Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Accepted Manuscript

Title: New Insights on the role of sulfur vacancies on


dibenzothiophene hydrodesulfurization over MoS2 edges

Authors: Seyyedmajid Sharifvaghefi, Bo Yang, Ying Zheng

PII: S0926-860X(18)30263-1
DOI: https://doi.org/10.1016/j.apcata.2018.05.033
Reference: APCATA 16682

To appear in: Applied Catalysis A: General

Received date: 10-4-2018


Revised date: 25-5-2018
Accepted date: 26-5-2018

Please cite this article as: Sharifvaghefi S, Yang B, Zheng Y, New Insights on the role of
sulfur vacancies on dibenzothiophene hydrodesulfurization over MoS2 edges, Applied
Catalysis A, General (2018), https://doi.org/10.1016/j.apcata.2018.05.033

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
New Insights on the role of sulfur vacancies on dibenzothiophene hydrodesulfurization over
MoS2 edges

Seyyedmajid Sharifvaghefia, Bo Yangc, Ying Zhenga,b,*

PT
aDepartment of Chemical Engineering, University of New Brunswick, 15 Dineen Drive, Fredericton,

NB, Canada E3B 5A3


bSchool

RI
of Engineering, University of Edinburgh, Mayfield Road, Edinburgh, UK EH9 3DW
E-mail: yzheng@unb.ca

SC
CSchool
of Physical Science and Technology, ShanghaiTech University, 393 Middle Huaxia Road,
Shanghai 201210, China

U
Graphical abstract
N
A
M
D
TE
EP
CC
A
Highlights
 Sulfur coverage at the edge of MoS2 influences the effect of H2S on HDS activity;
 Synergy of hematite and MoS2 leads to favored DDS pathway;
 The sigma adsorption of DBT is favored with vacancies present at edges of MoS2.
 Iron synergy with MoS2 leads to the formation of a vacancy site over Mo-edge.
 Acid sites are produced more readily by H2S dissociation than their H2 counterparts.

PT
Abstract

RI
Experimental and computational studies were performed to investigate the effect of varying

SC
sulfur coverage over MoS2 in hydrodesulfurization (HDS) of dibenzothiophene (DBT). Hematite
and iron particles were employed as H2S scavengers. MoS2 showed high hydrogenation (HYD)
preference in HDS of DBT which was explained by the DFT calculations showing that HYD pathway

U
is more favorable in brim adsorption. The addition of hematite drastically shifted this preference
to direct desulfurization (DDS) pathway. DFT calculations showed that MoS2 forms edges with
N
25% sulfur coverage in presence of hematite. The hydrogen dissociation was found to be
A
energetically unfavorable (deactivated) over the Mo-edge with such sulfur coverage and
becomes exothermic (reactivated) only after sigma adsorption of DBT on a vacancy site. In
M

agreement with experimental results, calculations also showed that in the presence of vacancies,
the sigma adsorption of DBT and therefore DDS pathway is more favorable. MoS2 with iron
particles present, favored the HYD reaction route, however, compared to MoS2 the biphenyl (BP)
D

selectivity increased while the selectivity for isomerized products was reduced. DFT calculations
showed that in the presence of iron, Mo-edge has 37% coverage with a vacancy site formed over
TE

this edge which explains the higher DDS activity for this catalyst compared to MoS2. A difference
was observed in the ability of H2 and H2S in creation of acid sites upon their dissociation at the
edges. This was attributed to the difference in S-H bond energies in each molecule and the change
EP

in oxidation sate of neighboring Mo atoms.


Keywords: MoS2, Density Functional Theory, Dibenzothiophene, Hydrodesulfurization, H2, H2S.
CC

1. Introduction
A

Hydrodesulfurization (HDS) is one of the most important reactions in a hydrotreating unit


employed in a refinery for removing sulfur from fuel fractions in order to meet the regulations
for sulfur levels of 15 ppm or lower in many countries [1]. Different types of catalysts have been
used based on types of fuels and desired product specifications. MoS2 is the most frequently used
catalyst that is usually promoted with Ni, Co and/or W and supported on alumina [2, 3]. The HDS
mechanism and its relation to catalyst active sites involved in the HDS process over MoS2 have
been extensively studied using various techniques. Imaging [4, 5], DFT studies [6, 7],
spectroscopic studies [8, 9], and use of poisons or inhibitors [10-12] are examples of these
techniques.

MoS2 consists of stacked layers in which each layer of molybdenum is sandwiched between two
layers of sulfur. These layers are joined together by weak van der Waals forces. A MoS2 particle
exposes two edges: Mo-edge (101̅0) and S-edge (1̅010). Traditional view advocates that the

PT
active sites of MoS2 are created by removing a sulfur from an edge forming a coordinatively
unsaturated site (CUS). More recently, scanning tunneling microscopy (STM) and DFT calculations
has illustrated the presence of the so-called brim sites with metallic like characteristics near the

RI
edges and on the basal plane of MoS2 [5, 13]. These sites were found to be active for
hydrogenation and possibly desulfurization. Molecules adsorption to these sites is through π-

SC
bonding and because of their position, even large molecules can easily adsorb to these sites
without any steric hindrance effect.

The sulfur coverage at each edge of MoS2 is considered to vary depending on reaction conditions

U
[14]. Earlier models assumed that naked Mo-edges (i.e. the 101̅0 Mo edge without terminal sulfur
N
atoms) possibly existed and were primarily active in HYD reaction [15, 16]. However, these
models were proposed mainly based on computational studies and lacked the experimental
A
evidence. Recent DFT studies demonstrated that the edges of MoS2 were fully or partially
covered by sulfur [17-20]. This is agreeable with the phase diagram profiles that show in normal
M

HDS reaction conditions the presence of naked edges is energetically unfavorable [21].

