Kinetic Assessment of The Simultaneous Hydrodesulfurization of Dibenzothiophene and The Hydrogenation of Diverse Polyaromatic Structures (2018)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Research Article

Cite This: ACS Catal. 2018, 8, 3926−3942 pubs.acs.org/acscatalysis

Kinetic Assessment of the Simultaneous Hydrodesulfurization of


Dibenzothiophene and the Hydrogenation of Diverse Polyaromatic
Structures
Edgar M. Morales-Valencia,† Carlos O. Castillo-Araiza,*,‡ Sonia A. Giraldo,†
and Víctor G. Baldovino-Medrano*,†,§

Centro de Investigaciones en Catálisis (@CICATUIS), Parque Tecnológico de Guatiguará (PTG), km 2 vía El Refugio,
Universidad Industrial de Santander, Piedecuesta (Santander) 681011, Colombia

Grupo de Procesos de Transporte y Reacción en Sistemas Multifásicos, Laboratorio de Ingeniería de Reactores Aplicada a Sistemas
Químicos y Biológicos, Departamento de IPH, Á rea de Ingeniería Química, Universidad Autónoma Metropolitana-Iztapalapa,
Av. San Rafael Atlixco No. 186, C.P. 09340 México D. F., México
§
Laboratorio de Ciencia de Superficies (@CSSS_UIS), Parque Tecnológico Guatiguará (PTG), Km. 2 vía El Refugio,
Universidad Industrial de Santander, Piedecuesta (Santander) 681011, Colombia
*
S Supporting Information

ABSTRACT: The elimination of sulfur from fossil fuels via


hydrodesulfurization (HDS) is paramount to produce cleaner
fuels. Ultradeep HDS refers to reducing sulfur in fuels below the
10 ppm level. Nevertheless, under such conditions, dibenzothio-
phenes (DBTs) are to be desulfurized in the presence of highly
complex aromatic structures that possibly exert inhibitory effects.
Therefore, this contribution presents a kinetic study of the
inhibition effect of diverse aromatic structures: naphthalene (NP),
fluorene (FL), and phenanthrene (PHE), on the HDS of
dibenzothiophene over a sulfided NiMo/γ-Al2O3 catalyst. Kinetic
modeling was based on the Langmuir−Hinshelwood−Hougen−
Watson (LHHW) formalism and was submitted to regression
analyses with the reparametrized form of the Arrhenius and van’t
Hoff equations. Before addressing inhibition effects, the kinetics of
the HDS of DBT was revisited. In this sense, observations were better fitted when considering that the two parallel pathways for
the HDS of DBT, i.e., the so-called direct desulfurization (DDS) and hydrogenation-mediated desulfurization (HYD) routes,
occur on two different types of active sites. The developed model was used as a basis for the kinetic modeling of the inhibition of
aromatics on the HDS of DBT. The kinetic parameters for the aromatics were estimated on both catalytic sites and exhibited
thermodynamic consistency. Kinetic modeling indicated the following: (i) aromatic compounds and their reaction products are
adsorbed on both DDS and HYD sites; (ii) the hydrogenation of naphthalene occurs on both sites while fluorene and
phenanthrene only react on HYD sites; (iii) the entropy values suggested that the mobility of the molecules is higher on HYD
sites than on DDS sites, except for dibenzothiophene; and (iv) fluorene strongly inhibits HYD sites, because of its structure
similarity with dibenzothiophene. These findings are important because they provide an insight into the inhibition effects of
polyaromatic compounds of different chemical structures on ultradeep HDS.
KEYWORDS: hydrodesulfurization, dibenzothiophene, polyaromatics, inhibition, kinetic models

1. INTRODUCTION of new catalysts and processes. HDS catalystsnormally, sulfided


Increasingly stringent environmental requirements for fuels CoMo/γ-Al2O3 and/or NiMo/γ-Al2O3face ultradeep HDS
result in operational and economic challenges for the petroleum aiming to remove sulfur from refractory dibenzothiophenes
refining industry.1,2 Hydrodesulfurization (HDS) of fuel frac- (DBTs).7−9
tions is one of the major catalytic processes coping with environ- It is rather well-established that DBTs react on the catalytic
mental statutes for sulfur content. For instance, the maximum surface via the two parallel pathways shown in Scheme 1, namely,
allowed concentration of sulfur in diesel fuels in the European via direct desulfurization (DDS), which yields biphenyl (BP),
Union is 10 ppm,3 15 ppm in the United States,4 10 ppm in Japan,5
and 50 ppm (on average) in South America.6 When the sulfur Received: February 14, 2018
content in fuels is ∼10 ppm, refiners refer to ultradeep HDS. Revised: March 19, 2018
Such demands drive constant research efforts for the development Published: March 21, 2018

© XXXX American Chemical Society 3926 DOI: 10.1021/acscatal.8b00629


ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Scheme 1. Reaction Network for the Hydrodesulfurization the HDS of DBT over NiMo/γ-Al2O3.29 Nevertheless because of
(HDS) of Dibenzothiophene (DBT) the high hydrogenation activity exhibited at high hydrogen
partial pressures, NiMo catalysts are preferred for ultralow sulfur
diesel production13,30 and for the pretreatment stage of the feed
for hydrocracking units.31,32 For that reason, more-rigorous
kinetic modeling of the HDS over a NiMo/γ-Al2O3 catalyst are
required to approach the molecular level description of hydro-
treating reactions.
In ultradeep HDS, DBTs must be desulfurized in the presence
of highly complex aromatic structures.33−35 Several authors have
studied the HDS of DBT under such conditions.34,36,37 Aromatic
compounds with two and three fused rings have been found to
inhibit HDS catalysts.36,38,39 Nevertheless, the inhibiting effect
on the aforementioned reaction routes is far from being
understood. The studies, which are undertaken to determine
possible inhibition effects of aromatics over the HDS of DBTs,
often limit themselves to describing relative impacts of the
presence of the aromatics on the conversion and selectivity of a
selected model molecule.34,36,40 Some contradictory trends have
been reported. Some authors observe that aromatics may inhibit
and via hydrogenation-mediated desulfurization (HYD) path- the HYD route of desulfurization to a larger extent, compared to
ways. HYD comprises intermediate steps in which one of the the DDS route,8,41 whereas some others report that both reaction
benzene rings of DBT is first hydrogenated to tetra-hydro- routes are inhibited to the same extent.7,40,42
dibenzothiophene (THDBT) and then to hexa-hydro-dibenzo- Kinetic studies for the HDS of DBTs in the presence of
thiophene (HHDBT). These hydrogenation reactions are aromatic compounds are limited. Pseudo-first-order models have
followed by the scission of the C−S−C bond to yield cyclo- been frequently used to correlate the inhibition effect of the
hexylbenzene (CHB).10−13 aromatic compounds.8,36,42 Nevertheless, with this approach, the
The question of the number and nature of the active sites adsorption equilibrium constants for all the compounds involved
involved in DDS and HYD is not clear yet. Some authors14−17 in the reaction are not determined, which does not allow the
have proposed a common dihydro-dibenzothiophene elucidation of a reaction mechanism.38 A simplified LHHW
(DHDBT) intermediate for both the DDS and HYD reaction model that assumes one site on which DBTs, aromatics, and that
pathways. They suggest that this intermediate could be the reaction products competitively adsorb have been postulated
hydrogenated or desulfurized on the same active site. The by Koltai et al.34 However, physicochemical and statistical criteria
difference in reactivity between the two pathways would thus be were not employed to evaluate the estimated kinetic and
related to the different reaction rates for the transformation of adsorption parameters. To the best of our knowledge, a rigorous
DHDBT. Kinetic models based on this hypothesis have kinetic modeling that describes the competitive adsorption of
proposed one site on which DBT and its reaction products aromatic compounds and DBT, assuming the existence of two
competitively adsorb and a second site for H2 adsorption.14,18,19 types of active sitesone for the hydrogenation reactions and
However, a strong argument against this mechanism is that the other for the C−S−C bond scissionhas not been published
DHDBT has never been detected in HDS reactions, under yet for NiMo/γ-Al2O3.
standard operational conditions. Other studies have suggested Considering the above, the objectives of this work were (i) to
that the HDS pathway is determined by the adsorption mode of revisit the kinetic model for the HDS of DBT over a NiMo/
the DBT molecules. Particularly, two modes of adsorption can be γ-Al2O3 catalyst for elucidating the role of the type of the catalytic
taken into account. For the DDS and the final C−S−C bond centers involved in the reaction, and (ii) to propose kinetically
scission in the HYD reaction route, adsorption is considered to viable reaction pathways for the mutual inhibition effects present
occur via direct attachment of the sulfur heteroatom in a in the simultaneous HDS of DBT and hydrogenation of
σ adsorption mode. Meanwhile, HYD is considered to occur via aromatics with different chemical structures (naphthalene
π-type adsorption of the aromatic structure of the molecule on (NP), phenanthrene (PHE), and fluorene (FL)). To achieve
the active site.20,21 These two different modes of adsorption the first objective, three kinetic models were devel-
require different active sites and different adsorption and kinetic oped. The first one was based on the assumption that HYD
constants for the DDS and HYD routes.22−24 However, these and DDS occur on the same types of active sites. The second
adsorption modes have only been shown to be feasible via assumed that the HYD and DDS routes occur on two different
molecular simulations.15 active sites. And, the third model also assumed that hydrocarbons
Assuming more than one active site for the adsorption of are adsorbed on the two different active sites, but included the
hydrocarbons increases the complexity of kinetic modeling by hydrogenation of biphenyl to cyclohexylbenzene in the reaction
augmenting the number of kinetic and adsorption parameters to network. For the second objective, different kinetic models were
be estimated. Kinetic expressions for the HDS of DBT, assuming developed to elucidate the structural effects of the polyaromatics
that the hydrogenolysis and the hydrogenation reactions occur on the simultaneous reactions of hydrogenation and HDS of
on two different active sites, have been modeled scarcely. In this DBT. Naphthalene and phenanthrene were used as models of
regard, pseudo-first-order models25 and the Langmuir−Hinshel- aromatics with two and three fused rings, respectively. Fluorene
wood−Hougen−Watson (LHHW) formalism have been used was selected as a model for aromatics with a structure equivalent
for modeling kinetics mainly over CoMo/γ-Al2O3 catalysts.26−28 to that of DBT, i.e., aromatics with a five-carbon-membered
Less attention has been given to the study of kinetic models for thiophenic ring. All models were based on the LHHW formalism
3927 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

