Ethanol
Ethanol
Contents
1. Introduction 91
2. Methods for ethanol production 94
2.1 Production of ethanol via biomass fermentation 94
2.2 Production of ethanol from ethylene 98
2.3 Direct thermochemical conversion of syngas to ethanol 98
2.4 Fermentation of syngas to ethanol 108
2.5 Production of ethanol from syngas via acetic acid 111
2.6 Production of ethanol from syngas via dimethyl oxalate 119
2.7 Comparison of different methods for ethanol production 125
3. Catalytic conversion of ethanol to chemicals and advanced fuels 128
3.1 Catalytic conversion of ethanol to chemicals 128
3.2 Catalytic conversion of ethanol to advanced fuels 149
4. Conclusions and outlook 168
Acknowledgments 174
References 174
About the authors 190
Abstract
The declining petroleum resource and deteriorating environmental issues caused by
the excessive consumption of primary energy resources have stimulated the explora-
tion of renewable chemicals and fuels. Many countries have set short and/or long
term plans to cut the CO2 emission and increase the renewable portions. Ethanol is
regarded as a most promising chemical to fulfill these goals. In the past decades, great
achievements have been made in ethanol production, and the flourish of ethanol
also extends its application to fine chemicals and advanced fuels. In this chapter,
the recent advances and developments in ethanol production from biomass and
syngas are summarized and compared from the viewpoint of fundamental study
to industrial applications. Afterwards, the potential catalytic routes for upgrading
ethanol to value-added chemicals (acetaldehyde, ethyl acetate, propylene, acetone,
1,3-butadiene and isobutene) and advanced fuels (H2, butanol and hydrocarbons) are
introduced with respect to the catalyst developments and engineering integrations.
An outlook is finally provided to highlight the challenges and opportunities associated
with this research area.
Abbreviations
1,3-BD 1,3-butadiene
ABE acetone, butanol and ethanol
AC active carbon
ACK acetate kinase
ACS acetyl-CoA synthase
ADH alcohol dehydrogenase
ADHE butyraldehyde/butanol dehydrogenase
ADP adenosine diphosphate
AOR aldehyde oxidoreductase
ATP adenosine triphosphate
CNTs carbon nanotubes
CODH CO dehydrogenase
Co-FeS-P corrinoid iron-sulfur protein
DFT density functional theory
DME dimethyl ether
DMO dimethyl oxalate
EMP Embden–Meyerhof–Parnas
ENO enolase
FBPA fructose bisphosphate aldolase
FDH formate dehydrogenase
FTS formyl-THF synthetase
GAPDH glyceraldehydes-3-phosphate dehydrogenase
GHSV gas hourly space velocity
HK hexokinase
HT hydrotalcite
LDHs layered double hydroxide
MOR, ZSM-5, MFI zeolites
MPV Meerwein–Ponndorf–Verley
MTG methanol to gasoline
MTR methyltransferase
NaOEt sodium ethoxide
OMC ordered mesoporous carbon
PDC pyruvate decarboxylase
PFK phosphofructokinase
PGI phosphoglucoisomerase
PGK phosphoglycerate kinase
PGM phosphoglyceromutase
PYK pyruvate kinase
scCO2 supercritical CO2
STY time space yield
syngas synthesis gas
Synthesis of ethanol and its catalytic conversion 91
THF tetrahydrofolate
TOF turnover frequency
TOS time on stream
TPI triose phosphate isomerase
WHSV weight hourly space velocity
WL Wood–Ljungdahl
1. Introduction
In the past century, fossil resources made a great contribution to
the development of global economy, and significantly improved the
living conditions for human beings. However, the reserve of petroleum
sources is limited in contrast to the huge consumption demands. Even
worse, the excessive consumption of fossil resources has induced negative
effects on the environment, causing global warming and some extreme
climates. Hence, many countries have initiated extensive research on
renewable resources that can partially displace the prevailing fossil ener-
gies (1–3).
