You are on page 1of 18

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Review

pubs.acs.org/acscatalysis

Bifunctional Catalysts for Upgrading of Biomass-Derived


Oxygenates: A Review
Allison M. Robinson,† Jesse E. Hensley,‡ and J. Will Medlin*,†

Department of Chemical and Biological Engineering, University of ColoradoBoulder, UCB 596, Boulder, Colorado 80309, United
States

National Bioenergy Center, National Renewable Energy Laboratory, 15013 Denver West Pikeway, Golden, Colorado 80401, United
States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via 88.202.186.59 on October 16, 2019 at 04:38:12 (UTC).

ABSTRACT: Deoxygenation is an important reaction in the conversion of biomass-derived oxygenates to fuels and chemicals. A
key route for biomass refining involves the production of pyrolysis oil through rapid heating of the raw biomass feedstock.
Pyrolysis oil as produced is highly oxygenated, so the feasibility of this approach depends in large part on the ability to selectively
deoxygenate pyrolysis oil components to create a stream of high-value finished products. Identification of catalytic materials that
are active and selective for deoxygenation of pyrolysis oil components has therefore represented a major research area. One
catalyst is rarely capable of performing the different types of elementary reaction steps required to deoxygenate biomass-derived
compounds. For this reason, considerable attention has been placed on bifunctional catalysts, where two different active materials
are used to provide catalytic sites for diverse reaction steps. Here, we review recent trends in the development of catalysts, with a
focus on catalysts for which a bifunctional effect has been proposed. We summarize recent studies of hydrodeoxygenation
(HDO) of pyrolysis oil and model compounds for a range of materials, including supported metal and bimetallic catalysts as well
as transition-metal oxides, sulfides, carbides, nitrides, and phosphides. Particular emphasis is placed on how catalyst structure can
be related to performance via molecular-level mechanisms. These studies demonstrate the importance of catalyst bifunctionality,
with each class of materials requiring hydrogenation and C−O scission sites to perform HDO at reasonable rates.
KEYWORDS: bifunctional, catalyst, hydrodeoxygenation, bimetallic, metal−metal oxide, tethered bifunctional catalysts

1. INTRODUCTION compounds, making upgrading through deoxygenation critical


Increasing environmental, economic, and political motivation for improving heating value, stability, and physical properties
to develop renewable liquid fuels and chemicals has generated such as acidity and viscosity.
significant interest in using biomass as a sustainable carbon Conversion of the compounds making up pyrolysis oil
source. Upgrading of lignocellulosic biomass provides a unique requires new types of catalysts that can selectively remove
opportunity to utilize a renewable resource to create these fuels oxygen from the oil. 9−11 One common strategy for
and chemicals from waste materials and plants that do not deoxygenating whole pyrolysis oil is hydrodeoxygenation
compete with food production. Through fast pyrolysis, the (HDO), which involves adding hydrogen to accomplish C−O
biomass is rapidly heated in the absence of oxygen to produce a bond scission via removal of water. This is typically done
condensable mixture of oxygenates along with char and light sequentially, with a hydrogenation-dehydration-hydrogenation
gases.1−3 Pyrolysis oil composition varies depending on the route in which dehydration often occurs over an acid site.
feedstock and process conditions, but it contains high Alternatively, direct deoxygenation (DDO) does not require
concentrations of phenolic compounds from lignin decom- intermediate hydrogenation steps; instead, the C−O bond is
position and furanic compounds derived from cellulose, in directly cleaved from the reactant through hydrogenolysis. It
addition to organic acids and other low molecular weight
oxygenates.4 This complex mixture of oxygenates generally has Received: March 31, 2016
poor properties as a fuel.1,5,6 The undesirable fuel character- Revised: June 18, 2016
istics stem from a high concentration of highly oxygenated Published: June 21, 2016

© 2016 American Chemical Society 5026 DOI: 10.1021/acscatal.6b00923


ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

should be noted that while these definitions of HDO and DDO gaining an understanding of not only how the current state-of-
are used throughout this manuscript, some previous reports the-art deoxygenation catalysts perform but also how they
have used different terminology for the same reactions. Other relate to one another.
reactions can also take place under these conditions, including
decarboxylation/decarbonylation, cracking, and hydrogenation 2. BIFUNCTIONAL CATALYSIS
(Figure 1).7,8 It should be noted that while HDO reactions The general concept behind bifunctional heterogeneous
catalysis is that there are two distinct types of active sites that
function in tandem to perform a surface-catalyzed reaction.
Often, these two types of sites are expected to catalyze different
elementary steps within an overall reaction. In principle,
however, the two sites could participate in the same step, for
example through interaction of different parts of an adsorbed
reactant molecule with the catalyst as in a concerted reaction. In
this section, we highlight two major types of materials used as
bifunctional catalysts before considering the application of
similar types of materials in biomass refining reactions.
Proposed mechanisms for two major classes of catalysts
discussed in this review are shown schematically below in
Figure 2a,b. To point out the connections between these
examples for simple reactants of well-studied reactions and the
oxygenate upgrading reactions of primary interest for this
review, an example of bifunctional catalysts for deoxygenation
of phenolic oxygenates is shown in Figure 2c.
2.1. Examples of Bifunctional Catalysis. 2.1.1. Bimetallic
Catalysts. One of the most common methods for controlling
the catalytic performance of supported metal catalysts is via
control over the composition of bimetallic catalysts. In a large
number of cases, adding one metal to another is not intended
to introduce a new catalytic function, but instead, the intent is
to modify the activity of the original metal. A second metal may
influence the properties of the first via ligand effects, which is
often interpreted in terms of a shifting of the d-band density of
states of the first metal. A much-studied example of this
phenomenon is in the performance of Pt “skin” catalysts that
have been modified by 3d elements such as Ni, Co, Fe, and Ti
Figure 1. Example reactions discussed in this review. in the subsurface for improved performance in the electro-
catalytic oxygen reduction reaction.13−15 In many of the
often produce saturated compounds and DDO may lead systems dominated by this effect, the modifying metal is
directly to aromatic compounds, aromatic and saturated assumed to not be present in the surface layer and thus has little
products can also be interconverted subsequent to being opportunity to directly create a bifunctional effect.16 Another
formed. For example, increasing the temperature shifts the mechanism by which addition of a second metal can modify a
thermodynamic equilibrium toward aromatic products such as catalyst surface is via the so-called “ensemble” effect.17 For
toluene production over methylcyclohexane. example, addition of (relatively unreactive) Au to Pd olefin
A wide variety of materials have been studied for pyrolysis oil hydrogenation catalysts has yielded an improved activity that
upgrading, ranging from carbides and nitrides to supported has been attributed to a decrease in the average number of
transition metals, but despite their diversity, the most promising contiguous Pd atoms; contiguous surface Pd clusters have been
materials are bifunctional: they contain disparate active sites associated with accumulation of unreactive spectator species.18
that catalyze complementary reaction steps. In many cases, In some cases, however, the presence of two metals at the
upgrading reactions demand a combination of the different surface under reaction conditions has been associated with a
reaction types described in Figure 1; it is natural that each step bifunctional effect. Bifunctional effects have been invoked, for
would require its own catalyst. example, in some studies involving PdAu catalysts. Xu and co-
In this contribution, we review advances in catalysts currently workers found that biphasic PdAu nanoparticleswhere
used for HDO with a focus on bifunctional catalysts. There segregation of the two metals into relatively Pd-rich and Au-
have been a wide variety of studies attempting to better richwere highly active for low-temperature CO oxidation.10
understand how HDO proceeds over different catalytic The authors found that Pd-rich sites were needed to carry out
materials using various probe molecules or whole oils. the dissociative adsorption of oxygen, whereas Au-rich sites
Comparing these approaches may help elucidate which types were active for CO oxidation, as shown schematically in Figure
of materials are most promising for the end goal of making 2a. The dispersion of the two metals in an active catalyst has
pyrolysis oil a viable alternative to petroleum in fuel and been seen experimentally through STM images of PdAu alloy
chemical production. Although pyrolysis oil is known to electrodes. In this example, PdAu thin films were electro-
contain many different compounds and types of functional deposited on Au(111), resulting in Pd monomers, dimers, and
groups,9 the majority of the studies referenced in this report use trimers on the surface (Figure 3).19 Scanning tunneling
phenolic compounds as probe molecules. This will aid in microscopy was used to obtain TM images of different PdAu
5027 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

Figure 2. Schematic representation of proposed reaction mechanisms by (a) bimetallic catalysts,10 (b) metal−metal oxide catalysts,11 and (c)
bimetallic catalysts specifically for deoxygenation.12

large Pd clusters (or a pure Pd surface) had a high


concentration of bridge/3-fold bound CO.19 This high
dispersion creates many of the neighboring Pd−Au sites
responsible for the bifunctional reaction shown schematically in
Figure 2a.
Similar effects have been proposed in olefin hydrogenation,
where Pd sites are necessary for dissociative hydrogen
adsorption while Au sites are active for hydrogenation.20
Bimetallic catalysts have also been employed such that different
types of sites are effective for reactions under different
conditions. For example, PtAu nanoparticles have been applied
Figure 3. Surface coverage of monomers, dimers, and trimers of Pd for
as electrocatalysts for lithium−air batteries. The Au component
two surfaces along with experimental results for the coverage of H and was found to be more active for the oxygen reduction reaction
CO on these two electrodes. Reprinted with permission from ref 19. (which occurs during battery discharging), while Pt sites were
Copyright 2001 American Association for the Advancement of associated with the reverse oxygen evolution reaction (which
Science. occurs during charging).21 The utility of combining two active
sites in one material has also been shown for use in unitized
compositions, which were found to contain highly dispersed Pd regenerative fuel cells (URFCs). These electrochemical cells
atoms. This dispersion is correlated with activity in Figure 3; operate in one mode to electrolyze water, producing hydrogen
the coverage of hydrogen and CO on the surface was measured and oxygen, then operate in fuel cell mode to recombine them
along with the coverage of monomers, dimers, and trimers of and produce electricity.22,23 The challenge with this concept is
Pd.19 Because CO adsorbs on Pd sites, the coverage of CO designing an electrocatalyst that is active for all of the reactions
increases with Pd loading. Furthermore, IR results showed that involved; catalysts active for the oxygen reduction reaction
the CO adsorbed on a top site in these disperse samples, while (ORR) perform poorly for the oxygen evolution reaction
5028 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