and subjected to regression analyses with the reparametrized velocity of the aromatic compounds was changed between
form of the Arrhenius and van’t Hoff equations. Because of the 28 kgcat kmolDBT−1 h and 118 kgcat kmolDBT−1 h (i.e., between
latter, activation energies and pre-exponential factors, as well as 2.0 wt % and 6.0 wt %). Reaction temperatures were between
the adsorption enthalpies and entropies of each compound were 260−300 °C. Consequently, the total number of experiments for
estimated. Statistical tests were performed to determine the the HDS of DBT, in the absence and presence of aromatic
overall significance of the regression and for the individual compounds, amounted to 80. Therefore, ca. 446 observations
significance of the parameters.43−45 The study provided robust were examined during modeling.
kinetic modeling for the inhibition of aromatics on the HDS of Liquid products identification was made using gas chromatog-
DBT. raphy (GC) and mass spectroscopy (MS). GC analyses were
performed with a HP 6890 chromatograph equipped with an
2. CATALYTIC TESTS FID detector and an automatic injector. An HP-1 column
(Agilent J&W, 100 m × 0.25 mm × 0.5 μm) was used for both
Catalytic tests were carried out in two different reaction setups, GC and GC-MS. Analysis conditions were as follows: the GC
namely, Catatest (I) and Catatest (II) (vide Figure S1 in the oven temperature was programmed from 90 °C to 180 °C (17 min)
Supporting Information), provided with high-pressure fixed-bed at a rate of 60 °C min−1, then to 260 °C (10 min), using a temper-
continuous flow reactors operated in integral mode. Reactors ature ramp of 80 °C min−1. Helium (Linde Colombia S.A, 99.99%)
were packed with an 8 wt % Mo and 1.8 wt % Ni NiMo/γ-Al2O3 was used as a carrier gas, with a linear velocity of 19 cm s−1
commercial catalyst (Procatalyse, Brunauer−Emmett−Teller (1.1 mL min−1, at a constant flow). Compounds in each sample
(BET) specific surface area ≈ 150 m2 g−1, Barrett−Joyner− were identified by means of a computer matching method,
Halenda (BJH) pore volume ≈ 0.42 cm3 g−1, and average BJH comparing their spectra with those provided in the Wiley, NIST,
pore diameter ≈ 11.2 nm). Extrudates of the catalyst were and QUADLIB libraries. The experimental error in the mass
ground and sieved to obtain particles with diameters ranging balance was found to be ±5%. The carbon balance calculated for
from 300 μm to 600 μm. For the catalytic tests, ca. 0.30 g each reaction is given in Tables S1−S4, in Section S2 of the
(Catatest I) or ca. 0.15 g (Catatest II) of catalyst were dried in Supporting Information.
situ under N2 flow (100 mL min−1) at 120 °C for 1 h. Afterward, Catalytic results were expressed in terms of conversion (%Xi),
the catalyst was sulfided during 4 h using a volumetric flow rate of products selectivity (%Sj), and yield (%Yj) percentages.
100 mL min−1 containing 15% of H2S in H2 at atmospheric Conversion was calculated as follows:
pressure and 400 °C. After sulfidation, reactants were fed to the
reactor at a volumetric flow rate of 30 mL h−1. The reactor Fi0 − Fi
%Xi = × 100
pressure was increased with H2 to 5 MPa and a hydrogen/(liquid Fi0 (1)
feed) rate ratio of 500 NL L−1 was fixed for the exper-
iments. Reaction temperatures were programmed to start at where and Fi are the inlet and outlet molar flow rate of the
F0i
300 °C to stabilize the catalyst. Afterward, the temperature was hydrocarbon, respectively. Selectivity and yields were calculated
decreased to 260 °C, followed by its increase to 280 °C and as follows:
finally to 300 °C again. Such conditions were applied to rule out Fj
possible deactivation effects. Some reactions were performed %Sj = × 100
independently at each one of the aforementioned temperatures, Fi0 − Fi (2)
using fresh catalysts at every run. Such conditions were applied
for a second check for possible deactivation of the catalyst during Fj
the experiments. Reactions were conducted until reaching steady %Yj = × 100
Fi0 (3)
state, considered herein as measurements where catalytic
conversion and selectivity did not change more than ±2% over where Fj is the outlet molar flow rate of product j.
time on stream. The absence of heat- and mass-transport limi-
tations was verified as stated elsewhere.37,46 3. KINETIC MODELING
The reaction feedstock accounted for the following com- 3.1. Hydrodesulfurization of Dibenzothiophene. Differ-
pounds employed either individually or in blends: DBT (Sigma− ent kinetic models based on the LHHW formalism were
Aldrich, 98%), as a model sulfur compound in ultradeep HDS; developed. The first model, denoted as DBT1S, considered a
naphthalene (Laboratorios León, 98%); fluorene (Merck, 95%); single active site (i.e., * sites, over which DBT and its reaction
and phenanthrene (Sigma−Aldrich, 98%), as model aromatic products adsorb and react via either DDS or HYD routes). The
compounds. Hexadecane (Sigma−Aldrich, ≥99%) was used as second model, named DBT2S, assumed that the HYD and DDS
an internal standard for chromatography and cyclohexane routes occur on two different active sites (i.e., π sites, where the
(commercial grade) was used as solvent. molecules are hydrogenated, and σ sites, over which the
Two sets of experiments were carried out to evaluate the molecules are desulfurized). A third model, denoted as
kinetics and inhibiting effect of aromatic compounds on the HDS DBTBP2S, was based on the same considerations as those
of DBT. In the first set of experiments, tests were performed in made for the DBT2S model, but it included the hydrogenation of
the absence of aromatic compounds. Temperature was varied biphenyl to cyclohexylbenzene.11,27,29 The three kinetic models
between 260 °C and 330 °C, and space-time velocity (WcatF−1 DBT0) assumed that the adsorption and heterolytical dissociation of H2
was varied between 33 kgcat kmolDBT h and 242 kgcat kmolDBT−1 h,
−1
occurred on β surface sites, which are different from the sites over
and, hence, the concentration of DBT was between 1.0 wt % which hydrocarbons adsorb.47,48
and 3.7 wt %. The second set of experiments was conducted The reaction mechanism for DBT1S is presented in Table 1.
in the presence of naphthalene, fluorene, or phenanthrene. These Therein, υi is the stoichiometric number used to describe the
experiments were carried out at a fixed space-time velocity of DBT number of times that each adsorption, desorption, and reac-
of 122 kgcat kmolDBT−1 h (i.e., 2 wt % DBT). The space-time tion step must occur in order to complete one catalytic cycle,
3928 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Table 1. Reaction Mechanism and Catalytic Cycles for the β sites (step A). DDS occurs on a single reaction on σ sites
Kinetic Model DBT1S (step C), while HYD is accounted for on π sites (step F).26,27
Steps J−L were considered only for the model DBTBP2S where
step mechanism v1 v2
the hydrogenation of BP to CHB was assumed to occur as
A H 2 + 2β ⇄ 2Hβ 2 5 presented in step K. The global reaction represented in step 3
B DBT + * ⇄ DBT* 1 1 and the respective v3 were only considered for this model. The
C DBT* + 3Hβ → BP* + SHβ + 2β 1 0 adsorption and readsorption of H2S on σ and π sites were also
BP* ⇆ BP + *
taken into consideration in steps M−P.
D 1 0
Generally, the following assumptions were made: (i) an
E DBT* + 9Hβ → CHB* + SHβ + 8β 0 1 approximation to pseudo-equilibrium for adsorption and desorp-
F CHB* ⇄ CHB + * 0 1 tion steps; (ii) the concentration of hydrogen is constant during
G SHβ + Hβ ⇄ H 2Sβ + β 1 1 the reaction due to its large partial pressure. In consequence, H2
H H 2 Sβ ⇄ H 2 S + β 1 1 concentration and sites β were lumped into the kinetic constant
of the corresponding reaction rate; (iii) surface reactions are rate-
I H 2S + * + β ⇄ SH* + Hβ 1 1
determining steps (RDS); (iv) the hydrogenolysis of THDBT
J SH* + Hβ ⇄ H 2S + * + β 1 1 and HHDBT are fast enough to be immediately desulfurized
Global Reactions into cyclohexylbenzene and H2S; and (v) all products contribute
1 DBT + 2H 2 → BP + H 2S 1 0 to kinetic inhibition by competitive adsorption with DBT
2 DBT + 5H 2 → CHB + H 2S 0 1 (i.e., biphenyl, cyclohexylbenzene, and H2S). Based on these
assumptions, the corresponding kinetic models can be written as
according to the overall reaction i, the latter represented by follows:
global reactions 1 and 2. The dissociation of H2 was considered DBT1S:
to occur only on β sites (step A). Whereas, DBT, biphenyl, rDBT,1 = − kDBT,1KDBT, *PDBTθ (4)
*
and cyclohexylbenzene were considered to adsorb and react over
* sites (steps B−F). The direct desulfurization reaction and rDBT,2 = − kDBT,2KDBT, *PDBTθ
* (5)
hydrogenation pathway are represented by steps C and E,
respectively. The formation and later desorption of H2S from DBT2S:
β sites are represented in steps G and H, respectively. The rDBT, σ = − kDBT, σKDBT, σ PDBTθσ (6)
H2S desorbed from β sites may be readsorbed on * sites (steps I
and J). rDBT , π = − kDBT , πKDBT , π PDBTθπ (7)
Table 2 presents the reaction mechanism proposed for both
DBTBP2S:
DBT2S and DBTBP2S models. H2 is dissociatively adsorbed on
rDBT, σ = −kDBT, σKDBT, σ PDBTθσ (8)
Table 2. Reaction Mechanism and Catalytic Cycles for the
Kinetic Models DBT2S and DBTBP2S rDBT, π = −kDBT, πKDBT, π PDBTθπ (9)
step mechanism v1 v2 v3 rBP, π = −kBP, πKBP, π PBPθπ (10)
A H 2 + 2β ⇄ 2Hβ 2 5 3
where
B DBT + σ ⇄ DBTσ 1 0 0
DBTσ + 3Hβ → BPσ + SHβ + 2β
1
C 1 0 0 θj =
BPσ ⇄ BP + σ
1 + ∑ K n , jPn (11)
D 1 0 0
E DBT + π ⇄ DBTπ 0 1 0 Herein, θj denotes the fraction of free active sites j, Kn,j corre-
F DBTπ + 9Hβ + σ → CHBσ + SHβ + 8β + π 0 1 0 sponds to the adsorption equilibrium coefficient for component
G CHBσ ⇄ CHB + σ 0 1 0
n on site j, and Pn is the partial pressure of component n.
Generally, calculations showed a high coverage of DBT on
H SHβ + Hβ ⇄ H 2Sβ + β 1 1 0 active sites (vide Section S3 of the Supporting Information).
I H 2 Sβ ⇄ H 2 S + β 1 1 0 Therefore, kinetic expressions were assumed to keep a constant
J BP + π ⇄ BPπ 0 0 1 reaction order.
BPπ + 6Hβ → CHBπ + 6β
3.2. Hydrodesulfurization of Dibenzothiophene Simul-
K 0 0 1
taneous to the Hydrogenation of Aromatic Compounds.
L CHBπ ⇄ CHB + π 0 0 1 After analyzing results that will be discussed later (Section 4.2.1),
M H 2S + σ + β ⇄ SHσ + Hβ 1 0 0 the considerations proposed for model DBTBP2S were used for
N SHσ + Hβ ⇄ H2S + σ + β 1 0 0 the kinetic modeling of the HDS of DBT simultaneous to the
hydrogenation of aromatic compounds. Kinetic models were
O H 2S + π + β ⇄ SHπ + Hβ 0 1 0
developed assuming that DBT, biphenyl, cyclohexylbenzene, and
P SHπ + Hβ ⇄ H 2S + π + β 0 1 0 H2S are adsorbed on both sites, and aromatic molecules and their
Global Reaction reaction products react either only on π sites or on both π and
1 DBT + 2H 2 → BP + H 2S 1 0 0 σ sites. The reversibility of the hydrogenation of naphthalene,
2 DBT + 5H 2 → CHB + H 2S 0 1 0
fluorene, and phenanthrene was considered for kinetic modeling.
In what follows, only the kinetic models that led to the best
3a BP + 3H 2 → CHB 0 0 1 results are presented for the sake of brevity. Kinetic models
a
Global reaction 3 must only be considered for the DBTBP2S model. describing the inhibiting effect on the simultaneous HDS of DBT
3929 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

and hydrogenation of naphthalene, fluorene, and phenanthrene Table 3. Reaction Mechanism and Catalytic Cycles for the
were named DBTNP, DBTFL, and DBTPHE, respectively. Kinetic Model DBTNP To Describe the Hydrogenation of NP
Kinetic expressions for each model were based on the mech-
step mechanism v1 v2
anism of the HDS of DBT (i.e., Table 2, integrated with the
mechanism of the respective hydrogenation of each aromatic A H 2 + 2β ⇄ 2Hβ 2 2
molecule). Generally, assumptions for the proposed models were B NP + σ ⇄ NPσ 1 0
based on both experimental results and literature review. The C NPσ + 4Hβ ⇄ TTLσ + 4β 1 0
reader may refer to Section S4 of the Supporting Information for
D TTLσ ⇄ TTL + σ 1 0
data concerning other models tested in this work.
3.2.1. Hydrogenation of Naphthalene Simultaneous to the E NP + π ⇄ NPπ 0 1
Hydrodesulfurization of Dibenzothiophene (DBTNP). Scheme 2a F NPπ + 4Hβ ⇄ TTLπ + 4β 0 1
G TTLπ ⇄ TTL + π 0 1
Scheme 2. Reaction Network for the Hydrogenation of Global Reaction
Polyaromatic Compounds: (a) Naphthalene, (b) Fluorene, 1 NP + 2H 2 ⇄ TTL 1 1
and (c) Phenanthrene
rDBTNP, π = −kDBT, πKDBT, π PDBTθDBTNP, π (13)

rBPNP, π = −kBP, πKBP, π PBPθDBTNP, π (14)


F B
rNP, σ = −(kNP, σ KNP, σ PNP − kNP, σ KTTL, σ PTTL)θDBTNP, σ
(15)
F B
rNP, π = −(kNP, π KNP, π PNP − kNP, π KTTL, π PTTL)θDBTNP, π
(16)
where kFNP and kBNP are the reaction rate coefficients of the forward
and backward steps of naphthalene hydrogenation.
3.2.2. Hydrogenation of Fluorene Simultaneous to the
Hydrodesulfurization of Dibenzothiophene (DBTFL). A sim-
plified hydrogenation network of fluorene based on the
proposition by Lapinas et al.52 is presented in Scheme 2b. For
the formulation of the kinetic model for fluorene hydrogenation
in the presence of DBT, the following assumptions were made:
(i) only the hydrogenation of the first ring of fluorene to form
1,2,3,4,4a,9a-hexahydrofluorene (HHFL) takes place; (ii) the
hydrogenation of fluorene only occurs on π sites; and, (iii) the
reversibility of the reaction is neglected. Table 4 presents the

Table 4. Reaction Mechanism and Catalytic Cycles for the


Kinetic Model DBTFL To Describe the Hydrogenation of FL
step mechanism v1
A H 2 + 2β ⇄ 2Hβ 3
B FL + π ⇄ FLπ 1
C FLπ + 6Hβ → HHFLπ + 6β 1
shows a simplified reaction network for naphthalene hydro- D HHFLπ ⇄ HHFL + π 1
genation.49−51 In the presence of DBT, the modeling of Global Reaction
naphthalene hydrogenation was based on the following 1 FL + 3H 2 → HHFL 1
assumptions:
(i) naphthalene is only hydrogenated to tetralin (TTL); mechanism to describe the hydrogenation of fluorene. Hydrogen
(ii) naphthalene and tetralin competitively adsorb and react and fluorene are adsorbed on β and π sites, respectively (steps A
on both π and σ sites; and and B). Step C summarizes a series of sequential hydrogenation
(iii) naphthalene hydrogenation to tetralin is reversible. steps. Finally, in step D, 1,2,3,4,4a,9a-hexahydrofluo is desorbed.
The rate expressions for the simultaneous DBT HDS and fluor-
Table 3 presents the mechanism employed to describe
ene hydrogenation were described by the following equations:
naphthalene hydrogenation. The dissociation of H2 occurs over
β sites (step A). Naphthalene is adsorbed and hydrogenated on rDBTFL, σ = −kDBT, σKDBT, σ PDBTθDBTFL, σ (17)
σ (steps B and C) and π sites (steps E and F). Finally, tetralin is
desorbed (steps D and G). rDBTFL, π = −kDBT, πKDBT, π PDBTθDBTFL, π (18)
The rate expressions considered for the simultaneous HDS of
DBT and hydrogenation of naphthalene were rBPFL, π = −kBP, πKBP, π PBPθDBTFL, π (19)

rDBTNP, σ = −kDBT, σKDBT, σ PDBTθDBTNP, σ (12) rFL, π = −kFL, πKFL, π PFLθDBTFL, π (20)

