalyzed system because of the loss of aliphatic chains and low-molecular-weight hydrocarbons. In the catalyzed scenario, the LHC mixture is composed mainly of C2H4 (>60%vol), a highly calorific gas; while in the noncatalyzed scenario, this gas was produced to a lesser extent (<30%vol).
On the other hand, CO2 production decreased between 200–300 °C, whereas CO and CH4 increased in the same range. CO2 reacts with carbon atoms to produce CO while CH4 is released by the breaking of methyl, methylene, and other short aliphatic chains. Then, between 300 and 400 °C, a drop in CO and a slight increase in CO2 and H2 is observed. This behavior can be attributed to the occurrence of water–gas-shift reaction [105]. In addition, methane-reforming reactions appear to occur in the same range of temperature as a reduction of the CH4 content in the gas mixture. It is worth mentioning that hydrogen released was higher than the noncatalyzed system in all the temperature range. It can be produced by the reaction of C in asphaltenes and H2O(g) molecules. The higher hydrogen content was 20%vol between 300 and 400 °C.
In the thermic process, reactions of total or partial combustion are common (formation of CO); in the presence of steam, the methanation (formation of CH4) and steam reforming are common [106]. The increase of CH4 production in the presence of nanoparticles with respect to virgin asphaltenes suggests an increment in methanation reactions promoted by the action of nanoparticles; at the same time, in addition to the combustion reaction by the presence of O2, part of the formed CH4 and the reaction in water present in steam form CO. These reactions are aligned with the upgrading process in the reservoir as lighter compounds would lead to viscosity reduction and API gravity increase.

4. Conclusions

Asphaltene adsorption and subsequent catalytic decomposition on functionalized silica nanoparticles with tungsten oxides were studied using model solutions on batch adsorption experiments and TGA/MS system for evaluating thermocatalytic properties in a steam/air atmosphere. Dispersion and medium active phase size of supported oxides were shown to be essential factors in the catalytic and adsorptive behavior of nanoparticles. Sodium tungstate and tungsten oxide supported nanocatalysts were compared, and the results showed a similar adsorptive capacity but a marked difference in catalytic behavior.
XPS analysis demonstrated a direct relationship between calcination temperature and W/O and W/Si ratios, which indicated the surface dispersion and concentration of tungsten on a silica support, depending on the calcination temperature used. In addition, the tungsten nanoparticles contain a large quantity of defects that are diminishing with the increasing calcination temperature. All these properties affect the adsorptive behavior of nanoparticles and their ability to inhibit the self-association degree of asphaltenes on their surfaces.
The catalytic results show that synthesized nanoparticles reduce asphaltene decomposition temperature. In addition, synthesis parameters, such as temperature and impregnation dosage, play an important role in the catalytic activity of the materials due to the different WOX–support interactions. The mixture released during the catalyzed asphaltene decomposition in the wet air atmosphere reveals an increase in light hydrocarbons, methane, and hydrogen content. Hydrogen production was prioritized between 300 and 400 °C, where the reduction of CO, CH4 and the increase in CO2 content similarly occurs, associated with water–gas-shift and methane-reforming reactions, respectively.
The results obtained in this research project suggest that the SiO2 nanoparticles functionalized with tungsten oxides (WSN) can be used as catalysts for the in situ upgrading of heavy crude oil during wet in situ combustion

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr10020349/s1, Solid–liquid equilibrium model; Section S2. OFW model; Figure S1. XPS survey spectrum and high-resolution spectra for SiWNa5 650 sample.; Section S3. Vacancy formation and compensation charge into a tungsten oxide using Kröger–Vink notation.; Figure S2. Adsorption isotherms of asphaltenes onto supported nanoparticles using the same impregnation dosage and calcination temperature but different metal oxide precursors. Nanoparticle dose, 10 g·L−1; agitation speed, 200 rpm; T, 25 °C. The symbols are experimental data, and the solid lines are from the SLE model.; Figure S3. Chemical species around an oxygen vacancy. Model for adsorption of reactants.

