On the role of niobium in nanostructured Mo_Nb-MCM-41 and NiMo_Nb-MCM-41 catalysts for hydrodesulfurization of dibenzothiophene

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Fuel 280 (2020) 118550

Contents lists available at ScienceDirect

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

Full Length Article

On the role of niobium in nanostructured Mo/Nb-MCM-41 and NiMo/Nb- T


MCM-41 catalysts for hydrodesulfurization of dibenzothiophene
Franklin J. Méndeza,1, Oscar E. Franco-Lópeza, Gabriela Díazb, Antonio Gómez-Cortésb,
Xim Bokhimic, Tatiana E. Klimovaa,

a
Facultad de Química, Universidad Nacional Autónoma de México (UNAM), Cd. Universitaria, Coyoacán, Ciudad de México C.P. 04510, Mexico
b
Instituto de Física, Universidad Nacional Autónoma de México (UNAM), Cd. Universitaria, Coyoacán, Ciudad de México C.P. 04510, Mexico
c
Instituto de Física, Universidad Nacional Autónoma de México (UNAM), Apartado Postal 20-364, Ciudad de México C.P. 01000, Mexico

GRAPHICAL ABSTRACT

ARTICLE INFO ABSTRACT

Keywords: In the present work, Mo and NiMo catalysts were supported on the MCM-41 material modified with Nb2O5 and
MCM-41 compared to similar catalysts supported on unmodified MCM-41 silica. Catalysts were characterized by different
Mo catalysts techniques and tested in the hydrodesulfurization of dibenzothiophene. Our interest was in clarifying the role of
NiMo catalysts niobium in the above catalysts. It was observed that the Mo/Nb-MCM-41 and NiMo/Nb-MCM-41 catalysts were
Niobium oxide
more active than their analogs supported on the MCM-41 material without niobia, and also they had higher
Hydrodesulfurization
Dibenzothiophene
hydrogenation ability. Based on the results of the characterization of calcined (oxide) and sulfided catalysts, it is
proposed that niobium acts in the above catalysts as a support improving the dispersion of the MoS2 phase and
increasing the amount of active sites and as a promoter, increasing the acidity of the Mo and NiMo catalysts.
Although the promotion effect of niobium was not as strong as that of the nickel promoter. Also, the possibility of
decoration of the edge surface of the Ni-Mo-S active sites with some Nb atoms cannot be excluded.


Corresponding author.
E-mail address: klimova@unam.mx (T.E. Klimova).
1
Present address: Departamento de Estado Sólido, Instituto de Física, Universidad Nacional Autónoma de México, Ciudad de México 04510, Mexico.

https://doi.org/10.1016/j.fuel.2020.118550
Received 25 March 2020; Received in revised form 4 June 2020; Accepted 25 June 2020
Available online 10 July 2020
0016-2361/ © 2020 Elsevier Ltd. All rights reserved.
F.J. Méndez, et al. Fuel 280 (2020) 118550

1. Introduction were reported by Rocha et al. [47], who attributed the absence of sy-
nergy between nickel and molybdenum on niobia to the strong inter-
Actual international environmental legislations are quite stringent action of each metal with niobia at the expense of interaction with each
regarding quality specifications of fossil fuels, in particular their sulfur other. Recently, Kaluža and Zdražil [48] studied the effect of using
content. The purpose of removing sulfur for production of ultra-low- nitrilotriacetic acid (NTA) in the preparation of CoMo/Nb2O5 hydro-
sulfur fuels is to reduce the emissions of sulfur oxides (SOx, x = 2 or 3) desulfurization catalysts. The NTA-CoMo/Nb2O5 catalyst resulted to be
resulting from the use of fossil fuels in automotive vehicles, aircraft, 6.6 times more active in thiophene HDS than the corresponding CoMo/
railroad locomotives and others. The reaction between SOx and hu- Nb2O5 sample prepared without NTA, as well as more active than
midity produces acid rain, in addition to the fact that sulfur, even in low commercial CoMo/Al2O3.
concentrations, poisons the catalysts based on noble metals involved in The use of the niobium as a dopant of hydrotreating catalysts was
different processes in petroleum refining [3]. This explains the high also studied [49–51]. Gaborit et al. [49] reported that the doping of a
demand for improved hydrodesulfurization (HDS) catalysts and pre- conventional NiMo hydrotreating catalyst increased the catalytic ac-
sents new challenges for catalyst design and technology [1,2]. To tivity in both HDS and HYD model reactions. The highest activities were
achieve this goal, many approaches have been tried, such as changing obtained with an optimum Nb content of 5 wt%. Cedeño-Caero et al.
the active phase and promoter, varying the preparation method and [50] tested niobium sulfide as a dopant for Mo/TiO2 catalysts. They
modifying the support. found that the formation of Nb–Ti mixed oxides led to catalysts with
Industrial HDS catalysts are composed by a MoS2 (or WS2) phase poor activity in HDS of thiophene, while the deposition of Nb2O5 on the
promoted by Ni (or Co) and supported on alumina. The active phase of surface of the TiO2 support resulted in catalysts with larger HDS ac-
these catalysts is the so-called Co(Ni)-Mo(W)-S phase [3]. The active tivities than those of Mo/Al2O3 and Mo/TiO2 references. This increase
sites of this phase are coordinatively-unsaturated sites (CUS) located at in activity was attributed to the formation of a larger population of
the edges of the Mo(W)S2 slabs [4]. anionic vacancies on the MoS2 active phase supported on TiO2 with
The wide use of alumina as a support for heterogeneous catalysts surface-deposited niobium oxide. Bouadjadja-Rohan et al. [51] pre-
can be attributed to its attractive mechanical properties, a relatively pared NiW catalysts supported on alumina doped with Nb. These cat-
low cost and the ability to provide high dispersion to the active phases alysts were evaluated in toluene hydrogenation and cyclohexane iso-
[3]. Although alumina is the most common support used for HDS cat- merization and compared to Nb-free analogs. The most efficient catalyst
alysts [5–10], other supports such as silica [8,11], carbon [6,8,12–15], of the series, with a W content of 15 wt% WO3 and a Ni/W ratio of 0.9,
titania [10,16,17], zirconia [17–19], titania nanotubes [20–23], or- kept the hydrogenation properties of a Nb-free NiW catalyst, but its
dered mesoporous molecular sieves (HMS [24–26], SBA-15 [17,27–29] isomerization activity was greatly enhanced due to the presence of Nb
and MCM-41 [30–34]) were successfully employed. On the other hand, and its acidic character at the oxide state. This demonstrates the po-
new components (active phase, promoter or additive) different from tential of using of Nb-containing catalytic formulations in hydrotreat-
those of the conventional HDS catalysts have attracted much attention ment. In addition, the same authors (Bouadjadja-Rohan et al. [52])
in the past few years. Among them, the interest in niobium-containing proposed an original approach for introducing Nb in Ni-promoted
materials has systematically increased since the beginning of the 21st alumina-supported HDS catalysts consisting in the use of Lindqvist
century [35]. The catalytic performance of niobia as a support or cat- isopolyanions [NbxW6-xO19](2+x)- (x = 0–4 and 6) precursors. The use
alyst has been studied in some heterogeneous reactions [36–38], in- of NbW isopolyanions instead of the conventional ammonium meta-
cluding its application in hydrotreatment [39–52]. tungstate resulted in a better dispersion of the metallic species in cal-
Niobium sulfide (NbS2) was tested as an active phase. Allali et al. cined catalysts and up to 5 times superior catalytic activity in the cy-
[39,40] and Geantet et al. [41] reported that the activity of carbon- or clohexane isomerization into methylcyclopentane than that of the
alumina-supported and unsupported niobium sulfide catalysts was conventional NiW catalyst.
higher than that of molybdenum sulfide analogs [39–41]. The best From the above review of the available literature information, it can
catalytic performance was obtained with catalysts prepared using nio- be concluded that niobium oxide or sulfide seem to be promising for
bium oxalate as a precursor [39]. Recently, Mansouri and Semagina application in HDS catalysts. In the majority of cases, NbS2 was used as
[42] synthesized different NbS2 nanostructures and tested them in di- an active phase or Nb2O5 was used as an individual support or as a
benzothiophene hydrodesulfurization. NbS2 nanohexagons with the modifier of the conventional alumina support. Up to now, the use of
highest fraction of corner and edge active sites showed the best activity niobium in the HDS catalysts supported on novel silica-based mesos-
in HDS. The same authors also reported that the sulfidation of niobium tructured materials is limited to just a few works. Thus, Palcheva et al.
(V) oxide can be improved by the addition of copper [43]. [53] studied NiMo catalysts supported on Nb-modified mesoporous
Niobium oxide was also used as a support. Weissman [44] studied SBA-15 and HMS materials prepared with the addition of thioglycolic
the Mo-Ni catalysts supported on two types of niobia-alumina materials acid. In our group [54,55], we observed that the addition of small
(surface niobium-aluminum oxides or mixed niobium-aluminum amounts of Nb (3–5 wt%) to the MCM-41 support increased the cata-
oxides). He demonstrated that in both cases, the catalysts supported on lytic activity of NiMo and CoMo catalysts in HDS of dibenzotiophene.
Nb2O5-Al2O3 compositions having maximum surface acidity were the The aim of the present work is to achieve a deeper insight into the role
most active for sulfur and nitrogen removal from two different gas-oil of niobium in the NiMo/MCM-41 catalysts and its interaction with Ni,
feeds. Puello-Polo et al. [45] also studied Al2O3-Nb2O5 supports with Nb Mo and NiMo species in their oxide and sulfided forms.
loading from 0 to 8 wt%. They found that the presence of Nb decreased
the active metal-support interaction and improves the degree of sulfi- 2. Experimental
dation of Ni and Mo oxide species, leading to an increase in their ac-
tivity in the high-pressure HDS of dibenzothiophene. Faro Jr. and dos 2.1. Supports preparation
Santos [46] compared the catalytic activity of sulfided Ni, Mo and NiMo
catalysts supported on alumina and niobia in cumene hydrocracking MCM-41 silica with p6mm hexagonal structure was synthesized
and thiophene HDS. They reported that the niobia-supported nickel following the procedure described in [34]. Tetraethylammonium hy-
catalysts were more active in thiophene HDS than the alumina-sup- droxide (TEAOH) and cetyltrimethylammonium bromide (CTMAB)
ported ones, while the niobia-supported molybdenum catalysts were were used as the structure directing agents and sodium silicate solution
less active than the alumina-supported analogs. With the bimetallic (Na2SiO3) as the silica source. The sodium silicate solution was pre-
NiMo catalysts, little or no synergy was observed with the niobia-sup- pared by dissolving fumed silicon oxide in aqueous sodium hydroxide
ported catalysts, in contrast with the alumina support. Similar results (SiO2:NaOH molar ratio = 1:2). The chemicals were mixed in order to