Biomass is the only major carbon-containing renewable resource, which
has been considered as the key feedstock for producing renewable fuels and
chemicals. Accordingly, many countries have set targets to promote the
renewable fuel production. The US Department of Energy has made a sce-
nario to replace 30% of the gasoline with biofuels by 2030. Energy Indepen-
dence and Security Act in the United States also created a Renewable Fuels
Standard that requires a minimum volume of 36 billion gallons renewable
fuel to be blended into the petroleum fuel in 2022. European Union has
developed a plan to produce one-fourth of fuel from biomass by 2030
(4). Similarly, China has planned to increase the renewable fuel to 20%
by the end of its current 5-year plan in 2020 (5). To fulfill these plans, a vari-
ety of strategies have been used to produce biofuels such as refining the
vegetable oil from energy crops, upgrading the pyrolysis oil to fuels, synthe-
sis fuels from chemicals that derived from biomass, or blending the gasoline
with additives of ethanol or butanol.
After years of explorations, ethanol has been proved to be the key chem-
ical for achieving these plans even though great progresses have been made
on other kinds of bio fuels (6). The ethanol production in the United States
had increased dramatically to 40 million tons in 2014 (7). Correspondingly,
Gasohol with 10 v% ethanol (E10) and 85 v% ethanol (E85) has been widely
used in the United States. Additionally, gasoline sold in Brazil contains at
92 Jifeng Pang et al.
least 25% anhydrous ethanol. More than 20% of cars in Brazil are able to use
100% ethanol, including ethanol-only and flex-fuel engines. In China, the
ethanol production will reach 10 million tons/year in 2020 with ca. 10 v%
ethanol blended in gasoline. In 2016, the global ethanol production reached
ca. 65 million tons with an estimated increase of 5% for the next 5 years (8).
Owing to the huge consumption amount of ethanol, its productions and
applications have attached great attention from the academics, industrials
and politics (9).
The production of ethanol can be dated back to the Neolithic period. In
this period, ethanol was produced by the fermentation of fruits, and used as a
main component in wine and a biological fluid in disease prevention (10). In
the late 1800s, internal combustion engines were invented to replace the
steam ones in vehicles. Consequently, different liquid fuels were developed.
In 1908, ethanol was used as the fuel component for the first time in a vehicle
named Ford’s Model T (11,12). Afterwards, the demand for ethanol
increased rapidly, and this demand was again consolidated by the World
War I. Then, as gasoline produced from fairly cheap petroleum became
the primary choice, the interest in using ethanol as a fuel was diminished.
During the energy crisis in the 1970s, a revival in using ethanol was surfaced.
Owing to the abundant sugarcane resources, Brazil first developed a Brazil-
ian Alcohol Program (Proalcool) in response to the Organization of Petro-
leum Exporting Countries Arab oil embargo. Under this initiative, ethanol
was mandatorily used in vehicles at a blend ratio larger than 20 v/v%.
Although facing the challenges of sugarcane production declines and petro-
leum price dwindles, gasoline sold in Brazil still has at least a 25 v/v% of
anhydrous alcohol (13). The growth of ethanol production in the United
States was gradual before 2000s, mainly due to the low gasoline price as well
as lacking commercially viable technologies. At the beginning of the 2000s,
methyl tertiary butyl ether, a dominant fuel additive for gasoline, was reg-
ulated in the United States, and ethanol was reintroduced into the fuels (14).
Furthermore, a political incentive named “Biofuels Initiative” was proposed
in 2006, which stimulates the ethanol production, and make it the main eth-
anol production country (7). Since then, great efforts have been made to
improve the synthetic process and reduce the cost of investment in the
fermentation of crops to ethanol (15).