Author Contributions

Conceptualization, K.M.C., D.A.-M. and C.A.F.; methodology, K.M.C., D.A.-M. and C.A.F.; formal analysis, K.M.C., D.A.-M., O.E.M., L.E.C., F.B.C., C.A.F. and J.G.; investigation, K.M.C., D.A.-M., C.A.F. and J.G.; data curation, K.M.C., D.A.-M., O.E.M., C.A.F. and J.G.; writing—original draft preparation, K.M.C., O.E.M. and D.A.-M.; writing—review and editing, all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fondo Nacional de Financiamiento para la Ciencia, la Tecnología y la Innovación “FRANCISCO JOSÉ DE CALDAS”, Agencia Nacional de Hidrocarburos (ANH), COLCIENCIAS and Universidad Nacional de Colombia for their support provided in Agreement 272 of 2017.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors give thanks to Universidad Nacional de Colombia and Universidad del Valle for the support.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

APIAmerican Petroleum Institute
C0Initial concentration of asphaltenes
CEEquilibrium concentrations in solution after adsorption
DLSDynamic light scattering
EHOExtra-heavy oil
HOHeavy oil
HHenry’s law constant
HTRHigh-temperature reaction
KSelf-association degree
LTRLow-temperature reaction
LHCLight hydrocarbons
MSAPMean particle size of active phase
OFWOzawa–Flynn–Wall
PAHPolyaromatic hydrocarbons
qmMaximum adsorbed amount
TEORThermal enhanced oil recovery
TEOTransition element oxides
TGAThermogravimetric analysis
SLESolid–liquid equilibrium
SBET (m2)BET surface area
SSSSodium silicate solution
WSNTungsten functionalized silica nanoparticles
XPSX-ray photoelectron spectroscopy
R Universal gas constant
T Absolute temperature
E α Activation energy

References

  1. Zaid, A.M.; Al-Dousari, M.M. Forces Driving Oil and Gas Demand Cycles. In Proceedings of the International Petroleum Technology Conference, Dubai, United Arab Emirates, 4–8 December 2007. [Google Scholar]
  2. Alboudwarej, H.; Felix, J.J.; Taylor, S.; Badry, R.; Bremner, C.; Brough, B.; Skeates, C.; Baker, A.; Palmer, D.; Pattison, K. La importancia del petróleo pesado. Oilfield Rev. 2006, 18, 38–58. [Google Scholar]
  3. Medina, O.E.; Gallego, J.; Rodríguez, E.; Franco, C.A.; Cortés, F.B. Effect of pressure on the oxidation kinetics of Asphaltenes. Energy Fuels 2019, 33, 10734–10744. [Google Scholar] [CrossRef]
  4. Medina, O.E.; Olmos, C.; Lopera, S.H.; Cortés, F.B.; Franco, C.A. Nanotechnology applied to thermal enhanced oil recovery processes: A review. Energies 2019, 12, 4671. [Google Scholar] [CrossRef] [Green Version]
  5. Mitra-Kirtley, S.; Mullins, O.C. Sulfur chemical moieties in carbonaceous materials. In Asphaltenes, Heavy Oils, and Petroleomics; Springer: New York, NY, USA, 2007; pp. 157–188. [Google Scholar]
  6. Tissot, B.P.; Welte, D.H. Petroleum Formation and Occurrence; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  7. Groenzin, H.; Mullins, O.C. Asphaltene molecular size and structure. J. Phys. Chem. A 1999, 103, 11237–11245. [Google Scholar] [CrossRef]
  8. Mullins, O.C. The asphaltenes. Annu. Rev. Anal. Chem. 2011, 4, 393–418. [Google Scholar] [CrossRef]
  9. Strausz, O.P.; Lown, E.M. The Chemistry of Alberta Oil Sands, Bitumens and Heavy Oils; Alberta Energy Research Institute: Calgary, AB, Canada, 2003. [Google Scholar]
  10. Yudin, I.; Nikolaenko, G.; Gorodetskii, E.; Kosov, V.; Melikyan, V.; Markhashov, E.; Frot, D.; Briolant, Y. Mechanisms of asphaltene aggregation in toluene–heptane mixtures. J. Pet. Sci. Eng. 1998, 20, 297–301. [Google Scholar] [CrossRef]
  11. Murgich, J.; Abanero, J.A.; Strausz, O.P. Molecular recognition in aggregates formed by asphaltene and resin molecules from the Athabasca oil sand. Energy Fuels 1999, 13, 278–286. [Google Scholar] [CrossRef]
  12. Demirbas, A. Deposition and flocculation of asphaltenes from crude oils. Pet. Sci. Technol. 2016, 34, 6–11. [Google Scholar] [CrossRef]
  13. Medina, O.E.; Gallego, J.; Cespedes, S.; Nassar, N.N.; Montoya, T.; Corteś, F.B.; Franco, C.A. Effect of pressure on thermo-oxidative reactions of saturates, aromatics, and resins (S-Ar-R) from extra-heavy crude oil. Fuel 2021, 311, 122596. [Google Scholar] [CrossRef]
  14. Adams, J.J. Asphaltene adsorption, a literature review. Energy Fuels 2014, 28, 2831–2856. [Google Scholar] [CrossRef]
  15. Zahedi, G.; Fazlali, A.; Hosseini, S.; Pazuki, G.; Sheikhattar, L. Prediction of asphaltene precipitation in crude oil. J. Pet. Sci. Eng. 2009, 68, 218–222. [Google Scholar] [CrossRef]
  16. Nassar, N.N.; Hassan, A.; Pereira-Almao, P. Application of nanotechnology for heavy oil upgrading: Catalytic steam gasification/cracking of asphaltenes. Energy Fuels 2011, 25, 1566–1570. [Google Scholar] [CrossRef]
  17. Medina, O.E.; Gallego, J.; Redondo, J.D.; Corteś, F.B.; Franco, C.A. Effect of pressure on the thermo-oxidative behavior of saturates, aromatics, and resins (S-Ar-R) mixtures. Fuel 2021, 122787. [Google Scholar] [CrossRef]
  18. Allenson, S.J.; Walsh, M.A. A novel way to treat asphaltene deposition problems found in oil production. In Proceedings of the International Symposium on Oilfield Chemistry, Houston, TX, USA, 18 February 1997. [Google Scholar]
  19. Chang, J.; Ivory, J. Steam-Air Injection Process. In Proceedings of the SPE Canada Heavy Oil Technical Conference, Calgary, AB, Canada, 8 June 2016. [Google Scholar]
  20. Kokal, S.L.; Sayegh, S.G. Asphaltenes: The cholesterol of petroleum. In Proceedings of the Middle East Oil Show, Bahrain, Saudi Arabia, 15 March 1997. [Google Scholar]
  21. Saniere, A.; Hénaut, I.; Argillier, J. Pipeline transportation of heavy oils, a strategic, economic and technological challenge. Oil Gas Sci. Technol. 2004, 59, 455–466. [Google Scholar] [CrossRef]
  22. Medina, O.E.; Gallego, J.; Nassar, N.N.; Acevedo, S.A.; Cortés, F.B.; Franco, C.A. Thermo-Oxidative Decomposition Behaviors of Different Sources of n-C7 Asphaltenes at High-Pressure Conditions. Energy Fuels 2020, 34, 8740. [Google Scholar] [CrossRef]
  23. Bagci, A.S. Wet forward combustion for heavy oil recovery. Energy Sources Part A 2006, 28, 221–232. [Google Scholar] [CrossRef]
  24. Lapene, A.; Castanier, L.M.; Debenest, G.; Quintard, M.Y.; Kamp, A.M.; Corre, B. Effects of Water on Kinetics of Wet In-Situ Combustion. In Proceedings of the SPE Western Regional Meeting, San Jose, CA, USA, 1 January 2009; p. 12. [Google Scholar]
  25. Ward, G.D.; Ward, C.E. Insitu Wet Combustion Process for Recovery of Heavy Oils. Google Patents US4691773A, 8 September 1987. [Google Scholar]
  26. Sarathi, P.S. In-Situ Combustion Handbook—Principles and Practices; National Petroleum Technology Office: Sawfa, Saudi Arabia, 1999.
