Skip to content
Publicly Available Published by De Gruyter June 18, 2020

Ni–Mo sulfide nanosized catalysts from water-soluble precursors for hydrogenation of aromatics under water gas shift conditions

  • Anna Vutolkina EMAIL logo , Aleksandr Glotov , Ilnur Baygildin , Argam Akopyan , Marta Talanova , Maria Terenina , Anton Maximov and Eduard Karakhanov

Abstract

The unsupported catalysts were obtained during hydrogenation by in situ high-temperature decomposition (above 300 °C) of water-soluble metal precursors (ammonium molybdate and nickel nitrate) in water-in-oil (W/O) emulsions stabilized by surfactant (SPAN-80) using elemental sulfur as sulfiding agent. These self-assembly Ni–Mo sulfide nanosized catalysts were tested in hydrogenation of aromatics under CO pressure in water-containing media for hydrogen generation through a water gas shift reaction (WGSR). The composition of the catalysts was determined by XRF and active sulfide phase was revealed by XRD, TEM and XPS techniques. The calculations based on TEM and XPS data showed that the catalysts are highly dispersed. The surfactant was found to affect both dispersion and metal distribution for Ni and Mo species, providing shorter slab length in terms of sulfide particle formation and stacking within high content of NiMoS phase. Catalytic evaluation in hydrogenation of aromatics was performed in a high-pressure batch reactor at T = 380–420 °С, p(CO) = 5 MPa with water content of 20 wt.% and CO/H2O molar ratio of 1.8 for 4–8 h. As shown experimentally with unsupported Ni–Mo sulfide catalysts, the activity of aromatic rings depends on the substituent therein and decreases as follows: anthracene>>1-methylnaphthalene≈2-methylnaphthalene>1,8-dimethylnaphthale-ne>>1,3-di-methylnaphthalene>2,6-dimethylnaphthalene≈2,3-dimethylnaphthalene>2-ethyl-naphthalene. The anthracene conversion reaches up to 97–100% for 4 h over the whole temperature range, while for 1MN and 2MN it doesn’t exceed 92 and 86% respectively even at 420 °С for 8 h. Among dimethyl-substituted aromatics the higher conversion of 45% was achieved for 1,8-dimethylnaphthalene with 100% selectivity to tetralines at 400 °С for 6 h. Similar to 1- and 2-methylnaphtalenes, the hydrogenation of asymmetric dimethyl-substituted substrate carries out through the unsubstituted aromatic ring indicating that steric factors influence on the sorption mechanism over active metal sites. The catalysts were found to be reused for at least six cycles when the hydrogenation is sulfur-assisted preventing metal oxide formation. It was established, that at the first 2–3 h known as the induction period, the oxide catalyst precursors formed slowly by metal salt decomposition, which reveals that it is the rate-determining step. The sulfidation is rather fast based on high catalytic activity data on 2MN conversion retaining at 93–95% upon recycling.

Introduction

The rapid developments in technology, improving living standards, and the growth in world population have led to the global energy demand becoming a critical issue. The exhaustion of petroleum resources and fossil fuels, and deterioration of the environment has encouraged the development of unconventional feedstock refining technologies. The feeds generally are classified as heavy or extra heavy oil, oil sands and bitumen [1], [2], [3]. The heavy oils produced by steam injection into the reservoir followed by recovering as emulsions. Unlike fossil fuels heavy oils have a high content of aromatics, resins and asphaltenes with heavy metals, nitrogen, sulfur and oxygen containing compounds resulting in high density and viscosity thereof, so the emulsions should be treated before transportation [4], [5], [6]. Due to high stability of water-oil emulsions caused by mineral compounds [7], [8], the water separation is an energy consuming and non-viable technology [9], [10]. To solve these problems the single-stage technology combining emulsion breaking and upgrading should be developed. The promising way is water activation for in situ hydrogen generation through the water gas shift reaction (WGSR), that would also prevent hydrogen consumption which has become more important in heavy oil refinery operations. For this process the unsupported transition metal sulfides are the most suitable: being the main components of hydrotreatment catalysts and are also highly active in WGSR over a wide temperature range [11], [12], [13], [14], [15]. While for alumina or silica–alumina supported sulfide catalysts reactant having a high molecular weight deposits on the active sites and blocks the pores [16], [17], [18], dispersed catalysts provide contact of substrate molecules with active component excluding diffusion limitations and mass transfer [19]. Blocking carbocations, thereby preventing condensation and polymerization, leads to a decrease in deactivation and coking [20]. Among the variety of different techniques, the decomposition of water- or oil-soluble metal salts followed by optional sulfidation is simple and reproducible, leading to production of binary and ternary sulfides with controlled stoichiometry [21], [22].

Despite the similar structure, Mo-based catalysts were found to be more active compared to W-containing ones, predominantly because of difficulties to convert tungsten oxide to corresponding sulfide [21], [23]. Ni–Mo sulfides were reported to be the most active in the WGSR among transition metal catalysts [12], [24]. Their high activity also caused by synergistic effect of the bimetallic system. Recently we reported the high activity of the Ni–Mo sulfide catalysts obtained from oil-soluble precursors in hydrogenation of aromatics [25] and hydrodesulphurization of dibenzothiophene compounds under CO pressure in the presence of water providing in situ hydrogen generation through a WGSR [26], [27]. Nevertheless, oil-soluble precursors is often more expensive than water-soluble ones [23]. The latter requires surfactants to stabilize water-in-oil emulsions and to control particles size. On decomposition of water-soluble metal salts, water evaporates very fast leading to low dispersion caused by particle agglomeration [23]. This work summaries the properties of the self-assembled Ni–Mo sulfide nanosized catalysts obtained during hydrogenation of aromatics by in situ high-temperature decomposition (above 300 °C) of water-soluble metal precursors (ammonium molybdate and nickel nitrate) in W/O emulsions stabilized by surfactant (SPAN-80) using elemental sulfur as sulfiding agent under CO pressure. Depending on the preparation and treatment conditions, the catalysts’ structure will be correlated with catalytic activity in hydrogenation of bi- and tricycle aromatics.

Experimental

Chemicals

The following reagents were used without additional purification: ammonium heptamolybdate tetrahydrate (NH4)6Mo7O24·4H2O (99%, “Alfa Aesar”),nickel nitrate hexahydrate Ni(NO3)2·6H2O (≥97%, “Sigma-Aldrich”), 1-methylnaphthalene С11Н10 (99%, “Sigma-Aldrich”), 2-methylnaphthalene С11Н10 (99%, “Sigma-Aldrich”), anthracene С14Н10 (98%, “Sigma-Aldrich”), 1,8-dimethylnaphtalene С12Н12 (95%, “Sigma-Aldrich”), 1,3-dimethylnaphtalene С12Н12 (96%, “Sigma-Aldrich”), 2,6-dimethylnaphtalene С12Н12 (99%, “Sigma-Aldrich”), 2,3-dimethylnaphtalene С12Н12 (97%, “Sigma-Aldrich”), 2-ethylnaphthalene С12Н12 (99%, “Sigma-Aldrich”), sorbitan monooleate (SPAN-80) C24H44O6 (98%, “Merck”), elemental sulfur S (98%, “Chemical reactant”), toluene C7H8 (99%, “Khimmed”).

Characterization techniques

X-ray powder diffraction (XRD) patterns were recorded between 2θ = 1.5°–70.0° by 0.05, using nickel filter and Cu Kα = 1.54 Å (λ = 0.154 nm) radiation on the Bruker D2 PHASER diffractometer at room temperature. XRD data obtained were interpreted using Rigaku PDXL software with a powder database.