(OER).22,23 Finding a catalyst that is both stable and active for studied for biomass deoxygenation is their bifunctionality. As
OER, which involves oxidizing water to molecular oxygen in discussed later in this review, the importance of having sites
acidic media, has represented a major research focus in capable of activating hydrogen in close proximity to sites that
electrocatalysis. This is due in part to the limited studies on are effective for C−O scission has been proposed many times
the relationship between catalyst active site structure and OER throughout the literature for materials ranging from supported
activity.24 While Pt is nearly ideal for ORR, it has shown poor transition metals to traditional hydrotreating catalysts such as
OER activity. The lower rates have been attributed to reduced sulfides.32−44 Often these two reaction steps cannot be
surface area and formation of a poorly conductive PtOx layer effectively catalyzed by one type of active site. As a result, the
that resists current flow under the conditions of OER. This may most commonly proposed mechanisms for deoxygenation of
be overcome by using a bifunctional catalyst consisting of the these compounds involve one active site (such as a transition
conventional Pt particles used for ORR supported on an oxide metal) that readily activates hydrogen and another type of site
such as IrO2 for the oxygen evolution reaction.22,23 nearby, such as an acid site on an oxide support, for C−O bond
2.1.2. Metal−Metal Oxide Catalysts. A frequently used type activation. Although this can be studied using whole pyrolysis
of bifunctional catalyst combines metal sites that are active for oil, the complex mixture of oxygenates present makes it very
hydrogenation/dehydrogenation and acid sites that are active difficult to elucidate reaction pathways to understand
for protonation/deprotonation. For example, catalytic metals mechanisms. For this reason, it can be useful to instead
may be supported on a zeolite. In the isomerization of alkanes, consider more detailed studies using model compounds.
the saturated hydrocarbons are dehydrogenated to alkenes on Phenolic compounds are useful probe molecules for pyrolysis
the metal sites; the alkenes can subsequently migrate to acid upgrading reactions; they make up a large fraction of the oil,
sites, which protonate the alkenes to form carbenium ions.25 and the C−O bonds in these molecules are very difficult to
Structural rearrangement of the carbenium ion can lead to break.33,45−47 Other processes more representative of alter-
formation of branched structures, which upon deprotonation native conversion approaches, such as aqueous-phase process-
yield branched alkenes. Finally, the branched alkenes are ing of polyols, are also described below.
hydrogenated on metal sites. While these reactions can be
catalyzed by physical mixtures of metal and zeolite catalysts, 3. BIFUNCTIONAL CATALYSTS FOR
achieving intimate contact between the different types of DEOXYGENATION
reactive sites significantly improves the kinetics. Producing 3.1. Supported Metals. Many supported metal catalysts
zeolite-supported metals with high dispersion is important for have been investigated for deoxygenation reactions. While it is
improving the intimacy of contact and achieving rapid difficult to directly compare the variety of different catalysts
isomerization kinetics.26 tested under varying reaction conditions, it may be informative
The intimacy of metal−metal oxide contact has also been to look at general trends in performance that has been observed
found to be important in other reactions, with considerable in the literature. Figure 4 shows the selectivity to deoxygenated
focus devoted to Au nanoparticles dispersed on reducible oxide products when feeding aromatic oxygenates including cresol
supports.27 For example, Au supported on ceria28,29 or titania30 and guaiacol.
has been found to be active for the water−gas shift reaction at
relatively low temperatures, ranging from approximately 200−
350 °C. The metal oxide support has been reported to be active
for the required water dissociation step, whereas CO oxidation
has been associated with sites at the Au−metal oxide interface.
A related mechanism has been demonstrated on Pt/ZrO2
catalysts for dry reforming of methane (Figure 2b). It was
shown that the Pt metal serves to activate methane,
dehydrogenating it to form CHx, while sites at the Pt-ZrO2
interface are responsible for CO2 dissociation. Achieving good
catalyst performance was found to require optimization of the
interfacial contact area.11 Furthermore, bifunctional metal−
metal oxide catalysts in which the oxide serves as a base have
also been reported. For example, isobutanol may be produced Figure 4. Selectivity to deoxygenated products from feeding vapor-
using syngas (CO and H2) on bifunctional catalysts consisting phase aromatic oxygenates (cresol or guaiacol) over Pt (diamonds),
Ru (crosses), Fe (squares), Pd (triangles), and Ni (circles) catalysts.
of Cu supported on basic metal oxides such as MgO.31 The
Individual data points are taken from a number of prior literature
syngas initially forms alcohols that may be coupled through reports, with the citation number indication by the color for that data
condensation reactions. In the proposed mechanism, the point. Dashed lines are used to connect points for each metal, but note
alcohol dissociatively adsorbs on the basic oxide surface, that results from each metal come from differently prepared catalysts,
creating hydrogen and alkoxide species.31 Hydrogen is removed as reported in Table 1.
by migration to Cu sites, where they recombine and desorb as
H2, while the alkoxide species may be further dehydrogenated
and eventually take part in condensation reactions to form
longer chain compounds.31 Chain growth rates are low due to Across a range of conversions, Pt catalysts appear to have the
the steric hindrance that occurs as branches are added to the highest deoxygenation selectivity of these monometallic
initially formed linear alcohols, resulting in production of samples. The data used in this figure can also be seen in
isobutanol at high selectivity.31 Table 1. However, the performance of these monometallic
2.2. Importance of Bifunctionality for Biomass catalysts can be dramatically changed through incorporation of
Reforming. A common theme across the classes of materials a second metal, as will be discussed below.
5029 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

Table 1. Data Used in Figure 4 To Compare Selectivity to Deoxygenation of m-Cresol or Guaiacol over Various Supported
Metals
catalyst temperature (K) P (atm) PH2 (atm) wt % metal reactant conversion (%) % deoxygenation ref
Pt/Al2O3 533 1 0.5 0.5 m-cresol 9.3 79.2 33
Pt/Al2O3 533 1 0.5 1 m-cresol 18 82.6 33
Pt/Al2O3 533 1 0.5 1.7 m-cresol 38.3 80.5 33
Pt/SiO2 533 1 0.5 1.6 m-cresol 54.9 82.1 33
Pt/Al2O3 533 1 0.5 1.7 m-cresol 30 76.1 32
Pt/TiO2 573 1 1 1 m-cresol 17 97 66
Pt/ZrO2 573 1 1 1 m-cresol 12 88.5 66
Ni/SiO2 523 1 1 5 m-cresol 55.8 7.2 68
Ni/C 623 39.5 37.6 5 guaiacol 30.7 17.5 60
Ni/SiO2 573 1 1 5 m-cresol 16.2 14.2 34
Pd/C 573 1 0.4 5 m-cresol 10 78 61
Pd/SiO2 523 1 1 1 m-cresol 54.9 24.1 68
Pd/C 623 39.5 37.6 1 guaiacol 15.5 23.3 60
Fe/SiO2 573 1 1 5 m-cresol 8.8 60.2 34
Fe/SiO2 623 1 0.9 not reported guaiacol 50 37 76
Ru/SiO2 573 1 1 9.4 m-cresol 4.5 38.5 67
Ru/C 623 39.5 37.6 1 guaiacol 34.2 39.6 60

3.1.1. Supported Noble Metals. One of the most widely reactions. This suggests that deoxygenation for reforming
used types of catalysts for upgrading of biomass-derived pyrolysis oil may be best accomplished using a bifunctional
oxygenates are supported metals. To understand reactivity catalyst that combines hydrogenation sites with sites that form
trends for these systems, it is often useful to study reactivity on strong metal−oxygen bonds.
model surfaces such as single crystals. Reactions of simple This hypothesis is also supported by studies of technical
organic oxygenates that can serve as probe molecules for catalysts composed of combinations of metal and acid sites.
components in deconstructed biomass have been extensively Late-transition-metal catalysts have several properties that make
studied using surface spectroscopies and model surfaces. There them potentially attractive for pyrolysis oil upgrading; they are
have been past reviews focusing on the decomposition of typically very active, are effective at hydrogenation, and have
simple oxygenates including alcohols, aldehydes, ketones, and well-characterized structures. Additionally, they do not suffer
carboxylic acids over transition-metal surfaces.48−51 For from some of the issues of traditional hydrotreating catalysts
example, formic acid decomposition has been studied as a discussed below, such as the need for cofeeding toxic H2S to
probe for organic acids. A past review of surface science studies maintain activity as for sulfides.56 When an active support
over single crystals focused on formic acid adsorption and material is used to impart bifunctionality, these catalysts can be
decomposition processes.48 Formic acid tends to decompose used for HDO of relatively recalcitrant compounds such as
through a formate intermediate. The formation of a carboxylate phenolics. Before discussing the bifunctional materials, it is
intermediate has been observed for larger carboxylic acids as useful to first discuss the individual components.
well, over a wide array of transition metals.51 Formate Noble metals including Ru, Rh, Pd, and Pt are a class of
decomposition can then occur either through dehydrogenation promising upgrading catalysts; they activate hydrogen, are not
to produce CO2 and H2, or C−O scission to form CO and deactivated by water, and are already widely used in commercial
H2O.48 The pathway that a given transition metal will favor catalytic processes.57 Under typical upgrading conditions, the
depends on how strongly it binds oxygen: metals like Pt that differences in the chemistry catalyzed by the various noble
form relatively weak metal−oxygen bonds favor dehydrogen- metals are subtle,8 and these materials have been highlighted in
ation52 while metals further left on the periodic table such as Fe a number of reviews discussing pyrolysis oil upgrading.1,7,8,39,58
favor C−O scission.53 On Pd(111) both pathways were As an illustration of their HDO activity as monofunctional
observed, although selectivity was greater toward dehydrogen- catalysts, Pd/C, Pt/C, and Ru/C were studied for vapor-phase
ation.54 On Ni(111) the binding configuration was proposed to guaiacol (2-methoxyphenol) HDO.59 While the hydrogenation
be different, creating formic anhydride as an intermediate, but activity was high, the paper concluded that carbon-supported
the dominate decomposition pathway was C−O scission.55 A precious metals are poor HDO catalysts because they favor ring
major conclusion from these studies is that C−O bond scission saturation. At 250 °C, the C−O bond between the ring and
is often not favored on metals like Pt and Pd that do not methoxy group was broken, but C−O scission of the strong
strongly bind oxygen, even when the C−O bond is relatively aromatic-hydroxyl bond did not readily occur, producing
weak. Similar observations have been made for reactions of phenol as the major product.59 Additionally, ring hydro-
other oxygenates such as alcohols and aldehydes. On noble genation to form cyclohexanone and cyclohexanol occurred,
metals such as Pt and Pd, decarbonylation via C−C scission is and a small amount of benzene was formed as well.59 Even at
often a dominant pathway, and C−O scission is generally not harsher reaction conditions (350 °C and 39.5 atm), noble-
observed.49 An exception is the reaction of unsaturated alcohols metal catalysts show poor selectivity to complete deoxygena-
and aldehydes such as prenal, furfural, and benzyl alcohol, tion.60
where the C−O bond is much weaker than in aliphatic or Supports with acid/base activity have been widely used to
phenolic reactants. However, noble metals are attractive for achieve better deoxygenation yields from phenolic reactants.
their ability to activate hydrogen and to conduct hydrogenation The aromatic-O bonds of phenolic compounds are weakened
5030 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

following ring hydrogenation, so it has been proposed that the form toluene.32 This hypothesis was further supported by
deoxygenation mechanism over noble metals involves hydro- reacting the fully hydrogenated alcohol intermediate (3-
genation followed by dehydration rather than direct C−O methylcyclohexanol) on a bare Al2O3 support, where it was
scission.32−34 In this model, one site is proposed to activate C− deoxygenated to form methylcyclohexene isomers.32 The rate
O bonds. This component is often proposed to be a Brønsted was much higher when the saturated intermediate 3-
acid site on the oxide support materials. In addition, sites are methylcyclohexanol was fed compared to the rate when m-
needed to supply atomic hydrogen to the reaction; this role is cresol or an alkene was fed. This result suggested that at 533 K
played by the noble metal.39 Proposed mechanisms from model ring hydrogenation over Pt sites to form the alcohol
compound studies on noble-metal catalysts consistently invoke intermediates was rate limiting. The proposed rate-limiting
hydrogenation followed by dehydration, after which further step of ring hydrogenation was used to explain the higher
hydrogenation steps can occur.8,32,33 This mechanism is shown selectivity of Pt/Al2O3 toward toluene than methylcyclohexane,
in Figure 5. because fewer steps are required to dehydrate the unsaturated
alcohol intermediates than to produce saturated alkanes. Note
that this mechanism requires intimate contact between the
hydrogenation and deoxygenation sites to prevent over-
hydrogenation of the relatively unstable intermediate. Also,
despite bare alumina easily catalyzing dehydration of the
saturated alcohol, no reaction was observed when feeding m-
cresol; at least partial hydrogenation was required to allow
cleavage of the C−O bond.33 To improve the HDO activity, Ni
and Co were added to form bimetallic catalysts. These 3d
metals have previously been shown to increase hydrogenation
activity of unsaturated compounds.62 By increasing the
hydrogenation activity through these bimetallic catalysts the
HDO activity increased, and the product distribution shifted to
produce more methylcyclohexane relative to the amount of
toluene formed. Total deoxygenation also increased over the
bimetallic catalysts, proposed to be a result of generating
aluminate species through interaction between the 3d metal
and the support, which are active for alcohol dehydration.32
Pt/SiO2 was also studied to compare the alumina-supported
material to a less acidic catalyst.33 Although both contain acid
sites, the Lewis and Brønsted acidity on Al2O363,64 is stronger
than that of SiO2.65 The less acidic Pt/SiO2 was found to be
ineffective at dehydrating 3-methylcyclohexanol to the alkane
product. A kinetic study of Pt/SiO2 also concluded that this
Figure 5. Schematic representation of metal-acid catalysis proposed dehydration pathway is limited.66 The dominant deoxygenation
for aromatic deoxygenation on noble-metal-based catalysts.
pathway was instead proposed to be dehydration of unsaturated
alcohol intermediates to form toluene. While SiO2 alone was
The deoxygenated product in Figure 5 is shown converting not active for dehydration of 3-methylcyclohexanol, in the
between the aliphatic and aromatic compound to suggest that, presence of Pt, its weakly acidic hydroxyl groups can dehydrate
depending on the catalyst and the reaction conditions, either partially hydrogenated alcohol intermediates present on the Pt
could be possible. Aromatic oxygenates may be fully saturated particles, as shown in Figure 5.33 Both the Pt and SiO2 sites are
over the metal particles, particularly if the reaction conditions necessary for this reaction. When the support was doped with
favor this (e.g., at lower temperatures).32,33 Bimetallic particles potassium to suppress the acidity, the product distribution was
may be used to eliminate the need for an acidic support by substantially altered. Rather than primarily toluene, m-cresol
providing oxophilic sites that catalyze dehydration. For formed 3-methylcyclohexanol with 84% selectivity.33 This is in
example, Fe can be used to promote Pd deoxygenation activity agreement with density functional theory (DFT) calculations
either by depositing Pd on an Fe2O3 support or by creating over Pt(111), which show that m-cresol HDO is less
PdFe particles on a carbon support.59,61 In other systems, an energetically favorable than ring hydrogenation.67 However,
acidic support such as alumina may still be needed for increasing the reaction temperature may increase the
dehydration activity. For example, PtCo and PtNi bimetallic favorability of the HDO pathway.68
particles were prepared to increase hydrogenation activity and Another study of support effects on noble metals also
therefore increase the rate of deoxygenation through sequential concluded that acidic supports facilitate phenolic HDO. Pt, Rh,
hydrogenation followed by dehydration on the support.32 Pd, and Ru supported on Al2O3, SiO2−Al2O3, and nitric-acid-
One report studying the vapor-phase HDO of m-cresol (3- treated carbon black (NAC) were studied for the liquid-phase
methylphenol) over Pt/Al2O3 found that a bifunctional batch upgrading of guaiacol.35 The support was found to have a
mechanism was required, involving hydrogenation activity to greater influence on the product distribution than the noble
form alcohol intermediates as well as acidity for dehydration.32 metal. Using ammonia TPD, the acidity of the catalysts was
It was proposed that ring hydrogenation first occurs to saturate found to decrease in the order SiO2-Al2O3 > Al2O3 > NAC.35
the phenyl ring and a partially or fully hydrogenated alcohol The same trend was found in production of cyclohexane, the
intermediate is formed. On Pt/Al2O3 the partially hydrogenated fully deoxygenated and hydrogenated product. The least acidic
alcohols may then deoxygenate on acidic sites on the support to catalysts, those supported on NAC, had the highest yields of 2-
5031 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