Other than the sugar fermentation, chemical methods have also been
employed for ethanol production. At first, ethanol was mainly manufactured
from ethylene hydration (16). Later, syngas to ethanol conversion has
achieved great progresses, and realized the ethanol synthesis via direct and
Synthesis of ethanol and its catalytic conversion 93
indirect ways. The direct conversion occurs in the one-pot process either on
metal catalysts or through the fermentation process, whereas the indirect
way uses acetic acid/ester or dimethyl oxalate as intermediates for the
ethanol production (17). Recently, more challenge works on the direct
catalytic conversion of cellulose to ethanol have been achieved.
Ethanol is a low volatile and flammable liquid with a boiling point of
351.3 K, which has a low vapor pressure, relatively high motor octane num-
ber (99.1) and energy density (19.6 MJ/L). All these properties make it an
important fuel additive to partially replace the modern gasoline without
affecting its performances. Besides the main application in fuels additive,
ethanol has been widely used in preparing alcoholic beverages, wines or
potent inhibitors of platelet aggregation and infection. Additionally, etha-
nol has several functional groups, e.g., hydroxyl and –CH3 groups, and has
been regarded as a platform chemical for the synthesis of some value-added
chemicals such as acetaldehyde, ethyl acetate, butadiene, and p-xylene or
some high energy fuels like hydrogen, n-butanol and hydrocarbons
(18–21). The methods for ethanol production and its main applications
were summarized in Fig. 1.
Since ethanol is the most important bulk chemical in bio-refinery, it has
been widely investigated from several aspects. Numerous papers and
reviews have dedicated to this area on some special topics. For instance,
Baeyens et al. analyzed different ways for ethanol production, and proposed
the challenges and opportunities in improving the ethanol production (22).
Fig. 2 Process for ethanol production from biomass via the fermentation method.
Glucose
Yeast Cell
Membrane
ATP
HK
NAD
ADP Yeast Cell
Ethanol
Glucose-6-P Cytoplasm Membrane
ADH NADH+H+
CO2 PGI ATP
Acetaldehyde
Fructose-6-P
PDC ADP
Pyruvate PFK
PYK Fructose-1,6-di-P
ATP P-Enolpyruvate FBPA
ADP
Dihydroxyacetone-P
ENO
H2O TPI
2-P-Glycerate
ATP Glyceraldehyde-3-P
ADP
3-P-Glycerate
NAD+HPO42•
PGK NADH+H+
1,3-di-P-Glycerate
Fig. 4 Catalytic pathway for hydrocarbon (methane), methanol and ethanol production
from syngas (61). Reprinted (adapted) with permission from Luk, H. T.; Mondelli, C.;
Ferre, D. C.; Stewart, J. A.; Perez-Ramirez, J. Chem. Soc. Rev. 2017, 46 (5), 1358–1426.
Copyright (2017) Royal Society of Chemistry.
100 Jifeng Pang et al.
of CO and the hydrogenation of CO. Hence, both the activity for CO con-
version and the selectivity to ethanol were increased (71,72). Lin et al. used
Mn as the promoter to modify Rh/SiO2 catalysts on the bench-scale syngas
conversion. The ethanol selectivity reached 54.2% with 62% CO conver-
sion after more than 1000 h TOS at 583 K and 6.0 MPa of total reaction
pressure (73). Hu et al. conducted this reaction in a micro-channel reactor,
and obtained 56.1% selectivity to ethanol with 24.6% CO conversion at
553 K over a Rh-Mn/SiO2 catalyst (74). Different from the promoter of
Mn to form new sites, Fe increased the ethanol selectivity by modifying
and/or blocking the Rh sites to increase the barrier for methane formation
and/or decrease the barrier for CO insertion (75). Wang et al. used different
methods to introduce Fe into Rh catalysts to maximize the interface
between Rh and FeOx species. Over the 5 wt% FeOxRh/SiO2 catalyst,
42% selectivity to ethanol with 12% conversion of CO was achieved at
523 K (76). Xue et al. modified Rh with Cr and Fe to get the moderate
levels of CO dissociation and high efficiency of CO insertion. However,
the ethanol selectivity was only 26.0% with 10.2% CO conversion at
553 K and a GHSV of 5000 h1 over the CrFeRh/SiO2 catalyst (77). Yang