2
F.J. Méndez, et al. Fuel 280 (2020) 118550

obtain a gel with molar composition: characterization, the samples (50 mg) were pretreated in situ at 400 °C
1Na2SiO3:0.2TEAOH:0.25CTMAB:40H2O. The pH of the gel was ad- for 2 h in an air flow and then cooled to room temperature in an Ar
justed to a value between 9 and 10 by dropwise addition of dilute stream. Then, the reduction of the samples was performed from room
sulfuric acid. After aging of the gel at room temperature for 24 h, the temperature to 1000 °C with a heating rate of 10 °C/min under a stream
obtained white product was separated by vacuum filtration, washed of a H2/Ar mixture (10:90 mol/mol, 50 mL/min flow). The consump-
several times with water and ethanol, dried at room temperature (12 h) tion of H2 at different temperatures was recorded. For the ammonia
and at 100 °C (12 h) and calcined at 550 °C for 8 h. TPD experiments, first the samples (50 mg) were pretreated in situ at
The niobium-containing support (Nb-MCM-41) was prepared by the 500 °C for 30 min in a helium flow to remove water and other con-
impregnation of calcined MCM-41 material with an excess of aqueous taminants. Then, the samples were cooled to 120 °C and contacted with
solution of ammonium niobate(V) oxalate hydrate. After impregnation, a NH3/He mixture (10:90 mol/mol, 20 mL/min flow) for 30 min and
the sample was dried at 100 °C for 6 h and calcined at 500 °C for 4 h in a then purged with pure He (20 mL/min flow) for 30 min. The desorption
static air atmosphere. The nominal composition of the Nb-MCM-41 step was performed in a He stream (50 mL/min) from 120 °C to 500 °C
support was 3 wt% Nb (4.3 wt% of Nb2O5). with a heating rate of 10 °C/min, keeping the final temperature until
the trace reached the baseline.
2.2. Catalysts preparation
2.4. Characterization of sulfided catalysts
The catalysts were prepared by successive impregnation with an
excess of aqueous solution of ammonium heptamolybdate tetrahydrate Sulfided catalysts were characterized by FT-IR of adsorbed NO and
and nickel(II) sulfate hexahydrate used as precursors of Mo and Ni high-resolution transmission electron microscopy (HRTEM). The NO
species, respectively. In all cases, molybdenum was deposited first on interactions with the surface of the sulfided catalysts were studied using
the MCM-41 support, followed by the deposition of nickel. After the a Thermo Scientific Nicolet iS50 spectrophotometer equipped with a
impregnation of each precursor, the samples were dried at 100 °C for Praying MantisTM DRIFT cell and reaction chamber CHC with KBr
6 h and calcined at 500 °C for 4 h in a static air atmosphere. In this windows of Harrick Scientific Products. The samples were put directly
stage, the temperature was increased at a linear rate of 1 °C/min. The in the reaction chamber and sulfided in situ using a H2S/H2 mixture
nominal composition of the catalysts was 2.4 wt% of Ni (3 wt% NiO) (15 vol% of H2S) upon the sulfidation conditions described in Section
and 10 wt% of Mo (15% wt. % MoO3). For comparison purposes, Ni 2.5. After the sulfidation was finished, the samples were purged with He
catalysts supported on MCM-41 and Nb-MCM-41 were prepared with (30 mL/min) for 30 min at 400 °C, cooled to 20 °C in the same gas
the same Ni loading. atmosphere and saturated with a NO/He mixture (2 vol% of NO,
30 mL/min flow). All spectra were collected using 256 scans with a
2.3. Characterization of supports and oxide catalysts resolution of 4 cm−1 using a MCT detector.
The high-resolution transmission electron microscopy (HRTEM)
The (Nb)-MCM-41 supports and calcined catalysts were character- characterization of the sulfided samples was performed using a JEOL
ized by scanning electron microscopy coupled to energy-dispersive X- 2010 microscope (resolving power 1.9 Å at 200 kV). The solids were
ray analyzer (SEM-EDS), X-ray diffraction (XRD), nitrogen physisorp- ultrasonically dispersed in heptane and the suspension was collected on
tion, UV–vis diffuse reflectance spectroscopy (UV–vis DRS), tempera- carbon coated grids. Average morphology of the MoS2 active phase in
ture-programmed reduction (TPR) and temperature-programmed des- the catalysts (slab length and layer stacking distributions) was de-
orption of ammonia (TPD-NH3). termined by measuring the characteristics of more than 300 crystallites
The chemical composition of the catalysts was measured by SEM- of MoS2 observed in several HRTEM images of sulfided catalysts taken
EDS using a Jeol 5900 LV microscope with Oxford ISIS microanalyzer. at different sites of the grids.
Small- and wide-angle X-ray diffraction (SA- and WA-XRD) patterns
were collected using a Bruker D8 Advance diffractometer at room 2.5. Catalytic activity tests
temperature and CuKα radiation (λ = 1.5406 Å). The SA-XRD char-
acterization was performed with a goniometer speed of 0.1°(2θ) per Catalytic activity of the synthesized catalysts was tested in hydro-
minute in the 1° ≤ 2θ ≤ 10° interval. The WA-diffraction patterns were desulfurization of dibenzothiophene (DBT). The reactions were per-
recorded with a goniometer speed of 1°(2θ) per minute in the 3° ≤ formed in a batch reactor under the following conditions:
2θ ≤ 80° interval. temperature = 300 °C, total pressure = 7.3 MPa and reaction
The textural properties of the (Nb)-MCM-41 supports and the cor- time = 8 h. Prior to the activity tests, catalysts were sulfided ex situ at
responding Mo and NiMo catalysts were determined with a 400 °C for 4 h in a stream of H2S/H2 (15 vol% of H2S) at atmospheric
Micromeritics ASAP 2020 automatic analyzer at liquid N2 temperature pressure. The sulfided catalysts were transferred in an inert atmosphere
(-197.5 °C). Previous to the measurements, the samples were degassed (Ar) to a batch reactor containing 40 mL of DBT solution in hexadecane
at 270 °C for 6 h (P less than 10-1 Pa). N2 adsorption–desorption iso- with initial sulfur concentration of 1300 ppm. The catalytic tests were
therms were obtained and specific surface areas were calculated by the carried out in the absence of external and internal mass transfer/dif-
BET method (SBET). The experimental error of the SBET determination fusion limitations. This was guaranteed by the high stirring rate
was ± 2%. The total pore volume (Vp) was determined by nitrogen (600 rpm) and the small catalyst’s particle size (less than 150 µm). The
adsorption at a relative pressure of 0.98. Pore size distributions were composition of the reaction mixture was followed by withdrawing ali-
obtained from the adsorption isotherms by the Barrett-Joyner-Halenda quots (0.5 mL) each hour and analyzing them in a gas chromatograph
(BJH) method. The mesopore diameters reported in the present work (model 6890, Agilent Technologies) equipped with a flame ionization
(Dp) correspond to the maxima of the pore size distributions. detector and a non-polar methyl siloxane HP-1 capillary column
UV–vis diffuse reflectance spectra (UV–vis DRS) were acquired on a (50 m × 0.32 mm inner diameter and 0.52 μm film thickness). The
Varian Cary 100 Conc spectrophotometer equipped with an integrating column was operated under the following temperature program: 20 min
diffuse reflectance sphere. Polytetrafluoroethylene was used as a re- at 90 °C, constant heating (20 °C/min) up to 200 °C, and finally
ference. 14.5 min isothermally at 200 °C. To corroborate product identification,
The temperature programmed reduction (TPR) and temperature chromatographic standards and an Agilent 7890A GC system equipped
programmed desorption of ammonia (TPD-NH3) experiments were with 5975C MS detector were used. The amount of gaseous products
carried out in a Micromeritics AutoChem II 2920 automatic analyzer was negligibly small under the present conditions. The identified re-
equipped with a thermal conductivity detector. Prior to the TPR action products were tetrahydrodibenzothiophene (THDBT),

