Вы находитесь на странице: 1из 12

Journal of Catalysis 321 (2015) 39–50

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Theoretical investigation of the reaction mechanism of the


hydrodeoxygenation of guaiacol over a Ru(0 0 0 1) model surface
Jianmin Lu, Andreas Heyden ⇑
Department of Chemical Engineering, University of South Carolina, 301 S. Main St., Columbia, SC 29208, USA

a r t i c l e i n f o a b s t r a c t

Article history: The reaction mechanism of the hydrodeoxygenation of guaiacol to aromatic products has been studied by
Received 31 July 2014 density functional theory calculations and microkinetic modeling over a Ru(0 0 0 1) model surface. Our
Revised 23 October 2014 model suggests that the dominant hydrodeoxygenation pathway proceeds via O–H bond cleavage of
Accepted 1 November 2014
guaiacol, C6H4(OH)(OCH3), to C6H4(O)(OCH3), followed by dehydrogenation of the methoxy group to C6-
Available online 24 November 2014
H4(O)(OC), decarbonylation to C6H4O, and finally hydrogenation to phenol. At the adsorbed C6H4(O)(OCH)
intermediate, a competitive deoxygenation pathway is identified, which involves methyne group removal
Keywords:
to C6H4O2, followed by hydrogenation to C6H4(OH)(O), dehydroxylation to C6H4O, and finally hydrogena-
Guaiacol
Ruthenium
tion to phenol. In agreement with experimental results, phenol is predicted to be the major product and
Density functional theory catechol is the most relevant minority side product. Further deoxygenation of phenol to benzene is found
Microkinetic modeling to be slow. Finally, computations predict the last dehydrogenation step of the methoxy species in
Deoxygenation guaiacol to be at least partially rate controlling over Ru(0 0 0 1).
Hydrodeoxygenation Ó 2014 Elsevier Inc. All rights reserved.
Lignin

1. Introduction cleavage reaction. Since bio-oils are a mixture of hundreds of com-


pounds, the only viable strategy for better understanding the cata-
With the continued depletion of fossil resources as a source for lytic upgrading process involves the mechanistic investigation of
fuels and chemicals, renewable raw materials such as biomass model compounds. Guaiacol (C6H4(OH)(OCH3)) is such a represen-
have drawn more and more attention as a possible source for fuels tative compound of phenol derivatives and lignin-derived oligo-
and chemicals. Bio-oils produced from fast pyrolysis and hydro- mers since it contains one phenyl ring and two oxygen-containing
thermal liquefaction of lignocellulosic biomass have been identi- functional groups, –OH and –OCH3. A considerable number of
fied as a promising alternative feedstock for fuels and chemical papers have reported experimental investigations on the catalytic
production with several environmental advantages. For example, hydrodeoxygenation of guaiacol on transition metals such as Fe
bio-oil production is CO2 neutral and its conversion emits no SOx [4,5], Ru [5–7], Rh [7–9], Pd [5,7,8], Pt [5,7,8,10], and bimetallic cat-
and 50% less of NOx than diesel oil combustion in a gas turbine alysts such as Co–Mo [11], Pt–Sn [12], Ni–Cu [13], Rh–Pt [8], and
[1]. The composition of bio-oils depends heavily on the type of bio- Pd–Fe [5]. Nimmanwudipong and coworkers [10] have mapped a
mass and the production conditions, but bio-oils are generally a reaction network for the hydrodeoxygenation of guaiacol over Pt/
complex mixture of water (15–30%) and many organic compounds, c-Al2O3 in the presence of H2 based on selectivity-conversion data,
including organic acids, aldehydes, alcohols, esters, phenol deriva- which is very insightful but lacks straightforward evidence. Sun and
tives (50–65%), and lignin-derived oligomers (20%) [2,3]. A disad- coworkers [5] have also proposed on the basis of their products a
vantage of bio-oils is their high viscosity, high corrosiveness possible reaction pathway for Ru, Pd, and Pt catalysts, guaiacol ?
(acidity), instability, and low heating value, etc., originating from catechol ? phenol ? benzene and a different one, guaiacol ?
the high oxygen content of most bio-oils. Therefore, upgrading of phenol ? benzene, for Fe and Pd–Fe catalysts.
bio-oils, that is, deoxygenation, is critical for its application as a In this study, we aimed at identifying the reaction mechanism
liquid fuel or chemical feedstock. and possible activity descriptors for the HDO of guaiacol over
Catalytic upgrading of bio-oils is commonly done by hydrodeox- Ru(0 0 0 1) model surfaces using density functional theory calcula-
ygenation (HDO), which involves the challenging C–O bond tions and mean-field microkinetic modeling. To the best of the
authors’ knowledge, no theoretical study on the catalytic hydro-
⇑ Corresponding author. Fax: +1 8037770973. deoxygenation of guaiacol over transition metal catalysts has yet
E-mail address: heyden@cec.sc.edu (A. Heyden). been published.

http://dx.doi.org/10.1016/j.jcat.2014.11.003
0021-9517/Ó 2014 Elsevier Inc. All rights reserved.
40 J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50

Fig. 1. Reaction network investigated for the hydrodeoxygenation of guaiacol to aromatic products over Ru(0 0 0 1). For clarity, duplicate structures are highlighted by
identical background colors. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

2. Methods plane-wave energy cutoff has been confirmed. The convergence


criterion for the total energy was set to 107 eV.
All calculations presented in this work were performed using Adsorption energies of all the surface intermediates reported in
the periodic DFT implementation in the Vienna Ab Initio Simula- this paper were calculated in their most favorable adsorption
tion Package (VASP) [14,15]. The electron–ion interactions were modes. The adsorption energies, Eads, were calculated by the fol-
described by the projector-augmented wave (PAW) method. The lowing equation:
PAW method is a frozen-core all-electron method that uses the
exact shape of the valence wave functions instead of pseudo-wave
Eads ¼ Eslabþadsorbate  Eslab  EadsorbateðgasÞ ð1Þ
functions [16]. The exchange correlation energy has been calcu-
where Eslab+adsorbate is the total energy of an adsorbate bound to the
lated within the generalized gradient approximation (GGA) using
Ru slab, Eslab is the total energy of the clean Ru slab, and Eadsorbate(gas)
the PBE functional form [17,18]. For dispersion interactions, we
is the total energy of the adsorbate in the gas phase.
used the DFT-D3 methodology [19]. An energy cutoff for plane
Finally, transition states for elementary reaction steps were
waves of 400 eV was employed throughout this study.
determined by a combination of the nudged elastic band (NEB)
The total energy of HCP-Ru bulk approaches a minimum when
method [21] and the dimer method [22–24]. In the NEB method,
its lattice constants are a = 2.7020 Å and c = 4.2741 Å, which is in
the path between the reactant and product is discretized into a
reasonable agreement with the experimental values (a = 2.7059 Å
series of structural images. The image that is closest to a likely
and c = 4.2815 Å [20]). The Ru(0 0 0 1) surface was constructed as
transition state structure was then employed as an initial guess
a periodic slab with four Ru layers separated by a vacuum layer
structure for the dimer method. All adsorption energies and activa-
of 15 Å in order to eliminate interactions between the slab and
p tion barriers reported in this study have been zero-point corrected
its images. Each Ru layer had 16 Pd atoms with a (4  2 3) period-
(DZPE).
icity, allowing for adsorbate coverages as low as 1/16 ML. The bot-
For the microkinetic modeling, we have employed the same
tom two Ru layers were fixed to their optimized bulk configuration
methodology as described in our previous paper [25]. The nonlin-
during all computations, while the top two layers were fully
ear steady state surface species equations have been solved using
relaxed. The adsorbates were free to relax in all directions. All
the BzzMath library [26] developed by Buzzi-Ferraris.
atomic coordinates of the adsorbates and the Ru atoms in the
relaxed layers were optimized to a force less than 0.03 eV/Å on
each atom. All self-consistent field (SCF) calculations were con- 3. Results and discussion
verged to 1  103 kJ/mol. Brillouin zone integration was per-
formed using a 4  4  1 Monkhorst–Pack grid and a Methfessel– Fig. 1 illustrates the reaction pathways included for the HDO of
Paxton smearing of 0.2 eV. In all cases, the convergence of total guaiacol to benzene over the Ru(0 0 0 1) surface model. In the fol-
energy with respect to the k-point mesh and with respect to lowing, we will first discuss the effects of dispersion corrections
J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50 41

