Anda di halaman 1dari 72

View Article Online

Green
View Journal

Chemistry
Cutting-edge research for a greener sustainable future
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: S. Kim, E. E. Kwon,
Y. T. Kim, S. Jung, H. J. Kim, G. W. Huber and J. Lee, Green Chem., 2019, DOI: 10.1039/C9GC01210A.

Volume 18 Number 7 7 April 2016 Pages 1821–2242 This is an Accepted Manuscript, which has been through the

Green Royal Society of Chemistry peer review process and has been
accepted for publication.
Chemistry
Cutting-edge research for a greener sustainable future Accepted Manuscripts are published online shortly after
www.rsc.org/greenchem

acceptance, before technical editing, formatting and proof reading.


Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. We will replace this Accepted Manuscript with the edited
and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the


author guidelines.

Please note that technical editing may introduce minor changes


to the text and/or graphics, which may alter content. The journal’s
ISSN 1463-9262 standard Terms & Conditions and the ethical guidelines, outlined
CRITICAL REVIEW
in our author and reviewer resource centre, still apply. In no
G. Chatel et al.
Heterogeneous catalytic oxidation for lignin valorization into valuable
chemicals: what results? What limitations? What trends?
event shall the Royal Society of Chemistry be held responsible
for any errors or omissions in this Accepted Manuscript or any
consequences arising from the use of any information it contains.

rsc.li/green-chem
Page 1 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

1 Recent Advances in Hydrodeoxygenation of Biomass-Derived Oxygenates

2 over Heterogeneous Catalysts

3
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

4 Soosan Kima,1, Eilhann E. Kwonb,1, Yong Tae Kimc,1, Sungyup Jungd,1, Hyung Ju Kimc,

5 George W. Hubere, Jechan Leea,*

Green Chemistry Accepted Manuscript


6

7 aDepartment of Environmental and Safety Engineering, Ajou University, Suwon 16499, Republic of Korea

8 bDepartment of Environment and Energy, Sejong University, Seoul 05006, Republic of Korea

9 cCarbon Resources Institute, Korea Research Institute of Chemical Technology, Daejeon 34114, Republic of Korea

10 dSchool of Mechatronic Systems Engineering, Simon Fraser University, Surrey, BC, V3T 0A3, Canada

11 eDepartment of Chemical and Biological Engineering, University of Wisconsin-Madison, Madison, WI 53706, USA

12

13 Abstract

14 Hydrodeoxygenation (HDO) using heterogeneous catalysts has received considerable attention as a

15 way of converting biomass-derived oxygenates into renewable fuels and chemicals. HDO involves

16 combinations of different reactions such as hydrogenation, hydrogenolysis, decarbonylation, and

17 dehydration. The reactions occur at different catalytic sites typically with heterogeneous catalysts (e.g.,

18 metal, acid, and bifunctional sites), making HDO reaction complex. Thus, the selection of active site type

19 is critical when designing effective heterogeneous catalysts for the HDO process. Catalyst stability is a

20 major issue of designing HDO catalysts. Traditional catalysts are not stable under HDO conditions (i.e., a

21 high partial pressure of water). Metal particles are leached and/or sintered. Solid-acid support undergoes

22 phase transformation and surface area loss. Condensation or polymerization of C=C or C=O bond within

23 biomass and its derived oxygenates is easy to conduct under high temperature in the HDO process,

1Theseauthors contributed equally to this work.


*Corresponding author: Prof. Jechan Lee (E-mail: jlee83@ajou.ac.kr)

1
Green Chemistry Page 2 of 71
View Article Online
DOI: 10.1039/C9GC01210A

24 causing carbon deposition on the catalyst. The poor catalyst stability is a major challenge that needs to be

25 overcome for innovation in HDO technologies. Therefore, herein, we focus on providing insight into how

26 to design effective catalysts for HDO. First, the roles of different catalytic sites in HDO and strategies to

27 stabilize these active sites are discussed. Current achievements in HDO of different biomass feedstocks
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

28 (e.g., model compounds and bio-oil) over various heterogeneous catalysts are highlighted. Recent

Green Chemistry Accepted Manuscript


29 developments in electrochemical HDO processes, which do not require gaseous hydrogen, are discussed.

30 Finally, it is discussed the points to be considered for further research into HDO over heterogeneous

31 catalysts and the pending challenges.

32

33 Keywords: biorefinery; bifunctional catalyst; electrochemical hydrodeoxygenation; metal; solid acid

34

35 1. Introduction

36 Global warming has become a significant environmental challenge in our society. To address this

37 challenge, biomass has attracted substantial attention in recent years as the only realistic renewable carbon

38 resource that represents an alternative to fossil fuels. Biomass can be transformed into transportation fuels

39 via different routes such as gasification to produce syngas, fast pyrolysis or hydrothermal liquefaction to

40 produce bio-oil, and hydrolysis to produce aqueous sugar.1

41 Hydrodeoxygenation (HDO) of biomass is a series of reactions that occurs during many different

42 biomass conversion technologies including hydrotreating of bio-oils,2 hydropyrolysis,3, 4 hydrogenolysis

43 of biomass into oxygenated chemicals,5 aqueous phase reforming of carbohydrates into fuels,6, 7 and

44 hydrotreating of organics acids.8, 9 The overall chemistry involves the reaction of hydrogen with an

45 oxygenated biomass-derived molecule to producing a more deoxygenated product and water. While

46 hydrotreating of petroleum derived products has been used for over 40 years,10 there are several new

47 challenges that are unique to biomass derived feedstocks that require new types of catalysts and catalyst

48 structures. The high water and organic content in these reactions can accelerate sintering and leaching of

2
Page 3 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

49 metals often requiring the use of precious metal catalysts. In addition, commonly used catalyst supports

50 can lose surface area and degrade under typical reaction conditions. Catalyst selectivity is often an issue

51 as a wide variety of undesired reactions can occur with these organic molecules often leading to undesired

52 humin formation. Other biogenic inhibitors can also poison the catalyst. Often expensive precious metal
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

53 catalysts are more stable than base metal catalysts and hence it would be highly desirable to have more

Green Chemistry Accepted Manuscript


54 active catalyst to decrease the catalyst cost.

55 There have recently been review articles in HDO of biomass feedstocks due to the importance of the

56 HDO reaction in biomass conversion and upgrading.11-15 For example, the HDO of diesel-range

57 oxygenates (e.g., vegetable oil) over different catalytic systems was reviewed by Dalai and co-workers.12

58 A review on catalytic upgrading of lignin-derived bio-oil via HDO with molecular sieves-based catalysts

59 was also recently reported.14 More recently, Lin et al. critically reviewed reaction pathways of HDO of

60 linear and ring-containing biomass oxygenates over metal carbide catalysts and the influence of reaction

61 conditions on the pathways.15

62 In this review, we first discuss the catalytic sites responsible for HDO. Current achievements in

63 catalytic HDO processes are then summarized. We also introduce efforts to improve catalyst stability

64 during HDO. In addition to discussions on the typical heterogeneous HDO catalysts, it is discussed recent

65 progress in electrochemical HDO which has been of great interest.

66

67 2. Catalytic sites for the reactions involved in hydrodeoxygenation of biomass-derived oxygenates

68 Figure 1 summarizes what needs to be considered when design a bifunctional catalyst for HDO.

69 Hydrogenation is reactions occurring at metal sites. Therefore, selection of the type of metallic sites (e.g.,

70 monometallic vs. bimetallic sites) is highly dependent on target HDO products. The stability of metal

71 particles under HDO conditions is also critical to select metal sites when designing a HDO catalyst. Solid-

72 acid support provides acid sites where dehydration takes place. Considering that the catalyst support with

73 a high surface area generally leads to a high metal surface area,16 surface area of the solid-acid support

3
Green Chemistry Page 4 of 71
View Article Online
DOI: 10.1039/C9GC01210A

74 should be taken into account to design a bifunctional catalyst. The solid-acid support also needs to be

75 hydrothermally stable under HDO conditions. The type of solid-acid support affects characteristics of

76 bifunctional metal–acid sites.

77
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

Green Chemistry Accepted Manuscript


78

79 Figure 1. HDO catalyst design variables


80

81 As depicted in Figure 1, metal sites can play a central role in HDO activity and selectivity. In this

82 section, we would like to provide information and discussions about performances of metal, solid acid,

83 and bifunctional catalysts for reactions that are potentially involved in HDO (hydrogenation,

84 decarbonylation/decarboxylation, dehydration, hydrogenolysis of C–O and C–O–C bonds, isomerization,

85 etc.). It would be helpful to selection of the parts of HDO catalysts to make highly active selective

86 catalysts for targeted HDO reactions. The effects of catalyst surface structure on performance of HDO of

87 biomass-derived feedstocks is not emphasized because a recent review focused on the relation between

88 catalyst surface structure and HDO performance and molecular-level mechanisms.13

89

90 2.1. Metal sites

4
Page 5 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

91 The reactions that occur on metal catalytic sites in HDO include the C–C bond cleavage via

92 decarbonylation and hydrogenation of various functionalities (e.g., carbonyl group (C=O), carboxylic acid

93 (–COOH), and C=C bond). The C–C bond cleavage reaction includes decarbonylation of aldehyde and

94 decarboxylation of carboxylic acid functionality at the end of the carbon chain17-19 with a
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

95 hydrogenation/dehydrogenation/water-gas shift reaction, and hydrocracking at the branched carbon

Green Chemistry Accepted Manuscript


96 atom.20

97 Table 1 summarizes the results of metal catalyst screening for hydrogenation of oxy functionalities, all

98 of which are available from previous literature. Metal carbide catalysts are not covered in this review

99 because HDO of biomass feedstocks over metal carbide catalysts has been reviewed very recently.15 Pd–

100 Fe (Fe/Pd molar ratio = 3) was the most active catalytic system among the tested catalysts for

101 hydrogenation of both furanic and non-furanic C=O functionalities.21 The high activity of the Pd–Fe

102 bimetallic catalyst is likely attributed to the interaction between Pd and Fe shifting the d-band center of

103 Pd.22 The change in electronic structure of Pd caused by adding Fe can weaken the adsorption strength of

104 reaction intermediates and products.23 In addition, Pd prevents the oxidation of Fe, leading to metallic Fe

105 remaining during HDO.24 Co has a high rate of hydrodeoxygenating carboxylic acids. A niobium

106 phosphate-supported Pd catalyst was active for HDO of triglycerides to C7-8 alkanes both by facilitating

107 hydrogenolysis of ester groups and by suppressing C–C bond cleavage.25 It was very recently reported

108 that Ru most favors deoxygenation while Pd most favors hydrogenation among Pt, Pd, Ru, Rh, Ni, and Cu

109 for HDO of eugenol.26

110

5
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 6 of 71

111 Table 1. Screening of metal sites based on their catalytic activities for hydrogenation reactions. All results are available from previous
112 literature.

Performance of the catalyst with the highest


Reaction Phase Feedstock Product Reaction conditions Activity order Ref.
activity

Hydrogenation of C=O Catalyst: Co–Pt/Al2O3


Gas Acetaldehyde Ethanol 70 °C; H2/acetaldehyde = 4:1 Co–Pt > Pt > Co 27

Green Chemistry Accepted Manuscript


bond Reaction rate: 3.75 × 10−3 min−1 μmolCO−1

Hydrogenation of C=O 100 °C; 5.4 MPa; 5 wt.% Catalyst: Ru/Al2O3


Aqueous Acetaldehyde Ethanol Ru ~ Ni > Pt > Pd > Co > Rh 28
bond feed concentration Reaction rate: 0.6 s−1 (normalized by H2 uptake)

Hydrogenation of C=O 100 °C; 5.4 MPa; 5 wt.% Catalyst: Ru/Al2O3


Aqueous Propionaldehyde n-Propanol Ru ~ Ni > Pt > Pd > Co > Rh 28
bond feed concentration Reaction rate: 0.5 s−1 (normalized by H2 uptake)

Hydrogenation of C=O 70 °C; Catalyst: Ni–Pt/Al2O3


Gas Propionaldehyde n-Propanol Ni–Pt > Ni > Pt 29
bond H2/propionaldehyde = 4:1 Reaction rate: 2.78 × 10−3 min−1 μmolCO−1

Pd–Fe > Pt–Co > Pt–Fe > Ru–Co > Pt–


Hydrogenation of C=O 100 °C; 5.4 MPa; 5 wt.% Catalyst: Pd–Fe/Al2O3
Aqueous Propionaldehyde n-Propanol Ni > Ru–Fe ~ Ru–Ni > Ru > Pd–Ni > 21
bond feed concentration Reaction rate: 1.8 s−1 (normalized by H2 uptake)
Pd–Co > Ni > Pt > Pd ~ Co

Hydrogenation of C=O 100 °C; 5.4 MPa; 5 wt.% Catalyst: Ru/Al2O3


Aqueous Acetone 2-Propanol Ru > Ni ~ Rh > Pt > Co > Pd 28
bond feed concentration Reaction rate: 4.7 s−1 (normalized by H2 uptake)

Hydrogenation of C=O Catalyst: Co–Pt/Al2O3


Gas Acetone 2-Propanol 35 °C; H2/acetone = 2:1 Co–Pt > Pt > Co 27
bond Reaction rate: 10.6 × 10−3 min−1 μmolCO−1

Catalyst: Ni/Al2O3; Co/Al2O3; Fe/Al2O3


Hydrogenation of C=O 100-250 °C; feed rate of
Gas Acetone 2-Propanol Ni > Co > Fe Conversion: 77%; 65%; 22% 30
bond 4.5 cm3 h-1
Selectivity: 64%; 69%; 4%

Hydrogenation of C=O Catalyst: Ru/activated C; Raney Ni


Aqueous Xylose Xylitol 100-125 °C; 1.4-5.5 MPa Ru > Ni > Rh > Pd 31, 32
bond Activation energy: 5 kcal mol−1; 6.5 kcal mol−1

Hydrogenation of C=O 100 °C; 5.4 MPa; 5 wt.% Catalyst: Ru/Al2O3


Aqueous Xylose Xylitol Ru > Ni ~ Co > Pt > Rh ~ Pd 28
bond feed concentration Reaction rate: 0.69 s−1 (normalized by H2 uptake)

6
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Page 7 of 71 Green Chemistry

Ru > Ru–Co > Pd–Ni > Ru–Ni > Ru–Fe


Hydrogenation of C=O 100 °C; 5.4 MPa; 5 wt.% Catalyst: Ru/Al2O3
Aqueous Xylose Xylitol > Pt-Ni > Pd–Co > Pd–Fe > Pt–Co > Ni 21
bond feed concentration Reaction rate: 0.69 s−1 (normalized by H2 uptake)
> Pt–Fe ~ Co > Pt > Pd

Hydrogenation of C=O 100 °C; 5.4 MPa; 4.8 Catalyst: Pd/Al2O3


Aqueous Furfural Furfuryl alcohol Pd ~ Ni > Co > Ru > Pt > Rh 28
bond wt.% feed concentration Reaction rate: 0.13 s−1 (normalized by H2 uptake)

Pd–Fe > Pd–Co > Pt–Fe > Ru–Fe > Ru–


Hydrogenation of C=O 100 °C; 5.4 MPa; 4.8 Catalyst: Pd–Fe/Al2O3
Aqueous Furfural Furfuryl alcohol Co > Pd–Ni > Pd > Pt–Co > Pt–Ni > Ni 21

Green Chemistry Accepted Manuscript


bond wt.% feed concentration Reaction rate: 1.3 s−1 (normalized by H2 uptake)
> Ru–Ni ~ Ru > Co > Pt

Hydrogenation of C=C Tetrahydrofurfuryl 80 °C; 5.4 MPa; 4.8 wt.% Catalyst: Pd/Al2O3
Aqueous Furfuryl alcohol Pd > Ni > Ru > Rh ~ Pt >> Co 28
bond alcohol (THFA) feed concentration Reaction rate: 0.08 s−1 (normalized by H2 uptake)

Catalyst: Ru/activated C
Hydrogenation of 110-290 °C; 5.17 MPa; 10
Aqueous Acetic acid Ethanol Ru > Pt ~ Rh > Pd ~ Ir > Ni > Cu Reaction rate: 120.7 h−1 at 175 °C (normalized 8
carboxylic acid wt.% feed concentration
by H2 uptake)

Catalyst: Co/Al2O3
Hydrogenolysis of C– 330 °C; 5 MPa; H2/oleic
Liquid Oleic acid Alkanes Co > Pd > Pt > Ni Reaction rate: 0.28 s−1 (normalized by CO 33
O bond acid = 1000:1
uptake)

Catalyst: Rh/ZrO2
Hydrogenolysis of C– Cyclohexanol, Rh ~ Rh–Pt > Rh–Pd > Pd > Pt >
Liquid Guaiacol 100 °C; 8 MPa; 5 h Conversion: 100% 34
O bond cyclohexane Pd–Pt
Selectivity: 12%

113

7
Green Chemistry Page 8 of 71
View Article Online
DOI: 10.1039/C9GC01210A

114 Apart from pristine metal catalysts discussed above, sulfided transition metal catalysts (e.g., Mo, Ni,

115 and Co) have been widely studied for HDO of bio-oil. For instance, Mo sulfide is well-known catalyst for

116 the removal of heteroatom from bio-oil.35 While a Ru catalyst produced a large amount of methane (i.e.,

117 product via C–C bond cleavage) for hydrotreating chemical mixtures as a model of bio-oil,36 a sulfided
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

118 Ru catalyst was effective at hydrogenating polyols and sugars without such a high methane formation.37

Green Chemistry Accepted Manuscript


119 Typical Al2O3-supported sulfided Ni–Mo and Co–Mo catalysts gave high yields of deoxygenated

120 compounds for different HDO reactions.38-43 For example, a Ni–Mo sulfide catalyst supported on Al2O3

121 was used for HDO of untreated microalgae oil to produce biodiesel.44 With the catalyst, 56.2% of C13-C20

122 hydrocarbons was obtained at 360 °C, 3.45 MPa H2, and WHSV = 0.65 h–1.

123 The metal sulfide catalysts are less expensive than precious metal catalysts, which is beneficial to their

124 applicability for HDO of biomass feedstocks.45 However, these sulfide catalysts can suffer from

125 deactivation attributed to the loss of sulfur and coking. Hydrogen sulfide (H2S) was used as a sulfiding

126 agent instead of carbon disulfide (CS2) as a way to reduce coking on sulfided catalysts, thereby increasing

127 catalytic activity of γ-Al2O3-supported NiMo and CoMo catalysts for HDO of aliphatic esters

128 (approximately 80% conversion using H2S but approximately 50% conversion using CS2).46 For Mo-

129 promoted sulfide catalysts (e.g., Mo–Raney Ni), Mo leaching can also be a problem.47

130

131 2.2. Solid acid support

132 Different reactions can occur during HDO on acid sites including dehydration (i.e., the removal of

133 oxygen from biomass feedstock in the form of water), isomerization, hydration, and hydrolysis.48-50 The

134 catalytic activity and selectivity for C–O bond scission via dehydration on the solid acid catalyst is a

135 function of the concentration of Brønsted and Lewis acid sites.51-53 For example, for dehydration of

136 carbohydrates (e.g., xylose, glucose), the selectivity toward furfural from xylose is a function of the ratio

137 of Brønsted to Lewis acid sites at a constant number of total acid sites.53 The selectivities toward 5-

138 hydroxymethylfurfural (HMF) and levulinic acid from glucose increased with an increase in the

8
Page 9 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

139 concentration of Brønsted acid sites54 and weak Lewis acid sites.55 The Brønsted and Lewis acid

140 concentration also plays an important role in isomerization reaction. Isomerization of glucose to fructose

141 predominantly occurred at lower fractions of Brønsted acid sites.54 Other study also reported that Lewis

142 acid sites promoted isomerization of glucose to fructose while Brønsted acid sites detrimentally affected
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

143 the isomerization.55

Green Chemistry Accepted Manuscript


144

145 2.3. Bifunctional metal–acid sites

146 HDO is often done with bifunctional catalysts which contain both metal and acid sites. In general,

147 such catalysts are made by dispersing metal particles on a solid acid support. Other types of catalysts that

148 have been used include bimetallic catalysts with a reducible metal combined with an oxophilic metal.56

149 For example, Chia et al. found that acid sites exist on the surface of the catalytic system containing both a

150 highly reducible metal (e.g., Rh) and an oxophilic metal (e.g., Re) and hydroxyl groups on Re atoms

151 associated with Rh are acidic (oxygen atoms strongly bind to Re).57 The hydroxyl groups are responsible

152 for the donation of a proton to reactants, which leads to carbenium ion formations. C–O–C bond

153 hydrogenolysis occurs through ring-opening–dehydration reactions catalyzed by the acid sites coupled

154 with hydrogenation catalyzed by the metal sites. To have a high activity for the C–O–C bond

155 hydrogenolysis, the Rh and Re sites must be in close proximity.58, 59 Characterization by XAS60 and

156 transmission electron microscopy (TEM)61 proved that reducible metal (e.g., Pt) particles are covered by

157 ReOx species, resulting in Pt sites directly interacting with ReOx sites. A recent report however claimed

158 there is optimal proximity of the metal–acid sites in bifunctional HDO catalysts (i.e., not “the closer the

159 better”).62

160 Other than the proximity of the reducible and oxophilic metals, a change in the kind of reducible and

161 oxophilic metals could affect catalytic activity for the hydrogenolysis of C–O–C bond. For instance, for

162 the catalyst consisting of two reducible metal species, Pd–Ir–ReOx used for the conversion of furfural into

163 1,5-pentanediol, the Pd species had a role in hydrogenation of furfural to THFA, and the Ir–ReOx species

9
Green Chemistry Page 10 of 71
View Article Online
DOI: 10.1039/C9GC01210A

164 had a role in hydrogenolysis of THFA to 1,5-pentanediol.63 A bifunctional Pt–WOx/TiO2 catalyst (not

165 ReOx-based catalysts that have been widely studied) was active for hydrogenolysis of tetrahydrofuran-

166 dimethanol (THFDM) to 1,6-hexanediol.64 The reducible support (e.g., WOx) could facilitate a synergistic

167 interaction between reducible metal and WOx which has an important role in hydrogenolysis of cyclic
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

168 ethers.