3930 DOI: 10.1021/acscatal.8b00629


ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

3.2.3. Hydrogenation of Phenanthrene Simultaneous to The rate expressions for both the simultaneous HDS of DBT
the Hydrodesulfurization of Dibenzothiophene (DBTPHE). and hydrogenation of phenanthrene are presented below:
Phenanthrene is a polycyclic aromatic hydrocarbon composed of rDBTPHE, σ = −kDBT, σKDBT, σ PDBTθDBTPHE, σ
three fused aromatic rings. For the hydrogenation of phen- (21)
anthrene, different networks have been proposed.39,53−56 rDBTPHE, π = −kDBT, πKDBT, π PDBTθDBTPHE, π (22)
Beltramone et al.39 and Schachtl et al.53 proposed two parallel
pathways. The first route is the hydrogenation of phenanthrene rBPPHE, π = −kBP, πKBP, π PBPθDBTPHE, π (23)
to 9,10-dihydrophenanthrene (DHPHE), and the second route
is the hydrogenation of phenanthrene to 1,2,3,4-tetrahydrophe- rPHE‐THPHE, π = −kPHE‐THPHE, πKPHE, π PPHEθDBTPHE, π (24)
nanthrene (THPHE). Subsequently, DHPHE and THPHE
are supposed to be hydrogenated to octahydrophenanthrene rPHE‐DHPHE, π = −kPHE‐DHPHE, πKPHE, π PPHEθDBTPHE, π (25)
(1,8-OHPHE and 1,10-OHPHE) and then to perhydrophenan-
threne (PHPHE). Conversely, Ishihara et al.54 proposed that rTHPHE, π = −k THPHE, πKTHPHE, π PTHPHEθDBTPHE, π (26)
phenanthrene is hydrogenated exclusively to DHPHE, from
where DHPHE is hydrogenated to THPHE and OHPHE, rDHPHE, π = −kDHPHE, πKDHPHE, π PDHPHEθDHPHE, π (27)
and finally to PHPHE. These two reaction networks are modeled 3.3. Estimation of Model Parameters. Kinetic parameters
herein, but the network proposing two parallel pathways were estimated by minimizing the objective function RSS(φ),
for the hydrogenation of phenanthrene in the presence of which includes the residual sum of squares of the concentration
DBT led to a better fit of the catalytic data. Details for the of the different species:57−59
other evaluated models are given in Section S4 of the nresp nexp
Supporting Information. Thus, in what follows, this reaction φ1, , φ2 , ..., φn
network and the corresponding kinetic model are presented. RSS(φ) = ∑ wn ∑ (Fk ,n − Fk̂ ,n)2 ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯→ min
n=1 k=1 (28)
Scheme 2c presents the reaction network for the hydrogenation
of phenanthrene. 1,8-octahydrophenanthrene and 1,10-octahy- where φ is the optimal parameter vector, nexp the number of
drophenanthrene were lumped together as octahydrophenan- independent experiments, nresp the number of responses, Fk,n the
threne. For the kinetic model, the following assumptions were molar flow of the nth experimental responses for the kth obser-
made: (i) the hydrogenation of phenanthrene, 9,10-dihydrophe- vation, and F̂k,n the molar flow of the nth predicted responses for
nanthrene, and 1,2,3,4-tetrahydrophenanthrene only occur on π the kth observation. In addition, wn is the weight factor assigned
sites; and (ii) the reversibility of the reactions was neglected. to the nth response.
The corresponding reaction mechanism is shown in Table 5. The subroutine VODE was used to solve the corresponding
set of ordinary differential equations (ODEs).60 The initial
Table 5. Reaction Mechanism and Catalytic Cycles for the minimization of the objective function, vide eq 28, in the model
Kinetic Model DBTPHE To Describe the Hydrogenation of regression was carried out using the Rosenbrock method.61 The
PHE ODRPACK subroutines then were called for fitting calculated
values to the corresponding experimental data point.62 These
step mechanism v1 v2 v3 v4
subroutines can perform either weight orthogonal distance regres-
A H 2 + 2β ⇄ 2Hβ 1 2 3 2 sion or nonlinear least-squares squares for explicit and implicit
B PHE + π ⇄ PHEπ 1 1 0 0 models using multiresponse data with an implementation of the
C PHEπ + 2Hβ → DHPHEπ + 2β 1 0 0 0
Levenberg−Marquard method.63
The reparameterization of Arrhenius and van’t Hoff expres-
D PHEπ + 4Hβ → THPHEπ + 4β 0 1 0 0 sions led to eqs 29 and 30, respectively. The resulting parameters
E DHPHEπ + 6Hβ → OHPHEπ + 6β 0 0 1 0 were used for the regression analysis. The activation energies and
F THPHEπ + 4Hβ → OHPHEπ + 4β 0 0 0 1 pre-exponential factors, as well as the adsorption enthalpies and
DHPHEπ ⇄ DHPHE + π
entropies, were calculated from the parameter estimation
G 1 0 0 0
procedure.
H THPHEπ ⇄ THPHE + π 0 1 0 0
OHPHEπ ⇄ H OHPHE + π
⎡ EA, i ⎛ 1 1 ⎟⎞⎤
I 0 0 1 1 ki = exp⎢Ai′ − ⎜ − ⎥
Global Reactions ⎣ R ⎝T T * ⎠⎦ (29)
1 PHE + H 2 → DHPHE 1 0 0 0
⎡ ΔS 0 ΔHn0 ⎛ 1 1 ⎟⎞⎤
2 PHE + 2H 2 → THPHE 0 1 0 0 K n = exp⎢ n − ⎜ − ⎥
3 DHPHE + 3H 2 → OHPHE 0 0 1 0 ⎣ R R ⎝T T * ⎠⎦ (30)
4 THPHE + 2H 2 → OHPHE 0 0 0 1 where, for the ith reaction, A′i is the natural logarithm of the pre-
exponential factor, EA,i the activation energy, T the reaction
Hydrogen and phenanthrene were assumed to adsorb on the temperature, T* the averaged reaction temperature, ΔS0n the
β and π sites, respectively (steps A and B). Steps C and standard adsorption entropy of component n, ΔH0n the standard
D represent the hydrogenation of phenanthrene to 9,10- adsorption enthalpy of component n, and R the universal gas
dihydrophenanthrene and 1,2,3,4-tetrahydrophenanthrene, constant.
respectively. The sequential hydrogenation to octahydrophenan- The significance of the overall regression for each model was
threne from 9,10-dihydrophenanthrene and 1,2,3,4-tetrahydro- tested after estimating the parameters of the model. This pro-
phenanthrene are represented in steps E and F, respectively. cedure was done by means of an F-test. The F-value for the global
Finally, the desorption of products is represented by significance of the regression was defined as the ratio of the
steps G−I. regression sum of squares to the residual sum of squares divided
3931 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Table 6. Conversion and Selectivity Observed under Different Conditions and Obtained By Evaluating the Conversion of DBT or
an Aromatic Compound Individually or in a Blenda
% Selectivity to DBT Products % Selectivity to PHE Products
feed WCatal.F0−1 %XDBT %SBP %SCHB %XNP %XFL %XPHE %STHPHE %SDHPHE %SOHPHE
DBT 46 33.4 87.9 12.1
91 57.9 87.5 12.5
122 70.8 86.1 13.9
182 88.7 85.1 14.9
NP 56 71.8
FL 73 12.7
PHE 73 77.0 19.0 28.1 52.9
DBT + NPb 56 72.2 86.7 13.3 38.1
28 70.1 86.5 13.5 29.2
DBT + FLb 73 65.3 89.3 10.7 4.5
37 58.6 89.1 10.9 3.4
DBT + PHEb 73 56.8 86.3 13.7 42.6 38.8 39.9 21.2
37 50.7 85.9 14.1 36.4 41.7 40.6 17.6
a
Reaction conditions: T = 300 °C, P = 5 MPa, liquid-flow rate of 30 mL h−1 and H2/liquid feed ratio of 500. bThe WcatF0−1 is given with respect to
the aromatic compound. WCatalFDBT0−1 is constant and equal to 122 kgcat kmolDBT−1 h.

by their respective degrees of freedom. The significance of the for the forward and backward reactions, respectively. Such a
individual parameters was evaluated by a t-test. Parity diagrams relationship is based on the potential energy diagram for the
were also built to visualize the agreement between experimental reaction.68
observations and model fit. The Bayesian information criterion
(BIC) was used to avoid model overfitting by introducing a 4. RESULTS
penalty for the number of parameters in the model.64−66 The 4.1. Experimental Evaluation. 4.1.1. Hydrodesulfurization
model leading to the lowest BIC was preferred among a finite set of Dibenzothiophene in the Absence and in the Presence of
of models. Aromatic Compounds. Table 6 displays the conversion of DBT
3.3.1. Physicochemical Tests on the Parameters of the and the selectivity to the reaction products as a function of space-
Models. Typically, testing the fitting of the rate equation to the time velocity; WcatFDBT0−1. During the HDS of DBT, only biphe-
experimental data and calculating the confidence intervals of the nyl and cyclohexylbenzene were detected. Partially hydrogenated
estimated parameters is considered to be sufficient when intermediates (i.e., THDBT and HHDBT) of the HYD pathway
developing a kinetic model. However, such a procedure does were only observed as traces. As expected, the conversion of
not provide physical meaning to the parameters of the model. DBT increased with space-time and the main reaction product
The kinetic and adsorption equilibrium constants contained in was biphenyl.11,15,27 The selectivity to the DDS route, that is, the
the rate expression must be evaluated to verify whether they are selectivity to biphenyl, decreased with space-time velocity, while
physically reasonable and thermodynamically consistent. Bou- the cyclohexylbenzene selectivity increased. This suggests that
dart et al.67 proposed several rules for evaluating kinetic param- biphenyl was probably hydrogenated to cyclohexylbenzene, as
eters that are described next. First, since adsorption is, with very has been mentioned by other authors14,27,69 and corroborated
few exceptions, exothermic, values for the estimated adsorption experimentally in this work, vide Section S5 of the Supporting
enthalpies must satisfy the following inequality: Information.
−ΔHn0 > 0 The HDS of DBT was studied in the presence of naphthalene,
(31)
fluorene, and phenanthrene at the same space time of DBT, i.e.,
Second, the adsorption entropy of the adsorbed species must 122 WcatFDBT0−1, but at different space times for the aromatic
be higher than zero and lower than the corresponding standard compounds. The conversion of DBT and the selectivity to
entropy of the corresponding species in the gas phase (ΔS0n,g): biphenyl and cyclohexylbenzene were not significantly affected
by the presence of naphthalene, since the standard deviation
0 < ΔSn0 < ΔSn0, g (32) calculated from these data was ca. 1.1%. On the other hand, the
The following limits for the adsorption entropy must be met: conversion of DBT decreased in the presence of fluorene and
phenanthrene. For example, at 37 WcatF0−1, DBT conversion
41.8 < −ΔSn0 < 51.04 − 1.4ΔHn0 (33) decreased from 70.8% to 58.6% and 50.7%, in the presence of
fluorene and phenanthrene, respectively. On the other hand, the
Finally, another thermodynamic criterion that must be con- HYD pathway of the HDS of DBT appeared to be diminished by
sidered is that the standard enthalpy of reaction must be equal to fluorene, since the selectivity to biphenyl increased from ca. 86%
the difference between the change in standard adsorption enthal- to 89%. However, phenanthrene did not seem to affect selec-
pies for reactants and products and the activation energies for the tivity. To this end, phenanthrene seems to inhibit the DDS and
forward and backward reactions (eq 34): HYD pathways to the same extent, while fluorene seems to
ΔHr° = (ΔHR° − ΔHP°) − (EaB − EaF) (34)
inhibit the HYD route stronger than the DDS route.
4.1.2. Hydrogenation of Naphthalene, Fluorene, And
where ΔH°r is the standard enthalpy of reaction, ΔH°R and ΔH°P Phenanthrene in the Absence and in the Presence of
are the standard adsorption enthalpies for reactants and Dibenzothiophene. Besides analyzing DBT reactivity in the
products, respectively, and EFa and EBa are the activation energies presence of the aromatics, it is also convenient to follow
3932 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Figure 1. Parity diagrams for comparing experimental with calculated conversion of DBT and the yields of BP and CHB, for the three kinetic models the
HDS of DBT: (a) DBT1S, (b) DBT2S, and (c) DBTBP2S. Conditions: p = 5 MPa, T = 260−300 °C, liquid flow rate of 30 mL h−1, and H2/liquid feed
ratio of 500.