Several experimental and DFT studies have focused on the impact of H2S on HDS reaction [11, 22,
D

23]. The findings show that presence or absence of H2S in the gas phase can have a significant
effect on the activity and selectivity of the bulk MoS2mainly by promoting the HYD pathway.
TE

However, these findings lack a comprehensive study to explore the reasons behind such
promotional effect. We believe that the presence or absence of H2S in the gas phase can alter
the sulfur decoration at the edges. MoS2 is known to favor the HYD pathway in normal HDS
EP

conditions where the edges are covered with at least 50% sulfur, however, the behavior of the
catalyst H2S absent conditions where the sulfur decoration of the edges is below 50% is not fully
investigated. Here, we have made an attempt to both experimentally and computationally
CC

investigate the effect of varying sulfur coverage at edges on the activity and selectivity of MoS2.
Hematite and iron particles were used as scavengers to adsorb H2S from the gas phase which
resulted in changing the sulfur decoration of the edges. Dibenzothiophene (DBT) was chosen as
the sulfur-containing model compound to study the HDS mechanism in the presence and absence
A

of H2S and also to help us understand the H2S effect on the activity and selectivity of the final
products. In doing so, we also examined the difference between the roles of H2 and H2S in
creating acidic sites over the edges of MoS2.
2. Experimental
The hydrothermal synthesis method used in this study to prepare the catalysts is reported
elsewhere [24, 25] with some modifications. Briefly, sodium sulfide (0.21 g of, nonahydrate
crystals) was dissolved in Deionized water (DIW, 5 ml). Molybdenum trioxide powder (54 mg)
was then dissolved in this solution. Then, DIW (10 ml), Cetyltrimethylammonium chloride (0.2
ml, 25% in water) as surfactant, and HCl (0.4 ml, 4 M) were added. The prepared mixtures were
transferred to a 25 ml autoclave reactor. The autoclave was mounted in a furnace horizontally
and heated to 200 °C in 30 minutes while rotated with a speed of 200 rpm. The catalysts were

PT
removed after the reaction and washed 3 times with deionized water and ethanol. The prepared
sample was named fresh MoS2.

RI
Treating procedure: To prepare the treated catalysts, the suspension of fresh catalysts in toluene
were dried in an oven in nitrogen atmosphere in two consecutive steps. First, the temperature

SC
was increased at 10 °C/min to 85 °C and held at that temperature for 2 hours. Second, the
temperature was then increased to 120 °C and held for 1 hour. The dried catalyst was mixed with
model oil (1 wt% DBT, 99 wt% Dodecane). The mixture was then added to a 25 ml autoclave
reactor and the reactor was purged of any residual air. The reactor was pressurized with H2 to 21

U
bar and mounted horizontally in a furnace and rotated at 200 rpm. The reactor was heated up to
N
340 °C and kept at that temperature for 1 hour. The catalysts were then removed and washed
several times with toluene. Two other samples were also prepared. Physically mix 100 mg Fe2O3
A
and 100 mg Fe powders with the treated MoS2 catalyst via sonication and denoted as MoS2-Fe2O3
and MoS2-Fe, respectively.
M

Characterization: Specific surface area and pore-size distribution of the catalysts were measured
by N2 isothermal adsorption at 77 K using a Quantachrome Autosorb-1-C.
D

Brunauer−Emmett−Teller (BET) [26] was applied to determine the specific surface area of
samples.
TE

Transmission electron microscopy (TEM) imaging was performed on a JEOL 2011 (200 keV). To
prepare the samples, a drop of the particles dispersed in ethanol was dried on a carbon-coated
copper grid. The length and average number of layers for crystals were determined using image
EP

analysis software and the averages were calculated based on at least 200 MoS2 crystals measured
from different particles according to Equation (1) and (2).
CC

∑𝑛
𝑖=1,2…𝑛 𝐿𝑖 𝑁𝑖
Average slab length: 𝐿̅ = ∑𝑛
(1)
𝑖=1,2…𝑛 𝑁𝑖
∑𝑛
Average layer number: 𝑁 ̅ = 𝑖=1,2…𝑛 𝑁𝑖 (2)
𝑛
A

where L, N and n stand for slab length, number of layers in each crystal, and the total number of
crystalline measured, respectively.

Quantitative elemental analysis was conducted with a CHNS-932 instrument (Leco Corporation)
and with a SEM instrument via EDX detector on selected areas of samples to measure the sulfur,
carbon, and molybdenum content in the samples. The weight percent of carbon in the sample
was used to determine the approximate amount of the remaining surfactant after the catalyst
synthesis.

Catalytic activity test: Prior to HDS reactions, the catalyst suspension in toluene was dried in the
same way as described in the treating procedure. The dried catalyst was added to model oil (1
wt% DBT, 99 wt% Dodecane) using the weight ratio of 1/250 for MoS2 to model oil. The mixture
was added to the autoclave reactor. For the first part of this study, the reactor was pressurized
with H2 to 41 bars and mounted horizontally in a furnace and rotated at 200 rpm. For the second

PT
part, different H2 pressures were used to pressurize the reactor. The HDS reaction was carried
out at 320 °C for 2 hours after the reactor was heated up to this temperature in 30 minutes. The
collected products were analyzed with GC-MS (Shimadzu GCMS-QP5000) instrument for

RI
qualitative and Varian 450-Gas Chromatography equipped with both FID and PFPD detectors for
quantitative analysis. The selectivity of the compounds was defined as the ratio of one specific

SC
product compound to the total product obtained at the end of the reaction.
DFT calculations: A single-layer periodic stripe model was used for the DFT calculations reported
in this study. To study hydrogen adsorption, a supercell was generated containing a stripe with

U
four Mo atoms in the x-direction and four Mo atoms in the z-direction as shown in Figure 1. A
sulfur layer was added to each side of Mo atoms to build the MoS2 structure with molybdenum
N
atoms in trigonal prismatic position with respect to sulfur atoms. For DBT adsorption, the model
contains four rows of six Mo atoms. At least 12 A° of vacuum was considered between the stripes
A
with an adsorbed molecule in each periodic image in the x and z direction. The lattice constant
of MoS2 was determined to be 3.19, which is in good agreement with the experimental value of
M

3.16 [27, 28].