The dispersions of the Ni–Mo sulfide catalysts were investigated with Transmission electron microscopy (TEM) using a JEOL JEM-2100 microscope with an accelerating voltage 200 kV. The samples were prepared by dispersing in n-hexane and then put onto formvar grids sample holder. The samples covered with liquid hexane for protection from oxidation by air were introduced into the vacuum chamber. More detailed information about the active sulfide phase (distributions of slab length and layers stacking) was revealed by the statistical evaluation of around 300 particles from different zones of a number of various TEM images using Image-Pro Plus 7.0 software. The characteristics of the active phase were calculated assuming that the MoS2 slabs were perfect hexagons and using the equations in Table 1.

Table 1:

The equations for calculation of the sulfide phase morphological characteristics from TEM micrographs [28], [29].

Parameter Equation
The average slab length of sulfide particles ( L ¯; )
L ¯ = l i n ,
l i – the length of slab i;

n – total number of slabs
(1)
The average stacking number ( N ¯; )
N ¯; = n i N i n ,
n i – the number of stacks in N i layers
(2)
The total number of Mo atoms along one side of MoS2 slab (n i )
n i = 10 L ¯; 3 , 2 + 1 2
(3)
The total number of Mo atoms at the edge surface of MoS2 (Mo e )
M o e = ( 6 n i 12 ) N ¯;
(4)
The total number of Mo corner sites for MoS2 (Mo c )
M o c = 6 N ¯;
(5)
The total number of Mo atoms (Mo T )
M o T = ( 3 n i 2 3 n i + 1 ) N ¯;
(6)
Active phase dispersion (D)
D = ( M o e + M o c ) M o T
(7)
The ratio of Mo atoms at the edge surface to Mo corner sites (f e /f c )
f e / f c = 10 L ¯; 3 , 2 3 2
(8)

The elemental composition of catalysts was determined by X-ray fluorescence (XRF) spectrometry on energy dispersive Thermo Scientific ARL Quant’X. The measurements were carried out in vacuum and results were processed using the standard-less UniQuant method.

The XPS spectra of the sulfide catalysts were recorded on a PHI5500VersaProbeII spectrometer equipped with a hemispherical electron analyzer and Al Kα (hν = 1486,6 eV) X-ray source. Binding energy (BE) scale was preliminary calibrated using the peaks position for the Au 4f7/2 (84.0 eV) and Cu 2p3/2 (932.6 eV) core levels. The pass energy was set at 117.4 eV for the survey scan and 23.5 eV for the narrow scan of Mo3d, S2p, O1s, C1s, while for the narrow scan of Ni2p the pass energy was set at 29.35 eV. The BE were referenced to the C1 s peak (284.9 eV) to account for the charging effects. The accuracy of the BE values is ± 0.1 eV. After fitting the instrument, peak areas were approximated by Gaussian/Lorentzian curves and removal of the background (Shirley function). The analysis of the individual spectral regions allowed identifying chemical state of the elements according to binding energies. Surface atomic ratios were calculated from the peak area ratios normalized by the corresponding atomic sensitivity factors. Metal distribution for Ni and Mo species at the surface of sulfided catalysts were calculated using the equations in Table 2.

Table 2:

The equations for calculation of the metal distribution for Ni and Mo species at the surface of sulfide catalysts from XPS data.

Parameter Equation
The effective Ni content in NiMoS phase
w N i M o S = [ N i M o S ] w N i 100   ,
[NiMoS] – relative concentration of NiMoS phase, %;

ωNi – effective Ni concentration , wt.%
(9)
The effective Mo content in MoS2 phase
w M o S 2 = [ M o S 2 ] w M o 100   ,
[MoS2] – relative concentration of MoS2 in sulfide state, %; ωMo – effective Mo concentration of, wt.%
(10)
Ni promotion of MoS2 slab
p ( Ni / Mo ) s = c N i M o S c M o S 2 ,
cNiMoS – absolute concentration of Ni in NiMoS species, at.%;

cMoS2 – absolute concentration of Mo in MoS2 phase, at.%
(11)

The Ni/Mo atomic ratio on the edges of NiMoS active phase was calculated from XPS and TEM as follows:

(12) p ( N i / M o ) e = p ( N i / M o ) s M o e + M o c M o T = p ( N i / M o ) s D

Catalytic experiments

The activity of the catalysts was evaluated in a batch reactor (50 cm3) equipped with a magnetic stirrer. In the typical experiment the catalyst precursors (ammonium molybdate and nickel nitrate) were dissolved in distilled water (20 wt.%) and placed in autoclave. Then, solution of aromatic hydrocarbon (10 wt.%) in toluene containing surfactant (SPAN-80) was added. Following this, elemental sulfur (2.5 wt.%) was suspended in the reaction mixture. The reactant ratios are shown in Table 3.

Table 3:

Reactant ratios.

ω Mo, wt.% S, mmol Mo:Ni molar ratio ω substrate, wt.% Gas mmol (25 °С) Mo:substrate molar ratio ω H2O, wt.% H2O mmol
0.1 3.9 3:1 10 100 1:65 20 55.6

The Mo content (wt.%) in the reaction mixture was calculated with the equation:

w ( N H 4 ) 6 M o 7 O 24 4 H 2 O =   m s u b s t r a t e M ( ( N H 4 ) 6 Mo 7 O 24 4 H 2 O ) 65 7 M s u b s t r a t e m s o l u t i o n 100 %

w N i ( N O 3 ) 2 6 H 2 O =   w ( ( NH 4 ) 6 Mo 7 O 24 4 H 2 O ) 7 ∗M ( Ni ( NO 3 ) 2 6 H 2 O ) 3 ∗M ( ( NH 4 ) 6 Mo 7 O 24 4 H 2 O )

where msubstrate – mass of aromatic hydrocarbon, g; M – molar mass of catalyst precursors or aromatic hydrocarbon (substrate), g/mol; m solution – solution mass, g; msolution  = mH2O + mtoluene  + mS + msubstrate  + m((NH4)6Mo7O24∙4H2O) + m(Ni(NO3)2∙6H2O) + msurfactant, msolution  = 5 g.

The autoclave was pressurized with carbon monoxide to 5.0 MPa, placed in a thermostatically controlled oven and heated up to reaction temperature with ramp of 15 °C/min. The stirring speed was kept at 400–500 rpm. We suppose these conditions and stirring rate free the system from diffusion limitations. When the temperature controlled at the bottom of the reactor by thermocouple reached the set point, the reaction was carried out for 4–8 h. After a predetermined time, the autoclave was cooled and disassembled. Under the reaction conditions the catalysts’ precursors, forming through the high-temperature decomposition of water-soluble metal salts, catalyze the WGSR to form hydrogen, followed by its interaction with elemental sulfur. Thus, the resulting hydrogen sulfide acts as a sulfiding agent to produce an active sulfide form of the catalyst. The catalyst was separated by filtration followed by step-by-step washing with toluene and hexane. Then, it was dried in inert atmosphere and evaluated by XRD, XRF, TEM and XPS techniques. The composition of liquid product was analyzed using GC (FID) provided with capillary column Petrocol™ (Supelco 0.25mmх50 m). The column temperature was varied depending on the reactant compound, and temperature programming was used to improve peak resolution. From the GC analysis the conversion of model substances (C, %) and was calculated as follows:

C = S P S s + S P 100 %

where S s – peak area of unreacted substrate, %; S p – sum of peak areas of products, %.

The selectivity (S, %) was determined from the equation:

S = S i S P 100 %

where S i – peak area of i-product, %.