aromatics reactivity in the absence and in the presence of DBT; pathwaysDBT2S (Figure 1b) and DBTBP2S (Figure 1c)
Table 6 shows the corresponding results. In the absence of DBT, fitted adequately the experimental observations within an error
aromatics conversion followed the order: %CPHE (77) > %CNP margin of <10%. Between DBT2S and DBTBP2S, the latter
(71.8) ≫ %CFL (12.7). Therefore, fluorene exhibited a much fitted the yield of cyclohexylbenzene better. The global
lower hydrogenation conversion than the other aromatics. The significance of the performed regressions for each model was
conversion of the aromatics was significantly affected by the assessed via calculation of the corresponding F and BIC values.
presence of DBT. The conversion of both naphthalene and Although all models led to a statistically significant regression;
phenanthrene decreased ca. 47% from the values observed in the i.e., F-value ≫ tabulated F-value ≈ 2.79, DBTBP2S presented the
absence of DBT. On the other hand, the conversion of fluorene largest F-value (29 274) and the lower BIC (1297), which
was affected more strongly as it decreased to ca. 65%. The selec- implies a better fitting of the experimental data.
tivity to phenanthrene products also varied in the presence of Considering the best fitting of the DBTBP2S model, the values
DBT. THPHE and DHPHE, i.e., the products with one hydroge- calculated for its parameters are presented in Table 7, together
nated aromatic ring, increased their selectivity significantly. with 95% probability confidence intervals and the corresponding
Particularly, selectivity to THPHE increased from 19% to 38.8%, t-values. Generally, all parameters were found to be statistically
whereas the selectivity to DHPHE increased from 28.1% to significant. Furthermore, the values of adsorption enthalpies and
39.9%. Meanwhile, the selectivity to OHPHE, the product with adsorption entropies on both σ and π sites satisfied Boudart
two hydrogenated aromatic rings, decreased ca. 60%. In conclu- et al.67 criteria (vide eqs 31−33). Therefore, values of −ΔH0i,σ and
sion, it is evident that the hydrogenation of naphthalene, − ΔH0i,π were indeed positive, as well as −ΔSi,σ 0
and −ΔSi,τ0
.
fluorene, and phenanthrene was strongly inhibited by the The latter were also lower than the corresponding standard
presence of DBT, which, in turn, affected the distribution of entropy of the gas-phase species, i.e., they were less than 316,
products. 384, and 205 J (mol K)−1 for DBT, BP, and H2S, respectively.
4.2. Kinetic Modeling. 4.2.1. Hydrodesulfurization of The second criterion for the adsorption entropy (eq 33) was also
Dibenzothiophene. Figure 1 depicts parity diagrams for the satisfied for values derived from the DBTBP2S model in all cases
HDS of DBT over the NiMo/γ-Al2O3, comparing the experi- except for H2S adsorption on π sites (−ΔS0H2S,τ), probably
mental results with the calculated conversion of DBT, and the because of poor adsorption of this compound on these sites.
yields of biphenyl and cyclohexylbenzene, for the three On the other hand, the activation energies obtained from the
developed kinetic models: DBT1S, DBT2S, and DBTBP2S. model were within the range proposed by Santacesaria,70 that is,
The DBT1S model, which assumes a single site for both HYD from 21 kJ mol−1 to ∼210 kJ mol−1. It is worth noting that
and DDS, showed the worst fitting (Figure 1a). Conversely, the activation energies of <21 kJ mol−1 suggest the presence of
models assuming two different sites for the HYD and DDS external diffusion limitations, whereas values of >210 kJ mol−1
3933 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Table 7. Values of the Kinetic Parameters and Corresponding naphthalene. The DBTNP kinetic model was found to predict
95% Probability Confidence Intervals for the Model the subtle effect of naphthalene on the HDS of DBT under the
DBTBP2S Used to Describe the Hydrodesulfurization of studied reaction conditions. Furthermore, the model was capable
Dibenzothiophenea of fitting experimental observations for the conversion of DBT
and for the production of BP, CHB, NP, and TTL with an error
estimated lower upper
parameter value limit limit t-value margin <10%. Similarly, the models DBTFL and DBTPHE
adequately fitted the experimental observation of the conversion
ADBT,σ (mmol (g h)−1) 0.533 0.531 0.536 458.6
ADBT,π (mmol (g h)−1) 0.179 0.177 0.180 299.7
of DBT and product yields, and also the conversion of fluorene
ABP,π (mmol (g h)−1) 0.012 0.012 0.013 104.4
and phenanthrene, and their respective product yields, as shown
Ea DBT,σ (kJ mol−1) 47.1 46.6 47.6 201.8
in the parity plots shown in Figures 3 and 4, respectively.
Ea DBT,π (kJ mol−1) 131.9 130.4 133.3 182.1
Ea BP,π (kJ mol−1) 162.4 159.7 165.2 115.5
−ΔS0DBT,σ (J (mol K)−1) 69.6 69.5 69.7 1259.0
−ΔS0BP,σ (J (mol K)−1) 89.5 88.3 90.7 141.8
−ΔS0CHB,σ (J (mol K)−1) 65.7 65.2 66.1 269.3
−ΔS0H2S,σ (J (mol K)−1) 126.1 119.6 132.5 38.6
−ΔS0DBT,π (J (mol K)−1) 105.8 105.2 106.3 363.6
−ΔS0BP,π (J (mol K)−1) 41.2 41.7 42.0 519.2
−ΔS0CHB,π (J (mol K)−1) 41.0 40.8 41.2 475.8
−ΔS0H2S,π (J (mol K)−1) 113.9 112.4 115.4 148.8
−ΔH0DBT,σ (kJ mol−1) 33.8 33.3 34.3 132.7
−ΔH0BP,σ (kJ mol−1) 34.8 34.2 35.4 114.0
−ΔH0CHB,σ (kJ mol−1) 18.2 17.9 18.5 117.4
−ΔH0H2S,σ (kJ mol−1) 203.1 198.3 207.8 84.6
−ΔH0DBT,π (kJ mol−1) 58.7 58.2 59.2 236.8 Figure 3. Parity diagram for comparing experimental with the calculated
−ΔH0BP,π (kJ mol−1) 42.0 41.4 42.6 143.4 conversion of DBT and FL, and the yields of their reaction products.
−ΔH0CHB,π (kJ mol−1) 2.6 2.5 2.6 147.7 The simulated values were calculated from the model DBTFL that
−ΔH0H2S,π (kJ mol−1) 1.4 1.4 1.5 63.1 describe the simultaneous HDS of DBT and hydrogenation of FL.
a Conditions: p = 5 MPa, T = 260−300 °C, liquid flow rate of 30 mL h−1,
Note: F-value = 29274. Ftab = 2.79, ttab = 1.963 at 1-α = 0.95 and
and H2/liquid feed ratio of 500.
922 degrees of freedom.

are related to the existence of thermal gradients during the


experimentation. To this end, activation energies obtained herein
are well-adjusted to intrinsic kinetic measurements.
4.2.2. Simultaneous HDS of DBT and Hydrogenation of
Aromatics. For the kinetic models DBTNP, DBTFL, and
DBTPHE both reactions are modeled simultaneous, i.e., the
HDS of DBT and the hydrogenation of respective aromatic
compound. Considering that DBTBP2S was the model that
better fitted the experimental results for the HDS of DBT, it was
chosen as a basis for further kinetic analysis of the HDS of DBT
in the presence of the aromatic compounds. Figure 2 presents the
parity plot for the reactions performed in the presence of

Figure 4. Parity diagram for comparing experimental with the calculated


conversion of DBT and PHE, and the yields of their reaction products.
The simulated values were calculated from the model DBTPHE that
describe the simultaneous HDS of DBT and hydrogenation of PHE.
Conditions: p = 5 MPa, T = 260−300 °C, liquid flow rate of 30 mL h−1,
and H2/liquid feed ratio of 500.

Generally, the developed kinetic models also fitted the results


for the hydrogenation of the studied aromatics. The BIC for the
global significance of the regressions amounted to 1401, 825, and
1499 for DBTNP, DBTFL, and DBTPHE, respectively, thus
evidencing a suitable statistical significance of these regressions.
Values for the parameters of DBT and its reaction products cor-
Figure 2. Parity diagram for comparing experimental with the calculated responding to the reactions under the presence of naphtha-
conversion of DBT and NP, and the yields of their reaction products. lene, phenanthrene, and fluorine are presented in Table 7.
The simulated values were calculated from the model DBTNP that Tables 8− 10 present the kinetic parameters related to the
describe the simultaneous HDS of DBT and hydrogenation of NP. studied aromatic compounds and its reaction products. The
Conditions: DBTNP and DBTFL models led to statistically significant
3934 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Table 8. Values of the Kinetic Parameters and Corresponding Table 10. Values of the Kinetic Parameters and
95% Probability Confidence Intervals for the Model DBTNP Corresponding 95% Probability Confidence Intervals for the
Used to Describe the HDS of DBT and Hydrogenation of NPa Model DBTPHE Used to Describe the Hydrodesulfurization
of Dibenzothiophene and Hydrogenation of PHEa
estimated lower upper
parameter value limit limit t-value estimated lower upper
AFNP,σ (mmol (g h)−1) 0.0067 0.0065 0.0069 70.6 parameter value limit limit t-value
AFNP,π (mmol (g h)−1) 0.0749 0.0743 0.0755 23.9 APHE‑THPHE,τ 0.9309 0.9004 0.9613 60.0
ARNP,σ (mmol (g h)−1) 0.1000 0.0972 0.1028 70.5 (mmol (g h)−1)
ARNP,π (mmol (g h)−1) 0.0094 0.0090 0.0098 44.1 APHE‑DHPHE,τ 1.7619 1.7280 1.7957 102.1
(mmol (g h)−1)
EFa NP,σ (kJ mol−1) 66.6 65.8 67.3 178.3
ADHPHE,τ (mmol (g h)−1) 3.8704 3.7774 3.9635 81.7
EFa NP,π (kJ mol−1) 50.7 50.1 51.3 161.2
ATHPHE,τ (mmol (g h)−1) 2.3959 2.1125 2.6793 16.6
ERa NP,σ (kJ mol−1) 97.8 93.9 101.6 49.8
Ea PHE‑THPHE,τ (kJ mol−1) 40.1 38.5 41.6 51.0
ERa NP,π (kJ mol−1) 107.8 105.2 110.5 79.6
Ea PHE‑DHPHE,τ (kJ mol−1) 30.0 29.1 30.9 69.0
−ΔS0NP,σ (J (mol K)−1) 126.1 125.0 127.3 210.6
Ea DHPHE,τ (kJ mol−1) 59.8 56.9 62.7 40.5
−ΔS0TTL,σ(J (mol K)−1) 108.8 106.6 111.0 96.8
Ea THPHE,τ (kJ mol−1) 40.0 37.8 42.2 35.8
−ΔS0NP,π(J (mol K)−1) 120.4 119.8 121.1 379.5
−ΔS0PHE,σ(J (mol K)−1) 104.8 95.7 114.0 22.5
−ΔS0TTL,π(J (mol K)−1) 73.0 72.5 73.5 305.5
−ΔS0DHPHE,σ (J (mol K)−1) 66.7 63.4 70.0 39.6
−ΔH0NP,σ (kJ mol−1) 31.6 31.0 32.2 100.4
−ΔS0THPHE,σ(J (mol K)−1) 70.5 52.1 88.9 7.5
−ΔH0TTL,σ(kJ mol−1) 28.5 27.4 29.5 51.8
−ΔS0PHE,τ (J (mol K)−1) 89.2 87.8 90.6 125.5
−ΔH0NP,π(kJ mol−1) 60.5 59.8 61.3 155.6
−ΔS0DHPHE,τ (J (mol K)−1) 63.4 61.5 65.3 65.2
−ΔH0TTL,π(kJ mol−1) 31.5 30.7 32.3 79.1
−ΔS0THPHE,τ (J (mol K)−1) 60.1 59.1 61.0 123.3
a
Note: F-value = 8848. Ftab = 2.79, ttab = 1.964 at 1 − α = 0.95 and −ΔH0PHE,σ (kJ mol−1) 78.2 59.3 97.1 8.1
560 degrees of freedom. −ΔH0DHPHE,σ (kJ mol−1) 23.8 21.4 26.1 19.7
−ΔH0THPHE,σ (kJ mol−1) 13.6 5.8 21.3 3.4
Table 9. Values of the Kinetic Parameters and Corresponding −ΔH0PHE,τ(kJ mol−1) 78.0 76.1 79.9 81.2
95% Probability Confidence Intervals for the Model DBTFL −ΔH0DHPHE,τ (kJ mol−1) 61.0 59.0 63.1 58.3
Used to Describe the Hydrodesulfurization of −ΔH0THPHE,τ(kJ mol−1) 73.3 68.9 77.6 32.8
Dibenzothiophene and Hydrogenation of FLa a
Note: = 9971. Ftab = 2.79, ttab = 1.963 at 1 − α = 0.95 and 792
estimated lower upper degrees of freedom.
parameter value limit limit t-value
AFL,π (mmol (g h)−1) 0.0792 0.0773 0.0787 232.7 studies have oversimplified modeling using mainly pseudo-first-
Ea FL,π (kJ mol−1) 76.5 77.3 78.7 179.8 order rate equations.11,16,19,33,72 This approach is capable of
−ΔS0FL,σ (J (mol K)−1) 94.8 94.2 95.3 340.7 fitting experimental observations well. However, pseudo-first-
−ΔS0HHFL,σ (J (mol K)−1) 65.7 65.2 66.2 272.2 order rate equations do not take into consideration elementary
−ΔS0FL,π (J (mol K)−1) 68.7 68.4 69.1 350.9 steps for the reaction mechanism and do not provide a descrip-
−ΔS0HHFL,π(J (mol K)−1) 42.0 41.3 42.7 126.0 tion of the adsorption−desorption steps for the different species
−ΔH0FL,σ (kJ mol−1) 89.7 86.5 93.0 53.7 on the catalyst.73,74 In order to understand the catalytic surface
−ΔH0HHFL,σ (kJ mol−1) 11.5 11.3 11.7 114.1 reactions for the HDS of DBT, some mechanisms have been
−ΔH0FL,π (kJ mol−1) 71.5 69.2 73.8 60.3 evaluated using the Langmuir−Hinshelwood-Hougen-Watson
−ΔH0HHFL,π (kJ mol−1) 69.4 66.7 72.2 49.8 (LHHW) model, as shown in Table 11. Singhal et al.14 and
a
Note: F-value = 42985. Ftab = 2.85, ttab = 1.966 at 1 − α = 0.95 and Vrinat18 modeled the HDS of DBT considering that the DDS
422 degrees of freedom. and HYD pathways occur on the same catalytic site. Meanwhile,
Broderick et al.26 and Vanrysselberghe et al.27 assumed that the
regressions (vide Tables 8 and 9, respectively), since all sites involved in hydrogenation and desulfurization are different.
parameters presented narrow confidence intervals and t-values All of them, using one or two sites for the adsorption of DBT and
larger than the tabulated t-values. The parameters estimated for its products on the catalyst gave a good fit to the experimental
the DBTPHE model, their corresponding 95% confidence data. Nevertheless, most of these kinetic models lack a proper
intervals, and the calculated t-values are shown in Table 10. The statistical and phenomenological analysis of the model itself and
thermodynamic consistency of these parameters was also ver- the estimated kinetic parameters. Kinetic parameters should be
ified, except for those related to the production and adsorption of not only capable of representing the physicochemical phenom-
OHPHE. This behavior is typically considered to originate from ena that occur adequately but also exhibiting statistical signif-
a relatively weak chemisorption of the molecule, in comparison icance and thermodynamic consistency.44
to other compounds.59 Adsorption equilibrium constants and 5.1.2. Nature of the Active Sites. As already discussed in the
rate coefficients at 280 °C for the reactions of HDS of DBT, Results section, among the three kinetic models developed in this
hydrogenation of naphthalene, fluorene, and phenanthrene are work, only those considering two active sitesone for
presented in the Supporting Information (Section S6). hydrogenolysis and the other for hydrogenationwere able to
properly describe experimental results. Few kinetic models for
5. DISCUSSION the HDS of DBT have been developed assuming two different
5.1. Kinetics of the Hydrodesulfurization of Dibenzo- sites for the adsorption of DBT and its products.25−27 The
thiophene. 5.1.1. Comparison of the Kinetic Model Presented development of this type of kinetic model does not seek to give
Herein with Previous Works. Although the HDS of DBT has information on the chemical structure of the active sites but
been thoroughly studied in the past,10,71 the majority of kinetic characterizes reaction rates relating them to both a macroscopic
3935 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Table 11. Kinetic Expressions for the Hydrodesulfurization of Dibenzothiophene, Based on the Langmuir−Hinshelwood−
Hougen−Watson Approach