Green Chemistry Accepted Manuscript


169 Understanding how the metal–acid interactions affect HDO reactions is more difficult than

170 monofunctional sites. Surface-level studies using model surfaces65 might help understand how specific

171 arrangements of metal and other atoms on the catalyst surface improve performance for HDO reactions.

172 For the surface-level investigations, the surface composition employed in modeling the HDO reactions

173 should be representative of those present under the reaction conditions. Therefore, catalyst

174 characterization using surface-sensitive operando spectroscopies (e.g., near-ambient pressure X-ray

175 photoelectron spectroscopy)66 would be informative about relating the real surface composition and

176 catalyst performance.

177

178 3. Hydrodeoxygenation of different biomass-derived oxygenates

179 3.1. Hydrodeoxygenation of carbohydrates

180 Biomass-derived carbohydrates have been widely used as a feedstock for producing renewable fuels

181 and chemicals (e.g., alkanes and alcohols) by HDO on bifunctional catalysts.67-69 Sorbitol, a sugar alcohol

182 that can be made via reduction of glucose,70 has been used as a surrogate compound for HDO over a range

183 of catalytic systems (Table 2).71-76 Recently, general reaction pathways for HDO of sorbitol has been

184 proposed.72 In the proposed pathways, sorbitol is converted into 1,4-sorbitan via acid-catalyzed

185 dehydration or two polyoxygenates (e.g., two glycerols) via metal-catalyzed dehydrogenation and

186 subsequent retro-aldol condensation. It should be preferred the former pathway (i.e., dehydration of

187 sorbitol) to synthesize liquid fuels for which maintaining energy density is important. These intermediate

188 compounds are subsequently deoxygenated, combined with different reactions such as dehydration,

10
Page 11 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

189 dehydrogenation, hydrogenation, decarbonylation, and ring-rearrangement, into various C1 to C6

190 monofunctional species and alkanes. The different deoxygenation routes require both metal and acid sites.

191 CO, produced via decarbonylation, is reacted with water to form CO2 (water-gas shift reaction) or

192 transformed into CH4 (methanation). Coupling reactions of C4-C6 monofunctional compounds produce
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

193 oligomers (C9+). The oligomers may be further degraded to coke and humins or hydrodeoxygenated into

Green Chemistry Accepted Manuscript


194 monofunctional species. It was also found that higher hydrogen pressure favors hydrogenation over

195 dehydrogenation.73

196 As an effort to use more realistic carbohydrate feedstocks than sorbitol, it has been tried to directly

197 convert cellulose into alkane (e.g., n-hexane) with a high yield in a one-pot process using a bifunctional

198 catalytic system,77, 78 although earlier studies reported low n-hexane yields from cellulose.79-81 For

199 example, Tomishige and co-workers used a combination of an Ir–ReOx/SiO2 catalyst and HZSM-5 co-

200 catalyst to synthesize n-hexane from cellulose in a biphasic system (water and n-dodecane).78 During the

201 process, 78% of microcrystalline cellulose and 83% of ball-milled cellulose were converted into n-hexane.

202 In the bifunctional system, the Ir–ReOx/SiO2 catalyst expedited hydrogenation reactions, while the

203 HZSM-5 promoted cellulose hydrolysis in hot water and hydrogenolysis of the C−O bond.

204 Sels and co-workers developed a strategic one-pot biphasic process for the production of straight-

205 chain alkanes (e.g., n-hexane) from cellulose with a catalytic system consisting of tungstosilicic acid

206 (TSA) and a bifunctional TSA–Ru/C catalyst.77 The TSA catalyst was responsible for hydrolysis of

207 cellulose and dehydration of reaction intermediates, and the Ru-based bifunctional catalyst was

208 responsible for hydrogenating the intermediates. Through this process, approximately 40% of n-hexane

209 yield was achieved from real biomass feedstock (e.g., raw softwood sawdust). Over the catalytic system,

210 dehydration of glucose (i.e., hydrolyzed from cellulose) to HMF was more kinetically favorable than

211 hydrogenation of glucose to sorbitol. Fructose, HMF, 2,5-dimethyltetrahydrofuran (DMTHF), 2,5-

212 hexanediol, 2,5-hexanedione, 1,2-hexanediol, 1-hexanol, and 2-hexanol were found to be the reaction

213 intermediates. The proposed process reaction pathways from cellulose to n-hexane are described in Figure

11
Green Chemistry Page 12 of 71
View Article Online
DOI: 10.1039/C9GC01210A

214 2. During the production of n-hexane via the HMF route, the ring-opening reaction of 2,5-

215 dihydroxymethylfuran (DHMF) via hydrolysis–dehydration–hydrogenation is crucial for forming 1-

216 hydroxy-2,5-hexanedione, which is further transformed into hexanols via hexanediols. The hexanols are

217 finally converted into alkanes (n-pentane and n-hexane) via HDO.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

218

Green Chemistry Accepted Manuscript


219

220 Figure 2. Reaction pathways for the conversion of cellulose into alkanes over the TSA/TSA–Ru/C
221 catalyst. The bold arrows indicate the most selective pathway to n-hexane from cellulose. Reproduced
222 from Op de Beeck et al.,77 with permission of The Royal Society of Chemistry.
223

12
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Page 13 of 71 Green Chemistry

224 Table 2. HDO of biomass-derived carbohydrates over different bifunctional catalysts. All results are available from recent literature.

Target
Feedstock Reactor type Catalyst Reaction conditions Performance Reference
compounds

Continuous Pt/Phosphated niobic 255 °C; 5.4 MPa; 0.5 h–1 WHSV;
Sorbitol C1-C6 alkanes 63% yield 82
flow reactor ncid (4 wt.% Pt) 25 wt.% aqueous feed solution

Green Chemistry Accepted Manuscript


Ni/HZSM-5 (2 wt.% Ni) 240 °C; 4.0 MPa; 1 h; 6.1 wt.%
Sorbitol C1-C6 alkanes Batch reactor 66.2% yield 83, 84
+ MCM-41 (40 wt.%) aqueous feed solution

Trickle-bed Ru/C (4 wt.% Ru) + 250 °C; 4 MPa; 0.75 h–1 LHSV; 20 80.8% yield (80.7% C5-C6
Sorbitol C1-C6 alkanes 74
reactor H3PO4 (pH = 1.5) wt.% aqueous feed solution alkane selectivity)

Gasoline Continuous 245 °C; 6.21 MPa; 0.73 h–1 WHSV;


Sorbitol Pt/Zr-P (4 wt.% Pt) 70% yield 68
products flow reactor 20 wt.% aqueous feed solution

Gasoline Continuous 245 °C; 6.21 MPa; 2.92 h–1 WHSV;


Sorbitol Pt–ReOx/C (Re/Pt = 1) 44.4% yield 58
products flow reactor 20 wt.% aqueous feed solution

Gasoline Continuous Pt–ReOx/TiO2 (Re/Pt = 245 °C; 6.21 MPa; 2.92 h–1 WHSV;
Sorbitol 24.7% yield 85
products flow reactor 1.5) 20 wt.% aqueous feed solution

Gasoline Continuous 245 °C; 6.21 MPa; 0.16 h–1 WHSV;


Sorbitol Pd–Fe/Zr-P (Fe/Pd = 3) 55.1% yield 21
products flow reactor 20 wt.% aqueous feed solution

Gasoline Continuous 245 °C; 6 MPa; 0.16 h–1 WHSV; 20


Sorbitol W–Pd/Zr-P (W/Pd = 3) 69.0% yield 75
products flow reactor wt.% aqueous feed solution

56.2% yield (40% C5-C6


C5-C6 alcohols Continuous 268 °C; 6.31 MPa; 0.7 h–1 WHSV;
Sorbitol Co/TiO2 (5 wt.% Co) alcohol & 23% 72
& heterocycles flow reactor 20 wt.% aqueous feed solution
heterocycles)

13
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 14 of 71

C4-C6 Continuous Pt–ReOx/Zr-P (Re/Pt = 180 °C; 6.21 MPa; 0.16 h–1 WHSV;
Sorbitol 28.1% yield 76
monoalcohols flow reactor 1) 20 wt.% aqueous feed solution

· Cellulose (pre-treated
in n-hexanol at
400 °C; 6.5 MPa; 12 h; 10 wt.%
Cellulose (commercial) C2-C9 alkanes Batch reactor Pt/HZSM-5 (1 wt.% Pt) 350 °C): 89% yield 80
aqueous feed solution
· Cellulose (no pre-
treatment): 35% yield

Green Chemistry Accepted Manuscript


Cellulose (ball-milled for Pt/H-beta zeolite (1 150 °C; 3 h; 1.2 wt.% aqueous feed
C3-C4 alkanes Batch reactor 48.1% selectivity 81
30 min) wt.% Pt) solution

· Ball-milled cellulose: 190 °C; 6


MPa; 12 h; 5 wt.% aqueous · Ball-milled cellulose
feed solution; water/n- (ball-milled for 2 h):
Ir–ReOx/SiO2 (Re/Ir =2)
Cellulose n-Hexane Batch reactor dodecane = 2.4 (v/v) 83% yield 78
+ HZSM-5 (28.6 wt.%)
· Microcrystalline cellulose: 24 · Microcrystalline
h; others are same as ball- cellulose: 78% yield
milled one
TSA modified Ru/C (2
220 °C; 5.0 MPa; 6 h; 9 wt.%
Cellulose mM TSA added to 5 75% yield (52% n-hexane
C5-C6 alkanes Batch reactor aqueous feed solution; water/n- 77
(microcrystalline) wt.% Ru/C) + TSA (49 yield)
decane = 1 (v/v)
mM)
Carbohydrates from · 1st reaction: Ru/C (5 · 1st reaction: 120 °C; 6.21 MPa;
Gasoline Continuous wt.% Ru) 1.2 h–1 WHSV
hydrolysis of maple 57% yield 79
products flow reactor · 2nd reaction: Pt/Zr-P · 2nd reaction: 245 °C; 6.21 MPa;
wood with oxalic acid
(4 wt.% Pt) 0.73 h–1 WHSV
225

14
Page 15 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

226 3.2. Hydrodeoxygenation of furanic compounds

227 Furfural, a representative furanic compound, is a major compound easily found in bio-oil.86-89 Thus,

228 furfural and its derivatives have been tested as a model feedstock for HDO of bio-oil. Many researches

229 have been conducted to efficiently produce 2,5-dimethylfuran (DMF) from biomass-derived furanic
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

230 compounds such as HMF90-94 since Dumesic group reported an improved DMF production strategy in

Green Chemistry Accepted Manuscript


231 2007,95 as summarized in Table 3. For example, Gyngazova et al. conducted HDO of HMF into DMF on

232 a Ni catalyst supported on carbon.94 A 70% yield of 2,5-dimethylfuran was achieved with 15 wt.% metal

233 loading at 180 °C and 4.5 MPa for 1 h. The authors also performed kinetic analysis of the HDO. The

234 results revealed that hydrogenation of HMF to 2,5-bishydroxymethylfuran (i.e., hydrogenation of C=O

235 group) is the rate determining step (estimated activation energy of 67 kJ mol–1) during HDO.

236 Hydrogenation of 2,5-dimethylfuran to 2,5-dimethyltetrahydrofuran was not kinetically favored, leading

237 to a high selectivity toward 2,5-dimethylfuran from furfural.

238 Catalytic transfer hydrogenation (CTH) reaction has been suggested for DMF production via HDO of

239 HMF (Table 3), which does not require molecular H2 as the hydrogen source.96-99 For the production of

240 DMF from HMF via CTH, instead of using pressurized H2 gas liquid-phase hydrogen donors (e.g.,

241 alcohols and formic acid) have been used. In 2010, Thananatthanachon and Rauchfuss reported the

242 transformation of HMF to DMF without using gaseous H2 with a 5% Pd/C catalyst in the presence of

243 formic acid and sulfuric acid (H2SO4).96 Later on, Vlachos and co-workers conducted CTH to synthesize

244 DMF from HMF using a 5 wt.% Ru/C catalyst and isopropanol as a hydrogen donor.97 The CTH process

245 gave rise to 81% DMF yield at 190 °C under nitrogen environment (2 MPa) for 6 h. The authors found

246 that the dominant phase of the Ru/C catalyst is RuO2 that is reduced to its metallic phase during the CTH

247 reaction by reacting with hydrogen generated from isopropanol.98 The reduction of RuO2 on the catalyst

248 surface was the main reason of the catalyst deactivation. It was also revealed that RuO2 catalyzes the

249 Meerwein–Ponndorf–Verley reduction of HMF to 2,5-bis(hydroxymethyl)furan and etherification of 2,5-

250 bis(hydroxymethyl)furan. The metallic Ru catalyzes C–O bond cleavage of 2,5-bis(hydroxymethyl)furan,

15
Green Chemistry Page 16 of 71
View Article Online
DOI: 10.1039/C9GC01210A

251 leading to the formation of DMF. The studies showed that Ru/C catalyst could act as a bifunctional

252 catalyst in the CTH conditions.

253 Other than the production of DMF, there have been efforts to produce renewable diesel and jet fuel

254 ranged alkanes from various furan derivatives (Table 3).100-108 For example, Zhang group have developed
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

255 catalytic processes for producing renewable diesel and jet fuel ranged alkanes from various furan

Green Chemistry Accepted Manuscript


256 derivatives.100-104 In their recent contribution, a FeOx-modified Pd/SiO2 catalyst (Pd–FeOx/SiO2) was

257 synthesized and used for HDO of three different furan derivatives (1-(furan-2-yl)-5-methylhex-1-en-3-one,

258 1-(furan-2-yl)hex-1-en-3-one, and 5,5’-(butane-1,1-diyl)bis(2-methylfuran)), which can be produced via

259 aldol condensation or alkylation of furfural into C9-C14 alkanes.104 HDO was performed at atmospheric

260 pressure with no solvent. With the bifunctional bimetallic catalyst, 100% conversion of the feedstocks

261 was achieved with high selectivities toward the long-chain alkanes (87 and 94%) at 350 °C and 300 °C,

262 respectively. Through a variety of characterizations, the authors showed that the addition of Fe species

263 hinders decarbonylation and promotes hydrogenation of the C=O group, thereby restraining short chain

264 alkane formation.

265 The ring-opening of tetrahydrofuran (THF) is more difficult than furfural or HMF-derived

266 condensation products. Due to its recalcitrance, harsh reaction conditions are required for the THF ring-

267 opening reaction. However, the harsh reaction conditions often lead to the C–C bond cleavage. Many

268 developments have been launched around this question. For instance, Xue et al. incorporated Nb with

269 SBA-15 (Nb–SBA-15) using a one-pot co-assembly method as an effort to make active catalyst for THF

270 ring-opening reaction.109 They found that the Nb–SBA-15 has two different Nb sites: Nb2O5 (octahedral)

271 and NbO4 (tetrahedral). As the content of Nb increased, the Nb2O5 particle size increased but the amount

272 of low-coordination Nb species decreased. A Pd catalyst supported on the Nb–SBA-15 was tested for

273 HDO of DMTHF to n-hexane, indicating that the HDO activity is highly associated with the state of the

274 Nb species. The catalyst with a low Nb coordination number was preferred to hydrodeoxygenate DMTHF

275 to n-hexane. However, the yield of n-hexane over the Pd/Nb–SBA-15 catalyst was low (~10%), so more

16
Page 17 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

276 researches need to be carried out for further developments of designing more active catalysts for HDO of

277 biomass-derived oxygenates containing THF ring.

278
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

Green Chemistry Accepted Manuscript

17
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 18 of 71

279 Table 3. HDO of biomass-derived furanic compounds over different bifunctional catalysts. All results are available from recent literature.

Target
Feedstock Reactor type Catalyst Reaction conditions Performance Reference
compounds

150 °C; 0.8 MPa; 8 h; 0.5 g feed


HMF DMF Batch reactor Pd–Zn/C (Zn/Pd = 3) 85% yield 90
in 15 mL THF

Green Chemistry Accepted Manuscript


130 °C; 0.7 MPa; 24 h; 0.25 g
HMF DMF Batch reactor Ru/Co3O4 (Ru/Co = 5) 93.4% yield 91
feed in 10 mL THF

Ni-W2C/C (Ni/W2C = 180 °C; 4 MPa; 3 h; 1 mmol feed


HMF DMF Batch reactor 96% yield 92
0.23) in 12 mL THF

Continuous 130 °C; 1 MPa; 3.3 h−1 WHSV;


HMF DMF Ni–Co/C (Ni/Co = 0.1) >90% yield 93
flow reactor 71 h TOS

180 °C; 4.5 MPa; 1 h; 40 mg


HMF DMF Batch reactor Ni/C (15 wt.% Ni) ~70% yield 94
mL–1 feed concentration in THF

190 °C; 2 MPa (N2); 6 h; 0.24 g


HMF DMF Batch reactor Ru/C (5 wt.% Ru) 81% yield 97
HMF in 24 mL 2-propanol

210 °C; 24 h; 0.25 g HMF in 10


HMF DMF Batch reactor Ni–Co/C (Ni/Co = 0.1) mL THF with 20 mmol formic 90% yield 99
acid

Continuous 450 °C; 0.5 MPa; 0.78 s contact


2-Methylfuran n-Pentane Ni–P/SiO2 (P/Ni = 0.1) ~69% yield 110
flow reactor time; 32.6 mol% feed in toluene

Furan derivative (4-(2-furyl)- Pd/Nb2O5/SiO2 (4 wt.% 170 °C; 2.5 MPa; 24 h; 0.2 g
Octane Batch reactor 95.5% yield 105
3-buten-2-one) Pd and 10 wt.% Nb2O5) feed in 6.46 g cyclohexane

C9-C13 Continuous 370 °C; 0.04 mL min–1 liquid


Furan derivatives Pd/C (5 wt.% Pd) 79.5% yield 101
alkanes flow reactor flow rate

18
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Page 19 of 71 Green Chemistry

C9-C15 Continuous Ni–Mo2C/SiO2 (Mo/Ni 300 °C; 6 MPa; 1.3 h–1 WHSV;
Furan derivatives 77% yield 102
alkanes flow reactor = 2.7) solvent-free

C9-C14 Continuous
Furan derivatives Ni/HZSM-5 (4 wt.% Ni) 260 °C; 6 MPa; 1.3 h–1 WHSV 90% yield 103
alkanes flow reactor

· 1-(Furan-2-yl)hex-1-en-3- · 1-(Furan-2-yl)hex-1-

Green Chemistry Accepted Manuscript


one: 350 °C; 0.1 MPa; en-3-one: 94% yield
C9-C14 Continuous Pt–FeOx/SiO2 (Pd/Fe = solvent-free · 5,5’-(Butane-1,1-
Furan derivatives 104
alkanes flow reactor 2) · 5,5’-(Butane-1,1-diyl)bis(2- diyl)bis(2-
methylfuran): 300 °C; 0.1 methylfuran): 87%
MPa; solvent-free yield
300 °C; 8.27 MPa; 0.02 cm3 min–
Hydrocycloadded products C12-C13 Continuous
Pt/SiO2–Al2O3 (4% Pt) 1 liquid flow (the feed in 60.1% selectivity 111
from furfural alkanes flow reactor
methanol)

Pd/C (10 wt.% Pd) and 225 °C; 5.07 MPa; 24 h; 1.23
Furylmethane oxygenates C14-C15 97% yield (93% C14 and C15
Batch reactor Hf(OTf)4 (Pd/Hf(OTf)4 mmol feed in 25 mL 106
(C15) alkanes alkane selectivity)
= 2) cyclohexane
280

19
Green Chemistry Page 20 of 71
View Article Online
DOI: 10.1039/C9GC01210A

281 3.3. Hydrodeoxygenation of aromatic compounds

282 Aromatic compounds have been widely used as a model feedstock for HDO of bio-oil, mainly derived

283 from lignin.112-116 Among different aromatic compounds, m-cresol, anisole, eugenol, and guaiacol have

284 been most widely used as the model feedstocks for HDO of bio-oil (Table 4).
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

285 As an example of m-cresol HDO, Shetty et al. compared site time yields (molar flow rate of toluene

Green Chemistry Accepted Manuscript


286 formation per moles of redox active Mo species) of 10 wt.% MoO3 catalysts supported on different oxides

287 such as TiO2, ZrO2, Al2O3, and SiO2.117 The site time yields increased with a decrease in electronegativity

288 of the oxide support cation (Si4+ > Al3+ > Ti4+ > Zr4+) for the HDO of m-cresol to toluene, increasing

289 electron density of Mo–O-support bridging oxygen. The increase in the electron density promotes redox

290 reaction, and the Mo–O-support bridging O atoms may play a key role in H abstraction that is required for

291 initiating the HDO reaction. For the HDO of m-cresol to toluene over MoOx-based catalysts, the

292 submonolayer of Mo species leads to the formation of polymeric molybdate regardless of the support

293 type.118 The poly-molybdate reacts with H2, creating MoOx species with the release of water. m-Cresol

294 can be adsorbed on the oxygen vacancy via its O atom, which favors C–O bond scission. Heterolytic

295 dissociation then occurs by H2 activation. A hydride species is added to the carbon that has the hydroxyl

296 group. The hydride addition results in C–O bond cleavage, forming toluene. The release of water again

297 recovers the oxygen vacancy.