Results and discussion

Surfactant influence

Nowadays more attention has been paid to bulk or unsupported transition metal catalysts. Unlike alumina or silica–alumina supported sulfide, dispersed catalysts provide contact of substrate molecules with active component excluding diffusion limitations and mass transfer (17) resulting in high activity. Unsupported catalysts due to its uniform dispersion in feeds acts as hydrogen transfer agent. Thus, these catalysts favor the rapid hydrogen consumption deactivating organic radicals and preventing thereby their condensation and polymerization resulting in suppression of coke formation. When hydrogen reacts with sulfur ions in MoS2 that are located in the corner and edge, the unsaturated sites and sulfur ion vacancies being active sites for hydrogenation are formed. Meanwhile, the Mo–H and S–H moieties are unstable and react rapidly with the free hydrocarbon radicals decreasing the rate of condensation or polymerization reactions [20], [30], [31], [32], [33], [34]. Among the variety of different techniques, the decomposition of water- or oil-soluble metal salts followed by sulfidation is simple and reproducible, allowing for preparation of binary and ternary sulfides with controlled stoichiometry [21], [22]. Generally, oil-soluble precursors are more active than water-soluble ones, but too expensive [23]. The lower activity of catalysts obtained from water-soluble precursors is caused by fast water evaporation leading to particle agglomeration and low active phase dispersion [23].

In this work water-soluble ammonium heptamolybdate tetrahydrate (NH4)6Mo7O24·4H2O and nickel nitrate hexahydrate Ni(NO3)2·6H2O were used as catalyst precursors and elemental sulfur as a sulfiding agent. The active phase formed in situ at high-temperature during hydrogenation of aromatics in W/O emulsions under CO pressure according to Scheme 1.

Scheme 1: 
Formation of Ni–Mo catalysts in W/O emulsion during hydrogenation of aromatics from water-soluble precursors under CO pressure.
Scheme 1:

Formation of Ni–Mo catalysts in W/O emulsion during hydrogenation of aromatics from water-soluble precursors under CO pressure.

Elevating the temperature of the reaction mixture (at the range of 45–420 °С) so as to decompose of ammonium molybdate and nickel nitrate, leads to formation of intermediates as well as metal oxides (Scheme 2) [35], [36], [37], [38], [39], [40], [41]. Interaction between Mo and promoter supposed to occurs spontaneously at ambient conditions [20]. The cationic species of Ni interact with molybdate anions [20] followed by sulfidation and leading to precipitation of dark solids. Thus, Ni2+ as well as Mo4O13 2– exists in solution simultaneously up to 200–240 °C that may provide formation of NiMo4O13 species. Moreover, (NH4)6Mo7O24 *4H2O precursor also may decomposed to MoO3*(H2O)2 at 300 °C [42] and reacts with NiNO3 to form hybrids containing NiMoO4*2H2O and MoO3 [43], [44]. When the temperature rises, the NiMo4O13 as well as NiMoO4 may acts with hydrogen sulfide to form NiMoxSy particles [45]. These mixed-metal oxides are easier to reduce due to it is easier to adsorb and dissociate H2S on Ni sites than on Mo sites of an oxide with water removal. Meanwhile, decomposition to form metal oxides also takes place.

Scheme 2: 
High-temperature decomposition-sulfidation of Ni and Mo salts [35], [36], [37], [38], [39], [40], [41].
Scheme 2:

High-temperature decomposition-sulfidation of Ni and Mo salts [35], [36], [37], [38], [39], [40], [41].

To control particle size and high active phase dispersion, the stabilizing of W/O emulsions by surfactant was applied. Sorbitan monooleate (SPAN-80) with a hydrophilic-lipophilic balance (HLB) value of 4–6 favoring the formation of invert emulsions was chosen [46]. The hydrogenation of 1- and 2-methylnaphtalenes was chosen as a model reaction to evaluate the surfactant influence on catalytic properties. The catalyst obtained through the surfactant-assisted synthesis in W/O emulsion was named NiMoS + SPAN, while counterpart formed without stabilizing agent was named NiMoS. According to the experimental data the former is more active in 2-methylnaphtalene (2MN) hydrogenation, about twice as high as the latter (by conversion for 4–6 h, see Fig. 1a). At higher reaction time of 8 h the conversion of 2MN over NiMoS and NiMoS + SPAN becomes comparable exceeding 75%. When the temperature rises, the conversion of 1-methylnaphtalene (1MN) also increases both for NiMoS and NiMoS + SPAN catalysts while the latter being more active (Fig. 1b).

Fig. 1: 
Conversion of (a) 2MN (Т = 380 °С) and (b) 1MN (t = 6 h) over Ni–Mo sulfide catalysts depending on surfactant presence. Reaction conditions: ω(SPAN-80) = 4 wt.%, p(CO) = 5 MPa, ω

substrate

 = 10 wt.%., ω(H

2

O) = 20 wt.% , СО:Н

2

О = 1.8 (molar ratio).
Fig. 1:

Conversion of (a) 2MN (Т = 380 °С) and (b) 1MN (t = 6 h) over Ni–Mo sulfide catalysts depending on surfactant presence. Reaction conditions: ω(SPAN-80) = 4 wt.%, p(CO) = 5 MPa, ω substrate  = 10 wt.%., ω(H 2 O) = 20 wt.% , СО:Н 2 О = 1.8 (molar ratio).

To evaluate the influence of surfactant concentration on catalytic activity the mass content of SPAN-80 was varied from 2 to 8 wt.% (Fig. 2). On rising surfactant content, the catalytic activity increases achieving maximum at 4 wt.% and was unaffected by higher concentration of SPAN-80. Such a significant difference in activity depending on surfactant is caused by its influence on sulfide phase morphological characteristics and metal distribution over Ni and Mo species at the surface of sulfide catalysts.

Fig. 2: 
Influence of surfactant concentration on 2MN conversion over Ni–Mo sulfide catalysts. Reaction conditions: Т = 380 °С, p(CO) = 5 MPa, ω(2MN) = 10 wt.%., ω(H

2

O) = 20 wt.%, t = 4 h, СО:Н

2

О = 1.8 (molar ratio).
Fig. 2:

Influence of surfactant concentration on 2MN conversion over Ni–Mo sulfide catalysts. Reaction conditions: Т = 380 °С, p(CO) = 5 MPa, ω(2MN) = 10 wt.%., ω(H 2 O) = 20 wt.%, t = 4 h, СО:Н 2 О = 1.8 (molar ratio).

To evaluate the characteristics of active phase species, the catalyst formed during hydrogenation from water-soluble metal precursors was separated by filtration followed by step-by-step washing with toluene and hexane. Then, it was dried in inert atmosphere and evaluated by XRD, XRF, TEM and XPS techniques.

Morphological characteristics of active phase species

The catalysts reveal the presence of a typical multilayer sulfide structure (Fig. 3) with interlayer distance approximately of 0.62–0.63 nm differs from basal plane of MoS2 crystalline [47]. According to calculation based on statistical evaluation of sulfide particles from TEM images, NiMoS + SPAN has a narrow distribution of slab length (Fig. 3) with average stacking number of 3.8 (Table 4a). For NiMoS the slabs are 6.0–18.0 nm in length within comparable content of sulfide particles having the length 6.0–12.0 nm and 12.0–18.0 nm. Meanwhile the number of stacks per particle is about 4.5. The formation of multilayered particles with stacking number of 6–9 also takes place from the presence of surfactant. Thus, the surfactant-assisted synthesis leads to increase the dispersion of active phase characterized by highly stacked of 3–4 layers in MoS2 particles with 14.5 nm lengths providing higher activity of this catalyst.

Fig. 3: 
TEM microphotographs and distribution of slab length for (a) NiMoS + SPAN and (b) NiMoS catalysts.
Fig. 3:

TEM microphotographs and distribution of slab length for (a) NiMoS + SPAN and (b) NiMoS catalysts.

Table 4a:

Characteristics of sulfide phase species calculated from TEM micrographs.

