You are on page 1of 9

Fuel 129 (2014) 138146

Contents lists available at ScienceDirect

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

Synthesis, characterization and hydrodesulfurization properties


of nickelcoppermolybdenum catalysts for the production
of ultra-low sulfur diesel
Huan Liu, Changlong Yin , He Li, Bin Liu, Xuehui Li, Yongming Chai, Yanpeng Li, Chenguang Liu
State Key Laboratory of Heavy Oil Processing, Key Laboratory of Catalysis, CNPC, China University of Petroleum, Qingdao, Shandong 266555, China

h i g h l i g h t s
 A series of nickelcoppermolybdenum catalyst precursors was synthesized by chemical precipitation.
 The particle sizes of the MoS2 nanoclusters and sulded nickel-containing compounds varied with the introduction of copper.
 A Ni9.5Cu0.5Mo10 catalyst exhibited high hydrodesulfurization activity.

a r t i c l e

i n f o

Article history:
Received 17 February 2014
Received in revised form 18 March 2014
Accepted 25 March 2014
Available online 5 April 2014
Keywords:
Nickelcoppermolybdenum catalyst
Hydrodesulfurization
4,6-Dimethyldibenzothiophene
Ultra-low sulfur diesel
Chemical precipitation

a b s t r a c t
A series of nickelcoppermolybdenum catalyst precursors was synthesized through chemical precipitation method, and highly loaded catalysts were prepared by mixing the precursors with alumina sol. The
nickelcoppermolybdenum catalysts were tested in the hydrodesulfurization (HDS) of the model
compound 4,6-dimethyldibenzothiophene (4,6-DMDBT) and uid catalytic cracking (FCC) diesel. The catalysts were characterized by means of N2-physisorption, XRD, FT-IR, SEM, TGDSCMS, TPR and HRTEM.
The characterizations results indicated that nickelcoppermolybdenum catalyst precursors could be
synthesized due to the similarities of the lattice parameters and ionic radii of nickel and copper, and
the HRTEM micrographs and XRD results of the sulded nickelcoppermolybdenum catalysts revealed
that the introduction of copper affected the particle sizes of the stacked MoS2 and sulded nickelcontaining compounds. The catalytic results showed that the Ni9.5Cu0.5Mo10 catalyst exhibited high
catalytic activity in the HDS of 4,6-DMDBT and FCC diesel, and might be suitable for the production of
ultra-low sulfur diesel.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Hydrotreating is one of the most important reactions in petroleum rening processes for the removal of heteroatoms from
feedstocks. Especially the removal of sulfur from sulfur-containing
compounds, the contents of which typically ranges between 1 and
3 wt.% (10,00030,000 ppmw) in middle distillates or raw distillates, is of importance [1,2]. The sulfur-containing compounds
cause several severe problems, such as the pollution of the environment through the emission of SO2, the corrosion of materials
by acid rain, and the poisoning of most petrochemical catalysts
[24]. In recent years, more stringent environmental legislations
and growing concerns about the environment, coupled with the
Corresponding author. Tel.: +86 053286984686.
E-mail addresses: yincl@upc.edu.cn (C. Yin), cgliu.upc.edu.cn@gmail.com
(C. Liu).
http://dx.doi.org/10.1016/j.fuel.2014.03.055
0016-2361/ 2014 Elsevier Ltd. All rights reserved.

dwindling fossil fuel reserves, have accelerated research of economically effective hydrodesulfurization (HDS) catalysts for producing cleaner fuels from feedstocks of progressively lower
quality [2]. The industrial HDS process is carried out over sulded
Mo or W-based catalysts promoted by Ni or Co and supported on
c-Al2O3 [3,57]. To lower the sulfur content of fuels to 10
15 ppmw, the urgent need for a drastic improvement of the HDS
catalyst activity motivated the search for novel and more efcient
catalysts [1,8].
Signicant and promising improvements in HDS catalysts have
been achieved for the production of ultra-low sulfur diesel (ULSD).
The introduction of organic additives during the preparation of
HDS catalysts, like ethyleneglycol and triethyleneglycol [9,10],
might promote the formation of the so-called NiMoS phase or
CoMoS phase, which are supposed to be catalytically active in
HDS [5,11]. Pt and Rh suldes were more active than MoS2 due
to electronic effects [12,13], and were more suitable for the

139

H. Liu et al. / Fuel 129 (2014) 138146

production of ULSD, because of the steric effects of the refractory


sulfur-containing molecules, like dibenzothiophenes (DBTs) alkylated at the 4 and 4,6 positions. The concentration of these molecules
can predominantly be reduced through hydrogenation [1,12,14].
Additives like uoride or phosphate are also used to prepare highly
active supported HDS catalysts. They modify and optimize the interactions between the active phase and support [8,9,15,16].
Copper-containing compounds are known to be catalytically
active for hydrogenation processes such as methanol synthesis
and selective hydrogenation [1719], although few papers deal
with copper in HDS catalysts. Harris and Chianelli found that the
introduction of Cu in unsupported MoS2 (through the reaction of
CuCl2 with (NH4)2MoS4) acted as a poison, because Cu decreased
the electron density on Mo, thus leading to oxidization of Mo in
MoS2 [20]. Recently, the formation of mixed-metal CuMoS
type structures shaped as single-layer hexagonally truncated triangular MoS2-like nanoclusters was revealed by Kibsgaard et al. [5].
Also, Cu+ ions are efcient adsorbents for the removal of sulfur,
because they can form p-complexes with thiophenic aromatic
rings, as in DBT and its derivates [21,22].
More recently, we reported the synthesis of novel crystalline
ammonium nickel molybdate as efcient HDS catalyst precursor
towards the production of ULSD, which was attributed to the efcient formation of the NiMoS structure [23]. As the lattice
parameters as well as the ionic radii of nickel and copper are very
close [24], it can be expected that nickelcoppermolybdenum
compounds could be prepared, which might have novel physicochemical and catalytic properties. To explore if such compounds
are efcient HDS catalysts, a series of nickelcoppermolybdenum
catalysts was synthesized by chemical precipitation and evaluated
in the HDS of 4,6-dimethyldibenzothiophene (4,6-DMDBT), and
uid catalytic cracking (FCC) diesel.