  27. Greaves, M.; Young, T.; El-Usta, S.; Rathbone, R.; Ren, S.; Xia, T. Air injection into light and medium heavy oil reservoirs: Combustion tube studies on West of Shetlands Clair oil and light Australian oil. Chem. Eng. Res. Des. 2000, 78, 721–730. [Google Scholar] [CrossRef]
  28. Franco, C.A.; Montoya, T.; Nassar, N.N.; Pereira-Almao, P.; Cortés, F.B. Adsorption and Subsequent Oxidation of Colombian Asphaltenes onto Nickel and/or Palladium Oxide Supported on Fumed Silica Nanoparticles. Energy Fuels 2013, 27, 7336–7347. [Google Scholar] [CrossRef]
  29. Medina, O.E.; Gallego, J.; Acevedo, S.; Riazi, M.; Ocampo-Pérez, R.; Cortés, F.B.; Franco, C.A. Catalytic Conversion of n-C7 Asphaltenes and Resins II into Hydrogen Using CeO2-Based Nanocatalysts. Nanomaterials 2021, 11, 1301. [Google Scholar] [CrossRef]
  30. Cardona, L.; Medina, O.E.; Céspedes, S.; Lopera, S.H.; Cortés, F.B.; Franco, C.A. Effect of Steam Quality on Extra-Heavy Crude Oil Upgrading and Oil Recovery Assisted with PdO and NiO-Functionalized Al2O3 Nanoparticles. Processes 2021, 9, 1009. [Google Scholar] [CrossRef]
  31. Medina, O.E.; Gallego, J.; Pérez-Cadenas, A.F.; Carrasco-Marín, F.; Cortés, F.B.; Franco, C.A. Insights into the Morphology Effect of Ceria on the Catalytic Performance of NiO–PdO/CeO2 Nanoparticles for Thermo-oxidation of n-C7 Asphaltenes under Isothermal Heating at Different Pressures. Energy Fuels 2021, 35, 18170–18184. [Google Scholar] [CrossRef]
  32. Mateus, L.; Moreno-Castilla, C.; López-Ramón, M.V.; Cortés, F.B.; Álvarez, M.Á.; Medina, O.E.; Franco, C.A.; Yebra-Rodríguez, Á. Physicochemical characteristics of calcined MnFe2O4 solid nanospheres and their catalytic activity to oxidize para-nitrophenol with peroxymonosulfate and n-C7 asphaltenes with air. J. Environ. Manag. 2021, 281, 111871. [Google Scholar] [CrossRef]
  33. Arias-Madrid, D.; Medina, O.E.; Gallego, J.; Acevedo, S.; Correa-Espinal, A.A.; Cortés, F.B.; Franco, C.A. NiO, Fe2O3, and MoO3 Supported over SiO2 Nanocatalysts for Asphaltene Adsorption and Catalytic Decomposition: Optimization through a Simplex–Centroid Mixture Design of Experiments. Catalysts 2020, 10, 569. [Google Scholar] [CrossRef]
  34. Hashemi, R.; Nassar, N.N.; Almao, P.P. Nanoparticle technology for heavy oil in-situ upgrading and recovery enhancement: Opportunities and challenges. Appl. Energy 2014, 133, 374–387. [Google Scholar] [CrossRef]
  35. Hosseinpour, N.; Mortazavi, Y.; Bahramian, A.; Khodatars, L.; Khodadadi, A.A. Enhanced pyrolysis and oxidation of asphaltenes adsorbed onto transition metal oxides nanoparticles towards advanced in-situ combustion EOR processes by nanotechnology. Appl. Catal. A Gen. 2014, 477, 159–171. [Google Scholar] [CrossRef]
  36. Nassar, N.N.; Hassan, A.; Pereira-Almao, P. Metal Oxide Nanoparticles for Asphaltene Adsorption and Oxidation. Energy Fuels 2011, 25, 1017–1023. [Google Scholar] [CrossRef]
  37. Cortés, F.B.; Chejne, F.; Carrasco-Marín, F.; Pérez-Cadenas, A.F.; Moreno-Castilla, C. Water sorption on silica-and zeolite-supported hygroscopic salts for cooling system applications. Energy Convers. Manag. 2012, 53, 219–223. [Google Scholar] [CrossRef]
  38. Montoya, T.; Coral, D.; Franco, C.A.; Nassar, N.N.; Cortés, F.B. A novel solid–liquid equilibrium model for describing the adsorption of associating asphaltene molecules onto solid surfaces based on the “chemical theory”. Energy Fuels 2014, 28, 4963–4975. [Google Scholar] [CrossRef]
  39. Krishnamoorti, R. Extracting the benefits of nanotechnology for the oil industry. J. Pet. Technol. 2006, 58, 24–26. [Google Scholar] [CrossRef]
  40. Medina, O.E.; Galeano-Caro, D.; Ocampo-Pérez, R.; Perez-Cadenas, A.F.; Carrasco-Marín, F.; Franco, C.A.; Corteś, F.B. Development of a monolithic carbon xerogel-metal composite for crude oil removal from oil in-saltwater emulsions: Evaluation of reuse cycles. Microporous Mesoporous Mater. 2021, 327, 111424. [Google Scholar] [CrossRef]