Table 1 Table 3
Zero-point energy corrected adsorption energies (in eV) of selected reaction Adsorption sites occupied, zero-point energy corrected adsorption energies (E0,ads),
intermediates in the hydrodeoxygenation of guaiacol computed with PBE and PBE- and zero-point energy corrections (DZPE) of various reaction intermediates in the
D3 functional. Herein, a-carbon means the aromatic ring carbon binding to the –OH hydrodeoxygenation of guaiacol computed with PBE-D3.
group, and b-carbon means the aromatic ring carbon binding to the –OCH3 group. Ha
symbolizes a hydrogen atom binding to the a–carbon, and Hb symbolizes a hydrogen Formula No. of sites E0,ads (eV) DZPE (eV)
atom binding to the b-carbon. C6H4(OH)(OCH3), guaiacol 4 2.46 0.02
C6H4Hb(OH)(OCH3) 4 4.10 0.05
Formula PBE PBE-D3 DE0 = PBE-D3  PBE
C6H4Ha(OH)(OCH3) 4 4.04 0.03
C6H4(OH)(OCH3) 1.33 2.46 1.13 C6H4OCH3 4 4.16 0.04
C6H4Hb(OH)(OCH3) 2.80 4.10 1.30 C6H4OH 4 4.27 0.06
C6H4Ha(OH)(OCH3) 2.77 4.04 1.27 C6H4(OH)(OCH2) 5 4.12 0.03
C6H4OCH3 3.10 4.16 1.06 C6H4(OH)(O) 4 3.74 0.04
C6H4OH 3.34 4.27 0.94 C6H4(O)(OCH3) 4 4.03 0.05
C6H4(OH)(OCH2) 2.88 4.12 1.25 C6H4Hb(OH)(OCH2) 5 5.32 0.02
C6H4(OH)(O) 2.71 3.74 1.03 C6H5OCH3, anisole 4 2.54 0.05
C6H4(O)(OCH3) 2.85 4.03 1.18 C6H4(OH)(OCH) 5 5.72 0.04
C6H5OCH3 1.43 2.54 1.11 C6H4(OH)2, catechol 4 2.60 0.03
C6H4(OH)2 1.60 2.60 1.00 C6H4O 4 5.04 0.12
C6H5OH 1.59 2.51 0.92 C6H4(O)(OCH2) 5 4.84 0.00
C6H6 1.53 2.42 0.89 C6H4Hb(OH)(O) 4 3.41 0.01
CO 1.86 2.06 0.20 C6H5OCH2 4 4.09 0.04
H 2.73 2.78 0.05 C6H5O, phenoxy 4 4.40 0.01
OH 3.42 3.56 0.14 C6H5OH, phenol 4 2.51 0.02
H2O 0.38 0.59 0.21 C6H5(OH)2 4 4.28 0.03
CH3 2.09 2.37 0.28 C6H4O2 5 4.56 0.03
CH3O 2.63 2.92 0.29 C6H5 4 4.11 0.01
CH3OH 0.40 0.75 0.35 C6H6OH 4 4.37 0.04
C6H6, benzene 3 2.42 0.01
C6H4Ha(O)(OH) 4 3.59 0.02
C6H4(O)(OCH) 5 6.66 0.01
C6H4(O)(OC) 5 4.64 0.09
in Section 3.1. Then, we report the adsorption behavior of reac- H 1 2.78 0.17
tants, products, and surface intermediates in Section 3.2. Finally, OH 1 3.56 0.12
we will investigate each reaction pathway shown in Fig. 1 individ- H2O 1 0.59 0.05
CH 1 6.94 0.18
ually in Sections 3.3.
CH2 1 4.52 0.12
CH3 1 2.37 0.08
CH4 1 0.22 0.01
3.1. Effects of dispersion corrections CO 1 2.06 0.07
CHO 2 2.65 0.11
It is well known that conventional DFT fails to capture disper- CH2O 2 1.36 0.05
sion forces (van der Waals forces) that often dominate the interac- CH3O 1 2.92 0.17
CH3OH 1 0.75 0.03
tions between alkanes and aromatics with transition metal
surfaces [27]. For example, the adsorption energy of benzene on
a Ag(1 1 1) surface has experimentally been determined to be
0.69 eV [28], while conventional DFT predicts a weak adsorption
(0.09 eV [29]) or even no adsorption [19]. DFT with dispersion [30]. PBE slightly overestimates the adsorption energy by 0.05 eV
corrections predicts adsorption energies closer to the experimental (DE0 = 1.86 eV), while PBE-D3 significantly overestimates it by
values (0.75 eV [29] by PBE + vdWsurf and 0.95 eV [19] by rev- about 0.25 eV (DE0 = 2.06 eV). Fortunately, Table 2 illustrates that
PBE-D3). Table 1 illustrates a similar trend in adsorption energies reaction energies for surface reactions are fairly independent of
of various reaction intermediates in the HDO of guaiacol over functional/dispersion interaction. For the test set of elementary
Ru(0 0 0 1) computed with PBE-D3 and PBE functional. For aromatic surface reactions, the difference in surface reaction energy
surface intermediates, PBE-D3 predicts a 0.9–1.3 eV stronger between PBE and PBE-D3 is less than 0.1 eV. As a result, we con-
adsorption energy than PBE. It is noticeable that PBE-D3 also pre- clude that PBE-D3 will likely provide a meaningful description of
dicts a higher adsorption energy for the smaller surface species the surface chemistry of the HDO of guaiacol; although, predicted
such as H, OH, H2O, and CO, which is likely an overestimation of surface coverages of, for example, CO and H will likely be overesti-
the adsorption energy. For example, the CO adsorption energy mated, primarily due to the neglect of electronic screening at the
has experimentally been estimated to be 1.81 eV on Ru(0 0 0 1) metal surface [31].

Table 2
Zero-point energy corrected reaction energies (in eV) of selected surface reactions in the hydrodeoxygenation of guaiacol computed with PBE and PBE-D3 functional.

Step no. Formula PBE PBE-D3 DE0 = PBE-D3  PBE


1 C6H4(OH)(OCH3) + H ? C6H4Hb(OH)(OCH3) 0.44 0.38 0.06
2 C6H4(OH)(OCH3) + H ? C6H4Ha(OH)(OCH3) 0.52 0.49 0.03
3 C6H4(OH)(OCH3) ? C6H4OCH3 + OH 0.24 0.21 0.03
4 C6H4(OH)(OCH3) ? C6H4OH + CH3O 0.48 0.44 0.04
5 C6H4(OH)(OCH3) ? C6H4(OH)(OCH2) + H 0.33 0.42 0.09
6 C6H4(OH)(OCH3) ? C6H4(OH)(O) + CH3 1.32 1.37 0.05
7 C6H4(OH)(OCH3) ? C6H4(O)(OCH3) + H 0.79 0.80 0.01
48 H + OH ? H2O 0.57 0.57 0.00
51 H + CH3O ? CH3OH 0.74 0.77 0.03
42 J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50