methoxycyclohexanol, a result of ring hydrogenation without This illustrates the importance of better understanding the
any C−O scission and an indication of low HDO activity.35 metal−support interface, especially for reducible metal oxides
Therefore, HDO was again proposed to need a bifunctional like TiO2 that may act as either an acid or a base. Similar results
catalyst where acid sites are required to produce a have been observed for Pt/TiO2. The combination of a metal
deoxygenated product rather than stop at a ring-saturated and a reducible oxide support were shown to enhance m-cresol
oxygenate. deoxygenation by introducing additional energetically favorable
In addition to HDO through hydrogenation reactions reaction pathways.71 While Pt/C preferred ring hydrogenation,
followed by dehydration, direct deoxygenation (DDO) has Pt/TiO2 catalyzed tautomerization and DDO. This synergistic
been observed over supported metal catalysts. As shown in effect was particularly important at elevated temperatures (623
Figure 1, DDO does not require intermediate hydrogenation K).71
steps; instead, the C−O bond is directly cleaved from the There has also been some interest in using zeolites as the
aromatic reactant. This may be desirable for reducing the solid acid component of these bifunctional systems. While their
hydrogen demand of the deoxygenation process. One example low activity toward phenolic reforming and high rates of coke
of DDO is using noble metals supported on reducible oxides generation make zeolites potentially less attractive than the
such as TiO2. For instance, Ru/TiO2 has been shown to other upgrading catalysts highlighted here, these issues can be
catalyze DDO of phenol to benzene through a bifunctional addressed by adding metals to the zeolite. Using guaiacol as a
mechanism at the interface between the metal and oxide.69,70 model compound, only transalkylation was found to occur over
Using both experimental and computational techniques, it was HY zeolite, resulting in no deoxygenated products.72 Similar
demonstrated that at 573 K and 3.8 MPa, water that is present results were observed for Pt/Al2O3 in the absence of hydrogen.
on the TiO2 may either accept or donate protons across the However, phenol could be hydrogenated and deoxygenated
metal−metal oxide interface. This allows the phenolic OH over several zeolites that had been doped with Pt.73 Pt/HY has
group to be directly removed through donation of a proton, also been used for HDO of phenolic compounds, with
shown in Figure 6. Although this is typically difficult to achieve
deoxygenation as well as coupling reactions forming larger
molecules.74
In addition to these purely heterogeneous examples,
supported metal catalysts have also been combined with liquid
acids to achieve high deoxygenation selectivity. For example, an
aqueous-phase phenolic mixture of pyrolysis oil model
compounds was upgraded over carbon-supported noble-metal
catalysts and a mineral acid.38 In neutral solution, the noble-
metal catalysts favored hydrogenation of phenol to cyclo-
hexanol. However, in the presence of phosphoric acid, the
alcohol was easily dehydrated. The proposed pathway for this
process is initial metal-catalyzed hydrogenation of the ring
followed by acid-catalyzed dehydration to form cyclohexene
and subsequent metal-catalyzed hydrogenation to ultimately
produce cyclohexane.38 This combination of a carbon-
supported noble metal with a mineral acid was found to be
effective for selective deoxygenation of a variety of phenolic
compounds in the aqueous phase, performing multistep
reactions involving hydrogenation, hydrolysis, and dehydration
Figure 6. Proposed mechanism of DDO at the bifunctional Ru/TiO2 steps. For instance, in neutral solution, all catalysts produced no
interface. (a) H2 is activated on Ru particles and heterolytically
dissociates across the interface to protonate a bridging hydroxyl group deoxygenated products at 100% conversion; however, upon
on TiO2, generating a Brønsted acid site. (b) This then assists in addition of phosphoric acid to provide bifunctionality to the
phenol deoxygenation. (c) Side view of DFT calculation of water- catalytic mixture in the form of acid sites, the deoxygenated
assisted phenol DDO on Ru/TiO2, including initial state (left panel), alkane product was formed with about 90% selectivity over all
transition state (center panel), and final state (right panel). Color metals.38
code: hydrogen, white; oxygen, red; carbon, dark gray; ruthenium, teal; 3.1.2. Supported Base Metals. While supported noble-metal
titanium, light gray. Reproduced with permission from ref 69. catalysts are promising for biomass HDO, their high cost has
Copyright 2015 American Chemical Society. sparked interest in studying less-expensive base metals such as
Ni, Co, and Fe. As with the noble-metal catalysts, it is useful to
due to the strong aromatic−oxygen bond, the bifunctional sites first investigate the monofunctional behavior of these metals.
present at the interface containing both metal and acid Fe/SiO2 was used for the gas-phase upgrading of guaiacol as a
functionalities were calculated to have a lower energy barrier model compound to study iron’s ability to break hydroxyl and
than hydrogenation. On large Ru particles, where the interfacial methoxyl bonds on an aromatic ring.75 The catalyst was found
area is much smaller, ring hydrogenation was found to be most to be selective for production of aromatic hydrocarbons over a
favorable as expected from previous studies of noble metals. range of temperatures.75 Kinetic studies of Fe/SiO2 for guaiacol
Isotopic labeling experiments in combination with DFT upgrading have also been performed.76 Results from Fe
calculations indicated that the Ru particles bind the phenol catalysts were compared to the same reaction over a Co
and activate hydrogen to catalyze the hydride attack (which catalyst. Cobalt was found to be more active than Fe but was
regenerates the water cocatalyst on the support), whereas shown to catalyze decomposition of the desirable products.76 A
Brønsted acidic water on TiO2 catalyzes C−O scission. simplified pathway is shown in Figure 7. Transalkylation
5032 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

reactions were proposed to occur over the acid sites, explaining


the formation of toluene from guaiacol.

Figure 7. Simplified pathway of guaiacol HDO on iron and cobalt


catalysts. Both iron and cobalt catalysts follow black arrows; only
cobalt catalyzes additional reactions, as shown in blue. Adapted from
ref 76.
Figure 8. Proposed reaction network for guaiacol HDO on precious
metals (Pd, Pt, Ru, pathways shown with black arrows) and base
A study comparing carbon-supported Fe and Cu to noble metals (Cu, Fe, pathways shown with red, dashed arrows). Adapted
metals obtained similar results for guaiacol HDO.59 Guaiacol from ref 59.
HDO was found to proceed through a phenol intermediate
over all catalysts, but unlike the noble metals, these base metals
did not lead to ring hydrogenation and C−C scission products.
For example, Fe/C was able to convert some of the phenol to
aromatic deoxygenated products such as benzene and
toluene.59 However, the activity of the base-metal catalysts
was lower than the noble metals. On Pd, Pt, and Ru the
intermediate phenol was primarily hydrogenated to form
cyclohexanone and cyclohexanol as expected in the absence
of an acid site for deoxygenation. In addition, ring opening to
form light gases also occurred. Meanwhile, Fe and Cu were not
active for ring saturation and ring opening reactions,
predominately forming benzene.59 This proposed network
can be seen in Figure 8.
These base metals show even more promising HDO activity
when used as part of a bifunctional metal-acid catalytic system.
For example, Raney Ni and Nafion on SiO2 were used for the
aqueous-phase HDO of a series of phenolic compounds.36 The
Ni was found to act as a hydrogenation catalyst, and the
Nafion/SiO2 provided acid sites for hydrolysis and dehydration,
resulting in high selectivity to hydrocarbon production.36 The Figure 9. Proposed reaction pathways of phenolic compounds over
proposed reaction pathway for phenolics over this catalyst Raney Ni and Nafion/SiO2. Adapted from ref 36.
system is depicted in Figure 9.
With Nafion/SiO2 still present in the aqueous mixture to hydrogen at the metal−support interface. The TPD trace for
provide acid sites, Ni/SiO2 has also shown activity for the smallest Ni loading had a small shoulder at this temperature
deoxygenation of phenol with the properly selected reaction along with a larger peak and additional high-temperature
temperature.77 Temperatures below 523 K did not catalyze shoulders at 720 and 760 K.77 The peak at 720 K was attributed
hydrogenolysis to form benzene; this was attributed to to hydrogen that is associated with the surface Ni sites, which
thermodynamic limitations. Instead, cyclohexanone and cyclo- has also been proposed to be responsible for hydrogenation
hexanol were produced. Higher Ni loading was shown to selectivity, while the peak at 760 K was attributed to hydrogen
increase activity as well as selectivity to benzene production, that had spilled over onto the support. Because the interfacial
while lower loadings of Ni produced more cyclohexanol and area was much higher for the catalysts with high Ni loading, it
cyclohexanone.77 This was proposed to be due to a greater was suggested that increases in the Ni loading increased the
population of hydrogen on the surface that can be used for availability of weakly bound hydrogen accessible for hydro-
hydrogenolysis.77 It was further proposed through H2 TPD genolysis to benzene.
results that on the larger particles the majority of the hydrogen A different bifunctional approach to deoxygenation using Ni
available is at metal−support interfacial sites, which are most is to utilize MgO as a basic support. Liquid-phase guaiacol
active for hydrogenolysis. Hydrogen TPD showed a sharp peak upgrading was studied over Ni/MgO at 433 K and 29.6 atm
at 663 K on large particles, which was proposed to come from H2.78 At 98% conversion, there was 100% selectivity to
5033 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

demethoxylation to produce cyclohexanol. This high selectivity metal−oxygen bond strengths for some commonly used
was attributed to the acid−base interaction between the oxophilic modifiers.
phenolic OH group of guaiacol and the MgO support, with
Ni particles providing hydrogenation sites.78 Guaiacol deme- Table 2. Metal−Oxygen Bond Strengths for Oxophilic
thoxylation has also been observed over Ru/MgO79 and Ru- Metals Used in Bimetallic Catalystsa
MnOx/C, where in the latter case, the MnOx provides
basicity.80 Selective demethoxylation and hydrogenation may metal−oxygen bond bond strength (kJ mol−1) ref
be desirable as a renewable method of producing cyclohexanol oxophilic metals
as a feedstock for polymer production or as an intermediate for Co−O 397 90
fuel production. Cr−O 461 90
Although the previous literature convincingly demonstrates Fe−O 407 90
that a combination of acid and metal sites leads to more W−O 653 91
complete deoxygenation, it is important to point out that Mo−O 607 91
metals themselves have some activity toward HDO reactions. Re−Ob 644 92
For reactants with relatively weak C−O bonds such an allyl hydrogenating metals
alcohol, furfuryl alcohol, and benzyl alcohol, deoxygenation can Ni−O 366 90
even be observed on single crystals in ultrahigh vacuum.81−83 Pd−O 234 91
Pt−O 347 91
Deoxygenation of these compounds on supported metals has
Rh−O 377 91
often not been interpreted in terms of the support having a a
role. Studies of such reactants on supported Pd catalysts have Values represent the bond dissociation energies measured at 298 K.
shown that selective poisoning of different types of Pd sites can Both modifier sites used as the oxophilic site and hydrogenating metals
are included for comparison. bSpecifically for the Re−O bond in ReO3
be used to favor selectivity for deoxygenation.84,85 However, as
the studies investigating support effects show, the interaction
between the metal particle and the support can significantly The hydrogenating sites provide H while the oxophilic metal
impact the surface chemistry. Going forward, a major challenge more strongly binds the reactant to the surface through its
will be to develop better models for the structure of metal− oxygen functionality, allowing easier C−O scission and
support interface and to learn how to control interfacial active potentially creating new reaction pathways unavailable on a
sites. This is discussed in greater detail in the final section. monometallic catalyst. The majority of bimetallic catalysts
3.2. Bimetallics. The previous examples of transition-metal- investigated have involved 3d metals as a modifier to noble
catalyzed HDO of phenolic compounds show that these metals. These systems are known to create subsurface alloys
catalysts may be promising materials for pyrolysis oil upgrading with the surface noble-metal atoms significantly electronically
if acid sites are present. This mechanism (as shown in Figure 5) modified.93,94 Subsurface transition-metal alloys have been
may also take place over bimetallic catalysts. Oxide supports shown to increase HDO activity as well as selectivity toward
such as alumina are often used as the source of acid sites, but saturated hydrocarbons.32,95−97 DFT results show that the Pt-
they also can strongly bind oxygenates like phenols, leading to terminated alloys bind adsorbates less strongly than pure Pt,
intermediates that block sites and can be precursors to coke resulting in a more optimized adsorption strength which
formation.33,86,87 Phenols and polyphenols have been shown to increases hydrogenation.32 An investigation of the hydro-
be coke precursors over a wide range of catalysts, including genation of the unsaturated compounds benzene and 1,3-
supported metals, oxides, and zeolites.88 This has prompted butadiene over Pt and PtNi catalysts also observed that
hydrogenation rates increased with the Ni loading.97 DFT
studies of the effects of bimetallic catalysts containing a
calculations have suggested this could be due to shifting the d-
hydrogenating metal and an oxophilic metal (an element that
band center of surface Pt atoms by subsurface Ni, resulting in
forms strong metal−oxygen bonds) to provide intimate contact
weaker binding of the unsaturated compounds.98 The hydro-
between diverse sites for pyrolysis oil upgrading.89 Many of
genation activity was shown to have a volcano-curve-type
these have been noble-metal-based alloys. Figure 10 shows one relationship with the binding energy of the reactant.98
proposed mechanism for the interaction of these bimetallic Therefore, tuning of this binding strength allows maximization
catalysts with an organic oxygenate. The metals used as a of the hydrogenation activity. In addition to influencing
modifier site may be present in their metallic state or as an hydrogenation rates, it was proposed that the bimetallic
oxide; regardless, their role is to introduce an additional active catalysts contained additional sites for dehydration. Metals
site to produce a bifunctional nanoparticle. Table 2 shows such as Co and Ni can create additional acid sites through
interaction with the alumina support to increase the rates of
dehydration reactions.32 This is proposed to be due to
formation of an aluminate species, which previous work has
shown increases the acidity on alumina.99 Because acid-
catalyzed dehydration steps are critical for C−O scission in
phenolic compounds, this increases HDO rates.
Many previous studies with bimetallic catalysts have been
performed using probe molecules such as glycerol. While the
precise role of the modifier is unknown, many studies have
shown increased activity and selectivity for C−O bond scission
Figure 10. Hypothesized mechanism for interaction of oxygenates reactions for a variety of probe molecules and metal
with bimetallic surfaces containing a metal capable of dissociating combinations.37,100−104 PtRe/C catalysts were shown to be
hydrogen and a metal that forms strong M−O bonds. more active than monometallic Pt for converting glycerol to
5034 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