  41. Medina, O.E.; Galeano-Caro, D.; Castelo-Quibén, J.; Ocampo-Pérez, R.; Perez-Cadenas, A.F.; Carrasco-Marín, F.; Franco, C.A.; Corteś, F.B. Monolithic carbon xerogels-metal composites for crude oil removal from oil in-saltwater emulsions and subsequent regeneration through oxidation process: Composites synthesis, adsorption studies, and oil decomposition experiments. Microporous Mesoporous Mater. 2021, 319, 111039. [Google Scholar] [CrossRef]
  42. Franco, C.A.; Montoya, T.; Nassar, N.N.; Cortés, F.B. NiO nd PdO supported on fumed silica nanoparticles for adsorption and catalytic steam gasification of colombian c7asphaltenes. In Handbook on Oil Production Research; Nova Science Hauppauge: New York, NY, USA, 2014. [Google Scholar]
  43. Nassar, N.N.; Franco, C.A.; Montoya, T.; Cortés, F.B.; Hassan, A. Effect of oxide support on Ni–Pd bimetallic nanocatalysts for steam gasification of n-C7 asphaltenes. Fuel 2015, 156, 110–120. [Google Scholar] [CrossRef]
  44. Medina, O.E.; Gallego, J.; Restrepo, L.G.; Cortés, F.B.; Franco, C.A. Influence of the Ce4+/Ce3+ Redox-Couple on the Cyclic Regeneration for Adsorptive and Catalytic Performance of NiO-PdO/CeO2±δ Nanoparticles for n-C7 Asphaltene Steam Gasification. Nanomaterials 2019, 9, 734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Medina, O.E.; Caro-Vélez, C.; Gallego, J.; Cortés, F.B.; Lopera, S.H.; Franco, C.A. Upgrading of Extra-Heavy Crude Oils by Dispersed Injection of NiO–PdO/CeO2±δ Nanocatalyst-Based Nanofluids in the Steam. Nanomaterials 2019, 9, 1755. [Google Scholar] [CrossRef] [Green Version]
  46. Franco-Ariza, C.A.; Guzmán-Calle, J.D.; Cortés, F.B. Adsorption and catalytic oxidation of asphaltenes in fumed silica nanoparticles: Effect of the surface acidity. Dyna 2016, 83, 171. [Google Scholar] [CrossRef]
  47. Nassar, N.N.; Hassan, A.; Carbognani, L.; Lopez-Linares, F.; Pereira-Almao, P. Iron oxide nanoparticles for rapid adsorption and enhanced catalytic oxidation of thermally cracked asphaltenes. Fuel 2012, 95, 257–262. [Google Scholar] [CrossRef]
  48. Nassar, N.N.; Hassan, A.; Pereira-Almao, P. Comparative oxidation of adsorbed asphaltenes onto transition metal oxide nanoparticles. Colloids Surf. A: Physicochem. Eng. Asp. 2011, 384, 145–149. [Google Scholar] [CrossRef]
  49. Nassar, N.N.; Hassan, A.; Pereira-Almao, P. Effect of surface acidity and basicity of aluminas on asphaltene adsorption and oxidation. J. Colloid Interface Sci. 2011, 360, 233–238. [Google Scholar] [CrossRef]
  50. Nassar, N.N.; Hassan, A.; Pereira-Almao, P. Effect of the particle size on asphaltene adsorption and catalytic oxidation onto alumina particles. Energy Fuels 2011, 25, 3961–3965. [Google Scholar] [CrossRef]
  51. Nassar, N.N.; Hassan, A.; Pereira-Almao, P. Thermogravimetric studies on catalytic effect of metal oxide nanoparticles on asphaltene pyrolysis under inert conditions. J. Therm. Anal. Calorim. 2012, 110, 1327–1332. [Google Scholar] [CrossRef]
  52. Barton, D.G.; Soled, S.L.; Iglesia, E. Solid acid catalysts based on supported tungsten oxides. Top. Catal. 1998, 6, 87–99. [Google Scholar] [CrossRef]
  53. Speight, J.G. Fouling in Refineries; Gulf Professional Publishing: Houston, TX, USA, 2015. [Google Scholar]
  54. J Spivey, J.; Dooley, K.M.; Han, Y.-F.; Meunier, F.; Tshentu, Z.R.; Dai, W.-L.; Jin, R.; Efstathiou, A.; Meunier, F.; Coppens, M.-O.; et al. Catalysis; Royal Society of Chemistry: Picadilly, UK, 2016; Volume 28. [Google Scholar]
  55. Kung, H.H. Transition Metal Oxides: Surface Chemistry and Catalysis; Elsevier: Amsterdam, The Netherlands, 1989; Volume 45. [Google Scholar]
  56. Wasmi, B.A.; Al-Amiery, A.A.; Kadhum, A.A.H.; Mohamad, A.B. Novel approach: Tungsten oxide nanoparticle as a catalyst for malonic acid ester synthesis via ozonolysis. J. Nanomater. 2014, 2014, 2. [Google Scholar] [CrossRef]
  57. Barre, L.; Espinat, D.; Rosenberg, E.; Scarsella, M. Colloidal structure of heavy crudes and asphaltene soltutions. Rev. De L’institut Français Du Pétrole 1997, 52, 161–175. [Google Scholar] [CrossRef] [Green Version]
  58. ASTM. D5236-13 Standard Test Method for Distillation of Heavy Hydrocarbon Mixtures (Vacuum Potstill Method). Annual Book of ASTM Standards; ASTM International: West Conshohocken, PA, USA, 2013. [Google Scholar]