with an activation barrier of 1.11 eV, which suggests that pathway


1 is unfavorable. C6H4Hb(OH)(OCH3) can subsequently go either
through dehydrogenation steps of the methoxy group to C6H4Hb(-
OH)(OCH2) (step 8) or through a –OCH3 removal step to produce
phenol (step 9). Considering that step 8 is exothermic by
0.16 eV and has an activation barrier of 0.88 eV, while step 9 is
exothermic by 0.97 eV and possesses a lower activation barrier
of 0.76 eV, step 9 seems to be preferred over step 8. Next, the
methylene group (–CH2) can be removed from C6H4Hb(OH)(OCH2)
to form C6H4Hb(OH)(O) (step 22), which is a relatively facile step
that is exothermic by 0.85 eV and possesses an activation barrier
of 0.37 eV. Hydrogenation of C6H4Hb(OH)(O) to C6H5(OH)2 (step
35), which is a hydrogenated catechol species, is a rather difficult
endothermic step (DE0 = 0.23 eV) with an activation barrier of
0.90 eV. The final step to produce phenol involves dehydroxylation
Fig. 2. Illustration of adsorption sites for guaiacol on the Ru(0 0 0 1) surface. There
of the phenyl ring (step 40) which is exothermic by 0.80 eV and
are four possible adsorption sites—atop, bridge, fcc, and hcp. At each site, guaiacol
has two orientations, 0° and 30°, referring to the angles of the C–C bond to the has a barrier of 0.63 eV.
nearest-neighboring Ru–Ru bond. Similar to the hydrogenation step 1 of the phenyl ring in b-car-
bon position in pathway 1, hydrogenation of guaiacol in a-carbon
position to form C6H4Ha(OH)(OCH3) (step 2) is also endothermic
3.2. Adsorbed intermediates by 0.49 eV and kinetically demanding with an activation barrier
of 1.15 eV, which makes pathway 2 similarly unfavorable. Dehydr-
Adsorption geometries of the reactants, products, and possible oxylation of C6H4Ha(OH)(OCH3) to form anisole (step 10) is exo-
intermediates involved in the reaction network are shown in Figs. 3 thermic by 0.88 eV and kinetically feasible with an activation
and 4. The number of adsorption sites, adsorption energies, and barrier of 0.58 eV. Anisole can either desorb or dehydrogenate its
zero-point energy corrections of these intermediates is listed in –OCH3 group to form C6H5OCH2 (step 23) which is mildly exother-
Table 3. mic by 0.38 eV and facile with an activation barrier of 0.44 eV.
For aromatic molecules, we generally find they can adsorb at Next, methylene group removal of C6H5OCH2 to phenoxy (step
four different adsorption sites (atop, bridge, fcc, and hcp) on the 36) is very exothermic by 1.29 eV and possesses a low barrier
Ru(0 0 0 1) surface. As shown in Fig. 3, all aromatic adsorbates bind of only 0.12 eV. Overall, the energetics of pathways 1 and 2 suggest
to the Ru(0 0 0 1) surface via their aromatic ring/carbon atoms, that that (except at very high hydrogen partial pressures) the HDO of
is, the aromatic rings are parallel to the Ru(0 0 0 1) surface. In addi- guaiacol does not proceed by partial hydrogenation of the phenyl
tion, they can adsorb in two orientations of 0° and 30° at each site, ring, which agrees with experimental observations [5].
referring to the angles of the C–C bond relative to the close-packed
Ru–Ru bond as shown in Fig. 2. Our calculations predict that all 26 3.3.2. Pathways with direct removal of functional groups
aromatic surface intermediates adsorb preferentially in the hcp30 Direct removal of functional groups such as –OH, –OCH3, and –
geometry (see Fig. 3). For comparison, Liu et al. found that benzene CH3 from guaiacol is found to be kinetically difficult. For pathway
prefers to chemisorb in bridge30 position on the (1 1 1) surfaces of 3, the direct removal of a hydroxyl group from guaiacol to form
Ir, Rh, Pd, and Pt [29]. Finally, all aromatic reactants and products adsorbed C6H4OCH3 (step 3) is slightly exothermic by 0.21 eV,
such as guaiacol, anisole, catechol, phenol, and benzene adsorb but requires overcoming an activation barrier of 1.07 eV. Subse-
strongly on the Ru(0 0 0 1) surface with an adsorption energies of quent hydrogenation to anisole (step 11) is again exothermic by
DE0  2.5 eV. 0.18 eV and involves overcoming a barrier of 0.67 eV. Pathways
3 and 2 merge at this point and will consequently not be discussed
3.3. Potential energy surface of various reaction pathways further.
In the fourth pathway, a methoxy group is directly removed
Various reaction pathways investigated for the HDO of guaiacol from guaiacol to form an adsorbed C6H4OH species (step 4). This
are shown in Fig. 1. Pathways leading to cyclic, saturated products process is exothermic by 0.44 eV, but again requires overcoming
such as cyclohexanone and cyclohexanol are not taken into consid- a relatively high barrier of 0.87 eV. To produce phenol, the C6H4OH
eration in this reaction network, primarily because these pathways species is hydrogenated (step 12) in an exothermic process of
are not favorable on Ru catalysts at low H2 partial pressure [5]. In 0.17 eV that involves an activation barrier of 0.60 eV.
the following, we labeled the reaction pathways (1)–(7) according The sixth pathway involves direct removal of a methyl group
to the first reaction step labeled steps 1–7 in Fig. 1. Pathways 1 and from guaiacol to form C6H4(OH)(O) (step 6) which is quite exother-
2 start with hydrogenation steps of the b-carbon and a-carbon of mic, DE0 = 1.37 eV, but also involves a large activation barrier of
guaiacol, respectively. Pathways 3, 4, and 6 start with –OH, –CH3, 1.41 eV. Next, C6H4(OH)(O) can be hydrogenated to catechol (step
and –OCH3 removal from guaiacol, respectively. Finally, pathway 17) in an endothermic process DE0 = 0.78 eV with a barrier of
5 starts with dehydrogenation of the –OCH3 group of guaiacol, 1.06 eV. Alternatively, it can be hydrogenated to C6H4Ha(OH)(O)
and pathway 7 starts with dehydrogenation of the –OH group. in an endothermic process of 0.87 eV with a high barrier of
All reaction energies and activation barriers for the elementary 1.42 eV (step 18). Finally, C6H4(OH)(O) can be dehydroxylated to
steps investigated are listed in Table 4. Snapshots for transition C6H4O (step 19) in an endothermic process, DE0 = 0.53 eV, with a
states of all elementary steps involved in the microkinetic model- high barrier of 1.14 eV.
ing are shown in Figs. 5 and 6. Once catechol is formed, its deoxygenation to phenol is slow
since both the direct hydroxyl group removal to C6H4OH
3.3.1. Pathways involving partial hydrogenations of the phenyl ring (Ea = 1.16 eV for step 26) and first hydrogenation to C6H5(OH)2
In pathway 1, adsorbed guaiacol is hydrogenated on the (Ea = 1.16 eV for step 27) followed by dehydroxylation (step 40)
b-carbon of the phenyl ring to form C6H4Hb(OH)(OCH3) (step 1). are kinetically slow. Overall, pathways 1, 2, 3, 4, and 6 are all kinet-
This endothermic step (DE0 = 0.38 eV) is kinetically demanding ically unfavorable due to the presence of large activation barriers
J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50 43

Fig. 3. Side (upper panel) and top view (lower panel) of preferred adsorption structure of various intermediates with aromatic rings in the reaction network of the
hydrodeoxygenation of guaiacol over Ru(0 0 0 1).

in the hydrogenation of the phenyl ring or direct removal of a Next, C6H4(OH)(OCH) decomposes by methyne removal to pro-
hydroxyl, methoxy, or methyl group. duce C6H4(OH)(O) (step 25) which is an exothermic reaction,
DE0 = 1.29 eV, with a low activation barrier of 0.31 eV. As a result,
3.3.3. Pathways involving dehydrogenation reactions of functional this step is favored over –OCH removal (step 24) which possesses a
groups barrier of Ea = 1.03 eV. Reaction steps involving the hydrogenation
In the following, we discuss reaction pathways involving initial and dehydroxylation of C6H4(OH)(O) have already been discussed
dehydrogenation steps to form more active adsorbate intermedi- in Section 3.3.2.
ates. In the fifth reaction pathway, the methoxy group of guaiacol Finally, in the seventh reaction pathway, the phenolic hydrogen
is dehydrogenated to produce C6H4(OH)(OCH2) (step 5) which we in guaiacol is initially removed to produce C6H4(O)(OCH3) (step 7)
find to be a facile exothermic reaction (DE0 = 0.42 eV; Ea = 0.46 - which calculations based on DFT predict to be even more facile
eV). On Ru(0 0 0 1), C6H4(OH)(OCH2) dehydrogenates essentially than dehydrogenation of the methoxy group. We compute for this
barrierless (Ea = 0.01 eV, DE0 = 0.55 eV) to C6H4(OH)(OCH) (step exothermic reaction (DE0 = 0.80 eV) an activation barrier of
15), such that the three competing reaction steps (step 13: – merely 0.29 eV. Then, the methoxy group in C6H4(O)(OCH3) is
OCH2 removal, Ea = 1.22 eV, DE0 = 0.22 eV; step 14: –OH dehy- dehydrogenated in an exothermic process (DE0 = 0.38 eV) with
drogenation, Ea = 0.28 eV, DE0 = 0.76 eV; and step 16: –CH2 an activation barrier of 0.49 eV to produce C6H4(O)(OCH2) (step
removal, Ea = 0.45 eV, DE0 = 1.25 eV) are all kinetically more 21). Alternatively, the methoxy group is removed; however, this
difficult. process involves overcoming an activation barrier of 1.08 eV. After
44 J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50

Table 4
Calculated zero-point corrected reaction energies (DE), activation barriers (E ), and rate parameters for all elementary reaction steps in the hydrodeoxygenation of guaiacol over
Ru(0 0 0 1).