298 A comparative study for HDO of anisole over USY zeolite-supported Pd, Ru, and Ni catalysts was

299 conducted.119 Initial TOF for the HDO reactions followed the order of Pd/USY zeolite > Ru/USY zeolite

300 > Ni/USY zeolite. The highest anisole conversion (75 mol.%) and cyclohexane selectivity (70%) was

301 achieved with the Ni/USY zeolite catalyst at 200 °C.

302 The HDO of more realistic aromatic feedstocks (e.g., lignin) has gained increasing attention.120-126 For

303 instance, Wang and co-workers have demonstrated some key advances for the HDO of lignin into arenes

304 using catalytic systems consisting of Ru and Nb2O5.125, 126 Catalytic transfer HDO of real lignin (from

305 birch) was conducted over a Ru/Nb2O5–SiO2 catalyst in the presence of isopropanol (hydrogen donor).125

20
Page 21 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

306 The process gave rise to 19.8 wt.% yield of total aromatic hydrocarbons including ethylbenzene,

307 propylbenzene, and toluene. Other Nb-based catalytic systems were used to decrease oxygen content of

308 lignin via HDO.127, 128 For the HDO of lignin over the catalyst, Caliphatic–O ether bond cleavage formed

309 phenolic intermediates. The phenolic compounds underwent hydrogenolysis to aromatic hydrocarbons via
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

310 selective Caromatic–OH bond cleavage more favored at higher temperatures. The conversion of lignin

Green Chemistry Accepted Manuscript


311 monomers was almost complete (71 wt.% selectivity towards aromatic hydrocarbons), ascribed to

312 Caromatic–OH bond dissociation energy reduction along with strong adsorption of the phenolics on the

313 Nb2O5.

314

21
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 22 of 71

315 Table 4. HDO of biomass-derived aromatic compounds over different bifunctional catalysts. All results are available from recent literature.

Feedstock Target compounds Reactor type Catalyst Reaction conditions Performance Reference

· Pd/USY zeolite: 57% selectivity at


100 mol.% conversion
· Ru/USY zeolite: 63% selectivity at

Green Chemistry Accepted Manuscript


Anisole Cyclohexane 119
· Pd/USY zeolite (5 100 mol.% conversion
wt.% Pd) 200 °C; 5.17 MPa; 140 · Ni/USY zeolite: 70% selectivity at 75
· Ru/USY zeolite (5 min; 53.9 mL of 4 mol.% conversion
Batch reactor
wt.% Ru) wt.% feedstock in · Pd/USY zeolite: 90% selectivity at
· Ni/USY zeolite (5 decane 85 mol.% conversion
wt.% Ni) · Ru/USY zeolite: 85% selectivity at
4-Ethylphenol Ethylcyclohexane 119
77 mol.% conversion
· Ni/USY zeolite: 92% selectivity at 55
mol.% conversion
230 °C; 5.5 MPa; 30
6.2 ×10–3 mol L–1 (100% phloroglucinol
Phloroglucinol Cyclohexanol Batch reactor Pt/C (5 wt.% Pt) min; 0.2 wt.% aqueous 129
conversion)
feed solution

170 °C; 2.5 MPa; 24


Pd/Nb2O5/SiO2 (4 wt.%
Diphenyl ether Cyclohexane Batch reactor h; 0.2 g feed in 6.46 g 98.2% yield 105
Pd and 10 wt.% Nb2O5)
cyclohexane

240 °C; 2 MPa; 4 h;


Pd/C (5 wt.% Pd) +
Eugenol Hydrocarbons Batch reactor 1.9 vol.% aqueous 36.5% yield 130
HZSM-5 (Si/Al = 25)
feed solution

240 °C; 5 MPa; 4 h;


Pd/C (5 wt.% Pd) +
Eugenol Hydrocarbons Batch reactor 0.0025 mol feed in 20 63.5% yield 131
HZSM-5 (Si/Al = 12.5)
mL water

250 °C; 4 MPa; 1 h;


Ru/SiO2–Al2O3 (5 wt.%
Guaiacol Cyclohexane Batch reactor 7.5 wt.% feedstock in 60% yield 132
Ru)
n-decane

400 °C; 4 MPa;


Continuous
Guaiacol Benzene Mo/C (10 wt.% Mo) 14.9 h–1 WHSV; 83.5% yield 133
flow reactor
H2/feed = 20

22
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Page 23 of 71 Green Chemistry

300 °C; 3 MPa; 3 h; 3


Guaiacol Phenol Batch reactor MoO2/C (20 wt.% Mo) g feed in 80 mL 70.6% yield 134

tetralin

Continuous Pt–Co/Al2O3 (Co/Pt = 260 °C; 0.05 MPa; 63


m-Cresol Methylcyclohexane 35% yield 135
flow reactor 2.9) g-cat h mol–1

Green Chemistry Accepted Manuscript


Continuous Pt/Fluorine-doped Al2O3 260 °C; 0.05 MPa; 63
m-Cresol Toluene 51.8% yield 136
flow reactor (1.7 wt.% Pt) g-cat h mol–1

200 °C; 2 MPa; 2 h;


Phenol 0.005 mol feed in 20 99.9% yield
mL water

170 °C; 2 MPa; 2 h;


Pd/C (5 wt.% Pd) +
Anisole Cyclohexane Batch reactor 0.005 mol feed in 20 84.3% yield 137
HZSM-5
mL water

240 °C; 2 MPa; 2 h;


Guaiacol 0.005 mol feed in 20 4.6% yield
mL water

300 °C; 0.1 MPa;


Continuous
Phenol Benzene Pd/Nb2O5 (2 wt.% Pd) 18.9 h–1 WHSV; 43% yield 138
flow reactor
H2/phenol = 60

Ru/Al2O3 (5 wt.% Ru) + 280 °C; 4 MPa; 4 h;


Corn stover C12-C18 cyclic 70.4 wt.% production distribution at
Batch reactor zeolite Y (form of H+ 100 mg feed in 30 mL 139
lignin hydrocarbons 92.7% lignin conversion
zeolite Y) water
316

23
Green Chemistry Page 24 of 71
View Article Online
DOI: 10.1039/C9GC01210A

317 3.4. Hydrodeoxygenation of bio-oil

318 As well as the HDO studies of model compounds discussed above, efforts have been made to upgrade

319 real bio-oil via HDO to improve its fuel quality. For HDO of bio-oil, multi-stage processes are often

320 used.140-144 In most cases, the last stage of the HDO system has two zones, into which different catalysts
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

321 are loaded.145-147 In the two-stage processes, the first low-temperature stage is used to saturate reactive

Green Chemistry Accepted Manuscript


322 functional groups, which stabilizes the bio-oil for subsequent high-temperature deoxygenation. Figure 3

323 shows an example of the two-zone reactor for HDO of bio-oil.148 Olarte et al. proved that deactivation and

324 low yields occur in one-stage HDO of bio-oil.145 For example, plugging was observed during single-stage

325 HDO without a pretreatment step for upgrading bio-oil after 48 h TOS operation. Ha and co-workers also

326 found that coking causes plugging for a one-stage HDO process with a Ru catalyst supported on

327 tungstate–zirconia (Ru/W–Zr).149

328

329

330 Figure 3. Scheme of a two-stage reactor for HDO of bio-oil. Reprinted from Sanna et al.,148 Copyright
331 (2014), with permission from Elsevier.

24
Page 25 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

332

333 A continuous HDO of bio-oil (liquid phase pyrolysis oil) in a plug flow reactor has very recently been

334 introduced.150 HDO was performed with a sulfided Co–Mo/Al2O3 catalyst at 400 °C, 12 MPa H2, and high

335 LHSVs from 0.5 to 3 h–1. The upgraded bio-oil evolved from the HDO process was separated from the
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

336 aqueous stream. The process produced fuel with an oxygen content from 0 to 1.2 wt.% with a H/C ratio

Green Chemistry Accepted Manuscript


337 between 1.7 and 2, consisting mostly of gasoline and diesel range alkanes. It was also claimed that coking

338 is suppressed in HDO processes with a high-water content, indicating that high LHSVs are a feasible

339 parameter for HDO of liquid-phase pyrolysis oil. Meyer et al. also mentioned that an increase in LHSV

340 can reduce the minimum fuel selling price and size of HDO reactors.151 Table 5 is a summary of recently

341 developed HDO processes for upgrading bio-oil. Historical developments in HDO of bio-oil prior to 2007

342 can be found in a review by Elliott.152

343 There are issues when collecting data on HDO of bio-oil in a batch reactor. Selectivities and yields

344 can be lower than those obtained in a continuous flow reactor.153 Catalyst stability is also difficult to study

345 in a batch reactor.153 Therefore, the type of reactor must be carefully chosen for upgrading of bio-oil

346 through HDO. Furthermore, it would be highly informative that catalyst stability is tested under HDO

347 conditions for longer TOS operation using a bench-scale continuous flow reactor. Characterization of

348 catalytic performance can become more convenient when using a flow reactor because it allows to easily

349 vary contact time and space velocity with different solvents at a wide range of temperatures.153

350

25
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 26 of 71

351 Table 5. Recently developed multi-stage HDO processes for bio-oil upgrading.
Entry Feedstock Reactor type Stage Catalyst Reaction conditions C/H/O contents in product Reference
340 °C; 13.8 MPa; 0.25
1 Pd/C 75.5/9.4/12.3
Fast pyrolysis oil from Continuous h–1 LHSV; 102 h TOS
1 141
mixed wood flow reactor Sulfided form of conventional 405 °C; 10.3 MPa; 0.2 h–
2 86.6/12.9/0.4
hydrocracking catalyst 1 LHSV; 102 h TOS

Liquid phase pyrolysis 1 Raney Ni 250 °C; 8.5 MPa; 2 h 62.5/8.3/28.7 (wt.%)

Green Chemistry Accepted Manuscript


2 Batch reactor 154
oil 2 Raney Ni 400 °C; 15 MPa; 2 h 85.5/12.1/1.9 (wt.%)

1 Sulfided Ru/C (7.8 wt.% Ru) 220 °C; 10.7 MPa -


Fast pyrolysis oil from Continuous
3 Sulfided Co–Mo/Al2O3
142
whole-tree pine flow reactor 2 400 °C; 10.7 MPa 87.5/12.7/1.0 (wt.%)
(MoO3/CoO = 3.3)
Dewatered liquid phase Continuous Sulfided Co–Mo/Al2O3 400 °C; 12.1 MPa; 0.2 h–
4 1 85.6/13.2/1.2 (wt.%) 155
pyrolysis oil flow reactor (MoO3/CoO = 4) 1 LHSV; 62 h TOS

Continuous 230-370 °C; 14 MPa; 0.6


5 Pyrolysis oil 1 Sulfided Ni–Mo/Al2O3 86.2/13.0/<0.8 (%) 156
flow reactor h–1 LHSV
140 °C; 8.4 MPa; 0.5 h–1
1 Ru/C (7.8 wt.% Ru) -
LHSV
Continuous
6 Pyrolysis oil from pine 145
flow reactor Sulfided Ru/C (7.8 wt.% Ru) + 170 °C + 406 °C; 0.1 h–1
2 55.3/6.3/38.1 (wt.%)
sulfided commercial HDO/HC LHSV; 130 h TOS

75 °C; 20 MPa; 0.3 h–1


1 -
WHSV
180 °C; 20 MPa; 0.3 h–1
Fast pyrolysis oil from Continuous 2 Ni–Cu/SiO2–ZrO2 (NiO/CuO -
7 WHSV 147
pine wood flow reactor = 9.2)
180 °C; 20 MPa; 0.3 h–1 · Organic product oil yield: 86
3 wt.%
WHSV; 40 h TOS
· Water content: 15.9 wt.%
140 °C; 8.2 MPa; 0.5 h–1
1 Ru/C (7.8 wt.% Ru) 56.9/6.7/36.3 (wt.%)
LHSV
Sulfided Ru/C + sulfided 170-190 °C + 400 °C;
Fast pyrolysis oil from Continuous 2-1 commercial hydrotreating 12.3 MPa; 0.22 h–1 84.9/13.3/1.8 (wt.%)
8 146
oak flow reactor catalyst LHSV
170-180 °C + 395-
Ru/C (7.8 wt.% Ru) + Pd/C
2-2 405 °C; 13.5 MPa; 0.27 81.9/12.3/5.9 (wt.%)
(2.5 wt.% Pd)
h–1 LHSV

26
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Page 27 of 71 Green Chemistry

130 °C; 13.8 MPa; 55 h


1 Ru/C (6.7% Ru dispersion) -
TOS
Pyrolysis oil from white Continuous
9 Gasoline, diesel, and jet fuel 143
oak wood flow reactor 400 °C; 13.8 MPa; 55 h
2 Pt/Zr-P (4 wt.% Pt) range compounds: 69 wt.%
TOS
carbon yield
125 °C; 5.17-10 MPa;
1 Ru/C (5 wt.% Ru) -
Aqueous fraction of 1.5-3 h–1 WHSV
Continuous
10 pyrolysis oil from dry 220-275 °C; 5.17-10 148
flow reactor Gasoline products and diols: up
pine wood 2 Pt/C (5 wt.% Pt) MPa; 3 h–1 WHSV; 80 h

Green Chemistry Accepted Manuscript


to 45% carbon yield
TOS
100-190 °C; 10 MPa; 0.4
1 Pd/C (5 wt.% Pd) -
Pyrolysis oil from pine Continuous h–1 LHSV
11 144
sawdust flow reactor 300-390 °C; 10 MPa; 0.4
2 Ru/W–Zr (3 wt.% Ru) 83.9/12.7/0.07 (wt.%)
h–1 WHSV
352

27
Green Chemistry Page 28 of 71
View Article Online
DOI: 10.1039/C9GC01210A

353 3.5. Hydrodeoxygenation of bio-lipid

354 Conversion of lipid mixtures, most energy dense molecules, has become a key route to economical

355 production of drop-in fuels such as diesel and kerosene through transesterification, cracking, and

356 deoxygenation.157, 158 Fatty acids generally have varying carbon distributions and hydrocarbon backbone
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

357 depending on the bio-lipid feedstock (i.e., vegetable oils, non-edible oils, animal fat, and microalgal

Green Chemistry Accepted Manuscript


358 oils).159, 160 Transesterification of the oils (triglycerides) produces oxygenated hydrocarbons (fatty acid

359 methyl esters), but has limitations such as high cloud points, pour points, and engine compatibility.161

360 Therefore, HDO of the oxygenated compounds is of importance to meet the European diesel fuel standard

361 (EN590).162 The HDO of such oils has advantages in producing the target linear hydrocarbons.12 In

362 common with other HDO processes discussed above, the lipid mixtures ideally undergo a sequence of

363 hydrogenolysis, hydrogenation, and dehydration, where the kind of catalytic sites and balance of metal

364 and acid sites is important to produce their corresponding saturated hydrocarbons (also refer Section 2).163

365 For instance, metallic sites catalyzed C–C bond cleavage from stearic acid to C17 hydrocarbons at 300 °C,

366 which decreased as following order: Pd > Pt > Ni > Rh > Ir > Ru > Os.164 By optimizing surface acidity

367 on the catalyst, the catalyst promotes the activation of C-O bond, promote the isomerization reaction

368 while inhibit cracking reaction for the HDO of lipid mixtures.165

369 Table 6 summarizes the results of HDO of bio-lipid and fatty acids over various heterogeneous

370 catalytic systems consisting of metal and solid acid. As an example, Lercher and co-workers investigated

371 the effect of bifunctional catalysts on the stearic acid conversion and found that 5wt% Ni/H-BEA

372 (Si/Al=75) led to a 96% conversion with selectivities of 82% for the C18 and of 18% for the C17 alkanes,

373 almost eliminating the cracking reaction.166 Modulation of metal-support interaction is also important for

374 increasing catalytic activity, especially for the active sites having high oxygen affinity (i.e., Ru). For

375 instance, nitrogen modification in carbon nanotube reduces the adsorption property of the carboxylic acids

376 on Ru and increased C17 alkane selectivity for conversion of stearic acid.167

28
Page 29 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

377 The nature of metallic sites can be influenced by the surface acid sites. For instance, Qian and co-

378 workers have studied the detail characterizations of core@shell structural ZSM-22@SiO2 zeolite

379 supported Pt catalysts based on FT-IR of CO adsorption combined with TEM and found that the Pt active

380 crystal plane changed from (111) to (100) with decreasing surface acid sites.168 Xia et al. showed the
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

381 synergistic effect of hydrogen activity from Pd and strong Lewis acidity on the reduced NbOPO4, in

Green Chemistry Accepted Manuscript


382 which glyceryl trioctanoate is converted to diesel-range alkanes while suppressing the C–C bond cleave

383 reaction with 96.4% C7-C8 alkane yield from glyceryl trioctanoate at 160 °C and 30 MPa.169

384

29
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 30 of 71

385 Table 6. HDO of bio-lipid and fatty acids over different bifunctional catalysts. All results are available from recent literature.

Target Reactor
Feedstock Catalyst Reaction conditions Performance Reference
compounds type

Ni/HBEA
Batch 166
Stearic acid C18 alkanes (5 wt.% Ni, 260 °C; 4.0 MPa; 8 h; 1.0 wt.% feedstock in dodecane solution 78.7% yield
reactor
Si/Al=75)

Green Chemistry Accepted Manuscript


Pd/Al-SBA-
Sunflower C15-C18 Batch 15 (5 wt.% 170
250 °C; 2.0 MPa; 5 h; 1.6 wt.% feedstock in dodecane solution 72.9% yield
oil hydrocarbons reactor Pd,
Si/Al=300)

Glyceryl C7-C8 Batch Pd/NbOPO4 160 °C; 3.0 MPa; 48 h; 2.0 wt.% feedstock in cyclohexane 169
96.4% yield
trioctanoate alkanes reactor (3 wt.% Pd) solution

Pt-Re/H-
ZSM-5 (5
C18 Batch 270 °C; 2.0 MPa; 12 h; 10 wt.% feedstock in cyclohexane 171
Jatropha oil wt.% Pt, 56% yield
hydrocarbons reactor solution
Re/Al=0.8,
Si/Al=11.5)
Pt/H-ZSM-
Trickle-bed 357 °C; 4.0 MPa; LHSV = 1h-1; pure soybean oil 172
Soybean oil alkanes 22 (1wt.% 100% yield
reactor (palmitic/stearic/oleic/linoleic/linolenic=10.2:3.7:22.8:53.7:8.6)
Pt)

Ru/NCNT-
C17-18 Batch 167
Stearic acid 800 (5wt.% 225 °C; 4.0 MPa; 4 h; 1.5 wt.% feedstock in dodecane solution 89.5% C17, 8.5% C18
alkanes reactor
Ru)

Microalgae C17-18 Batch Ni/HBEA 173


260 °C; 4.0 MPa; 4 h; 1.0 wt.% stearic acid in dodecane solution 34.0% C17, 61.0% C18
oil alkanes reactor (5wt.% Ni)

Pt-Re/C
Batch (5.2wt.% 300 °C; 0.35 MPa; 9 h; 19 wt.% oleic acid, 4.7 wt.% glycerol in 174
Oleic acid C17 alkane 37% yield
reactor Pt, 4.2 wt.% solution
Re)
Cu-
Ni/Al2O3
Batch 175
Oleic acid C17 alkane (20wt.% 330 °C; 1h; 8.9 wt.% stearic acid, 1.8 wt.% methanol in solution 92.7%
reactor
Cu, 40
wt.% Ni)

30
Page 31 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

387 4. Catalyst stability for hydrodeoxygenation of biomass-derived oxygenates

388 4.1. Catalyst deactivation

389 As HDO of biomass-derived oxygenates is conducted under harsh reaction conditions (e.g., high

390 temperature and pressure), heterogeneous HDO catalysts generally suffer from irreversible deactivation.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

391 Deactivation often ensures that the catalyst is not reusable, potentially increasing the operational costs of

Green Chemistry Accepted Manuscript


392 the HDO process. Therefore, the stability of HDO catalysts must be carefully considered. The different

393 heterogeneous catalytic sites (i.e., metal and acid sites) are deactivated differently; hence, different

394 methods have been applied to stabilize each type of active site, as discussed in Section 4.2.