et al. disclosed that the selectivity to C2+ oxygenates over Rh catalysts is
structure dependent, and acetaldehyde is preferable produced over the
Rh(111) surface. In the presence of Fe, the ethanol selectivity significantly
improved even though the total C2+ oxygenate selectivity remained similar
to the un-promoted Rh counterparts. The authors proposed that Fe facili-
tated the hydrogenation of acetaldehyde to ethanol, and caused the forma-
tion of more thermodynamically stable ethanol (78). Huang et al. revisited
the silica supported Rh, Rh-Mn, and Rh-Mn-Fe catalysts, and found that
an Rh carbide phase was prone to be formed during the reaction over the
Rh–Mn/SiO2 catalyst. The addition of Fe induced the formation of bime-
tallic Rh-Fe alloys, which prevented the carbide formation and then
improved the ethanol selectivity (79). Although only slight increase was
observed, alkali metals were also used to suppress the methanation reaction
and enhance the CO insertion reaction. Li et al. reported a synergistic effect
between Ce and Rh at a Ce/Rh weight ratio of 1. At 573 K, 3 MPa and a
GHSV of 2400 h1, 33.3% selectivity to ethanol was achieved at 32.4% CO
conversion (80).
Since the dispersion and structure of Rh are crucial to its reactivity and
selectivity for CO conversion, a wide range of solid materials have been used
to adjust the status of active metals and improve the interaction between
metal and supports. Katzer et al. compared the activity of Rh over SiO2,
Al2O3, TiO2, CeO2 and MgO, and found that the TiO2 supported Rh gave
Synthesis of ethanol and its catalytic conversion 103
Catalysts
Syngas Supports (CNT) (multi-component) Ethanol
Fig. 6 Conversion of syngas to ethanol over CNT confined multiple catalysts (such as
Rh-Mn sites) (87). Reprinted (adapted) with permission from Pan, X.; Fan, Z.; Chen, W.;
Ding, Y.; Luo, H.; Bao, X., Nat. Mater. 2007, 6 (7), 507–511. Copyright (2007) Springer Nature.
104 Jifeng Pang et al.
of Co and Ga, and obtained 93% fraction of ethanol and high carbon alco-
hols in the final products. Additionally, the CoGa particles were trapped in
the LDHs, which showed highly stable catalytic performance (92). Ding
et al. incorporated K into CuFeMnZnO catalysts, which suppressed the
CH4 formation and increased the ethanol selectivity. At 0.5 wt% K loading,
533 K and a GHSV of 6000 h1, the ethanol selectivity reached 20.7% at
27.3% CO conversion. Interestingly, the C2OH/C1OH ratio reached
1.7, much higher than other K doped metal catalysts (93).
To control the adsorption of CO and H2 on catalysts, different carriers
were used to adjust the size and distribution of metal particles and the inter-
action between metals and supports. Feng et al. found that the CNTs favored
the distribution and interaction between Cu and Co species, which pro-
moted the reaction between CO and CHX species, and caused the depressed
methanol formation. Over the CuCo/CNT catalyst, 43.5% ethanol selec-
tivity was achieved with 39.9% CO conversion at 573 K (94). Wang
et al. supported Cu–Co–Ce on CNTs for high alcohols synthesis. The high
concentrations of Co/Ce in the narrowest CNT channels confined the car-
bon chain growth, and induced an enhancement in space–time yield and
selectivity to high alcohols (291.9 mg/gcath, 39%) (95). In addition, catalyst
preparation method is crucial to the structure of catalysts and its product
selectivity. Gao et al. prepared CuFe-based catalysts via the conventional
and LDHs based co-precipitation method. The CuFeMg catalyst derived
from CuFeMg-LDHs showed highly dispersed active species with the syn-
ergistic effect between the Cu and the Fe species, which exhibited higher
CO conversion (56.9%) and selectivity toward higher alcohols (49.1%) (96).
in a long term run (4000 h). It was found that H2S with ca. 100 ppm was
sufficient for sulfur maintenance over K-CoMoSx catalysts. Under non-
H2S conditions, the catalysts would be oxidized and carburized, leading
to the formation of acidic surface that changes its selectivity from alcohols
to hydrocarbons (119).