syngas.37 Here, the role of the oxophilic metal may not be EXAFS and STEM/EDS results, was used to propose that Fe is
directly related to its affinity for oxygen. Rather, CO-TPD the active site for HDO and the role of Pd is facilitating
showed that CO binding energies were reduced on PtRe reduction of FeOx as well as modifying the Fe sites through
compared to either monometallic sample, suggesting that electronic effects due to alloy formation.59 STEM images of the
reduced site blocking could be at least partially responsible for morphologies of these materials can be seen in Figure 11.
the increased activity.37 Another glycerol reforming study in
which PtRe/C was used for hydrogenolysis to form propane-
diols reported both increased rate and selectivity toward
deoxygenation of the secondary alcohol.37 Similar results were
obtained for RhRe/SiO2 catalysts.100 In trying to understand
the nature of these modifier sites, two main hypotheses have
been proposed. First, it was suggested that hydroxylated Re−
OH groups may be responsible for C−O activation by acting as
Brønsted acid sites, with the strong metal−oxygen bond
weakening the hydroxyl’s oxygen−hydrogen bond, thus
allowing the hydrogen to be more easily donated to a
neighboring oxygenate.37 NH3 TPD has shown increased Figure 11. Representative aberration-corrected high-angle annular
acidity on PtRe catalysts in the presence of water relative to dark field scanning transmission electron microscopy (AC-HAADF-
what is observed over the monometallic Pt sample.105 The STEM) images of reduced and (left) and spent (right) Pd/Fe2O3
second hypothesis is that these oxophilic sites are strongly studied for m-cresol HDO. Brighter spots show Pd atoms, and darker
anchoring the reactant to the catalyst surface through its oxygen spots show Fe atoms. Reproduced with permission from ref 59.
functionality, allowing surrounding sites to donate hydrogen for Copyright 2013 Elsevier.
deoxygenation.100 This strong interaction of the oxophilic site
and the oxygenated reactant can result in altered reaction
pathways. For example, m-cresol was shown to preferentially Bright regions with a diameter of 1−2 nm are Pd clusters on
follow the ring hydrogenation route to deoxygenation over pure the Fe surface; the darker spots underneath these are Fe
group VIII metals such as Ni, Pd, and Pt (shown by the atoms.59 Both reduced and spent catalysts were imaged; the
saturated compounds in Figure 5).34,71 However, when an comparison between these was used to suggest that there is
oxophilic modifier was added to the catalyst a tautomerization negligible change in morphology during reaction conditions,
pathway became accessible, creating a keto intermediate which and therefore, the Pd−Fe structure is robust.59
may then be dehydrated to form toluene. By making this Similar results were obtained from a Pd/Fe2O3 catalyst for
additional deoxygenation pathway energetically favorable the HDO of m-cresol: Pd was proposed to activate hydrogen and
deoxygenation activity is increased. This has been shown for m- allow reduction of the Fe2O3 as well as prevent oxidation by
cresol HDO over PtMo/Al2O3, as illustrated in Figure 2c.12 water formed during the HDO reaction.61 Addition of Pd was
Although more work is needed to fully understand the role of also found to increase the activity compared to monometallic
oxophilic modifiers, model compound studies show their Fe while maintaining a similar selectivity toward the production
promise for biomass deoxygenation. In addition to improving of aromatic deoxygenated products.61
deoxygenation selectivity, some studies using bimetallic Cyclic oxygenates have also been studied for selective
catalysts also reported DDO mechanisms. This is advantageous hydrogenolysis on bimetallic catalysts. A common conclusion
if the goal is to remove oxygen with minimal hydrogen among many of these reports is that the combination of a noble
consumption. Increased selectivity to C−O scission products metal and an oxophilic metal such as Re, Mo, or W can
has been observed over noble-metal catalysts modified with dramatically increase selectivity to cleaving more sterically
oxophilic metals such as Fe, Re, Mo, and W; the modifier sites hindered C−O bonds. For example, Rh catalysts were modified
could potentially bind oxygenates more strongly to the surface with Re to improve selectivity to 1,6-hexanediol production
or create acid sites when they are hydroxylated by water present from tetrahydropyran-2-methanol (THPM).108,109 While Rh/C
under typical reaction conditions.37,58,101,102,104,106,107 had only 15% selectivity to this desired product, Rh-ReOx/C
In one study, PdFe alloys were used for the vapor-phase had 96% selectivity as well as increasing the reaction rate from 2
upgrading of guaiacol.59 PdFe/C showed better selectivity to 180 mmol h−1 g catalyst−1. This improved activity was
toward production of deoxygenated aromatics than either of the proposed to be due to strong Re−OH interactions that allowed
monometallic catalysts, and no ring-saturation or ring-opening an alkoxide species to form, strongly tethering the THPM to
products were observed. This study led to the hypothesis that the surface. Hydrogenolysis of the C−O bond adjacent to the
phenolics prefer to adsorb on the Pd-modified Fe sites. alkoxide group is favored and therefore results in high
Compared to the monometallic Fe surface, these sites selectivity to 1,6-hexanediol.108 A similar mechanism was
hypothetically have altered electronic properties and are in a proposed using Rh-ReOx/SiO2 for tetrahydrofurfuryl alcohol110
more reduced state due to increased hydrogen activation by Pd. and glycerol100,111 hydrogenolysis. Another study using Rh-
This results in a catalyst that can activate phenolic C−O bonds ReOx/C for hydrogenolysis of cyclic ethers proposed that this
without first needing to saturate the ring. This mechanism was increase in selectivity to C−O bond scission at the more
explored through DFT calculations by substituting Pd atoms sterically hindered carbon may be due to the ability of OH
into a Fe(110) surface. Phenol was adsorbed with varying groups on ReOx sites to stabilize the intermediate carbenium
distances from the Pd atoms. The results indicated that ions.112 To illustrate this mechanism, Figure 12 shows the
adsorption and C−O cleavage are both more favorable on the proposed reactant and intermediate adsorption states for
Fe sites.59 TPR results on the PdFe/C catalysts showed that Pd tetrahydrofurfuryl alcohol hydrogenolysis on Rh(111) modified
assists in the reduction of FeOx.59 This evidence, along with with an Re−OH site.
5035 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

been studied for m-cresol HDO.34 Over monometallic Ni,


methylcyclohexanone was found to be the major product, while
on the bimetallic catalyst, toluene was dominant. A physical
mixture of Ni/SiO2 and Fe/SiO2 did not show this change in
the reaction pathway, indicating that interactions between the
two metals are important. It was proposed that alloying with
oxophilic metals such as Fe weakened the interaction between
the aromatic ring and the surface, favoring strong binding with
the carbonyl group and leading to production of unsaturated
compounds.34 Studies like these have recently generated a great
deal of interest in better understanding bimetallic catalysts
Figure 12. DFT-calculated reactant and transition-state adsorption containing less-expensive base metals for deoxygenation
structures for tetrahydrofurfuryl alcohol hydrogenolysis. The mecha- applications.
nism is proposed to be a concerted protonation, hydride transfer, and A summary of these studies and the improvement of
ring opening over Re−OH modified Rh(111). Reproduced with deoxygenation over the monometallic catalyst can be seen in
permission from ref 112. Copyright 2011 American Chemical Society. Table 3. For both noble-metal and base-metal-containing
bimetallic catalysts, one of the main challenges is understanding
Increased hydrogenolysis selectivity has been observed for and controlling the active sites at the interface of the two
acyclic oxygenates as well.112 Ir-ReOx/SiO2 showed higher metals. Continuing to work toward a better, atomic-scale
selectivity to 1,4-butanediol production from erythritol hydro- picture of how the two metals interact under reaction
genolysis than unmodified Ir.113 This increase in selectivity conditions will be a key challenge to overcome going forward.
from 0% to 33% was again proposed to be due to formation of 3.3. Oxides. The discussion above mainly considers metal
an alkoxide species on ReOx sites followed by hydrogenolysis oxides that serve as a carrier for supported metal nanoparticles,
by a hydride species on a neighboring Ir site. Similar results and which require the presence of the metal to effectively
were found for glycerol hydrogenolysis on Ir-ReOx/SiO2, with catalyze deoxygenation. However, some metal oxide surfaces
increased C−O scission attributed to formation of an contain diverse types of active sites so that they themselves may
alkoxide.114−116 Characterization work on Ir-ReOx surfaces
be considered bifunctional without the addition of other metals.
has supported the hypothesis that interaction between the two
sites is important in the enhanced hydrogenolysis; a In particular, reducible metal oxides have shown deoxygenation
combination of XAS, TPR, XPS, and CO chemisorption activity. For example, benzoic acid deoxygenation was studied
suggested that there is significant interaction between the Ir over ZnO and ZrO2.119 Hydrogen was used to create oxygen
sites and the low-valent ReOx.117 A similar investigation for Rh- vacancies in the oxide surface which could then be reoxidized
ReOx also found that the two sites were well mixed and by the benzoic acid reactant to form benzaldehyde.119
interacting.118 However, these vacancies were susceptible to poisoning by
Although bimetallics containing noble metals have been the compounds like CO2 and water. In addition to model
most extensively studied, another promising area is in the use of compound studies, oxides have been studied for catalytic
bimetallics formed from base metals. NiFe/SiO2 catalysts have upgrading of whole pyrolysis vapors. A ZrO2 and TiO2 mixed

Table 3. Summary of Bimetallic Catalysts Studied for Deoxygenationa


PH2 conversion % increase over
catalyst, wt % metals temperature (K), phase (atm) reactant (%) % deoxygenation monometallic ref
Pt−Co/Al2O3 533 0.5 m-cresol 57 97.1 21 32
1.3% Pt 4.3% Co vapor (1 atm pressure)
Pt−Ni/Al2O3 533 0.5 m-cresol 63 94.2 18.1 32
1.5% Pt 4.9% Ni vapor (1 atm pressure)
Pt−Re/C 443 39.5 glycerol 20 26b 26 37
5.7% Pt 4.6% Re aqueous
Rh-ReOx/SiO2 393 39.5 glycerol 10 22b 17 100
4% Rh, 0.5 molar ratio Rh/ReOx aqueous
Pt−Mo/Al2O3 523 4.6 m-cresol 95 96 76 12
5.1% Rh 1.8% ReOx vapor (5 atm pressure)
Pd−Fe/C 623 0.4 guaiacol 90 29 26 59
2% Pd 10% Fe vapor (1 atm pressure)
Pd−Fe/C 723 0.4 guaiacol 95 85 80 59
2% Pd 10% Fe vapor (1 atm pressure)
Rh-ReOx/C 373 39.5 THPM 36 97 58 108
4% Rh, 0.5 molar ratio Rh/ReOx aqueous
Rh-ReOx/SiO2 393 39.5 THFA 57 94c 76 110
4% Rh, 0.5 molar ratio Rh/ReOx aqueous

a
Reactants include m-cresol, glycerol, guaiacol, tetrahydropyran-2- methanol (THPM), and tetrahydrofurfuryl alcohol (THFA). bSelectivity to 1,3-
propanediol. cSelectivity to 1,5-pentanediol.