Catalytic Sites

DDS HID activation energy


catalyst route route H2 rate expression (kJ mol−1) ref
kKDBTPDBT K H2PH2
rDBT = ·
1 + KDBTPDBT + KPRODPPROD 1 + K H2PH2 163.3

CoMo/γ-Al2O3 × × Singhal et al.14


kKBPPBP K H2PH2
rBP = ·
1 + KBPPBP + K CHBPCHB 1 + K H2PH2 210.6

kKDBTPDBT K H2PH2
rDBT = ·
CoMo/γ-Al2O3 × × 1 + KDBTPDBT + K H2SPH2S 1 + K H2PH2 96.3 Vrinat18

kDBT, σKH, σ KDBT, σ C H2C DBT


rDBTσ =
(1 + (KH, σ C H2)0.5 + KDBT, σ C DBT + KBP, σ C BP + K H2S, σ C H2S)3 122

kDBT, πKH, π KDBT, π C H2C DBT


rDBTπ =
CoMo/γ-Al2O3 × × (1 + (KH, π C H2)0.5 + KDBT, π C DBT + KBP, π C BP)3 186 Vanrysselberghe et al.27

kBP, πKH, π KBP, π C H2C BP


rBPπ =
(1 + (KH, π C H2)0.5 + KDBT, π C DBT + KBP, π C BP)3 255.7

kDBT, σKDBT, σ C DBT K H2C H2


rDBTσ = ·
(1 + KDBT, σ C DBT + K H2S, σ C H2S)2 1 + K H2C H2 125.6

CoMo/γ-Al2O3 × × × Broderick et al.26


kDBT, πKDBT, π K H2C DBTC H2
rDBTπ = 117.2
1 + KDBT, π CDDBT

reaction mechanism and macroscopic process variables such as flat on hydrogenation sites. Therefore, π-flat adsorption of DBT
concentration, pressure, and temperature. Nevertheless, several and of other aromatic structures is feasible on brim sites.
theoretical studies and DFT calculations have been carried out in As mentioned earlier, this work is not aimed at determining the
recent years aiming to elucidate the atomic structure and nature nature of the active sites. Nevertheless, based on the literature,
of the active sites of hydrotreatment catalysts.24,75−78 Most and as a matter of speculation, the developed models hint that
authors accept that coordinately unsaturated sites (CUS) (sulfur σ and π sites might well correspond to CUS and brim sites,
vacancy sites) at the edges and corners of molybdenum sulfide respectively. Consequently the results obtained in this work can
(MoS2) play an important role for the DDS route. In this respect, be interpreted considering that DBT would preferably be
the MoS2 particles expose two types of edges: Mo-edge and adsorbed on corner CUS sites of a mixed NiMoS phase.
S-edge. Under typical sulfidation conditions, Ni atoms might Experimental evidence shows that, in these sites,84 the S atom
replace the Mo atoms from the Mo-edge to form the so-called from DBT might replace the S atom from the CUS, hence
NiMoS phase, hence promoting the formation of CUS at the leading to direct hydrogenolysis for producing biphenyl.
edges and corners, because of the reduction of the sulfur binding Furthermore, and provided the strongest kinetic viability of the
energy.79,80 The desulfurization of DBTs via DDS would model including the subsequent hydrogenation of biphenyl to
preferably occur on vacancies at corner sites via perpendicular cyclohexylbenzene, the former product should readsorb on
adsorption through the S atom.20,24,79,81 On the other hand, the brim sites. On the other hand, the HYD route might be accounted
nature of the sites for the HYD pathway remains yet unclarified. to proceed via the adsorption of DBT in a π-flat bonding between
Topsøe and co-workers,22−24,77 based on scanning tunneling the benzene ring and the brim site, where partially hydrogenated
microscopy (STM) images of the top of MoS2 and CoMoS slabs, intermediates are to be produced. These intermediates would
proposed that the so-called brim sites are involved in hydro- desorb to react on CUS sites for the C−S−C bond scission step.
genation reactions. Brim sites are modeled as bridge sites This proposition is summarized in Scheme 3. It is interesting to
presenting a metallic character and located at the cluster top and remark that the mechanism postulated herein resembles previous
perimeter sites along the cluster edges. They are supposed to not considerations for the reactivity of DBT over noble-metal-based
be sterically hindered, because the adsorption of aromatic catalysts.37,46,85
molecules does not require the formation of sulfur vacancies.82,83 5.1.3. Considerations about the Reaction Mechanism.
As expected, on brim sites, the adsorption of H2S is negligibly Given the scarcity of literature reports dealing with kinetic stud-
weak, explaining the low inhibition effect of H2S on the HYD ies over NiMo catalysts, it was necessary to compare the present
pathway.82 Moreover, it is often theorized that aromatics adsorb results qualitatively with reports for CoMo catalysts26,27 and
3936 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

Scheme 3. Illustration of the Adsorption of DBT on the Two established between the sulfur atom (η1S), thiophene (η5), and
Routes of HDS on NiMoS/γ-Al2O3a the aromatic rings (η6) of DBT with the corresponding active
sites of the catalyst. It is worth stressing that these theoretical
analyses are in agreement with our calculated values, which indi-
cate that the adsorption enthalpy of DBT is slightly larger on π
hydrogenation sites than on σ hydrogenolysis sites. To this respect,
the adsorption enthalpy for DBT on π and σ sites reported in the
literature agrees with values estimated herein.26−28 For biphenyl,
unlike that observed for cyclohexylbenzene, it was found that this
product adsorbs more strongly on π sites: −ΔH0BP,π = 42 kJ mol−1.
Literature reports for adsorption enthalpies of biphenyl over
HDS catalysts range from 38 kJ mol−1 to 50 kJ mol−1.27,28 For
cyclohexylbenzene, the relatively low values of adsorption
enthalpy on π sites point to a high mobility under the conditions
a
The rectangular box represents brim regions, and the oval marks the of the reaction atmosphere. Unfortunately, values derived from
CUS-like sites. Color scheme: green, Ni; blue, Mo; yellow, S. the present study could not be compared to theoretical or
experimental data from the literature, since they were not found
with theoretical studies performed via molecular simula- in our survey.
tion.24,76,78,82,83 According to Table 7, activation energies for Finally, with regard to H2S, values calculated for its adsorp-
DBT on σ and π sites and of its DDS product over π sites tion entropy and enthalpy on σ sites, 126 J (mol K)−1 and
followed the trend: Ea BP,π > Ea DBT,π ≫ Ea DBT,σ. This trend is in 203 kJ mol−1, respectively, were in good agreement with liter-
good agreement with literature reports (vide Table 11). As an ature.27,51,59 Furthermore, it has been postulated11,16,76,87,88 that
exception, Broderick et al.26 found that Ea DBT,σ > Ea DBT,π. this compound selectively inhibits σ sites. In contrast, a low value
However, their report may be criticized because they raised the of enthalpy (1.24 kJ mol−1) was found for the adsorption H2S on
denominator of their rate expression for hydrogenolysis to the π sites, suggesting a weak adsorption on brim sites. In this sense,
square power without providing a justification. Furthermore, Lauritsen et al.76 using high-resolution STM studies on hydro-
these authors did not check their results for thermodynamic treating HDS model systems, claimed that H2S does not compete
consistency. The order found for activation energies coincides for adsorption on brim sites. In addition, other reports11,16,87 have
with the high selectivity of NiMo sulfides to the DDS route. presented evidence of an increment of the selectivity to partially
In addition, the highest value found for Ea BP,π also agrees with the hydrogenated intermediates of DBT and 4,6 DMDBT with the
experimental evidence on the low rate of the hydrogenation of partial pressure of H2S. This agrees with the mechanism
biphenyl to cyclohexylbenzene. Theoretical studies have proposed for the HDS of DBT in Table 2. On the HYD
determined that this reaction is hindered by the fact that the pathway, the molecules are first hydrogenated on π sites and then
two phenyl rings of biphenyl are not coplanar, making its moved to σ sites for the C−S−C bond scission step. Therefore,
adsorption on π sites difficult.69 H2S inhibits both the DDS pathway and the final desulfurization
On the other hand, calculated entropy values (Table 7) step in the HYD pathway, generating an increase in the concen-
suggest that the mobility of DBT on σ sites is greater than on tration of the partially hydrogenated intermediates of DBTs.
π sites. Conversely, biphenyl, cyclohexylbenzene, and H2S might Thermodynamic calculations performed from the kinetic
present a higher mobility on π sites. Rangarajan et al.,82 using modeling of this contribution are deemed relevant for developing
density functional theory (DFT) for evaluating the adsorption of a better understanding of the mechanism of HDS.
different hydrocarbons on a CoMoS formulation, reported an 5.2. Kinetics of the Hydrodesulfurization of Dibenzo-
apparent opposite result to the obtained herein (i.e., DBT and its thiophene and Hydrogenation of Aromatics. 5.2.1. Simul-
alkyl-substituted aromatics presented a larger entropy when taneous Hydrodesulfurization of Dibenzothiophene and
adsorbed on brim sites than on CUS). Nevertheless, the authors Hydrogenation of Naphthalene. 5.2.1.1. Reactivity of
did not probe the adsorption of DBT on a CUS located in corner Dibenzothiophene and Naphthalene. Results showed that
sites. In fact, in a later work, these authors78 considered the the presence of naphthalene did not modify the reactivity of DBT
adsorption of DBT on these corner-type sites and suggested that over the tested sulfided NiMo/γ-Al2O3 catalyst. However, the
C−S−C bond scission is more feasible to occur on them. Tuxen conversion of naphthalene significantly decreased under the
et al.24 carried out a similar study but used scanning tunneling presence of DBT. From the different models evaluated for this
microscopy to investigate the atomic-scale adsorption of DBT simultaneous reaction, the one assuming that the hydrogenation
and 4,6-DMDBT on MoS2 and CoMoS nanoclusters. These of naphthalene can be carried out on both σ and π sites showed
authors, based on chemisorption calculations, also elucidated the the best fitting of the experimental data. In this regard, the fact
preference of DBT to adsorb on corner-type sites. According to that the DDS/HYD selectivity was not affected suggests that
Ding et al.,80 the Co atom on a corner exhibits a square-planar- naphthalene does not have a preferential adsorption site.
coordinated structure with four S atoms, which allow for the high Considering the latter, the values of the Arrhenius and van’t
mobility of DBT and the high hydrogenolysis activity. Given the Hoff parameters for the selected kinetic model were fixed for
values of entropy calculated herein, the same kind of explanation DBT and its reaction products (vide Table 7), whereas the
could apply to Ni-promoted sulfides. corresponding values for naphthalene and tetralin were
Concerning the calculated values for adsorption enthalpies, estimated from the model (vide Table 8). In this instance, the
Yang et al.86 observed that a flat π-bonding on sites related to calculated values for entropy indicate that naphthalene and
HYD route leads to larger values of this thermodynamic param- tetralin might present more mobility on π sites than on σ sites.
eter, compared to the perpendicular adsorption on σ sites. The Conversely, entropy values for tetralin suggest a larger mobility
authors related this observation to the type of interactions of this compound, compared to naphthalene at both types of
3937 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