Step Reaction Constant T (K)


473 523 573 623
0 C6H4(OH)(OCH3)(gas) + 4⁄ ? C6H4(OH)(OCH3)⁄⁄⁄⁄ K0 7.35  108 5.23  105 1.06  103 4.78  100
DE0 = 2.46 eV k+0, s1 bar1 6.88  107 6.54  107 6.25  107 5.99  107
1 C6H4(OH)(OCH3)⁄⁄⁄⁄ + H⁄ ? C6H4Hb(OH)(OCH3)⁄⁄⁄⁄ + ⁄ K1 1.24  104 2.94  104 5.96  104 1.08  103
DE1 = 0.38 eV, E 1 = 1.11 eV k+1, s1 1.36  101 1.99  102 1.83  103 1.18  104
2 C6H4(OH)(OCH3)⁄⁄⁄⁄ + H⁄ ? C6H4Ha(OH)(OCH3)⁄⁄⁄⁄ + ⁄ K2 4.59  106 1.36  105 3.32  105 6.96  105
DE2 = 0.49 eV, E 2 = 1.15 eV k+2, s1 6.01  100 9.66  101 9.57  102 6.56  103
3 C6H4(OH)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H4OCH⁄⁄⁄⁄ 3 + OH⁄ K3 2.17  102 1.40  102 9.87  101 7.39  101
DE3 = 0.21 eV, E 3 = 1.07 eV k+3, s1 7.11  101 9.70  102 8.45  103 5.23  104
4 C6H4(OH)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H4OH⁄⁄⁄⁄ + CH3O⁄ K4 1.22  105 4.68  104 2.14  104 1.11  104
DE4 = 0.44 eV, E 4 = 0.87 eV k+4, s1 1.34  104 1.01  105 5.39  105 2.22  106
5 C6H4(OH)(OCH3)⁄⁄⁄⁄ + 2⁄ ? C6H4(OH)(OCH2)⁄⁄⁄⁄⁄ + H⁄ K5 1.10  104 4.02  103 1.76  103 8.86  102
DE5 = 0.42 eV, E 5 = 0.46 eV k+5, s1 6.46  107 2.08  108 5.51  108 1.26  109
6 C6H4(OH)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H4(OH)(O)⁄⁄⁄⁄ + CH⁄3 K6 2.80  1014 1.18  1013 8.76  1011 9.93  1010
DE6 = 1.37 eV, E 6 = 1.41 eV k+6, s1 1.19  102 3.57  101 5.98  100 6.42  101
7 C6H4(OH)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H4(O)(OCH3)⁄⁄⁄⁄ + H⁄ K7 1.43  108 2.11  107 4.37  106 1.17  106
DE7 = 0.80 eV, E 7 = 0.29 eV k+7, s1 5.41  109 1.14  1010 2.11  1010 3.56  1010
8 C6H4Hb(OH)(OCH3)⁄⁄⁄⁄ + 2⁄ ? C6H4Hb(OH)(OCH2)⁄⁄⁄⁄⁄ + H⁄ K8 1.35  101 9.38  100 7.00  100 5.52  100
DE8 = 0.16 eV, E 8 = 0.88 eV k+8, s1 1.43  103 1.21  104 7.14  104 3.19  105
9 C6H4Hb(OH)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H5OH⁄⁄⁄⁄ + CH3O⁄ K9 1.05  1011 1.20  1010 2.02  109 4.54  108
DE9 = 0.97 eV, E 9 = 0.76 eV k+9, s1 1.08  105 7.33  105 3.57  106 1.36  107
10 C6H4Ha(OH)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H5OCH⁄⁄⁄⁄ 3 + OH⁄ K10 5.08  109 7.16  108 1.43  108 3.73  107
DE10 = 0.88 eV, E 10 = 0.58 eV k+10, s1 1.47  107 6.52  107 2.24  108 6.36  108
11 C6H4OCH⁄⁄⁄⁄
3 + H⁄ ? C6H5OCH⁄⁄⁄⁄ 3 +⁄ K11 1.07  102 6.94  101 4.81  101 3.52  101
DE11 = 0.18 eV, E 11 = 0.67 eV k+11, s1 1.15  106 6.09  106 2.42  107 7.71  107
12 C6H4OH⁄⁄⁄⁄ + H⁄ ? C6H5OH⁄⁄⁄⁄ + ⁄ K12 8.83  101 5.81  101 4.08  101 3.02  101
DE12 = 0.17 eV, E 12 = 0.60 eV k+12, s1 7.84  106 3.56  107 1.25  108 3.58  108
13 C6H4(OH)(OCH2)⁄⁄⁄⁄⁄ + ⁄ ? C6H4(OH)⁄⁄⁄⁄ + CH2O⁄⁄ K13 4.03  102 2.54  102 1.74  102 1.27  102
DE13 = 0.22 eV, E 13 = 1.22 eV k+13, s1 4.90  101 9.48  100 1.10  102 8.66  102
14 C6H4(OH)(OCH2)⁄⁄⁄⁄⁄ + ⁄ ? C6H4(O)(OCH2)⁄⁄⁄⁄⁄ + H⁄ K14 3.89  107 6.24  106 1.38  106 3.90  105
DE14 = 0.76 eV, E 14 = 0.28 eV k+14, s1 4.42  109 8.96  109 1.62  1010 2.66  1010
15 C6H4(OH)(OCH2)⁄⁄⁄⁄⁄ + ⁄ ? C6H4(OH)(OCH)⁄⁄⁄⁄⁄ + H⁄ K15 4.79  105 1.33  105 4.65  104 1.93  104
DE15 = 0.55 eV, E 15 = 0.01 eV k+15, s1 4.84  1012 5.32  1012 5.79  1012 6.25  1012
16 C6H4(OH)(OCH2)⁄⁄⁄⁄⁄ ? C6H4(OH)(O)⁄⁄⁄⁄ + CH⁄2 K16 2.47  1013 1.39  1012 1.31  1011 1.80  1010
DE16 = 1.25 eV, E 16 = 0.45 eV k+16, s1 1.39  108 4.33  108 1.11  109 2.47  109
17 C6H4(OH)(O)⁄⁄⁄⁄ + H⁄ ? C6H4(OH)⁄⁄⁄⁄ 2 +⁄ K17 5.39  108 3.41  108 1.56  107 5.57  107
DE17 = 0.78 eV, E 17 = 1.06 eV k+17, s1 8.46  101 1.15  103 9.96  103 6.12  104
18 C6H4(OH)(O)⁄⁄⁄⁄ + H⁄ ? C6H4Ha(OH)(O)⁄⁄⁄⁄ + ⁄ K18 3.05  1010 2.21  109 1.12  108 4.39  108
DE18 = 0.87 eV, E 18 = 1.42 eV k+18, s1 7.62  103 2.35  101 3.56  100 4.33  101
19 C6H4(OH)(O)⁄⁄⁄⁄ + ⁄ ? C6H4O⁄⁄⁄⁄ + OH⁄ K19 3.11  106 1.13  105 3.32  105 8.23  105
DE19 = 0.53 eV, E 19 = 1.14 eV k+19, s1 3.59  100 5.57  101 5.40  102 3.65  103
20 C6H4(O)(OCH3)⁄⁄⁄⁄ + ⁄ ? C6H4O⁄⁄⁄⁄ + CH3O⁄ K20 7.15  103 1.40  102 2.45  102 3.93  102
DE20 = 0.26 eV, E 20 = 1.08 eV K+20, s1 3.00  101 4.18  102 3.71  103 2.33  104
21 C6H4(O)(OCH3)⁄⁄⁄⁄ + 2⁄ ? C6H4(O)(OCH2)⁄⁄⁄⁄⁄ + H⁄ K21 1.24  103 2.94  103 5.96  102 1.08  102
DE21 = 0.38 eV, E 21 = 0.49 eV k+21, s1 1.36  101 1.99  102 1.83  103 1.18  104
22 C6H4Hb(OH)(OCH2)⁄⁄⁄⁄⁄?C6H4Hb(OH)(O)⁄⁄⁄⁄ + CH⁄2 K22 1.21  109 1.77  108 3.63  107 9.66  106
DE22 = 0.85 eV, E 22 = 0.37 eV k+22, s1 1.24  109 3.21  109 7.04  109 1.37  1010
23 C6H5OCH⁄⁄⁄⁄
3 + ⁄ ? C6H5OCH⁄⁄⁄⁄ 2 + H⁄ K23 1.72  103 6.76  102 3.14  102 1.66  102
DE23 = 0.38 eV, E 23 = 0.44 eV k+23, s1 7.26  107 2.19  108 5.50  108 1.20  109
24 C6H4(OH)(OCH)⁄⁄⁄⁄⁄ + ⁄ ? C6H4(OH)⁄⁄⁄⁄ + CHO⁄⁄ K24 5.59  102 3.67  102 2.60  102 1.96  102
DE24 = 0.22 eV, E 24 = 1.03 eV k+24, s1 2.32  102 2.89  103 2.34  104 1.36  105
25 C6H4(OH)(OCH)⁄⁄⁄⁄⁄ ? C6H4(OH)(O)⁄⁄⁄⁄ + CH⁄ K25 6.83  1013 3.46  1012 2.96  1011 3.77  1010
DE25 = 1.29 eV, E 25 = 0.31 eV k+25, s1 3.83  109 8.46  109 1.63  1010 2.84  1010
26 C6H4(OH)⁄⁄⁄⁄
2 + ⁄ ? C6H4(OH)⁄⁄⁄⁄ + OH⁄ K26 6.89  101 5.27  101 4.26  101 3.58  101
DE26 = 0.15 eV, E 26 = 1.16 eV k+26, s1 3.96  100 6.52  101 6.62  102 4.66  103
27 C6H4(OH)⁄⁄⁄⁄
2 + H⁄ ? C6H5(OH)⁄⁄⁄⁄⁄ 2 K27 3.33  106 1.06  105 2.72  105 5.98  105
DE27 = 0.51 eV, E 27 = 1.16 eV k+27, s1 3.90  100 6.40  101 6.46  102 4.51  103
28 C6H4Ha(OH)(O) ⁄⁄⁄⁄
+ ? C6H5O⁄⁄⁄⁄ + OH⁄