Catalyst L ¯ , nm1 N ¯ 2 D 3 Distribution of stacking number (rel.%)
1 2 3 4 5 6 >6
NiMoS 15.0 4.5 0.08 2 11 20 24 20 10 13
NiMoS + SPAN 14.5 3.8 0.09 2 21 27 19 14 6 11

1The average slab length of sulfide particles (Eq. 1).

2The average stacking number (Eq. 1).

3Active phase dispersion (Eq. 7).

XRD pattern of the NiMoS + SPAN catalyst are shown in Fig. 4. The peaks at 2θ = 14.5° (hkl = 002), 33.0° (hkl = 101) and 58.4° (hkl = 110) correspond to MoS2 [48]. Nevertheless, the XRD pattern shows poor crystallinity with a weak peak at 2θ ≈ 14°, characterizing a small stacking of MoS2 layers caused by synergistic effect of Ni and Mo atoms. The second reason is the intercalation of organic and inorganic molecules into the interplanar space of MoS2 phase leading to violation of the coplanarity of crystallographic planes. The (101) and (310) reflections at 2θ = 31.8° and 58.4° respectively are ascribed to molybdenum oxides [49], while peaks at 2θ = 30.0° and 34.4° (hkl = 101) correspond to NiS phase [50]. The reflections at 2θ = 45.5° and 53.3° indicates NiS2 or NixMoSy phase [50]. The reflections around 27°, 30°, and 44° attributed to α-NiMoO4 phase are not quite resolved, nevertheless the nickel molybdate formation should not to exclude [51], [52].

Fig. 4: 
XRD pattern of NiMoS + SPAN catalyst.
Fig. 4:

XRD pattern of NiMoS + SPAN catalyst.

According to XRF data (Table 4b), the composition of NiMoS is similar to NiMoS + SPAN catalysts. When the surfactant is used the Ni content is slightly lower compared to NiMoS catalyst. These results are in a full accordance with XPS data (Table 4b) revealing the active phase composition and chemical state of Mo, Ni and S.

Table 4b:

The composition of sulfide catalysts (from XRF and XPS data).

Catalyst Elemental composition, rel.% (from XRF) Surface atomic concentration, at.% (from XPS) Surface atomic ratio (from XPS)
Mo Ni S Mo 3d Ni 2p S 2p Ni/Mo S/(Ni + Mo)
NiMoS 82.0 12.2 5.8 13.3 3.9 25.4 0.29 1.5
NiMoS + SPAN 83.7 9.7 6.6 12.3 2.2 28.3 0.18 2.0

The Mo3d-S2s, Ni2p and S2p core level spectra are shown in Fig. 5. The peak at 226.0 eV is characteristic for S2− ions of S2s spectra [19], [20]. For both catalysts the most intense photoelectron Mo3d doublet with binding energy of 228.9 and 232.0 eV corresponds to Mo4+ species of the MoS2 phase, while two peaks at 232.8 and 235.9 eV is attributed to NiMoO4 as well as Mo6+ oxidation state [20], [25], [53], [54]. Quantitative XPS analysis of the Mo 3d5/2 core level spectra shows that Mo atoms are predominantly in sulfide with relative content of 71.2 and 85.0 for NiMoS and NiMoS + SPAN catalysts respectively (Table 5).

Fig. 5: 
Mo3d-S2, Ni 2p and S 2p core level spectra of (a) NiMoS + SPAN and (b) NiMoS catalysts.
Fig. 5:

Mo3d-S2, Ni 2p and S 2p core level spectra of (a) NiMoS + SPAN and (b) NiMoS catalysts.

Table 5:

The quantitative XPS analysis of the Mo 3d 5/2 , Ni 2p 3/2 and S 2p 3/2 core levels.

Catalyst Mo 3d 5/2 , eV (content, %) Ni 2p 3/2 , eV (content, %) S 2p 3/2 , eV (content, %)
MoS2 Mo6+ NiMoS Ni+2 NiMoO4 S2– S2 2– SO4 2–
NiMoS 228.9 (71.2) 232.8 (28.8) 854.5 (25.2) 856.3 (20.8) 857.1 (54.0) 162.2 (52.8) 163.0 (27.9) 168.9 (19.3)
NiMoS + SPAN 228.9 (85.0) 232.9 (15.0) 854.0 (49.8) 856.1 (17.2) 857.4 (33.0) 162.0 (72.3) 163.0 (14.7) 168.7 (13.0)

For Ni 2p core level spectra a strong difference is observed. The binding energy region between 854.0 and 854.5 eV corresponds to NiMoS phase [19], [20], [25], whose the relative content for NiMoS + SPAN catalyst reaches almost 50%, while Ni 2p core level spectra for NiMoS counterpart indicates the formation of NiMoO4 (54.0%) predominantly. It is confirmed by doublet peaks at 857.1 and 874.3 eV [53]. Thus, Ni is fully in mixed oxide or sulfide phases. The XPS data also confirm that NiS with characteristic binding energy of 852.9–853.0 eV doesn’t form. Meanwhile peaks at 856.1–856.3 eV is assigned to Ni+2 species either in oxide or salts form [19], [20]. The peak attributed to NiMoS phase in Ni 2p core level spectrum for NiMoS + SPAN catalyst slightly shifted to lower binding energies at 0.5 eV caused by the residual presence of carbon species. In the S2p3/2 core level spectra two peaks at 162.0–162.2 and 163.0 eV are assigned to S2− and (S2)2− sulfur species respectively [19], [20]. The photoelectron peak with binding energies at the range of 168.7–168.9 eV is attributed to sulfur in SO4 2– form [19], [20]. While for NiMoS + SPAN the S percentage in sulfide form is close to 72%, in the case of NiMoS it doesn’t exceed 53%. The latter is characterized by higher content of (S2)2− sulfur species (27.9 vs. 14.7%) and sulfate state (19.3 vs. 13.0%). The surfactant-assisted synthesis probably provides better dispersion of catalysts precursors leading to formation of nickel molybdate. Besides, surfactant is favorable for sulfidation and presumably provides the organic species that remain in the precipitate and are transformed to carbonaceous matter on further sulfidation [20]. The sulfate species (SO4 2–) may be formed during migration of sulfur atoms from the metal centers onto the O sites during nickel molybdate sulfidation [55].

Based on calculation results, the effective amount of Ni in active phase species for NiMoS is close to 2.5 wt.%, while those obtained for NiMoS + SPAN catalyst is 2.9 wt.%. When the surfactant is used, the effective Mo content in MoS2 phase reaches up to 45%, which is 6% higher compared to NiMoS counterpart (Table 6). The Ni/Mo ratios in the NiMoS slabs as well as on edges are equal for both catalysts.

Table 6:

Characteristics for active phase of sulfide catalysts (from XPS).

Catalyst ωNiMoS 1, wt.% ωMoS 2 , wt.% p(Ni/Mo) s 2 p(Ni/Mo) e 3
NiMoS 2.5 39.2 0.10 1.20
NiMoS + SPAN 2.9 45.3 0.10 1.11

1effective Ni content in NiMoS and Mo amount in MoS2 phases (Eq. 9, Eq. 10).

2Ni/Mo atomic ratios in the NiMoS slabs (Eq. 11).

3Ni/Mo atomic ratios on the edges of active phase (Eq. 12).