5036 DOI: 10.1021/acscatal.6b00923


ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

oxide catalyst increased the detected percent of hydrocarbons occur; metal oxides that are not reducible and therefore lack the
to 13.1%, compared to 0.1% in the raw vapor.120 metal cations for binding the oxygen functional group do not
One of the most promising uses for oxide catalysts in show the same reactivity. These bifunctional surfaces are
biomass upgrading is for ketonization. The carboxylic acids effective because they provide a surface containing Lewis acid
present in pyrolysis oil are especially important to remove due sites for binding the intermediate in close proximity to Lewis
to their corrosivity and reactivity. One approach to eliminating basic sites for deprotonation. More information on ketonization
these functional groups while also increasing carbon chain catalysts and mechanisms can be found in past reviews.51,123
length is ketonization, or ketonic decarboxylation, which 3.4. Sulfides. The most commonly used industrial hydro-
converts two carboxylic acid molecules into a ketone with treating catalysts are sulfided molybdenum or tungsten
water and CO2 formed as byproducts.121,122 promoted by cobalt or nickel on an alumina support; these
R1COOH + R 2COOH → R1COR 2 + CO2 + H 2O materials are widely used in petroleum refining for hydro-
(1)
desulurization (HDS) and hydrodenitrogation (HDN).8 Due to
This is also beneficial because ketones can undergo aldol this large-scale production, these catalysts can already be made
condensation reactions to form larger molecules in the gasoline economically and are well-characterized and understood in the
to diesel range. The majority of studies for ketonization have context of HDS and HDN. The wide variety of past studies has
been done using oxide catalysts. Reducible metal oxides such as shown that the edge area of these materials is important for
CeO2, ZrO2, and TiO2 have been found to be more active and HDS activity, leading to the conclusion that edges are the active
selective than purely acidic or basic oxides.121 It has been sites for this reaction.124−127 However, the nature of the active
proposed that reducible oxides allow the carboxylic acids to sites for HDO is still under debate. It is typically proposed that
dissociatively adsorb to form carboxylates and hydroxyls on the the materials are bifunctional, with one site for activation of
surface. These oxides also have available cation sites that heteroatom-C bonds and another for hydrogenation.39,40
promote the adsorbed acids to interact and form the required Oxygen can adsorb to coordinatively unsaturated sites (sulfur
carboxylate intermediates.121 These two distinct types of active anion vacancies) at the catalyst edges.39,40 An example of this is
sites were proposed to create a bifunctional surface capable of shown schematically in Figure 14.
improving activity for ketonization. An example of this
mechanism is shown in Figure 13.

Figure 14. Proposed mechanism of deoxygenation over transition-


metal sulfides.

Modifiers including Ni and Co donate electrons to the metal,


weakening the metal−S bond and allowing more vacancies to
be created.39,128 Hydrogen is thought to come from S−H
groups on the surface (formed either through activation of
molecular hydrogen or from decomposition of molecules such
as water or alcohols).8 This hydrogen can be used for HDO by
breaking the C−O bond or for decarbonylation/decarbox-
ylation by using acid sites to remove end groups from
Figure 13. Ketonization mechanism proposed over reducible metal aldehydes or carboxylic acids, respectively128 (see Figure 1).
oxides. M = metal site. Adapted from ref 121. Reduced M+ sites act as The selectivity to a given pathway depends on reaction
Lewis acids to bind the carboxylic acid and oxygen anions provide conditions as well as the transition metal being used. For the
Lewis basicity to deprotonate the adsorbed acid. HDO pathway, unsaturated carbon bonds as well as carbonyl
groups are often first hydrogenated to form alcohols, which are
A combination of surface science and reactor studies has led then dehydrated. An example of this can be seen for phenolics
to proposing that the active site for these reactions contains and ketones in Figure 15.
coordinatively unsaturated metal cations that act as Lewis acid Recent work has focused on materials that more clearly fit
sites for binding of intermediates and oxygen anions, acting as the definition of bifunctional (i.e., incorporating additional
Lewis bases for deprotonation of the acid. The availability of metals to create sites that assist in HDO of compounds relevant
these sites as well as the redox properties of the material can be to pyrolysis oil reforming).129−133 Previous investigations have
influenced by adding dopants such as transition metals and shown that Co and Ni can promote the HDO pathway over
metal oxides.121,122 This can improve catalyst activity because hydrogenation or decarboxylation/decarbonylation for Mo2S
both types of active sites are important for ketonization to catalysts.130,131,134 Experimental and DFT work has suggested
5037 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

compared to a Pd/Al2O3 catalyst at the same conditions.138 A


proposed reaction scheme for guaiacol HDO on these catalysts
can be seen in Figure 16. The Pd catalyst had a much higher

Figure 15. Example mechanism proposed for the HDO pathway over
sulfides.

that incorporation of Ni−Mo sites lowers the activation barrier


for breaking aldehyde CO bonds and alcohol C−OH bonds
by altering the preferred adsorption orientation, enhancing
HDO over decarbonylation/decarboxylation or hydrogenation.
Instead of binding with just the oxygen group on top of the Mo
site, it was proposed that the bimetallic surface allows bidentate
binding where the carbon atom of the functional group (for
example, in a carbonyl) is in close proximity to an Ni−H site.
This lowers the energy of the transition state for HDO.
However, if too much Ni is incorporated, Ni3S2 is formed,
which increases selectivity to decarbonylation over HDO.131,134
These studies have also expanded the composition of
transition-metal sulfides being investigated, incorporating
multiple metals and modifiers, as well as studying support
effects. For example, on a W−Mo−S catalyst promoted by Ni,
tuning the W/Mo ratio was found to change the morphology of Figure 16. Proposed guaiacol HDO reaction pathways on transition-
the catalyst, influencing the number of coordinatively metal phosphides. Proposed intermediates shown in parentheses.
Adapted from ref 138.
unsaturated sites that have been previously proposed as the
active site for activating C−O bonds.129 This resulted in
increased activity as well as selectivity to deoxygenation. activity than the phosphides under the experimental conditions
3.5. Carbides, Nitrides, and Phosphides. Transition- but formed primarily catechol, an undesired product that also is
metal carbides, nitrides, and phosphides have also been a precursor to coke formation.138 Ni2P was the most active of
explored for HDO, showing promising hydrotreating abilities the phosphides tested and avoided this pathway, favoring
because they combine the stability of ceramics with the production of phenol, benzene, and methoxybenzene.138 The
electronic properties of a metal.135,136 These materials are well- decrease in catechol formation may be attributed to the P−OH
characterized HDS and HDN catalysts due to their use in groups on the phosphide that are proposed to add acidity to the
petroleum refining, but less attention has been given to their surface.41−43 The suppression of catechol production on
applications for HDO until recently. phosphides is promising; however, additional studies are
A current hypothesis for the mechanism of deoxygenation needed to improve selectivity to fully deoxygenated products.
over transition-metal phosphides relies on the presence of both The selectivity toward phenol and methoxybenzene seen in this
metallic and acidic groups.41−43 For example, it has been study are not ideal for a final upgrading catalyst.
proposed that for Ni2P, P−OH groups could contribute Carbides have been compared to platinum group metals,
Brønsted acid sites while Ni sites add metallic behavior.41−43 showing similar catalytic properties; reactions that would not
Nickel phosphide has received the most attention as an HDO occur on W or Mo alone are observed over the carbide surface,
catalyst due to its activity and selectivity.136 For example, because binding of the metal to carbon atoms shifts the energy
anisole HDO over supported Ni2P has been investigated; the of the transition-metal d electrons to lower energies.137,139−141
proposed reaction scheme involves demethylation to form The ability of tungsten and molybdenum carbides to remove
phenol followed by hydrogenolysis and hydrogenation.42 The heteroatoms suggests they could be promising catalysts for
active sites on this catalyst are proposed to be Ni, modified by biomass deoxygenation. For example, through DFT calculations
its P neighbors.8 Ni sites with 4 neighboring P atoms, or Ni(1) as well as surface science and gas-phase reactor studies using
sites, have a tetrahedral geometry and are active for propanol and propanal as reactants, WC was shown to activate
decarbonylation; Ni(2) sites have 5 P neighbors in a square C−O bonds but not C−C; the dominant product formed was
pyramidal geometry and catalyze HDO.8,136,137 Bulk Ni2P propene.139 The catalysts have again been suggested to be
contains equal amounts of Ni(1) and Ni(2), but the surface is bifunctional in nature. The presence of oxygen on the WC
enriched with Ni(2), and this enrichment increases as the surface was proposed to introduce acidic sites, on which
particle size decreases. This has been confirmed experimentally dehydration reactions are faster than when oxidized W or WC
using a range of particle sizes and measuring both coordination is not present.139 A similar effect of surface oxygen has been
numbers as well as Ni/P ratios.136 Increasing the number of proposed for other carbide surfaces as well. For the HDO of
Ni(2) sites correlated with an increase in hydrotreating activity model compounds over NiMo carbide, it was proposed that
and a decrease in decarbonylation.136 metallic sites as well as metal−OH sites are generated under
A series of phosphides including Ni2P, Co2P, Fe2P, WP, and reaction conditions, leading to enhanced dehydration activity.44
MoP supported on SiO2 were studied for guaiacol HDO and Mo2C has also been studied for HDO and has shown
5038 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