N2 adsorptiondesorption experiments were carried out on a


ChemBET 3000 (Quantachrome, USA) instrument. X-ray powder
diffraction (XRD) characterization was performed on a Rigaku D/
max-IIA diffractometer using graphite-ltered Cu Ka radiation.
Combined thermogravimetry, differential scanning calorimetry,
and mass spectrometry analysis (TGDSCMS) was carried out at
a heating rate of 10 oC/min under high-purity nitrogen atmosphere. Fourier-transform infrared (FT-IR) spectra were collected
on a Nexus spectrometer (Nicolet, USA) using KBr disks. Morphological information was collected on a FEI Quanta200 scanning
electron microscope (SEM) instrument. Temperature-programmed
reduction (TPR) analysis of nickelcoppermolybdenum catalyst
precursors and sulded catalysts was performed on a Micromeritics RS 232 instrument. The H2 consumption of the catalyst precursor or sulded catalyst reduction was recorded by a thermal
conductivity detector (TCD). High-resolution transmission electron
microscopy (HRTEM) was carried out on a JEM 2100 microscope
operated at 200 kV. More detailed information about the characterization of the nickelcoppermolybdenum catalyst precursors
can be found in Ref. [23].

2. Experimental

2.4. Catalytic activity

2.1. Preparation of nickelcoppermolybdenum catalyst precursors

2.4.1. HDS of 4,6-DMDBT using sulded catalyst precursors


Catalytic activity measurements were carried out in a 500 mL
micro-autoclave reactor. 4,6-DMDBT (purchased from Aladdin
Co., purity 97%), was chosen as representative component for the
HDS reaction. The sulded catalyst was made from the precursor
in the following way. The reactor was charged with 2.0 g nickel
coppermolybdenum catalyst precursor of 400460 mesh size
and 1.0 mL CS2 in petroleum ether solvent (boiling ranging 90
120 C). Before the suldation reaction, the reactor was ushed ve
times with high-purity hydrogen followed by adjusting the hydrogen pressure to 3.0 MPa at room temperature. The reactor was then
heated at a ramping rate of 3 C/min, and stirred at about 60 r/min,
so that the catalytic reaction was inhibited until reaching the
desired reaction conditions. When the temperature was approaching the evaluated temperature (320 C for suldation), the stirring
rate was set as 700 r/min. At that time, the catalytic reaction began.
After 6 h suldation, the reactor was opened at room temperature
and atmospheric pressure, then 0.10 g 4,6-DMDBT was added to
the reactor, and the hydrodesulfurization reactions were carried
out at 8.0 MPa and 315 C during 3 h. Then the products were
cooled to room temperature and analyzed with a Varian 3800

Nickelcoppermolybdenum catalyst precursors were prepared


by the chemical precipitation method in the following way. An
amount of 0.10 mol ammonium heptamolybdate and the required
amount of nickel nitrate and/or copper nitrate (Sinopharm Chemical Reagent Company, China, Grade AR) were dissolved in deionized water with the desired molar ratio (shown in Table 1). The
solution was added to a ask, heated, and stirred at 90 C. The addition of aqueous ammonia precipitated a light green solid that dissolved in excess ammonia, giving a solution with various colors
(shown in Table 1 as Color of dissolved precipitation) depending
on the composition and molar ratio. The solution was stirred and
heated for 12 h, leading to the formation of a precipitate. The products were isolated by vacuum ltration, washed with deionized
water, and dried overnight at 80 C and atmospheric pressure.
2.2. Preparation of highly loaded catalyst
The preparation of a highly loaded catalyst has been described
in detail elsewhere [23]. First, an alumina sol was made from

boehmite and nitric acid. Then the nickelcoppermolybdenum


catalyst precursor was mixed with the alumina sol with the weight
ratio of precursor/alumina 75/25, and extruded into rods 1.2 mm
in diameter. The rods were dried at 100 C for 12 h and then calcined at 350 C for 4 h.
An industrial NiMo/Al2O3 catalyst containing 24 wt.% MoO3 and
4 wt.% NiO supported on alumina (BET specic surface area of
225 m2/g) was used as reference catalyst in the HDS of FCC diesel.
2.3. Characterization of nickelcoppermolybdenum catalyst
precursors

Table 1
Experimental conditions and textural properties.
Label

Ni/mol

Cu/mol

Color of dissolved precipitate

BET surface area (m2/g)

Pore volume (cm3/g)

Pore size (nm)