3
F.J. Méndez, et al. Fuel 280 (2020) 118550

Table 1 reflection was preserved in the Nb-modified MCM-41 support and all
Chemical composition of the prepared catalysts determined by SEM-EDX.a prepared catalysts, indicating that the interplanar distance
Samples Chemical composition (wt. %)
d100 = 40.2 ± 0.3 Å was maintained in all samples (Table 2), as well
as the lattice parameter a0 = 46.5 ± 0.5 Å and the pore wall thickness
NiO MoO3 Nb2O5 δ = 21 ± 1 Å. On the other hand, the intensity of the main (1 0 0)
signal decreased with the deposition of metal oxides (Fig. 2A, curves
Nb-MCM-41 – – 4.4 ± 0.3
from b to f). This decrease can be attributed, on the one side, to the
Mo/MCM-41 – 15.3 ± 1.0 –
Mo/Nb-MCM-41 – 15.0 ± 1.2 4.6 ± 0.4 modification of the MCM-41 pore wall by the deposited species that can
NiMo/MCM-41 2.8 ± 0.3 16.1 ± 1.4 – produce a reduction of the scattering contrast between the pores and
NiMo/Nb-MCM-41 3.1 ± 0.4 14.4 ± 1.5 4.3 ± 0.3 the walls of the molecular sieve decreasing the relative intensities of the
a
diffraction signals [57]. On the other side, this decrease can be due to
The nominal composition of the catalysts was 4.3 wt% of Nb2O5, 3.0 wt%
some loss in the long-range pore order of the support after the de-
of NiO and 15 wt% of MoO3.
position of Nb, Ni and Mo oxide species, which can partially or com-
pletely block some pores or pore entrances of the MCM-41 support.
cyclohexylbenzene (CHB), dicyclohexyl (DCH) and biphenyl (BP).
Some partial destruction of the MCM-41 support’s pore structure by the
ChemStation Plus software was used for the integration of detected GC
deposited metal oxides also cannot be excluded due to a small wall
signals. The quantitative analysis of the reactant and products was
thickness (~21 Å).
performed using the corresponding calibration curves and hexadecane
Wide-angle diffractograms of the supports and catalysts (Fig. 2B)
as an internal standard in all cases.
showed a broad band between 15° and 40° (2θ), which is characteristic
of the amorphous silica material of the MCM-41 support. The Nb-MCM-
3. Results and discussion 41 sample (Fig. 1B, (b)) showed the diffractogram without additional
signals of niobium oxide crystalline phase pointing out to its good
3.1. Supports and oxide catalysts characterization dispersion on the MCM-41 surface. In contrast, for the Mo/MCM-41
catalyst (Fig. 2B, (c)), new diffraction signals were observed: three in-
3.1.1. Chemical composition tense signals at 23.4°, 25.7° and 27.3° (2θ) and other lower intensity
The chemical composition of the Nb-MCM-41 support and the pre- ones at 33.6°, 35.4°, 38.8°, 45.9° and 49.2° (2θ). All these signals cor-
pared Mo and NiMo catalysts was determined by a SEM-EDS analysis. respond to the orthorhombic MoO3 crystalline phase (PDF #76–1003)
The results (Table 1) point out that all the samples presented metal [58], suggesting poor dispersion of the Mo6+ oxide species on the
oxide loadings similar to the theoretically expected ones. The elemental MCM-41 silica surface. A rough estimation of the MoO3 crystallite size
mapping results of the prepared Mo and NiMo catalysts supported on with the Scherrer equation and the most intense (0 2 1) reflection gave
MCM-41 and Nb-MCM-41 materials are shown in Fig. 1. It can be seen us a value of about 220 Å. This result indicates that the MoO3 crys-
that Ni and Mo species exist in dispersed form and they are more tallites detected by XRD are located on the external surface of the MCM-
homogeneously distributed in the catalysts supported on the Nb-con- 41 support. The intensities of the MoO3 peaks decreased in the XRD
taining material. pattern of the corresponding Mo catalyst supported on Nb-containing
material (Mo/Nb-MCM-41, Fig. 2B, (d)). In line with this, MoO3 crys-
3.1.2. Small- and wide-angle X-ray diffraction tallite size decreased to 180–200 Å. Regarding the NiMo/MCM-41
Fig. 2A and 2B show the small- and wide-angle diffraction patterns catalyst, only one weak signal was observed at 26.7° (2θ) (Fig. 1B, (e))
of the prepared materials. The small angle XRD profile of the parent attributed to the (2 2 0) plane of the NiMoO4 crystalline phase (PDF
MCM-41 (Fig. 2A, curve a) shows the peaks characteristic of its ordered #45–0142) [58]. The poor resolution of this signal allowed us to esti-
regular pore structure: the main peak at 2.2° (2θ) and low-intensity mate only the approximate size of nickel molybdate crystallites being
signals at 3.8° and 4.6°(2θ). These signals are attributed to the (1 0 0), between 100 and 150 Å. Finally, the NiMo/Nb-MCM-41 catalyst
(1 1 0) and (2 0 0) planes of the structure with long-range two-dimen- (Fig. 2B, curve f) did not show the presence of any crystalline metal
sional hexagonal honeycomb arrangement of cylindrical pores and oxide phase detectable by powder XRD technique. These results reveal
p6mm space group symmetry [56]. The position of the main (1 0 0) that niobium oxide deposited on the surface of the MCM-41 support

Fig. 1. SEM-EDS mapping of the prepared Mo and NiMo catalysts.

4
F.J. Méndez, et al. Fuel 280 (2020) 118550

Fig. 2. (A) Small- and (B) wide-angle X-ray diffraction patterns of (a) MCM-41, (b) Nb-MCM-41, (c) Mo/MCM-41, (d) Mo/Nb-MCM-41, (e) NiMo/MCM-41 and (f)
NiMo/Nb-MCM-41 samples. The curves were shifted in a vertical direction to be seen better.

Table 2 catalysts, the Mo/MCM-41 one showed the lowest surface area and pore
Textural and structural properties of the supports and catalysts. volume values that, in line with the XRD results, can be due to the
Samples SBET (m2/ Vp (cm3/ Dp (Å) c
d100 (Å) a0 (Å) e
δ (Å) f agglomeration of Mo6+ oxide species and the formation of the MoO3
g) a g) b d crystalline phase.