The structures and electronic energies were calculated using the plane wave DFT code
D

QUANTUM ESPRESSO [29] with ultra-soft pseudopotentials. The vdW-DF based optB86b-vdW
exchange correlation functional [30, 31] was used for all the calculations to consider for the effect
TE

of dispersion interactions. A Gaussian smearing of 0.05 eV was used, and all energies were
extrapolated to 0 K. Plane and density wave cutoff values were set at 30 and 240 Rydberg, with
the convergence criterion for geometric relaxation being 10^-3 Ry/Bohr. The Brillouin zone was
EP

sampled by a Monkhorst-Pack set of 1 × 1 × 2 k-point grid for the unit cells [32]. The calculations
for molecules in gas phase was carried out with the same parameters and settings as in the slab
models using a cubic box with at least 12 A° of vacuum in each direction employing the gamma
CC

point sampling.
To model Fe and Fe2O3, (110) surfaces were cleaved from their crystal structures and consisted
of six layers for Fe (Figure 2 a,b) and four layers for Fe2O3 (Figure 2, c,d). The XRD results showed
A

that Fe2O3 and Fe crystals expose the (110) face. Similar settings were employed as for MoS2 slab
beside introducing spin polarization and changing the K-point sampling to 6 × 6 × 1. Lattice
constants for Fe and Fe2O3 were calculated as 2.89 and 5.08, respectively. The values are
comparable to the experimental values of 2.87 and 5.04. To model the sulfur adsorption, sulfur
atom was placed on one side of the slab and three bottom layers for Fe and two bottom layers
for Fe2O3 were allowed to relax. The sulfur adsorption was obtained via the dissociative
adsorption of H2S on the MoS2 catalyst, Fe, and Fe2O3 particles represented by the following
equation:
H2S(g) + X = X-S + H2 (g)
Where X = MoS2 slab, Fe, and Fe2O3
Sulfur adsorption energy was found by the difference in total energy of the products minus the
reactants:

PT
Eads,S = E (products) – E (reactants) = [E (X-S) + E (H2)] – [E(X) + E (H2S)]
A positive value for Eads,S ,S indicates an unfavorable adsorption, while a negative value shows

RI
adsorption to be energetically favorable, with the lowest value being the most stable.

SC
Simplified steps were taken to find the preference for the HYD or DDS pathways in HDS of DBT
over MoS2 by studying mono-hydrogenated intermediates of DBT. Each product represents one
reaction pathway as shown in Figure 3. The following equation was used to find the formation

U
energy of the intermediates:
Eform,A/B = Eads,A/B - Eads,DBT

3. Results and discussion


N
A
M

3.1. Properties of catalysts


Table 1 shows the properties of the fresh and treated MoS2 and MoS2-Fe2O3 catalysts. The freshly
prepared catalysts were put through a treating procedure described in the experimental section
D

to remove the surfactant from the catalysts, remove any excess sulfur at catalyst edges and to
TE

further stabilize the structure of the catalysts. The high carbon content in the fresh catalysts is
due to the surfactant used during the catalyst synthesis. After the catalysts were treated, the
surfactant was mostly removed from their structure. The specific surface areas of the samples
EP

show a great improvement for the treated catalysts compared to the fresh ones with 19 and 12
times increase for MoS2 and MoS2-Fe2O3, respectively. The S/Mo ratios of the catalysts were
found to be below the stoichiometric value for MoS2 (ca. 1.9) indicative of presence of sulfur
CC

vacancies at the edges of these catalysts.


Figure 4 shows the TEM images of MoS2 before (a,b) and after the treatment procedure (c,d). As
the images show, some of the small crystals of the fresh catalyst join together to form longer
A

multilayered slabs. After the treating procedure, the length of the slabs of the MoS2
approximately increased by 60% and the average number of layers increased from 1.3 to 2.2.
After the HDS activity tests the structure of the catalysts remained unchanged. This means that
the amount of catalyst edges available for the HDS reactions is similar. Therefore, studying any
difference in the activity of the catalysts, the difference in the structure of MoS2 catalyst as an
interfering factor can be neglected.
The X-ray diffraction (XRD) patterns of MoS2 and MoS2-Fe2O3 catalysts after the treating
procedure and HDS reaction is shown in figure 5. Characteristic peaks of MoS2 shows the
presence of crystalline structure after the treating procedure. For the treated MoS2-Fe2O3 and
MoS2-Fe, the main XRD peaks can be assigned to Fe2O3 and Fe. MoS2 peaks are also present in
the spectra of both samples.

PT
3.2. Reaction routes over the catalysts

DBT is desulfurized through two main pathways, i.e. DDS and hydrogenation (HYD) [33]. Figure 6

RI
represents the reaction network for these HDS routes. In the DDS route, DBT undergoes direct C-
S bond cleavage by hydrogenolysis [34] or by elimination [35] to form bipheny (BP). In the HYD
route, cyclohexylbenzene (CHB) is formed via first hydrogenation of DBT to

SC
tetrahydrodibenzothiophen (THDBT) and/or hexahydrodibenzothiophene (HHDBT) and further
breakage of C-S bond in these molecules. CHB can then be isomerized to form
(cyclopentylmethyl)benzene (CPMB). Bicyclohexyl (BCH) is mainly formed through

U
hydrogenation and desulfurization of HHDBT. Both BCH and CPMB can isomerize to cyclohexane-
cyclopentylmethyl (CHCPM) which can undergo further isomerization to form
N
Dicyclopentylethane (DCPE). Preference of HDS reaction between these two routes mainly
depends on type of the active catalytic material, additives or promoters to the catalyst and
A
support [36].
M

Figure 7 shows the DBT conversion and product distribution of MoS2, MoS2-Fe2O3, and MoS2-Fe
catalysts. The mass ratio of MoS2-to-model oil is maintained the same for all the catalysts. The
selectivity of the isomerized products is shown in Figure 8. The DBT conversion for MoS2, MoS2-
D

Fe, and MoS2-Fe2O3 and MoS2 are 64, 67, and 47 percent, respectively. Overall, the results show
TE

that the increase in the amount of hematite leads to increase in the selectivity of BP but decrease
in the selectivity of CHB and BCH and therefore causing a great change in the DDS/HYD ratio. The
selectivity of BP decreases from nearly 80% for MoS2-Fe2O3 to 11% for MoS2 which contains no
hematite core. A reverse trend can be observed for hydrogenated products (CHB, BCH, and their
EP

isomerized forms). The selectivity for CHB showed an increase with the introduction of hematite.
No BCH was detected for MoS2-Fe2O3 while 5.1% of selectivity for BCH was observed over MoS2.
This clearly shows that the presence of hematite particles in the samples moves the preference
CC

of the reaction toward DDS. In the absence of hematite, the selectivity of the isomerized
compounds (CPMB, CHCPM, and DCPE) was increased.
A