To further improve the catalytic activity and stability of K-CoMoSx cat-
alysts, different methods were used to modify the dispersion of active sites
and the structure of catalysts. Yang et al. used La to inhibit the formation
of CoSx crystal particles in catalysts, by which the dispersion of Co species
were enhanced and the concentration of active sites on surface (Mo–Co–S
mixture phases) were improved. Over the La0.2MoCo0.5K0.6 catalyst, 75%
selectivity to alcohols at ca. 20% CO selectivity was achieved in the >600 h
TOS at 603 K and a GHSV of 2225 h1 (120). Xie et al. incorporated Mn
oxides into the K-CoMoSx catalysts by a modified sol-gel method. The Mn
species enhanced the interaction between Co and Mo, adjusted the reduc-
tion degrees of Co and Mo precisely, and accordingly promoted the alcohols
selectivity from 4.2% to ca. 60% with 20% CO conversion at 723 K and a
GHSV of 6000 h1 (121).
Fig. 7 The WL pathway of acetogens for producing ethanol from syngas (125,129).
110 Jifeng Pang et al.
SA1, the ethanol and acetic acid production were 2.2 and 0.9 g L1, respec-
tively, in a semi-continuous fermentation process (132). Martin et al.
compared Clostridium ljungdahlii PETC, C. ljungdahlii ERI-2 and Clostrid-
ium autoethanogenum JA1-1 for the syngas fermentation. Over Clostridium
ljungdahlii PETC, the highest ethanol production rate of 0.301 g L1 h1
with a 5.5:1 ethanol/acetate molar ratio was achieved at pH 4.5 (129).
Liu et al. used the Alkalibaculum bacchi for the fermentation of syngas at
310 K and a pH of 8.0. The maximum ethanol concentration reached
1.7 g L1 after 360 h reaction (133). Mohammadi et al. employed the Clos-
tridium ljungdahlii to convert syngas in a 2 L continuous flow stirred tank
bioreactor. At a gas flow rate of 14 mL/min (55% CO) and an agitation rate
of 500 rpm, the maximum cell concentration was 2.34 g L1 at 93% CO
conversion and an ethanol concentration of 6.50 g L1 (134). In spite of
these achievements, most of the acetogens results are incomparable due
to the different reaction conditions. More parallel reactions should be con-
ducted to evaluate different bacterials.
Generally, the concentration of produced ethanol is ca. 2% with the
molar ratio of ethanol to acetic acid of ca. 1–20 (135). To increase the eth-
anol selectivity, several operating parameters need to be optimized. An
increased level of stress within bioreactor is one way to benefit high ethanol
selectivity. Under the stressed conditions, a severe growth limitation is con-
structed that redirected the reducing equivalents toward ethanol other than
acetic acid (136). To minimize the acetic acid yield, the produced acetic acid
must be converted to ethanol, which is thermodynamically promoted by a
high ratio of available reducing equivalents. A high CO partial pressure
enhances reduction of the ferredoxin via CO-dehydrogenase, and has been
reported to be a favorable condition for ethanol production (137). Some two
stage continuous fermentation systems were also used to balance the growth
and alcohol formation stages to obtain promising ethanol productivity of
0.374 g/(L h) (135). Additives were also used to change the ratio of ethanol
to acetic acid. For instance, Abubackar et al. added 0.75 μM of tungsten to
the solution. As a result, no acetic acid was produced with ethanol production
of 4260 mg L1 due to a shift in pH to a lower level in solution (from 6
to 4.75) (138,139). To reduce the operation cost, some inexpensive nutrients
were developed to replace the standard yeast. For instance, Gao et al. used ten
different media for fermentation over the Clostridium ragsdalei, and optimized
the ethanol yield at low-cost solution (named M9) (140).