promising selectivity similar to that of WC.142−144 In addition materials are important for deoxygenation is the necessity of
to these C3 oxygenates, studies have also been performed using both hydrogenation and acid/base sites. While many bifunc-
larger, aromatic oxygenates. For instance, a study of group 4, 5, tional combinations have demonstrated promising deoxygena-
and 6 metal carbides and nitrides investigated the HDO of tion activity, key questions about the mechanism remain. In
benzofuran at 643 K and 3.1 MPa.145 The carbides tested particular, there is a critical need to understand the role of the
(Mo2C, WC, NbC, and VC) showed poor selectivity to HDO interface between the different active sites in the reaction
(about 9, 10, 5, and 3% respectively). As a comparison, sulfided mechanism. In most studies to date, the density of interfacial
Ni−Mo/Al2O3, which is a common industrial hydrotreating sites has not been explicitly controlled, so it is difficult to
catalyst, showed over 60% selectivity to HDO. determine the role proximity between distinct types of sites
Nitrides have been studied extensively, particularly for plays in the reaction mechanism. It is important to resolve these
hydrogenation/dehydrogenation reactions. As discussed for questions because design of improved catalysts would ideally
carbides, the electronic effects induced by alloying can create a involve tailoring the interface between the two types of catalyst
material that has some catalytic properties similar to noble functionalities.
metals.146 They have been explored as hydrotreating catalysts, Certain materials lend themselves to studies in which the
primarily for HDS and HDN, but there has been increasing interface structure can be more systematically varied. Bimetallic
interest in using these materials for HDO.146,147 Nitrides are catalysts in which different metals at the surface play different
also thought to be bifunctional, with active sites being both catalytic roles (e.g., PdFe catalysts in which Pd serves as a
coordinatively unsaturated metal atoms as well as 4-fold hydrogenation center and Fe serves as a deoxygenation center)
nitrogen vacancies.148,149 Through studying guaiacol HDO, it hypothetically allow the density of one metal to be varied with
was proposed that demethoxylation and dehydroxylation occur respect to another and for parameters such as average ensemble
over the unsaturated Mo sites and demethylation is promoted size to be varied in a controlled manner. Surface-level
by nitrogen vacancies.148 When Mo2N was studied for guaiacol investigations with model surfaces have previously been
HDO, similar results were obtained for that of the phosphides employed to understand how specific arrangements of metal
reported above, avoiding the pathway to produce catechol but atoms lead to improved performance for other types of
producing mainly phenol rather than achieving complete reactions.152−155 Such combinations of metals are also relatively
deoxygenation.150 Mo2N also showed low selectivity to HDO straightforward to investigate using computational methods. In
in the study by Ramanathant and Oyama using benzofuran.145 all these investigations, care must be taken to ensure that the
All nitrides tested had low selectivity (<10%) to deoxygenation surface composition used in modeling the reaction is
with the exception of VN, which had about 50% HDO representative of that present under reaction conditions.
selectivity. Although this is lower than the control case of Thus, use of in situ and operando spectroscopies that are
sulfided Ni−Mo/Al2O3 (which had over 60% selectivity to surface-sensitive, such as near-ambient pressure X-ray photo-
HDO), VN was relatively selective to the aromatic product electron spectroscopy, will provide important information for
ethylbenzene, whereas the traditional hydrotreating catalyst tracing catalyst performance back to the true surface
favored ethylcyclohexane. It was proposed that the much more composition.156−158
promising activity of VN than any other nitride or carbide Metal−metal oxide interfaces are significantly more difficult
tested was due to its optimized binding strength of the reactant; to model but are clearly of interest for combining metal and
a volcano curve plotting HDO activity as a function of acid/base activity. There have been many recent advances in
benzofuran binding energy peaked near VN.145 the computational modeling of such interfaces that show the
While carbides, nitrides, and phosphides are effective importance of interfacial sites in facilitating catalysis.69,159,160
hydrotreating catalysts for HDS and HDN, overall they do Employing materials’ synthesis techniques that enable better
not seem to perform as well as other materials for HDO, such control over interfacial sites is one likely method for
as transition-metal-based catalysts. Most of the studies establishing structure−property relations in deoxygenation
summarized here report low activity and/or selectivity to catalysis. For example, techniques such as atomic layer
deoxygenation. However, they have not been thoroughly deposition could be employed to generate well-defined
investigated, and their low cost compared to noble-metal- coverages of different metal and metal oxide layers on both
based catalysts makes them attractive to study further. While model surfaces and high-surface area catalysts.161−167 Another
hypotheses exist regarding the nature of active sites for HDO opportunity for generating control over the interface (and
on these materials, a more fundamental understanding of their being able to employ reliable computational modeling
surface, particularly how the surface is influenced by the approaches) is to employ metal nanoparticles that are very
presence of oxygen, would aid in designing better catalysts. A well-defined in their structure,168−174 including the use of single
good comparison of a variety of carbide, nitride, and phosphide metal atoms as catalysts.175−178
catalysts for deoxygenation of various oxygenates can be found Finally, work using tethered organocatalysts suggests an
in a recent review.8 One promising area to explore in greater interesting path for designing bifunctional catalysts for biomass
depth is supported nanoparticles of these materials, which may conversions. In addition to inorganic catalysts, tethered organic
be expected to have different surface structures and facets acids and bases are a promising route for acid−base chemistry.
exposed than bulk phosphides, carbides, or nitrides. For A significant research effort has been devoted to studies of
example, nanoparticle Rh2P/SiO2 and Ni2P/SiO2 were studied these bifunctional (or “cooperative”) catalysts; one of the key
for guaiacol deoxygenation and showed similar levels of concepts in this area is controlling the spatial arrangement of
deoxygenation selectivity to transition-metal catalysts.151 the two functionsfor example, an acid and a baseto allow
simultaneous interaction of the two functions with different
4. CONCLUSIONS AND OUTLOOK positions on a reactant molecule or a transition state. Part of
Bifunctional materials have been widely studied for use as the motivation for this approach is to mimic the behavior of
deoxygenation catalysts. The key reason that bifunctional enzymes, where spatial arrangement of multiple catalytic
5039 DOI: 10.1021/acscatal.6b00923
ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

domains can be used to enable specific catalysis.179 This For example, Figure 17 shows the aldol condensation of 4-
approach can be used to combine catalyst functional groups nitrobenzaldehyde and acetone using a bifunctional tethered
that would be incompatible in solution. For example, acidic catalyst consisting of a sulfonic acid and an amine immobilized
sulfonic acid groups have been cotethered with basic amine on a support.191 In this example, the acetone is deprotonated to
groups on a silica support, shown in Figure 17.180 The create a nucleophilic enolate which may then attack the
electron-deficient aldehyde to form the aldol product.191 Proper
design and synthesis prevents the acid and base from
neutralizing each other and creates a unique bifunctional
catalyst.190 A similar approach is taken by many enzymes, which
operate at high selectivity.189 By being able to stabilize
incompatible functional groups in a way that allows them to
maintain their individual functionalities, enzymes such as
adolases are able to catalyze multistep reactions that would
not take place in solution.189 Mimicking this, the field of
heterogeneous catalysis has adopted a similar approach for
these acid−base-catalyzed reactions.
There is the potential to integrate the use of well-defined
tethered organocatalysts with reactive surfaces such as metal
oxides. Bifunctional surfaces have already been employed for
such reactions. For example, this has been demonstrated using
solid base catalysts in single-phase, mixed oxides containing
both acidic and basic sites, as well as on novel catalysts at the
water/oil interface of an emulsion.186−188,192 The high water
content of pyrolysis oil can cause phase separation; taking
advantage of differing partition coefficients of these reactant
molecules could lead to improved selectivites to targeted
products. Smaller aldehydes have also been used as probe
Figure 17. Example of the catalytic cycle proposed for a tethered molecules to better understand condensation reaction mecha-
bifunctional catalyst in the aldol condensation of 4-nitrobenzaldehyde nisms over novel materials. For example, the mixed oxide
and acetone at 323 K. Adapted from ref 191. Ce0.5Zr0.5O2 has been studied for propanal conversion to C6−
C9 products.192
bifunctional catalyst was found to be more active for a model Overall, the approach of using bifunctional (or multifunc-
aldol condensation reaction than the monofunctional controls. tional) catalysts is an important direction for catalysis in general
Interestingly, it was also more active than a physical mixture of and particularly for reactions that require significantly different
acid- and base-functionalized materials, indicating that prox- types of elementary steps as part of an overall reaction. For this
imity of the acid and base sites is important in maximizing reason, efforts to employ rational design of interfaces will
catalyst rate. Whereas in this example the proximity of the become important for supplementing the promising initial work
functional groups was not controlled in a direct way, other in demonstrating bifunctional catalysis for biomass conversion.
examples of tethered bifunctional catalysts have sought to tailor
the proximity of groups with high precision, for example
through molecular imprinting approaches.181 Other work has
■ AUTHOR INFORMATION
Corresponding Author
exploited the bifunctional effect of a tethered organic catalyst *E-mail: will.medlin@colorado.edu.
combined with acidic groups on the support.182 A related
example is the use of molecular catalysts for homogeneously Notes
catalyzing chiral syntheses. Using transition-metal-based The authors declare no competing financial interest.
complexes that contain both a Lewis acid and Lewis base,
C−H, C−C, C−O, and C−N bonds may be formed.183,184
Through proper tuning of the bifunctional complex issues with
■ ACKNOWLEDGMENTS
This work was supported by the Department of Energy
acid−base neutralization can be mitigated and high degrees of Bioenergy Technologies Office (BETO) under Contract no.
enantioselective reduction and C−C bond formation may be DE-AC36-08-GO28308. A.M.R. acknowledges support from
achieved.185 the National Science Foundation (Award CHE-1464979) and
As an example of the promise of these catalysts, aldol partial support from the Department of Education Graduate
condensation reactions may be considered. These reactions of Assistantships in Areas of National Need (GAANN).


aldehydes such as furfurals with short ketones followed by
hydrogenation can yield gasoline/diesel range hydrocarbons REFERENCES
(C8−C13).186−188 Coupling reactions like aldol condensation
may be improved by utilizing a bifunctional catalyst rather than (1) Bridgwater, A. V. Biomass Bioenergy 2012, 38, 68−94.
(2) Di Blasi, C. Prog. Energy Combust. Sci. 2009, 35, 121−140.
a single acidic or basic material.189 The two components of this (3) Bridgwater, A.; Peacocke, G. Renewable Sustainable Energy Rev.
bifunctional catalyst can each activate one of the substrates to 2000, 4, 1−73.
be reacted, resulting in a lower energy pathway.190 These (4) Pham, T. N.; Shi, D.; Resasco, D. E. Appl. Catal., B 2014, 145,
catalysts are often composed of an acid tuned to activate one of 10−23.
the substrates for electrophilic addition and a base that is (5) Mettler, M. S.; Vlachos, D. G.; Dauenhauer, P. J. Energy Environ.
selected to increase the nucleophilicity of the other substrate. Sci. 2012, 5, 7797−7809.