MCM-41 1027 1.06 25 40.7 47.0 22


3.1.4. UV–vis diffuse reflectance spectroscopy
Nb-MCM-41 810 0.78 25 40.5 46.8 21
Mo/MCM-41 495 0.51 26 39.9 46.1 20 The coordination state and the degree of agglomeration of the
Mo/Nb-MCM-41 637 0.60 26 39.9 46.1 20 transition metal oxides in the prepared catalysts were characterized by
NiMo/MCM-41 713 0.76 25 40.3 46.5 22 UV–vis diffuse reflectance spectroscopy (Fig. 4). As reported previously
NiMo/Nb-MCM- 655 0.63 25 40.1 46.3 21 [62], the MCM-41 support does not exhibit any absorption of UV or
41
visible light (Fig. 4, curve a). On the contrary, the Nb-MCM-41 support
a
Specific surface area determined by the Brunauer-Emmett-Teller method. showed an absorption band between 200 and 300 nm with the max-
b
Pore volume obtained at a relative pressure of 0.98. imum at 242 nm (Fig. 4, spectrum b). This signal is due to the ligand-to-
c
Pore diameter determined from the adsorption branch of the isotherms by metal charge transfer (LMCT) between the O2– anion and Nb5+ cations
the Barrett-Joyner-Halenda method. of the Nb2O5 deposited on the silica surface. According to literature,
d
Interplanar spacing values (d100). this absorption band can be ascribed to isolated pentahedrally-co-
e
Lattice parameter (a0) for a hexagonal pore structure: a0 = 2d100/ 3 . ordinated Nb5+ or to highly dispersed octahedral Nb5+ oxide species
f
Pore wall thickness (δ) estimated as δ = a0 – DP [35,54]. The DRS results of the Ni/MCM-41 and Ni/Nb-MCM-41 sam-
ples (Fig. 4, spectra (c) and (d)) exhibited similar absorption bands
improved the dispersion of nickel and molybdenum oxide species in between 200 and 300 nm characteristic of the LMCT O2– → Ni2+ of
calcined catalysts. octahedrally coordinated Ni2+ oxide species, Ni(Oh) [63] and two less
intense absorption signals at 405 and 740 nm attributed to d-d transi-
3.1.3. Nitrogen physisorption tions 3A2g → 3T1g(P) and 3A2g → 3T1g(F) [64], respectively, of the Ni
Nitrogen adsorption–desorption isotherms and pore size distribu- (Oh) species. No signals of tetrahedrally-coordinated Ni2+ oxide species
tions of the samples are shown in Fig. 3. All samples presented Type-IV were observed in the spectra. Regarding the Mo catalysts, the DRS
isotherms (Fig. 3A), typical of mesoporous materials [34,59–61]. The spectra of the Mo/MCM-41 and Mo/Nb-MCM-41 catalysts (Fig. 4, (e)
uniformity of the pore size of the (Nb)-MCM-41 supports and prepared and (f)) showed two overlapped absorption bands. The first signal lo-
catalysts can be observed in the pore size distributions, where only one cated at about 260–270 nm is characteristic of tetrahedrally-co-
well-defined peak appears between 20 and 30 Å with the maximum at ordinated Mo6+ oxide species, Mo(Td), or of well-dispersed Mo6+
25 Å (Fig. 3B and Table 2). The amount of the adsorbed N2 decreased oxide species in octahedral coordination, Mo(Oh). The second signal
with the loading of the metal oxide species on the MCM-41 surface with the maximum at 295–310 nm is generally ascribed to agglomer-
(Fig. 3A). As a result, catalysts showed lower values of specific surface ated polymeric octahedral Mo6+ oxide species. Both of the above ab-
area and pore volume than the starting MCM-41 support (Table 2) that sorption bands are associated with the LMCT O2– → Mo6+ [65–67]. The
can be ascribed to an increase in the materials' density and/or pore diffuse reflectance spectra of the NiMo/MCM-41 and NiMo/Nb-MCM-
plugging caused by the deposited metal oxide species. Among all the 41 catalysts (Fig. 4, (g) and (h)) also showed a similar shape with two

5
F.J. Méndez, et al. Fuel 280 (2020) 118550

Fig. 3. (A) Nitrogen adsorption–desorption isotherms and (B) pore size distributions of (a) MCM-41, (b) Nb-MCM-41, (c) Mo/MCM-41, (d) Mo/Nb-MCM-41, (e)
NiMo/MCM-41 and (f) NiMo/Nb-MCM-41 samples. The curves were shifted in a vertical direction to be seen better.

Ni2+ oxide. Resuming the above DRS results, we could not observe a
clear effect of the incorporation of niobium on the type of Ni and Mo
oxide species in the catalysts.

3.1.5. Temperature-programmed reduction


Fig. 5 presents the temperature-programmed reduction (TPR) pro-
files of the prepared catalysts. The starting MCM-41 support and the
Nb-MCM-41 sample (Fig. 5, (a) and (b)) did not show any reduction
peak. Therefore, no reduction of the niobium oxide (Nb2O5) was

Fig. 4. UV–vis DRS spectra of (a) MCM-41, (b) Nb-MCM-41, (c) Ni/MCM-41,
(d) Ni/Nb-MCM-41, (e) Mo/MCM-41, (f) Mo/Nb-MCM-41, (g) NiMo/MCM-41
and (h) NiMo/Nb-MCM-41 samples.

well-defined absorption bands with the maxima at 245–250 nm and


330 nm and a small shoulder at 405 nm. The first band centered at
245–250 nm seems to be the result of the overlapping of the LMCT
signal of octahedrally-coordinated Ni oxide species (maximum at
230 nm) with the LMCT absorption band of the tetrahedral or well-
dispersed octahedral Mo6+ oxide species (signal with the maximum at
260–270 nm). The second signal can be clearly assigned to the ag- Fig. 5. Temperature-programmed reduction profiles of a) MCM-41, (b) Nb-
glomerated octahedrally-coordinated Mo oxide species, while the MCM-41, (c) Ni/MCM-41, (d) Ni/Nb-MCM-41, (e) Mo/MCM-41, (f) Mo/Nb-
shoulder at 405 nm is due to d-d transitions of octahedrally-coordinated MCM-41, (g) NiMo/MCM-41, and (h) NiMo/Nb-MCM-41 samples.

6
F.J. Méndez, et al. Fuel 280 (2020) 118550

detected. On the contrary, the reduction profiles of the Ni catalysts


(Fig. 5, (c) and (d)) revealed just one well-defined symmetrical reduc-
tion peak with the maxima at about 400 °C. This signal is assigned to
the one-step reduction of Ni oxide species from Ni2+ to Ni0 weakly
bound to silica sites [68–70]. The presence of niobium in the catalyst
increased the temperature of reduction of the NiO species from 380 °C
(Ni/MCM-41) to 405 °C (Ni/Nb-MCM-41), as well as the amount of
hydrogen consumed during this reduction. The reduction profiles of the
Mo/MCM-41 and Mo/Nb-MCM-41 catalysts (Fig. 5, (e) and (f)) ex-
hibited two reduction peaks. The first intense reduction signal between
400 and 600 °C is associated to the first reduction step of polymeric
octahedral Mo oxide species from Mo6+ to Mo4+. In the presence of
niobium in the catalyst, the maximum of this signal appears at a higher
temperature (552 °C) than in the Mo/MCM-41 catalyst (538 °C) that can
be due to the stronger interaction of the Mo oxide species with the
niobium-containing support. The second peak of H2 consumption ob-
served at 675 °C for the Mo/MCM-41 sample can be associated with the
first step of reduction of crystalline MoO3 phase detected in this sample
by powder XRD (Fig. 2B) or to the second step of reduction from Mo4+
to Mo0 of polymeric octahedral Mo oxide species [71,72]. In the Mo/
Nb-MCM-41 catalyst, the second reduction signal became less defined,
probably because of a better dispersion of the MoO3 phase as detected
by XRD characterization (Fig. 2B). Unlike the TPR results of the un-
promoted Mo catalysts, the reduction profiles of the NiMo/MCM-41 and
NiMo/Nb-MCM-41 catalysts (Fig. 5, (g) and (h), respectively) presented
only one well-defined intense reduction signal between 300 and 500 °C
and low-intensity hydrogen consumption at 500–1000 °C region. The Fig. 6. Profiles of the temperature-programmed desorption of ammonia (TPD-
low-temperature intense signal can be assigned to the first step of re- NH3) of (a) MCM-41, (b) Nb-MCM-41, (c) Mo/MCM-41, (d) Mo/Nb-MCM-41,
duction of the dispersed octahedral Mo oxide species. When reduction (e) NiMo/MCM-41 and (f) NiMo/Nb-MCM-41 samples. WAS, weak acid sites
(120–200 °C), MAS, medium acid sites (200–400 °C) and SAS, strong acid sites
patterns of the Ni-promoted Mo catalysts are compared to those of the
(400–500 °C).
unpromoted Mo samples, it can be noted that the addition of nickel led
to a significant decrease in the reduction temperature of octahedral Mo
oxide species (from 538 to 552 °C to 416–436 °C). This result is in line Table 4
with the increase in the dispersion of Mo oxide species detected by XRD Surface acidity of supports and catalysts determined by TPD-NH3.
after the incorporation of Ni (Fig. 2B). It is interesting to note that in Samples Amount of acid sites (μmol NH3/gcat) a