  59. ASTM. D2892 Standard Test Method for Distillation of Crude Petroleum (15-Theoretical Plate Column). Annual Book of ASTM Standards; ASTM International: West Conshohocken, PA, USA, 2016. [Google Scholar]
  60. Franco, C.; Patiño, E.; Benjumea, P.; Ruiz, M.A.; Cortés, F.B. Kinetic and thermodynamic equilibrium of asphaltenes sorption onto nanoparticles of nickel oxide supported on nanoparticulated alumina. Fuel 2013, 105, 408–414. [Google Scholar] [CrossRef]
  61. Franco, C.A.; Nassar, N.N.; Montoya, T.; Ruíz, M.A.; Cortés, F.B. Influence of asphaltene aggregation on the adsorption and catalytic behavior of nanoparticles. Energy Fuels 2015, 29, 1610–1621. [Google Scholar] [CrossRef]
  62. Vert, M.; Hellwich, K.-H.; Hess, M.; Hodge, P.; Kubisa, P.; Rinaudo, M.; Schué, F. Terminology for biorelated polymers and applications (IUPAC Recommendations 2012). Pure Appl. Chem. 2012, 84, 377–410. [Google Scholar] [CrossRef]
  63. Medina, O.E.; Gallego, J.; Arias-Madrid, D.; Cortés, F.B.; Franco, C.A. Optimization of the load of transition metal oxides (Fe2O3, Co3O4, NiO and/or PdO) onto CeO2 nanoparticles in catalytic steam decomposition of n-C7 asphaltenes at low temperatures. Nanomaterials 2019, 9, 401. [Google Scholar] [CrossRef] [Green Version]
  64. Diao, W.; Tengco, J.M.M.; Regalbuto, J.R.; Monnier, J.R. Preparation and characterization of Pt–Ru bimetallic catalysts synthesized by electroless deposition methods. ACS Catal. 2015, 5, 5123–5134. [Google Scholar] [CrossRef]
  65. Geyer, R.; Hunold, J.; Keck, M.; Kraak, P.; Pachulski, A.; Schödel, R. Methods for determining the metal crystallite size of Ni supported catalysts. Chem. Ing. Tech. 2012, 84, 160–164. [Google Scholar] [CrossRef]
  66. Shastri, A.G.; Schwank, J. Metal dispersion of bimetallic catalysts via stepwise chemisorption and surface titration: I. Ru_AuSiO2. J. Catal. 1985, 95, 271–283. [Google Scholar] [CrossRef]
  67. Socrates, G. Infrared and Raman Characteristic Group Frequencies: Tables and Charts; John Wiley & Sons: Hoboken, NJ, USA, 2004. [Google Scholar]
  68. Yao, H.; Dai, Q.; You, Z. Fourier Transform Infrared Spectroscopy characterization of aging-related properties of original and nano-modified asphalt binders. Constr. Build. Mater. 2015, 101, 1078–1087. [Google Scholar] [CrossRef]
  69. Leofanti, G.; Padovan, M.; Tozzola, G.; Venturelli, B. Surface area and pore texture of catalysts. Catal. Today 1998, 41, 207–219. [Google Scholar] [CrossRef]
  70. Guzmán, J.D.; Betancur, S.; Carrasco-Marín, F.; Franco, C.A.; Nassar, N.N.; Cortés, F.B. Importance of the Adsorption Method Used for Obtaining the Nanoparticle Dosage for Asphaltene-Related Treatments. Energy Fuels 2016, 30, 2052–2059. [Google Scholar] [CrossRef]
  71. Giraldo, J.; Nassar, N.N.; Benjumea, P.; Pereira-Almao, P.; Cortés, F.B. Modeling and prediction of asphaltene adsorption isotherms using Polanyi’s modified theory. Energy Fuels 2013, 27, 2908–2914. [Google Scholar] [CrossRef]
  72. Cortés, F.B.; Chejne, F.; Carrasco-Marín, F.; Moreno-Castilla, C.; Pérez-Cadenas, A.F. Water adsorption on zeolite 13X: Comparison of the two methods based on mass spectrometry and thermogravimetry. Adsorption 2010, 16, 141–146. [Google Scholar] [CrossRef]
  73. 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, 561–571. [Google Scholar] [CrossRef] [Green Version]
  74. Franco Ariza, C.A. Synthesis and Application of Supported Metallic and Multi-Metallic Oxides Nanoparticles for In-Situ Upgrading and Inhibition of Formation Damage; Universidad Nacional de Colombia: Bogota, Columbia, 2015. [Google Scholar]
  75. López, D.; Giraldo, L.J.; Salazar, J.P.; Zapata, D.M.; Ortega, D.C.; Franco, C.A.; Cortés, F.B. Metal Oxide Nanoparticles Supported on Macro-Mesoporous Aluminosilicates for Catalytic Steam Gasification of Heavy Oil Fractions for On-Site Upgrading. Catalysts 2017, 7, 319. [Google Scholar] [CrossRef] [Green Version]
  76. Chen, Y.-F.; Lee, C.-Y.; Yeng, M.-Y.; Chiu, H.-T. The effect of calcination temperature on the crystallinity of TiO2 nanopowders. J. Cryst. Growth 2003, 247, 363–370. [Google Scholar] [CrossRef]