Cu10Mo10
Ni5Cu5Mo10
Ni8Cu2Mo10
Ni9.5Cu0.5Mo10
Ni10Mo10

0
0.05
0.08
0.095
0.10

0.10
0.05
0.02
0.005
0

Dark blue
Blackish green
Dark blue
Dark blue
Dark blue

9
38
53
28
88

0.07
0.10
0.10
0.09
0.13

37.3
7.7
5.4
8.6
4.5

140

H. Liu et al. / Fuel 129 (2014) 138146

gas chromatograph equipped with ame ionization detector (GCFID) and a 50 m OV101 capillary column, coupled with a quadrupole mass spectrometer (Finnigan SSQ710) (GCMS).
2.4.2. HDS of FCC diesel on highly loaded catalyst
HDS of uid catalytic cracking (FCC) diesel (Sinopec Qingdao
Rening & Chemical, Co., LTD.) was carried out in a high pressure
xed-bed down-ow micro-reactor with an inner diameter of
10 mm, 40 cm in length. Ten milliliter of the highly loaded catalyst
was placed in the center of the reactor. The catalyst was rst treated at 110 C for 1 h and then at 330 C for 12 h by a liquid stream
containing 3.0 wt.% CS2 in cyclohexane. The reaction was carried
out at a pressure of 6.0 MPa, a H2/feed ratio of 600, a temperature
of 350 C, and a liquid hourly space velocity (LHSV) of 2.0 h1. The
sulfur compositions of the liquid products were analyzed using a
Varian 3800 gas chromatograph equipped with a pulse ame photometric detector (GC-PFPD) and a 30 m fused-silica capillary
column.
3. Results and discussion
3.1. Characterization of nickelcoppermolybdenum catalyst
precursors
3.1.1. Textural properties
The textural properties (BET surface area, pore volume, and pore
size) of the nickelcoppermolybdenum catalyst precursors are

reported in Table 1. The Cu10Mo10 catalyst precursor exhibits a


lower BET surface area than the other precursors, while the
Ni10Mo10 catalyst precursor has the highest BET surface area and
pore volume (88 m2/g and 0.13 cm3/g, respectively). The N2 adsorptiondesorption isotherms of the Ni10Mo10, Cu10Mo10,
Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalyst precursors are depicted
in Fig. 1. According to the International Union of Pure and Applied
Chemistry (IUPAC) classication, the N2 adsorptiondesorption isotherm of the Ni10Mo10 catalyst precursor is of type IV, typical for
mesoporous materials [3,16]. The isotherm of the Cu10Mo10 catalyst precursor is of type III [25]. The isotherms of the Ni5Cu5Mo10
and Ni9.5Cu0.5Mo10 catalyst precursors are conrmed as type IV
and the hysteresis loop of the isotherm ios of H2 type, which is
ascribed to randomly folded sheets with inkbottle pore shape [16].
3.1.2. XRD
The nickelcoppermolybdenum catalyst precursors were characterized by XRD. Fig. 2 shows that the diffraction peaks of the
nickelcoppermolybdenum catalyst precursors are approximately
identical, with 2h peaks detected at 12.1, 17.4, 23.5, 26.5, 29.6
and 33.4. These peaks belong to the structure of ammonium nickel
molybdate (NH4)HNi2(OH)2(MoO4)2 (PDF card No. 50-1414)
[23,26]. Copper can replace nickel in the nickelmolybdenum compound, due to the similarities of their lattice parameters and ionic
radii, and therefore homogeneous nickelcoppermolybdenum
catalyst precursors might be formed, as suggested by the fact that
all XRD patterns are nearly the same [24].

Fig. 1. N2 adsorptiondesorption isotherms of the Ni10Mo10, Cu10Mo10, Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalyst precursors.

H. Liu et al. / Fuel 129 (2014) 138146

Fig. 2. XRD patterns of the nickelcoppermolybdenum catalyst precursors.

3.1.3. FT-IR
The FT-IR spectra of the Cu10Mo10, Ni5Cu5Mo10,
Ni9.5Cu0.5Mo10, and Ni10Mo10 catalyst precursors are shown in
Fig. 3, they are nearly the same. The adsorption peaks from 3300
to 3050 cm1 are attributed to the m3 NH asymmetric stretching
vibration [26], and those from 2000 to 1600 cm1 to the torsional
vibration of NH+4. The FT-IR peak at 1410 cm1 is due to the HNH
deformation vibration [26], and the peaks between 1000 and
700 cm1 with different intensities are assigned to the stretching
vibrations of bridging oxygen in MoOMo [27]. Therefore, it can
be concluded that molybdate and ammonium ions exist in the
nickelcoppermolybdenum catalyst precursors.
3.1.4. SEM
The SEM micrographs of the Ni10Mo10, Cu10Mo10,
Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalyst precursors are presented in Fig. 4. The four catalyst precursors exhibit various morphologies. The Ni10Mo10 catalyst precursor contains irregular
stacked particles with sizes ranging from 200 to 1000 nm, whereas
the Cu10Mo10 and Ni5Cu5Mo10 catalyst precursors form smaller
particles, with particle size around 50 and 30 nm, respectively.
The Ni9.5Cu0.5Mo10 catalyst precursor displays disorderly stacked
particles with a size of 100300 nm.