Catalytic testing

The hydrogenation of aromatics was chosen as a test reaction to evaluate the behavior of Ni–Mo sulfide nanosized catalysts assembled in situ from water-soluble metal precursors in W/O emulsions stabilized by surfactant. The catalytic evaluation of more active NiMoS + SPAN catalyst was performed in a high-pressure batch reactor at T = 380–420 °С, p(CO) = 5 MPa with water content of 20 wt.% and CO/H2O molar ratio of 1.8 for 4–8 h. At a temperature above 420–430 °С, the thermal cracking or dehydrogenation of hydrocarbons takes place, so the temperature range was limited to 420 °С. At a temperature below 320 °С the aromatics ring hydrogenation is affected by kinetic factor [56], [57], [58]. This range is also determined by conditions, when the metal precursors are decomposed to oxides followed by sulfidation forming active phase. The lower the temperature, the longer time is needed for active phase formation. Recently we established the reaction induction period where the catalysts assembly depends on temperature and varies from 2 to 3 h at 360–400 °C [25], [26], [27], so the reaction time was not lower than 4 h.

The temperature-dependent activity of Ni–Mo sulfide catalysts in hydrogenation of model substrates is shown in Fig. 6. The anthracene conversion reaches to 97–100% over the whole temperature range, while for 1MN and 2MN it doesn’t exceed 92 and 86% respectively even at 420 °С. This is in a full accordance with data reported [59], where the reactivity of polycyclic aromatics in hydrogenation increases in contrast with that of monocyclic aromatics.

Fig. 6: 
Temperature-dependent conversion of model substrates over Ni-Mo sulfide catalysts. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10wt.%., ω(H2O) = 20 wt.% , t = 6 h, ω(SPAN-80) = 4wt.%, CO:H2O = 1.8 (molar ratio).
Fig. 6:

Temperature-dependent conversion of model substrates over Ni-Mo sulfide catalysts. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10wt.%., ω(H2O) = 20 wt.% , t = 6 h, ω(SPAN-80) = 4wt.%, CO:H2O = 1.8 (molar ratio).

Anthracene hydrogenation (Scheme 3) proceeds faster in comparison with methyl-substituted naphthalenes. Thus, the 97% conversion was achieved at 380 °C after 4 h. The reaction products are comprised of dihydroanthracene (2HA), tetrahydroanthracene (4HA), octahydroanthracene (8HN) as well as trace amount of perhydride (14HA). At the temperature above 420 °C thermal cracking occurs.

Scheme 3: 
The pathway for anthracene hydrogenation [59].
Scheme 3:

The pathway for anthracene hydrogenation [59].

At 380 °C the anthracene hydrogenation results in 2HA mainly (Fig. 7a). It is due to reactivity of the conjugated double bonds in the anthracene molecule, when the carbon atoms at positions 9 and 10 have high delocalized electron density facilitating the hydrogen consumption [59], [60]. Nevertheless, 2HA content decreases with reaction time accompanied by increase of 4HA and 8HA amounts. On rising temperature to 400 °C, the 4HA becomes the main reaction product with selectivity of 50% at 8 h. The 2HA content decreases while 8HA accumulates in reaction products (Fig. 7b). At higher temperature up to 420 °C, the products distribution is an analog of the one at 380 °C with higher 8HA initial content being time independent (Fig. 7c). Thus, at 400 °С hydrogen consumption is maximal and minimal at 420 °С (Fig. 7d). At higher temperatures and hydrogen pressure deficiency, dehydrogenation may proceed leading to formation of 4HA or 2HA.

Fig. 7: 
Product distribution for anthracene hydrogenation at (a) 380 °C, (b) 400 °C and 420 °C over Ni–Mo sulfide catalysts and (d) hydrogen consumption depending on reaction temperature. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10 wt.%., ω(H

2

O) = 20 wt.%, ω(SPAN-80) = 4 wt.%, СО:Н

2

О = 1.8 (mol).
Fig. 7:

Product distribution for anthracene hydrogenation at (a) 380 °C, (b) 400 °C and 420 °C over Ni–Mo sulfide catalysts and (d) hydrogen consumption depending on reaction temperature. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10 wt.%., ω(H 2 O) = 20 wt.%, ω(SPAN-80) = 4 wt.%, СО:Н 2 О = 1.8 (mol).

In the case of methyl-substituted naphthalenes the lower conversion may be caused by steric hindrances during the adsorption of substrate molecules on the active catalytic sites. The time and temperature dependencies of catalytic activity in 2MN and 1MN hydrogenation are presented in Fig. 8. To achieve higher conversions, the reaction temperature should be at least 380 °С and time at least 4 h. On increasing the temperature by 20 °С (from 360 to 380 °С), the 2MN conversion at 6 h increase up to 27%. With the lapse of reaction time to 8 h at 400–420 °С the conversion rises exceeding 90% conversions. Under reaction conditions 1MN and 2MN transform into 5- and 6-methyltetralines (5MT and 6MT), i. e., unsubstituted aromatic rings are hydrogenated probably due to steric effect of methyl group during the sorption. Thus, at 380 °С and 4 h the 6MT:2MT and 5MT:1MT ratios are approximately 1:1. After increasing temperature and reaction time the 6MT and 5MT contents are 1.2–1.5 times higher than 2MT and 1MT respectively. Meanwhile, the selectivity to 6MT and 5MT varies in the range of 50–53%. The higher the temperature and reaction time, the higher the concentration of decalines: 21% for 1MN and 19% for 2MN at 420 °С after 8 h. The Ni–Mo sulfide catalysts were found to be slightly more active in 2MN hydrogenation at least at 380 °С. Despite of comparable 1MN and 2MN conversions, the decalines’ content in hydrogenation products of 2MN is 2–3% lower.

Fig. 8: 
Product distribution for hydrogenation reaction of (a) 1MN and (b) 2MN over Ni–Mo sulfide catalysts. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10 wt.%., ω(H

2

O) = 20 wt.%, ω(SPAN-80) = 4 wt.%, СО:Н

2

О = 1.8 (mol).
Fig. 8:

Product distribution for hydrogenation reaction of (a) 1MN and (b) 2MN over Ni–Mo sulfide catalysts. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10 wt.%., ω(H 2 O) = 20 wt.%, ω(SPAN-80) = 4 wt.%, СО:Н 2 О = 1.8 (mol).

The substituent effect in aromatic naphthalene rings on catalytic activity is presented in Fig. 9. The experimental results showed that hydrogenation activity decreases following trend 1,8-dimethylnaphthalene>>1,3-dimethylnaphthalene>2,6-dimethylnaphthalene≈2,3-dimethylnaphthalene>2-ethylnaphthalene.

Fig. 9: 
Product distribution for hydrogenation reaction of (a) 1MN and (b) 2MN over Ni–Mo sulfide catalysts. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10 wt.%., ω(H

2

O) = 20 wt.%, ω(SPAN-80) = 4 wt.%, СО:Н

2

О = 1.8 (mol).
Fig. 9:

Product distribution for hydrogenation reaction of (a) 1MN and (b) 2MN over Ni–Mo sulfide catalysts. Reaction conditions: p(CO) = 5 MPa, ω(substrate) = 10 wt.%., ω(H 2 O) = 20 wt.%, ω(SPAN-80) = 4 wt.%, СО:Н 2 О = 1.8 (mol).

The highest conversion of 45% was achieved for 1,8-dimethylnaphthalene with 100% selectivity to tetralines. For 1,3-dimethyl-substituted substrate the conversion decreases dramatically to 10% being comparable to 2,6-dimethylnaphthalene. The conversion to tetralines not higher than 5% was found for 2,3-dimethyl- and 2-ethyl-naphthalenes. Similar to 1- and 2-methyl-naphthalenes the hydrogenation of asymmetric dimethyl-substituted substrate carries out through the unsubstituted aromatic ring indicating that steric factors probably influence on the sorption mechanism over active metal sites.