K28 1.13  101 1.08  101 1.06  101 1.04  101
DE28 = 0.06 eV, E 28 = 0.81 eV k+28, s1 5.76  104 4.42  105 2.40  106 1.00  107
29 C6H4O⁄⁄⁄⁄ + H⁄ ? C6H5O⁄⁄⁄⁄ + ⁄ K29 5.39  109 6.29  108 1.06  108 2.37  107
+
DE29 = 0.91 eV, E29 
= 0.54 eV k29 , s1 1.72  107 6.77  107 2.11  108 5.50  108
30 C6H4(O)(OCH2)⁄⁄⁄⁄⁄ + ⁄ ? C6H4O⁄⁄⁄⁄ + CH2O⁄ K30 2.49  108 1.26  107 4.78  107 1.46  106
DE30 = 0.70 eV, E 30 = 1.29 eV k+30, s1 3.42  102 7.20  101 8.91  100 7.35  101
31 C6H4(O)(OCH2)⁄⁄⁄⁄⁄ + ⁄ ? C6H4O⁄⁄⁄⁄⁄ 2 + CH⁄2 K31 3.46  106 7.17  105 1.96  105 6.60  104
DE31 = 0.67 eV, E 31 = 0.79 eV k+31, s1 1.08  104 6.98  104 3.27  105 1.19  106
32 C6H4(O)(OCH2)⁄⁄⁄⁄⁄ + ⁄ ? C6H4(O)(OCH)⁄⁄⁄⁄⁄ + H⁄ K32 2.82  105 8.44  104 3.14  104 1.38  104
DE32 = 0.52 eV, E 32 = 0.01 eV k+32, s1 5.39  1012 6.04  1012 6.68  1012 7.31  1012
33 C6H4(O)(OCH)⁄⁄⁄⁄⁄ + ⁄ ? C6H4O2⁄⁄⁄⁄⁄ + CH⁄ K33 5.67  1010 5.73  109 8.67  108 1.79  108
DE33 = 1.00 eV, E 33 = 0.47 eV k+33, s1 1.56  108 5.16  108 1.40  109 3.23  109
34 C6H4(O)(OCH)⁄⁄⁄⁄⁄ + ⁄ ? C6H4(O)(OC)⁄⁄⁄⁄⁄ + H⁄ K34 5.36  103 2.35  103 1.19  103 6.79  102
DE34 = 0.35 eV, E 34 = 0.39 eV k+34, s1 7.76  108 2.11  109 4.83  109 9.75  109
35 C6H4Hb(OH)(O)⁄⁄⁄⁄⁄ + H⁄ ? C6H5(OH)⁄⁄⁄⁄⁄ 2 +⁄ K35 6.34  103 1.09  102 1.71  102 2.47  102
DE35 = 0.23 eV, E 35 = 0.90 eV k+35, s1 3.25  103 3.00  104 1.89  105 8.90  105
J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50 45

Table 4 (continued)

Step Reaction Constant T (K)


473 523 573 623
36 C6H5OCH⁄⁄⁄⁄2 + ⁄ ? C6H5O⁄⁄⁄⁄ + CH⁄2 K36 9.70  1013 5.14  1012 4.59  1011 6.09  1010
DE36 = 1.29 eV, E 36 = 0.12 eV k+36, s1 4.50  1011 6.56  1011 9.01  1011 1.18  1012
37 C6H5O⁄⁄⁄⁄ + H⁄ ? C6H5OH⁄⁄⁄⁄ + ⁄ K37 8.91  1010 7.18  109 4.01  108 1.70  107
DE37 = 0.88 eV, E 37 = 1.05 eV k+37, s1 7.19  101 9.24  102 7.63  103 4.51  104
38 C6H5OH⁄⁄⁄⁄ + ⁄ ? C6H⁄⁄⁄⁄
5 + OH⁄ K38 1.87  101 1.47  101 1.20  101 1.03  101
DE38 = 0.13 eV, E 38 = 1.15 eV k+38, s1 3.85  100 6.26  101 6.30  102 4.40  103
39 C6H5OH⁄⁄⁄⁄ + H⁄ ? C6H6OH⁄⁄⁄⁄ + ⁄ K39 9.44  107 3.10  106 8.20  106 1.85  105
DE39 = 0.53 eV, E 39 = 1.25 eV k+39, s1 4.39  101 8.92  100 1.08  102 8.69  102
40 C6H5(OH)⁄⁄⁄⁄
2 + ⁄ ? C6H5OH⁄⁄⁄⁄ + OH⁄ K40 8.59  108 1.47  108 3.43  107 1.02  107
DE40 = 0.80 eV, E 40 = 0.63 eV k+40, s1 4.99  106 2.48  107 9.36  107 2.87  108
41 C6H4O⁄⁄⁄⁄⁄
2 + H⁄ ? C6H4(OH)(O)⁄⁄⁄⁄ + 2⁄ K41 5.26  105 1.53  104 3.66  104 7.60  105
DE41 = 0.44 eV, E 41 = 0.96 eV k+41, s1 1.26  103 1.34  104 9.54  104 4.96  105
42 C6H4(O)(OC)⁄⁄⁄⁄⁄ ? C6H4O⁄⁄⁄⁄ + CO⁄ K42 2.06  105 8.48  104 4.09  104 2.22  104
DE42 = 0.41 eV, E 42 = 0.68 eV k+42, s1 9.47  105 5.19  106 2.13  107 6.97  107
43 C6H⁄⁄⁄⁄
5 + H⁄ ? C6H⁄⁄⁄
6 +2

K43 1.02  102 6.83  101 4.89  101 3.68  101
DE43 = 0.17 eV, E 43 = 0.55 eV k+43, s1 2.02  107 8.23  107 2.64  108 7.02  108
44 C6H6OH⁄⁄⁄⁄ ? C6H⁄⁄⁄ 6 + OH