141

3.1.5. TGDSCMS
TGDSCMS curves of the Ni10Mo10, Cu10Mo10, Ni5Cu5Mo10,
and Ni9.5Cu0.5Mo10 catalyst precursors are shown in Fig. 5. The
curves of the Ni10Mo10 catalyst precursor are similar to the curves
shown in a previous paper [23], showing an endothermic peak at
407 C (DSC curve), and a total weight loss between 350 and
500 C of about 11.0% (TG curve). This is expected for the loss of
NH3 and H2O (10.8%) [23]. The entire thermal effect is accompanied by the release of H2O (m/z = 18, not shown here) and NH+3
(m/z = 17) that is detected by MS analysis at 411 C. The Cu10Mo10
catalyst precursor shows one endothermic peak at 225 C. This
indicates that it has poor thermostability and possibly becomes
sintered after calcination at higher temperature. The total weight
loss is about 11.6% from 195 to 265 C. NH3 and H2O are also
detected by MS analysis for the Cu10Mo10 catalyst precursor.
The TGDSCMS curves of the Ni5Cu5Mo10 catalyst precursor
are different from the curves of the Ni10Mo10 and Cu10Mo10 catalyst precursors. The total weight loss ratio from 280 to 365 C is
similar (11.3%), but four main endothermic peaks are observed at
305, 342, 522 and 545 C. The intensity of the endothermic peak
at 305 C is higher than that at 342 C, with weight losses of 6.8%
and 4.5%, respectively. The two higher temperature peaks without
any release of H2O or NH3 need further investigated. The
Ni9.5Cu0.5Mo10 catalyst precursor exhibits one endothermic peak
at 280 C, which is in between the ones detected for the Cu10Mo10
and Ni10Mo10 catalyst precursors. The weight loss from 270 to
306 C is 11.2% (TG curve), accompanied by the release of NH3
and H2O (DSC curve). Higher temperature endothermic peaks at
517 and 542 C without release of NH3 or H2O are also detected
by DSC, which is similar to that of Ni5Cu5Mo10 catalyst precursor.
3.1.6. TPR
The reducibility of the nickelcoppermolybdenum catalyst
precursors was studied by TPR and the proles are presented in
Fig. 6. The Cu10Mo10 catalyst precursor exhibits one principal
reduction peak at 420 C, which might be attributed to the reduction of Mo6+ species to Mo4+ [3,8], and the reduction peak ends
before 520 C, indicating incomplete reduction of the Mo species
[3,8,25]. The major peak at 420 C is preceded by two lower hydrogen consumption peaks centered at 325 and 360 C, due to the
reduction of Cu2+ species to Cu+, and Cu+ to Cu0, respectively
[28]. One major broad peak around 500 C is detected for the
Ni10Mo10 catalyst precursor due to the reduction of Mo species,
and the reduction of Ni species is conrmed at about 350 C [25].
For the Ni8Cu2Mo10 and Ni5Cu5Mo10 catalyst precursors, the
TPR peak detected around 375 C might be attributed to the reduction of a mixed CuxNi1xO phase [29], followed by hydrogen consumption originating from the reduction of Mo species in the
range 420590 C. For the Ni9.5Cu0.5Mo10 catalyst precursor,
three hydrogen consumption peaks are located at 340, 430 and
540 C. The rst broad peak is attributed to the reduction processes
of Ni or Cu species, similar to the reduction in the Ni10Mo10 and
Cu10Mo10 catalyst precursors [25,28], and the last two peaks are
due to the reduction of Mo species [25]. Compared with the
Ni10Mo10 catalyst precursor, the reduction of Mo6+ species in
Ni9.5Cu0.5Mo10 precursor shifts to lower temperature at 430 C,
possibly due to the formation of the reduced Ni or Cu species.
3.2. Characterizations of sulded nickelcoppermolybdenum
catalysts

Fig. 3. FT-IR spectra of the Cu10Mo10, Ni5Cu5Mo10, Ni9.5Cu0.5Mo10, and


Ni10Mo10 catalyst precursors.

3.2.1. HRTEM of sulded catalysts


Typical HRTEM micrographs of the sulded catalysts are presented in Fig. 7. The typical stacked MoS2 slabs characterized by
black thread-like fringes are observed in the sulded nickel
coppermolybdenum catalysts, and the interplanar distance

142

H. Liu et al. / Fuel 129 (2014) 138146

Fig. 4. SEM photographs of the Ni10Mo10, Cu10Mo10, Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalyst precursors.

Fig. 5. TGDSCMS curves of the Ni10Mo10, Cu10Mo10, Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalyst precursors.

(about 0.61 nm) of the (0 0 2) basal planes of the crystalline MoS2 is


conrmed. Some smaller fringes with a spacing of about 0.29 nm
are also detected for the Ni10Mo10 catalyst, which might be

attributed to crystalline sulded nickel species [23]. The number


of stacked layers and the slab lengths of MoS2 are different for
the samples shown in Fig. 7. The stacking number of MoS2

H. Liu et al. / Fuel 129 (2014) 138146

143

A separate sulded Ni compound is conrmed in the Ni10Mo10


catalyst, with the sharp diffraction peaks of Ni3S2 (PDF card No. 441418) at 21.8, 31.1, 37.8, 44.3, 50.1 and 55.2. Weak Ni3S2
peaks at 21.8, 31.1, 44.3 and 55.2 are also observed in the
Ni9.5Cu0.5Mo10 catalyst, but with much lower intensities. Narrow
sharp diffraction peaks at 30.1, 32.3, 34.6, 35.7, 45.9, 48.9 and
53.4 in the Ni5Cu5Mo10 catalyst are detected and attributed to
various NiS phases (PDF card No. 75-0612 and No. 75-0613).

Fig. 6. TPR proles of the nickelcoppermolybdenum catalyst precursors.

decreases in the order Ni9.5Cu0.5Mo10 > Ni5Cu5Mo10 >


Ni10Mo10 > Cu10Mo10. The slab length of the Ni9.5Cu0.5Mo10
catalyst is in the range of 825 nm, longer than the other catalysts.
3.2.2. XRD of sulded catalysts
The nickelcoppermolybdenum catalysts were characterized
by XRD after suldation and the results are exhibited in Fig. 8.
The characteristic XRD peaks attributed to MoS2 at 14.4, 33.0,
38.2 and 58.2 are detected on the sulded nickelcoppermolybdenum catalysts with different intensities. The intensities of the
MoS2 peaks of the Ni9.5Cu0.5Mo10 catalyst are higher than those
of the other sulded catalysts, especially the intensity of the
14.4 peak is higher. This might be due to a higher stacking of
the MoS2 slabs, in accordance with the HRTEM results. For the
Cu10Mo10 catalyst, the XRD peaks of CuS phase (PDF card No.
03-1090) at 29.6, 31.7, 32.9, 47.8 and 59.2 are detected.