Reuse and recyclability

The recycling tests were carried out for 2MN hydrogenation at 380 °С for 6 h with elemental sulfur (the curve «+S» in the Fig. 10) for active phase formation and without it (the curve «-S» in the Fig. 10). At the first cycle, during 2MN hydrogenation, active sulfide particles formed by decomposition of metal precursors and sulfidation of oxides. Then, the reactor was cooled, depressurized and disassembled, reaction products were separated by decantation and in situ assembled catalyst was used without regeneration many times in 2MN hydrogenation. The catalyst was found to be more active in the second cycle without the induction period for the formation of active phase taking place for the first time (Fig. 10). For the next four cycles in the sulfur-assisted process the 2MN conversion slightly decreased, being maintained at 93–95% level. However in the absence of sulfur, the catalyst activity decreases dramatically with cycle number. Thus, the 2MN conversion after six cycles doesn’t exceed 25% that may be caused by the influence of water on active sulfide phase as well as a lack of hydrogen sulfide in the reaction system leading to metal oxides formation. The H2S generated under reaction conditions exits the system when the reactor is depressurized and disassembled after second reaction cycle thereby sulfiding agent is absent when catalyst is reused.

Fig. 10: 
Recycling tests of Ni–Mo sulfide catalysts in 2MN hydrogenation. Reaction conditions: p(CO) = 5 MPa, t = 6 ч, ω(substrate) = 10 wt.%., ω(H

2

O) = 20 wt.% , ω(SPAN-80) = 4 wt.%, ω(S) = 2.5 wt.

%
 (when the sulfur added)

,
 СО:Н

2

О = 1.8 (molar ratio).
Fig. 10:

Recycling tests of Ni–Mo sulfide catalysts in 2MN hydrogenation. Reaction conditions: p(CO) = 5 MPa, t = 6 ч, ω(substrate) = 10 wt.%., ω(H 2 O) = 20 wt.% , ω(SPAN-80) = 4 wt.%, ω(S) = 2.5 wt. % (when the sulfur added) , СО:Н 2 О = 1.8 (molar ratio).

To evaluate the characteristics of active phase species for the NiMoS + SPAN catalyst after six cycles of 2MN sulfur-assisted hydrogenation, it was separated by filtration followed by step-by-step washing with toluene and hexane. The XPS spectra of reused catalyst revealed the Ni and Mo atomic concentration was close to initial composition (Table 8). The Mo content decreased by 0.3at.% while Ni content was maintained at 2.2at.%. The S/(Ni + Mo) ratio is also comparable (2.0 vs. 1.9). It confirms, that at the first 2–3 h, which is ascribed to the induction period, the oxide catalyst precursors formed slowly by metal salt decomposition, which reveals that it is the rate-determining step. The sulfidation is rather fast based on high catalytic activity data upon recycling. Deconvolution of Mo3d-S2s core level spectra revealed that 75.1% of Mo is in sulfide state (Table 7) with doublet peaks at 228.7 and 231.8 eV (Fig. 11) [19], [20], [61]. The characteristic binding energy of 230.1 eV is ascribed to oxysulfide which contributed 11.4 % and 13.5% of Mo6+ [13], [19], [20], [61]. Taking into account the same reaction condition from cycle to cycle it is to be assumed that the minor difference in Mo valency state for initial and recycled catalysts is caused by the its contact with air before XPS analysis which resulted in MoS2 transformation to MoSxOy form. It is also confirmed by the decrease of NiMoS phase as well as sulfide to 29.3 and 68.9% respectively. Meanwhile, the NiMoO4 content is more than 22% higher compared to the initial composition of the catalysts. The effective Mo content in MoS2 phase decreases to 40.3%, while the ωNiMoS content is 1.8 wt.% and the Ni/Mo ratio in the NiMoS slabs is the same as the initial catalysts (Table 8).

Table 7:

The quantitative XPS analysis of the Mo 3d 5/2 , Ni 2p 3/2 and S 2p 3/2 core levels for NiMoS + SPAN catalyst after six cycles of 2MN sulfur-assisted hydrogenation.

Mo 3d 5/2 , eV (content, %) Ni 2p 3/2 , eV (content, %) S 2p 3/2 , eV (content, %)
MoS2 MoSxOy Mo6+ NiMoS Ni+2 NiMoO4 S2– S2 2– SO4 2–
228.7 (75.1) 230.1 (11.4) 232.6 (13.5) 853.7 (29.3) 855.9 (19.8) 857.8 (50.9) 161.5 (68.9) 163.0 (14.7) 168.7 (16.4)
Table 8:

Characteristics for active phase of NiMoS + SPAN catalyst after six cycles of 2MN sulfur-assisted hydrogenation (from XPS).

Surface atomic concentration, at.% (from XPS) Surface atomic ratio (from XPS) ωNiMoS 1, wt.% ωMoS 2 ,wt.% p(Ni/Mo) s 2
Mo 3d Ni 2p S 2p Ni/Mo S/(Ni + Mo)
12.0 2.2 27.1 0.18 1.9 1.8 40.3 0.07

1effective Ni content in NiMoS and Mo amount in MoS2 phases (Eq. 9, Eq.10).

2Ni/Mo atomic ratios in the NiMoS slabs (Eq. 11).

Fig. 11: 
Mo3d-S2s, Ni 2p and S 2p core level spectra for NiMoS + SPAN catalyst after six cycles of 2MN sulfur-assisted hydrogenation.
Fig. 11:

Mo3d-S2s, Ni 2p and S 2p core level spectra for NiMoS + SPAN catalyst after six cycles of 2MN sulfur-assisted hydrogenation.

Conclusions

The unsupported catalysts were obtained during hydrogenation by in situ high-temperature decomposition (above 300 °C) of water-soluble metal precursors in W/O emulsions stabilized by surfactant under CO pressure. Elemental sulfur was used as a sulfiding agent. Based on experimental results, the catalyst obtained by surfactant-assistant synthesis is more active in hydrogenation of aromatics and 4 wt.% is the optimal SPAN-80 content. The surfactant was found to be affective in both dispersion and metal distribution for Ni and Mo species, allowing for the formation of sulfide particles with shorter slab length and stacking (according to calculation based on TEM data) within high content of NiMoS phase revealed from XPS data. Thus, the surfactant-assisted synthesis leads to highly stacked layers (3–4) in MoS2 particles with 14.5 nm in length. XPS spectra reveal 85.0% of MoS2 within 49.8% of NiMoS phase. Moreover, NiS was not observed and Ni is fully in mixed oxide or sulfide phases. The effective Mo content in MoS2 phase reached up to 45%, which is 6% higher compared to its NiMoS counterpart. For surfactant-assisted synthesis Ni content is slightly lower within higher S/(Ni + Mo) ratio.

As shown experimentally with unsupported Ni–Mo sulfide catalysts, the activity of aromatic rings depends on the substituent therein and decreases as follows: anthracene>>1-methylnaphthalene≈2-methylnaphthalene>1,8-dimethylnaphthalene>>1,3-dimethylnaphthale-ne>2,6-dimethylnaphthalene≈2,3-dimethylnaphthalene>2-ethylnaphthalene. The anthracene conversion reaches up to 97–100% for 4 h over the whole temperature range, while for 1MN and 2MN it doesn’t exceed 92 and 86% respectively even at 420 °С for 8 h. Under reaction conditions 1MN and 2MN transform into 5- and 6-methyltetralines (5MT and 6MT) and selectivity to 6MT as well as 5MT varies in the range of 50–53%, i. e., unsubstituted aromatic rings are hydrogenated probably due to steric effect of methyl group during the sorption. The higher the temperature and reaction time, the higher decalines’ content observed: 21% for 1MN and 19% for 2MN at 420 °С for 8 h. For dimethyl-substituted aromatics the higher conversion of 45% was achieved for 1,8-dimethylnaphthalene with 100% selectivity to tetralines at 400 °C for 6 h. For 1,3-dimethyl-substituted substrate the conversion decreases dramatically to 10% being comparable to 2,6-dimethylnaphthalene. The conversion not higher than 5% was found for 2,3-dimethyl- and 2-ethyl-naphtalenes. Similar to 1- and 2-methylnaphtalenes the hydrogenation of asymmetric dimethyl-substituted substrate carries out through the unsaturated aromatic ring indicating steric factors influence on the sorption mechanism over active metal sites.