active sites. Such a result is reasonable, considering their dif- fluorene and its partially hydrogenated product, hexahydro-
ference in aromaticity. The resonance energy of the aromatic ring fluorene, adsorb on both π and σ sites, but that fluorene hydro-
of tetralin is greater than that of the naphthalene ring, since the genation can only be performed on π sites. On the other hand,
latter is first hydrogenated and has lower aromaticity than the the activation energy for the hydrogenation of fluorene
former.50,89 On the other hand, the adsorption enthalpy of amounted to 77 kJ mol−1, which is a value similar to that
naphthalene and tetralin on π sites was found to be larger than reported by Lapinas et al.,52 using a NiW/Al2O3 catalyst.
that on σ sites. This might be related to the fact that the π-flat Values for The chemisorption entropies indicate that fluorene
adsorption of naphthalene and tetralin on π sites involves has higher mobility on π sites than on σ sites. The low conversion
multipoint interactions, so adsorption would be stronger than on observed for fluorene is related to the strong adsorption of this
σ sites, where η6 is theorized to occur.50 Although, under the molecule on σ sites, since it cannot be hydrogenated therein.
reaction conditions of this work, no significant impact of Besides, HHFL presents a higher chemisorption enthalpy on
naphthalene on the HDS of DBT was reported, other authors π sites than on σ sites, indicating that HHFL produced on π sites
have reported otherwise. Egorova et al.40 investigated the mainly contributes to the inhibition of HDS of DBT on these
inhibition effect of naphthalene on the HDS of DBT and 4,6- sites.
DMDBT and found that naphthalene may inhibit the HDS of 5.2.3. Simultaneous Hydrodesulfurization of Dibenzothio-
DBT and 4,6-DMDBT under conditions in which both the DDS phene and Hydrogenation of Phenanthrene. 5.2.3.1. Reac-
and the HYD pathways were affected to the same extent. tivity of Dibenzothiophene and Phenanthrene. Phenanthrene
Consequently, these authors proposed that the hydrogenation of had the strongest inhibition effect on DBT HDS. In addition, this
naphthalene occurred on both DDS and HYD sites. Conversely, molecule affected both the DDS and the HYD pathways to the
Egorova et al.40 also reported that the hydrogenation of same extent. As in the case of the other aromatics, phenanthrene
naphthalene was inhibited by the presence of both organo-sulfur conversion was also inhibited by DBT. The hydrogenation of
compounds. phenanthrene has been studied by several authors, and different
5.2.1.2. Considerations about the DBTNP Reaction reaction networks have been proposed.53−56 The product
Mechanism. Some studies have elucidated how hydrogenation distribution for the hydrogenation of phenanthrene, in the
of small molecules such as naphthalene may occur on more than absence of DBT obtained herein, was similar to that reported in
one type of catalytic sites. Gutiérrez et al.90 proposed that the size literature.53,56 Particularly, a high production of aromatics with
of the molecules should dictate their accessibility to hydro- two hydrogenated aromatic rings were detected, namely,
genolysis CUS sites. They speculated that the hydrogenation of 1,8-OHPHE and 1,10-OHPHE. Schachtl et al.53 reported similar
o-propylaniline, a molecule with a single aromatic ring, occurs on results for Ni-promoted MoS2/ γ-Al2O3 catalysts. Authors found
both CUS and brim sites, in view of their weak effect on the that Ni favored the adsorption of phenanthrene leading to the
selectivities of hydrodenitrogenation and HDS reactions. production of mainly 1,8-OHPHE and 1,10-OHPHE. On the
Moreover, other authors have associated CUS sites not only other hand, the addition of DBT to the reaction feed both
with hydrogenolysis but also with hydrogenation reactions.51,91 decreased conversion and shifted selectivity to less OHPHE and
Kinetic studies38,49,51,92,93 published for the hydrogenation of more DHPHE and THPHE. Such an increase, particularly in the
naphthalene in the absence and presence of other compounds selectivity to THPHE, agrees with suggestions made by Schachtl
assumed that naphthalene is only adsorbed and hydrogenated on et al.,53 who postulated that the hydrogenation of the two
one type of active site. Cortés et al.51 developed a LHHW kinetic aromatic rings of phenanthrene essentially occurred via the
model for the hydrogenation of naphthalene over a NiMo/ THPHE intermediate.
γ-Al2O3, catalyst assuming that the chemisorption of naphthalene 5.2.3.2. Considerations about the DBTPHE Reaction
and tetralin only occurred on CUS sites. Since there is a lack of Mechanism. The kinetic model that best fitted the experimental
information related to kinetic parameters on brim sites, a proper values for the HDS of DBT and for the hydrogenation of
comparison of our results with those from literature is not phenanthrene assumed that both phenanthrene and its reaction
performed. However, our results are in agreement with those products adsorb on both σ and π sites, but only react on π sites.
reported elsewhere.50,51,92 The adsorption entropies of phenanthrene and its reaction
5.2.2. Simultaneous Hydrodesulfurization of Dibenzothio- products (Table 10) elucidated that aromatic molecules have a
phene and Hydrogenation of Fluorene. 5.2.2.1. Reactivity of higher mobility on π sites than on σ sites. At π sites, the molecule
Dibenzothiophene and Fluorene. The presence of fluorene mobility increased as the molecule became hydrogenated: PHE
affected the conversion of DBT and mainly inhibited the HYD (−ΔS0PHE,π = 89.2 J (mol K)−1) < DHPHE (−ΔS0DHPHE,π = 63.4 J
pathway during the HDS reaction. There are few reports on this (mol K)−1) < THPHE (−ΔS0THPHE,π = 60.1 J (mol K)−1). On the
effect in open literature. Koltai et al.34 reported that, among other hand, the chemisorption enthalpy of phenanthrene is
anthracene, phenanthrene, and fluorene, the latter had the similar on σ sites (78.2 kJ mol−1) and π sites (78.0 kJ mol−1). This
strongest inhibiting effect on the transformation of 4,6-DMDBT. result is in agreement with the large inhibition effect of this
Authors ascribed this trend to the structural similarity between molecule during the HDS of DBT. However, the selectivity to
4,6-DMDBT and fluorene. Unlike DBT, 4,6-DMDBT princi- HYD and DDS pathways were affected to the same extent.
pally reacts via the HYD route. They argued that both molecules According to results, phenanthrene would only be hydrogenated
were adsorbed in the same manner over the catalytic active sites, on π sites, probably because the π sites are clearly less sterically
noting a preference for hydrogenation sites. On the other hand, hindered.76 Regarding enthalpies for the primary products on π
the very low conversion of fluorene, compared with the other sites, THPHE showed a higher value than DHPHE. On σ sites,
aromatic compounds, was also reported by Koltai et al.34 results were in contraposition nonetheless, since the chem-
5.2.2.2. Considerations about the Reaction Mechanism isorption enthalpy for DHPHE was higher than that found for
DBTFL. The kinetic model proposed in this work effectively THPHE. Concerning σ sites, the geometrical configuration of
accounted for the effect of fluorene on the reactivity of DBT. the aromatic rings of phenanthrene dictate their accessibility.
Furthermore, results from the model justify the assumption that DHPHE and THPHE present their aromatic ring in different
3938 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

positions. According to Beltramone et al.,39 the central ring of the to all the kinetic models evaluated to support the selection
THPHE presents a steric constraint that limits its adsorption. of the models avoiding possible overfitting; surface
This might be related to the larger adsorption enthalpy of coverage of the σ- and π-sites on the reactions of hydrode-
DHPHE than that of THPHE on σ sites. Finally, it is worth men- sulfurization (HDS) of dibenzothiophene (DBT) in the
tioning that the chemisorption entropy and enthalpy of OHPHE absence and in the presence of the aromatic compounds;
was not found to be statistically significant, so a discussion on this adsorption equilibrium constants and rate coefficients at
parameter is not possible herein. Such an outcome might be 280 °C for the reactions of hydrodesulfurization of DBT,
related to the lumping procedure carried out; i.e., 1,8-OHPHE and hydrogenation of naphthalene, fluorene, and phenan-
and 1,10-OHPHE were lumped together as OHPHE. threne (PDF)

6. CONCLUSIONS
A kinetic study of the hydrodesulfurization of dibenzothiophene
■ AUTHOR INFORMATION
Corresponding Authors
was undertaken over a NiMo/γ-Al2O3 catalyst, in the absence *E-mail: coca@xanum.uam.mx (C. O. Castillo-Araiza).
and in the presence of aromatic compounds with different *E-mail: vicbaldo@uis.edu.co (V. G. Baldovino-Medrano).
chemical structures: naphthalene, phenanthrene, and fluorene. ORCID
Among three proposed models for the HDS of DBT in the
absence of the aromatics, the one considering two different types
Edgar M. Morales-Valencia: 0000-0003-3401-2811
of active sitesone for hydrogenation and another for C−S−C Víctor G. Baldovino-Medrano: 0000-0003-3227-0251
bond scissionand considering the subsequent hydrogenation Notes
of biphenyl to cyclohexylbenzene was found to best-fit the The authors declare no competing financial interest.
observations. The developed model presented both physico-
chemical and statistical significance. With respect to the nature of
the HYD and DDS mechanisms, σ and π sites proposed in this
■ ACKNOWLEDGMENTS
This work was possible thanks to the financial support of UIS-
work for performing the hydrogenolysis and hydrogenation VIE. Edgar M. Morales-Valencia thanks COLCIENCIAS for
reactions, might be interpreted to correspond CUS and brim sites, his Ph.D. fellowship. Authors thank K. Gómez, F. Gómez,
respectively. Under such a consideration, the DDS pathway J. Rodríguez, and Y. Barrera for their collaboration on performing
would mainly occur on CUS sites. Meanwhile, for the HYD path- some of the featured experiments. We are grateful to the
way, the molecules would be first hydrogenated on brim sites and reviewers for their insightful comments and valuable suggestions.
then move to a CUS sites for the hydrogenolysis of C−S−C We dedicate this paper to the loving memory of Professor
bond. On the other hand, the weak chemisorption of H2S on Aristóbulo Centeno.
π sites agreed with the hypothesis that the adsorption of H2S
occurs only on CUS sites, hence mainly inhibiting the
hydrogenolysis reaction.
■ NOMENCLATURE
Roman Letters
The inhibition effect of the polyaromatic compounds on the
HDS of DBT was found to be related to the size of the molecules. Ai′ = natural logarithm of pre-exponential factor, mm (g h)−1
Phenanthrene was the molecule that most affected the Pn = partial pressure of component n
conversion of DBT and naphthalene was the one that affected EA = activation energy, kJ mol−1
it the least. However, regardless of size, the model predicted that F0n = inlet molar flow rate of the component n, mmol h−1
these two molecules did not have a preference for being adsorbed Fn = molar flow rate of the component n, mmol h−1
on a specific site, hence inhibiting both the DDS and HYD kFn = reaction rate coefficients of the forward
pathways to the same extend. Conversely, fluorene was found to kBn = reaction rate coefficient of the backward
mostly inhibit the HYD pathway. Such a trend was explained by rn = specific reaction rate of reaction n, mmol (g h)−1
considering its structure and geometry. The similarity between R = universal gas constant, kJ (mol K)−1
DBT and fluorene and the results from kinetic modeling suggest RSS = objective function
that the two molecules would be adsorbed in the same manner Sn = selectivity to component n, %
on the π hydrogenation sites. T = temperature, K
Finally, regarding the hydrogenation of the tested aromatics, Yn = yield of component n, %
the model predicted that these compounds are adsorbed on both wn = objective function weight factor of each response
σ and π sites. However, it was postulated that the size of the W = catalyst weight, g
molecules dictates their accessibility to σ sites to be hydro- Greek Letters
genated. In general, the kinetic models that best-represented the * = active sites for the hydrogenation and hydrogenolysis
experimental observations of the simultaneous HDS of DBT and reactions
hydrogenation of aromatics considered that naphthalene is β = active sites for the H2 adsorption
hydrogenated on both σ and π sites, whereas fluorene and phen- ΔH0n = standard enthalpy of adsorption for component n,
anthrene are hydrogenated only on π sites. kJ mol−1