3.2.3. TPR of sulded catalysts


The TPR proles of sulded nickelcoppermolybdenum catalysts are depicted in Fig. 9. There exist distinguished differences
of the TPR patterns among the sulded catalysts. The hydrogen
consumption peak around 275 C in Ni10Mo10 catalyst might be
due to the reduction of nickel sulde located at the edges of
MoS2 slabs forming the NiMoS active phase [30], and the hydrogen consumption peak centered at 443 C is attributed to reduction
of MoS2 nanoclusters [30,31]. The TPR pattern of Ni9.5Cu0.5Mo10
catalyst is similar to that of Ni10Mo10 catalyst, but the rst hydrogen consumption peak shifts to lower temperature region at
235 C. This is also detected over Ni5Cu5Mo10 catalyst, and the
sharp peak at 504 C might be ascribed to the reduction of sulded
nickel compounds [32,33]. The reduction peak owing to MoS2
nanoclusters over Ni5Cu5Mo10 catalyst is about 20 C lower than
those of Ni9.5Cu0.5Mo10 and Ni10Mo10 catalysts, and the relative
intensity of hydrogen consumption is also lower. As for Cu10Mo10
catalyst, the hydrogen consumption peak at 275 C is due to the
reduction of sulded copper-containing compounds [32]. Due to
the limitations of the instruments and the common temperature
used for the production of ULSD (temperature generally no more
than 400 C), the TPR data were collected before 650 C. Other
reports cover the complete reduction of sulded Mo-containing
compounds ended centered at 10001200 C [3133].
3.3. Catalytic activity of the sulded catalysts
3.3.1. Catalytic evaluation of 4,6-DMDBT
The results of the HDS reactions of 4,6-DMDBT over the nickel
coppermolybdenum catalysts are shown in Table 2. Five main

Fig. 7. HRTEM micrographs of the sulded Ni10Mo10, Cu10Mo10, Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalysts.

144

H. Liu et al. / Fuel 129 (2014) 138146

more desulfurization products detected (69.6%) than over the other


catalysts. The DDS selectivity of the nickelcoppermolybdenum
catalysts decreased in the order: Cu10Mo10 > Ni10Mo10 > Ni8Cu2Mo10 > Ni5Cu5Mo10 > Ni9.5Cu0.5Mo10. The high SDDS of
Cu10Mo10 catalyst might be attributed to the hydrogenation ability of MoS2 [37,38], but the Cu10Mo10 catalyst exhibited poor
desulfurization performance due to the lack of synergetic effect .

Fig. 8. XRD patterns of the sulded Cu10Mo10, Ni5Cu5Mo10, Ni9.5Cu0.5Mo10, and


Ni10Mo10 catalysts.

Fig. 9. TPR of the sulded Ni10Mo10, Ni9.5Cu0.5Mo10, Ni5Cu5Mo10, and


Cu10Mo10 catalysts.

products
were
detected,
including
tetrahydro-4,6-dimethyldibenzothiophene (TH-DMDBT) and hexahydro-4,6-dimethyldibenzothiophene
(HH-DMDBT),
the
hydrogenated
intermediates of 4,6-DMDBT in the HDS reactions, 3,30 -dimethylbiphenyl (DMBP) produced through the direct desulfurization route
(DDS), and methylcyclohexyltoluene (MCHT) and 3,30 -dimethylbicyclohexyl (DMBCH), the hydrogenated products of the
hydrogenation route (HYD) [14,3436]. The selectivity of the DDS
route (SDDS) was calculated as DMBP/(100-4,6-DMDBT), and the
results are exhibited in Table 2.
Table 2 shows that the Cu10Mo10 catalyst exhibited poor HDS
ability towards 4,6-DMDBT, with a conversion of only 1.1%. The
detected desulfurization product was very low (0.3%). The
Ni10Mo10 catalyst showed higher hydrogenation ability. More
partially hydrogenated intermediates TH-DMDBT and HH-DMDBT
were produced on the Ni10Mo10 catalyst than on the other catalysts. The Ni9.5Cu0.5Mo10 catalyst exhibited high HDS ability
towards 4,6-DMDBT, with a conversion of 80.3%, and there were

3.3.2. HDS of FCC diesel on highly loaded catalyst


The properties of the FCC diesel used in this work are shown in
Table 3, and the GC-PFPD chromatography results of the FCC diesel
(partially the refractory sulfur compounds) before and after HDS
reactions are shown in Fig. 10 with arrows to indicate where the
peaks of 4-methyldibenzothiophene (4-MDBT), 4,6-DMDBT and
2,4,6-trimethyldibenzothiophene (2,4,6-TMDBT) are. These molecules are often used as standard samples to determine typical sulfur compounds. The sulfur content of the FCC diesel was
10,637 ppmw, and after HDS the sulfur content decreased by more
than 99.9% (to 9.6 ppmw) on the highly loaded Ni9.5Cu0.5Mo10
catalyst, while the sulfur content was 105 ppmw on the reference
catalyst. This demonstrates the excellent HDS activity of the highly
loaded Ni9.5Cu0.5Mo10 catalyst. Fig. 10 shows that the refractory
DBTs, especially DBT alkyl-substituted at the 4 and 6 positions,
such as 4,6-DMDBT and 2,4,6-TMDBT, were detected after HDS
reaction, even on the highly loaded Ni9.5Cu0.5Mo10 catalyst.
The synthesis of nickelcoppermolybdenum catalyst precursors through chemical precipitation leads to the formation of a
solid solution [23,26], which might be attributed to the similarities
of the lattice parameters and ionic radii of nickel and copper, so
that copper can replace nickel in the structure [24]. The TGDSC
MS results indicate that the interactions between binary
Cu10Mo10 and Ni10Mo10 catalyst precursors and ternary
Ni9.5Cu0.5Mo10 catalyst precursor are distinct. The decomposition of the structure, with a loss of weight and the release of H2O
and NH3, occurred in one step for the Ni10Mo10, Cu10Mo10, and
Ni9.5Cu0.5Mo10 catalyst precursors at 407, 225 and 280 C,
respectively. This indicates that the introduction of copper into
the nickelmolybdenum catalyst precursor changed the interactions among the nickelcoppermolybdenum compounds. Two
similar endothermic peaks at higher temperatures without any
release of NH3 or H2O were also detected over the Ni9.5Cu0.5Mo10
and Ni5Cu5Mo10 catalyst precursors, implying that at higher temperature the complex nickelcoppermolybdenum compounds
might undergo some analogous physicochemical changes. Therefore the formation of nickelcoppermolybdenum catalyst precursors could also be conrmed through the TGDSCMS results. The
TPR results of oxidic nickelcoppermolybdenum catalyst precursors indicated that the introduction of copper affected the reducibility of these precursors. As for Ni9.5Cu0.5Mo10 catalyst
precursor, the hydrogen consumption due to the reduction of
Mo6+ species shifted to lower temperature compared to that of
Ni10Mo10 catalyst [25].
The SEM micrographs of the Ni10Mo10, Cu10Mo10,
Ni5Cu5Mo10, and Ni9.5Cu0.5Mo10 catalyst precursors indicate
that the aggregated particles vary in size. The size of the Ni10Mo10
catalyst precursor is about 2001000 nm, and for the Cu10Mo10
and Ni5Cu5Mo10 catalyst precursors the sizes are centered at 50
and 30 nm, respectively, while the size of the Ni9.5Cu0.5Mo10 catalyst precursor is in the range of 100300 nm. The HRTEM results
showed distinct differences among the number of the stacked layers and slab lengths of MoS2 in the sulded nickelcopper
molybdenum catalysts. Crystalline sulded nickel species could
also be detected in the Ni10Mo10 catalyst (Fig. 7). This is in
accordance with XRD characterization results of the sulded catalysts. From Fig. 8, sulded nickel and copper species could be
detected in the Ni10Mo10 and Cu10Mo10 catalysts, respectively.