K44 2.01  109 3.23  108 7.19  107 2.05  107
DE44 = 0.83 eV, E 44 = 0.56 eV k+44, s1 2.38  107 1.01  108 3.37  108 9.29  108
45 CH⁄ + H⁄ ? CH⁄2 + ⁄ K45 7.55  107 3.03  106 9.50  106 2.47  105
DE45 = 0.59 eV, E 45 = 0.62 eV k+45, s1 2.36  106 1.09  107 3.85  107 1.11  108
46 CH⁄2 + H⁄ ? CH⁄3 + ⁄ K46 1.03  103 2.12  103 3.81  103 6.21  103
DE46 = 0.30 eV, E 46 = 0.62 eV k+46, s1 2.30  106 1.07  107 3.78  107 1.09  108
47 CH⁄3 + H⁄ ? CH⁄4 + ⁄ K47 1.36  105 6.89  105 2.61  104 7.91  104
DE47 = 0.64 eV, E 47 = 1.13 eV k+47, s1 5.88  101 9.50  102 9.43  103 6.48  104
48 OH⁄ + H⁄ ? H2O⁄ + ⁄ K48 5.25  106 2.15  105 6.85  105 1.81  104
DE48 = 0.57 eV, E 48 = 1.18 eV k+48, s1 8.57  100 1.53  102 1.67  103 1.24  104
49 CHO⁄⁄ + H⁄ ? CH2O⁄⁄ + ⁄ K49 1.92  106 6.39  106 1.72  105 3.92  105
DE49 = 0.52 eV, E 49 = 0.50 eV k+49, s1 3.20  107 1.11  108 3.09  108 7.34  108
50 CH2O⁄⁄ + H⁄ ? CH3O⁄ + 2⁄ K50 2.74  102 4.59  102 6.98  102 9.83  102
DE50 = 0.21 eV, E 50 = 0.73 eV k+50, s1 5.07  105 3.49  106 1.56  107 5.46  107
51 CH3O⁄ + H⁄ ? CH3OH⁄ + ⁄ K51 1.65  108 1.03  107 4.63  107 1.63  106
DE51 = 0.77 eV, E 51 = 1.23 eV k+51, s1 1.02  100 2.02  101 2.38  102 1.90  103
52 C6H5OCH⁄⁄⁄⁄3 ? C6H5OCH3(gas) + 4⁄ K52 2.23  1011 3.13  108 1.52  105 3.27  103
DE52 = 2.54 eV k+52, s1 1.64  103 2.19  100 1.02  103 2.10  105
53 C6H4(OH)⁄⁄⁄⁄
2 ? C6H4(OH)2(gas) + 4⁄ K53 2.49  1012 3.92  109 2.07  106 4.71  104
DE53 = 2.60 eV k+53, s1 1.82  104 2.73  101 1.37  102 3.00  104
54 C6H5OH⁄⁄⁄⁄ ? C6H5OH(gas) + 4⁄ K54 2.35  1012 2.49  109 9.16  107 1.55  104
DE54 = 2.51 eV k+54, s1 1.86  104 1.87  101 6.58  101 1.06  104
55 C6H⁄⁄⁄
6 ? C6H6(gas) + 3

K55 1.96  1013 1.43  1010 3.81  108 4.74  106
DE55 = 2.42 eV k+55, s1 1.70  105 1.18  102 3.00  100 3.59  102
56 CH⁄4 ? CH4(gas) + ⁄ K56 1.01  107 2.63  107 6.02  107 1.27  108
DE56 = 0.22 eV k+56, s1 1.94  1015 4.80  1015 1.05  1016 2.13  1016
57 CH4OH⁄ ? CH4OH(gas) + ⁄ K57 3.58  103 3.48  104 2.47  105 1.34  106
DE57 = 0.75 eV k+57, s1 4.85  1011 4.48  1012 3.04  1013 1.58  1014
58 H2O⁄ ? H2O(gas) + ⁄ K57 4.74  102 2.82  103 1.28  104 4.71  104
DE58 = 0.59 eV k+57, s1 8.57  1010 4.84  1011 2.09  1012 7.42  1012
59 CO⁄ ? CO(gas) + ⁄ K58 3.63  1013 6.48  1011 4.88  109 1.86  107
DE59 = 2.05 eV k+58, s1 5.26  105 8.92  103 6.41  101 2.34  101
60 0.5H2(g) + ⁄ ? H⁄ K59 8.24  103 1.51  103 3.65  102 1.10  102
DE60 = 0.65 eV k+59, s1 bar1 7.66  108 7.28  108 6.96  108 6.67  108

step 21, the C6H4(O)(OCH2) intermediate prefers again further Finally, deoxygenation steps by direct phenyl(C)–O bond scis-
dehydrogenation to C6H4(O)(OCH) (step 32: DE0 = 0.52 eV; sion are not included in our reaction network because these steps
Ea = 0.01 eV) over –OCH2 removal (step 30: Ea = 1.29 eV) and – have very high reaction barriers. For instance, the elementary step
CH2 removal (step 31: Ea = 0.79 eV). After reaction step 32, –CH C6H4(O)(OCH3) ? C6H4(OCH3) + O has a high barrier of 1.77 eV, the
removal (step 33: Ea = 0.47 eV) and –OCH removal (step 34: elementary step C6H4O2 ? C6H4O + O has a barrier of 1.52 eV, and
Ea = 0.39 eV) are in close competition and only a microkinetic C6H5O ? C6H5 + O has a barrier of 1.87 eV.
model can predict the preferred reaction pathway. Step 33 will
be followed by step 41 (hydrogenation to C6H4(OH)(O); 3.4. Mean-field microkinetic modeling
DE0 = 0.44 eV; Ea = 0.96 eV) which connects the seventh reaction
pathway with the fifth at this intermediate. Reaction step 34 (last To better understand the meaning of the reaction energies, we
dehydrogenation of the methylene group; DE0 = 0.35 eV; developed a microkinetic model that considers lateral interactions
Ea = 0.90 eV) will be followed by a decarbonylation (step 42) to approximately using a method similar to the one proposed by Gra-
form C6H4O and CO which is an exothermic process (DE0 = - bow et al. [32] We note that other functional forms for describing
0.41 eV) with a modest barrier (Ea = 0.68 eV). Next, C6H4O is lateral interactions have been proposed and that we selected the
hydrogenated to phenol in steps 29 (DE0 = 0.91 eV; Ea = 0.51 eV) method from Grabow et al. primarily for its simplicity. Microkinet-
and 37 (DE0 = 0.88 eV; Ea = 1.05 eV). Overall, the seventh pathway ic modeling without lateral interactions suggests that H, CO, and
is likely to proceed by C6H4(OH)(OCH3) ? C6H4(O)(OCH3) + phenoxy (C6H5O) are the three most abundant surface intermedi-
H ? C6H4(O)(OCH2) + 2H ? C6H4(O)(OCH) + 3H ? C6H4(O)(OC) + ates and it is most important to consider their lateral interactions
4H ? C6H4O + CO + 4H ?    ? C6H5OH + CO + 2H. approximately. From adsorption energy calculations of H at various
46 J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50

Fig. 4. Side (upper panel) and top view (lower panel) of preferred adsorption structure of various small intermediates in the reaction network of the hydrodeoxygenation of
guaiacol over Ru(0 0 0 1).