this case, Mo6+ oxide species of the Nb-containing NiMo catalyst are
reduced at a lower temperature than those of the NiMo/MCM-41 one. Weak (WAS) Medium (MAS) Strong (SAS) Total
This is an unexpected result since, in the case of monometallic Ni and
MCM-41 10 64 35 109
Mo catalysts, the addition of niobia to the MCM-41 support resulted in Nb-MCM-41 74 165 63 302
an increase in the reduction temperature of Ni and Mo oxides. Probably, Mo/MCM-41 306 336 33 675
the reduction behavior of the NiMo catalysts is different due to the Mo/Nb-MCM-41 386 451 38 875
NiMo/MCM-41 332 389 50 771
formation of a Ni-promoted Mo oxide phase, since the reduction signals
NiMo/Nb-MCM-41 432 530 67 1029
of the unpromoted Mo oxide species were not observed. The dispersion
of this Ni-Mo oxide phase is enhanced on the Nb-containing support as a
WAS = Weak Acid Sites (120–200 °C), MAS = Medium Acid Sites
observed by WA-XRD (Fig. 2B), and the reduction of Mo is promoted by (200–400 °C) and SAS = Strong Acid Sites (400–500 °C).
the presence of Ni, whose reduction in turn is improved by niobium.
Table 3 shows results from the quantification of the H2
consumptions of the catalysts during their reduction. Hydrogen con-
Table 3 sumptions were determined at two temperature intervals: a low tem-
Hydrogen consumption and degree of the reduction of metal oxide species in perature region (LT, 200–600 °C) and a high temperature region (HT,
calcined catalysts determined from TPR results. 600–1000 °C), in order to evaluate the amount of metal oxide species
a b reduced in each temperature interval. In the case of Ni/MCM-41 and
Samples H2 consumption (mmol/gcat) αR
Ni/Nb-MCM-41 catalysts, the latter showed higher H2 consumption at
LT HT Total the LT region (0.37 mmol/gcat) and higher degree of reduction
(αR = 0.93) than the Ni/MCM-41 sample (0.29 mmol/gcat and
Ni/MCM-41 0.29 – 0.29 0.73
αR = 0.73). Therefore, the addition of niobium to the MCM-41 support
Ni/Nb-MCM-41 0.37 – 0.37 0.93
Mo/MCM-41 1.16 1.46 2.62 0.85 improved the proportion of reduced NiO species almost reaching their
Mo/Nb-MCM-41 0.82 0.88 1.70 0.55 complete reduction. In contrast to the above results, the Mo/Nb-MCM-
NiMo/MCM-41 2.09 0.60 2.68 0.76 41 catalyst showed a significantly smaller H2 consumption than the
NiMo/Nb-MCM-41 2.35 0.40 2.75 0.78 Mo/MCM-41 one (Table 3). This behavior suggests the possibility of a
a relatively strong interaction between MoO3 and Nb2O5 species that
Hydrogen consumption at low temperature (LT, 200–600 °C) and at high
temperature (HT, 600–1000 °C) intervals determined from TPR profiles. inhibits the reduction of the Mo6+ oxide species. This MoO3-Nb2O5
b
Degree of the reduction of oxide species determined from the measured interaction can also be responsible for a better dispersion of the MoO3
total H2 consumption of each sample and the theoretical values corresponding species in the Mo/Nb-MCM-41 catalyst than in the Mo/MCM-41 one, as
to its complete reduction (0.4, 3.1 and 3.5 mmol/gcat for Ni, Mo and NiMo observed by WA-XRD (Fig. 2B). Finally, the addition of a Ni promoter to
catalysts, respectively). the Mo/MCM-41 and Mo/Nb-MCM-41 catalysts resulted in both cases

7
F.J. Méndez, et al. Fuel 280 (2020) 118550

Fig. 7. HRTEM micrographs of sulfided catalysts: (a) Mo/MCM-41, (b) Mo/Nb-MCM-41, (c) NiMo/MCM-41 and (d) NiMo/Nb-MCM-41. Insets show representative
MoS2 slabs.

in an increase in the H2 consumption at the LT interval and a decrease considerable amount of weak and medium acid sites appeared (Fig. 6,
at the HT interval, being the degrees of reduction similar for both NiMo (b): signals at 249 and 394 °C, respectively). This is in line with the
catalysts (αR ≈ 0.8, Table 3). Nevertheless, it can be observed that the well-known fact that niobium species are acidic in the oxide and sulfide
effect of the Ni promoter on the ease of reduction of the Mo oxide states [43,51]. All the prepared Mo and NiMo catalysts showed sig-
species was enhanced in the presence of niobium in the catalyst. Thus, nificantly larger amounts of desorbed ammonia than the (Nb)-MCM-41
the H2 consumption at a low-temperature interval increased three-fold supports that can be attributed to the presence of Ni and Mo oxide
after the addition of Ni to the Mo/Nb-MCM-41 catalyst. For compar- species. Once again, the acidity of the catalysts supported on the Nb-
ison, for the catalysts without niobium, this increase was smaller (less containing MCM-41 material was larger than that of the MCM-41-
than twice). It seems that a more complete reduction of the Ni oxide supported counterparts. Weak and medium strength acid sites were
species in the presence of niobium oxide is favorable for the reduction predominant in all the catalysts (Fig. 6, (c-e)). For the Mo/Nb-MCM-41
of Mo oxide species at low temperature. Probably, the reduced Ni catalyst, a new signal of desorbed ammonia was observed at 338 °C
species were able to activate hydrogen molecules and supply them to (Fig. 6, (d)). In addition, Nb2O5 incorporation to the MCM-41 support
facilitate the reduction of the MoO3 species. resulted in a slight shift of the maxima of the TPD-NH3 signals to higher
temperatures: from 185 °C for the Mo/MCM-41 to 197 °C for the Mo/
Nb-MCM-41 catalysts (Fig. 6, (c) and (d)) and from 187 °C for the
3.1.6. Temperature-programmed desorption of ammonia
NiMo/MCM-41 to 192 °C for the NiMo/Nb-MCM-41 one (Fig. 6, (e) and
Fig. 6 shows the TPD-NH3 profiles of the MCM-41 and Nb-MCM-41
(f)). This points out to the increase in the strength of acid sites in the
supports and corresponding Mo and NiMo catalysts. The amounts of
presence of niobia.
ammonia thermodesorbed at different temperatures are listed in
Table 4 summarizes the results of the quantification of the total
Table 4. The amounts of acid sites were quantified at three temperature
amount of acid sites, as well as the amounts of acid sites of different
intervals: 120–200 °C (weak acid sites, WAS), 200–400 °C (medium acid
strength (WAS, MAS and SAS) in the prepared catalysts. The results
sites, MAS) and 400–500 °C (strong acid sites, SAS). The total amount of
from this table confirm that the addition of niobium oxide to the MCM-
acid sites was also determined.
41 support resulted in a considerable increase in the acidity of the
The parent MCM-41 presented negligible acidity (Fig. 6, (a)).
catalysts, especially of the weak and medium strength acid sites. The
However, after the addition of niobium oxide to MCM-41, a

8
F.J. Méndez, et al. Fuel 280 (2020) 118550

[43]. Previously NbS2 was prepared at a high temperature (950 °C)


followed by quenching from an annealing temperature of 750 °C [73].
Fig. 8 shows the distributions of slab length and number of layers of
the MoS2 crystallites in the sulfided catalyst. The length and stacking
distributions were broader for the catalysts supported on Nb-free MCM-
41 material than for those supported on the Nb-MCM-41 counterparts.
In the Mo/MCM-41 and NiMo/MCM-41 catalysts, the majority of the
MoS2 crystallites showed a length between 20.1 and 80 Å and stacking
between 2 and more than 5 layers. In comparison, the Mo/Nb-MCM-41
and NiMo/Nb-MCM-41 catalysts had MoS2 crystallites of a shorter
length (20.1–60 Å) and lower stacking degree (1 to 4 layers). This result
indicates that the incorporation of Nb to the MCM-41 support improved
the dispersion of the catalytically active MoS2 phase. This conclusion
was confirmed by the calculation of the average slab length (L) and
average stacking number (N) of MoS2 crystallites in different catalysts
(Table 5). It was observed that both average characteristics (L and N) of
the MoS2 phase were smaller in the catalysts supported on the Nb-
containing MCM-41 support.
The average slab lengths and stacking numbers were used to de-
termine the average fraction of Mo atoms located on the edge surface of
MoS2 particles (fMo, Table 5), assuming that the MoS2 crystals are
perfect hexagons [74,75]. It has been reported that these values can be
Fig. 8. Slabs length (A) and stacking degree (B) distributions of MoS2 crystal-
considered an indicator of the amount of sulfided Mo species located on
lites in sulfided catalysts: (a) Mo/MCM-41, (b) Mo/Nb-MCM-41, (c) NiMo/
MCM-41 and (d) NiMo/Nb-MCM-41. the catalytically active surface of the MoS2 particles [74–77]. As ex-
pected, the catalysts supported on Nb-MCM-41 showed higher fMo value
(0.26) than the MCM-41 supported counterparts (fMo = 0.22). There-
total amount of acid sites in the Mo/Nb-MCM-41 and NiMo/Nb-MCM- fore, the presence of Nb in the MCM-41 support resulted in the im-
41 catalysts (771 μmol/gcat and 1029 μmol/gcat, respectively) was proved dispersion of the MoS2 active phase. It is worth to mention that
higher than that in the corresponding samples without niobium the effect of niobium on the characteristics of the oxide and sulfide Mo
(675 μmol/gcat in the Mo/MCM-41 and 875 μmol/gcat in the NiMo/ and NiMo phases supported on MCM-41 silica is different from that
MCM-41). So, the total acidity of the Nb-containing catalysts was about observed previously for the alumina-supported NiMo catalysts. Thus, it
10–20% higher than that of the catalysts without niobium. These results was shown for the alumina support that the presence of Nb decreases
show clearly the effect of niobium on the acidity of the prepared cat- the active metal-support interaction [45]. In line with this, in another
alysts. work [78] it was observed that the introduction of Nb to the NiMo/
Al2O3 catalysts led to the agglomeration of the NiMo phase in the
3.2. Sulfided catalysts characterization oxidation state and to an increase in the length and stacking number of
the NiMoS phase. However, for the silica support MCM-41, we observed
3.2.1. High resolution transmission electron microscopy that the addition of Nb improved the dispersion of Mo and NiMo spe-
Sulfided (Ni)Mo/(Nb)-MCM-41 catalysts were characterized by cies.
HRTEM in order to acquire more information about the effect of the
incorporation of Nb2O5 on the morphology of the catalytically active
MoS2 phase. The obtained HRTEM images are shown in Fig. 7. The 3.2.2. In situ Fourier-transformed infrared spectroscopy of chemisorbed
micrographs show some dark lines separated by interplanar distances of nitric oxide
6.2 Å (Fig. 7, the insets) characteristic of layered MoS2 crystal structure, The analysis of the FT-IR spectra of chemisorbed NO on the hy-
in which molybdenum layers are located between two sulfur layers drotreating catalysts renders information about their dispersion, oxi-
forming hexagonal shape particles. A large number of MoS2 particles dation state, and the degree of unsaturation of the active sites [79]. The
was observed on the external surface of the mesoporous catalysts. advantage of this technique is that it provides independent information
Previously, it was reported that the niobium sulfide's species, NbS2, on the NO molecules adsorbed on the adsorption sites associated to Mo
have a crystallographic structure similar to that of the MoS2, but with a and Ni in the sulfided catalysts. The spectra of NO chemisorbed on the
slightly larger interplanar distance (6.3 Å). However, the formation of sulfided (Ni)Mo/(Nb)-MCM-41 catalysts prepared in the present work
such sulfided NbS2 phase is not possible in the case of our catalysts, are shown in Fig. 9. A single peak observed at 1820 cm−1 in the
because its formation from niobium oxide is thermodynamically limited spectrum of the Ni/MCM-41 catalyst (Fig. 9 (a), solid line), can be

Table 5
Morphology of the MoS2 active phase determined by HRTEM and pseudo-first order rate constants (k) for the hydrodesulfurization of dibenzothiophene.
a
Samples Average MoS2 characteristics Pseudo-first order rate constants

L (Å) N fMo k, h−1 k’ × 105, L/(s·gcat)

Mo/MCM-41 55 3.6 0.22 0.05 ± 0.005 0.37 ± 0.04


Mo/Nb-MCM-41 46 3.1 0.26 0.10 ± 0.01 0.71 ± 0.08
Mo/γ-Al2O3 ref.b 0.03 ± 0.004 0.20 ± 0.03
NiMo/MCM-41 54 3.7 0.22 0.19 ± 0.01 1.38 ± 0.10
NiMo/Nb-MCM-41 46 2.3 0.26 0.34 ± 0.04 2.51 ± 0.29
NiMo/γ-Al2O3 ref.b 0.28 ± 0.03 2.05 ± 0.19

a
L (average length) and N (average stacking degree) of MoS2 crystallites; fMo, estimated fraction of Mo atoms on the edge surface of MoS2 particles.
b
Reference Mo/γ-Al2O3 and NiMo/γ-Al2O3 catalysts with 3.0 wt% of NiO and 15 wt% of MoO3 loadings.

9
F.J. Méndez, et al. Fuel 280 (2020) 118550

and at 1700–1715 cm−1 for the anti-symmetric one. However, for the
sulfided Mo catalysts, these bands appear shifted to lower wave-
numbers [79]. Therefore, the positions of these bands in the spectrum
of the Mo/MCM-41 catalyst point out to the adsorption of NO on the
sulfided MoS2 phase.
The IR spectrum of chemisorbed NO on the NiMo/MCM-41 catalyst
(Fig. 9(c), solid line) presented a combination of the signals observed
for the monometallic catalysts (Fig. 9 (a) and 9 (b), solid lines). The
high frequency signal between 1800 and 1900 cm−1 and the low fre-
quency band between 1550 and 1750 cm−1 are assigned to NO che-
misorbed on Ni and Mo vacancies, respectively, while a shoulder be-
tween 1750 and 1800 cm−1 corresponds to overlapped signals of NO
adsorbed on both Ni and Mo sites. Therefore, in this case, only the
higher and lower frequency bands are useful for the analysis of the Ni
and Mo species, respectively. An increase in the relative intensity of the
NO-Ni signal (1810 cm−1) compared to the intensity of the NO-Mo
signal (1692–1624 cm−1) was found for the NiMo/MCM-41 catalyst
(Fig. 9 (c), solid line). This result suggests an interaction of Ni with Mo,
and it is reasonable to assume that Ni atoms in the NiMo/MCM-41
catalyst occupy edge positions in the MoS2 slabs decreasing the amount
of Mo vacancies [80]. This effect was more marked for the Nb-modified
NiMo catalyst (Fig. 9 (c), dashed line), where the ratio of the intensities
of the signals of NO-Mo (1672 cm−1) and NO-Ni (1836 cm−1) was even
smaller than in the NiMo/MCM-41 catalyst, indicating a smaller
amount of Mo vacancies for NO adsorption. Therefore, it can be con-
cluded that in the presence of Nb, a larger amount of Ni atoms dec-
Fig. 9. DRIFTS spectra of nitric oxide chemisorbed on in situ sulfided catalysts:
orates the edge surface of the MoS2 slabs. Another possible explanation
(a) Ni, (b) Mo and (c) NiMo supported on MCM-41 (solid lines) and Nb-MCM-41
(dashed lines).
is related to the possibility of the incorporation of Nb on the Ni-Mo-S
active sites as recently proposed [78]. The DFT calculations showed
that the Nb could be stable on the S-edge and Mo-edge with various
substitution rates. A moderate Nb introduction would enhance the HDS
conversion rate and the C-S bond cleavage, whereas the full Nb sub-
stitution hinders desulfurization through the C-S bond cleavage.
On the other hand, the Nb incorporation (Fig. 9, dashed lines) to the
MCM-41 support resulted in a shift of the signals of the adsorbed NO
towards higher wavelength values as compared with unmodified cata-
lysts (Fig. 9, solid lines). This indicates a decrease in the electron
density at the metal adsorption site, leading to weaker Ni-NO and Mo-
NO bonds.

3.3. Hydrodesulfurization of dibenzothiophene

Mo and NiMo catalysts supported on MCM-41 and Nb-MCM-41


were tested in the hydrodesulfurization of dibenzothiophene for 8 h at
300 °C and 7.3 MPa. The catalytic activity results are shown in Fig. 10
and Table 5. Fig. 10 shows the DBT conversions obtained with different
catalysts. It can be observed that in all cases DBT conversions increased
with the reaction time. The catalysts can be ordered in the following
way of increasing DBT conversions: Mo/MCM-41 < Mo/Nb-MCM-
Fig. 10. DBT conversions obtained at different reaction times with: (a) Mo/ 41 < NiMo/MCM-41 < NiMo/Nb-MCM-41. It can be concluded from
MCM-41, (b) Mo/Nb-MCM-41, (c) NiMo/MCM-41 and (d) NiMo/Nb-MCM-41 this order that Ni-promoted catalysts were more active than their un-
catalysts. Reaction conditions: temperature = 300 °C, total pressure = 7.3 MPa promoted analogs, as well as Nb-containing Mo and NiMo catalysts
and reaction time = 8 h. Error bars are shown. were more active than the MCM-41-silica supported ones. The pseudo-
first order rate constants (Table 5) confirm this conclusion. Thus, the
introduction of Ni to the Mo/MCM-41 resulted in almost a four-fold
assigned to the stretching vibration of the N-O bond in the NO increase of the rate constant. Similar results have been reported pre-
monomer adsorbed on Ni2+ cations of the Ni sulfide species [80]. This viously by other authors for CoMo/MCM-41 [30] and NiMo/MCM-41
band is slightly asymmetric toward the low frequency side due to the [31] catalysts. Regarding the Mo and NiMo catalysts supported on Nb-
presence of dinitrosyl species. For the Mo/MCM-41 catalyst (Fig. 9 (b), containing MCM-41 material, the rate constant value increased three-
solid line), two NO vibration modes were observed at 1775 and fold after the incorporation of Ni. The NiMo/Nb-MCM-41 catalyst was
1645 cm−1 associated to a pair of NO molecules coupled in the ad- the most active between the synthesized and tested catalysts. For both
sorbed state as dinitrosyl or dimeric (NO)2 structures on the same unpromoted and Ni-promoted Mo catalysts, the presence of niobia in
molybdenum site [80]. It is well known that these species are adsorbed the support resulted in a 1.8–2.0 increase in the rate constants values. A
on the coordinatively unsaturated sites (CUS) of the sulfided Mo cata- comparison between the activities of the Mo/MCM-41 and Mo/Nb-
lysts. In the case of reduced catalysts, the vibration modes of NO appear MCM-41 catalysts points out that the addition of Nb to the MCM-41
at about 1800–1820 cm−1 for the symmetric stretching vibration mode support produced a two-fold increase in the catalytic activity (k value

10
F.J. Méndez, et al. Fuel 280 (2020) 118550

unpromoted Mo catalysts supported on MCM-41 and Nb-MCM-41


showed substantially higher catalytic activity than the Mo/γ-Al2O3
sample. The Ni-promoted Mo catalysts showed a different trend,
namely, the NiMo/MCM-41 sample was less active than the NiMo/γ-
Al2O3 reference, while the Nb-containing sample was ~ 20% more
active than the alumina-supported analog.
Regarding the reaction products obtained in the HDS of DBT, the
main reaction products were biphenyl (BP), produced through the di-
rect desulfurization route and cyclohexylbenzene (CHB) obtained
through the hydrogenation route (Scheme 1). These products were
detected in all cases (Fig. 11). At the beginning, appreciable amounts of
tetrahydrodibenzothiophene (THDBT) were also detected for the Mo/
Nb-MCM-41, NiMo/MCM-41 and NiMo/Nb-MCM-41 catalysts. This
Scheme 1. Reaction mechanism of DBT hydrodesulfurization. The DDS route, intermediate product gradually disappeared at higher DBT conversions
the direct desulfurization route of the reaction leading to the formation of the (≈ 60%). The formation of THDBT was not detected for the Mo/MCM-
biphenyl (BP) product; the HYD route, the hydrogenation route of the reaction 41 catalyst. In addition, sulfided NiMo/MCM-41 and NiMo/Nb-MCM-
resulting in the formation of the tetrahydrodibenzothiophene (THDBT) inter- 41 catalysts formed small amounts of dicyclohexyl (DCH) at DBT con-
mediate and cyclohexylbenzene (CHB) and dicyclohexyl (DCH) products. versions above 50%. The NiMo catalyst supported on Nb-modified
MCM-41 presented also a small amount (below 2%) of undesired
increased from 0.05 h−1 to 0.10 h−1, respectively, Table 5). This result cracking products that can be attributed to the highest acidity of this
is in line with HRTEM observations suggesting that niobium increases catalyst (Table 4).
the dispersion of the MoS2 active phase and the amount of active sites. To characterize the hydrogenation/hydrogenolysis abilities of the
Finally, reaction rate constants obtained for the prepared Mo and NiMo catalysts, we calculated the HYD/DDS ratios (Fig. 12). These ratios
catalysts supported on (Nb)-MCM-41 were compared to those obtained represent the ratio of the sum of the amount of each product obtained
with the Mo/γ-Al2O3 and NiMo/γ-Al2O3 references (Table 5). Both through the HYD route (THDBT + CHB + DCH) to the amount of the

Fig. 11. Reaction product compositions obtained with different catalysts in the hydrodesulfurization of dibenzothiophene: (A) Mo/MCM-41, (B) Mo/Nb-MCM-41,
(C) NiMo/MCM-41 and (D) NiMo/Nb-MCM-41 catalysts. Reaction products: tetrahydrodibenzothiophene (THDBT), biphenyl (BP), cyclohexylbenzene (CHB), di-
cyclohexyl (DCH) and other products (cracking). Error bars are shown.

11
F.J. Méndez, et al. Fuel 280 (2020) 118550

dispersed (shorter average lengths and lower average stacking num-


bers) on the Nb-containing MCM-41 support. The FT-IR characteriza-
tion of NO chemisorbed on the sulfided catalysts pointed out to a better
promotion of the MoS2 phase with Ni on the Nb-containing support or
to the possibility of the incorporation of some Nb atoms on the edge
surface of the Ni-Mo-S phase. On the other hand, Nb incorporation in
the MCM-41 support resulted in an increase of surface acidity of Mo and
NiMo catalysts (TPD-NH3). According to the catalytic activity tests,
addition of Nb to the MCM-41 support improved activity of the Mo and
NiMo catalysts in the HDS of dibenzothiophene, as well as their hy-
drogenation ability. The role of niobium in the above catalysts can be
described as a support and a promoter.

CRediT authorship contribution statement

Franklin J. Méndez: Investigation, Data curation, Writing - original


draft, Validation. Oscar E. Franco-López: Visualization, Investigation,
Validation. Gabriela Díaz: Investigation, Data curation. Antonio
Gómez-Cortés: Investigation, Validation. Xim Bokhimi: Investigation.
Fig. 12. The HYD/DDS ratios at different conversion values obtained with the Tatiana E. Klimova: Supervision, Conceptualization, Methodology,
following catalysts: (a) Mo/MCM-41, (b) Mo/Nb-MCM-41, (c) NiMo/MCM-41 Writing - review & editing, Funding acquisition.
and (d) NiMo/Nb-MCM-41 catalysts. The HYD/DDS ratio was defined as: HYD/
DDS = (THDBT + CHB + DCH)/BP. Error bars are shown. Declaration of Competing Interest

main product of the DDS route (BP). The HYD/DDS ratios shown in The authors declare that they have no known competing financial
Fig. 12 indicate that the addition of both niobia and nickel promoter interests or personal relationships that could have appeared to influ-
increased the hydrogenation ability of the Mo catalysts. ence the work reported in this paper.
We think that the above activity and selectivity trends are related to
a number of facts. Thus, the effect of the nickel promoter can be re- Acknowledgment
sumed, for the calcined catalysts, as an increase in the dispersion of the
Mo oxide species and a decrease in the temperature of their reduction, F.J. Méndez acknowledges DGAPA-UNAM for the postdoctoral
while for the sulfided catalysts, Ni addition leads to the formation of a grant. Financial support from DGAPA-UNAM (PAPIIT project IN-
so-called Ni-Mo-S phase, in which Ni atoms decorate the edge surface of 115218) is gratefully acknowledged. The authors thank C. Salcedo
the MoS2 particles. On the other side, the addition of niobium to the Luna, A. Morales, I. Puente Lee and R. Ruiz Trejo for technical assis-
unpromoted Mo/MCM-41 catalyst resulted in the improvement of the tance.
dispersion of the Mo oxide precursor and sulfide MoS2 particles, as well
as in the change in their morphology (formation of shorter, but more References
stacked MoS2 particles). This explains the higher activity and higher
HYD selectivity of the Mo/Nb-MCM-41 sample compared to the Mo/ [1] Knudsen KG, Cooper BH, Topsøe H. Appl Catal A 1999;189:205–15.
MCM-41 one. For the NiMo/MCM-41 catalyst, niobium addition to the [2] Song C, Ma X. Appl Catal B 2003;41:207–38.
[3] Topsøe H, Clausen BS, Massoth FE. Hydrotreating catalysis: Science and technology.
support was favorable for the further decrease in the temperature of Berlin: Springer-Verlag; 1996.
reduction of Mo oxide species. For the sulfided NiMo/Nb-MCM-41 [4] Sun M, Adjaye J, Nelson AE. Appl Catal A 2004;263:131–43.
catalyst, it seems that the presence of Nb improved the promotion of the [5] Topsøe H, Clausen BS. Appl Catal 1986;25:273–93.
[6] Vissers JPR, Scheffer B, de Beer VHJ, Moulijn JA, Prins R. J Catal 1987;105:277–84.
MoS2 phase by Ni or possibly some Nb atoms were also incorporated on
[7] van Veen JAR, Colijn HA, Hendriks PAJM, van Welsenes AJ. Fuel Proc Techn
the edge surface of the Ni-Mo-S phase. Finally, the niobium in- 1993;35:137–57.
corporation to the MCM-41 support increased the acidity of the Mo and [8] Bouwens SMAM, Vanzon FBM, Vandijk MP, Vanderkraan AM, Debeer VHJ,
Vanveen JAR, et al. J Catal 1994;146:375–93.
NiMo catalysts that could also affect the HYD ability of the catalysts and
[9] Chianelli RR, Daage M, Ledoux MJ. Adv Catal 1994;40:177–232.
the distribution of the obtained products. Resuming all the above ob- [10] Shimada H. Catal Today 2003;86:17–29.
servations, we consider that niobium acts not only as a support, but also [11] Coulier L, de Beer VHJ, van Veen JAR, Niemantsverdriet JW. J Catal
as a promoter in the Mo and NiMo catalysts supported on MCM-41. 2001;197:26–33.
[12] Calafat A, Laine J, López-Agudo A, Palacios JM. J Catal 1996;162:20–30.
[13] Farag H, Whitehurst DD, Mochida I. Ind Eng Chem Res 1998;37:3533–9.
[14] Pawelec B, Mariscal R, Fierro JLG, Greenwood A, Vasudevan PT. Appl Catal A
4. Conclusions 2001;206:295–307.
[15] Lee JJ, Han S, Kim H, Koh JH, Hyeon T, Moon SH. Catal Today 2003;86:141–9.
[16] Araki Y, Honna K, Shimada H. J Catal 2002;207:361–70.
In the present work, Mo and NiMo catalysts supported on the MCM- [17] Klimova TE, Gutiérrez OY, Lizama L, Amezcua J. Micro Meso Mater 2010;133:91–9.
41 silica and the Nb-containing MCM-41 material were synthesized, [18] Maity SK, Rana MS, Srinivas BN, Bej SK, Murali-Dhar G, Prasada-Rao TSR. J Molec
characterized and tested in the hydrodesulfurization of dibenzothio- Catal A 2000;153:121–7.
[19] Jia M, Afanasiev P, Vrinat M. Appl Catal A 2005;278:213–21.
phene. Obtained results show that the addition of a small amount [20] Cortés-Jácome MA, Escobar J, Angeles-Chávez C, López-Salinas E, Romero E, Ferrat
(about 4 wt%) of Nb2O5 to the MCM-41 support improved the disper- G, et al. Catal Today 2008;130:56–62.
sion of the MoO3 and NiMoO4 crystalline phases in the calcined Mo and [21] Toledo JA, Cortés-Jácome MA, Angeles-Chávez C, Escobar J, Barrera MC, López-
Salinas E. Appl Catal B 2009;90:213–23.
NiMo catalysts. TPR and UV–vis DRS characterizations showed that [22] Ortega-Domínguez RA, Mendoza-Nieto JA, Hernández-Hipólito P, Garrido-Sánchez
octahedrally coordinated Mo6+ oxide species were predominant in all F, Escobar-Aguilar J, Barri SAI, et al. J Catal 2015;329:457–70.
prepared catalysts, and that the presence of niobia in the NiMo/Nb- [23] Palcheva R, Dimitrov L, Tyuliev G, Spojakina A, Jiratova K. Appl Surf Sci
2013;265:309–16.
MCM-41 catalyst resulted in a decrease in the temperature of the first [24] Zepeda TA, Halachev T, Pawelec B, Nava R, Klimova TE, Fuentes GA, et al. Catal
step of reduction of Mo6+ oxide species. Characterization of the sul- Commun 2006;7:33–41.
fided catalysts by HRTEM showed that the MoS2 active phase was better [25] Zepeda TA, Pawelec B, Fierro JLG, Halachev T. J Catal 2006;242:254–69.

12
F.J. Méndez, et al. Fuel 280 (2020) 118550

[26] Nava R, Pawelec B, Morales J, Ortega RA, Fierro JLG. Micro Meso Mater Klimova TE. Appl Catal B 2017;219:479–91.
2009;118:189–201. [55] F.J. Méndez, G. Bravo-Ascención, M. González-Mota, I. Puente-Lee, X. Bokhimi, T.E.
[27] Gutiérrez OY, Fuentes GA, Salcedo C, Klimova TE. Catal Today 2006;116:485–97. Klimova, Catal. Today, in press, https://doi.org/10.1016/j.cattod.2018.03.039.
[28] Rayo P, Rana MS, Ramírez J, Ancheyta J, Aguilar-Elguézabal A. Catal Today [56] Beck JS, Vartuli JC, Roth WJ, Leonowicz ME, Kresge CT, Schmitt KD, et al. JACS
2008;130:283–91. 1992;114:10834–43.
[29] Rayo P, Ramírez J, Rana MS, Ancheyta J, Aguilar-Elguézabal A. Ind Eng Chem Res [57] Marler B, Oberhagemann U, Vortmann S, Gies H. Micro Mater 1996;6:375–83.
2009;48:1242–8. [58] International Center for Diffraction Data, PCPDFWIN v.2.02. PDF-2 Data Base,
[30] Wang A, Wang Y, Kabe T, Chen Y, Ishihara A, Qian W. J Catal 2001;199:19–29. Newtown Philadelphia, 1995.
[31] Wang A, Wang Y, Kabe T, Chen Y, Ishihara A, Qian W, et al. J Catal [59] Leofanti G, Padovan M, Tozzola G, Venturelli B. Catal Today 1998;41:207–19.
2002;210:319–27. [60] Thommes M, Kaneko K, Neimark AV, Olivier JP, Rodriguez-Reinoso F, Rouquerol J,
[32] Klimova TE, Calderón M, Ramı́rez J. Appl Catal A 2003;240:29–40. et al. Pure Appl Chem 2015;87:1051–69.
[33] Turaga UT, Song C. Catal Today 2003;86:129–40. [61] Méndez FJ, Llanos A, Echeverría M, Jáuregui R, Villasana Y, Díaz Y, et al. Fuel
[34] Méndez FJ, Bastardo-González E, Betancourt P, Paiva L, Brito JL. Catal Letters 2013;110:249–58.
2013;143:93–100. [62] Shen S, Chen J, Koodali RT, Hu Y, Xiao Q, Zhou J, et al. Appl Catal B
[35] Ziolek M, Sobczak I. Catal Today 2017;285:211–25. 2014;150–151:138–46.
[36] García-Sancho C, Moreno-Tost R, Mérida-Robles JM, Santamaría-González J, [63] P. Carraro, V. Elías, A. García Blanco, K. Sapag, S. Moreno, M. Oliva, G. Eimer,
Jiménez-López A, Maireles-Torres P. Appl Catal B 2011;108–109:161–7. Micro Meso Mater 191 (2014) 103-111.
[37] Chagas P, Oliveira HS, Mambrini R, Le Hyaric M, de Almeida MV, Oliveira LCA. [64] Hadjiivanov K, Mihaylov M, Klissurski D, Stefanov P, Abadjieva N, Vassileva E,
Appl Catal A 2013;454:88–92. et al. J Catal 1999;185:314–23.
[38] C. Tiozzo, C. Bisio, F. Carniato, A. Phys. Chem. Chem. Phys. Gallo, S.L. Scott, R. [65] Giordano N, Bart JCJ, Vaghi A, Castellan A, Martinotti G. J Catal 1975;36:81–92.
Psaro, M. Guidotti, Phys. Chem. Chem. Phys. 15 (2013) 13354-13362. [66] Liu Z, Chen Y. J Catal 1998;177:314–24.
[39] Allali N, Marie AM, Danot M, Geantet C, Breysse M. J Catal 1995;156:279–89. [67] Williams CC, Ekerdt JG, Jehng JM, Hardcastle FD, Turek AM, Wachs IE. J Phys
[40] Allali N, Yacoubi A, Nadiri A, Geantet C, Danot M. Ann Chim Sci Mat Chem 1991;95:8781–91.
1998;23:209–12. [68] Liu D, Cheo WNE, Lim YWY, Borgna A, Lau R, Yang Y. Catal Today
[41] Geantet C, Afonso J, Breysse M, Allali N, Danot M. Catal Today 1996;28:23–30. 2010;154:229–36.
[42] Mansouri A, Semagina N. ACS Appl Nano Mater 2018;1:4408–12. [69] Wojcieszak R, Monteverdi S, Mercy M, Nowak I, Ziolek M, Bettahar MM. Appl Catal
[43] Mansouri A, Semagina N. ACS Catal 2018;8:7621–32. A 2004;268:241–53.
[44] Weissman JG. Catal Today 1996;28:159–66. [70] Yang Y, Lim S, Du G, Chen Y, Ciuparu D, Haller GL. J Phys Chem B
[45] Puello-Polo E, Marquez E, Brito JL. J Sol-Gel Sci Techn 2018;88:90–9. 2005;109:13237–46.
[46] Faro Jr AC, dos Santos ACB. Catal Today 2006;118:402–9. [71] López-Cordero R, Gil-Llambias FJ, López-Agudo A. Appl Catal 1991;74:125–36.
[47] Rocha AS, Faro Jr AC, Oliviero L, Van Gestel J, Maugé F. J Catal 2007;252:321–34. [72] Jibril BY, Ahmed S. Catal Commun 2006;7:990–6.
[48] Kaluža L, Zdražil M. Catal Comm 2018;107:62–7. [73] Fisher WG, Sienko MJ. Inorg Chem 1980;19:39–43.
[49] Gaborit V, Allali N, Geantet C, Breysse M, Vrinat M, Danot M. Catal Today [74] Kasztelan S, Toulhoat H, Grimblot J, Bonnelle JP. Appl Catal 1984;13:127–59.
2000;57:267–73. [75] Gutiérrez OY, Klimova TE. J Catal 2011;281:50–62.
[50] Cedeño-Caero L, Romero AR, Ramirez J. Catal Today 2003;78:513–8. [76] Hayden TF, Dumesic JA. J Catal 1987;103:366–84.
[51] Bouadjadja-Rohan K, Bonduelle-Skrzypczak A, Hugon A, Fournier M, Lamonier C, [77] Hensen EJM, Kooyman PJ, van der Meer Y, van der Kraan AM, de Beer VHJ, van
Lancelot C. ChemCatChem 2015;7:297–302. Veen JAR, et al. J Catal 2001;199:224–35.
[52] Bouadjadja-Rohan K, Lancelot C, Fournier M, Bonduelle-Skrzypczak A, Hugon A, [78] Ding S, Li A, Jiang S, Zhou Y, Wei Q, Zhou W, et al. Fuel 2019;237:429–41.
Mentré O, et al. Eur J Inorg Chem 2015;12:2067–75. [79] J. Ramı́rez, R. Contreras, P. Castillo, T. Klimova, R. Zárate, R. Luna, Appl Catal A
[53] Palcheva R, Kaluža L, Dimitrov L, Tyuliev G, Avdeev G, Jirátová K, et al. Appl Catal 197 (2000) 69-78.
A 2016;520:24–34. [80] Navarro R, Pawelec B, Fierro JLG, Vasudevan PT. Appl Catal A 1996;148:23–40.
[54] Méndez FJ, Franco-López OE, Bokhimi X, Solís-Casados DA, Escobar-Alarcón L,

13

You might also like