145

H. Liu et al. / Fuel 129 (2014) 138146


Table 2
Product distribution of 4,6-DMDBT catalytic evaluation.
Catalyst

4,6-DMDBT

TH-DMDBT

HH-DMDBT

DMBP

MCHT

DMBCH

SDDS,%

Cu10Mo10
Ni5Cu5Mo10
Ni8Cu2Mo10
Ni9.5Cu0.5Mo10
Ni10Mo10

98.9
80.0
60.8
19.7
31.1

0.7
4.2
5.4
3.8
5.2

0.1
2.4
3.9
6.9
7.7

0.2
7.3
16.7
37.1
21.3

0.1
4.3
9.6
17.9
12.0

0
1.8
3.6
14.6
22.7

18
36
43
46
31

Table 3
Properties of the FCC diesel.
Density (g/mL)

0.8809
a
b

Total sulfur (ppm)

10,637

Boiling range (C)


IBPa

10

30

50

70

90

FBPb

231

261

285

298

319

350

377

IBP: initial boiling point.


FBP: nal boiling point.

Fig. 10. GC-PFPD chromatography of the sulfur distribution in FCC diesel before and
after HDS reaction.

The Ni9.5Cu0.5Mo10 catalyst showed distinct characteristic peaks


attributed to the MoS2 phase and some weak diffraction peaks
owing to the Ni3S2 phase. The TPR proles of sulded nickel
coppermolybdenum catalysts exhibited that due to the introduction of copper, the reduction of nickel sulde, which was supposed
to be located at the edges of MoS2 slabs forming the NiMoS
active phase, shifted to lower temperature region with comparison
of Ni10Mo10 and Cu10Mo10 catalysts [30].
The catalytic results show that the Ni9.5Cu0.5Mo10 catalyst
exhibits high HDS activity both in the conversion of 4,6-DMDBT
and in that of FCC diesel. The 4,6-DMDBT conversion of the
Ni9.5Cu0.5Mo10 catalyst reached 80.3%, and the desulfurized
product was 69.6%, higher than those over the other catalysts.
The DDS selectivity over the Ni9.5Cu0.5Mo10 catalyst reached
46%, higher than those over the other catalysts except the
Cu10Mo10 catalyst, which could only convert 1.1% 4,6-DMDBT.
This indicates that the direct desulfurization route was preferred
over the Ni9.5Cu0.5Mo10 catalyst, possibly due to the inhibition
effect caused by the formed and accumulated H2S [37]. These
results are in accordance with the HRTEM and XRD results of sulded catalysts. The highly stacked MoS2 slabs together with much
smaller nickel-containing compound on Ni9.5Cu0.5Mo10 catalyst
might help fulll the synergetic effect.
The Rim-Edge model developed by Chianelli and coworkers
[39] explains the selectivity to DBT by geometrical considerations.
The stacked MoS2 particles are supposed to possess two catalytic