5040 DOI: 10.1021/acscatal.6b00923


ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

(6) Lu, Q.; Li, W.-Z.; Zhu, X.-F. Energy Convers. Manage. 2009, 50, (41) Chen, J.; Shi, H.; Li, L.; Li, K. Appl. Catal., B 2014, 144, 870−
1376−1383. 884.
(7) Zacher, A. H.; Olarte, M. V.; Santosa, D. M.; Elliott, D. C.; Jones, (42) Li, K.; Wang, R.; Chen, J. Energy Fuels 2011, 25, 854−863.
S. B. Green Chem. 2014, 16, 491−515. (43) Lee, Y.-K.; Oyama, S. T. J. Catal. 2006, 239, 376−389.
(8) Ruddy, D. A.; Schaidle, J. A.; Ferrell, J. R., III; Wang, J.; Moens, (44) Zhang, W.; Zhang, Y.; Zhao, L.; Wei, W. Energy Fuels 2010, 24,
L.; Hensley, J. E. Green Chem. 2014, 16, 454−490. 2052−2059.
(9) Ingram, L.; Mohan, D.; Bricka, M.; Steele, P.; Strobel, D.; (45) Talukdar, A. K.; Bhattacharyya, K. G.; Sivasanker, S. Appl. Catal.,
Crocker, D.; Mitchell, B.; Mohammad, J.; Cantrell, K.; Pittman, C. U., A 1993, 96, 229−239.
Jr Energy Fuels 2008, 22, 614−625. (46) Park, C.; Keane, M. A. J. Colloid Interface Sci. 2003, 266, 183−
(10) Xu, J.; White, T.; Li, P.; He, C.; Yu, J.; Yuan, W.; Han, Y.-F. J. 194.
Am. Chem. Soc. 2010, 132, 10398−10406. (47) Zhao, C.; He, J.; Lemonidou, A. A.; Li, X.; Lercher, J. A. J. Catal.
(11) Bitter, J.; Seshan, K.; Lercher, J. J. Catal. 1998, 176, 93−101. 2011, 280, 8−16.
(12) Robinson, A. M.; Ferguson, G. A.; Gallagher, J. R.; Cheah, S.; (48) Columbia, M.; Thiel, P. J. Electroanal. Chem. 1994, 369, 1−14.
Beckham, G. T.; Schaidle, J. A.; Hensley, J. E.; Medlin, J. W. ACS (49) Mavrikakis, M.; Barteau, M. A. J. Mol. Catal. A: Chem. 1998,
Catal. 2016, 6, 4356−4368. 131, 135−147.
(13) Kitchin, J.; Nørskov, J. K.; Barteau, M.; Chen, J. J. Chem. Phys. (50) Ferrin, P.; Simonetti, D.; Kandoi, S.; Kunkes, E.; Dumesic, J. A.;
2004, 120, 10240−10246. Nørskov, J. K.; Mavrikakis, M. J. Am. Chem. Soc. 2009, 131, 5809−
(14) Stamenkovic, V. R.; Mun, B. S.; Mayrhofer, K. J.; Ross, P. N.; 5815.
Markovic, N. M. J. Am. Chem. Soc. 2006, 128, 8813−8819. (51) Rajadurai, S. Catal. Rev.: Sci. Eng. 1994, 36, 385−403.
(15) Stamenkovic, V. R.; Mun, B. S.; Arenz, M.; Mayrhofer, K. J.; (52) Columbia, M.; Crabtree, A.; Thiel, P. J. Am. Chem. Soc. 1992,
Lucas, C. A.; Wang, G.; Ross, P. N.; Markovic, N. M. Nat. Mater. 2007, 114, 1231−1237.
6, 241−247. (53) Hung, W.-H.; Bernasek, S. L. Surf. Sci. 1996, 346, 165−188.
(16) Chen, J. G.; Menning, C. A.; Zellner, M. B. Surf. Sci. Rep. 2008, (54) Davis, J.; Barteau, M. Surf. Sci. 1991, 256, 50−66.
63, 201−254. (55) Erley, W.; Sander, D. J. Vac. Sci. Technol., A 1989, 7, 2238−2244.
(17) Liu, P.; Nørskov, J. K. Phys. Chem. Chem. Phys. 2001, 3, 3814− (56) Wildschut, J.; Mahfud, F. H.; Venderbosch, R. H.; Heeres, H. J.
3818. Ind. Eng. Chem. Res. 2009, 48, 10324−10334.
(18) Rebelli, J.; Detwiler, M.; Ma, S.; Williams, C. T.; Monnier, J. R. J. (57) Sheu, Y.-H. E.; Anthony, R. G.; Soltes, E. J. Fuel Process. Technol.
Catal. 2010, 270, 224−233. 1988, 19, 31−50.
(19) Maroun, F.; Ozanam, F.; Magnussen, O.; Behm, R. Science 2001, (58) Bu, Q.; Lei, H.; Zacher, A. H.; Wang, L.; Ren, S.; Liang, J.; Wei,
293, 1811−1814. Y.; Liu, Y.; Tang, J.; Zhang, Q.; Ruan, R. Bioresour. Technol. 2012, 124,
(20) Neurock, M.; Mei, D. Top. Catal. 2002, 20, 5−23. 470−477.
(21) Lu, Y.-C.; Xu, Z.; Gasteiger, H. A.; Chen, S.; Hamad-Schifferli, (59) Sun, J.; Karim, A. M.; Zhang, H.; Kovarik, L.; Li, X. S.; Hensley,
K.; Shao-Horn, Y. J. Am. Chem. Soc. 2010, 132, 12170−12171. A. J.; McEwen, J.-S.; Wang, Y. J. Catal. 2013, 306, 47−57.
(22) Zhang, Y.; Zhang, H.; Ma, Y.; Cheng, J.; Zhong, H.; Song, S.; (60) Chang, J.; Danuthai, T.; Dewiyanti, S.; Wang, C.; Borgna, A.
Ma, H. J. Power Sources 2010, 195, 142−145. ChemCatChem 2013, 5, 3041−3049.
(23) Kong, F.-D.; Zhang, S.; Yin, G.-P.; Zhang, N.; Wang, Z.-B.; Du, (61) Hong, Y.; Zhang, H.; Sun, J.; Ayman, K. M.; Hensley, A. J.; Gu,
C.-Y. J. Power Sources 2012, 210, 321−326.
M.; Engelhard, M. H.; McEwen, J.-S.; Wang, Y. ACS Catal. 2014, 4,
(24) Dau, H.; Limberg, C.; Reier, T.; Risch, M.; Roggan, S.; Strasser,
3335−3345.
P. ChemCatChem 2010, 2, 724−761.
(62) Lu, S.; Lonergan, W. W.; Bosco, J. P.; Wang, S.; Zhu, Y.; Xie, Y.;
(25) Chiusoli, G. P.; Maitlis, P. M. Metal-catalysis in Industrial Organic
Chen, J. G. J. Catal. 2008, 259, 260−268.
Processes, 3rd ed.; Chiusoli, G. P. M., Peter, M, Eds.; Royal Society of
(63) Chapman, K. W.; Sava, D. F.; Halder, G. J.; Chupas, P. J.;
Chemistry: Cambridge, U.K., 2008.
(26) Ono, Y. Catal. Today 2003, 81, 3−16. Nenoff, T. M. J. Am. Chem. Soc. 2011, 133, 18583−18585.
(27) Gong, J. Chem. Rev. 2012, 112, 2987−3054. (64) Medema, J.; Van Bokhoven, J.; Kuiper, A. J. Catal. 1972, 25,
(28) Fu, Q.; Saltsburg, H.; Flytzani-Stephanopoulos, M. Science 2003, 238−244.
301, 935−938. (65) Auroux, A. Top. Catal. 1997, 4, 71−89.
(29) Fu, Q.; Deng, W.; Saltsburg, H.; Flytzani-Stephanopoulos, M. (66) Nie, L.; Resasco, D. E. J. Catal. 2014, 317, 22−29.
Appl. Catal., B 2005, 56, 57−68. (67) Tan, Q.; Wang, G.; Nie, L.; Dinse, A.; Buda, C.; Shabaker, J.;
(30) Sakurai, H.; Ueda, A.; Kobayashi, T.; Haruta, M. Chem. Resasco, D. E. ACS Catal. 2015, 5, 6271−6283.
Commun. 1997, 271−272. (68) Chen, C.; Chen, G.; Yang, F.; Wang, H.; Han, J.; Ge, Q.; Zhu, X.
(31) Xu, M.; Gines, M. J.; Hilmen, A.-M.; Stephens, B. L.; Iglesia, E. J. Chem. Eng. Sci. 2015, 135, 145−154.
Catal. 1997, 171, 130−147. (69) Nelson, R. C.; Baek, B.; Ruiz, P.; Goundie, B.; Brooks, A.;
(32) Do, P. T.; Foster, A. J.; Chen, J.; Lobo, R. F. Green Chem. 2012, Wheeler, M. C.; Frederick, B. G.; Grabow, L. C.; Austin, R. N. ACS
14, 1388−1397. Catal. 2015, 5, 6509−6523.
(33) Foster, A. J.; Do, P. T.; Lobo, R. F. Top. Catal. 2012, 55, 118− (70) Newman, C.; Zhou, X.; Goundie, B.; Ghampson, I. T.; Pollock,
128. R. A.; Ross, Z.; Wheeler, M. C.; Meulenberg, R. W.; Austin, R. N.;
(34) Nie, L.; de Souza, P. M.; Noronha, F. B.; An, W.; Sooknoi, T.; Frederick, B. G. Appl. Catal., A 2014, 477, 64−74.
Resasco, D. E. J. Mol. Catal. A: Chem. 2014, 388-389, 47−55. (71) Griffin, M. B.; Ferguson, G. A.; Ruddy, D. A.; Biddy, M. J.;
(35) Lee, C. R.; Yoon, J. S.; Suh, Y.-W.; Choi, J.-W.; Ha, J.-M.; Suh, Beckham, G. T.; Schaidle, J. A. ACS Catal. 2016, 6, 2715−2727.
D. J.; Park, Y.-K. Catal. Commun. 2012, 17, 54−58. (72) Nimmanwudipong, T.; Runnebaum, R. C.; Block, D. E.; Gates,
(36) Zhao, C.; Kou, Y.; Lemonidou, A. A.; Li, X.; Lercher, J. A. Chem. B. C. Energy Fuels 2011, 25, 3417−3427.
Commun. 2010, 46, 412−414. (73) Horácě k, J.; Št’ávová, G.; Kelbichová, V.; Kubička, D. Catal.
(37) Daniel, O. M.; DeLaRiva, A.; Kunkes, E. L.; Datye, A. K.; Today 2013, 204, 38−45.
Dumesic, J. A.; Davis, R. J. ChemCatChem 2010, 2, 1107−1114. (74) Hong, D.-Y.; Miller, S. J.; Agrawal, P. K.; Jones, C. W. Chem.
(38) Zhao, C.; Kou, Y.; Lemonidou, A. A.; Li, X.; Lercher, J. A. Commun. 2010, 46, 1038−1040.
Angew. Chem. 2009, 121, 4047−4050. (75) Olcese, R. N.; François, J.; Bettahar, M.; Petitjean, D.; Dufour,
(39) Mortensen, P. M.; Grunwaldt, J.-D.; Jensen, P. A.; Knudsen, K.; A. Energy Fuels 2013, 27, 975−984.
Jensen, A. D. Appl. Catal., A 2011, 407, 1−19. (76) Olcese, R.; Bettahar, M.; Petitjean, D.; Malaman, B.; Giovanella,
(40) Furimsky, E. Appl. Catal., A 2000, 199, 147−190. F.; Dufour, A. Appl. Catal., B 2012, 115-116, 63−73.

5041 DOI: 10.1021/acscatal.6b00923


ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

(77) Shin, E.-J.; Keane, M. A. Ind. Eng. Chem. Res. 2000, 39, 883− (112) Chia, M.; Pagan-Torres, Y. J.; Hibbitts, D.; Tan, Q.; Pham, H.
892. N.; Datye, A. K.; Neurock, M.; Davis, R. J.; Dumesic, J. A. J. Am. Chem.
(78) Long, J.; Shu, S.; Wu, Q.; Yuan, Z.; Wang, T.; Xu, Y.; Zhang, X.; Soc. 2011, 133, 12675−12689.
Zhang, Q.; Ma, L. Energy Convers. Manage. 2015, 105, 570−577. (113) Amada, Y.; Watanabe, H.; Hirai, Y.; Kajikawa, Y.; Nakagawa,
(79) Nakagawa, Y.; Ishikawa, M.; Tamura, M.; Tomishige, K. Green Y.; Tomishige, K. ChemSusChem 2012, 5, 1991−1999.
Chem. 2014, 16, 2197−2203. (114) Tomishige, K.; Tamura, M.; Nakagawa, Y. Chem. Rec. 2014, 14,
(80) Ishikawa, M.; Tamura, M.; Nakagawa, Y.; Tomishige, K. Appl. 1041−1054.
Catal., B 2016, 182, 193−203. (115) Nakagawa, Y.; Ning, X.; Amada, Y.; Tomishige, K. Appl. Catal.,
(81) Pang, S. H.; Medlin, J. W. ACS Catal. 2011, 1, 1272−1283. A 2012, 433-434, 128−134.
(82) Pang, S. H.; Román, A. M.; Medlin, J. W. J. Phys. Chem. C 2012, (116) Amada, Y.; Shinmi, Y.; Koso, S.; Kubota, T.; Nakagawa, Y.;
116, 13654−13660. Tomishige, K. Appl. Catal., B 2011, 105, 117−127.
(83) Davis, J.; Barteau, M. J. Mol. Catal. 1992, 77, 109−124. (117) Amada, Y.; Watanabe, H.; Tamura, M.; Nakagawa, Y.;
(84) Pang, S. H.; Schoenbaum, C. A.; Schwartz, D. K.; Medlin, J. W. Okumura, K.; Tomishige, K. J. Phys. Chem. C 2012, 116, 23503−
Nat. Commun. 2013, 4, 1−6. 23514.
(85) Lien, C.-H.; Medlin, J. W. J. Phys. Chem. C 2014, 118, 23783− (118) Koso, S.; Watanabe, H.; Okumura, K.; Nakagawa, Y.;
23789. Tomishige, K. J. Phys. Chem. C 2012, 116, 3079−3090.
(86) Huber, G. W.; Iborra, S.; Corma, A. Chem. Rev. 2006, 106, (119) De Lange, M.; Van Ommen, J.; Lefferts, L. Appl. Catal., A
2001, 220, 41−49.
4044−4098.
(120) Lu, Q.; Zhang, Y.; Tang, Z.; Li, W.-z.; Zhu, X.-f. Fuel 2010, 89,
(87) Popov, A.; Kondratieva, E.; Goupil, J. M.; Mariey, L.; Bazin, P.;
2096−2103.
Gilson, J.-P.; Travert, A.; Maugé, F. J. Phys. Chem. C 2010, 114,
(121) Pham, T. N.; Sooknoi, T.; Crossley, S. P.; Resasco, D. E. ACS
15661−15670. Catal. 2013, 3, 2456−2473.
(88) Meloni, D.; Monaci, R.; Solinas, V.; Berlier, G.; Bordiga, S.; (122) Gaertner, C. A.; Serrano-Ruiz, J. C.; Braden, D. J.; Dumesic, J.
Rossetti, I.; Oliva, C.; Forni, L. J. Catal. 2003, 214, 169−178. A. J. Catal. 2009, 266, 71−78.
(89) Alonso, D. M.; Wettstein, S. G.; Dumesic, J. A. Chem. Soc. Rev. (123) Renz, M. Eur. J. Org. Chem. 2005, 2005, 979−988.
2012, 41, 8075−8098. (124) Iwata, Y.; Araki, Y.; Honna, K.; Miki, Y.; Sato, K.; Shimada, H.
(90) Haynes, W. M. CRC Handbook of Chemistry and Physics; CRC Catal. Today 2001, 65, 335−341.
Press: New York, 2014. (125) Chianelli, R. R.; Siadati, M. H.; De la Rosa, M. P.; Berhault, G.;
(91) Dean, J. A. Lange’s Handbook of Chemistry; McGraw-Hill: New Wilcoxon, J. P.; Bearden, R., Jr; Abrams, B. L. Catal. Rev.: Sci. Eng.
York, 1985. 2006, 48, 1−41.
(92) Gisdakis, P.; Antonczak, S.; Rösch, N. Organometallics 1999, 18, (126) Tauster, S.; Pecoraro, T.; Chianelli, R. J. Catal. 1980, 63, 515−
5044−5056. 519.
(93) Skoplyak, O.; Barteau, M. A.; Chen, J. G. G. Catal. Today 2009, (127) Bahl, O.; Evans, E.; Thomas, J. Proc. R. Soc. London, Ser. A
147, 150−157. 1968, 306, 53−65.
(94) Skoplyak, O.; Barteau, M. A.; Chen, J. G. G. J. Phys. Chem. B (128) Romero, Y.; Richard, F.; Brunet, S. Appl. Catal., B 2010, 98,
2006, 110, 1686−1694. 213−223.
(95) Lu, S.; Lonergan, W. W.; Bosco, J. P.; Wang, S.; Zhu, Y.; Xie, Y.; (129) Wang, W.; Zhang, K.; Qiao, Z.; Li, L.; Liu, P.; Yang, Y. Catal.
Chen, J. G. J. Catal. 2008, 259, 260−268. Commun. 2014, 56, 17−22.
(96) Lu, S.; Menning, C. A.; Zhu, Y.; Chen, J. G. ChemPhysChem (130) Bui, V. N.; Laurenti, D.; Afanasiev, P.; Geantet, C. Appl. Catal.,
2009, 10, 1763−1765. B 2011, 101, 239−245.
(97) Lonergan, W. W.; Vlachos, D. G.; Chen, J. G. J. Catal. 2010, (131) Ruinart de Brimont, M.; Dupont, C.; Daudin, A.; Geantet, C.;
271, 239−250. Raybaud, P. J. Catal. 2012, 286, 153−164.
(98) Humbert, M. P.; Chen, J. G. J. Catal. 2008, 257, 297−306. (132) Yoosuk, B.; Tumnantong, D.; Prasassarakich, P. Fuel 2012, 91,
(99) Stanislaus, A.; Absi-Halabi, M.; Al-Dolama, K. Appl. Catal. 1989, 246−252.
50, 237−245. (133) Popov, A.; Kondratieva, E.; Mariey, L.; Goupil, J. M.; El Fallah,
(100) Shinmi, Y.; Koso, S.; Kubota, T.; Nakagawa, Y.; Tomishige, K. J.; Gilson, J.-P.; Travert, A.; Maugé, F. J. Catal. 2013, 297, 176−186.
Appl. Catal., B 2010, 94, 318−326. (134) Dupont, C.; Lemeur, R.; Daudin, A.; Raybaud, P. J. Catal.
(101) Kunkes, E. L.; Simonetti, D. A.; Dumesic, J. A.; Pyrz, W. D.; 2011, 279, 276−286.
Murillo, L. E.; Chen, J. G.; Buttrey, D. J. J. Catal. 2008, 260, 164−177. (135) Oyama, S. Catal. Today 1992, 15, 179−200.
(102) Simonetti, D.; Kunkes, E.; Dumesic, J. J. Catal. 2007, 247, (136) Oyama, S. T.; Gott, T.; Zhao, H.; Lee, Y.-K. Catal. Today 2009,
143, 94−107.
298−306.
(137) Hwu, H. H.; Chen, J. G. Chem. Rev. 2005, 105, 185−212.
(103) Lee, J.; Kim, Y. T.; Huber, G. W. Green Chem. 2014, 16, 708−
(138) Zhao, H.; Li, D.; Bui, P.; Oyama, S. Appl. Catal., A 2011, 391,
718. 305−310.
(104) Chia, M.; Pagán-Torres, Y. J.; Hibbitts, D.; Tan, Q.; Pham, H. (139) Ren, H.; Chen, Y.; Huang, Y.; Deng, W.; Vlachos, D. G.; Chen,
N.; Datye, A. K.; Neurock, M.; Davis, R. J.; Dumesic, J. A. J. Am. Chem. J. G. Green Chem. 2014, 16, 761−769.
Soc. 2011, 133, 12675−12689. (140) Weigert, E. C.; Stottlemyer, A. L.; Zellner, M. B.; Chen, J. G. J.
(105) Zhang, L.; Karim, A. M.; Engelhard, M. H.; Wei, Z.; King, D. Phys. Chem. C 2007, 111, 14617−14620.
L.; Wang, Y. J. Catal. 2012, 287, 37−43. (141) Levy, R.; Boudart, M. Science 1973, 181, 547−549.
(106) Pallassana, V.; Neurock, M. J. Catal. 2002, 209, 289−305. (142) Ren, H.; Yu, W.; Salciccioli, M.; Chen, Y.; Huang, Y.; Xiong,
(107) Ma, L.; He, D. Top. Catal. 2009, 52, 834−844. K.; Vlachos, D. G.; Chen, J. G. ChemSusChem 2013, 6, 798−801.
(108) Chen, K.; Koso, S.; Kubota, T.; Nakagawa, Y.; Tomishige, K. (143) Han, J.; Duan, J.; Chen, P.; Lou, H.; Zheng, X.; Hong, H. Green
ChemCatChem 2010, 2, 547−555. Chem. 2011, 13, 2561−2568.
(109) Koso, S.; Nakagawa, Y.; Tomishige, K. J. Catal. 2011, 280, (144) Jongerius, A. L.; Gosselink, R. W.; Dijkstra, J.; Bitter, J. H.;
221−229. Bruijnincx, P. C.; Weckhuysen, B. M. ChemCatChem 2013, 5, 2964−
(110) Koso, S.; Furikado, I.; Shimao, A.; Miyazawa, T.; Kunimori, K.; 2972.
Tomishige, K. Chem. Commun. 2009, 2035−2037. (145) Ramanathan, S.; Oyama, S. J. Phys. Chem. 1995, 99, 16365−
(111) Amada, Y.; Koso, S.; Nakagawa, Y.; Tomishige, K. 16372.
ChemSusChem 2010, 3, 728−736. (146) Hargreaves, J. Coord. Chem. Rev. 2013, 257, 2015−2031.