The main challenge of the syngas fermentation process is the low solu-
bility of CO and H2 in water. To overcome this, various kinds of bioreactors
Synthesis of ethanol and its catalytic conversion 111
- I CO I C
I –
O
HI H2O Rh Rh
CH3 I CO I CO
I CO
I CO Ir I
Ir I CO
I CO
CO
O O
CH3 CH3 –
H3C I H3C I CO
C O I C
Rh O
I CO CO
Rh I CO
O I
I CO
H 3C OH
I-
+ CO
+ H2O
CH3OH MOR-Acylium cation
Methyl acetate
form methyl group saturated catalyst surfaces. CO was then inserted into the
methyl-saturated surfaces, forming the key intermediate of acetyl species
(164,165). The same group also found that the formation of carbon–carbon
bonds was preferably occurred within eight-membered ring (8-MR) zeolite
channels of H-MOR, and the reaction rates were proportional to the OdH
group numbers in the 8-MR channel (166). To improve the stability of cat-
alysts, Liu et al. modified H-MOR with pyridine, which enhanced the cat-
alysts stability to 48 h with a ca. 30% yield of methyl acetate at 473 K. Using
NMR characterizations, they found methyl acetate synthesis occurs within
the 8-MR, and the coke generates in the 12-MR (167,168). By the high-
temperature steam treatments to selectively remove the framework Al in the
12-MR of H-MOR, the stability of catalysts was improved albeit a slightly
decreased in methyl acetate selectivity (169). Rasmussen et al. presented an
in-depth understanding of the acetyl species formation over MOR catalyst
by DFT calculations. A ketene intermediate was found to be involved in the
reaction, which was the key issue for CdC bonds formation and coking
(170). Recently, a breakthrough was achieved by Liu et al. in the conversion
of DME into methyl acetate. Over some zeolite catalysts, the DME con-
version reached 68% with 98% selectivity to methyl acetate. The catalyst
was rather stable, which showed no deactivation after 6000 h TOS. This
process has been incorporated with methyl acetate hydrogenation for etha-
nol conversion at a plant scale of 100,000 ton/year. Meanwhile, a million
ton/year factory is being designed for further ethanol production (171).
116 Jifeng Pang et al.
were inhibited, and obtained 64% selectivity to ethanol with 93% acetic acid
conversion at a low temperature of 548 K and a LHSV of 0.9 h1 (179). Xu
et al. investigated the effect of thermal pretreatments on the active sites of the
PtSn/SiO2 catalyst. When the catalyst pretreated at 373 K, 94.9% acetic acid
conversion and 89.1% ethanol selectivity was obtained. The reason should
be attributed to the high dispersion of Pt, large amounts of weak Lewis acid
sites and synergistic effect between the Pt and SnOx species (180). Owing to
the synergistic effect of bimetallic catalysts, Dong et al. used the bimetallic
Cu-In/SBA-15 for acetic acid hydrogenation. Over the optimized 9Cu1In/
SBA-15 catalyst, the acetic acid conversion and ethanol selectivity reached
99.1% and 90.9%, respectively, at 623 K, 2.5 MPa hydrogen and a LHSV of
1.25 h1. As shown in Fig. 10, the Cu–In alloy promoted the dissociation of
acetic acid to acetate, and inhibited the combination of acetyl and ethoxy
species to form ethyl acetate. Therefore, the conversion of acetic acid was
boosted with high ethanol selectivity (181).
Owing to the corrosive problems of acetic acid under high tempera-
tures and pressures, much easy handling feedstock of acetic ester has been
used for producing ethanol, and the catalysts shift from noble metal cata-
lysts to Cu-based catalysts. The Cr modified Cu-based catalysts are indus-
trially favored catalysts, which exhibit excellent catalytic activities for ester
hydrogenation (182–184). However, the leaching of Cr species causes
serious environmental problems, which limits its practical applications.