3.3. Changes in Sulfur edge coverage


H2S content measurement in gas phase was performed for each test. As expected, no H2S was in
the gas when MoS2-Fe2O3 and MoS2-Fe were used as catalyst while 1.8 µmol of H2S was detected
when MoS2 is used. The main reason for the difference between the H2S content in each sample
can be related to the ability of hematite and iron particles to adsorb the produced H2S from the
gas phase. This was also confirmed through TEM imaging (Figure 9) showing the formation of
hexagonal particles after several cycles of DBT HDS which were identified as iron sulfide (FeS).
Although both iron and hematite have the ability to adsorb H2S, a clear difference can be
observed between their HDS activity and selectivity of the final products. MoS2-Fe sample shows
high affinity toward HYD with CHB selectivity of 59.6% while the selectivity for the same
compound dropped to 17.4% over MoS2-Fe2O3. The BP selectivity, on the other hand, reaches as
high as 79% over MoS2-Fe2O3 catalyst, indicating that DDS pathway dominates the

PT
hydrodesulfurization reaction. It was reported that sulfur decoration at the edges of MoS2
crystalline has a direct impact on the activity and selectivity of the HDS reaction [22, 37].
Therefore, in the first step, the edge decoration of MoS2 samples were determined. When MoS2

RI
was tested, based on our reaction conditions, both edges were considered to consist of 50%
sulfur coverage with sulfurs in bridging position between molybdenum atoms (Figure 1). For

SC
catalysts MoS2-Fe2O3 and MoS2-Fe, the sulfur adsorption energy (Eads,S) on hematite and iron was
first calculated. Eads,S was also calculated for sulfur and molybdenum edges of MoS2 (Table 2).

U
Eads,S for Fe was found to be lower than Eads,S4 on the Mo-edge and higher than the adsorption
energy of sulfurs at the S-edge. The adsorption of sulfur on Fe surface is therefore preferred to
N
S4 sulfur adsorption on Mo-edge and the S4 site remains vacant. The lower energy of the sulfur
adsorption on the S-edge compared to Fe results in the adsorption of sulfurs on the S-edge before
A
Fe surface. This suggests that the Mo-edge in the presence of Fe is decorated with 3 sulfur atoms
and on the S-edge, it is covered with 4 sulfurs at bridging positions (Figure 10a). For Fe2O3, we
M

found Eads to be lower than Eads,S3 and Eads,S4 on both the Mo and S-edges. The stable
structure for Mo and S-edges on MoS2 are therefore both have two sulfurs adsorbed at each edge
(Figure 10b).
D

After determining the edge structures for MoS2, MoS2-Fe2O3, and MoS2-Fe catalysts, hydrogen
dissociation was investigated on both edges of MoS2 in each catalyst using the following
TE

equation:
H2 + MoS2 slab = MoS2 slab-H-H
EP

Ediss,H2 = E (MoS2 slab-H-H)-[E(MoS2 slab)+E(H2)]


CC

Different states for H2 dissociation over the edges of catalysts was considered and the results are
shown in Table 3. Over the Mo-edge, two different configurations for H2 dissociation was
considered: a) forming two S-H groups and b) forming one S-H and one Mo-H group. The first
A

structure was found to be the stable configuration and we found the second configuration
endothermic (Ediss,H2 = 0.05 ev). This can be attributed to the fact that at Mo-edge with 50% sulfur
coverage, the molybdenum atoms are already six-fold coordinated. Spectroscopic studies, have
also proven the existence of S-H groups [38-40] but no evidence of the presence of Mo-H has
been found. Our calculation shows the formation of Mo-H groups at 50% sulfur covered Mo-edge
to be energetically unfavorable.
For MoS2 with 50% sulfur coverage at each edge, H2 dissociation was found to be considerably
favorable on the Mo-edge with Ediss,H2 of -0.24 compared to -0.04 on the S-edge. Table 3 shows
the corresponding position of hydrogen atoms on the edges. For MoS2-Fe2O3 catalyst with 25%
sulfur coverage at the edges, Ediss,H2 was found to be endothermic over Mo-edge for both
dissociation configurations considered involving the formation of two S-H or one S-H and one
Mo-H groups. We also tested the naked Mo-edge (0% sulfur coverage) to further explore the
hydrogen dissociation on this edge and found the dissociation on the naked edge to be favorable.

PT
Therefore, the only configuration found in which the hydrogen dissociation was unfavorable was
25% sulfur covered Mo-edge. The CUS formation on the S-edge, on the other hand, was found to
considerably increase the energy release from H2 dissociation suggesting an easier pathway for

RI
formation of H atoms. Between the two different configurations that were studied (Table 1),
hydrogen dissociation resulting in the formation of S-H and Mo-H groups (hydrogen at a bridging

SC
position between the Mo atoms) above that was the most stable.
The ability of 25% sulfur-covered Mo-edge for hydrogen dissociation was further investigated by
sigma adsorption of a DBT molecule at a CUS site. Upon adsorption of DBT, it was found that the

U
dissociation of hydrogen at the adjacent site forming Mo-H and S-H groups (Table 3) becomes
thermodynamically stable. This indicates that after the sigma adsorption of DBT, the catalyst can
N
resume its normal activity in providing dissociated hydrogen for further reactions. This is also in
A
agreement with our experimental results showing high DDS preference in the presence of
hematite particles. For MoS2-Fe2O3 catalyst, the Mo-edge has 25% sulfur coverage which can only
M

be activated after sigma adsorption of the DBT molecule on one of the CUS sites.
D

3.4. HYD and DDS pathways at the edges


We made an effort to study different forms of DBT adsorption (sigma and brim) at both Mo and
TE

S-edge and its transformation to intermediates from each reaction route. Mono-hydrogenated
DBT intermediates from HYD and DDS routes were considered at each edge and the adsorption
energies were examined (Table 4). At S-edge, the most stable hydrogenated intermediate was
EP

from the DDS pathway at a vacancy site. The formation of this intermediate was found to be
highly exothermic (-1.14 ev). This high adsorption energy and the highly favorable dissociation of
CC

H2 on this edge as shown in previous section, suggests that CUS sites are highly active for
desulfurization reactions. The intermediates formed through brim adsorption were energetically
unfavorable, however, the hydrogenated intermediate through HYD route was stabilized to a
greater extent compared to the intermediate through DDS route indicating the possibility of HYD
A

reaction pathway on these sites.