sites, one called Rim sites, which are made of external stacked
MoS2 layers and catalytically active both in hydrogenation (HYD)
and direct desulfurization (DDS) reactions. The others are called
Edge sites, which are only present on the surface of internal
stacked MoS2 layers and are only active in the DDS reaction. Furthermore, the formation of the so-called NiMoS phase has
been revealed by Topse and coworkers with scanning tunneling
microscopy (STM), where nickel substitutes Mo atoms at the edge
sites of single-layer MoS2 nanoclusters, and exhibits distinct catalytic activity [5,7,11]. Therefore the formation of aggregated sulded nickel species would not be desired for HDS catalysts, possible
because of the lack of synergetic effect. Stacked MoS2 particles
with layer number in the range of 717 and smaller nickelcontaining compound have been detected on the Ni9.5Cu0.5Mo10
catalyst, which would possess more HDS active sites. The FCC
diesel hydrodesulfurization results also conrm that the
Ni9.5Cu0.5Mo10 catalyst exhibits high activity, which could
decrease the sulfur content from 10,637 to 9.6 ppmw in hydrogenated products.
4. Conclusions
Homogeneous nickelcoppermolybdenum catalyst precursors
were synthesized by chemical precipitation. The XRD results indicated that due to the similarities of lattice parameters and ionic
radii of nickel and copper, nickel/copper can displace copper/nickel
in the synthesized structure, and the TGDSCMS results revealed
that the interactions of binary Ni10Mo10, Cu10Mo10 and ternary
Ni9.5Cu0.5Mo10 catalyst precursors are varied due to the introduction of copper. The HRTEM photographs and XRD characterization results of the sulded catalysts showed that the particle sizes
of the MoS2 nanoclusters and the sulded nickel-containing species varied after the introduction of copper. Small and narrow
XRD peaks attributing to sulded nickel-containing species could
be detected in the Ni9.5Cu0.5Mo10 catalyst, even though the content of nickel was almost equal to that in the Ni10Mo10 catalyst.
TPR results of sulded catalysts indicated that after the introduction of copper into nickelmolybdenum catalysts, the reduction
of nickel sulde from the NiMoS active phase shifted to lower
temperature region. The catalytic results indicated that the
Ni9.5Cu0.5Mo10 catalyst exhibited high catalytic activity in the
HDS of 4,6-DMDBT and FCC diesel. This catalyst is composed of
stacked MoS2 particles with high layers and much smaller nickelcontaining compounds that could help fulll the synergetic effect,
and thus lead to the efcient formation of NiMoS structure.

146

H. Liu et al. / Fuel 129 (2014) 138146

Acknowledgements
This work was nancially supported by the National Key Fundamental Research Development Project of China (973 Project No.
2010CB226905), National Natural Science Foundation of China
(Grants No. 21106185 and 21006128), the Fundamental Research
Funds for the Central Universities (Grants No. 14CX06032A),
Research Fund for the Doctoral Program of Higher Education of
China (Grant No. 20100133120007), Shandong Provincial Natural
Science Foundation of China (Grants No. ZR2011BQ002), and Postdoctoral Science Foundation of China (Grant No. 2013M530923).
Financial support from PetroChina Corporation Limited is also
greatly appreciated.
References
[1] Ho TC. Deep HDS of diesel fuel: chemistry and catalysis. Catal Today
2004;98:318.
[2] Ozkan US, Zhang LP, Ni SY, Moctezuma E. Characterization and activity of
unsupported NiMo sulde catalysts in HDN/HDS reactions. Energ Fuel
1994;8:8308.
[3] Lai WK, Pang LQ, Zheng JB, Li JJ, Wu ZF, Yi XD, et al. Efcient one pot synthesis
of
mesoporous
NiMoAl2O3
catalysts
for
dibenzothiophene
hydrodesulfurization. Fuel Process Technol 2013;110:816.
[4] Liu BJ, Zhu YB, Liu SW, Mao JW. Adsorption equilibrium of thiophenic sulfur
compounds on the Cu-BTC metal-organic framework. J Chem Eng Data
2012;57:132630.
[5] Kibsgaard J, Tuxen A, Knudsen KG, Brorson M, Topse H, Lgsgaard E, et al.
Comparative atomic-scale analysis of promotional effects by late 3d-transition
metals in MoS2 hydrotreating catalysts. J Catal 2010;272:195203.
[6] Castillo-Villaln P, Ramirez J, Castaeda R. Relationship between the
hydrodesulfurization of thiophene, dibenzothiophene, and 4,6-dimethyl
dibenzothiophene and the local structure of Co in CoMoS sites: infrared
study of adsorbed CO. J Catal 2012;294:5462.
[7] Tuxen AK, Fchtbauer HG, Temel B, Hinnemann B, Topse H, Knudsen KG, et al.
Atomic-scale insight into adsorption of sterically hindered dibenzothiophenes
on MoS2 and CoMoS hydrotreating catalysts. J Catal 2012;295:14654.
[8] Gonzlez-Corts SL, Xiao TC, Costa PM, Fontal B, Green ML. Ureaorganic
matrix method: an alternative approach to prepare CoMoS2/c-Al2O3 HDS
catalyst. Appl Catal A: Gen 2004;270:20922.
[9] Nguyen TS, Loridant S, Chantal L, Cholley T, Geantet C. Effect of glycol on the
formation of active species and suldation mechanism of CoMoP/Al2O3
hydrotreating catalysts. Appl Catal B: Environ 2011;107:5967.
[10] Escobar J, Barrera MC, Toledo JA, Corts-Jcome MA, Angeles-Chvez C, Nez
S, et al. Effect of ethyleneglycol addition on the properties of P-doped NiMo/
Al2O3 HDS catalysts: Part I. Materials preparation and characterization. Appl
Catal B: Environ 2009;88:56475.
[11] Lauritsen JV, Kibsgaard J, Olesen GH, Moses PG, Hinnemann B, Helveg S, et al.
Location and coordination of promoter atoms in Co- and Ni-promoted MoS2based hydrotreating catalysts. J Catal 2007;249:22033.
[12] Baldovino-Medrano VG, Giraldo SA, Centeno A. The functionalities of Pt/cAl2O3 catalysts in simultaneous HDS and HDA reactions. Fuel 2008;87:
191726.
[13] Pecoraro TA, Chianelli RR. Hydrodesulfurization catalysis by transition metal
suldes. J Catal 1981;67:43045.
[14] Rthlisberger A, Prins R. Intermediates in the hydrodesulfurization of 4,6dimethyl-dibenzothiophene over Pd/c-Al2O3. J Catal 2005;235:22940.
[15] Guzmn MA, Huirache-Acua R, Loricera CV, Hernndez JR, Daz de Len JN, de
los Reyes JA, et al. Removal of refractory S-containing compounds from liquid
fuels over P-loaded NiMoW/SBA-16 sulde catalysts. Fuel 2013;103:32133.
[16] Rashidi F, Sasaki T, Rashidi AM, Kharat AN, Jozani KJ. Ultra-deep
hydrodesulfurization of diesel fuels using highly efcient nanoaluminasupported catalysts: impact of support, phosphorus, and/or boron on the
structure and catalytic activity. J Catal 2013;299:32135.