ΔH°r = standard enthalpy of reaction, kJ mol−1
ASSOCIATED CONTENT ΔS0n = standard entropy of adsorption for component n,
J (mol K)−1
*
S Supporting Information
ΔS0n = standard entropy of adsorption for component n used
The Supporting Information is available free of charge on the in eq 30, kJ (mol K)−1
ACS Publications website at DOI: 10.1021/acscatal.8b00629. φ = vector of parameters accounted for in the objective
Description of the reaction systems used in the catalytic function
tests, information and results of the statistical test applied π = active sites for the hydrogenation reactions
3939 DOI: 10.1021/acscatal.8b00629
ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

σ = active sites for the hydrogenolysis reactions (17) Mijoin, J.; Pérot, G.; Bataille, F.; Lemberton, J. L.; Breysse, M.;
θi = fraction coverage of vacant sites on i sites Kasztelan, S. Mechanistic Considerations on the Involvement of
vj = Horiuti number Dihydrointermediates in the Hydrodesulfurization of Dibenzothio-
phene-Type Compounds over Molybdenum Sulfide Catalysts. Catal.
Subscripts Lett. 2001, 71, 139−145.
cat = catalyst (18) Vrinat, M. L. The Kinetics of the Hydrodesulfurization Process
exp = experiment A Review. Appl. Catal. 1983, 6, 137−158.
g = gas phase (19) Wang, Y.; Sun, Z.; Wang, A.; Ruan, L.; Lu, M.; Ren, J.; Li, X.; Li,
n = component n C.; Hu, Y.; Yao, P. Kinetics of Hydrodesulfurization of Dibenzothio-
obs = observed phene Catalyzed by Sulfided Co−Mo/MCM-41. Ind. Eng. Chem. Res.
tab = tabulated 2004, 43, 2324−2329.
(20) Cristol, S.; Paul, J.-F.; Payen, E.; Bougeard, D.; Hutschka, F.;
Superscripts/Accent Symbols
Clémendot, S. DBT Derivatives Adsorption over Molybdenum Sulfide
∧ = calculated Catalysts: A Theoretical Study. J. Catal. 2004, 224, 138−147.
° = inlet, standard


(21) Valencia, D.; Peña, L.; García-Cruz, I. Reaction Mechanism of
Hydrogenation and Direct Desulfurization Routes of Dibenzothio-
REFERENCES phene-like Compounds: A Density Functional Theory Study. Int. J.
(1) European Environment Agency. Air pollution. Available via the Quantum Chem. 2012, 112, 3599−3605.
Internet at: http://www.eea.europa.eu/themes/air/intro (accessed (22) Besenbacher, F.; Brorson, M.; Clausen, B. S.; Helveg, S.;
Dec. 16, 2016). Hinnemann, B.; Kibsgaard, J.; Lauritsen, J. V.; Moses, P. G.; Nørskov, J.
(2) Chevron Corporation. Diesel Fuel Technical Review. Available via K.; Topsøe, H. Recent STM, DFT and HAADF-STEM Studies of
the Internet at: http://www.chevronwithtechron.ca/products/diesel. Sulfide-Based Hydrotreating Catalysts: Insight into Mechanistic,
aspx (accessed June 16, 2017). Structural and Particle Size Effects. Catal. Today 2008, 130, 86−96.
(3) European Commission. Legislation: Transport and Environment. (23) Moses, P. G.; Hinnemann, B.; Topsøe, H.; Nørskov, J. K. The
Available via the Internet at: http://ec.europa.eu/environment/air/ Hydrogenation and Direct Desulfurization Reaction Pathway in
legis.htm#transportb (accessed June 16, 2017). Thiophene Hydrodesulfurization over MoS2 Catalysts at Realistic
(4) United States Environmental Protection Agency. Diesel Fuel Conditions: A Density Functional Study. J. Catal. 2007, 248, 188−203.
Standards and Rulemakings. Available via the Internet at: https://www. (24) Tuxen, A. K.; Füchtbauer, H. G.; Temel, B.; Hinnemann, B.;
epa.gov/diesel-fuel-standards/diesel-fuel-standards-and-rulemakings Topsøe, H.; Knudsen, K. G.; Besenbacher, F.; Lauritsen, J. V. Atomic-
(accessed June 16, 2017). Scale Insight into Adsorption of Sterically Hindered Dibenzothiophenes
(5) Tani, M. Japan’s Environmental Policy. Available via the Internet at: on MoS2 and Co-Mo-S Hydrotreating Catalysts. J. Catal. 2012, 295,
http://www.rieti.go.jp/en/special/policy-update/059.html (accessed 146−154.
June 16, 2016). (25) Farag, H. Kinetic Analysis of the Hydrodesulfurization of
(6) Fuels: Brazil Automotive Diesel Fuel. Available via the Internet at: Dibenzothiophene: Approach Solution to the Reaction Network. Energy
https://www.dieselnet.com/standards/br/fuel_automotive.php#anp_ Fuels 2006, 20, 1815−1821.
2009 (accessed June 16, 2017). (26) Broderick, D. H.; Gates, B. C. Hydrogenolysis and Hydrogenation
(7) Van Looij, F.; van der Laan, P.; Stork, W. H. J.; DiCamillo, D. J.; of Dibenzothiophene Catalyzed by Sulfided CoO-MoO3/γ-Al2O3: The
Swain, J. Key Parameters in Deep Hydrodesulfurization of Diesel Fuel. Reaction Kinetics. AIChE J. 1981, 27, 663−673.
Appl. Catal., A 1998, 170, 1−12. (27) Vanrysselberghe, V.; Froment, G. F. Hydrodesulfurization of
(8) Whitehurst, D. D.; Isoda, T.; Mochida, I. Present State of the Art Dibenzothiophene on a CoMo/Al2O3 Catalyst: Reaction Network and
and Future Challenges in the Hydrodesulfurization of Polyaromatic Kinetics. Ind. Eng. Chem. Res. 1996, 35, 3311−3318.
Sulfur Compounds. Adv. Catal. 1998, 42, 345−471. (28) Edvinsson, R.; Irandoust, S. Hydrodesulfurization of Dibenzo-
(9) Chandak, N.; Al Hamadi, A.; Yousef, M.; Mohamed, A.; Inamura, thiophene in a Monolithic Catalyst Reactor. Ind. Eng. Chem. Res. 1993,
K.; Berthod, M. A Pilot Reactor Study to Determine Operational Factors
32, 391−395.
of the Commercial Hydrodesulphurization (HDS) Catalyst to Produce
(29) Ho, T. C.; Sobel, J. E. Kinetics of Dibenzothiophene
Ultra-Low Sulphur Diesel (ULSD). Fuel 2014, 138, 37−44.
Hydrodesulfurization. J. Catal. 1991, 128, 581−584.
(10) Houalla, M.; Nag, N. K.; Sapre, A. V.; Broderick, D. H.; Gates, B.
(30) Choi, K.-H.; Kunisada, N.; Korai, Y.; Mochida, I.; Nakano, K.
C. Hydrodesulfurization of Dibenzothiophene Catalyzed by Sulfided
Facile Ultra-Deep Desulfurization of Gas Oil through Two-Stage or
CoO−MoO3/γ-Al2O3: The Reaction Network. AIChE J. 1978, 24,
1015−1021. -Layer Catalyst Bed. Catal. Today 2003, 86, 277−286.
(11) Egorova, M.; Prins, R. Hydrodesulfurization of Dibenzothiophene (31) Sahu, R.; Song, B. J.; Im, J. S.; Jeon, Y.-P.; Lee, C. W. A Review of
and 4,6-Dimethyldibenzothiophene over Sulfided NiMo/γ-Al2O3, Recent Advances in Catalytic Hydrocracking of Heavy Residues. J. Ind.
CoMo/γ-Al2O3, and Mo/γ-Al2O3 Catalysts. J. Catal. 2004, 225, 417− Eng. Chem. 2015, 27, 12−24.
427. (32) Rayo, P.; Ramírez, J.; Torres-Mancera, P.; Marroquín, G.; Maity,
(12) Shafi, R.; Hutchings, G. J. Hydrodesulfurization of Hindered S. K.; Ancheyta, J. Hydrodesulfurization and Hydrocracking of Maya
Dibenzothiophenes: An Overview. Catal. Today 2000, 59, 423−442. Crude with P-Modified NiMo/Al2O3 Catalysts. Fuel 2012, 100, 34−42.
(13) Stanislaus, A.; Marafi, A.; Rana, M. S. Recent Advances in the (33) Wang, H.; Prins, R. Hydrodesulfurization of Dibenzothiophene,
Science and Technology of Ultra Low Sulfur Diesel (ULSD) 4,6-Dimethyldibenzothiophene, and Their Hydrogenated Intermedi-
Production. Catal. Today 2010, 153, 1−68. ates over Ni−MoS2/γ-Al2O3. J. Catal. 2009, 264, 31−43.
(14) Singhal, G. H.; Espino, R. L.; Sobel, J. E.; Huff, G. A., Jr. (34) Koltai, T.; Macaud, M.; Guevara, A.; Schulz, E.; Lemaire, M.;
Hydrodesulfurization of Sulfur Heterocyclic Compounds: Kinetics of Bacaud, R.; Vrinat, M. Comparative Inhibiting Effect of Polycondensed
Dibenzothiophene. J. Catal. 1981, 67, 457−468. Aromatics and Nitrogen Compounds on the Hydrodesulfurization of
(15) Meille, V.; Schulz, E.; Lemaire, M.; Vrinat, M. Hydro- Alkyldibenzothiophenes. Appl. Catal., A 2002, 231, 253−261.
desulfurization of Alkyldibenzothiophenes over a NiMo/Al2O3 Catalyst: (35) Lecrenay, E.; Sakanishi, K.; Mochida, I. Catalytic Hydro-
Kinetics and Mechanism. J. Catal. 1997, 170, 29−36. desulfurization of Gas Oil and Model Sulfur Compounds over
(16) Bataille, F.; Lemberton, J.-L.; Michaud, P.; Pérot, G.; Vrinat, M.; Commercial and Laboratory-Made CoMo and NiMo Catalysts: Activity
Lemaire, M.; Schulz, E.; Breysse, M.; Kasztelan, S. Alkyldibenzothio- and Reaction Scheme. Catal. Today 1997, 39, 13−20.
phenes Hydrodesulfurization-Promoter Effect, Reactivity, and Reaction (36) Song, T.; Zhang, Z.; Chen, J.; Ring, Z.; Yang, H.; Zheng, Y. Effect
Mechanism. J. Catal. 2000, 191, 409−422. of Aromatics on Deep Hydrodesulfurization of Dibenzothiophene and