At Mo-edge, both brim and sigma adsorption were found equally exothermic (-0.88 ev). The most
stable hydrogenated intermediate was from the DDS pathway at the CUS site which was found
to be highly exothermic (-1.02 ev). However, the energy difference is lower when compared to
DDS pathway at the CUS site on S-edge (-1.14 ev) which indicates higher ability of S-edge in S-C
bond scission. Other DFT findings also support our findings, suggesting that S-edge exhibit a lower
barrier for S–C cleavage [41]. It is also found that the electron back-donation from the adsorption
sites to an adsorbed molecule is much stronger on S-edge than Mo-edge. Such electron back-
donation is suggested to play a role in the activity of the edge sites since it can strongly activate
the adsorbed molecule toward the following reactions [9, 42]. With brim adsorption over this
edge, the calculations showed that the hydrogenated intermediate via HYD pathway was,
although endothermic, yet more energetically favorable than the DDS intermediate. These
observations suggest that at both edges, hydrogenation may be favored over the brim sites and

PT
direct desulfurization at CUS sites.
3.5. Activity and selectivity
Considering the above discussion regarding the preferred adsorption sites, the difference

RI
between the activity and selectivity of the catalysts tested in this study can be understood. For
MoS2 catalyst and without the presence of any vacancy, DBT adsorbs over the brim sites and

SC
results in the formation of the mainly hydrogenated products (CHB and BCH). DBT adsorption on
MoS2-Fe is mainly over the brim sites which favor HYD pathway as found by the high selectivity
of CHB for this catalyst. Presence of iron in MoS2-Fe was found to facilitate the formation of a

U
vacancy over the Mo-edge which is largely active in direct desulfurization of DBT to BP. The higher
selectivity of BP over MoS2-Fe compared to MoS2 (25.9 vs 11.1, Figure 7) can be attributed to the
facile formation of this CUS site. N
A
We showed (Figure 10) the possibility of formation of four CUS sites over MoS2-Fe2O3 (two at
each edge,). The most stable adsorption mode for DBT is via sigma bonding with its sulfur atom
M

which results in the formation of mainly directly desulfurized product (BP) as evidenced through
the experiments. The higher activity of the other catalysts compared to MoS2-Fe2O3 can be
D

ascribed to their superior hydrogenation ability since DBT hydrogenated intermediates (THDBT
and HHDBT) are known to be more reactive than DBT in HDS [43]. The presence of H2S over MoS2
TE

therefore, has a minor effect on the activity of the catalyst since the main reaction sites for MoS2
are the brim sites. Our calculations agreeable to the previous findings showed that H2S can only
weakly adsorb on brim sites. Eads,H2S was calculated as -0.122 ev and -0.16 ev over the sulfur and
EP

molybdenum edges, respectively.


3.6. H2 and H2S dissociation and acid sites
CC

Both MoS2 and MoS2-Fe show high affinity toward the HYD pathway. However, the 9.9%
difference in the selectivity of the hydrogenated products (CHB and BCH) between MoS2 and
MoS2-Fe, suggests that H2S in addition to H2 may also play a role in hydrogenation. Moreover,
the selectivity of the isomerized compounds shows a significant increase with the increase in the
A

H2S content (MoS2 vs MoS2-Fe/MoS2-Fe2O3). H2S can dissociate into S-H and –H species over CUS
sites converting them into Brønsted acid sites [44, 45]. These acid sites can therefore increase
the hydrogenation ability of the catalyst. The role of H2S in hydrogenation is more profound in
the selectivity of BCH and the isomerized products. Either the increase in hydrogen over the
surface of the catalyst or presence of more H2S is the reason for this increase. Considering the
fact that no trace of BCH was found over MoS2-Fe2O3 catalyst where H2S is absent and the low
selectivity of BCH over MoS2-Fe catalyst where H2S is also absent, it can be concluded that H2S is
the main participant for the production of BCH. The process for BCH production can be both
through the production of higher amounts of HHDBT which further desulfurizes to BCH and/or
through direct hydrogenation of CHB. Following the same reasoning, it is clear that CUS sites
covered with dissociated H2S are the main sites for isomerization. The only exception can be the
production of CPMB which is detected even over MoS2-Fe2O3 and with high selectivity over MoS2-
Fe. The main sites for the formation of CPMB are therefore the same sites responsible for HYD

PT
pathway.
Although no H2S is available over MoS2-Fe catalyst, the HDS results show the presence of
isomerized products. Beside the dissociation of H2S, the Brønsted acid sites can be created

RI
through dissociation of H2 over the edges of MoS2. This will create the same –SH groups that H2S
dissociation at the edges would produce. However, there is a noticeable difference in the

SC
selectivity of the isomerized products over MoS2 (23.5%) and MoS2-Fe (8.5%). Considering the
8.5% selectivity through H2-produced acid sites, 15% of the selectivity for isomerized products
can be attributed to the H2S-produced acid sites which is almost twice the amount for H2-

U
produced acid sites. The first reason for this can be found in the different dissociation energies
of H-H bond (436 KJ/mol) and H-S bond (339 KJ/mol). Higher H-H bond energy makes the
N
dissociation of H2 more difficult. The second reason is that upon dissociation of H2S, the oxidation
A
state of the adjacent Mo atoms will remain unchanged. On the contrary, the dissociation of H2
will reduce the oxidation state of the adjacent Mo to +2 which is less favorable. These facts
M

contribute to the difference in the formation of acid sites through H2 and H2S dissociation (figure
11).
D

4. Conclusions
TE

To better understand the effects of varying sulfur coverages at the MoS2 edges, a series of
experimental and computational studies were performed. HDS of DBT was tested over MoS2 and
EP

MoS2 with and without the addition of iron and hematite particles as H2S scavengers. MoS2
showed high affinity toward HYD pathway. The addition of hematite to MoS2 (MoS2-Fe2O3) was
found to significantly shift the reaction to the DDS route and BP is a dominant product, up to
CC