  77. Colque, S.; Payen, E.; Grange, P. Novel preparation of highly dispersed tungsten oxide on silica. J. Mater. Chem. 1994, 4, 1343–1348. [Google Scholar] [CrossRef]
  78. Elkins, T.W.; Hagelin-Weaver, H.E. Characterization of Mn–Na2WO4/SiO2 and Mn–Na2WO4/MgO catalysts for the oxidative coupling of methane. Appl. Catal. A: Gen. 2015, 497, 96–106. [Google Scholar] [CrossRef]
  79. Salmaoui, S.; Sediri, F.; Gharbi, N.; Perruchot, C.; Jouini, M. Hexagonal hydrated tungsten oxide nanomaterials: Hydrothermal synthesis and electrochemical properties. Electrochim. Acta 2013, 108, 634–643. [Google Scholar] [CrossRef]
  80. Terohid, S.A.A.; Heidari, S.; Jafari, A.; Asgary, S. Effect of growth time on structural, morphological and electrical properties of tungsten oxide nanowire. Appl. Phys. A 2018, 124, 567. [Google Scholar] [CrossRef] [Green Version]
  81. Biesinger, M.C. X-ray Photoelectron Spectroscopy (XPS) Reference Pages. Available online: http://www.xpsfitting.com/ (accessed on 1 May 2020).
  82. Agilent Scientific Instruments, Thermo Scientific XPS Simplified. Available online: https://xpssimplified.com/ (accessed on 1 May 2020).
  83. Susanti, D.; Diputra, A.A.G.P.; Tananta, L.; Purwaningsih, H.; Kusuma, G.E.; Wang, C.; Shih, S.; Huang, Y. WO3 nanomaterials synthesized via a sol-gel method and calcination for use as a CO gas sensor. Front. Chem. Sci. Eng. 2014, 8, 179–187. [Google Scholar] [CrossRef]
  84. Bursill, L.A. Structure of small defects in nonstoichiometric WO3−x. J. Solid State Chem. 1983, 48, 256–271. [Google Scholar] [CrossRef]
  85. Fukushi, D.; Sasaki, A.; Hirabayashi, H.; Kitano, M. Effect of oxygen vacancy in tungsten oxide on the photocatalytic activity for decomposition of organic materials in the gas phase. Microelectron. Reliab. 2017, 79, 1–4. [Google Scholar] [CrossRef]
  86. Tanaka, D.; Oaki, Y.; Imai, H. Enhanced photocatalytic activity of quantum-confined tungsten trioxide nanoparticles in mesoporous silica. Chem. Commun. 2010, 46, 5286–5288. [Google Scholar] [CrossRef]
  87. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051. [Google Scholar] [CrossRef] [Green Version]
  88. Zhang, H.; Feng, M.; Liu, F.; Liu, L.; Chen, H.; Gao, H.; Li, J. Structures and defects of WO3−x nanorods grown by in-situ heating tungsten filament. Chem. Phys. Lett. 2004, 389, 337–341. [Google Scholar] [CrossRef]
  89. Gayapan, K.; Sripinun, S.; Panpranot, J.; Praserthdam, P.; Assabumrungrat, S. Effect of pretreatment atmosphere of WOX/SiO2 catalysts on metathesis of ethylene and 2-butene to propylene. RSC Adv. 2018, 8, 11693–11704. [Google Scholar] [CrossRef] [Green Version]
  90. Medina, O.E.; Gallego, J.; Olmos, C.M.; Chen, X.; Cortés, F.B.; Franco, C.A. Effect of multifunctional nanocatalysts on n-C7 asphaltene adsorption and subsequent oxidation under high-pressure conditions. Energy Fuels 2020, 34, 6261–6278. [Google Scholar] [CrossRef]
  91. Ramirez-Corredores, M.M. The Science and Technology of Unconventional Oils: Finding Refining Opportunities; Academic Press: Cambridge, MA, USA, 2017. [Google Scholar]
  92. Zimmer, A.K.; Becker, C.; Chambliss, C.K. Exploiting metal oxide nanoparticle selectivity in asphaltenes for identification of pyridyl-containing molecules. Energy Fuels 2013, 27, 4574–4580. [Google Scholar] [CrossRef]
  93. Hosseinpour, N.; Khodadadi, A.A.; Bahramian, A.; Mortazavi, Y. Asphaltene adsorption onto acidic/basic metal oxide nanoparticles toward in situ upgrading of reservoir oils by nanotechnology. Langmuir 2013, 29, 14135–14146. [Google Scholar] [CrossRef]
  94. Fritschy, G.; Papirer, E. Interactions between a bitumen, its components and model fillers. Fuel 1978, 57, 701–704. [Google Scholar] [CrossRef]
  95. Andreini, A.; Mol, J.C. Activity of supported tungsten oxide catalysts for the metathesis of propene. J. Chem. Soc. Faraday Trans. 1: Phys. Chem. Condens. Phases 1985, 81, 1705–1714. [Google Scholar] [CrossRef]
  96. Dudášová, D.; Simon, S.; Hemmingsen, P.V.; Sjöblom, J. Study of asphaltenes adsorption onto different minerals and clays: Part 1. Experimental adsorption with UV depletion detection. Colloids Surf. A: Physicochem. Eng. Asp. 2008, 317, 1–9. [Google Scholar] [CrossRef]