395 Sintering and leaching of metal particles are major problems for liquid-phase HDO reactions. For

396 example, Co particles dispersed on TiO2 or Al2O3 leached during aqueous-phase hydrogenation of

397 furfuryl alcohol at 140 °C and 2.34 MPa H2 (~10% cobalt leached during 35 h).176 An Al2O3-supported

398 Cu nanoparticles were sintered (i.e., the particle size was increased from 3 to 5 nm) after liquid-phase

399 hydrogenation of furfural at 140 °C and 2.2 MPa H2.177 Both leaching and sintering led to the loss of

400 active sites, and the leached or sintered catalyst could not be regenerable. Also, leaching contaminates the

401 product stream with the leached metal.

402 Sintered metal particles can lower catalytic activity for HDO and cause more coking on the catalyst.

403 The stability of bifunctional Ni/ZrO2 catalysts (one calcined then reduced and one only reduced) for HDO

404 of a mixture of guaiacol and n-octanol to cyclohexane and heptane, respectively, at 250 °C and 10 MPa

405 was investigated for 100-h continuous operation using a trickle bed reactor.178 Calcination of the Ni/ZrO2

406 catalyst at 400 °C increased its average Ni particle size from 9 to 18 nm. Both activity and stability of the

407 Ni/ZrO2 catalyst were related to Ni particle size. Larger Ni particles led to a lower activity for HDO and

408 caused more coking on the catalyst. Leaching of Ni particles was negligible for the catalytic system.

409 Hydrothermal stability of solid acid support is important when making HDO catalyst (Figure 1). For

410 instance, γ-Al2O3 support experiences structural changes (i.e., boehmite formation) in the presence of

411 liquid water at 200 ºC under autogenic pressure.179 A high level of coke formation was also observed in

31
Green Chemistry Page 32 of 71
View Article Online
DOI: 10.1039/C9GC01210A

412 the support; thus, carbon supports are considered a substitute for alumina.180 Water leads not only to loss

413 of the support surface area but also to phase changes of the support to more crystalline material. As an

414 example, phase transformations were reported for the γ-Al2O3 support into hydrated boehmite under hot

415 water conditions (>120 °C), significantly lowering its acidity and surface area.179, 181-183 For microporous
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

416 solid acid (e.g., zeolite), deposition of insoluble oligomeric species could form coke in the micropores,

Green Chemistry Accepted Manuscript


417 leading to a rapid deactivation.184 In addition to coking, structural degradation caused by destannation and

418 desilication rapidly deactivated the zeolite catalyst.184

419 Impurities in biomass feedstocks could deactivate HDO catalyst. The effects of impurities in the

420 feedstocks such as sulfur (S), chlorine (Cl), and potassium (K) on the stability of Ni particles were also

421 evaluated.178 Among the impurities, S deactivated the catalyst fastest, probably by forming a NiS phase,

422 leading to loss of the metal catalytic sites. Deactivation by Cl could be regenerable because Cl adsorbs on

423 the surface Ni sites, forming an equilibrium layer on the surface that can be readily removed without co-

424 feeding of Cl. Nevertheless, Cl potentially causes irreversible deactivation such as sintering of Ni particles.

425 The addition of K to the feed stream notably decreased deoxygenation activity while not affecting

426 hydrogenation activity. This may be because low-coordinate sites associated with deoxygenation are

427 blocked by K.

428

429 4.2. Methods for catalyst stabilization

430 A thin film can be deposited on a substrate via discrete pulsing of vapor phase chemical precursors by

431 a method called atomic layer deposition (ALD).185 ALD of oxide layers on the surface of supported metal

432 catalysts was employed to stabilize metal particles dispersed on the oxide support for hydrogenation of

433 biomass-derived oxygenates.176, 177, 186-189 For instance, O’Neill et al. coated Al2O3 films on a Cu/Al2O3

434 catalyst (Al2O3/Cu/A12O3) using 45 cycles of ALD.177 The Al2O3/Cu/A12O3 catalyst (after calcination at

435 700 °C) was fully regenerable after a regeneration treatment, indicating that the ALD catalyst does not

436 undergo leaching or sintering of Cu nanoparticles. The ALD overcoat on the catalyst surface was thought

32
Page 33 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

437 to selectively block under-coordinated metal atoms existing in the form of surface defects and step edges.

438 Considering that leaching and sintering occur at under-coordinated metal atoms,190 the selective

439 interaction between the ALD Al2O3 overcoat and under-coordinated sites prevents metal leaching and

440 sintering.176, 177 Although ALD Al2O3 overcoat effectively prevented leaching and sintering of Cu
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

441 nanoparticles in liquid phase, it failed to stabilized Co nanoparticles because the high-temperature

Green Chemistry Accepted Manuscript


442 calcination generated cobalt aluminate that is not reducible at temperatures below 800 °C.176 Lee et al.

443 successfully stabilized a TiO2 supported Co catalyst by ALD of TiO2 layers (30 cycles) on the Co/TiO2

444 catalyst surface.176 Reducibility of the catalyst at 600 °C was proven by hydrogen TPR (temperature-

445 programmed reduction). It was revealed that the non-ALD Co/TiO2 catalyst lost approximately 10% of

446 cobalt via leaching during 35-h hydrogenation reaction in aqueous phase, but the ALD catalyst did not

447 lose cobalt during hydrogenation. TEM results showed that no sintering was observed for the ALD

448 TiO2/Co/TiO2 catalyst.

449 Marshall and co-workers compared ALD TiO2 and Al2O3 overcoats on Cu chromite for gaseous phase

450 hydrogenation of furfural.188 Through in situ TPR/TPO (temperature programmed oxidation) and X-ray

451 absorption fine structure (XAFS) characterization techniques, the authors concluded that TiO2 layers by

452 ALD slightly modify the redox properties of Cu via electronic interactions between Cu particles and the

453 TiO2 layers. For the ALD TiO2/Cu chromite catalyst, copper oxide (Cu2+) existed at the interface between

454 TiO2 and Cu particles, while for the ALD Al2O3/Cu chromite catalyst, copper aluminate existed. Given

455 that copper aluminate is a lot harder to be reduced than copper oxide, the ALD TiO2/Cu chromite catalyst

456 was more active (~80% furfural conversion) than the ALD Al2O3/Cu chromite catalyst (~12% furfural

457 conversion) for 700 min time on stream (TOS). The ALD TiO2 layer (20 cycles) stabilized Cu chromite

458 catalyst (a 15% activity loss after 10-h run while >80% activity loss for pure Cu chromite).

459 ALD of Nb2O5 layers (10, 19, or 30 cycles) on a synthesized mesoporous SiO2 resulted in highly

460 ordered Nb2O5 catalysts with high hydrothermal stability.191 The ALD-synthesized material did not

461 experience a change in pore structure or loss of surface area and pore size after treatment in water at

33
Green Chemistry Page 34 of 71
View Article Online
DOI: 10.1039/C9GC01210A

462 200 °C and 28 MPa Ar for 12 h. The ALD Nb2O5 material (19 ALD cycles) showed higher activities than

463 commercial Nb2O5 material for dehydration of 2-propanol and 2-butanol. In addition, the ALD Nb2O5-

464 supported Pd catalyst was tested for the production of valeric acid from γ-valerolactone via a ring-opening

465 reaction followed by hydrogenation, showing a higher stability than the commercial Nb2O5-supported Pd
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

466 catalyst during 70-h TOS operation.

Green Chemistry Accepted Manuscript


467 Recently, there have been efforts to exploit the so called “strong metal-support interaction (SMSI)”

468 phenomenon as an alternative method of ALD. The SMSI causes migration of metal oxide species onto

469 the surface of metal particles in a supported metal catalyst.192-194 The SMSI effect was used to tune a

470 Co/TiO2 catalyst for hydrogenation of furfuryl alcohol in the aqueous phase.195 The SMSI was caused by

471 high-temperature treatments (calcination and reduction), proven by in situ Raman spectroscopy. The

472 SMSI not only stabilized Co particles against sintering and leaching but also increased selectivity toward

473 C–O bond scission of the furan ring. The extent of the selectivity increase was associated with calcination

474 treatments. However, a similar SMSI did not properly stabilize a Co/TiO2 catalyst for aqueous-phase

475 HDO of sorbitol (20 wt.% solution).72 The reasons for this difference in stability between the two studies

476 may be related to the concentration of the feed, oxygen content of the feed, and reaction temperature. The

477 ALD and SMSI catalysts introduced above were only stable at feed concentrations of less than 10 wt.%

478 and temperatures less than 160 °C. ALD should be applied to stabilize supported metal catalysts with

479 more concentrated feedstock at high temperatures by considering different factors: thickness of ALD

480 overcoat, overcoating with mixing different oxides, etc.

481 Other than ALD and SMSI, pyrolysis of carbonaceous substances was used to stabilize catalyst by

482 forming a carbon film onto the catalyst. For example, Datye and co-workers hydrothermally stabilized

483 commercially available SiO2 and Al2O3 by depositing sucrose-derived carbon (approximately 10 wt.%) on

484 the surface of oxide materials.196 The deposition procedure involved the addition of an aqueous sucrose

485 solution to SiO2 or Al2O3, followed by stirring overnight at ambient temperature. The dried product was

486 then pyrolyzed in N2 at 400 °C for 2 h. The carbon-coated oxides were stable after hydrothermal

34
Page 35 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

487 treatment in water at 200 °C and 2.2 MPa for 12 h (i.e., no loss of surface area or structural integrity). In

488 addition, Pd nanoparticles dispersed on the carbon-coated oxides showed higher performance than those

489 supported on normal oxides for selective hydrogenation of a mixture of acetylene and ethylene to ethane.

490 Apart from deposition methods, precisely controlled surface functionalization could be applied to
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

491 improve the stability of solid acid catalysts. Zapata et al. generated a hydrophobic surface on HY zeolite

Green Chemistry Accepted Manuscript


492 by functionalizing with octadecyltrichlorosilane.197 The increase in hydrophobicity stabilized the zeolite

493 for conversion of biomass-derived oxygenates (e.g., 2-propanol and m-cresol) at 200 °C in a biphasic

494 mixture of water and decalin (50:50 v/v), preventing the zeolite structure from degradation by limiting the

495 contact between the functionalized hydrophobic zeolite surface and water. Table 7 summarizes the

496 catalyst stabilization methods and the area of usage discussed in this section.

497

35
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 36 of 71

498 Table 7. Catalyst stabilization methods for reactions involved in HDO of biomass-derived oxygenates
Method for catalyst
Catalyst Reaction Deactivation Reference
stabilization
Hydrogenation of furfural (liquid 177
Cu/γ-Al2O3 (Al2O3 overcoat) Sintering; leaching
phase)
Cu/γ-Al2O3 (MgOx−Al2O3 Hydrogenation of furfural (liquid 186
Coking
overcoat) phase)

Green Chemistry Accepted Manuscript


ALD
Hydrogenation of furfuryl alcohol 176
Co/TiO2 (TiO2 overcoat) Sintering; leaching
(aqueous phase)
Production of valeric acid from γ- 191
Pd/SBA-15 (Nb2O5 overcoat) Support dissolution
valerolactone
Hydrogenation of furfuryl alcohol 195
SMSI Co/TiO2 (TiOy overlayer) Sintering; leaching
(aqueous phase)
Pd/SiO2 and Pd/Al2O3 (carbon 196
Pyrolysis of sugar Hydrogenation of acetylene/ethylene Support dissolution
film)
Chemical vapor Hydrogenation of lactic acid (aqueous Sintering; support 198
Ru/γ-Al2O3 (carbon coating)
deposition phase) dissolution
Non-hydrolytic sol-gel
chemistry-based SBA-15 (TiO2 film) Dehydration of styrallyl alcohol Carbon poisoning 199

deposition
Surface functionalization
using silylation with HY zeolite Dehydration Dissolution 197

octadecyltrichlorosilane
499

36
Page 37 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

500 5. Electrochemical hydrodeoxygenation

501 Electrochemical reduction including hydrogenation and hydrogenolysis has been considered a way to

502 upgrade biomass-derived oxygenates. Electrochemical HDO of biomass-derived chemicals can be

503 performed at ambient conditions without externally supplied hydrogen gas,200 because the hydrogen can
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

504 be obtained from aqueous solutions. Electrochemical processes can potentially be available in remote

Green Chemistry Accepted Manuscript


505 locations at small scale and for pairing with extra electricity from renewable energy sources during off-

506 peak times. These benefits led to recent studies of electrochemical upgrading of biomass-derived

507 oxygenates. However, current studies of electrochemical HDO of biomass-derived chemicals have been

508 focused on the relationships between electrochemical reaction conditions and resulting products with

509 reaction efficiency to investigate the feasibility of electrochemical biomass upgrading. Monometallic

510 catalysts were mainly used for the current studies, and electrochemical HDO requires development of

511 advanced heterogeneous catalysts to improve the reaction efficiency.

512 As depicted in Figure 1 and discussed above, the roles of different catalytic sites (metal, acid,

513 bifunctional metal-acid) in HDO have been investigated by a variety of researchers, and the information

514 gives advice to design advanced and efficient heterogeneous catalysts for HDO of biomass-derived

515 oxygenates. In the same way, understanding the effects of current HDO electrocatalysts and reaction

516 conditions for biomass-derived oxygenates would help to reveal the roles of electrocatalysts. Then, the

517 information will give the direction to develop advanced and efficient electrocatalysts for electrochemical

518 HDO at ambient condition without externally supplied hydrogen gas.

519 This section reviews the electrochemical HDO of biomass-derived oxygenates to summarize the

520 current efforts, challenges and needs in this field, comparing to non-electrochemical HDO.

521

522 5.1. Fundamentals of electrochemical HDO

523 Key difference of electrochemical HDO with conventional chemical HDO is a source of hydrogen.

524 Without dissociation of externally supplied hydrogen gas, hydrogen for electrochemical HDO can come

37
Green Chemistry Page 38 of 71
View Article Online
DOI: 10.1039/C9GC01210A

525 from aqueous solutions. For example, electrochemical hydrogenation and hydrogenolysis has two main

526 routes: (1) electroreduction of proton (in acidic) or water (in neutral or basic) leading to the adsorption of

527 hydrogen atoms on a cathode electrode surface (Volmer reaction), and then electrocatalytic hydrogenation

528 and hydrogenolysis of an organic compound with the adsorbed hydrogen on the electrode surface, or (2)
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

529 direct electroreduction of an organic compound uses protons in solution to make proton-electron pairs,

Green Chemistry Accepted Manuscript


530 which forms a radical intermediate (C-OH) due to participation of carbonyls in electron transfer (Figure

531 4).201 The adsorbed hydrogen atoms from Volmer reaction for electrocatalytic hydrogenation and

532 hydrogenolysis can produce hydrogen gas via Heyrovsky and Tafel reactions. This illustrates that

533 adsorbed hydrogen atoms via Volmer reaction has either hydrogen evolution reaction (HER) or

534 hydrogenation and hydrogenolysis (ECH) of the organic molecule.

535

536

537 Figure 4. Proposed pathways of electrochemical reduction of carbonyls in acidic electrolytes. Reprinted
538 with permission from Chadderdon et al.,201 Copyright (2017) American Chemical Society.
539

540 Electrochemical efficiency, also known as the faradaic efficiency, accounts for the number of

541 electrons used for the desired electrochemical reactions per total number of electrons applied to the entire

542 reaction processes. High electrochemical efficiency for the electrochemical HDO is required, avoiding the

38
Page 39 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

543 side HER. As shown in Figure 5, chemisorbed hydrogen atoms from the Volmer reaction can be used for

544 either the electrochemical hydrogenation and hydrogenolysis or the HER following the Heyrovsky or

545 Tefel reactions. For selective electrochemical HDO of biomass-derived chemicals over HER, high

546 probability for Volmer reaction with less Heyrovsky or Tafel reactions (i.e., HER) is required.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

547 Figure 6 illustrates the onset potentials of both the HER and the electrochemical hydrogenation and

Green Chemistry Accepted Manuscript


548 hydrogenolysis of HMF, which shows the tendency for HER and HDO of HMF on different types of

549 electrodes at the broad range of pH.202 Pd, Co, Pt, Ni and Au had 0.3 to 0.9V lower onset potential for

550 HER, compared to other transition or poor metals in 0.5 M H2SO4. Their onset potentials for HER were

551 before the onset potentials for electrochemical HDO of HMF in the acidic condition. In general, noble

552 metals are recognized as stable and active electrocatalysts, because they have high resistance to corrosion

553 of metal and are active for various electrochemical reactions. However, these metals (Pd, Pt, and Au) have

554 high possibility to have HER over electrochemical hydrogenation and hydrogenolysis of HMF in acidic

555 condition. In neutral solution, the onset potentials for HDO of HMF were lower than HER (Figure 5) at

556 most of electrodes.

557

558

559 Figure 5. Comparison of onset potentials of HER and HMF product formation on (a) transition d metals
560 and (b) post-transition sp metals under acidic (0.5 M H2SO4) and neutral (0.1 M Na2SO4) conditions. The
561 onset potentials for HER are obtained at a current density of −0.5 mA cm−2. Reproduced with permission
562 from Kwon et al.,202 Copyright 2015 John Wiley and Sons.

39
Green Chemistry Page 40 of 71
View Article Online
DOI: 10.1039/C9GC01210A

563

564 According to the Biddinger and Li groups, the onset potentials of both the HER and electrochemical

565 hydrogenation and hydrogenolysis of furfural was close on Cu in 0.5 M H2SO4.201, 203, 204 Electrochemical

566 HDO near the onset potentials for both HER and HDO is one of the option that can suppress the HER, but
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

567 the reaction rate at the near onset potential could be slow. This means that development of electrocatalysts,

Green Chemistry Accepted Manuscript


568 having lower onset potential for HDO than the onset potential for HER, is needed to have high

569 electrochemical efficiency and reaction rate.

570

571 5.2. Electrochemical hydrogenation and hydrogenolysis of furanic compounds

572 In the initial studies of electrochemical hydrogenation and hydrogenolysis of furfural, the primary

573 product was furfuryl alcohol from electrochemical hydrogenation of C=O group on furfural. The Belgsir

574 group reported electrochemical redox reaction of furfural on different electrodes and reaction

575 conditions.205, 206 Their primary product was furfuryl alcohol from electrochemical hydrogenation of

576 furfural in the broad range of pH (1 to 10) and different electrodes (Cu, Pt, Pb, and graphite felt). The

577 highest yield of furfuryl alcohol was 72% on Cu at pH 10 and –0.65 V (vs. SCE) in their study.206 They

578 also qualitatively reported the production of 2-methylfuran, 2-methyltetrahydrofuran, other furfural

579 derivatives as side products of ECH of furfural.206 Among the electrochemical hydrogenation of furfural

580 in aqueous electrolytes, the highest yield of furfuryl alcohol was 82% (99% selectivity) at –0.5 V (vs.

581 SCE) on Pt supported on activated carbon fiber in 0.1 M H2SO4.207 Highest yield of furfuryl alcohol in

582 non-aqueous electrolyte was 89% on La-doped TiO2 electrode with the current density of –50 mA cm–2 in

583 0.1 M tetrabutylammonium bromide in dimethylformamide.208

584 Recent studies showed the significant production of 2-methylfuran from electrochemical

585 hydrogenolysis of furfuryl alcohol or furfural in strongly acidic conditions (pH ≤ 1). Saffron209,

586 Schröder210, Li201, and Biddinger203, 204, 211 groups studied the electrochemical hydrogenation and

587 hydrogenolysis of furfural in an electrochemical batch cell reactor with different acidities and electrodes

40
Page 41 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

588 (Table 8). The selectivity toward 2-methylfuran on a Cu electrode was higher than 40% in highly acidic

589 solutions (pH ≤ 0.5),201, 203, 204, 210, 211 while that on a Ni electrode was between 10%209 and 30%210 with

590 pH ≤ 1. When pH was higher than 1, the selectivity of 2-methylfuran was significantly decreased, and

591 furfuryl alcohol became a primary product on both Cu201, 210, 211 and Ni209 electrodes. These observations
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

592 showed that strongly acidic conditions (pH ≤ 1) were required to produce significant amount of 2-

Green Chemistry Accepted Manuscript


593 methylfuran (selectivity > 40%), and the highest yield of 2-methylfuran among previous studies was near

594 55% on a Cu electrode in 0.5 M H2SO4 in water/acetonitrile (3/1 by volume).201 Also, a dimer product,

595 hydrofuroin, was quantitatively shown over several electrodes.201, 210 The Li group distinguished two

596 reaction pathways for electroreduction of furfural: (1) electrocatalytic hydrogenation and hydrogenolysis

597 of furfural to furfuryl alcohol and 2-methylfuran with adsorbed hydrogen on the electrode surface and (2)

598 direct electroreduction of furfural to a dimer product with proton-electron pairs (Figure 6).