79%. The addition of iron (MoS2-Fe) increased the selectivity of BP and reduced the selectivity of
the isomerized products while maintaining the HYD as the main reaction pathway.
DFT calculations showed that in the presence of hematite the sulfur and molybdenum edges are
A

covered with 25% sulfur. It was found that at this level of coverage, the Mo-edge becomes
inactive in dissociating hydrogen and it can only be reactivated by sigma adsorption of DBT at a
CUS site. Studying the hydrogenated intermediates, DDS was found to be the more favorable
reaction pathway with sigma adsorption and hydrogenation with brim adsorption over both
edges. In presence of iron particles, S-edge remains at its 50%sulfur coverage and one CUS site is
formed on the Mo-edge of MoS2 increasing the DDS activity of the catalyst, also evidenced by
increase in BP selectivity from the experimental results.
It was shown that in the presence of H2S, the desulfurization sites have the ability to transform
to hydrogenation/isomerization sites. Both H2 and H2S showed acid properties over MoS2 capable
of isomerizing the main HDS products. However, the results showed that the creation of acid sites
were more facile through H2S dissociation. This was assigned to the difference in the bond
energies for each molecule and the change in oxidation state of the adjacent Mo atom upon their

PT
dissociation.

RI
SC
U
N
A
M
D
TE
EP
CC
A
References

[1] A. Stanislaus, A. Marafi, M.S. Rana, Recent advances in the science and technology of ultra low sulfur
diesel (ULSD) production, Catalysis Today, 153 (2010) 1-68.
[2] H. Shimada, T. Sato, Y. Yoshimura, J. Hiraishi, A. Nishijima, Support effect on the catalytic activity and
properties of sulfided molybdenum catalysts, Journal of Catalysis, 110 (1988) 275-284.
[3] F. Van Looij, P. Van der Laan, W. Stork, D. DiCamillo, J. Swain, Key parameters in deep
hydrodesulfurization of diesel fuel, Applied Catalysis A: General, 170 (1998) 1-12.

PT
[4] Y. Zhu, Q.M. Ramasse, M. Brorson, P.G. Moses, L.P. Hansen, C.F. Kisielowski, S. Helveg, Visualizing the
Stoichiometry of Industrial‐Style Co‐Mo‐S Catalysts with Single‐Atom Sensitivity, Angewandte Chemie
International Edition, 53 (2014) 10723-10727.
[5] A.K. Tuxen, H.G. Füchtbauer, B. Temel, B. Hinnemann, H. Topsøe, K.G. Knudsen, F. Besenbacher, J.V.

RI
Lauritsen, Atomic-scale insight into adsorption of sterically hindered dibenzothiophenes on MoS 2 and
Co–Mo–S hydrotreating catalysts, Journal of Catalysis, 295 (2012) 146-154.

SC
[6] T.F. Jaramillo, K.P. Jørgensen, J. Bonde, J.H. Nielsen, S. Horch, I. Chorkendorff, Identification of active
edge sites for electrochemical H2 evolution from MoS2 nanocatalysts, science, 317 (2007) 100-102.
[7] P.G. Moses, L.C. Grabow, E.M. Fernandez, B. Hinnemann, H. Topsøe, K.G. Knudsen, J.K. Nørskov, Trends
in hydrodesulfurization catalysis based on realistic surface models, Catalysis Letters, 144 (2014) 1425-

U
1432.
[8] N. Topsoe, H. Topsoe, FTIR studies of Mo/Al 2 O 3-based catalysts: II. Evidence for the presence of SH
N
groups and their role in acidity and activity, Journal of Catalysis, 139 (1993) 641-651.
[9] J. Chen, F. Maugé, J. El Fallah, L. Oliviero, IR spectroscopy evidence of MoS 2 morphology change by
A
citric acid addition on MoS 2/Al 2 O 3 catalysts–A step forward to differentiate the reactivity of M-edge
and S-edge, Journal of Catalysis, 320 (2014) 170-179.
M

[10] H. Farag, A.-N.A. El-Hendawy, K. Sakanishi, M. Kishida, I. Mochida, Catalytic activity of synthesized
nanosized molybdenum disulfide for the hydrodesulfurization of dibenzothiophene: Effect of H 2 S partial
pressure, Applied Catalysis B: Environmental, 91 (2009) 189-197.
D

[11] H. Farag, K. Sakanishi, M. Kouzu, A. Matsumura, Y. Sugimoto, I. Saito, Investigation of the influence
of H2S on hydrodesulfurization of dibenzothiophene over a bulk MoS2 catalyst, Industrial & engineering
TE

chemistry research, 42 (2003) 306-310.


[12] P. Wiwel, B. Hinnemann, A. Hidalgo-Vivas, P. Zeuthen, B.O. Petersen, J.Ø. Duus, Characterization and
identification of the most refractory nitrogen compounds in hydroprocessed vacuum gas oil, Industrial &
Engineering Chemistry Research, 49 (2010) 3184-3193.
EP

[13] J.V. Lauritsen, J. Kibsgaard, G.H. Olesen, P.G. Moses, B. Hinnemann, S. Helveg, J.K. Nørskov, B.S.
Clausen, H. Topsøe, E. Lægsgaard, Location and coordination of promoter atoms in Co-and Ni-promoted
MoS 2-based hydrotreating catalysts, Journal of Catalysis, 249 (2007) 220-233.
CC

[14] M. Bollinger, K.W. Jacobsen, J.K. Nørskov, Atomic and electronic structure of MoS 2 nanoparticles,
Physical Review B, 67 (2003) 085410.
[15] S. Cristol, J.-F. Paul, E. Payen, D. Bougeard, F. Hutschka, S. Clémendot, DBT derivatives adsorption
over molybdenum sulfide catalysts: a theoretical study, Journal of Catalysis, 224 (2004) 138-147.
A