The catalysts were found to be reused at least for six cycles when the hydrogenation is sulfur-assisted preventing metal oxides formation. Based on the recycling tests and XPS data, it was confirmed, that in the first 2–3 h referred to as the induction period, the oxide catalyst precursors formed slowly by metal salts decomposition, which reveals that it is the rate-determining reaction step. The sulfidation is rather fast based on high catalytic activity data on 2MN conversion retaining at 93–95% upon recycling.


Corresponding author: Anna Vutolkina, Faculty of Chemistry, Department of Petroleum Chemistry and Organic Catalysis, Lomonosov Moscow State University, GSP-1, 1-3 Leninskiye Gory, 119991, Moscow, Russia, E-mail:

Article note: A collection of papers from the 18th IUPAC International Symposium Macromolecular-Metal Complexes (MMC-18), held at the Lomonosov Moscow State University, 10–13 June 2019.


Award Identifier / Grant number: № 19-79-00259

Acknowledgments

We thank Dr. Yusuf Darrat from Louisiana Tech University (USA) for English editing.

  1. Funding: The reported study was financially supported by Russian Science Foundation (RSF) grant (project No 19-79-00259).

References

[1] H. Luik, L. Luik, I. Johannes, L. Tiikma, N. Vink, V. Palu, M. Bitjukov, H. Tamvelius, J. Krasulina, K. Kruusement, I. Nechaev. Fuel Process. Technol. 124, 115 (2014).10.1016/j.fuproc.2014.02.018Search in Google Scholar

[2] A. Marafi, H. Albazzaz, M. S. Rana. Catal. Today 329, 125 (2019).10.1016/j.cattod.2018.10.067Search in Google Scholar

[3] R. Mythili, P. Venkatachalam, P. Subramanian, D. Uma. Bioresour. Technol. 138, 71 (2013).10.1016/j.biortech.2013.03.161Search in Google Scholar PubMed

[4] A. Alvarez-Majmutov, R. Gieleciak, J. Chen. Fuel. 241, 744 (2019).10.1016/j.fuel.2018.12.096Search in Google Scholar

[5] P. C. S. Zorzenão, R. M. Mariath, F. E. Pinto, L. V. Tose, W. Romão, A. F. Santos, A. P. Scheer, S. Simon, J. Sjöblom, C. I. Yamamoto. J. Petrol. Sci. Eng. 160, 1 (2018).10.1016/j.petrol.2017.10.014Search in Google Scholar

[6] E. Lazzari, T. Schena, M. C. A. Marcelo, C. T. Primaz, A. N. Silva, M. F. Ferrão, T. Bjerk, E. B. Caramão. Ind. Crops Prod. 111, 856 (2018).10.1016/j.indcrop.2017.11.005Search in Google Scholar

[7] S. V. Egazar’yants, V. A. Vinokurov, A. V. Vutolkina, M. Y. Talanova, V. I. Frolov, E. A. Karakhanov. Chem. Technol. Fuels Oils. 51, 506 (2015).10.1007/s10553-015-0632-7Search in Google Scholar

[8] O. V. Klimov, K. A. Nadeina, P. P. Dik, G. I. Koryakina, V. Y. Pereyma, M. O. Kazakov, S. V. Budukva, E. Y. Gerasimov, I. P. Prosvirin, D. I. Kochubey, A. S. Noskov. Catal. Today 271, 56 (2016).10.1016/j.cattod.2015.11.004Search in Google Scholar

[9] E. A. Mazlova, S. V. Meshcheryakov. Chem. Technol. Fuels Oils. 35, 49 (1999).10.1007/BF02694263Search in Google Scholar

[10] L. Silva, F. C. Alves, F. França. Waste Manag. Res.: J. Int. Solid Wastes Pub. Clean. Assoc., ISWA. 30, 1016 (2012).10.1177/0734242X12448517Search in Google Scholar PubMed

[11] J. Chen, J. Zhang, J. Mi, E. D. Garcia, Y. Cao, L. Jiang, L. Oliviero, F. Maugé. Int. J. Hydrogen Energy 43, 7405 (2018).10.1016/j.ijhydene.2018.02.194Search in Google Scholar

[12] Y.-Y. Chen, M. Dong, J. Wang, H. Jiao. J. Phys. Chem. C. 116, 25368 (2012).10.1021/jp308383rSearch in Google Scholar

[13] R. Z. Lee, F. T. T. Ng. Catal. Today 116, 505 (2006).10.1016/j.cattod.2006.06.032Search in Google Scholar

[14] G. K. Reddy, P. G. Smirniotis. in Water Gas Shift Reaction, G. K. Reddy, P. G. Smirniotis (Eds.), p.101, Elsevier (2015).10.1016/B978-0-12-420154-5.00004-8Search in Google Scholar

[15] Y. Sun, S. S. Hla, G. J. Duffy, A. J. Cousins, D. French, L. D. Morpeth, J. H. Edwards, D. G. Roberts. Int. J. Hydrogen Energy 36, 79 (2011).10.1016/j.ijhydene.2010.09.093Search in Google Scholar

[16] A. V. Mozhaev, P. A. Nikulshin, A. A. Pimerzin, K. I. Maslakov, A. A. Pimerzin. Catal. Today 271, 80 (2016).10.1016/j.cattod.2015.11.002Search in Google Scholar

[17] A. A. Pimerzin, P. A. Nikul’shin, A. V. Mozhaev, A. Pimerzin. A. Petrol. Chem. 53, 245 (2013).10.1134/S0965544113030067Search in Google Scholar

[18] A. A. Pimerzin, A. A. Roganov, P. A. Nikul’shin, A. A. Pimerzin. Chem. Technol. Fuels Oils. 53, 654 (2017).10.1007/s10553-017-0847-xSearch in Google Scholar

[19] A. N. Varakin, A. V. Mozhaev, A. A. Pimerzin, P. A. Nikulshin. Catal. Today. 2019), in press.Search in Google Scholar

[20] D. Genuit, P. Afanasiev, M. Vrinat. J. Catal. 235, 302 (2005).10.1016/j.jcat.2005.08.016Search in Google Scholar

[21] M. Kniazeva, A. Maksimov. Catalysts. 8, 644 (2018).10.3390/catal8120644Search in Google Scholar

[22] M. I. Knyazeva, D. I. Panyukova, A. L. Maximov. Petrol. Chem. 59, 504 (2019).10.1134/S0965544119050049Search in Google Scholar

[23] I. A. Sizova, A. B. Kulikov, M. I. Onishchenko, S. I. Serdyukov, A. L. Maksimov. Petrol. Chem. 56, 44 (2016).10.1134/S0965544115080174Search in Google Scholar

[24] S. Harris, R. R. Chianelli. J. Catal. 98, 17 (1986).10.1016/0021-9517(86)90292-7Search in Google Scholar

[25] A. V. Vutolkina, D. F. Makhmutov, A. V. Zanina, A. L. Maximov, A. P. Glotov, N. A. Sinikova, E. A. Karakhanov. Petrol. Chem. 58, 528 (2018).10.1134/S0965544118070095Search in Google Scholar

[26] A. V. Vutolkina, A. V. Akopyan, A. P. Glotov, M. S. Kotelev, A. L. Maksimov, E. A. Karakhanov. Russ. J. Appl. Chem. 91, 981 (2018).10.1134/S1070427218060162Search in Google Scholar