599 Huber group studied electrochemical hydrogenation and hydrogenolysis of furfural on a continuous

600 electrocatalytic membrane reactor. They produced protons and electrons from water splitting on a Pt–

601 Ru/C anode, and the protons and electrons were used for ECH of furfural on a Pd/C or Pt/C cathode.212

602 Furfuryl alcohol was the only product from ECH of furfural at lower cell potentials (< 1.3 V), but further

603 reduced products, 2-methylfuran, tetrahydrofurfuryl alcohol and 2-methyltetrahydrofuran, were produced

604 when the cell potential was higher than 1.45 V (Table 8).212 This showed that the furfuryl alcohol was an

605 intermediate for the production of 2-methylfuran, tetrahydrofurfuryl alcohol and 2-methyltetrahydrofuran

606 on Pd/C or Pt/C (Figure 6).212

607

41
Green Chemistry Page 42 of 71
View Article Online
DOI: 10.1039/C9GC01210A
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

Green Chemistry Accepted Manuscript


608

609 Figure 6. Possible reaction pathways for electrochemical HDO of furfural.201, 204, 212
610

611 According to Li and Biddinger groups, 2-methylfuran was directly produced from electrochemical

612 hydrogenolysis of furfural,201, 204 showing that furfuryl alcohol was not the intermediate for 2-methylfuran

613 production from electrochemical HDO of furfural. Unlike Pt/C and Pd/C electrocatalysts from Huber

614 group study,212 electrochemical hydrogenolysis of furfuryl alcohol was negligibly shown on a Cu

615 electrode in 0.5 M H2SO4, regardless of applied potential.204 These results were inconsistent with the

616 reaction pathways of thermal catalytic HDO of furfural on Cu-based catalysts.213 Thermal catalytic HDO

617 had two-step hydrogenation of furfural (C=O bond) to furfuryl alcohol and hydrogenolysis of furfuryl

618 alcohol (cleavage of C–O bond) to 2-methylfuran on Cu-based catalysts.213

619 Electrochemical hydrogenation and hydrogenolysis of HMF has been also studied, and the studies

620 have been primarily focused on the relationship between reaction conditions and resulting products. Kwon

621 et al. investigated a broad range of electrode materials in neutral214 and acidic202 solutions, applying

622 different reaction potentials. The product distributions were categorized into three electrode groups in

623 both solutions. In neutral solution (0.1 M Na2SO4), DHMF from direct hydrogenation of HMF was the

624 primary product on Fe, Ni, Ag, Zn, In and Cd electrodes. However, Co, Au, Cu, Sn, and Sb electrodes

625 resulted in further reduced products.214 On Pd, Al, Bi, and Pb electrodes, product distribution toward

42
Page 43 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

626 DHMF and further reacted products depended on the applied potentials.214 The qualitatively reported

627 further reduced products include 5-methylfurfural (MFF), 5-methylfurfuryl alcohol (MFA), 2,5-

628 dimethylfuran (DMF), and 2,5-dimethyl-2,3-dihydrofuran (DMDHF).214 Figure 7 shows the proposed

629 reaction pathways for electrochemical hydrogenation, hydrogenation and ring opening of HMF. In acidic
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

630 solution (0.5 M H2SO4), the product distributions were also categorized into three groups of electrodes.202

Green Chemistry Accepted Manuscript


631 DHMF was the main product on Fe, Ni, Cu, and Pb, while DMDHF was the main product on Pd, Pt, Al,

632 Zn, In, and Sb. Selectivity toward DHMF and DMDHF relied on the applied potentials over Co, Ag, Au

633 Cd, Sn, and Bi electrodes in acidic solution.202 However, Schröder group showed the highest selectivity

634 (36%) of DMF on a monometallic Cu electrode, followed by DHMF (34%) and MFA (11%) in highly

635 acidic solution (0.5 M H2SO4 in 1:1 mixture of water and ethanol) (Table 8).210

636

637

638 Figure 7. Possible reaction pathways for electrochemical HDO of HMF.202, 214-216
639

640 Choi group developed an Ag electrode with a nanocrystalline dendritic and fractal morphology, and

641 they had 99% selectivity and 99% electrochemical efficiency for DHMF production at pH 9.2 (0.5 M

642 borate buffer solution.215 Choi group also observed the formation of 2,5-hexanedione (HD) and hydroxy-

643 2,5-hexanedione (HHD) from the ring opening of the HMF on a Zn electrode at pH 2.0 (0.2M sulfate

644 buffer solution) having highest selectivity (82%) of 2,5-hexanedione (Table 8).216 When pH was 7.2

43
Green Chemistry Page 44 of 71
View Article Online
DOI: 10.1039/C9GC01210A

645 (0.2M phosphate solution), however, DHMF was a primary product on a Zn electrode. On Au, Cu, and Pd

646 electrodes, DHMF was produced primarily at pH 2.0, showing ring opening products.216 However, the

647 product yields from electroreduction of HMF were less than 3% or were not reported with monometallic

648 electrocatalysts.202, 210, 214-216


Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

649 In recent work, Zhang et al. investigated electrochemical hydrodeoxygenation of HMF to produce

Green Chemistry Accepted Manuscript


650 DMF with bimetallic catalyst, composed of Cu and Ni.217 They used electrodeposition of Cu precursor to

651 make high surface area Cu on Cu foil, followed by electrodeposition of Ni on the Cu substrate. The

652 faradaic efficiency for DMF production was 88.0%, and its selectivity was 88.8% selectivity at pH 2

653 buffer solution and −0.8 V vs Ag/AgCl. The yield of DMF was 40.5% with 150 C charge passed. Side

654 products observed were DHMF (1.5%) and MFA (0.8%), respectively (Table 8). The selectivity, yield,

655 and faradaic efficiency for DMF production on CuNi bimetallic catalyst were the highest values among

656 other electrochemical HDO of HMF. In the study, they argued that Cu site could have high affinity to

657 oxygen groups of HMF, and hydrogen atom (Hads) could be readily adsorbed on the surface. Also, they

658 mentioned that this resulted in the promotion of hydrogenation and hydrogenolysis of HMF to DMF,

659 compared to monometallic Cu or Ni catalysts.217

660 In electrochemical reduction of FF and HMF, hydrogenation of C=O on furfural and HMF was shown

661 in a broad range of pH and electrocatalysts. While, strongly acidic electrolytes have been required to

662 significantly promote electrochemical hydrogenolysis of furfural or furfuryl alcohol to 2-methylfuran201,

663 203, 204, 209-211 and to convert HMF into further reduced DMF,202, 210 DMDHF,202 or HD.216

664 In the acidic conditions, Zn catalyst showed the ring opening of HMF,216 and Cu catalysts showed the

665 hydrogenolysis of aldehyde or alcohol to produce 2-methylfuran and DMF.201, 203, 204, 210, 211, 217 Among the

666 Cu-based catalysts, CuNi bimetallic catalyst showed the highest performance to produce DMF from

667 electrochemical HDO of HMF.217 However, still there is lack of mechanism understanding how the Cu-

668 based catalyst promoted C-O cleavage, and CuNi bimetallic catalyst showed superior performance for

669 electrochemical HDO of furan oxygenates over Cu monometallic catalyst. The development of advanced

44
Page 45 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

670 electrocatalysts and understanding of mechanisms for electrochemical HDO of oxygenates will help to

671 further improve the reaction efficiency and widen the applications of electrochemical upgrading of

672 biomass-derived oxygenates. Following section will deal with the electrochemical HDO of other biomass-

673 derived oxygenates.


Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

674

Green Chemistry Accepted Manuscript

45
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Green Chemistry Page 46 of 71

675 Table 8. Electrochemical HDO of furanic compounds over different catalysts. All results are available from recent literature.

Selectivity, %
Feedstock Products a Reactor type Catalyst Reaction conditions Reference
(Faradaic efficiency, %)

Two compartments −10 mA cm-2 (19.8 cm2) for 2 to 4 h; 8 °C; ~ 80 for 2-MF
Furfural 2-MF, FA Cu 210
cell 0.5M H2SO4 in water/acetonitrile mixture ~ 10 for FA

Green Chemistry Accepted Manuscript


Two compartments −10 mA cm-2 (19.8 cm2) for 2 to 4 h; 8 °C; ~ 30 for 2-MF
Furfural 2-MF, FA Ni 210
cell 0.5M H2SO4 in water/acetonitrile mixture ~30 for FA
Two compartments −10 mA cm-2 (19.8 cm2) for 2 to 4 h; 8 °C; ~ 20 for 2-MF
Furfural 2-MF, FA Pt 210
cell 0.5M H2SO4 in water/acetonitrile mixture ~60 for FA
Two compartments −0.8V vs RHE for 1.5 h; 25 °C; 0.1M FF in 51.0 (50.0) for 2-MF
Furfural 2-MF, FA Cu 213
cell 0.5M H2SO4 water/acetonitrile (4/1) mixture 11.2 (5.6) for FA
Two compartments −0.5V vs RHE for 1.5 h; 25 °C; 0.1M FF in 52.8 (74.8) for 2-MF
Furfural 2-MF, FA Cu/Cu 214
cell 0.5M H2SO4 water/acetonitrile (4/1) mixture 24.4 (19.9) for FA
Two compartments −0.55V vs RHE for 1 h; 0.05M FF in 0.5M ~ 70 (~ 58) for 2-MF
Furfural 2-MF, FA Cu 201
cell H2SO4 water/acetonitrile (3/1) mixture ~ 10 (~ 4) for FA
Single −6mA cm−2 for 5 h; 0.1M FF in water/methanol 9.7 (9.6) for 2-MF
Furfural 2-MF, FA Ni 209
compartment cell (4/1) mixture, pH 1 59.1 (29.4) for FA
~ 8 for 2-MF
1.8 V (applied voltage); 111.7 h−1 WHSV b
2-MF, FA, THFA, Continuous ~ 55 for FA
Furfural Pd/C furfural; 30 °C; PtRu/C anode for water 212
MTHF membrane reactor ~ 25 for THFA
electrolysis
~ 10 for MTHF
35.6 for DMF
DMF, DHMF, Two compartments −10 mA cm-2 (19.8 cm2); 8 °C; 33.8 for DHMF
HMF Cu 210
MFF, MFA cell 0.5M H2SO4 in water/ethanol (1/1) mixture 11.1 for MFF
0.5 for MFA
Single −0.89 V vs RHE, 20 °C charge passed; 20mM 81.6 (72.4) for HD
HMF HD, HHD Zn 216
compartment cell HMF in 0.2M sulfate buffer solution (pH 2) 9.0 (5.6) for HHD
Two compartments CuNi 88.8 (88.0) for DMF
DMF, DHMF, −0.8 V vs Ag/AgCl, 150 °C charge passed;
HMF divided cell, linked (82/17 1.5 (0.7) for DHMF 217
MFA 0.635 mM HMF in 0.2M sulfate buffer (pH 2)
by salt bridge at%) 0.8 (0.6) for MFA
a 2-MF (2-methylfuran), FA (furfuryl alcohol), THFA (tetrahydrofurfuryl alcohol), MTHF (2-methyltetrahydrofuran), HMF (5-
hydroxymethylfurfural), DHMF (2,5-dihydroxymethylfuran), MFF (5-methylfurfural), MFA (5-methylfurfuryl alcohol), DMF (2,5-
dimethylfuran), HD (2,5-hexanedione), HHD (Hydroxy-2,5-hexanedione)
b WHSV (weight hourly space velocity), defined as the ratio of furfural mass flow rate to the mass of metal catalyst present

46
Page 47 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

677 5.3. Electrochemical hydrogenation and hydrogenolysis of other biomass-derived oxygenates

678 Besides the electrochemical hydrogenation and hydrogenolysis of furanic compounds, there have been

679 electrochemical HDO of other biomass-derived oxygenates. The Saffron and the Jackson groups produced

680 cyclohexanol (C6H12O) from guaiacol (C7H8O2) or its analogues by electrochemical reduction on Ru/C218
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

681 and RANEY® Nickel (Ra-Ni)219 electrocatalysts. Seven different Ru/C catalysts were used as a cathode

Green Chemistry Accepted Manuscript


682 for electrochemical conversion of a lignin monomer, guaiacol, in different catholytes (0.2 M HCl, 0.2 M

683 NaCl, or 0.2 M NaOH catholytes).218 Ru/C catalysts were prepared by different combinations of the

684 preparation method (incipient wetness impregnation or cation exchange) and Ru precursor (RuCl3 or

685 Ru(NO)(NO3)3). The catalyst preparation method, precursor and reaction temperature affected guaiacol

686 conversion and the electrochemical efficiency. Maximum cyclohexanol selectivity (72%) at 53% guaiacol

687 conversion was achieved at 80 °C and in 0.2 M HCl with the Ru/C catalyst synthesized by incipient

688 wetness impregnation of Ru(NO)(NO3)3 (Table 9).218

689 Ra-Ni electrocatalyst was used for the electrochemical conversion of guaiacol or its analogues.219 Ra-

690 Ni cathode was prepared by electroplating in a nickel ammonia plating solution, and a

691 cobalt-phosphate electrode was prepared as an anode for water splitting to produce protons and electrons

692 for the reaction.219 From the electrochemical reduction of guaiacol (2-methoxyphenol) and its analogues

693 such as 2-ethoxyphenol, 2-isoproxyphenol, 3-methoxyphenol and 4-methoxyphenol, the highest yield

694 toward cyclohexanol was 91% from the 2-ethoxyphenol conversion at pH 8 and 75 °C (Table 9).219 They

695 studied the reaction mechanism and suggested the hydrodemethoxylation of the alkoxyphenols on the Ra-

696 Ni219 and Ru/C218 electrodes. This process was also recently patented by the authors.220

697 Since the Ru/C catalysts showed the capability to reduce biomass-derived oxygenates, they expanded

698 this work to 16 aromatic rings, which were lignin-derived oxygenates and their analogues.221 From the

699 study of lignin pyrolysis, they revealed various aromatic monomers, exhibiting different oxygenated

700 functional groups such as hydroxyl, methoxy, alkyl, allyl, and carbonyl groups. Figure 8 shows

701 electrochemical hydrogenation and further reductions of different aromatic rings on Ru/C catalyst.

47
Green Chemistry Page 48 of 71
View Article Online
DOI: 10.1039/C9GC01210A

702 Regardless of functional groups attached to the aromatic rings, ring saturation with hydrogenation was

703 occurred. Phenol (1) and catechol (14) have one and two hydroxyl functional groups, respectively. Phenol

704 (1) was exclusively converted to ring saturated cyclohexanol (17) with only hydrogenation. However,

705 catechol (14) produced ring saturated 1,2-cyclohexanediol (18), and further C-O scission of the second
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

706 hydroxyl group to catechol (14) produced cyclohexanol (17) (Figure 8). Anisole (13) and 3-

Green Chemistry Accepted Manuscript


707 methoxyanisole (15), which were model aromatic compounds with only methoxy groups, had dominant

708 ring saturation reaction with hydrogenation. Further demethoxylation and hydrogenation of methoxy

709 groups on 3-methoxyanisole (15) resulted in the formation of deoxygenated products, cyclohexanol (17)

710 and methoxycyclohexane (19) (Figure 8). Guaiacol (2a) and syringol (10), having hydroxyl and methoxy

711 groups, also showed ring saturation. Additional demethoxylation of methoxy groups on them produced

712 deoxygenated products, cyclohexanol (17) and 2-methoxycyclohexanol (21) (Figure 8 and Table 9).221

713 No trace amounts of fully hydrodeoxygenated products such as cyclohexanone and benzene were

714 detected from phenolic oxygenates. The only hydroxyl group on the aromatic rings did not have further

715 reduction. However, the second and third oxygen functional groups attached to aromatic rings had

716 hydrogenolysis and demethoxylation reactions, resulting in the production of deoxygenated products

717 (Figure 8 and Table 9).221

718 Schröder and the Brushett groups reported the electrochemical hydrogenation and hydrogenolysis of

719 hydroxyacetone on different metals electrodes. The Schröder group tested 11 different cathode metals (Ti,

720 Cr, Mo, W, Fe, Ni, Cu, Zn, Al, Sn, and Pb) for electrochemical HDO of hydroxyacetone.222 The process

721 was performed at ambient temperature and pressure with a constant potential of –1.8 V (vs. Ag/AgCl) in

722 neutral electrolytes. At comparable conditions with a NaHSO4 electrolyte, Zn showed the highest

723 selectivity toward acetone (C3H6O) for electrochemical hydrogenolysis of hydroxyacetone (C3H6O2)

724 among the tested cathode materials (~75%). They claimed that the high selectivity on the Zn catalyst

725 might be due to the fully loaded d-orbitals of Zn.222 However, Cu was most active (highest reaction rate)

48
Page 49 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

726 during the formation of acetone from hydroxyacetone in sodium chloride and hydrochloric electrolyte

727 solutions.

728
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

Green Chemistry Accepted Manuscript


729

730 Figure 8. Observed products from electrochemical HDO of lignin-based model compounds. Reprinted
731 with permission from Garedew et al.,221 Copyright (2019) American Chemical Society.
732

733 Brushett groups prepared monometallic and bimetallic electrocatalysts on Vulcan carbon support by

734 incipient wetness impregnation: Cu/C, Ni/C, CuNi/C, Pt/C, Ru/C, and PtRu/C electrocatalysts.223 They

735 studied the electroreduction of hydroxyacetone at pH 2 and at –1.5 V (vs. Ag/AgCl), and the highest

736 selectivity toward acetone was ~ 18% on Cu/C. However, the primary product was propylene glycol from

737 the electrochemical hydrogenation of hydroxyacetone on the six electrocatalysts tested.223

49
Green Chemistry Page 50 of 71
View Article Online
DOI: 10.1039/C9GC01210A

738 Electrochemical hydrogenation and hydrogenolysis of levulinic acid have also been reported. Nilges

739 et al. reported the production of n-octane from the levulinic acid via two step electrochemical reactions:

740 electrochemical hydrogenolysis and Kolbe electrolysis (decarboxylation and dimerization).224 Also, they

741 reported valeric acid as a reaction intermediate, which can be obtained from the electrochemical
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

742 hydrogenolysis of levulinic acid at –1.8 V (vs. Ag/AgCl) with Pb and 0.5 M H2SO4.224

Green Chemistry Accepted Manuscript


743 Xin et al. also studied the electrochemical hydrogenation and hydrogenolysis of levulinic acid on Pb at

744 pH 0 and pH 7.5, applying different potentials from –0.35 to –1.6V (vs. RHE).225 At –1.3V (vs. RHE)

745 with Pb electrode, valeric acid had 95% selectivity at pH 0, while gamma valerolactone (gVL) had 100%

746 selectivity at pH 7.5. Also, the selectivity toward valeric acid increased as the magnitude of the electrode

747 increased at pH 0. Figure 9 illustrates the reaction pathways for electrochemical hydrogenation and

748 hydrogenolysis of levulinic acid over a Pb electrode.

749

750

751 Figure 9. Proposed reaction pathway for the electrochemical hydrogenation and hydrogenolysis (ECH) of
752 levulinic acid to form valeric acid and γ-valerolactone in acidic electrolyte on Pb. Modified reproduced
753 with permission from Xin et al.,225 Copyright 2013 John Wiley and Sons.
754

755 When the magnitude of the applied potential increased in acidic environment, the C–O bond scission

756 was occurred for the hydrogenolysis of reactant. Electrochemical hydrogenolysis of levulinic acid in

757 acidic solution was consistent with electrochemical hydrogenolysis of other biomass-derived oxygenates

758 in acidic solutions mentioned above in this section.

50
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:
Page 51 of 71 Green Chemistry

759 Table 9. Electrochemical HDO of other biomass-derived oxygenates over different catalysts. All results are available from recent literature

Selectivity, %
Feedstock Products a Reactor type Catalyst Reaction conditions Reference
(Faradaic efficiency, %)

Two
−100mA; 80 °C; 20mM guaiacol in 0.2M HCl 218
Guaiacol Cyclohexanol compartments Ru/C 70
water/isopropanol (30/1)
cell

Green Chemistry Accepted Manuscript


Two
Cyclohexanol, −100mA for 2 h; 80 °C; 20mM guaiacol in 62 for Cyclohexanol 221
Guaiacol compartments Ru/C
2-Methoxycyclohexanol 0.2M HCl water/isopropanol (30/1) 38 for 2-Methoxycyclohexanol
cell
−8 mA cm−2 for 6 h; 75 °C; 0.1 M pH 8
Two
potassium borate buffer and 0.5mM 219
2-ethoxyphenol Cyclohexanol compartments Ra-Ni Yield 91% (23)
triethylammonium bromide; CoP anode for
cell
water electrolysis
Two
−1.8V vs Ag/AgCl for 4 h; 0.09M 222
Hydroxyacetone Acetone compartments Zn 75 (~5)
hydroxyacetone in 0.5M NaHSO4 (pH 1.15)
cell
Two ~18 for acetone
Acetone, propylene −1.5V vs Ag/AgCl; 0.05M hydroxyacetone in 223
Hydroxyacetone compartments Cu/C ~50 for propylene glycol
glycol, isopropanol Na2SO4 (pH 2)
cell ~3 for isopropanol
Single or two
−1.8V vs Ag/AgCl for 4 h; 8 °C; 0.1M 224
Levulinic acid Valeric acid compartments Pb 97.2 (27.0)
levulinic acid in 0.5M H2SO4
cell
Single or two
−1.8V vs Ag/AgCl for 23 h; 8 °C; 0.1M 2,7- 224
2,7-Octanedione n-Octane compartments Pb 27.5 (11.2)
octanedione in 0.5M H2SO4
cell
Single cell
with glass frit −1.3 V vs RHE for 1 h; room temperature; 225
Levulinic acid Valeric acid Pb 94.1 (83.8)
for reference 0.2M levulinic acid in pH 0 solution (H2SO4)
electrode
−1.3 V vs RHE for 1 h; room temperature; 225
Levulinic acid Valeric acid Flow cell Pb 95.0 (86.5)
0.2M levulinic acid in pH 0 solution (H2SO4)
−1.3 V vs RHE for 1 h; room temperature;
Levulinic acid Gamma valerolactone Flow cell Pb 0.2M levulinic acid pH 7.5 solution (buffer 100 (18.2) 225
solution)
760

51
Green Chemistry Page 52 of 71
View Article Online
DOI: 10.1039/C9GC01210A

761 5.4. Current challenges of electrochemical HDO

762 Although acidic environments have been required to promote electrochemical hydrogenolysis for

763 HDO, in general, there have been reported concurrent side reactions during electrochemical HDO of

764 biomass-derived oxygenates in these conditions.201-204, 211, 225 One is the hydrogen evolution reaction
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

765 (HER) that consumes electrons and decreases electrochemical efficiency, and the other is the

Green Chemistry Accepted Manuscript


766 polymerization or degradation that decreases mass balance of the reaction.203, 204, 209, 211, 219

767 As described in Section 5.1, the onset potentials of both the HER and electrochemical hydrogenation

768 and hydrogenolysis of furanic compounds were close on various electrodes in 0.5 M H2SO4.201-204 Their

769 studies reported that high electrochemical efficiency for electrochemical hydrogenation and

770 hydrogenolysis of furfural was observed at near the onset potential of the reaction. However, the slow

771 reaction rate was still challenge.201, 203, 204 While the reaction rate for both the electrochemical

772 hydrogenation and hydrogenolysis of furfural and HER could be improved, the HER became dominant as

773 the applied potential increased. According to the study of Lercher group, surface coverage of chemisorbed

774 hydrogen increases as the applied potential increases, and binding energy of aromatic oxygenate, phenol,

775 weakens on Pt electrode, leading to HER.226 Other biomass-derived oxygenates had low electrochemical

776 efficiency for the hydrogenolysis at high applied potentials due to the HER.219, 222, 224, 225

777 Because noble metals prefer HER in acidic condition, non-noble metals should be considered as

778 electrocatalysts for electrochemical HDO in the acidic condition. This may lead to stability and durability

779 issues for long time use of electrocatalysts in acidic condition.

780 Recent development of electrocatalysts by using electrodeposition showed stability of HDO reactions

781 and improve electrochemical efficiency near the onset potentials. The Fan group lowered onset potential

782 for electrochemical HDO of HMF at pH 2 with the development of electrodeposited Ni on Cu bimetallic

783 catalyst.217 This catalyst led to the large reaction potential gap between HDO and HER, and the preference

784 of HDO over HER. This catalyst also showed the stability for 5 running of electrochemical HDO of HMF

785 to DMF. The Biddinger group reported the electrodeposited Cu on Cu catalyst, which had four cycles of

52
Page 53 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

786 stability for electrochemical HDO of furfural to furfuryl alcohol and 2-methylfuran. 214 Although the

787 onset potentials for both HDO and HER were not remarkably changed, the higher surface area

788 electrodeposited Cu on Cu catalyst showed 2-3 times higher reaction rate for electrochemical HDO of

789 furfural near the onset potentials, comparing to bear Cu electrode. Electrochemical efficiency of HER was
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

790 less than 3.5% at the near onset potentials in 0.5M H2SO4.214

Green Chemistry Accepted Manuscript


791 Another concern in electrochemical hydrogenation and hydrogenolysis of biomass-derived oxygenates

792 is the possibility of polymerization or/and degradation in acidic conditions. In previous reports, a poor

793 mass balance from the electrochemical hydrogenation and hydrogenolysis of biomass-derived oxygenates

794 in acidic electrolytes was observed.203, 204, 209, 211, 219 The degraded biomass-derived chemicals covered the

795 surface of electrocatalysts, decreasing catalyst activity.204, 227 The reaction activity was improved when

796 high potential was applied227 and surface area electrocatalyst was increased.186

797 Neutral and mildly acidic conditions can avoid significant HER, polymerization, degradation of the

798 chemicals, or/and catalyst stability issues. However, hydrogenolysis of biomass-derived oxygenate was

799 not significantly shown in these conditions. The development of active electrocatalysts for hydrogenolysis

800 in mild conditions will be important. The lifetime of electrocatalysts can be extended, avoiding corrosion

801 and surface coverage of degraded oxygenates on the electrocatalyst in mild conditions. HER will be

802 spontaneously suppressed, because HDO can be favored than HER.

803 Current studies for electrochemical hydrogenation and hydrogenolysis of biomass-derived oxygenates

804 have been primarily focused on the relationships between the reaction conditions and resulting products

805 and efficiency on monometallic or carbon electrodes. This suggests the need of advanced electrocatalysts

806 for selective hydrogenolysis of biomass-derived oxygenates for electrochemical HDO.

807 Advanced heterogeneous catalysts or modification from the catalysts can be applied for

808 electrochemical HDO. Bifunctional catalysts can have possibility to reduce oxygen contents more and to

809 help hydrogenolysis of biomass-derived oxygenates. As depicted in Figure 1, metal-acid catalysts can

810 help both the hydrogenation and scission of C−O bond on both metal and acid sites. However, the

53
Green Chemistry Page 54 of 71
View Article Online
DOI: 10.1039/C9GC01210A

811 bifunctional catalysts have not been applied to electrochemical reaction significantly. The development of

812 electrocatalysts, composed of metal, metal oxide with solid acid and oxophilic sites may be expected to

813 help electrochemical hydrogenation, hydrogenolysis, decarboxylation, decarbonylation or other reactions

814 of biomass-derived oxygenates. Because some bifunctional catalysts have low electrical conductivity,
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

815 they should be electrically conductive themselves for electron transfer or supported on electrically

Green Chemistry Accepted Manuscript


816 conductive materials.

817 Although the examples of electrochemical HDO showed promises in each reactant, the studies

818 primarily rely on empirical studies with a slight understanding of fundamental mechanism for

819 electrocatalysis. There are many required studies to improve electrochemical HDO, but not limited to

820 these: (1) Physical and chemical properties of electrocatalysts that control reaction pathways and product

821 selectivity, (2) orientation of organic molecules on electrocatalyst surface and their active sites, (3)

822 reaction kinetics, mass transfer and reaction mechanisms, (4) corrosion and degradation phenomena of

823 electrocatalysts during electrochemical HDO, and (5) simulation for all the approaches above. Comparing

824 to heavily studied other electrocatalysis (i.e. CO2 electroreduction and oxygen reduction reaction),

825 electrochemical HDO of biomass-derived oxygenates is rarely studied and requires more attention and

826 effort to improve this process.

827 In recent work, the Koper group reported a comprehensive study of electrochemical hydrogenation of

828 aliphatic ketone, acetone, at platinum single-crystal electrodes.228 They used online electrochemical mass

829 spectroscopy and in situ Fourier transform infrared spectroscopy for experimental understanding of

830 reaction mechanisms of electrochemical hydrogenation of acetone. Theoretical study with DFT

831 calculations was also supported to understand this reaction. Activity and product distribution of acetone

832 electroreduction on Pt electrode relied on the various factors such as surface orientation, reaction

833 potential, and coordination number of active sites. Under-coordinated step site of acetone on Pt (510) was

834 preferred to have C-O scission to produce propane. This finding supported to understand the reaction

835 mechanisms of other aliphatic ketones.228

54
Page 55 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

836 The Lercher group reported that the activity for electrochemical hydrogenation of phenol and

837 benzaldehyde is sensitivity to structure of Pt catalyst.229 The reaction rates for the hydrogenation were

838 directly proportional to the size of Pt particles (between 3.1 and 12.1 nm), because the larger particles had

839 more accessible sites for the adsorption of aromatic molecules on Pt (100) or Pt (111) surfaces.229 This
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

840 study helps to understand the competition between electrochemical hydrogenation and HER, but

Green Chemistry Accepted Manuscript


841 electrochemical hydrogenation of phenol and benzaldehyde did not have further hydrogenolysis. More

842 comprehensive hydrodeoxygenation studies on various electrocatalysts will be required to understand

843 electrochemical HDO and to design of advanced electrocatalysts.

844 The comparison of electrochemical HDO to thermal catalytic HDO is important, and it is able to help

845 to understand the feasibility of electrochemical HDO in broader applications. Lercher and co-workers

846 studied the electrochemical hydrogenation and the thermal catalytic hydrogenation of phenol230 and

847 benzaldehyde231 using monometallic catalysts on carbon support. Their studies showed the turn over

848 frequency (TOF) of electrochemical hydrogenation and thermal catalytic hydrogenation with same metal

849 catalysts, pH and pressure. In phenol hydrogenation to cyclohexanol, thermal catalytic hydrogenation had

850 four times higher TOF than electrochemical hydrogenation.230 While, electrochemical hydrogenation had

851 higher TOF than thermal catalytic hydrogenation for hydrogenation of benzaldehyde to benzyl alcohol in

852 their reaction conditions.231 However, TOF highly depends on the reaction conditions such as applied

853 potential, current, temperature or pressure.230, 231 Applied potential or current is one of key factors to

854 decide the reaction rate and selectivity in electrochemical hydrogenation, while the reaction efficiency and

855 results highly rely on temperature in thermal catalytic hydrogenation when same pressure and catalyst are

856 used. To quantitatively assess the reaction efficiency between electrochemical and thermal catalytic HDO,

857 energy demand and cost for electron transfer in electrochemical HDO should be compared to energy

858 demand and cost for high temperature and pressure applied in thermal catalytic HDO. To our best

859 knowledge, the direct comparison of energy efficiency of electrochemical HDO to thermal catalytic HDO

860 has not been clearly studied.

55
Green Chemistry Page 56 of 71
View Article Online
DOI: 10.1039/C9GC01210A

861 Electrochemical HDO is an alternative way to upgrade biomass-derived oxygenates without externally

862 supplied hydrogen gas at ambient conditions. Investigation of advanced electrocatalysts and reaction

863 systems, having high selectivity, electrochemical efficiency and reaction rate for desired reactions will

864 have opportunities to develop renewable and sustainable processes for the production of biofuels and fine
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

865 chemicals by electrochemical HDO. However, there have been many challenges and needs in this field as

Green Chemistry Accepted Manuscript


866 reviewed in this section.

867 The successful lessons from heterogeneous catalysis for HDO of biomass-derived chemicals may help

868 to develop advanced electrocatalysts. Both experimental and theoretical studies for electrochemical HDO

869 using advanced electrocatalysts with fundamental understanding of the reactions are expected to tackle

870 current challenges of electrochemical HDO of biomass-derived oxygenates.

871

872 6. Future prospects and perspectives

873 Even though the potential to employ the HDO with heterogeneous catalysts as a way to upgrade

874 biomass-derived fuels and produce value-added chemicals from biomass is clear, the continuous

875 development of practical applications of HDO is important to sustaining interest in the process. Despite

876 praiseworthy achievements in HDO of biomass-derived oxygenates, there are still disadvantages

877 including the needs for harsh reactions conditions (e.g., high temperatures and pressures in the presence of

878 water), the employment of molecular H2 as a hydrogen donor, a high cost of developing catalyst (i.e.,

879 generally precious metal catalysts), a low selectivity toward target products, and the low catalyst stability

880 under HDO conditions.

881 There is a clear need for improved HDO bifunctional catalysts that has high stability under reaction.

882 Both surface and bulk properties for heterogeneous catalysts should be resistant to lots of water present

883 (up to 95% H2O) during the reaction. Finding phase transformation inhibitors that interact with active

884 sites helps improve structural stability during aqueous-phase processing. Non-precious metal sites that are

885 stable under the reaction have a better potential than precious metallic sites. Particular concern about the

56
Page 57 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

886 stability of bifunctional HDO catalysts is that the different catalytic sites contained in bifunctional

887 catalysts (i.e., metal and acid sites) are deactivated due to different mechanisms.232 Metal particles suffer

888 from leaching and/or sintering, but solid-acid supports suffer from phase transformation and/or loss of

889 surface area. However, current methods to stabilize each type of catalytic sites are limited. For instance,
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

890 ALD overcoat or SMSI could be used for the stabilization of metal particles, and carbon deposition or

Green Chemistry Accepted Manuscript


891 surface functionalization could be used for the stabilization of solid acids. However, preparing

892 bifunctional catalysts with these different methods is difficult. Considering that the improvement on

893 stability can help address the issues of high catalyst costs, more research is needed on the stabilization of

894 the bifunctional sites at HDO reaction conditions. To this end, long-term operation studies in continuous

895 reactor systems need to be done, because batch reactors (e.g., autoclave), which are typically used for

896 HDO reactions in most studies, are not suitable option to evaluate catalyst stability.

897 One of the major drawbacks of most HDO processes that large amount of molecular H2 that is

898 typically originated from non-renewable resources (e.g., steam reforming of coal). The use of gaseous H2

899 creates challenges of process economy and transport. More importantly, it has low solubility in most

900 solvents, resulting in high pressure. The high H2 pressure can raise serious safety issues.233 The

901 employment of CTH reactions may be options to overcome this problem (also refer Section 3).234, 235

902 Electrochemical HDO is an alternative way to upgrade biomass-derived oxygenates into more valuable

903 chemicals. This process does not require externally supplied H2 gas, because the source of hydrogen can

904 be from the reaction solutions. Thus, it can allow small and on-site upgrading of biomass-derived

905 chemicals in remote locations where H2 gas is not readily available. Significant development of renewable

906 electricity will make electrochemical HDO more attractive with the cost down of electricity, though the

907 cost of electricity may be considered a challenge in electrochemical HDO. In addition, electrochemical

908 processes are available in ambient conditions and can be paired with renewable electricity during the off-

909 peak times. These advantages recently led to emerging studies of electrochemical upgrading of biomass-

910 derived oxygenates, but still there have been challenges and needs in this field such as techno-economic

57
Green Chemistry Page 58 of 71
View Article Online
DOI: 10.1039/C9GC01210A

911 feasibility, fundamental understanding of electrochemical HDO of biomass-derived chemicals, undesired

912 side reactions, relatively lower yield compared to thermal catalytic HDO, and others.

913 As mentioned above, electrochemical HDO does not require externally supplied H2 gas and works at

914 ambient condition. However, this reaction requires electricity for electroreduction. The estimation of total
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

915 cost for electrochemical HDO of biomass-derived chemicals and techno-economic feasibility of the

Green Chemistry Accepted Manuscript


916 electrochemical processes will be required. The estimated cost also needs to compare with that of catalytic

917 HDO of biomass-derived feedstocks and fossil fuel-based processes.

918 The experimental studies for electrochemical hydrogenation and hydrogenolysis of biomass-derived

919 oxygenates have primarily focused on the reaction conditions (electrocatalyst, pH, applied potential,

920 current, etc.) and reaction efficiency (selectivity, yield, electrochemical efficiency, etc.) using

921 monometallic or carbon electrodes. Current improvements in reaction efficiency for electrochemical HDO

922 of biomass-derived chemicals have been from the experiences of thermal catalytic HDO and the

923 relationships between experimental reaction conditions and results. However, fundamental understanding

924 of each reaction and role of electrocatalysts will be significantly helpful to develop the electrochemical

925 HDO with high efficiency.

926 A side reaction in electrochemical HDO experimentally proven was the HER that consumes electrons

927 and decreases electrochemical efficiency. For HDO from biomass-derived oxygenates, electrochemical

928 hydrogenolysis, or other reductions were required in highly acidic solutions or at highly negative

929 potentials. However, the HER was readily available in the conditions, and the electrochemical efficiency

930 for electrochemical hydrogenation and hydrogenolysis decreased, while increasing the electrochemical

931 efficiency of the HER. The other challenge was the polymerization or degradation of biomass-derived

932 chemicals in acidic conditions, and it decreased the mass balance of the reactions. The development of

933 advanced electrocatalysts that have dominant electrochemical HDO over the HER in mild conditions will

934 be required. Also, the electrocatalysts should have high reaction rate and stability during the long time of

935 electrochemical HDO. Reaction mechanisms in electrochemical hydrogenation and HDO are rarely

58
Page 59 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

936 studied, and the strong understanding of reaction mechanisms is required to suggest the efficient reaction

937 pathway and electrocatalysts. Advanced electrocatalysts, used for thermal catalytic HDO (bifunctional,

938 bimetallic, or others) can be considered a strong candidate for electrochemical hydrogenation,

939 hydrogenolysis, decarboxylation or other reactions. The lessons from both experimental and the density
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

940 functional theory studies in thermal catalytic HDO will help to understand the reaction mechanisms and to

Green Chemistry Accepted Manuscript


941 suggest proper electrocatalysts for electrochemical HDO.

942 Overall, there have been great achievement in biomass conversion via HDO reactions over

943 bifunctional catalysts. However, there are still rooms to understand more about designing bifunctional

944 catalysts to selectively product target products from biomass. Therefore, efforts to employ rational design

945 of bifunctional catalysts with a high stability will become more important for innovating biomass

946 conversion technologies.

947

948 Acknowledgements

949 This work was supported by a National Research Foundation of Korea (NRF) Grant funded by the

950 Korean Government (Ministry of Education) (No. 2018R1D1A1A09082841).

951

952 References

953 1. G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006, 106, 4044-4098.

954 2. H. Pourzolfaghar, F. Abnisa, W. M. A. Wan Daud and M. K. Aroua, Journal of Analytical and
955 Applied Pyrolysis, 2018, 133, 117-127.

956 3. T. L. Marker, L. G. Felix, M. B. Linck and M. J. Roberts, Environ. Prog. Sustain. Energy, 2012,
957 31, 191-199.

958 4. B. Balagurumurthy, T. S. Oza, T. Bhaskar and D. K. Adhikari, Journal of Material Cycles and
959 Waste Management, 2013, 15, 9-15.

960 5. Y. N. Regmi, J. K. Mann, J. R. McBride, J. Tao, C. E. Barnes, N. Labbé and S. C. Chmely,


961 Catalysis Today, 2018, 302, 190-195.

962 6. R. D. Cortright, ACS National Meeting Book of Abstracts, 2007.

59
Green Chemistry Page 60 of 71
View Article Online
DOI: 10.1039/C9GC01210A

963 7. R. D. Cortright, ACS National Meeting Book of Abstracts, 2008.

964 8. H. Olcay, L. J. Xu, Y. Xu and G. W. Huber, ChemCatChem, 2010, 2, 1420-1424.

965 9. H. Olcay, Y. Xu and G. W. Huber, Green Chem., 2014, 16, 911-924.

966 10. P. Grange and X. Vanhaeren, Catalysis Today, 1997, 36, 375-391.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

967 11. S. De, B. Saha and R. Luque, Bioresource Technology, 2015, 178, 108-118.

968 12. N. Arun, R. V. Sharma and A. K. Dalai, Renewable and Sustainable Energy Reviews, 2015, 48,

Green Chemistry Accepted Manuscript


969 240-255.

970 13. A. M. Robinson, J. E. Hensley and J. W. Medlin, ACS Catal., 2016, 6, 5026-5043.

971 14. X. Li, G. Chen, C. Liu, W. Ma, B. Yan and J. Zhang, Renewable and Sustainable Energy Reviews,
972 2017, 71, 296-308.

973 15. Z. Lin, R. Chen, Z. Qu and J. G. Chen, Green Chem., 2018, 20, 2679-2696.

974 16. T. M. Onn, M. Monai, S. Dai, E. Fonda, T. Montini, X. Pan, G. W. Graham, P. Fornasiero and R.
975 J. Gorte, J. Am. Chem. Soc., 2018, 140, 4841-4848.

976 17. A. Wawrzetz, B. Peng, A. Hrabar, A. Jentys, A. A. Lemonidou and J. A. Lercher, Journal of
977 Catalysis, 2010, 269, 411-420.

978 18. N. Li and G. W. Huber, Journal of Catalysis, 2010, 270, 48-59.

979 19. B. Liu and J. Greeley, The Journal of Physical Chemistry C, 2011, 115, 19702-19709.

980 20. Y. Jing, Q. Xia, X. Liu and Y. Wang, ChemSusChem, 2017, 10, 4817-4823.

981 21. J. Lee, Y. T. Kim and G. W. Huber, Green Chem., 2014, 16, 708-718.

982 22. A. J. R. Hensley, R. Zhang, Y. Wang and J.-S. McEwen, The Journal of Physical Chemistry C,
983 2013, 117, 24317-24328.

984 23. J. Nørskov, T. Bligaard, A. Logadottir, S. Bahn, L. B. Hansen, M. Bollinger, H. Bengaard, B.


985 Hammer, Z. Sljivancanin, M. Mavrikakis, Y. Xu, S. Dahl and C. J. H. Jacobsen, Journal of Catalysis,
986 2002, 209, 275-278.

987 24. A. J. R. Hensley, Y. Hong, R. Zhang, H. Zhang, J. Sun, Y. Wang and J.-S. McEwen, ACS Catal.,
988 2014, 4, 3381-3392.

989 25. Q. Xia, X. Zhuang, M. M.-J. Li, Y.-K. Peng, G. Liu, T.-S. Wu, Y.-L. Soo, X.-Q. Gong, Y. Wang
990 and S. C. E. Tsang, Chemical Communications, 2016, 52, 5160-5163.

991 26. A. Bjelić, M. Grilc, M. Huš and B. Likozar, Chemical Engineering Journal, 2019, 359, 305-320.

992 27. R. Y. Zheng, Y. X. Zhu and J. G. G. Chen, ChemCatChem, 2011, 3, 578-581.

993 28. J. Lee, Y. Xu and G. W. Huber, Appl. Catal. B-Environ., 2013, 140–141, 98-107.

60
Page 61 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

994 29. R. Y. Zheng, M. P. Humbert, Y. X. Zhu and J. G. Chen, Catalysis Science & Technology, 2011, 1,
995 638-643.

996 30. S. Narayanan and R. Unnikrishnan, J. Chem. Soc., Faraday Trans., 1998, 94, 1123-1128.

997 31. J. Wisniak, Hershkow.M, Leibowit.R and S. Stein, Ind. Eng. Chem. Prod. Res. Dev., 1974, 13, 75-
998 79.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

999 32. J. Wisniak, M. Hershkowitz and S. Stein, Ind. Eng. Chem. Prod. Res. Dev., 1974, 13, 232-236.

33. A. Srifa, K. Faungnawakij, V. Itthibenchapong and S. Assabumrungrat, Chemical Engineering

Green Chemistry Accepted Manuscript


1000
1001 Journal, 2015, 278, 249-258.

1002 34. A. Gutierrez, R. K. Kaila, M. L. Honkela, R. Slioor and A. O. I. Krause, Catalysis Today, 2009,
1003 147, 239-246.

1004 35. E. Furimsky, Applied Catalysis A: General, 2000, 199, 147-190.

1005 36. D. C. Elliott and T. R. Hart, Energy Fuels, 2009, 23, 631-637.

1006 37. C. Montassier, J. C. Ménézo, L. C. Hoang, C. Renaud and J. Barbier, Journal of Molecular
1007 Catalysis, 1991, 70, 99-110.

1008 38. E. G. Baker and D. C. Elliott, in Research in Thermochemical Biomass Conversion, eds. A. V.
1009 Bridgwater and J. L. Kuester, Springer Netherlands, Dordrecht, 1988, DOI: 10.1007/978-94-009-2737-
1010 7_67, pp. 883-895.

1011 39. W. Baldauf and U. Balfanz, 1991.

1012 40. D. C. Elliott, T. R. Hart, G. G. Neuenschwander, L. J. Rotness, M. V. Olarte, A. H. Zacher and Y.


1013 Solantausta, Energy Fuels, 2012, 26, 3891-3896.

1014 41. D. C. Elliott, G. G. Neuenschwander and T. R. Hart, ACS Sustainable Chemistry & Engineering,
1015 2013, 1, 389-392.

1016 42. M. Grilc and B. Likozar, Chemical Engineering Journal, 2017, 330, 383-397.

1017 43. B. Hočevar, M. Grilc, M. Huš and B. Likozar, Chemical Engineering Journal, 2019, 359, 1339-
1018 1351.

1019 44. L. Zhou and A. Lawal, Energy Fuels, 2015, 29, 262-272.

1020 45. J. Wildschut, I. Melián-Cabrera and H. J. Heeres, Appl. Catal. B-Environ., 2010, 99, 298-306.

1021 46. O. İ. Şenol, T. R. Viljava and A. O. I. Krause, Applied Catalysis A: General, 2007, 326, 236-244.

1022 47. J.-P. Mikkola, H. Vainio, T. Salmi, R. Sjöholm, T. Ollonqvist and J. Väyrynen, Applied Catalysis
1023 A: General, 2000, 196, 143-155.

1024 48. R. M. West, D. J. Braden and J. A. Dumesic, Journal of Catalysis, 2009, 262, 134-143.

1025 49. G. Yang, E. A. Pidko and E. J. M. Hensen, Journal of Catalysis, 2012, 295, 122-132.

61
Green Chemistry Page 62 of 71
View Article Online
DOI: 10.1039/C9GC01210A

1026 50. G. Li, C. Liu, R. Rohling, E. J. M. Hensen and E. A. Pidko, in Modelling and Simulation in the
1027 Science of Micro- and Meso-Porous Materials, eds. C. R. A. Catlow, V. Van Speybroeck and R. A. van
1028 Santen, Elsevier, 2018, DOI: https://doi.org/10.1016/B978-0-12-805057-6.00007-7, pp. 229-263.

1029 51. D. Stošić, S. Bennici, J.-L. Couturier, J.-L. Dubois and A. Auroux, Catalysis Communications,
1030 2012, 17, 23-28.

1031 52. V. V. Ordomsky, J. van der Schaaf, J. C. Schouten and T. A. Nijhuis, ChemSusChem, 2012, 5,


Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1032 1812-1819.

53. R. Weingarten, G. A. Tompsett, W. C. Conner Jr. and G. W. Huber, Journal of Catalysis, 2011,

Green Chemistry Accepted Manuscript


1033
1034 279, 174-182.

1035 54. R. Weingarten, Y. T. Kim, G. A. Tompsett, A. Fernández, K. S. Han, E. W. Hagaman, W. C.


1036 Conner Jr., J. A. Dumesic and G. W. Huber, Journal of Catalysis, 2013, 304, 123-134.

1037 55. X. Li, K. Peng, X. Liu, Q. Xia and Y. Wang, ChemCatChem, 2017, 9, 2739-2746.

1038 56. Y. Nakagawa, S. Liu, M. Tamura and K. Tomishige, ChemSusChem, 2015, 8, 1114-1132.

1039 57. M. Chia, Y. J. Pagán-Torres, D. Hibbitts, Q. Tan, H. N. Pham, A. K. Datye, M. Neurock, R. J.


1040 Davis and J. A. Dumesic, J. Am. Chem. Soc., 2011, 133, 12675-12689.

1041 58. Y. T. Kim, J. A. Dumesic and G. W. Huber, Journal of Catalysis, 2013, 304, 72-85.

1042 59. P. U. Karanjkar, S. P. Burt, X. Chen, K. J. Barnett, M. R. Ball, M. D. Kumbhalkar, X. Wang, J. B.


1043 Miller, I. Hermans, J. A. Dumesic and G. W. Huber, Catalysis Science & Technology, 2016, 6, 7841-7851.

1044 60. O. M. Daniel, A. DeLaRiva, E. L. Kunkes, A. K. Datye, J. A. Dumesic and R. J. Davis,


1045 ChemCatChem, 2010, 2, 1107-1114.

1046 61. E. L. Kunkes, D. A. Simonetti, J. A. Dumesic, W. D. Pyrz, L. E. Murillo, J. G. Chen and D. J.


1047 Buttrey, Journal of Catalysis, 2008, 260, 164-177.

1048 62. C. Ju, M. Li, Y. Fang and T. Tan, Green Chem., 2018, 20, 4492-4499.

1049 63. S. Liu, Y. Amada, M. Tamura, Y. Nakagawa and K. Tomishige, Green Chem., 2014, 16, 617-626.

1050 64. J. He, S. P. Burt, M. Ball, D. Zhao, I. Hermans, J. A. Dumesic and G. W. Huber, ACS Catal., 2018,
1051 8, 1427-1439.

1052 65. M. Ruff, N. Takehiro, P. Liu, J. K. Nørskov and R. J. Behm, ChemPhysChem, 2007, 8, 2068-2071.

1053 66. N. J. Divins, I. Angurell, C. Escudero, V. Pérez-Dieste and J. Llorca, Science, 2014, 346, 620-623.

1054 67. J. R. Regalbuto, Science, 2009, 325, 822-824.

1055 68. N. Li, G. A. Tompsett and G. W. Huber, ChemSusChem, 2010, 3, 1154-1157.

1056 69. C. Chatterjee, F. Pong and A. Sen, Green Chem., 2015, 17, 40-71.

62
Page 63 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

1057 70. P. Gallezot, N. Nicolaus, G. Flèche, P. Fuertes and A. Perrard, Journal of Catalysis, 1998, 180,
1058 51-55.

1059 71. L. Vilcocq, A. Cabiac, C. Especel, S. Lacombe and D. Duprez, Journal of Catalysis, 2014, 320,
1060 16-25.

1061 72. N. M. Eagan, J. P. Chada, A. M. Wittrig, J. S. Buchanan, J. A. Dumesic and G. W. Huber, Joule,
1062 2017, 1, 178-199.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1063 73. B. M. Moreno, N. Li, J. Lee, G. W. Huber and M. T. Klein, RSC Adv., 2013, 3, 23769-23784.

Green Chemistry Accepted Manuscript


1064 74. Y. Weng, T. Wang, S. Qiu, C. Wang, L. Ma, Q. Zhang, L. Chen, Y. Li, F. Sun and Q. Zhang,
1065 ChemCatChem, 2017, 9, 774-781.

1066 75. E. E. Kwon, Y. T. Kim, H. J. Kim, K.-Y. Andrew Lin, K.-H. Kim, J. Lee and G. W. Huber,
1067 Journal of Environmental Management, 2018, 227, 329-334.

1068 76. J. Lee, I. Ro, H. J. Kim, Y. T. Kim, E. E. Kwon and G. W. Huber, Process Safety and
1069 Environmental Protection, 2018, 115, 2-7.

1070 77. B. Op de Beeck, M. Dusselier, J. Geboers, J. Holsbeek, E. Morre, S. Oswald, L. Giebeler and B. F.
1071 Sels, Energy & Environmental Science, 2015, 8, 230-240.

1072 78. S. Liu, M. Tamura, Y. Nakagawa and K. Tomishige, ACS Sustainable Chemistry & Engineering,
1073 2014, 2, 1819-1827.

1074 79. N. Li, G. A. Tompsett, T. Y. Zhang, J. A. Shi, C. E. Wyman and G. W. Huber, Green Chem., 2011,
1075 13, 91-101.

1076 80. K. Murata, Y. Liu, M. Inaba and I. Takahara, Catalysis Letters, 2010, 140, 8-13.

1077 81. Y. Kato and Y. Sekine, Catalysis Letters, 2013, 143, 418-423.

1078 82. R. M. West, M. H. Tucker, D. J. Braden and J. A. Dumesic, Catalysis Communications, 2009, 10,
1079 1743-1746.

1080 83. Q. Zhang, T. Jiang, B. Li, T. Wang, X. Zhang, Q. Zhang and L. Ma, ChemCatChem, 2012, 4,
1081 1084-1087.

1082 84. Q. Zhang, T. Wang, Y. Xu, Q. Zhang and L. Ma, Energy Conversion and Management, 2014, 77,
1083 262-268.

1084 85. J. Duan, Y. T. Kim, H. Lou and G. W. Huber, Catalysis Today, 2014, 234, 66-74.

1085 86. C. J. Barrett, J. N. Chheda, G. W. Huber and J. A. Dumesic, Appl. Catal. B-Environ., 2006, 66,
1086 111-118.

1087 87. R. He, X. P. Ye, B. C. English and J. A. Satrio, Bioresource Technology, 2009, 100, 5305-5311.

1088 88. N. Lohitharn and B. H. Shanks, Catalysis Communications, 2009, 11, 96-99.

1089 89. S. Sitthisa and D. E. Resasco, Catalysis Letters, 2011, 141, 784-791.

63
Green Chemistry Page 64 of 71
View Article Online
DOI: 10.1039/C9GC01210A

1090 90. B. Saha, C. M. Bohn and M. M. Abu-Omar, ChemSusChem, 2014, 7, 3095-3101.

1091 91. Y. Zu, P. Yang, J. Wang, X. Liu, J. Ren, G. Lu and Y. Wang, Appl. Catal. B-Environ., 2014, 146,
1092 244-248.

1093 92. Y.-B. Huang, M.-Y. Chen, L. Yan, Q.-X. Guo and Y. Fu, ChemSusChem, 2014, 7, 1068-1072.

1094 93. P. Yang, Q. Xia, X. Liu and Y. Wang, Journal of Energy Chemistry, 2016, 25, 1015-1020.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1095 94. M. S. Gyngazova, L. Negahdar, L. C. Blumenthal and R. Palkovits, Chemical Engineering


Science, 2017, 173, 455-464.

Green Chemistry Accepted Manuscript


1096

1097 95. Y. Román-Leshkov, C. J. Barrett, Z. Y. Liu and J. A. Dumesic, Nature, 2007, 447, 982-985.

1098 96. T. Thananatthanachon and T. B. Rauchfuss, Angew. Chem. Int. Ed., 2010, 49, 6616-6618.

1099 97. J. Jae, W. Zheng, R. F. Lobo and D. G. Vlachos, ChemSusChem, 2013, 6, 1158-1162.

1100 98. J. Jae, W. Zheng, A. M. Karim, W. Guo, R. F. Lobo and D. G. Vlachos, ChemCatChem, 2014, 6,
1101 848-856.

1102 99. P. Yang, Q. Xia, X. Liu and Y. Wang, Fuel, 2017, 187, 159-166.

1103 100. G. Li, N. Li, J. Yang, A. Wang, X. Wang, Y. Cong and T. Zhang, Bioresource Technology, 2013,
1104 134, 66-72.

1105 101. G. Li, N. Li, S. Li, A. Wang, Y. Cong, X. Wang and T. Zhang, Chemical Communications, 2013,
1106 49, 5727-5729.

1107 102. S. Li, N. Li, G. Li, A. Wang, Y. Cong, X. Wang and T. Zhang, Catalysis Today, 2014, 234, 91-99.

1108 103. S. Li, N. Li, G. Li, L. Li, A. Wang, Y. Cong, X. Wang, G. Xu and T. Zhang, Appl. Catal. B-
1109 Environ., 2015, 170-171, 124-134.

1110 104. J. Yang, S. Li, L. Zhang, X. Liu, J. Wang, X. Pan, N. Li, A. Wang, Y. Cong, X. Wang and T.
1111 Zhang, Appl. Catal. B-Environ., 2017, 201, 266-277.

1112 105. Y. Shao, Q. Xia, X. Liu, G. Lu and Y. Wang, ChemSusChem, 2015, 8, 1761-1767.

1113 106. S. Dutta and B. Saha, ACS Catal., 2017, 7, 5491-5499.

1114 107. S. Liu, S. Dutta, W. Zheng, N. S. Gould, Z. Cheng, B. Xu, B. Saha and D. G. Vlachos,
1115 ChemSusChem, 2017, 10, 3225-3234.

1116 108. Q.-N. Xia, Q. Cuan, X.-H. Liu, X.-Q. Gong, G.-Z. Lu and Y.-Q. Wang, Angew. Chem. Int. Ed.,
1117 2014, 53, 9755-9760.

1118 109. F. Xue, D. Ma, T. Tong, X. Liu, Y. Hu, Y. Guo and Y. Wang, ACS Sustainable Chemistry &
1119 Engineering, 2018, 6, 13107-13113.

1120 110. J. Zhang, K. Matsubara, G.-N. Yun, H. Zheng, A. Takagaki, R. Kikuchi and S. T. Oyama, Applied
1121 Catalysis A: General, 2017, 548, 39-46.

64
Page 65 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

1122 111. H. Olcay, A. V. Subrahmanyam, R. Xing, J. Lajoie, J. A. Dumesic and G. W. Huber, Energy &
1123 Environmental Science, 2013, 6, 205-216.

1124 112. C. Zhao, J. He, A. A. Lemonidou, X. Li and J. A. Lercher, Journal of Catalysis, 2011, 280, 8-16.

1125 113. A. L. Jongerius, R. Jastrzebski, P. C. A. Bruijnincx and B. M. Weckhuysen, Journal of Catalysis,


1126 2012, 285, 315-323.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1127 114. P. M. Mortensen, J.-D. Grunwaldt, P. A. Jensen and A. D. Jensen, ACS Catal., 2013, 3, 1774-1785.

115. B. Güvenatam, O. Kurşun, E. H. J. Heeres, E. A. Pidko and E. J. M. Hensen, Catalysis Today,

Green Chemistry Accepted Manuscript


1128
1129 2014, 233, 83-91.

1130 116. M. Huš, A. Bjelić, M. Grilc and B. Likozar, Journal of Catalysis, 2018, 358, 8-18.

1131 117. M. Shetty, K. Murugappan, T. Prasomsri, W. H. Green and Y. Román-Leshkov, Journal of


1132 Catalysis, 2015, 331, 86-97.

1133 118. V. O. O. Gonçalves, C. Ciotonea, S. Arrii-Clacens, N. Guignard, C. Roudaut, J. Rousseau, J.-M.


1134 Clacens, S. Royer and F. Richard, Appl. Catal. B-Environ., 2017, 214, 57-66.

1135 119. D. P. Gamliel, S. Karakalos and J. A. Valla, Applied Catalysis A: General, 2018, 559, 20-29.

1136 120. N. Yan, C. Zhao, P. J. Dyson, C. Wang, L.-t. Liu and Y. Kou, ChemSusChem, 2008, 1, 626-629.

1137 121. A. L. Jongerius, P. C. A. Bruijnincx and B. M. Weckhuysen, Green Chem., 2013, 15, 3049-3056.

1138 122. J. Kong, M. He, J. A. Lercher and C. Zhao, Chemical Communications, 2015, 51, 17580-17583.

1139 123. Z. Luo, Y. Wang, M. He and C. Zhao, Green Chem., 2016, 18, 433-441.

1140 124. Z. Luo and C. Zhao, Catalysis Science & Technology, 2016, 6, 3476-3484.

1141 125. T. Guo, Q. Xia, Y. Shao, X. Liu and Y. Wang, Applied Catalysis A: General, 2017, 547, 30-36.

1142 126. L. Dong, L.-L. Yin, Q. Xia, X. Liu, X.-Q. Gong and Y. Wang, Catalysis Science & Technology,
1143 2018, 8, 735-745.

1144 127. Q. Xia, Z. Chen, Y. Shao, X. Gong, H. Wang, X. Liu, S. F. Parker, X. Han, S. Yang and Y. Wang,
1145 Nature Communications, 2016, 7, 11162.

1146 128. Y. Shao, Q. Xia, L. Dong, X. Liu, X. Han, S. F. Parker, Y. Cheng, L. L. Daemen, A. J. Ramirez-
1147 Cuesta, S. Yang and Y. Wang, Nature Communications, 2017, 8, 16104.

1148 129. J. Yang, C. L. Williams, A. Ramasubramaniam and P. J. Dauenhauer, Green Chem., 2014, 16,
1149 675-682.

1150 130. X. Li, J. Xing, M. Zhou, H. Zhang, H. Huang, C. Zhang, L. Song and X. Li, Catalysis
1151 Communications, 2014, 56, 123-127.

1152 131. C. Zhang, J. Xing, L. Song, H. Xin, S. Lin, L. Xing and X. Li, Catalysis Today, 2014, 234, 145-
1153 152.

65
Green Chemistry Page 66 of 71
View Article Online
DOI: 10.1039/C9GC01210A

1154 132. C. R. Lee, J. S. Yoon, Y.-W. Suh, J.-W. Choi, J.-M. Ha, D. J. Suh and Y.-K. Park, Catalysis
1155 Communications, 2012, 17, 54-58.

1156 133. J. Chang, T. Danuthai, S. Dewiyanti, C. Wang and A. Borgna, ChemCatChem, 2013, 5, 3041-
1157 3049.

1158 134. Z. Cai, F. Wang, X. Zhang, R. Ahishakiye, Y. Xie and Y. Shen, Molecular Catalysis, 2017, 441,
1159 28-34.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1160 135. P. T. M. Do, A. J. Foster, J. Chen and R. F. Lobo, Green Chem., 2012, 14, 1388-1397.

Green Chemistry Accepted Manuscript


1161 136. A. J. Foster, P. T. M. Do and R. F. Lobo, Topics in Catalysis, 2012, 55, 118-128.

1162 137. C. Zhang, J. Qi, J. Xing, S.-F. Tang, L. Song, Y. Sun, C. Zhang, H. Xin and X. Li, RSC Adv., 2016,
1163 6, 104398-104406.

1164 138. A. M. Barrios, C. A. Teles, P. M. de Souza, R. C. Rabelo-Neto, G. Jacobs, B. H. Davis, L. E. P.


1165 Borges and F. B. Noronha, Catalysis Today, 2018, 302, 115-124.

1166 139. H. Wang, H. Ruan, H. Pei, H. Wang, X. Chen, M. P. Tucker, J. R. Cort and B. Yang, Green
1167 Chem., 2015, 17, 5131-5135.

1168 140. D. Carpenter, T. Westover, D. Howe, S. Deutch, A. Starace, R. Emerson, S. Hernandez, D.


1169 Santosa, C. Lukins and I. Kutnyakov, Biomass and Bioenergy, 2017, 96, 142-151.

1170 141. D. C. Elliott, T. R. Hart, G. G. Neuenschwander, L. J. Rotness and A. H. Zacher, Environ. Prog.
1171 Sustain. Energy, 2009, 28, 441-449.

1172 142. D. Howe, T. Westover, D. Carpenter, D. Santosa, R. Emerson, S. Deutch, A. Starace, I.


1173 Kutnyakov and C. Lukins, Energy Fuels, 2015, 29, 3188-3197.

1174 143. R. Kamalakanta, K. J. Barnett and G. W. Huber, Energy Technology, 2017, 5, 80-93.

1175 144. G. Kim, J. Seo, J.-W. Choi, J. Jae, J.-M. Ha, D. J. Suh, K.-Y. Lee, J.-K. Jeon and J.-K. Kim,
1176 Catalysis Today, 2018, 303, 130-135.

1177 145. M. V. Olarte, A. H. Zacher, A. B. Padmaperuma, S. D. Burton, H. M. Job, T. L. Lemmon, M. S.


1178 Swita, L. J. Rotness, G. N. Neuenschwander, J. G. Frye and D. C. Elliott, Topics in Catalysis, 2016, 59,
1179 55-64.

1180 146. M. V. Olarte, A. B. Padmaperuma, J. R. Ferrell, E. D. Christensen, R. T. Hallen, R. B. Lucke, S. D.


1181 Burton, T. L. Lemmon, M. S. Swita, G. Fioroni, D. C. Elliott and C. Drennan, Fuel, 2017, 202, 620-630.

1182 147. W. Yin, A. Kloekhorst, R. H. Venderbosch, M. V. Bykova, S. A. Khromova, V. A. Yakovlev and


1183 H. J. Heeres, Catalysis Science & Technology, 2016, 6, 5899-5915.

1184 148. A. Sanna, T. P. Vispute and G. W. Huber, Appl. Catal. B-Environ., 2015, 165, 446-456.

1185 149. I. Kim, A. A. Dwiatmoko, J.-W. Choi, D. J. Suh, J. Jae, J.-M. Ha and J.-K. Kim, Journal of
1186 Industrial and Engineering Chemistry, 2017, 56, 74-81.

66
Page 67 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

1187 150. K. Treusch, N. Schwaiger, K. Schlackl, R. Nagl, A. Rollett, M. Schadler, B. Hammerschlag, J.


1188 Ausserleitner, A. Huber, P. Pucher and M. Siebenhofer, Reaction Chemistry & Engineering, 2018, 3, 258-
1189 266.

1190 151. P. A. Meyer, L. J. Snowden-Swan, K. G. Rappé, S. B. Jones, T. L. Westover and K. G. Cafferty,


1191 Energy Fuels, 2016, 30, 9427-9439.

1192 152. D. C. Elliott, Energy Fuels, 2007, 21, 1792-1815.


Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1193 153. J. Luo, Publicly Accessible Penn Dissertations, 2016, 2449.

Green Chemistry Accepted Manuscript


1194 154. H. Pucher, N. Schwaiger, R. Feiner, L. Ellmaier, P. Pucher, B. S. Chernev and M. Siebenhofer,
1195 Green Chem., 2015, 17, 1291-1298.

1196 155. N. Schwaiger, D. C. Elliott, J. Ritzberger, H. Wang, P. Pucher and M. Siebenhofer, Green Chem.,
1197 2015, 17, 2487-2494.

1198 156. J. Neumann, N. Jäger, A. Apfelbacher, R. Daschner, S. Binder and A. Hornung, Biomass and
1199 Bioenergy, 2016, 89, 91-97.

1200 157. S. Lestari, P. Maki-Arvela, J. Beltramini, G. Q. M. Lu and D. Y. Murzin, ChemSusChem, 2009, 2,


1201 1109-1119.

1202 158. B. Smith, H. C. Greenwell and A. Whiting, Energy & Environmental Science, 2009, 2, 262-271.

1203 159. I. M. Atadashi, M. K. Aroua, A. R. Abdul Aziz and N. M. N. Sulaiman, Renewable and
1204 Sustainable Energy Reviews, 2012, 16, 3275-3285.

1205 160. C. Zhao, T. Brück and J. A. Lercher, Green Chem., 2013, 15, 1720-1739.

1206 161. E. Santillan-Jimenez and M. Crocker, Journal of Chemical Technology & Biotechnology, 2012, 87,
1207 1041-1050.

1208 162. M. Snåre and D. Y. Murzin, Industrial & Engineering Chemistry Research, 2006, 45, 6875-6875.

1209 163. J. C. Serrano-Ruiz, R. M. West and J. A. Dumesic, Annual Review of Chemical and Biomolecular
1210 Engineering, 2010, 1, 79-100.

1211 164. M. Snåre, I. Kubickova, P. Maki-Arvela, K. Eranen and D. Y. Murzin, Industrial & Engineering
1212 Chemistry Research, 2006, 45, 5708-5715.

1213 165. S. Chen, G. Zhou and C. Miao, Renewable and Sustainable Energy Reviews, 2019, 101, 568-589.

1214 166. B. Peng, Y. Yao, C. Zhao and J. A. Lercher, Angew. Chem. Int. Ed., 2012, 51, 2072-2075.

1215 167. J. Li, S. Wang, H.-Y. Liu, H.-j. Zhou and Y. Fu, ChemistrySelect, 2017, 2, 33-41.

1216 168. N. Chen, N. Wang, Y. Ren, H. Tominaga and E. W. Qian, Journal of Catalysis, 2017, 345, 124-
1217 134.

1218 169. Q. Xia, X. Zhuang, M. M. Li, Y. K. Peng, G. Liu, T. S. Wu, Y. L. Soo, X. Q. Gong, Y. Wang and
1219 S. C. E. Tsang, Chemical Communications, 2016, 52, 5160-5163.

67
Green Chemistry Page 68 of 71
View Article Online
DOI: 10.1039/C9GC01210A

1220 170. J. Duan, J. Han, H. Sun, P. Chen, H. Lou and X. Zheng, Catalysis Communications, 2012, 17, 76-
1221 80.

1222 171. K. Murata, Y. Liu, M. Inaba and I. Takahara, Energy Fuels, 2010, 24, 2404-2409.

1223 172. C. Wang, Z. Tian, L. Wang, R. Xu, Q. Liu, W. Qu, H. Ma and B. Wang, ChemSusChem, 2012, 5,
1224 1974-1983.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1225 173. W. Song, C. Zhao and J. A. Lercher, Chemistry - A European Journal, 2013, 19, 9833-9842.

174. D. R. Vardon, B. K. Sharma, H. Jaramillo, D. Kim, J. K. Choe, P. N. Ciesielski and T. J.

Green Chemistry Accepted Manuscript


1226
1227 Strathmann, Green Chem., 2014, 16, 1507-1520

1228 175. Z. Zhang, Q. Yang, H. Chen, K. Chen, X. Lu, P. Ouyang, J. Fu and J. G. Chen, Green Chem.,
1229 2018, 20, 197-205.

1230 176. J. Lee, D. H. K. Jackson, T. Li, R. E. Winans, J. A. Dumesic, T. F. Kuech and G. W. Huber,
1231 Energy & Environmental Science, 2014, 7, 1657-1660.

1232 177. B. J. O'Neill, D. H. K. Jackson, A. J. Crisci, C. A. Farberow, F. Shi, A. C. Alba-Rubio, J. Lu, P. J.


1233 Dietrich, X. Gu, C. L. Marshall, P. C. Stair, J. W. Elam, J. T. Miller, F. H. Ribeiro, P. M. Voyles, J.
1234 Greeley, M. Mavrikakis, S. L. Scott, T. F. Kuech and J. A. Dumesic, Angew. Chem. Int. Ed., 2013, 52,
1235 13808-13812.

1236 178. P. M. Mortensen, D. Gardini, H. W. P. de Carvalho, C. D. Damsgaard, J.-D. Grunwaldt, P. A.


1237 Jensen, J. B. Wagner and A. D. Jensen, Catalysis Science & Technology, 2014, 4, 3672-3686.

1238 179. R. M. Ravenelle, J. R. Copeland, W. G. Kim, J. C. Crittenden and C. Sievers, ACS Catal., 2011, 1,
1239 552-561.

1240 180. A. Centeno, O. David, C. Vanbellinghen, R. Maggi and B. Delmon, in Developments in


1241 Thermochemical Biomass Conversion, eds. A. V. Bridgwater and D. G. B. Boocock, Springer-Science &
1242 Business Media, 1997, pp. 589-601.

1243 181. H. T. Li, Y. X. Zhao, C. G. Gao, Y. Z. Wang, Z. J. Sun and X. Y. Liang, Chemical Engineering
1244 Journal, 2012, 181, 501-507.

1245 182. R. M. Ravenelle, J. R. Copeland, A. H. Van Pelt, J. C. Crittenden and C. Sievers, Topics in
1246 Catalysis, 2012, 55, 162-174.

1247 183. R. M. Ravenelle, F. Z. Diallo, J. C. Crittenden and C. Sievers, ChemCatChem, 2012, 4, 492-494.

1248 184. W. N. P. van der Graaff, C. H. L. Tempelman, F. C. Hendriks, J. Ruiz-Martinez, S. Bals, B. M.


1249 Weckhuysen, E. A. Pidko and E. J. M. Hensen, Applied Catalysis A: General, 2018, 564, 113-122.

1250 185. B. J. O’Neill, D. H. K. Jackson, J. Lee, C. Canlas, P. C. Stair, C. L. Marshall, J. W. Elam, T. F.


1251 Kuech, J. A. Dumesic and G. W. Huber, ACS Catal., 2015, 5, 1804-1825.

1252 186. B. J. O'Neill, C. Sener, D. H. K. Jackson, T. F. Kuech and J. A. Dumesic, ChemSusChem, 2014, 7,
1253 3247-3251.

68
Page 69 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

1254 187. B. J. O'Neill, J. T. Miller, P. J. Dietrich, F. G. Sollberger, F. H. Ribeiro and J. A. Dumesic,


1255 ChemCatChem, 2014, 6, 2493-2496.

1256 188. H. Zhang, C. Canlas, A. Jeremy Kropf, J. W. Elam, J. A. Dumesic and C. L. Marshall, Journal of
1257 Catalysis, 2015, 326, 172-181.

1258 189. H. Zhang, Y. Lei, A. J. Kropf, G. Zhang, J. W. Elam, J. T. Miller, F. Sollberger, F. Ribeiro, M. C.
1259 Akatay, E. A. Stach, J. A. Dumesic and C. L. Marshall, Journal of Catalysis, 2014, 317, 284-292.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1260 190. J. Greeley, Electrochimica Acta, 2010, 55, 5545-5550.

Green Chemistry Accepted Manuscript


1261 191. Y. J. Pagán-Torres, J. M. R. Gallo, D. Wang, H. N. Pham, J. A. Libera, C. L. Marshall, J. W. Elam,
1262 A. K. Datye and J. A. Dumesic, ACS Catal., 2011, 1, 1234-1245.

1263 192. S. J. Tauster, S. C. Fung and R. L. Garten, J. Am. Chem. Soc., 1978, 100, 170-175.

1264 193. S. J. Tauster, Acc. Chem. Res., 1987, 20, 389-394.

1265 194. J. Santos, J. Phillips and J. A. Dumesic, Journal of Catalysis, 1983, 81, 147-167.

1266 195. J. Lee, S. P. Burt, C. A. Carrero, A. C. Alba-Rubio, I. Ro, B. J. O’Neill, H. J. Kim, D. H. K.


1267 Jackson, T. F. Kuech, I. Hermans, J. A. Dumesic and G. W. Huber, Journal of Catalysis, 2015, 330, 19-
1268 27.

1269 196. H. N. Pham, A. E. Anderson, R. L. Johnson, K. Schmidt-Rohr and A. K. Datye, Angew. Chem. Int.
1270 Ed., 2012, 51, 13163-13167.

1271 197. P. A. Zapata, J. Faria, M. P. Ruiz, R. E. Jentoft and D. E. Resasco, J. Am. Chem. Soc., 2012, 134,
1272 8570-8578.

1273 198. H. Xiong, T. J. Schwartz, N. I. Andersen, J. A. Dumesic and A. K. Datye, Angew. Chem. Int. Ed.,
1274 2015, 54, 7939-7943.

1275 199. F. Héroguel, L. Silvioli, Y.-P. Du and J. S. Luterbacher, Journal of Catalysis, 2018, 358, 50-61.

1276 200. S. K. Green, G. A. Tompsett, H. J. Kim, W. B. Kim and G. W. Huber, ChemSusChem, 2012, 5,
1277 2410-2420.

1278 201. X. H. Chadderdon, D. J. Chadderdon, J. E. Matthiesen, Y. Qiu, J. M. Carraher, J.-P. Tessonnier


1279 and W. Li, J. Am. Chem. Soc., 2017, 139, 14120-14128.

1280 202. Y. Kwon, Y. Y. Birdja, S. Raoufmoghaddam and M. T. M. Koper, ChemSusChem, 2015, 8, 1745-
1281 1751.

1282 203. S. Jung, A. N. Karaiskakis and E. J. Biddinger, Catalysis Today, 2019, 323, 26-34.

1283 204. S. Jung and E. J. Biddinger, Energy Technology, 2018, 6, 1370-1379.

1284 205. G. Chamoulaud, D. Floner, C. Moinet, C. Lamy and E. M. Belgsir, Electrochimica Acta, 2001, 46,
1285 2757-2760.

69
Green Chemistry Page 70 of 71
View Article Online
DOI: 10.1039/C9GC01210A

1286 206. P. Parpot, A. P. Bettencourt, G. Chamoulaud, K. B. Kokoh and E. M. Belgsir, Electrochimica Acta,
1287 2004, 49, 397-403.

1288 207. B. Zhao, M. Chen, Q. Guo and Y. Fu, Electrochimica Acta, 2014, 135, 139-146.

1289 208. F. Wang, M. Xu, L. Wei, Y. Wei, Y. Hu, W. Fang and C. G. Zhu, Electrochimica Acta, 2015, 153,
1290 170-174.
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1291 209. Z. Li, S. Kelkar, C. H. Lam, K. Luczek, J. E. Jackson, D. J. Miller and C. M. Saffron,
1292 Electrochimica Acta, 2012, 64, 87-93.

Green Chemistry Accepted Manuscript


1293 210. P. Nilges and U. Schröder, Energy & Environmental Science, 2013, 6, 2925-2931.

1294 211. S. Jung and E. J. Biddinger, ACS Sustainable Chemistry & Engineering, 2016, 4, 6500-6508.

1295 212. S. K. Green, J. Lee, H. J. Kim, G. A. Tompsett, W. B. Kim and G. W. Huber, Green Chem., 2013,
1296 15, 1869-1879.

1297 213. Y. Nakagawa, M. Tamura and K. Tomishige, ACS Catal., 2013, 3, 2655-2668.

1298 214. Y. Kwon, E. de Jong, S. Raoufmoghaddam and M. T. M. Koper, ChemSusChem, 2013, 6, 1659-
1299 1667.

1300 215. J. J. Roylance, T. W. Kim and K.-S. Choi, ACS Catal., 2016, 6, 1840-1847.

1301 216. J. J. Roylance and K.-S. Choi, Green Chem., 2016, 18, 2956-2960.

1302 217. Y.-R. Zhang, B.-X. Wang, L. Qin, Q. Li and Y.-M. Fan, Green Chem., 2019, 21, 1108-1113.

1303 218. Z. Li, M. Garedew, C. H. Lam, J. E. Jackson, D. J. Miller and C. M. Saffron, Green Chem., 2012,
1304 14, 2540-2549.

1305 219. C. H. Lam, C. B. Lowe, Z. Li, K. N. Longe, J. T. Rayburn, M. A. Caldwell, C. E. Houdek, J. B.


1306 Maguire, C. M. Saffron, D. J. Miller and J. E. Jackson, Green Chem., 2015, 17, 601-609.

1307 220. J. E. Jackson, C. H. Lam, C. M. Saffron and D. J. Miller, US Pat., US 9951431 B2, 2018.

1308 221. M. Garedew, D. Young-Farhat, J. E. Jackson and C. M. Saffron, ACS Sustainable Chemistry &
1309 Engineering, 2019, 7, 8375-8386.

1310 222. W. Sauter, O. L. Bergmann and U. Schröder, ChemSusChem, 2017, 10, 3105-3110.

1311 223. K. J. Carroll, T. Burger, L. Langenegger, S. Chavez, S. T. Hunt, Y. Román-Leshkov and F. R.


1312 Brushett, ChemSusChem, 2016, 9, 1904-1910.

1313 224. P. Nilges, T. R. dos Santos, F. Harnisch and U. Schröder, Energy & Environmental Science, 2012,
1314 5, 5231-5235.

1315 225. L. Xin, Z. Zhang, J. Qi, D. J. Chadderdon, Y. Qiu, K. M. Warsko and W. Li, ChemSusChem, 2013,
1316 6, 674-686.

70
Page 71 of 71 Green Chemistry
View Article Online
DOI: 10.1039/C9GC01210A

1317 226. N. Singh, M.-T. Nguyen, D. C. Cantu, B. L. Mehdi, N. D. Browning, J. L. Fulton, J. Zheng, M.
1318 Balasubramanian, O. Y. Gutiérrez, V.-A. Glezakou, R. Rousseau, N. Govind, D. M. Camaioni, C. T.
1319 Campbell and J. A. Lercher, Journal of Catalysis, 2018, 368, 8-19.

1320 227. N. Singh, Y. Song, O. Y. Gutiérrez, D. M. Camaioni, C. T. Campbell and J. A. Lercher, ACS
1321 Catal., 2016, 6, 7466-7470.

1322 228. C. J. Bondue, F. Calle-Vallejo, M. C. Figueiredo and M. T. M. Koper, Nature Catalysis, 2019, 2,
Published on 10 May 2019. Downloaded by California Institute of Technology on 5/13/2019 6:04:43 PM.

1323 243-250.

229. U. Sanyal, Y. Song, N. Singh, J. L. Fulton, J. Herranz, A. Jentys, O. Y. Gutiérrez and J. A.

Green Chemistry Accepted Manuscript


1324
1325 Lercher, ChemCatChem, 2019, 11, 575-582.

1326 230. Y. Song, O. Y. Gutiérrez, J. Herranz and J. A. Lercher, Appl. Catal. B-Environ., 2016, 182, 236-
1327 246.

1328 231. Y. Song, U. Sanyal, D. Pangotra, J. D. Holladay, D. M. Camaioni, O. Y. Gutiérrez and J. A.


1329 Lercher, Journal of Catalysis, 2018, 359, 68-75.

1330 232. S. Kim, Y. F. Tsang, E. E. Kwon, K.-Y. A. Lin and J. Lee, Korean Journal of Chemical
1331 Engineering, 2019, 36, 1-11.

1332 233. M. J. Gilkey and B. Xu, ACS Catal., 2016, 6, 1420-1436.

1333 234. W. Hao, W. Li, X. Tang, X. Zeng, Y. Sun, S. Liu and L. Lin, Green Chem., 2016, 18, 1080-1088.

1334 235. B. Li, L. Li, H. Sun and C. Zhao, ACS Sustainable Chemistry & Engineering, 2018, 6, 12096-
1335 12103.
1336

71

Anda mungkin juga menyukai