[16] H. Yang, C. Fairbridge, Z. Ring, Adsorption of Dibenzothiophene Derivatives over a MoS2 Nanocluster
A Density Functional Theory Study of Structure− Reactivity Relations, Energy & fuels, 17 (2003) 387-398.
[17] Á. Logadóttir, P.G. Moses, B. Hinnemann, N.-Y. Topsøe, K.G. Knudsen, H. Topsøe, J.K. Nørskov, A
density functional study of inhibition of the HDS hydrogenation pathway by pyridine, benzene, and H 2 S
on MoS 2-based catalysts, Catalysis Today, 111 (2006) 44-51.
[18] J.-F. Paul, E. Payen, Vacancy formation on MoS2 hydrodesulfurization catalyst: DFT study of the
mechanism, The Journal of Physical Chemistry B, 107 (2003) 4057-4064.
[19] B. Hinnemann, J.K. Nørskov, H. Topsøe, A density functional study of the chemical differences
between type I and type II MoS2-based structures in hydrotreating catalysts, The Journal of Physical
Chemistry B, 109 (2005) 2245-2253.
[20] L.P. Hansen, Q.M. Ramasse, C. Kisielowski, M. Brorson, E. Johnson, H. Topsøe, S. Helveg, Atomic‐Scale
Edge Structures on Industrial‐Style MoS2 Nanocatalysts, Angewandte Chemie International Edition, 50
(2011) 10153-10156.
[21] P.-Y. Prodhomme, P. Raybaud, H. Toulhoat, Free-energy profiles along reduction pathways of MoS 2
M-edge and S-edge by dihydrogen: a first-principles study, Journal of catalysis, 280 (2011) 178-195.

PT
[22] P. Afanasiev, The influence of reducing and sulfiding conditions on the properties of unsupported
MoS 2-based catalysts, Journal of Catalysis, 269 (2010) 269-280.
[23] H. Farag, K. Sakanishi, M. Kouzu, A. Matsumura, Y. Sugimoto, I. Saito, Dual character of H 2 S as
promoter and inhibitor for hydrodesulfurization of dibenzothiophene, Catalysis Communications, 4 (2003)

RI
321-326.
[24] S. Sharifvaghefi, Y. Zheng, Development of a Magnetically Recyclable Molybdenum Disulfide Catalyst
for Direct Hydrodesulfurization, ChemCatChem, 7 (2015) 3397-3403.

SC
[25] S. Sharifvaghefi, Y. Zheng, Dispersed Ni and Co Promoted MoS2 Catalysts with Magnetic Greigite as a
Core: Performance and Stability in Hydrodesulfurization, ChemistrySelect, 2 (2017) 4678-4685.
[26] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers, Journal of the

U
American Chemical Society, 60 (1938) 309-319.
[27] P. Raybaud, G. Kresse, J. Hafner, H. Toulhoat, Ab initio density functional studies of transition-metal
N
sulphides: I. Crystal structure and cohesive properties, Journal of Physics: Condensed Matter, 9 (1997)
11085.
A
[28] P. Raybaud, J. Hafner, G. Kresse, S. Kasztelan, H. Toulhoat, Structure, energetics, and electronic
properties of the surface of a promoted MoS2 catalyst: an ab initio local density functional study, Journal
M

of Catalysis, 190 (2000) 128-143.


[29] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G.L. Chiarotti, M.
Cococcioni, I. Dabo, QUANTUM ESPRESSO: a modular and open-source software project for quantum
D

simulations of materials, Journal of physics: Condensed matter, 21 (2009) 395502.


[30] J. Klimeš, D.R. Bowler, A. Michaelides, Chemical accuracy for the van der Waals density functional,
Journal of Physics: Condensed Matter, 22 (2009) 022201.
TE

[31] J. Klimeš, D.R. Bowler, A. Michaelides, Van der Waals density functionals applied to solids, Physical
Review B, 83 (2011) 195131.
[32] H.J. Monkhorst, J.D. Pack, Special points for Brillouin-zone integrations, Physical review B, 13 (1976)
EP

5188.
[33] K.G. Knudsen, B.H. Cooper, H. Topsøe, Catalyst and process technologies for ultra low sulfur diesel,
Applied Catalysis A: General, 189 (1999) 205-215.
CC

[34] T. Todorova, R. Prins, T. Weber, A density functional theory study of the hydrogenolysis reaction of
CH 3 SH to CH 4 on the catalytically active (100) edge of 2H-MoS 2, Journal of Catalysis, 236 (2005) 190-
204.
[35] F. Bataille, J.-L. Lemberton, P. Michaud, G. Pérot, M. Vrinat, M. Lemaire, E. Schulz, M. Breysse, S.
A

Kasztelan, Alkyldibenzothiophenes hydrodesulfurization-promoter effect, reactivity, and reaction


mechanism, Journal of catalysis, 191 (2000) 409-422.
[36] D. Valencia, T. Klimova, Citric acid loading for MoS 2-based catalysts supported on SBA-15. New
catalytic materials with high hydrogenolysis ability in hydrodesulfurization, Applied Catalysis B:
Environmental, 129 (2013) 137-145.
[37] W. Lai, Y. Xu, Y. Ren, L. Yang, J. Zheng, X. Yi, W. Fang, Insight into the effect of non-stoichiometric
sulfur on a NiMoS hydrodesulfurization catalyst, Catalysis Science & Technology, 6 (2016) 497-506.
[38] J. Maternova, Sulfhydryl groups on the surface of molybdenum desulfurization catalysts, Applied
Catalysis, 3 (1982) 3-11.
[39] M. Lacroix, H. Jobic, C. Dumonteil, P. Afanasiev, M. Breysse, S. Kasztelan, Role of adsorbed hydrogen
species on ruthenium and molybdenum sulfides. Characterization by inelastic neutron scattering,
thermoanalysis methods and model reactions, Studies in Surface Science and Catalysis, 101 (1996) 117-
126.
[40] H. Topsøe, B. Clausen, N.Y. Topsøe, J. Nørskov, C. Ovesen, C. Jacobsen, The bond energy model for
hydrotreating reactions: theoretical and experimental aspects, Bulletin des Societes Chimiques Belges,

PT
104 (1995) 283-291.
[41] P.G. Moses, B. Hinnemann, H. Topsøe, J.K. Nørskov, The hydrogenation and direct desulfurization
reaction pathway in thiophene hydrodesulfurization over MoS 2 catalysts at realistic conditions: a density
functional study, Journal of Catalysis, 248 (2007) 188-203.

RI
[42] J. Lauritsen, M. Nyberg, J.K. Nørskov, B. Clausen, H. Topsøe, E. Lægsgaard, F. Besenbacher,
Hydrodesulfurization reaction pathways on MoS 2 nanoclusters revealed by scanning tunneling
microscopy, Journal of Catalysis, 224 (2004) 94-106.