5042 DOI: 10.1021/acscatal.6b00923


ACS Catal. 2016, 6, 5026−5043
ACS Catalysis Review

(147) Dolce, G.; Savage, P.; Thompson, L. Energy Fuels 1997, 11, (176) Boucher, M. B.; Zugic, B.; Cladaras, G.; Kammert, J.;
668−675. Marcinkowski, M. D.; Lawton, T. J.; Sykes, E. C. H.; Flytzani-
(148) Ghampson, I. T.; Sepúlveda, C.; Garcia, R.; Radovic, L. R.; Stephanopoulos, M. Phys. Chem. Chem. Phys. 2013, 15, 12187−12196.
Fierro, J.; DeSisto, W. J.; Escalona, N. Appl. Catal., A 2012, 439-440, (177) Flytzani-Stephanopoulos, M.; Gates, B. C. Annu. Rev. Chem.
111−124. Biomol. Eng. 2012, 3, 545−574.
(149) Frapper, G.; Pélissier, M.; Hafner, J. J. Phys. Chem. B 2000, 104, (178) Zhai, Y.; Pierre, D.; Si, R.; Deng, W.; Ferrin, P.; Nilekar, A. U.;
11972−11976. Peng, G.; Herron, J. A.; Bell, D. C.; Saltsburg, H.; Mavrikakis, M.;
(150) Ghampson, I.; Sepulveda, C.; Garcia, R.; Frederick, B.; Flytzani-Stephanopoulos, M. Science 2010, 329, 1633−1636.
Wheeler, M.; Escalona, N.; DeSisto, W. Appl. Catal., A 2012, 413-414, (179) Margelefsky, E. L.; Zeidan, R. K.; Davis, M. E. Chem. Soc. Rev.
78−84. 2008, 37, 1118−1126.
(151) Griffin, M. B.; Baddour, F. G.; Habas, S. E.; Ruddy, D. A.; (180) Zeidan, R. K.; Hwang, S. J.; Davis, M. E. Angew. Chem., Int. Ed.
2006, 45, 6332−6335.
Schaidle, J. A. Top. Catal. 2016, 59, 124−137.
(181) Katz, A.; Davis, M. E. Nature 2000, 403, 286−289.
(152) Ruff, M.; Takehiro, N.; Liu, P.; Nørskov, J. K.; Behm, R. J.
(182) Notestein, J. M.; Katz, A. Chem. - Eur. J. 2006, 12, 3954−3965.
ChemPhysChem 2007, 8, 2068−2071.
(183) Sawamura, M.; Ito, Y. Chem. Rev. 1992, 92, 857−871.
(153) Sachtler, J.; Somorjai, G. J. Catal. 1983, 81, 77−94.
(184) Takamura, M.; Funabashi, K.; Kanai, M.; Shibasaki, M. J. Am.
(154) Chen, M.; Kumar, D.; Yi, C.-W.; Goodman, D. W. Science
Chem. Soc. 2000, 122, 6327−6328.
2005, 310, 291−293. (185) Ikariya, T. In Bifunctional Molecular Catalysis; Springer: Tokyo,
(155) Scott, R. W.; Sivadinarayana, C.; Wilson, O. M.; Yan, Z.; 2011.
Goodman, D. W.; Crooks, R. M. J. Am. Chem. Soc. 2005, 127, 1380− (186) Chheda, J. N.; Dumesic, J. A. Catal. Today 2007, 123, 59−70.
1381. (187) Zapata, P. A.; Faria, J.; Ruiz, M. P.; Resasco, D. E. Top. Catal.
(156) Divins, N. J.; Angurell, I.; Escudero, C.; Pérez-Dieste, V.; 2012, 55, 38−52.
Llorca, J. Science 2014, 346, 620−623. (188) Crossley, S.; Faria, J.; Shen, M.; Resasco, D. E. Science 2010,
(157) Tao, F.; Grass, M. E.; Zhang, Y.; Butcher, D. R.; Aksoy, F.; 327, 68−72.
Aloni, S.; Altoe, V.; Alayoglu, S.; Renzas, J. R.; Tsung, C.-K.; Zhu, Z.; (189) Brunelli, N. A.; Jones, C. W. J. Catal. 2013, 308, 60−72.
Liu, Z.; Salmeron, M.; Somorjai, G. A. J. Am. Chem. Soc. 2010, 132, (190) Brunelli, N. A.; Venkatasubbaiah, K.; Jones, C. W. Chem. Mater.
8697−8703. 2012, 24, 2433−2442.
(158) Tao, F.; Dag, S.; Wang, L.-W.; Liu, Z.; Butcher, D. R.; Bluhm, (191) Zeidan, R. K.; Davis, M. E. J. Catal. 2007, 247, 379−382.
H.; Salmeron, M.; Somorjai, G. A. Science 2010, 327, 850−853. (192) Gangadharan, A.; Shen, M.; Sooknoi, T.; Resasco, D. E.;
(159) Saavedra, J.; Doan, H. A.; Pursell, C. J.; Grabow, L. C.; Mallinson, R. G. Appl. Catal., A 2010, 385, 80−91.
Chandler, B. D. Science 2014, 345, 1599−1602.
(160) Green, I. X.; Tang, W.; Neurock, M.; Yates, J. T. Science 2011,
333, 736−739.
(161) Stair, P.; Marshall, C.; Xiong, G.; Feng, H.; Pellin, M.; Elam, J.;
Curtiss, L.; Iton, L.; Kung, H.; Kung, M.; Wang, H.-H. Top. Catal.
2006, 39, 181−186.
(162) Lu, J.; Stair, P. C. Angew. Chem., Int. Ed. 2010, 49, 2547−2551.
(163) Lu, J.; Fu, B.; Kung, M. C.; Xiao, G.; Elam, J. W.; Kung, H. H.;
Stair, P. C. Science 2012, 335, 1205−1208.
(164) Pellin, M.; Stair, P.; Xiong, G.; Elam, J.; Birrell, J.; Curtiss, L.;
George, S.; Han, C.; Iton, L.; Kung, H.; Kung, M.; Wang, H.-H. Catal.
Lett. 2005, 102, 127−130.
(165) Wank, J. R.; George, S. M.; Weimer, A. W. J. Am. Ceram. Soc.
2004, 87, 762−765.
(166) Li, J.; Liang, X.; King, D. M.; Jiang, Y.-B.; Weimer, A. W. Appl.
Catal., B 2010, 97, 220−226.
(167) Gould, T. D.; Lubers, A. M.; Neltner, B. T.; Carrier, J. V.;
Weimer, A. W.; Falconer, J. L.; Medlin, J. W. J. Catal. 2013, 303, 9−15.
(168) Lei, Y.; Mehmood, F.; Lee, S.; Greeley, J.; Lee, B.; Seifert, S.;
Winans, R. E.; Elam, J. W.; Meyer, R. J.; Redfern, P. C.; Teschner, D.;
Schlogl, R.; Pellin, M. J.; Curtiss, L. A.; Vajda, S. Science 2010, 328,
224−228.
(169) Vajda, S.; Pellin, M. J.; Greeley, J. P.; Marshall, C. L.; Curtiss, L.
A.; Ballentine, G. A.; Elam, J. W.; Catillon-Mucherie, S.; Redfern, P. C.;
Mehmood, F.; Zapol, P. Nat. Mater. 2009, 8, 213−216.
(170) Yin, Y.; Lu, Y.; Gates, B.; Xia, Y. J. Am. Chem. Soc. 2001, 123,
8718−8729.
(171) Xu, Z.; Xiao, F.-S.; Purnell, S.; Alexeev, O.; Kawi, S.; Deutsch,
S.; Gates, B. Nature 1994, 372, 346−348.
(172) Gates, B. C. J. Mol. Catal. A: Chem. 2000, 163, 55−65.
(173) Bratlie, K. M.; Lee, H.; Komvopoulos, K.; Yang, P.; Somorjai,
G. A. Nano Lett. 2007, 7, 3097−3101.
(174) Kwon, G.; Ferguson, G. A.; Heard, C. J.; Tyo, E. C.; Yin, C.;
DeBartolo, J.; Seifert, S. n.; Winans, R. E.; Kropf, A. J.; Greeley, J.;
Johnston, R. L.; Curtiss, L. A.; Pellin, M. J.; Vajda, S. ACS Nano 2013,
7, 5808−5817.
(175) Matsubu, J. C.; Yang, V. N.; Christopher, P. J. Am. Chem. Soc.
2015, 137, 3076−3084.

5043 DOI: 10.1021/acscatal.6b00923


ACS Catal. 2016, 6, 5026−5043

You might also like