  97. Contreras, J.L.; Fuentes, G.A. Sintering of Supported Metal Catalysts. In Sintering-Methods and Products; InTech: London, UK, 2012. [Google Scholar]
  98. Siddiqui, M.N. Catalytic pyrolysis of Arab Heavy residue and effects on the chemistry of asphaltene. J. Anal. Appl. Pyrolysis 2010, 89, 278–285. [Google Scholar] [CrossRef]
  99. Hamedi Shokrlu, Y.; Babadagli, T. In-situ upgrading of heavy oil/bitumen during steam injection by use of metal nanoparticles: A study on in-situ catalysis and catalyst transportation. SPE Reserv. Eval. Eng. 2013, 16, 333–344. [Google Scholar] [CrossRef]
  100. Breysse, M.; Furimsky, E.; Kasztelan, S.; Lacroix, M.; Perot, G. Hydrogen activation by transition metal sulfides. Catal. Rev. 2002, 44, 651–735. [Google Scholar] [CrossRef]
  101. Kawai, T.; Goto, H.; Yamazaki, Y.; Ishikawa, T. Metathesis of n-alkenes over a CsNO3-Re2O7-Al2O3 catalyst. J. Mol. Catal. 1988, 46, 157–172. [Google Scholar] [CrossRef]
  102. Spamer, A.; Dube, T.; Moodley, D.; Van Schalkwyk, C.; Botha, J. The reduction of isomerisation activity on a WO3/SiO2 metathesis catalyst. Appl. Catal. A: Gen. 2003, 255, 153–167. [Google Scholar] [CrossRef]
  103. Van Roosmalen, A.; Mol, J. Active centers for the metathesis and isomerization of alkenes on tungsten-oxide/silica catalysts. J. Catal. 1982, 78, 17–23. [Google Scholar] [CrossRef]
  104. Nassar, N.N.; Hassan, A.; Vitale, G. Comparing kinetics and mechanism of adsorption and thermo-oxidative decomposition of Athabasca asphaltenes onto TiO2, ZrO2, and CeO2 nanoparticles. Appl. Catal. A: Gen. 2014, 484, 161–171. [Google Scholar] [CrossRef]
  105. Gradisher, L.; Dutcher, B.; Fan, M. Catalytic hydrogen production from fossil fuels via the water gas shift reaction. Applied Energy 2015, 139, 335–349. [Google Scholar] [CrossRef]
  106. Kök, M.V.; Pamir, M.R. Pyrolysis and combustion studies of fossil fuels by thermal analysis methods. J. Anal. Appl. Pyrolysis 1995, 35, 145–156. [Google Scholar] [CrossRef]
Figure 1. (a) XPS high-resolution spectra of the W4f doublet 4f7/2 and 4f5/2. (b) Binding energy of W surface species on selected nanoparticles. (c) Atomic ratio of each species in the selected WSN.
Figure 1. (a) XPS high-resolution spectra of the W4f doublet 4f7/2 and 4f5/2. (b) Binding energy of W surface species on selected nanoparticles. (c) Atomic ratio of each species in the selected WSN.
Processes 10 00349 g001
Figure 2. Adsorption isotherms of n-C7 asphaltenes onto supported nanoparticles: Effect of calcination temperature at the same impregnation dosage; (ac) effect of calcination temperature at the same impregnation dosage; effect of impregnation dosage at the same calcination temperature (df); effect of precursor (g). Nanoparticle dose, 10 g·L−1; agitation speed, 200 rpm; T, 25 °C. The symbols are experimental data, and the solid lines are from the SLE model (see Equation (2)).
Figure 2. Adsorption isotherms of n-C7 asphaltenes onto supported nanoparticles: Effect of calcination temperature at the same impregnation dosage; (ac) effect of calcination temperature at the same impregnation dosage; effect of impregnation dosage at the same calcination temperature (df); effect of precursor (g). Nanoparticle dose, 10 g·L−1; agitation speed, 200 rpm; T, 25 °C. The symbols are experimental data, and the solid lines are from the SLE model (see Equation (2)).
Processes 10 00349 g002
Figure 3. Trends for K and H parameters of the adjustment model of the adsorption isotherms (a) with nanoparticles functionalized from sodium tungstate at different calcination rates and temperatures, and (b) with functionalized nanoparticles from ammonium tungstate and sodium tungstate at 1% (mass fraction) and different calcination temperatures.
Figure 3. Trends for K and H parameters of the adjustment model of the adsorption isotherms (a) with nanoparticles functionalized from sodium tungstate at different calcination rates and temperatures, and (b) with functionalized nanoparticles from ammonium tungstate and sodium tungstate at 1% (mass fraction) and different calcination temperatures.