SC
[43] H. Wang, R. Prins, Hydrodesulfurization of dibenzothiophene and its hydrogenated intermediates
over sulfided Mo/γ-Al 2 O 3, Journal of Catalysis, 258 (2008) 153-164.
[44] M. Vrinat, R. Bacaud, D. Laurenti, M. Cattenot, N. Escalona, S. Gamez, New trends in the concept of

U
catalytic sites over sulfide catalysts, Catalysis today, 107 (2005) 570-577.
[45] E. Schachtl, E. Kondratieva, O.Y. Gutiérrez, J.A. Lercher, Pathways for H2 Activation on (Ni)-MoS2
N
Catalysts, The journal of physical chemistry letters, 6 (2015) 2929-2932.
A
M
D
TE
EP
CC
A
S-edge

PT
Mo-edge

RI
SC
Mo S C H

Figure 1: A typical periodic slab model for MoS2 with both S-edge and Mo-edge (top) and DBT

U
molecule with color scheme (bottom). Cyan: molybdenum; Yellow: Sulfur; Grey: Carbon;
White: Hydrogen.
N
A
M
D

a b
TE
EP
CC
A

Figure 2: Top and side views of a) Fe and b) Fe2O3 cells with sulfur adsorbed on the top layer.
A

PT
RI
Figure 3: Reaction pathways for the formation of mono-hydrogenated intermediates of DBT.

SC
U
a
N b
A
) )
M
D
TE
EP
CC
A
c d
) )

PT
RI
Figure 4: TEM images of MoS2, a, b) fresh and c, d) treated for 2 h.

SC
U
N
A
M
D
TE
EP
CC
A

Figure 5: XRD patterns of a) treated MoS2, b) MoS2-Fe2O3, and c) MoS2-Fe.


PT
RI
Figure 6: Possible reaction pathways for hydrodesulfurization of DBT.

SC
90 90

U
79
80 80

70
67 N 70
A
61
59.6
60 60
DBT conversion (%)

Selectivity (%)
M

47
50 47.1 50

40 40
D

30 25.9 30
TE

20 17.4 20
13.6
10.0 11.1
10 6.1 10
EP

5.1 4.5
3.2 3.8 3.4
2.3 1.7 2.5 1.9
1.2 0.6
0 0
DBT THDBT HHDBT CHB BCH BP CPMB DCPE CHCPM
conversion
CC

MoS₂ MoS₂-Fe MoS₂-Fe₂O₃

Figure 7: DBT conversion after 2h of reaction and product selectivity at 50% conversion for the
synthesized catalysts.
A
16 13.6

14
12
Selectivity (%)

10
8 6.1

PT
4.5
3.8
4 3.4
1.9

RI
0.6

0
CPMB DCPE CHCPM

SC
MoS₂ MoS₂-Fe MoS₂-Fe₂O₃

Figure 8: The selectivity of the isomerized products over the synthesized catalysts at 50%

U
conversion of DBT.

N
A
a b
M

) )
D
TE
EP
CC

Figure 9: TEM images of a) treated MoS2-Fe2O3 (circle shows hematite) and b) MoS2-Fe2O3 after
6h of HDS reaction (circled is the hexagonal iron sulfide formed from the hematite sulfidation).
A
(a) (b)

S-edge S-edge

Mo-edge

PT
Mo-edge

Figure 10: Edge configuration for MoS2 in the presence of a) iron (MoS2-Fe, left) and b) hematite

RI
(MoS2-Fe2O3, right).

SC
H
H
H S

U
H

N
A
+3 +3 +3 +3
M
D
TE
EP

H H H H
+2 +2 +3 S +3
CC

Figure 11: Dissociation of H2 and H2S on MoS2 and the oxidation states of MoS2.
A
Table 1: Properties of the fresh, treated and spent MoS2-Fe2O3 and MoS2 catalysts.

Sulfur Carbon S/Mo Specific Average Average


Sample content (± content (± ratio (± surface area slab length number
0.2 wt %) 0.2 wt %) 0.04) (m2/g) (nm) of layers
Fresh MoS2 21.0 34.0 1.90 24.3 3.6±1.5 1.3±0.2

PT
Treated MoS2 30.8 2.1 1.87 474.7 6.1±1.7 2.2±0.8
Spent MoS2 34.2 2.6 1.94 469.9 6.8±1.2 2.4±1.1
Spent MoS2-Fe2O3 18.2 2.1 1.88 207.1 6.9±1.5 2.4±0.8

RI
SC
U
Table 2: Sulfur adsorption energies (ev) at edges of MoS2, Fe, and Fe2O3.

Sulfur
number
Mo-edge S-edge Fe N
Fe2O3 3
2
A
S1 -2.72 -3.17 -1.39 -2.56 2
2

4
S2 -2.82 -2.80 4
M

1
S3 -1.43 -2.34 1
S4 -0.75 -2.49 3
D
TE
EP
CC
A
Table 3: Hydrogen dissociation energy over the molybdenum and sulfur edges of MoS2.

Adsorption site
Formed groups Mo-edge (50% S-edge (50% Mo-edge (25% S-edge (25%
coverage) coverage) coverage) coverage)
S-H & S-H -0.24 0.20 * - -

PT
RI
SC
Mo-H & S-H 0.05 -0.04 0.23 -0.50

U
N
A
M

Mo-H & Mo-H - - 0.39 -0.26


D
TE
EP
CC

* Hydrogen on the second sulfur moves to the bridging positions between the two Mo atoms.
A
Table 4: Formation energies of mono-hydrogenated intermediates of DBT on molybdenum and
sulfur edges of MoS2.

Site Initial structure Pathway Final structure Eformation


Mo- Brim HYD 0.39
edge

PT
RI
SC
DDS 0.83

U
N
A
M
D
TE

Edge HYD 0.38


EP
CC
A
PT
RI
DDS -1.02

SC
U
N
A
S-edge Brim HYD 0.19
M
D
TE
EP
CC

DDS 0.34
A
PT
Edge HYD 0.26

RI
SC
U
N
A
M

DDS -1.14
D
TE
EP
CC
A

You might also like