[17] Zhao YF, Yang Y, Mims C, Peden C, Li J, Mei D. Insight into methanol synthesis
from CO2 hydrogenation on Cu(111): complex reaction network and the
effects of H2O. J Catal 2011;281:199211.
[18] Grabow LC, Mavrikakis M. Mechanism of methanol synthesis on Cu through
CO2 and CO hydrogenation. ACS Catal 2011;1:36584.
[19] lvarez-Rodrguez J, Guerrero-Ruiz A, Rodrguez-Ramos I, Arcoya-Martn A.
Surface and structural effects in the hydrogenation of citral over RuCu/KL
catalysts. Micropor Mesopor Mater 2006;97:12231.
[20] Harris S, Chianelli RR. Catalysis by transition metal suldes: a theoretical and
experimental study of the relation between the synergic systems and the
binary transition metal suldes. J Catal 1986;98:1731.
[21] Hernndez-Maldonado AJ, Yang RT. Desulfurization of diesel fuels by
adsorption via p-complexation with vapor-phase exchanged Cu(I)Y
zeolites. J Am Chem Soc 2004;126:9923.
[22] Dai W, Zhou YP, Li SN, Li W, Su W, Sun Y, et al. Thiophene capture with
complex adsorbent SBA-15/Cu(I). Ind Eng Chem Res 2006;45:78926.
[23] Yin CL, Zhao LY, Bai ZJ, Liu H, Liu YQ, Liu CG. A novel porous ammonium nickel
molybdate as the catalyst precursor towards deep hydrodesulfurization of gas
oil. Fuel 2013;107:8738.
[24] El Khadiri A, Astier MP. Characterisation by test reactions of nickelcopper
molybdenum catalysts prepared by reduction of well-dened complex
precursors. Appl Catal A: Gen 1997;193:17786.
[25] Eswaramoorthi I, Sundaramurthy V, Das N, Dalai AK, Adjaye J. Application of
multi-walled carbon nanotubes as efcient support to NiMo hydrotreating
catalyst. Appl Catal A: Gen 2008;339:18795.
[26] Levin D, Soled SL, Ying JY. Crystal structure of an ammonium nickel molybdate
prepared by chemical precipitation. Inorg Chem 1996;35:41917.
[27] Shaheen WM. Thermal behaviour of pure and binary basic nickel carbonate
and ammonium molybdate systems. Mater Lett 2002;52:27282.
[28] Aboul-Gheit AK, Awadallah AE, Aboul-Enein AA, Mahmoud ALH. Molybdenum
substitution by copper or zinc in H-ZSM-5 zeolite for catalyzing the direct
conversion of natural gas to petrochemicals under non-oxidative conditions.
Fuel 2011;90:30406.
[29] Rao GR, Meher SK, Mishra BG, Charan PHK. Nature and catalytic activity of
bimetallic CuNi particles on CeO2 support. Catal Today 2012;198:1407.
[30] Rodrguez-Castelln E, Jimnez-Lpez A, Eliche-Quesada D. Nickel and cobalt
promoted tungsten and molybdenum sulde mesoporous catalysts for
hydrodesulfurization. Fuel 2008;87:1195206.
[31] Cordero RL, Agudo AL. Effect of water extraction on the surface properties of
Mo/Al2O3 and NiMo/Al2O3 hydrotreating catalysts. Appl Catal A: Gen
2000;202:2335.
[32] Mangnus PJ, Riezebos A, Langeveld AD, Moulijin JA. Temperature-programmed
reduction and HDS activity of sulded transition metal catalysts: formation of
nonstoichiometric sulfur. J Catal 1995;151:17891.
[33] Puello-Polo E, Gutirrez-Alejandre A, Gonzlez G, Brito JL. Relationship
between suldation and HDS catalytic activity of activated carbon supported
Mo, FeMo. CoMO and NiMo Carbides. Catal Lett 2010;135:2128.
[34] Romero-Galarza A, Gutirrez-Alejandre A, Ramrez J. Analysis of the promotion
of CoMoP/Al2O3 HDS catalysts prepared from a reduced H-P-Mo
heteropolyacid Co salt. J Catal 2011;280:2308.
[35] Gutirrez OY, Klimova T. Effect of the support on the high activity of the
(Ni)Mo/ZrO2-SBA-15 catalystin the simultaneous hydrodesulfurization of DBT
and 4,6-DMDBT. J Catal 2011;281:5062.
[36] Nava R, Infantes-Molina A, Castao P, Guil-Lpez R, Pawelec B. Inhibition of
CoMo/HMS catalyst deactivation in the HDS of 4,6-DMDBT by support
modication with phosphate. Fuel 2011;90:272637.
[37] Egorova M, Prins R. Hydrodesulfurization of dibenzothiophene and 4,6dimethyldibenzothiophene over sulded NiMo/c-Al2O3, CoMo/c-Al2O3, and
Mo/c-Al2O3 catalysts. J Catal 2004;225:41727.
[38] Wang HM, Prins R. Hydrodesulfurization of dibenzothiophene and its
hydrogenated intermediates over sulded Mo/c-Al2O3. J Catal 2008;258:
15364.
[39] Alonso G, Berhault G, Aguilar A, Collins V, Ornelas C, Fuentes S, et al.
Characterization and HDS activity of mesoporous MoS2 catalysts prepared by
in situ activation of tetraalkylammonium thiomolybdates. J Catal 2002;208:
35969.

You might also like