Processes 10 00349 g003
Figure 4. The plot of the rate of mass loss as a function of temperature in the steam/air atmosphere and conversion for asphaltenes in the presence of nanoparticles: (ac) effect of calcination temperature at the same impregnation dosage; effect of impregnation dosage at the same calcination temperature (df); effect of precursor (g). Amount adsorbed = 0.2 mg·m−2; air flow rate = 100 mL·min−1; H2O (g) flow rate = 6.30 mL·min−1.
Figure 4. The plot of the rate of mass loss as a function of temperature in the steam/air atmosphere and conversion for asphaltenes in the presence of nanoparticles: (ac) effect of calcination temperature at the same impregnation dosage; effect of impregnation dosage at the same calcination temperature (df); effect of precursor (g). Amount adsorbed = 0.2 mg·m−2; air flow rate = 100 mL·min−1; H2O (g) flow rate = 6.30 mL·min−1.
Processes 10 00349 g004
Figure 5. Effective activation energies calculated by OFW method as a function of the % conversion for catalytic decomposition of virgin asphaltenes in the presence and absence of selected WSN with both precursor salts at 1% (in mass fraction).
Figure 5. Effective activation energies calculated by OFW method as a function of the % conversion for catalytic decomposition of virgin asphaltenes in the presence and absence of selected WSN with both precursor salts at 1% (in mass fraction).
Processes 10 00349 g005
Figure 6. Selectivity distribution of light gases produced from the nonisothermal conversion of asphaltenes (a) without and (b) with SiWNa 350 nanoparticles between 100 and 600 °C. Air flow rate = 100 mL·min−1; H2O(g) flow rate = 6.30 mL·min−1; asphaltene load = 0.0002 mmol·m−2; heating rate = 10 °C·min−1.
Figure 6. Selectivity distribution of light gases produced from the nonisothermal conversion of asphaltenes (a) without and (b) with SiWNa 350 nanoparticles between 100 and 600 °C. Air flow rate = 100 mL·min−1; H2O(g) flow rate = 6.30 mL·min−1; asphaltene load = 0.0002 mmol·m−2; heating rate = 10 °C·min−1.
Processes 10 00349 g006
Table 1. Measured values of tungsten nanoparticle diameter (MSAP), dispersion, and surface area of hybrid nanoparticles, WSN.
Table 1. Measured values of tungsten nanoparticle diameter (MSAP), dispersion, and surface area of hybrid nanoparticles, WSN.
Material* SBET ± 0.1 m2/g * MSAP ± 0.1 nmTungsten Dispersion (%)
SiO2126.1--
SiWNa1 350122.31.626.1
SiWNa1 45092.91.919.5
SiWNa1 650101.12.616.5
SiWNa3 350108.57.613.6
SiWNa3 45093.58.610.8
SiWNa3 65094.310.25.6
SiWNa5 35094.79.55.6
SiWNa5 45094.213.73.1
SiWNa5 65093.715.42.9
SiW1 35081.11.815.2
SiW1 45079.63.412.6
SiW1 65066.24.79.1
* Uncertainty is obtained by triplicate measurements.
Table 2. SLE model parameters of asphaltenes adsorption by studied nanoparticles at 25 °C with the effect of calcination temperature at the same impregnation dosage, the effect of impregnation dosage at the same calcination temperature, and effect of the precursor.
Table 2. SLE model parameters of asphaltenes adsorption by studied nanoparticles at 25 °C with the effect of calcination temperature at the same impregnation dosage, the effect of impregnation dosage at the same calcination temperature, and effect of the precursor.
MaterialH (mg·g−1)K × 10−4 (g/g)qm (g·g−1)R2χ2
Si2.276.611.290.990.49
SiWNa1 3500.761.733.210.990.19
SiWNa1 4501.901.894.410.990.62
SiWNa1 6502.602.024.800.990.48
SiWNa3 3501.902.034.370.990.19
SiWNa3 4502.074.354.950.990.62
SiWNa3 6502.185.374.660.990.48
SiWNa5 3501.751.194.200.991.00
SiWNa5 4502.251.464.780.991.25
SiWNa5 6502.672.405.510.990.18
SiW1 3501.241.646.020.990.46
SiW1 4501.581.845.950.990.50
SiW1 6501.752.303.140.990.14
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cerón, K.M.; Arias-Madrid, D.; Gallego, J.; Medina, O.E.; Chinchilla, L.E.; Cortés, F.B.; Franco, C.A. Catalytic Decomposition of n-C7 Asphaltenes Using Tungsten Oxides–Functionalized SiO2 Nanoparticles in Steam/Air Atmospheres. Processes 2022, 10, 349. https://doi.org/10.3390/pr10020349

AMA Style

Cerón KM, Arias-Madrid D, Gallego J, Medina OE, Chinchilla LE, Cortés FB, Franco CA. Catalytic Decomposition of n-C7 Asphaltenes Using Tungsten Oxides–Functionalized SiO2 Nanoparticles in Steam/Air Atmospheres. Processes. 2022; 10(2):349. https://doi.org/10.3390/pr10020349

Chicago/Turabian Style

Cerón, Karen M., Daniela Arias-Madrid, Jaime Gallego, Oscar E. Medina, Lidia E. Chinchilla, Farid B. Cortés, and Camilo A. Franco. 2022. "Catalytic Decomposition of n-C7 Asphaltenes Using Tungsten Oxides–Functionalized SiO2 Nanoparticles in Steam/Air Atmospheres" Processes 10, no. 2: 349. https://doi.org/10.3390/pr10020349

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop