You are on page 1of 10

Journal of Sol-Gel Science and Technology

https://doi.org/10.1007/s10971-018-4792-x

REVIEW PAPER: SOL–GEL AND HYBRID MATERIALS FOR CATALYTIC,


PHOTOELECTROCHEMICAL, AND SENSOR APPLICATIONS

One-pot synthesis of Nb-modified Al2O3 support for NiMo


hydrodesulfurization catalysts
Esneyder Puello-Polo1 Edgar Marquez2 J. L. Brito3,4
● ●

Received: 11 April 2018 / Accepted: 24 August 2018


© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
The effect of Nb as a support modifier on the NiMo6/Al2O3–Nb2O5(x) (x = 0, 1, 4, and 8 wt% Nb) catalysts was studied. The
supports were prepared by one-pot coprecipitation from soluble precursors. The XRF analysis of the catalysts showed that
the contents of Mo and Ni increased slightly with the presence of Nb. Micropore area and pore volume augmented
importantly with Nb content, resulting in pore diameters between 5.3 and 9.3 nm. XPS analysis showed that the presence of
Nb decreases the active metal–support interaction, improving the Mo and Ni sulfidation degree. The Raman spectra of
1234567890();,:
1234567890();,:

sulfided catalysts suggested an increase in the number of layers of MoS2 in the presence of Nb. Generally, the thiophene
HDS activity at normal pressure of sulfided NiMo6/Al2O3–Nb2O5(8) was greater than that of the sulfided catalysts with x =
0, 1, and 4 wt% Nb, which can be attributed to the Nb promotion that would have an effect on the type of active site
(Brønsted or Lewis acidic sites), since the number of sites by CO chemisorption for sulfided NiMo6/Al2O3–Nb2O5(x) did not
show correlation with the catalytic activity. The high-pressure HDS activity of dibenzothiophene was also greater in the
presence of Nb, and the hydrogenation route was preferred for the Nb-promoted solid, while the unpromoted one showed a
larger yield of direct desulfurization products.

Graphical Abstract

* Esneyder Puello-Polo Norte, Barranquilla, Colombia


snypollqco@yahoo.com 3
Laboratorio de Fisicoquímica de Superficies, Centro de Química,
1 Instituto Venezolano de Investigaciones Científicas, IVIC.
Grupo de Investigación en Oxi/Hidrotratamiento Catalítico y
Apartado 20632, Caracas 1020-A, Venezuela
Nuevos Materiales, Programa de Química-Ciencias Básicas
4
Universidad del Atlántico, Barranquilla, Colombia Yachay Tech University, Urcuqui 100119 Imbabura, Ecuador
2
Grupo de investigación en Química y Biología, Universidad del
Journal of Sol-Gel Science and Technology

Highlights
● The niobium-containing catalyst presented the best catalytic activity.
● Niobium promotion influenced the type of active site.
● The presence of Nb improves the sulfidation degree of Mo and Ni.
● Nb increases the number of Lewis acid sites related to HDS activity.
Keywords Anderson-type heteropolyoxomolybdate Hydrodesulfurization Niobia–alumina One-pot synthesis
● ● ●

1 Introduction niobium–aluminum oxide catalysts containing 5 wt%


Nb2O5 (the highest Lewis acidity) [15]. Rocha et al. com-
Hydrodesulfurization (HDS) catalysts are fundamental to pared the activity of NiMoS catalysts supported on niobia,
decrease pollutant emissions, because sulfur compounds in alumina, and niobia/alumina in HDS and HDN reactions. It
fuels cause environmental problems due to the production was shown that on a niobia/alumina support, the formation
of SOx [1, 2]. Alumina is the most used support for of the NiMoS phase can be directly linked to alumina not
hydrotreatment (HDT) catalysts, since it has excellent covered by niobia [16].
mechanical and dispersion properties [3, 4]. However, alu- Palcheva et al. [17] and Cedeño et al. [18] have reported
mina shows relatively strong metal–support interactions the preparation and study of Nb-modified SBA-15, HMS,
with the components of the active phase, which leads to the and MCM-41 as supports of NiMo. The HDS activity of the
formation of compounds non-active for HDT (Ni(Co) NiMo catalysts supported on Nb-SBA-15, Nb-HMS, and
Al2O4) and consequently to partial sulfiding of metal spe- Nb-MCM-41 was assigned to a better dispersion of Ni and
cies [5, 6]. Although this can be improved by first covering Mo promoted by Nb. Recently, Mendez et al. studied HDS
the alumina surface with a MoO3 monolayer before Ni or over Nb-modified MCM-41-supported NiMo and CoMo
Co impregnation [5], during activation by sulfiding, the catalysts [19]. The results show that the incorporation of
MoO3 monolayer is disrupted and thus direct interaction small amounts of Nb (3–5 wt%) increased the catalytic
promoter-support becomes almost unavoidable. By con- activity of both types of catalysts and affected their selec-
sidering these limitations, recent studies have shown that it tivity. In all these cases, the incorporation of Nb is carried
is convenient to modify the surface properties of the alu- out by applying a conventional impregnation method. On
mina to decrease the metal–support interaction and, in this the other hand, the incorporation of additional metal com-
regard, the acid–base characteristics of the support would ponents during the synthesis of mesoporous materials by a
play a fundamental role in the structure and activity of these one-pot process possesses special advantages because the
catalysts [4, 6]. Niobium-containing solids have attracted metal components can be highly dispersed and better resist
special attention due to their applications as active phases, against sintering even on suffering from a high-temperature
supports, or promoters [7, 8]. However, the use of pure treatment, as has been shown in a recent study by Lai [20].
niobium oxide as support for hydroprocessing catalysts is In this regard, the present work reports the effect of the
limited since it also presents strong metal–support interac- addition of niobium during the synthesis of mesoporous
tion, which affects the catalytic properties [9, 10]. On the alumina supports by a one-pot process on some physico-
other hand, the addition of niobia as a support promoter to chemical characteristics of the resulting NiMo/Al2O3–
the catalysts used in HDT could be of interest, as it Nb2O5 catalysts and their HDS activity.
apparently leads to increase of the catalytic activities [11].
One of the reasons for this could be attributed to the
occupation of tetrahedral sites by niobium, inhibiting the 2 Experimental
formation of Ni(Co) aluminates and resulting in an increase
in the amount of nickel (cobalt) species that can participate 2.1 Preparation of alumina modified with niobium
in the decoration of MoS2 slabs by forming “Ni(Co)–Mo–
S” phase, as has been demonstrated for gallium, silicon, and The Al2O3–Nb2O5 catalytic supports with different Nb
other alumina-modifier elements [12–14]. The addition of contents were prepared by a modification of the one-pot
niobia for the preparation of NiMo/Al catalysts, either as synthesis method of Lai [20]. Appropriate amounts of
surface niobium–aluminum oxide phases or as bulk nio- AlCl3·6H2O (Sigma-Aldrich, 99%) and urea (Merck,
bium–aluminum oxides has been reported by Weissman, 99.5%) were dissolved in deionized water under constant
who suggested that the enhanced activity with the addition stirring at pH between 5 and 6, and then a given quantity of
of niobia was due to the increase in the support acidity. starch (Sigma-Aldrich, ACS reagent) and the appropriate
Maximum rates of sulfur and nitrogen removal occurred for amount of ammonium niobate(V) oxalate hydrate (Sigma-
Journal of Sol-Gel Science and Technology

Aldrich, 99.9%) dissolved in a citric acid (Merck, 99.5%)

wt% percentage by weight, SBET BET surface area, Smicro micropores surface area, Sext external surface area, Smicro/SBET fraction of the micropore area to the total surface area, VT total volume of
Mo (wt%) Ni (wt%) Nb (wt%) SBET (m2/g) Sext (m2/g) Smicro (m2/g) Dp (nm) Vmeso (cm3/g) Vmicro (cm3/g) VT (cm3/g) Smicro/SBET (%) Vmicro/VT (%)
solution were slowly added to form a uniform pasty mix-
ture. Finally, the paste was stirred at 363 K for 2 h. Then,

11.9
2.8
6.9
4.7

8.7
the as-made sample was further dried at 423 K for 12 h
overnight and the sample was moved to a tubular furnace
and treated with air at a flow rate of 50 cm3 min−1. Calci-
nation of the precursor was carried out at 873 K for 6 h. The
obtained solid sample was milled and the particles of 100–

13
21
19
30
24
140 mesh were collected for use (denoted as Al2O3–Nb2O5
(x), x = 0, 1, 4, or 8 wt% Nb). The mass ratio of
AlCl3·6H2O:urea:starch:citric acid: (ammonium niobate

0.470
0.282
0.396
0.243
0.222
oxalate) was 1:0.17:0.84:0.1: (0; 0.16; 0.65; or 1.3) so that
the final catalyst contained nominally 0, 1, 4, or 8 wt% Nb.

pores, Dp mean pore diameter, Vmeso mesopore volume, Vmicro micropore volume, Vmicro/VT fraction of the micropore volume to the total volume
2.2 Preparation of catalyst precursors supported on

0.0132
0.0194
0.0190
0.0289
0.0193
Table 1 Composition and textural properties of NiMo6/Al2O3–Nb2O5(x) catalyst varying the content of niobium (x = 0, 1, 4, and 8 wt%)
niobium-modified alumina

The Anderson ammonium salt (NH4)4[NiMo6O24H6]•5H2O


(hereafter named NiMo6) was obtained as described else-
where [21]. The supported precursor from hetero-

0.457
0.263
0.377
0.214
0.203
Specific surface area and porous structure characteristics
polyoxomolybdate complexes was impregnated in excess of
pore volume, adding dropwise an aqueous solution of
NiMo6 (15 wt% Mo and 1.6 wt% Ni) to a flask containing 5

8.1
7.2
9.3
5.9
5.3
g of support, under stirring at 323 K and pH around 5–6.
The impregnation step lasted until removal of the solvent by
evaporation. Finally, the mass obtained was further dried at
423 K for 12 h.
The (NH4)4[NiMo6O24H6]•5H2O supported on alumina
34
49
50
72
48
modified with niobium will be identified as NiMo6/Al2O3–
Nb2O5(x), where x is the weight content of niobium (x = 0,
1, 4, or 8 wt%).
227
184
218
170
150

2.3 Catalyst characterization

The elemental analysis for NiMo6/Al2O3–Nb2O5(x) was


261
233
268
242
198

determined by X-ray fluorescence using a MagixPro PW-


2440 Philips instrument. The results for all compounds were
in agreement, within experimental accuracy, with the
0.920
4.30
7.70

nominal metal contents (Table 1). The textural properties


Experimental composition


0

were determined by means of the physisorption of N2 at 77


K using a Micromeritics 3FLEXTM instrument. Prior to the
1.30
1.80
1.90
2.10

measurements, the samples were degassed at 573 K for 16 h


in vacuum of 10−6 mmHg. The surface areas of samples


were calculated by the Brunauer–Emmett–Teller (BET)
multipoint method, and the total pore volume and pore size
NiMo6/Al2O3–Nb2O5(0) 13.0
NiMo6/Al2O3–Nb2O5(1) 14.4
NiMo6/Al2O3–Nb2O5(4) 16.2
NiMo6/Al2O3–Nb2O5(8) 16.4

distribution were determined from the adsorption branch of


the isotherm using the Barret–Joyner–Halenda (BJH)
model. The total mesopore volume (Vme) and mean pore
diameter (Dp) were provided by the BJH method, while the
Al2O3–Nb2O5(0)

micropore volume (Vmicro) and external surface area (Sext)


were evaluated by the t-plot method. Hence, the total pore
volume was estimated by the addition of microporous and
Solid

mesoporous volumes. The surface area due to micropores


Journal of Sol-Gel Science and Technology

(Smicro) was obtained from the equation: Smicro = SBET−Sext considered to calculate the number of active sites was 1:1.
[21–23]. XRD analysis of the samples was carried out using Likewise, an additional correction for chemisorption on
a SIEMENS D-5005 diffractometer with a Cu Kα radiation pure alumina or alumina modified with niobium was
source (λ = 1.5418 Å) and Ni filter, within the range 5° ≤ determined under conditions identical to those of the cata-
2θ ≤ 90°, step size of 0.02°, and acquisition speed of 0.08°/ lyst [25].
s. For the identification of the crystallographic phase(s), the
JCPDS data files were preferentially used of Powder Dif- 2.4 Catalytic tests
fraction File, ICDD, Newtown Square, Philadelphia (1995).
The surface composition of both the most active catalyst Prior to the HDS catalytic reactions, NiMo6/Al2O3–Nb2O5
and of the unpromoted one (NiMo6/Al2O3–Nb2O5(x) with x (x) catalysts were activated by sulfiding in situ under a
= 8 and 0, respectively) was determined by means of X-ray feedstock of CS2 (2% in heptane) and 70 cm3 min−1 of
photoelectron spectroscopy (XPS) with a Thermo Scientific hydrogen at 573 K for 4 h to attain a reproducible and stable
K-Alpha spectrometer, equipped with a dual (non-mono- initial state of the catalytic surface. The thiophene HDS
chromatic) Mg/Al anode, operated at 400 W and under a reaction was carried out in a fixed-bed tubular stainless-steel
vacuum better than 10−9 torr. Calibration of the instrument reactor (30 cm long, 20 mm I.D.) with an axial thermowell
was done employing the Au 4f7/2 line at 83.9 eV. The containing a thermocouple centered in the catalyst bed, at
internal referencing of binding energies was made using the 623 K and atmospheric pressure. The HDS test conditions
dominating Al 2p band of the support at 74.4 eV [21]. were 500 mg of catalyst, flow of 0.333 cm3 min−1 of the
Before XPS analysis, samples were sulfided ex situ as thiophene (2%)/C7H16 mixture, 70 cm3 min−1 of H2, 0.1
described below, afterward immersed in purified n-hexane, MPa, and kept for 4.5 h of reaction time.
and quickly transferred to the XPS instrument, where the The HDS of dibenzothiophene (DBT) of NiMo6/Al2O3–
solvent was slowly removed under vacuum. Curve fitting of Nb2O5(8), the most active catalyst for thiophene HDS, and
the spectra was carried out by means of the XPSPEAK of NiMo6/Al2O3–Nb2O5(0), the unpromoted one, was car-
version 4.1 and XPS GRAPH programs, using a standard ried out in a high-pressure batch reactor. In order to mini-
nonlinear least-squares curve-fitting routine, employing mize internal diffusion limitations, all catalysts were
mixed Gaussian/Lorentzian peak shapes of variable pro- thoroughly ground in a mortar to a fine powder. After
portion, after baseline subtraction by the Shirley method. presulfiding ex situ as described above, the sample was
Raman spectra of the catalysts after sulfiding at 593 K cooled down to room temperature under a hydrogen flow.
were obtained in a Horiba Scientific Raman micro- The sulfided catalyst (ca. 500 mg) was then introduced into
spectrometer LabRAM HR Evolution at room temperature the batch reactor containing 100 mL of DBT in cyclohexane
using the 532-nm line and between 41.1 and 56.1 μW He– (0.032 mol/L) under the following reaction conditions: the
Ne laser as the excitation source through an Olympus TM reactor was pressurized to 8.2 MPa with hydrogen and
BX41 optical microscope. The pyridine adsorption experi- heated up to 593 K, and kept for 6 h of reaction time at 700
ments (FTIR Py) of sulfided NiMo6/Al2O3–Nb2O5(x) were rpm. The total DBT conversion was calculated as DBT
performed with the aim to monitor both the type (Brønsted disappearance, and direct desulfurization (DDS) and
or Lewis) of surface acid sites and their relative abundance hydrogenation (HYD) selectivities were defined as Eqs. (1)
[24]. Spectra were recorded in 256 scans using a Nicolet and (2), respectively:
FTIR spectrometer and the data analysis was performed
with the spectrometer software Omnic 8.1. The spectrum DDS : ðBPÞ  100%=ðCHB þ BP þ THDBT þ BCHÞ
recorded before adsorption of pyridine was subtracted from ð1Þ
spectra taken after pyridine adsorption. Band intensities
were corrected for slight differences in the weight of the HYD : ðCHB þ THDBT þ BCHÞ  100%=
ð2Þ
wafer and normalized to a disk of 8 mg cm−2. ðCHB þ BP þ THDBT þ BCHÞ;
The CO chemisorption analyses of sulfided NiMo6/
Al2O3–Nb2O5(x) were carried out under static volumetric where BP, CHB, THDBT, and BCH are biphenyl,
conditions in a Micromeritics ASAP 2020 apparatus. cyclohexylbenzene, tetrahydrodibenzothiophene, and bicy-
Samples were sulfided ex situ and transferred in an inert clohexyl, respectively.
atmosphere. Prior to measurement, samples were treated The course and products of the thiophene and DBT HDS
in situ in H2 at 673 K (30 min) and evacuated at 308 K and reaction was followed using a GC–MS Agilent Technolo-
105 kPa for 10 h. The chemisorption isotherm was obtained gies 7890B GC System-5977A MSD and a Shimadzu GC-
by measuring the amount of CO adsorbed between 13.3 and 2014 gas chromatograph, with sampling of the liquid
93.3 kPa at 308 K. The chemisorption stoichiometry effluent occurring at 60-min intervals, and the catalytic
Journal of Sol-Gel Science and Technology

350

300

250
Adsorbed volume (cm 3 g-1)

200

150

100

50

0
0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1
P/P0

Fig. 1 Adsorption/desorption isotherms of N2 at 77 K on the γ-Al2O3


and NiMo6/Al2O3–Nb2O5(x) catalyst varying the content of niobium
(x = 0, 1, 4, and 8 wt%). (□) γ-Al2O3; (▲) NiMo6/Al2O3–Nb2O5(0);
(♦) NiMo6/Al2O3–Nb2O5(1); (■) NiMo6/Al2O3–Nb2O5(4); (●)
NiMo6/Al2O3–Nb2O5(8)

activities of the catalysts were determined. The alumina


support and the Nb-modified alumina showed negligible
thiophene and DBT conversion. Absence of mass and heat-
flow transport effects was verified according to established
procedures [26].
All experiments reported in this work (synthesis proto-
cols, characterizations, and catalytic activity measurements)
were carried out at least in duplicate. Good reproducibility
was verified, better than 10% in all quantitative Fig. 2 XRD of NiMo6/Al2O3–Nb2O5(x) catalyst varying the content of
measurements. niobium (x = 0, 1, 4, and 8 wt%). (●) γ-Al2O3

3 Results and discussion the total pore volume follows roughly the same trend. From
the corresponding pore size distributions, the precursors and
3.1 Textural properties supports mostly locate in the range of mesopores (2–50
nm), with values of BJH mesopores mean size between 5.3
The N2 physisorption isotherms of γ-Al2O3 and NiMo6/ and 9.3 nm.
Al2O3–Nb2O5(x) are shown in Fig. 1 and the textural data As listed in Table 1, the presence of Nb causes a more
derived from the isotherms are summarized in Table 1. All pronounced decrease of the surface area for samples con-
the isotherms belong to type IV in the IUPAC classification. taining Mo and Ni with respect to the bare support (except
In their initial part, they show a fast increase that reflects NiMo6/Al2O3–Nb2O5(1) which can be related to the large
significant micropore surface area and volume which aug- generation of micropores in this case). Such behavior can be
ment importantly with Nb content. Thus, when comparing related to the migration of the metallic phases into the
the undoped γ-Al2O3 with NiMo6/Al2O3–Nb2O5(x), Smicro/ support pores during the impregnation process and/or
SBET varies between 13 and 30% and Vmicro/VT between 2.8 synthesis of the material that decreases their pore volume
and 11.9%. These trends can be attributed to the generation and therefore the surface area [28].
of microporosity induced by Nb (see Table 1). The hys-
teresis loops can be observed at intermediate and high 3.2 XRD analysis
relative pressures (P/P0 > 0.5), which is mainly due to the
presence of ink-bottle type of mesopores [27]. The overall The XRD analysis of NiMo6/Al2O3–Nb2O5(x) with x = 0, 1,
surface area was found to increase as follows: NiMo6/ 4, and 8 wt% Nb, irrespective of the presence of Nb,
Al2O3–Nb2O5(8) < NiMo6/Al2O3–Nb2O5(0) < NiMo6/ revealed no lines other than those characteristics of γ-Al2O3
Al2O3–Nb2O5(4) < Al2O3 < NiMo6/Al2O3–Nb2O5(1), while at 2θ = 67.10; 45.90; and 37.48 (card No. 10-0425, see Fig.
Journal of Sol-Gel Science and Technology

Fig. 3 X-ray photoelectron spectra Mo 3d, Ni 2p, and S 2p regions of sulfided NiMo6/Al2O3–Nb2O5(x) catalyst varying the content of niobium
(x = 0 and 8 wt%)

2) and it did not show appreciable changes on XRD pat- important for the Nb-containing sample, which showed no
terns. Diffraction peaks due to (NH4)4[NiMo6O24H6]•5H2O presence of sulfates.
or Nb2O5 were absent probably because the crystallites are The Ni 2p region spectra show four Ni 2p3/2–1/2 peaks at
too small to give XRD signals or the particles of POMs and 853.1–853.6; 854.2, 856.1–856.7; and 862 eV [31]. These
Nb2O5 were well dispersed on the support. signals suggest the presence of NixSy sulfide phases (Ni2S3,
Ni9S8, or NiS), the so-called NiMoS phase, the NiMoO4
3.3 XPS analysis species, and the strong shake-up lines characteristic of Ni2+
species in a Ni–Mo–O matrix. Note that the signal assigned
Figure 3 presents the XPS spectra of the Mo 3d5/2–3/2 region, to NiMoS is more prominent for the Nb-containing catalyst,
showing signals that confirm the presence on the surface of at the expense of the NiMoO4 one. In this regard, the (Ni +
Mo4+ (229 eV) attributable to sulfide-phase Mo5+ (230.3 Mo)/S atomic ratio is higher for NiMo6/Al2O3–Nb2O5(0)
eV) that can be assigned to oxysulfide species and Mo6+ than for NiMo6/Al2O3–Nb2O5(8) (Table 2).
(232.5 eV) which can be identified as MoO3-like com- The XPS spectra of Nb 3d5/2–3/2 (not shown), reported a
pounds [29, 30]. It can be seen that Nb presence is reflected doublet centered at 207.6 and 210.3 eV which can be
in a higher amount of Mo4+ (sulfided) species and lower assigned to Nb5+ as in Nb2O5 [32].
concentrations of surface oxidic phases (Mo5+ and Mo6+). In spite of the variation of the types of species of Mo, Ni,
Likewise, in the XPS spectral Mo 3d5/2 region, a quite and S at the surface by the presence of Nb, their total atomic
visible shoulder is found at 226.5 eV, corresponding to the percentages do not have significant differences. This is
2-s signal of sulfur. This can be confirmed by the presence possibly because the synthesis of both catalysts from an
of three bands in the S 2p3/2 region [30]: a signal at 161.7 Anderson–Evans-type heteropolyanion guarantees uni-
eV due to terminal disulfide and/or sulfide (S2–) ligands in formity in the deposition of the precursor as it has been
MoS2 or NiS phases, another signal at 163.1 eV corre- reported by Puello et al. [13, 21]; hence, the Ni/Mo atomic
sponding to bridging disulfide (S22−) ligands, and the signal ratios were in agreement (see Table 2). On the other hand,
at 168.9 eV which can be assigned to SO42−. The signals the species of Mo and Ni partially sulfided of these catalysts
due to sulfide and disulfide species are much more have been evidenced in a variety of studies [33]. These
Journal of Sol-Gel Science and Technology

Atomic ratio—

SO42− 168.9 eV Mo/Ni (Mo +


Ni)/S

1.5

1.3
XPS

6.8

6.8
NiMo6/Al2O3-Nb2O5(8)

36.8
(%)

NiMo6/Al2O3-Nb2O5(4)

S22−163.1 eV
Table 2 Distribution of Mo, Ni, and S surface species in sulfided NiMo6/Al2O3–Nb2O5(x) catalyst varying the content of niobium (x = 0 and 8 wt%)

12.4

15.5
(%)
Ni2+ 856.7 eV S2− 161.7 eV
S 2p3/2–2p1/2

NiMo6/Al2O3-Nb2O5(1)
50.8

84.5
(%)

NiMo6/Al2O3-Nb2O5(0)

320 350 380 410 440 470 500


40.0

30.9
(%)

Raman shift (cm-1)


Mo4+ 229 eV Mo5+ 230.3 eV Mo6+ 232.5 eV NixSy 853.3 eV NiMoS 854.2 eV

Fig. 4 Raman spectra of sulfided NiMo6/Al2O3–Nb2O5(x) catalyst


varying the content of niobium (x = 0, 1, 4, and 8 wt%)
26.7

32.3
(%)

works showed that it is possible that the catalysts react with


the gaseous H2S produced during the HDS reaction to
Ni 2p3/2–2p1/2

become completely sulfided.

3.4 Raman analysis and FTIR of adsorbed pyridine


19.5

18.4
(%)

The Raman spectra (see Fig. 4) collected after sulfidation at


593 K for NiMo6/Al2O3–Nb2O5(x) catalysts display two
main Raman active modes characteristic of the bulk or
25.7

24.6

multilayered MoS2, the E2g1 (in-plane vibration) and A1g


(%)

(out-of-plane vibration) [34]. It has been demonstrated that


the frequency difference between E2g1 and A1g modes can
be used as a robust and convenient preliminary diagnostic
of the layer thickness of MoS2 samples [34]. In this regard,
11.9
(%)

9.7

the Raman spectra of the catalysts independently of the


Mo 3d5/2–3d3/2

presence of Nb exhibit this doublet characteristic of MoS2


with Δω = 26 to Δω = 24, suggesting an increase in the
number of layers of MoS2 (around 3–4 layers) [34]. Figure
62.3

65.8
(%)

4 also shows that the E2g1 and A1g bands in the spectrum of
NiMo6/Al2O3–Nb2O5(x) with x = 0, 4, and 8 are broader
than the corresponding bands of NiMo6/Al2O3–Nb2O5(1),
NiMo6/Al2O3–

NiMo6/Al2O3–

which is likely induced by the growth of the MoS2 crystal


Nb2O5 (0)

Nb2O5 (8)

size.
Figure 5 shows the IR bands characteristics of the pyr-
Solid

idine adsorbed on Lewis acid sites (LAS) and Brønsted acid


Journal of Sol-Gel Science and Technology

sites (BAS) of the sulfided NiMo6/Al2O3–Nb2O5(x) [35]. more intense envelope in the FT–IR spectra of the OH
LAS sites are evidenced by a ν8a band at around 1610 cm−1 region for these catalysts (not shown). On the other hand,
and a ν19b band at around 1450 cm−1. Adsorption of Py on NiMo6/Al2O3+Nb(x) with x = 1, 4, and 8 wt% exhibit a
BAS leads to a ν8a band at around 1630 cm−1 and to a ν19b larger amount of LAS than NiMo6/Al2O3+Nb(0). The
band at around 1545 cm−1, and a peak at 1490 cm−1 is amount of LAS and total acidity was found to increase as
assignable to a mixture of LAS and BAS. Quantitative follows: NiMo6/Al2O3–Nb2O5(0) < NiMo6/Al2O3–Nb2O5
results are shown in Table 3, and the amount of Brønsted (1) < NiMo6/Al2O3–Nb2O5(4) < NiMo6/Al2O3–Nb2O5(8).
sites is high in NiMo6/Al2O3–Nb2O5(1) and NiMo6/Al2O3– This steady increase in acidity and consequent improved
Nb2O5–Nb2O5(4) catalysts, which is in agreement with the interaction with the NiMo6 precursor could explain the
slight augment of Mo and Ni concentrations with increasing
LAS
Nb content (Table 1).
LAS
3.5 Catalytic test

BAS Thiophene HDS activities (Table 3) are reported as con-


version percent after ~4.5 h of reaction time (steady state).
BAS
LAS + BAS No correlation was found between textural properties (sur-
face area and pore volume) and the conversion of thio-
phene. On the other hand, the amount of Nb had a large
NiMo6/Al2O3-Nb2O5(8) effect as shown in Table 3; hence, the overall activities of
the thiophene HDS (sulfided precursors) correlated well
with the Nb content, as follows: NiMo6/Al2O3–Nb2O5(0) =
41% < NiMo6/Al2O3–Nb2O5(1) = 47% < NiMo6/Al2O3–
Nb2O5(4) = 50% < NiMo6/Al2O3–Nb2O5(8) = 54%.
NiMo6/Al2O3-Nb2O5 (4)
The CO chemisorption values of the freshly sulfided
catalysts are included in Table 3. Clearly, a significant
correlation between thiophene conversion and the CO
chemisorption capacity is not observed. This means that the
differences in thiophene conversion observed for these
NiMo6/Al2O3-Nb2O5(1) catalysts are not due to increasing concentrations of surface
active sites, but to differences due to the nature of the sites
depending on the Nb content in the catalysts. One clear
difference seems to be the concentration of LAS (Table 3)
NiMo6/Al2O3-Nb2O5(0) which would be related to the adsorption of thiophene on or
close to the active sites. Another one would be the optimal
sulfiding of the surface, guaranteeing the formation of the
1700 1650 1600 1550 1500 1450 1400
NiMoS active phase as suggested from the XPS results.
wavenumber (cm-1) Thus, the results of catalytic activity obtained in this work
Fig. 5 FTIR spectra of pyridine adsorbed over sulfided NiMo6/Al2O3– may be correlated with the presence at the surface of
Nb2O5(x) catalyst varying the content of niobium (x = 0, 1, 4, and 8 wt molybdenum and nickel species, such as Mo4+ and NiMoS
%)

Table 3 Steady-state thiophene HDS activity, proportion of Lewis/Brönsted surface acid sites, and CO uptake of NiMo6/Al2O3–Nb2O5(x) catalyst
varying the content of niobium (x = 0, 1, 4, and 8 wt%)
Catalysts Activity × 106 (mol Th conv/g Cat min) Pyridine adsorption(μmol Py/ CO uptake
gCatalyst)
LAS BAS Total acidity μmol/gCatalyst μmol/m2

NiMo6/Al2O3–Nb2O5(0) 16.74 2.93 0.260 3.19 31.91 0.137


NiMo6/Al2O3–Nb2O5(1) 19.18 3.07 2.07 5.14 32.91 0.131
NiMo6/Al2O3–Nb2O5(4) 20.41 5.45 1.48 6.93 21.67 0.090
NiMo6/Al2O3–Nb2O5(8) 22.04 6.33 3.67 10.0 27.02 0.136
LAS Lewis acid sites, BAS Bronsted acid sites
Journal of Sol-Gel Science and Technology

Table 4 Steady-state
Catalysts Product distribution at 40% of DBT conversion DBT conversion (%) at
dibenzothiophene HDS
6h
conversion, product distribution, DDS HYD HYD/
and HYD/DDS ratio (at 40% of DDS
DBT conversion) of NiMo6/ BP (%) CHB (%) BCH (%) THDBT (%)
Al2O3–Nb2O5(x) catalyst
varying the content of niobium NiMo6/Al2O3– 29.6 3.15 0.10 1.46 0.16 71
(x = 0 and 8 wt%) Nb2O5(0)
NiMo6/Al2O3-Nb2O5 17.0 20.6 0.00 3.10 1.36 83
(8)

(Table 2). It is clear that the presence of Nb induces a MoS2 and crystal sizes of around 3–4 layers for the latter. It
greater catalytic activity, suggesting that the surface species was found that the overall thiophene HDS conversion par-
of NiMo6/Al2O3–Nb2O5(x) are important for the HDS allels the Nb content of the catalysts, but did not show any
reaction. correlation with the textural properties (surface area or pore
To further test the catalysts in more real conditions, high- volume) nor with the total number of sites, as measured by
pressure experiments of DBT HDS were carried out with CO chemisorption. Instead, a good correlation with the
the Nb-unpromoted catalyst and the promoted one showing amount of LAS, probably arising from the presence of Nb,
higher thiophene HDS conversion. Table 4 shows the DBT was found. Thus, the effect of Nb seems to be modifying
conversions obtained with NiMo6/Al2O3–Nb2O5(0) and the nature and not the number of active sites. Results of
NiMo6/Al2O3–Nb2O5(8) at 6 h of reaction time and the HDS of DBT confirmed that the catalyst containing Nb is
product distributions observed at 40% of DBT conversion. more active than the unpromoted one, and showed pre-
The HDS reaction occurs through two different routes: DDS ference for the HYD route against the DDS route observed
yielding BP and HYD that yields THDBT, CHB, and for the catalyst supported on alumina.
dicyclohexyl (DCH). Nb-containing catalysts proved to be
more active than catalysts without Nb and relative amounts Acknowledgements The authors would like to acknowledge financial
support to Universidad del Atlántico-Colombia (through Project
of the obtained products depended on the Nb presence. The
CB21-FGI2016 “7ma Convocatoria 2016”). E.P.P. dedicates this work
main desulfurized reaction products obtained were BP, to Ofelia Polo (R.I.P.).
CHB, and THDBT. The DDS product was formed in a
larger proportion for the NiMo6/Al2O3–Nb2O5(0) catalyst, Compliance with ethical standards
while the products of the HYD route (THDBT + CHB +
DCH) were predominantly formed with the NiMo6/Al2O3– Conflict of Interest The authors declare that they have no conflicts of
Nb2O5(8) catalyst, which is reflected in the HYD/DDS ratio. interest.
This result is in good agreement with previous literature
reports for Nb-modified NiMo/Al2O3 catalysts [19].
References
1. Srivastava VC (2012) RSC Adv 2:759–783
4 Conclusions 2. Hayyan M, Ibrahim MH, Hayyan A, AlNashef IM, Alakrach AM,
Hashim MA (2015) Ind Eng Chem Res 54:12263–12269
One-pot synthesis method was successful in introducing Nb 3. Zuo D, Vrinat M, Nie H, Maugé F, Shi Y, Lacroix M, Li D (2004)
into Al2O3 as a fast preparation way to supports for HDS Catal Today 93–95:751–760
4. Breysse M, Afanasiev P, Geantet C, Vrinat M (2003) Catal Today
catalysts. The characterization techniques demonstrated that
86:5–16
Nb presence influences the physicochemical properties of 5. Cabello CI, Botto IL, Thomas HJ (2000) Appl Catal A Gen
the NiMo-impregnated catalysts. Ni and Mo contents 197:79–86
slightly increased in Nb presence, but no crystalline phases 6. Debecker PD, Stoyanova M, Rodemerck U, Gaigneaux EM
(2011) J Mol Catal A Chem 340:65–76
such as the heteropolymolybdate precursor or Nb2O5 were
7. Ziolek M (2003) Catal Today 78:47–64
detected. XPS analysis showed for the sulfided NiMo6/ 8. Abdel-Rehim MA, Dos Santos ACB, Camorim VLL, Faro Jr. AC
Al2O3–Nb2O5(0) and NiMo6/Al2O3–Nb(8) similar species (2006) Appl Catal A Gen 305:211–218
of Mo and Ni on the surface; however, the S species type 9. Faro Jr AC, Grange P, dos Santos ACB (2002) Phys Chem Chem
Phys 4:3997–4002
and relative proportions of Mo, Ni, and S species depended 10. Sankar S, Vasudevan CNR, Rao J (1988) Phys Chem 92:1878–
on the presence of Nb, suggesting less active metal–support 1882
interaction that improves the sulfidation degree of Mo and 11. Gaborit V, Allali N, Geantet C, Breysse M, Vrinat M, Danot M
Ni. The Raman spectra for sulfided NiMo6/Al2O3–Nb2O5(x) (2000) Catal Today 57:267–273
suggest an increase with Nb of the S/Mo atomic ratio of
Journal of Sol-Gel Science and Technology

12. Altamirano E, de los Reyes JA, Murrieta F, Vrinat M (2008) Catal 24. Crépeau G, Montouillout V, Vimont A, Mariey L, Cseri T, Maugé
Today 133–135:292–298 F (2006) J Phys Chem B 110:15172–15185
13. Polo A, Puello-Polo E, Diaz-Uribe C (2017) Prospectiva 15 25. Infantes-Molina A, Romero-Pérez A, Finocchio E, Busca G,
(1):74–82 Jiménez-López A, Rodríguez-Castellón E (2013) J Catal 305:101–
14. Palcheva R, Kaluza L, Spojakina A, Jiratova K, Tyuliev G (2012) 117
Chin J Catal 33:952–961 26. Froment GF, Bischoff KB (1990) Chemical reactor analysis and
15. Weissman JG (1996) Catal Today 28:159–166 design, 2nd ed. Wiley, New York, NY, pp 61–197
16. Rocha AS, Faro Jr AC, Oliviero L, Van Gestel J, Maugé F (2007) 27. Juang R-S, Wu F-C, Tseng R-L (2002) Coll Surf A 201:191–199
J Catal 252:321–334 28. Sampieri A, Pronier S, Brunet S, Carrier X, Louis C, Blanchard
17. Palcheva R, Kaluza L, Dimitrov L, Tyuliev G, Avdeev G, Jirátová J, Fajerwerg K, Breysse M (2010) Micro Mesop Mater 130:
K, Spojakina A (2016) Appl Catal A Gen 520:24–34 130–141
18. Cedeño L, Hernandez D, Klimova T, Ramirez J (2003) Appl Catal 29. Th. Weber JC, Muijsers JHMC, van Wolput CPJ, Verhagen JW
A Gen 241:39–50 (1996) Niemantsverdriet. J Phys Chem 100:14144–14150
19. Méndez FJ, Franco-López OE, Bokhimi X, Solís-Casados DA, 30. Aigler J, Brito JL, Leach PA, Houalla M, Proctor A, Cooper NJ,
Escobar-Alarcón L, Klimova TE (2017) Appl Catal B Env Hall WK, Hercules DM (1993) J Phys Chem 97:5699–5702
219:479–491 31. Wang X, Ozkan US (2005) J Phys Chem B 109:1882–1890
20. Lai W, Pang L, Zheng J, Li J, Wu Z, Yi X, Fang W, Jia L (2013) 32. Shi C, Xiang K, Zhu Y, Chen X, Zhou W, Chen H (2017)
Fuel Proc Technol 110:8–16 Electrochim Acta 246:1088–1096
21. Ayala-G M, Puello-Polo E, Quintana P, González-García G, Diaz 33. Roukoss C, Laurenti D, Devers E, Marchand K, Massin L, Vrinat
C (2015) RSC Adv 5:102652–102662 M (2009) C R Chimie 12:683–691
22. Barrett EP, Joyner LG, Halenda PP (1951) J Am Chem Soc 34. Lee C, Yan H, Brus LE, Heinz TF, Hone J, Ryu S (2010) ACS
73:373–380 Nano 4:2695–2700
23. Sing KSW, Everett DH, Haul RAW, Moscou L, Pierotti RA, 35. Kung MC, Kung HH (1985) Catal Rev Sci Eng 27:425–460
Rouquerol J, Siemieniewska T (1985) Pure Appl Chem 57:603–619

You might also like