coverages (hH = 1/4 ML, 2/4 ML, 3/4 ML, 4/4 ML), we determined a ECO ðhCO ;hphenoxy Þ ¼ 2:126 þ 2  0:843  ðhCO  0:092Þ þ 0:851hphenoxy
hydrogen coverage-dependent differential adsorption energy, þ 1:5  7:503  hphenoxy ðhphenoxy hCO Þ0:5 ð7Þ
EH(hH), for the reaction 0.5H2(gas)+⁄ ? H⁄ shown in Eq. (2).
EH ðhH Þ ¼ 0:649 þ 2  0:075  ðhH  0:100Þ ð2Þ Ephenoxy ðhCO ; hphenoxy Þ ¼ 4:438 þ 2  4:508  ðhphenoxy  0:041Þ
From adsorption energy calculations of CO at various coverages þ 0:851hCO þ 1:5  7:503  hCO ðhphenoxy hH Þ0:5
(hCO = 1/4 ML, 2/4 ML, 3/4 ML, 4/4 ML), we determined a CO cover-
ð8Þ
age-dependent differential adsorption energy, ECO(hCO) as shown in
Eq. (3). Eqs. (5) and (7) are implemented in our microkinetic model.
ECO ðhCO Þ ¼ 2:126 þ 2  0:843  ðhCO  0:092Þ ð3Þ Considering that we do not have a meaningful water–gas shift
or methanation model for Ru catalysts, we fixed in our microkinet-
Similarly, from adsorption energy calculations of phenoxy at ic model the partial pressures of the reactants, guaiacol, and H2, to
various coverages (hphenoxy = 1/20 ML, 1/16 ML, 1/12 ML, 1/9 ML), 1 bar. For the reaction products, phenol, catechol, benzene, and
we determined a phenoxy coverage-dependent differential water, we choose low-conversion conditions of 106 bar. For CO,
adsorption energy of phenoxy, Ephenoxy(hphenoxy) shown in Eq. (4). we choose a slightly higher pressure of 104 bar to observe the
Ephenoxy ðhphenoxy Þ ¼ 4:438 þ 2  4:508  ðhphenoxy  0:041Þ ð4Þ effect of CO to the reaction mechanism. In such a reaction environ-
ment, we observe an H coverage of 4.8%, a CO coverage of 2.2% and
Note that we built p(5  4)  4, p(3  4)  4 slabs, and a phenoxy coverage of 22.4%, that is, phenoxy occupies 89.6% of the
p(3  3)  4 slabs to get adsorption energies of phenoxy at cover- surface sites since one phenoxy occupies 4 Ru atoms. Other surface
ages of hphenoxy = 1/20 ML, 1/12 ML, and 1/9 ML. We used a k-point intermediates with noticeable coverage are CH (0.73%) and C6H4(-
mesh of 4  5  1 for p(5  4)  4 slabs, 5  5  1 for p(3  4)  4 OH)(O) (0.50%). The fraction of free sites is 0.17%.
slabs, and 6  6  1 for the p(3  3)  4 slabs. All other parameters Our microkinetic model shown in Fig. 7 suggests that hydroge-
of the DFT calculations are the same as those for the (4  4)  4 nation steps of the aromatic ring are not favorable at a hydrogen
slabs. pressure of 1 bar. Hydrogenations of the b-carbon (pathway 1)
Coadsorption of H and phenoxy at various coverages on and a-carbon (pathway 2) are slow with TOFs of 1.3  109 s1
p(3  3)  4 slabs (hphenoxy = 1/9 ML and hH = 1/9 ML, 2/9 ML, 3/ and 5.9  1010 s1, respectively. The model also shows that path-
9 ML, 4/9 ML, 5/9 ML) leads to EH(hH, hphenoxy) and Ephenoxy(hH, ways starting with direct removal of –OH (pathway 3), –OCH3
hphenoxy), respectively. (pathway 4), and –CH3 (pathway 6) groups are very slow with TOFs
EH ðhH ; hphenoxy Þ ¼ 0:649 þ 2  0:075  ðhH  0:100Þ  1:124hphenoxy of 1.9  1010 s1, 4.3  109 s1, and 1.4  1013 s1, respectively.
Next, the fifth pathway has a TOF of 1.9  108 s1, which is 4
þ 1:5  11:043  hphenoxy ðhphenoxy hH Þ0:5 ð5Þ orders of magnitude slower than pathway 7 (3.6  104 s1). We
conclude that demethylation of guaiacol to form catechol is much
Ephenoxy ðhH ; hphenoxy Þ ¼ 4:438 þ 2  4:581  ðhphenoxy  0:041Þ slower than HDO to phenol. The observation that pathway 7 is
 1:124hH þ 1:5  11:043  hH ðhphenoxy hH Þ0:5 dominant suggests that the O–H bond in guaiacol is quite easily
broken when adsorbed on a Ru surface. The intermediate thus cre-
ð6Þ
ated, C6H4(O)(OCH3), will proceed by –OCH3 dehydrogenation to
Finally, coadsorption of H and phenoxy at various coverages on form C6H4(O)(OCH2). Alternative removal of –OCH3 (step 20) is
p(3  3)  4 slabs (hphenoxy = 1/9 ML and hCO = 1/9 ML, 2/9 ML, 3/ slow (TOF = 3.2  106 s1). C6H4(O)(OCH2) follows further dehy-
9 ML, 4/9 ML, 5/9 ML) leads to ECO(hCO, hphenoxy) and Ephenoxy(hCO, drogenation to form C6H4(O)(OCH) because neither –OCH2 removal
hphenoxy), respectively. (9.8  1016 s1) nor –CH2 removal (3.6  1011 s1) are favorable.
J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50 47

Fig. 5. Transition state structures of elementary steps 1–24 on the Ru(0 0 0 1) surface. Upper panels are for side views and lower ones for top views.

C6H4(O)(OCH) then reacts by –CH removal to form C6H4O2 an appreciable amount of benzene (5% yield) and aromatic ring-
(7.5  105 s1) and –CHO dehydrogenation to form C6H4(O)(OC) saturated products such as cyclohexanone (10% yield) and cyclo-
(2.8  104 s1), which are two competitive pathways. C6H4O2 is hexanol (5% yield) were observed. They also observed that the
first hydrogenated to C6H4(O)(OH) and then dehydroxylated to selectivity to benzene, cyclohexanone, and cyclohexanol increased
C6H4O, before it is hydrogenated to phenol. C6H4(O)(OC) will follow at the expense of phenol, suggesting that benzene was formed by
a decarbonylation step to form C6H4O, which is then hydrogenated hydrodeoxygenation of phenol, and cyclohexanone and cyclohexa-
to phenol. Phenol will desorb since both the –OH removal pathway nol were formed by ring saturation of phenol. This is partly con-
(TOF = 2.0  107 s1) and the a-carbon hydrogenation followed firmed by our calculations since phenyl ring saturation of
by –OH removal (5.4  107 s1) are unfavorable. It is interesting guaiacol (pathways 1 and 2) before dehydrogenation and phenol
to note though that the –OH removal by hydrogenation of the phe- formation is found to be unfavorable.
nyl ring is slightly preferred to the direct –OH removal even at low Elliott et al. [6] investigated the HDO of guaiacol over Ru/C cat-
hydrogen partial pressures. We predict that at higher hydrogen alysts at 250 °C and high H2 partial pressure and observed aromatic
partial pressures benzene production will primarily occur by par- ring-saturated products such as cyclohexanol, 2-methoxycyclo-
tial phenyl ring hydrogenation which can of course also lead to hexanol, and cyclohexane as the major products. In agreement to
cyclohexane production not investigated in this study. Next, calcu- our calculations, they also observed phenol as a major early prod-
lations based on DFT suggest that hydrogenation of C6H4(O)(OH) to uct that is later converted to cyclohexanone and cyclohexanol.
catechol is less favorable than its dehydroxylation to C6H4O Also, similar reaction mechanisms have been found on precious
(3.9  106 s1 versus 7.5  105 s1). Overall, the major product metal catalysts such as Pd- and Pt-based catalysts [5,10]. For
is predicted to be phenol, and catechol is the most important side example, Sun et al. [5] proposed that on Pd/C and Pt/C, the favor-
product. able pathway should be first demethoxylation to form phenol fol-
Sun et al. [5] recently observed experimentally that phenol was lowed by hydrogenation to cyclohexanone/cyclohexanol. These
the major product (61.6% yield) of guaiacol HDO on a Ru/C cata- results stand in contrast to experiments by Lin et al. [9] who pro-
lysts at 250 °C at low guaiacol and H2 partial pressures. In addition, posed that on Rh-based catalysts, the HDO of guaiacol proceeds
48 J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50

Fig. 6. Transition state structures of elementary steps 25–51 on the Ru(0 0 0 1) surface. Upper panels are for side views and lower ones for top views.

by hydrogenation of the aromatic ring to form 2-methoxycyclo- 3.5.2. Degree of thermodynamic rate control
hexanone and 2-methoxycyclohexanol intermediates, followed The degree of thermodynamic rate control [35], XTRC, is used to
by demethoxylation and/or dehydroxylation to yield cyclohexa- analyze the sensitivity of the microkinetic model with regard to
none/cyclohexanol and cyclohexane products. adsorbed intermediates

3.5. Sensitivity analysis 0 1


1 @ @r A
X TRC;n ¼   ð10Þ
3.5.1. Degree of rate control r @ G0n
RT
We have used Campbell’s degree of rate control [33,34], XRC, to G0m–n ;G0;TS
i

determine the rate-controlling steps in the mechanism. The degree


of rate control is defined as where r is the overall rate of reaction; Gn0 is the free energy of
 
ki @r adsorbate n. Under the reaction conditions mentioned above, the
X RC;i ¼ ð9Þ surface intermediate C6H4(OH)(O) has a degree of thermodynamic
r @ki K i ;kj –ki
rate control of 0.4. This relatively large thermodynamic rate con-
where r is the overall rate of reaction. Under the reaction conditions trol for C6H4(OH)(O) can be understood by our observation that sta-
mentioned above, we found the largest degree of rate control of bilizing it deceases the already small surface coverage of various
0.5 for the OC–H cleavage of C6H4(O)(OCH)), suggesting that dehy- species on the dominant reaction pathway. Next, CO has a XTRC of
drogenation steps are at least partially rate controlling. Next, we 0.25 indicating its ability to poison the Ru surface. Finally, we
observed for a few other steps, such as the guaiacol adsorption observe some sensitivity of our results to the adsorption strength
and possibly hydrogenation of phenoxy to phenol, some sensitivity of H and CH. Unfortunately, no meaningful thermodynamic rate
for the turnover frequency likely due to the relatively small free site control values can be reported due to the large numerical noise in
and hydrogen coverage. However, due to numerical inaccuracies in the calculations of these XTRC values. However, the XTRC values are
solving the large network of elementary steps and our approximate likely small and positive which can be explained by a higher hydro-
approach for describing lateral interactions, no reliable rate control gen coverage and a reduced reverse reaction for step 33 if H and CH
values could be computed. are stabilized on the Ru surface.
J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50 49

Fig. 7. Turnover frequencies (s1) of all elementary steps for a reactant pressure of Pguaiacol = PH2 = 1 bar and a temperature of 573 K. For clarity, duplicate structures are
highlighted by identical background colors. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

3.6. Computed apparent reaction orders further deoxygenation of phenol to benzene is found to be very
challenging over a Ru(0 0 0 1) catalyst surface. Catechol is predicted
Finally, we calculated the order of reactions (ai) at 573 K with to be the most relevant side product. Computations predict that
respect to guaiacol and H2, respectively, to understand the sensitiv- the final dehydrogenation step of the methoxy species in guaiacol
ity of our results to changes in partial pressure. is at least partially rate controlling over Ru catalysts. Interestingly,
  direct removal of an oxygen atom bound to a phenyl ring is unli-
@ lnðrÞ kely to occur over the Ru(0 0 0 1) as deoxygenation only occurs by
ai ¼ ð11Þ
@ lnðpi Þ T;P decarbonylation and dehydroxylation. Finally, hydrogenation of
the phenyl ring is energetically demanding such that aromatic
Our model predicts a small guaiacol and H2 reaction order of
products can be produced over Ru catalysts at mild hydrogenation
0.1 in the range of 0.2–1.6 bar for guaiacol and 0.2–1.0 bar for
conditions. At higher hydrogen partial pressures, we predict –OH
H2, indicating that in this low pressure regime, our results are
removal from phenol to occur by partial phenyl ring hydrogena-
not very sensitive to the reactant partial pressures. Finally, we
tion, which can lead to benzene and cyclohexane production.
compute a CO reaction order of 0.4 in the pressure range of
105–103 bar, indicating that a good water–gas shift or methana-
tion activity is required for a good HDO catalyst to keep the CO par- Acknowledgments
tial pressure low and avoid catalyst poisoning by CO.
We acknowledge Dr. Aravind Asthagiri from the Ohio State Uni-
versity for helping us obtain and install the vasp D3 corrections for
4. Conclusions dispersion interactions. This work has been supported by the U.S.
Department of Energy, Office of Basic Energy Sciences, Chemical
We have built a microkinetic model based on parameters Sciences Division under Contract DE-FG02-11ER16268 (DE-
obtained from density functional theory calculations to investigate SC0007167). Computational resources have been provided by the
the hydrodeoxygenation mechanism of guaiacol over Ru(0 0 0 1) National Energy Research Scientific Computing Center (NERSC)
model surfaces. We find that at low hydrogen partial pressure, which is supported by the Office of Science of the U.S. Department
the kinetically most favorable pathway proceeds via dehydrogena- of Energy and in part by XSEDE under Grant Number TG-
tion of the hydroxyl group of guaiacol [C6H4(OH)(OCH3)] to yield CTS090100. Finally, computing resources from the USC NanoCenter
C6H4(O)(OCH3), followed by full dehydrogenation of the methoxy and USC’s High Performance Computing Group are gratefully
group to yield C6H4(O)(OC). Next, C6H4(O)(OC) decarbonylates to acknowledged.
produce C6H4O and CO, followed by hydrogenation to produce
phenol (C6H5OH). At the adsorbed C6H4(O)(OCH) intermediate, a
competitive deoxygenation pathway is identified that involves References
methyne group removal to C6H4O2, followed by hydrogenation to
[1] S. Xiu, A. Shahbazi, Ren., Sust. Energy Rev. 16 (2012) 4406–4414.
C6H4(OH)(O), dehydroxylation to C6H4O, and finally hydrogenation [2] A. Oasmaa, E. Kuoppala, A. Ardiyanti, R.H. Venderbosch, H.J. Heeres, Energy
to phenol. Phenol is found to be the major reaction product, and Fuels 24 (2010) 5264–5272.
50 J. Lu, A. Heyden / Journal of Catalysis 321 (2015) 39–50

[3] T. Sfetsas, C. Michailof, A. Lappas, Q. Li, B. Kneale, J. Chromatogr. A 1218 (2011) [20] V.A. Finkel, G.P. Kovtun, M.I. Palatnik, Phys. Met. Metallogr. 32 (1971) 231–235.
3317–3325. [21] G. Henkelman, B.P. Uberuaga, H. Jonsson, J. Chem. Phys. 113 (2000) 9901–
[4] R.N. Olcese, M. Bettahar, D. Petitjean, B. Malaman, F. Giovanella, A. Dufour, 9904.
Appl. Catal. B: Environ. 115–116 (2012) 63–73. [22] G. Henkelman, H. Jonsson, J. Chem. Phys. 111 (1999) 7010–7022.
[5] J.M. Sun, A.M. Karim, H. Zhang, L. Kovarik, X.S. Li, A.J. Hensley, J.-S. McEwen, Y. [23] R.A. Olsen, G.J. Kroes, G. Henkelman, A. Arnaldsson, H. Jonsson, J. Chem. Phys.
Wang, J. Catal. 306 (2013) 47–57. 121 (2004) 9776–9792.
[6] D.C. Elliott, T.R. Hart, Energy Fuels 23 (2009) 631–637. [24] A. Heyden, A.T. Bell, F.J. Keil, J. Chem. Phys. 123 (2005) 224101.
[7] C. Zhao, J. He, A.A. Lemonidou, X. Li, J.A. Lercher, J. Catal. 280 (2011) 8–16. [25] J.M. Lu, S. Behtash, M. Faheem, A. Heyden, J. Catal. 305 (2013) 56–66.
[8] A. Gutierrez, R.K. Kaila, M.L. Honkela, R. Slioor, A.O.I. Krause, Catal. Today 147 [26] G. Buzzi-Ferraris, BzzMath: Numerical libraries in C++, Politecnico di Milano
(2009) 239–246. <www.chem.polimi.it/homes/gbuzzi>.
[9] Y. Lin, C. Li, H. Wan, H. Lee, C. Liu, Energy Fuels 25 (2011) 890–896. [27] A. Antony, C. Hakanoglu, A. Asthagiri, J.F. Weaver, J. Chem. Phys. 136 (2012)
[10] T. Nimmanwudipong, R.C. Runnebaum, D.E. Block, B.C. Gates, Energy Fuels 25 054702.
(2011) 3417–3427. [28] X.L. Zhou, M.E. Castro, J.M. White, Surf. Sci. 238 (1990) 215–225.
[11] V.N. Bui, G. Toussaint, D. Laurenti, C. Mirodatos, C. Geantet, Catal. Today 143 [29] W. Liu, V.G. Ruiz, G.X. Zhang, B. Santra, X.G. Ren, M. Scheffler, A. Tkatchenko,
(2009) 172–178. New J. Phys. 15 (2013) 053046.
[12] M.A. Gonzalez-Borja, D.E. Resasco, Energy Fuels 25 (2011) 4155–4162. [30] O. Dulaurent, M. Nawdali, A. Bourane, D. Bianchi, Appl. Catal. A – Gen. 201
[13] M.V. Bykova, D.Y. Ermakov, V.V. Kaichev, O.A. Bulavchenko, A.A. Saraev, M.Y. (2000) 271–279.
Lebedev, V.A. Yakovlev, Appl. Catal. B: Environ. 113–114 (2012) 296–307. [31] G. Mercurio, E. McNellis, I. Martin, S. Hagen, F. Leyssner, S. Soubatch, J. Meyer,
[14] G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558–561. M. Wolf, P. Tegeder, F. Tautz, Phys. Rev. Lett. 104 (2010) 036102.
[15] G. Kresse, J. Furthmuller, Comput. Mater. Sci. 6 (1996) 15–50. [32] L.C. Grabow, B. Hovlbak, J.K. Norskov, Top. Catal. 53 (2010) 298–310.
[16] G. Kresse, D. Joubert, Phys. Rev. B 59 (1999) 1758–1775. [33] C.T. Campbell, Top. Catal. 1 (1994) 353–366.
[17] J.P. Perdew, Y. Wang, Phys. Rev. B 33 (1986) 8800–8802. [34] C.T. Campbell, J. Catal. 204 (2001) 520–524.
[18] J.P. Perdew, Y. Wang, Phys. Rev. B 45 (1992) 13244–13249. [35] C. Stegelmann, A. Andreasen, C. Callaghan, J. Am. Chem. Soc. 131 (2009) 8077–
[19] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, J. Chem. Phys. 132 (2010) 154104. 8083.

Вам также может понравиться