3940 DOI: 10.1021/acscatal.8b00629


ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

4,6-Dimethyldibenzothiophene over NiMo/Al2O3 Catalyst. Energy (57) Castillo-Araiza, C. O.; Chávez, G.; Dutta, A.; de los Reyes, J. A.;
Fuels 2006, 20, 2344−2349. Nuñez, S.; García-Martínez, J. C. Role of Pt−Pd/γ-Al2O3 on the HDS of
(37) Baldovino-Medrano, V. G.; Giraldo, S. A.; Centeno, A. The 4,6-DMBT: Kinetic Modeling & Contribution Analysis. Fuel Process.
Functionalities of Pt/γ-Al2O3 Catalysts in Simultaneous HDS and HDA Technol. 2015, 132, 164−172.
Reactions. Fuel 2008, 87, 1917−1926. (58) Che-Galicia, G.; Ruiz-Martínez, R. S.; López-Isunza, F.; Castillo-
(38) Stanislaus, A.; Cooper, B. H. Aromatic Hydrogenation Catalysis: Araiza, C. O. Modeling of Oxidative Dehydrogenation of Ethane to
A Review. Catal. Rev.: Sci. Eng. 1994, 36, 75−123. Ethylene on a MoVTeNbO/TiO2 Catalyst in an Industrial-Scale Packed
(39) Beltramone, A. R.; Resasco, D. E.; Alvarez, W. E.; Choudhary, T. Bed Catalytic Reactor. Chem. Eng. J. 2015, 280, 682−694.
V. Simultaneous Hydrogenation of Multiring Aromatic Compounds (59) Raghuveer, C. S.; Thybaut, J. W.; Marin, G. B. Pyridine
over NiMo Catalyst. Ind. Eng. Chem. Res. 2008, 47, 7161−7166. Hydrodenitrogenation Kinetics over a Sulphided NiMo/γ-Al2O3
(40) Egorova, M.; Prins, R. Competitive Hydrodesulfurization of 4,6- Catalyst. Fuel 2016, 171, 253−262.
Dimethyldibenzothiophene, Hydrodenitrogenation of 2-Methylpyri- (60) Brown, P. N.; Byrne, G. D.; Hindmarsh, A. C. VODE: A Variable-
dine, and Hydrogenation of Naphthalene over Sulfided NiMo/γ-Al2O3. Coefficient ODE Solver. SIAM J. Sci. Stat. Comput. 1989, 10, 1038−
J. Catal. 2004, 224, 278−287. 1051.
(41) Logadóttir, Á .; Moses, P. G.; Hinnemann, B.; Topsøe, N.-Y.; (61) Rosenbrock, H. H. An Automatic Method for Finding the
Knudsen, K. G.; Topsøe, H.; Nørskov, J. K. A Density Functional Study Greatest or Least Value of a Function. Comput. J. 1960, 3, 175−184.
of Inhibition of the HDS Hydrogenation Pathway by Pyridine, Benzene, (62) Boggs, P. T.; Donaldson, J. R.; Byrd, R. h.; Schnabel, R. B.
and H2S on MoS2-Based Catalysts. Catal. Today 2006, 111, 44−51. Algorithm 676: ODRPACK: Software for Weighted Orthogonal
(42) Farag, H.; Sakanishi, K.; Mochida, I.; Whitehurst, D. D. Kinetic Distance Regression. ACM Trans. Math. Softw. 1989, 15, 348−364.
Analyses and Inhibition by Naphthalene and H2S in Hydrodesulfuriza- (63) Marquardt, D. W. An Algorithm for Least-Squares Estimation of
tion of 4,6-Dimethyldibenzothiophene (4,6-DMDBT) over CoMo- Nonlinear Parameters. J. Soc. Ind. Appl. Math. 1963, 11, 431−441.
Based Carbon Catalyst. Energy Fuels 1999, 13, 449−453. (64) Kass, R. E.; Raftery, A. E. Bayes Factors. J. Am. Stat. Assoc. 1995,
(43) Vannice, M. A.; Hyun, S. H.; Kalpakci, B.; Liauh, W. C. Entropies 90, 773−795.
of Adsorption in Heterogeneous Catalytic Reactions. J. Catal. 1979, 56, (65) Myung, J.; Tang, Y.; Pitt, M. A. Evaluation and Comparison of
358−362. Computational Models. Methods Enzymol. 2009, 454, 287−304.
(44) Froment, G. F.; Bischoff, K. B.; De Wilde, J. Chemical Reactor (66) Schwarz, G. Estimating the Dimension of a Model. Ann. Stat.
Analysis and Design, 3rd Edition; Wiley: New York, 2010; p 126. 1978, 6, 461−464.
(45) Jae Lee, W.; Froment, G. F. Ethylbenzene Dehydrogenation into (67) Boudart, M.; Djéga-Mariadassou, G. Kinetics of Heterogeneous
Styrene: Kinetic Modeling and Reactor Simulation. Ind. Eng. Chem. Res. Catalytic Reactions; Princeton University Press: Princeton, NJ, 1984.
2008, 47, 9183−9194. (68) Mhadeshwar, A. B.; Wang, H.; Vlachos, D. G. Thermodynamic
(46) Baldovino-Medrano, V. G.; Eloy, P.; Gaigneaux, E. M.; Giraldo, S. Consistency in Microkinetic Development of Surface Reaction
Mechanisms. J. Phys. Chem. B 2003, 107, 12721−12733.
A.; Centeno, A. Development of the HYD Route of Hydrodesulfuriza-
(69) Egorova, M. Mutual Influence of the HDS of Dibenzothiophene
tion of Dibenzothiophenes over Pd−Pt/γ-Al2O3 Catalysts. J. Catal.
and HDN of 2-Methylpyridine. J. Catal. 2004, 221, 11−19.
2009, 267, 129−139.
(70) Santacesaria, E. Kinetics and Transport Phenomena. Catal. Today
(47) Schachtl, E.; Kondratieva, E.; Gutiérrez, O. Y.; Lercher, J. A.
1997, 34, 393−400.
Pathways for H2 Activation on (Ni)-MoS2 Catalysts. J. Phys. Chem. Lett.
(71) Kilanowski, D. R.; Teeuwen, H.; Beer, V. H. J. de; Gates, B. C.;
2015, 6, 2929−2932.
Schuit, G. C. A.; Kwart, H. Hydrodesulfurization of Thiophene,
(48) García-Martínez, J. C.; Castillo-Araiza, C. O.; De los Reyes
Benzothiophene, Dibenzothiophene, and Related Compounds Cata-
Heredia, J. A.; Trejo, E.; Montesinos, A. Kinetics of HDS and of the lyzed by Sulfided CoO-MoO3/γ-Al2O3: Low-Pressure Reactivity
Inhibitory Effect of Quinoline on HDS of 4,6-DMDBT over a Ni-Mo-P/ Studies. J. Catal. 1978, 55, 129−137.
Al2O3 Catalyst: Part I. Chem. Eng. J. 2012, 210, 53−62. (72) Kagami, N.; Vogelaar, B. M.; van Langeveld, A. D.; Moulijn, J. A.
(49) Girgis, M. J.; Gates, B. C. Reactivities, Reaction Networks, and Reaction Pathways on NiMo/Al2O3 Catalysts for Hydrodesulfurization
Kinetics in High-Pressure Catalytic Hydroprocessing. Ind. Eng. Chem. of Diesel Fuel. Appl. Catal., A 2005, 293, 11−23.
Res. 1991, 30, 2021−2058. (73) Alsolami, B. H.; Berger, R. J.; Makkee, M.; Moulijn, J. A. Catalyst
(50) Rautanen, P. A.; Lylykangas, M. S.; Aittamaa, J. R.; Krause, A. O. I. Performance Testing in Multiphase Systems: Implications of Using
Liquid-Phase Hydrogenation of Naphthalene and Tetralin on Ni/ Small Catalyst Particles in Hydrodesulfurization. Ind. Eng. Chem. Res.
Al2O3: Kinetic Modeling. Ind. Eng. Chem. Res. 2002, 41, 5966−5975. 2013, 52, 9069−9085.
(51) Cortés Romero, C. M.; Thybaut, J. W.; Marin, G. B. Naphthalene (74) De Oliveira, L. P.; Hudebine, D.; Guillaume, D.; Verstraete, J. J. A
Hydrogenation over a NiMo/γ-Al2O3 Catalyst: Experimental Study and Review of Kinetic Modeling Methodologies for Complex Processes. Oil
Kinetic Modelling. Catal. Today 2008, 130, 231−242. Gas Sci. Technol. 2016, 71, 45.
(52) Lapinas, A. T.; Klein, M. T.; Gates, B. C.; Macris, A.; Lyons, J. E. (75) Grønborg, S. S.; Šarić, M.; Moses, P. G.; Rossmeisl, J.; Lauritsen, J.
Catalytic Hydrogenation and Hydrocracking of Fluorene: Reaction V. Atomic Scale Analysis of Sterical Effects in the Adsorption of 4,6-
Pathways, Kinetics, and Mechanisms. Ind. Eng. Chem. Res. 1991, 30, 42− Dimethyldibenzothiophene on a CoMoS Hydrotreating Catalyst. J.
50. Catal. 2016, 344, 121−128.
(53) Schachtl, E.; Zhong, L.; Kondratieva, E.; Hein, J.; Gutiérrez, O. Y.; (76) Lauritsen, J. V.; Besenbacher, F. Atom-Resolved Scanning
Jentys, A.; Lercher, J. A. Understanding Ni Promotion of MoS2/γ-Al2O3 Tunneling Microscopy Investigations of Molecular Adsorption on
and Its Implications for the Hydrogenation of Phenanthrene. MoS2 and CoMoS Hydrodesulfurization Catalysts. J. Catal. 2015, 328,
ChemCatChem 2015, 7, 4118−4130. 49−58.
(54) Ishihara, A.; Lee, J.; Dumeignil, F.; Takashi, M.; Qian, E. W.; Kabe, (77) Lauritsen, J. V.; Nyberg, M.; Nørskov, J. K.; Clausen, B. S.;
T. Elucidation of Retarding Effects of Sulfur and Nitrogen Compounds Topsøe, H.; Lægsgaard, E.; Besenbacher, F. Hydrodesulfurization
on Aromatic Compounds Hydrogenation. Energy Fuels 2003, 17, 1338− Reaction Pathways on MoS2 Nanoclusters Revealed by Scanning
1345. Tunneling Microscopy. J. Catal. 2004, 224, 94−106.
(55) Korre, S. C.; Klein, M. T.; Quann, R. J. Polynuclear Aromatic (78) Rangarajan, S.; Mavrikakis, M. Adsorption of Nitrogen- and
Hydrocarbons Hydrogenation. 1. Experimental Reaction Pathways and Sulfur-containing Compounds on NiMoS for Hydrotreating Reactions:
Kinetics. Ind. Eng. Chem. Res. 1995, 34, 101−117. A DFT and vdW-corrected Study. AIChE J. 2015, 61, 4036−4050.
(56) Yang, H.; Wang, Y.; Jiang, H.; Weng, H.; Liu, F.; Li, M. Kinetics of (79) Rangarajan, S.; Mavrikakis, M. On the Preferred Active Sites of
Phenanthrene Hydrogenation System over CoMo/Al2O3 Catalyst. Ind. Promoted MoS2 for Hydrodesulfurization with Minimal Organo-
Eng. Chem. Res. 2014, 53, 12264−12269. nitrogen Inhibition. ACS Catal. 2017, 7, 501−509.

3941 DOI: 10.1021/acscatal.8b00629


ACS Catal. 2018, 8, 3926−3942
ACS Catalysis Research Article

(80) Ding, S.; Jiang, S.; Zhou, Y.; Wei, Q.; Zhou, W. Catalytic
Characteristics of Active Corner Sites in CoMoS Nanostructure
HydrodesulfurizationA Mechanism Study Based on DFT Calcu-
lations. J. Catal. 2017, 345, 24−38.
(81) Chen, W.; Long, X.; Li, M.; Nie, H.; Li, D. Influence of Active
Phase Structure of CoMo/Al2O3 Catalyst on the Selectivity of
Hydrodesulfurization and Hydrodearomatization. Catal. Today 2017,
292, 97−109.
(82) Rangarajan, S.; Mavrikakis, M. DFT Insights into the Competitive
Adsorption of Sulfur- and Nitrogen-Containing Compounds and
Hydrocarbons on Co-Promoted Molybdenum Sulfide Catalysts. ACS
Catal. 2016, 6, 2904−2917.
(83) Kibsgaard, J.; Tuxen, A.; Knudsen, K. G.; Brorson, M.; Topsøe,
H.; Lægsgaard, E.; Lauritsen, J. V.; Besenbacher, F. Comparative
Atomic-Scale Analysis of Promotional Effects by Late 3d-Transition
Metals in MoS2 Hydrotreating Catalysts. J. Catal. 2010, 272, 195−203.
(84) Kabe, T.; Qian, W.; Ishihara, A. Elucidation of Hydro-
desulfurization Mechanism Using 35S Radioisotope Pulse Tracer
Methods. Catal. Today 1997, 39, 3−12.
(85) Baldovino-Medrano, V. G.; Giraldo, S. A.; Centeno, A. Reactivity
of Dibenzothiophene Type Molecules over Pd Catalysts. J. Mol. Catal.
A: Chem. 2009, 301, 127−133.
(86) Yang, H.; Fairbridge, C.; Ring, Z. Adsorption of Dibenzothio-
phene Derivatives over a MoS2 NanoclusterA Density Functional
Theory Study of Structure−Reactivity Relations. Energy Fuels 2003, 17,
387−398.
(87) Kabe, T.; Aoyama, Y.; Wang, D.; Ishihara, A.; Qian, W.; Hosoya,
M.; Zhang, Q. Effects of H2S on Hydrodesulfurization of Dibenzothio-
phene and 4,6-Dimethyldibenzothiophene on Alumina-Supported
NiMo and NiW Catalysts. Appl. Catal., A 2001, 209, 237−247.
(88) Lamure-Meille, V.; Schulz, E.; Lemaire, M.; Vrinat, M. Effect of
Experimental Parameters on the Relative Reactivity of Dibenzothio-
phene and 4-Methyldibenzothiophene. Appl. Catal., A 1995, 131, 143−
157.
(89) Weitkamp, A. W. Stereochemistry and Mechanism of Hydro-
genation of Naphthalenes on Transition Metal Catalysts and
Conformational Analysis of the Products. Adv. Catal. 1968, 18, 1−110.
(90) Gutiérrez, O. Y.; Singh, S.; Schachtl, E.; Kim, J.; Kondratieva, E.;
Hein, J.; Lercher, J. A. Effects of the Support on the Performance and
Promotion of (Ni)MoS2 Catalysts for Simultaneous Hydrodenitroge-
nation and Hydrodesulfurization. ACS Catal. 2014, 4, 1487−1499.
(91) Rana, M. S.; Navarro, R.; Leglise, J. Competitive Effects of
Nitrogen and Sulfur Content on Activity of Hydrotreating CoMo/Al2O3
Catalysts: A Batch Reactor Study. Catal. Today 2004, 98, 67−74.
(92) Monteiro-Gezork, A. C. A.; Natividad, R.; Winterbottom, J. M.
Hydrogenation of Naphthalene on NiMo- Ni- and Ru/Al2O3 Catalysts:
Langmuir−Hinshelwood Kinetic Modelling. Catal. Today 2008, 130,
471−485.
(93) Sapre, A. V.; Gates, B. C. Hydrogenation of Aromatic
Hydrocarbons Catalyzed by Sulfided Cobalt Oxide-Molybdenum
Oxide/α-Aluminum Oxide. Reactivities and Reaction Networks. Ind.
Eng. Chem. Process Des. Dev. 1981, 20, 68−73.

3942 DOI: 10.1021/acscatal.8b00629


ACS Catal. 2018, 8, 3926−3942

You might also like