[27] A. V. Vutolkina, D. F. Makhmutov, A. V. Zanina, A. L. Maximov, D. S. Kopitsin, A. P. Glotov, S. V. Egazar’yants, E. A. Karakhanov. Petrol. Chem. 58, 1227 (2018).10.1134/S0965544118140141Search in Google Scholar

[28] G. Berhault, M. Perez De la Rosa, A. Mehta, M. J. Yácaman, R. R. Chianelli. Appl. Catal., A: Gen. 345, 80 (2008).10.1016/j.apcata.2008.04.034Search in Google Scholar

[29] S. Kasztelan, H. Toulhoat, J. Grimblot, J. P. Bonnelle. Appl. Catal. 13, 127 (1984).10.1016/S0166-9834(00)83333-3Search in Google Scholar

[30] M. T. Nguyen, N. T. Nguyen, J. Cho, C. Park, S. Park, J. Jung, C. W. Lee. J. Ind. Eng. Chem. 43, 1 (2016).10.1016/j.jiec.2016.07.057Search in Google Scholar

[31] Y. Villasana, M. A. Luis-Luis, F. Mendez, H. Labrador, J. Brito. in Oil and natural Gas in Energy Science and Technlogy Series, U. C. Sharma, R. Prasad, S. Sivakumar (Eds.), Vol. 3, p. 304, Studium Press LLC (2015).Search in Google Scholar

[32] R. R. Chianelli, M. H. Siadati, M. P. De la Rosa, G. Berhault, J. P. Wilcoxon, R. Bearden, B. L. Abrams. Catal. Rev. 48, 1 (2006).10.1080/01614940500439776Search in Google Scholar

[33] J. Ancheyta, M. S. Rana, E. Furimsky. Catal. Today 109, 3 (2005).10.1016/j.cattod.2005.08.025Search in Google Scholar

[34] J. G. Speight. Heavy and Extra-heavy Oil Upgrading Technologies, Elsevier Science (2013).10.1016/B978-0-12-404570-5.00001-6Search in Google Scholar

[35] J. Dang, G.-H. Zhang, K.-C. Chou, R. G. Reddy, Y. He, Y. Sun. Int. J. Refract. Met. Hard Mater. 41, 216 (2013).10.1016/j.ijrmhm.2013.04.002Search in Google Scholar

[36] T. N. Kovács, D. Hunyadi, A. L. A. de Lucena, I. M. Szilágyi. J. Therm. Anal. Calorim. 124, 1013 (2016).10.1007/s10973-015-5201-0Search in Google Scholar

[37] P. J. Mangnus, A. Bos, J. A. Moulijn. J. Catal. 146, 437 (1994).10.1006/jcat.1994.1081Search in Google Scholar

[38] P. J. Mangnus, E. K. Poels, A. D. van Langeveld, J. A. Moulijn. J. Catal. 137, 92 (1992).10.1016/0021-9517(92)90141-4Search in Google Scholar

[39] I. B. Sharma, S. Batra. J. Therm. Anal. 34, 1273 (1988).10.1007/BF01914351Search in Google Scholar

[40] A. Steinbrunn, C. Lattaud. Surf. Sci. 155, 279 (1985).10.1016/0039-6028(85)90418-2Search in Google Scholar

[41] Ž. D. Živković, D. T. Živković, D. B. Grujičić. J. Therm. Anal. Calorim. 53, 617 (1998).10.1023/A:1010170231923Search in Google Scholar

[42] Y. Zhang, Z. Zhang, T. Zhou, P. Lu, Y. Gao, F. Yu, A. Umar, Q. Wang. RSC Adv. 6, 54553 (2016).10.1039/C6RA12437BSearch in Google Scholar

[43] H. Oudghiri Hassani, F. Wadaani. Molecules. 23, 273 (2018).10.3390/molecules23020273Search in Google Scholar PubMed PubMed Central

[44] H. Oudghiri Hassani. Catal. Commun. 60 (2014).10.1016/j.catcom.2014.11.019Search in Google Scholar

[45] J. n. L. Brito, A. L. Barbosa. J. Catal. 171, 467 (1997).10.1006/jcat.1997.1796Search in Google Scholar

[46] W. C. Griffin. J. Soc. Cosmet. Chem. 1, 311 (1949).Search in Google Scholar

[47] P. Liu, J. Zhu, J. Zhang, P. Xi, K. Tao, D. Gao, D. Xue. ACS Energy Lett. 2, 745 (2017).10.1021/acsenergylett.7b00111Search in Google Scholar

[48] F. Yang, Z. Zhang, Y. Wang, M. Xu, W. Zhao, J. Yan, C. Chen. Mater. Res. Bull. 87, 119 (2017).10.1016/j.materresbull.2016.11.029Search in Google Scholar

[49] M. C. Morris, H. F. McMurdie, E. H. Evans, B. Paretzkin, H. S. Parker, N. C. Panagiotopoulos, R. C. Hubbard. Standard X-Ray Diffraction Powder Patterns: Section 18. Data for 58 Substances, U.S. Department of Commerce, National Bureau of Standards (1981).10.6028/NBS.MONO.25-18Search in Google Scholar

[50] D. Zhang, H. Mou, F. Lu, C. Song, D. Wang. Appl. Catal., B: Environmental. 254, 471 (2019).10.1016/j.apcatb.2019.05.029Search in Google Scholar

[51] P. K. Bankar, S. Ratha, M. A. More, D. J. Late, C. S. Rout. Appl. Surf. Sci. 418, 270 (2017).10.1016/j.apsusc.2017.02.177Search in Google Scholar

[52] R. S. Arshadi, R. Mamoory, F. Dabir, N. Blomquist, M. Phadatare, H. Olin. Crystals. 9, 31 (2019).10.3390/cryst9010031Search in Google Scholar

[53] S. Kumar Ray, D. Dhakal, G. Gyawali, B. Joshi, A. Raj Koirala, S. Wohn Lee. Chem. Eng. J. 373, 259 (2019).10.1016/j.cej.2019.05.041Search in Google Scholar

[54] Y. Yi, C. T. Williams, M. Glascock, G. Xiong, J. Lauterbach, C. Liang. Mater. Res. Bull. 56, 54 (2014).10.1016/j.materresbull.2014.04.028Search in Google Scholar

[55] J. A. Rodriguez, S. Chaturvedi, J. C. Hanson, J. L. Brito. J. Phys. Chem. B 103, 770 (1999).10.1021/jp983115mSearch in Google Scholar

[56] S. A. Ali. Pet. Sci. Technol. 25, 1293 (2007).10.1080/10916460500528607Search in Google Scholar

[57] A. Stanislaus, B. H. Cooper. Catal. Rev. 36, 75 (1994).10.1080/01614949408013921Search in Google Scholar

[58] R. P. Dutta, H. H. Schobert. Catal. Today 31, 65 (1996).10.1016/0920-5861(96)00084-3Search in Google Scholar

[59] M. Li, D. Wang, J. Li, Z. Pan, H. Ma, Y. Jiang, Z. Tian, A. Lu. Chinese J. Catal. 38, 597 (2017).10.1016/S1872-2067(17)62779-7Search in Google Scholar

[60] M. Li, D. Wang, J. Li, Z. Pan, H. Ma, Y. Jiang, Z. Tian. RSC Adv. 6, 71534 (2016).10.1039/C6RA16084KSearch in Google Scholar

[61] A. V. Vutolkina, A. P. Glotov, A. V. Zanina, D. F. Makhmutov, A. L. Maximov, S. V. Egazar’yants, E. A. Karakhanov. Catal. Today 329, 156 (2019).10.1016/j.cattod.2018.11.030Search in Google Scholar

Published Online: 2020-06-18
Published in Print: 2020-06-25

© 2020 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2019-1115/html
Scroll to top button