So, Cr-free catalysts have been widely explored for the conversion of
acetic ester to ethanol.
Catalyst preparation methods are crucial to the catalysts activity. Wang
et al. used impregnation, deposition precipitation, and homogeneous depo-
sition precipitation methods for the synthesis of Cu/SBA-15 catalysts. Over
the catalysts prepared by the homogeneous deposition precipitation method,
both the methyl acetate conversion and ethanol selectivity reached 99.5%
at 493 K and 2 MPa of hydrogen. Based on catalyst characterizations, the
high activity of the catalysts was attributed to the small particle size of Cu
and an appropriate molar ratio of Cu+/(Cu+ + Cu0) (53.5%) (185). Simi-
larly, by adjusting the Cu species, Huang et al. prepared a core-shell structure
Cu/SiO2 catalyst for methyl acetate hydrogenation. At 523 K, the conver-
sion of methyl acetate reached 95% with a total selectivity of 95% for ethanol
and methanol in the 100 h TOS (186). To prevent Cu particles sintering
and improve the mechanical stability of catalysts, additives such as ZnO,
MgO, CeO2 and FeOx were extensively introduced (187,188). Among
these oxides, ZnO showed significant promotion effect in the activity
CH3CHO O
CH3COO
CH3COO
acetic acid H H H H H H
hydrogen H H H H H H CuO
In2O3
CH3CH2O
Culn alloy
CH3CHO
C H3CO
H2 Heat
H H H H H H Cu
H H H H H H
SBA-15
Acetaldehyde
Main-reaction pathway
O
CH3CH2O
CH3COC2H5 CH3CH2OH
By-reaction pathway
H H H H H H
H H H H H H
Fig. 10 Schematic diagram of acetic acid hydrogenation on Cu–In bimetallic catalysts (181). Reprinted (adapted) with permission from Dong, X.;
Lei, J.; Chen, Y.; Jiang, H.; Zhang, M. Appl. Catal. B Environ. 2019, 244, 448–458. Copyright (2019) Elsevier.
Synthesis of ethanol and its catalytic conversion 119
and stability for ethanol production. Zhu et al. investigated the influence of
Zn on Cu/SiO2 and Cu/Al2O3 catalysts. The addition of Zn affected the
structural and chemical state of Cu, which decreased the particle size of
Cu and increased the interaction between Cu and supports, and in turn
improved the activity and stability (>90 h TOS) of catalyst for acetic ester
conversion (189,190). Additionally, Di et al. used a co-precipitation method
for preparing Cu-Zn/SiO2 catalysts. Addition NH4+ and Zn2+ during the
catalysts preparation process generated more active Cu species on the surface
of catalysts with particle size <5 nm, which increased the methyl acetate
conversion to 98.0% with 99% selectivity to ethanol at 553 K and a LHSV
of 1.24 h1 (191). Lu et al. screened several supports (SiO2, Al2O3, and
ZrO2) to load Cu–Zn by a co-precipitation method, and found that the
Cu/ZnO/Al2O3 catalyst possessed the highest activity (86.5% ethyl acetate
conversion, 95.9% ethanol selectivity), due to the lower activation energy of
ethyl acetate on this catalyst (115 kJ/mol) (192).
Recently, Zhou et al. coupled the syngas conversion and intermediates
hydrogenation together. Over ZnAl2O4 j H-MORj ZnAl2O4 catalysts, 52%
ethanol selectivity was achieved at 6% CO conversion with methyl acetate
and dimethyl ether as the key intermediates, which provides a new concept
for the selective conversion of syngas to ethanol (193).
(CH3OCO)2
Pd0
DMO
2CH3ONO NO O + CH3OH
O O
CO linear
CO adsorption
Fig. 11 Proposed catalytic mechanism for CO oxidative coupling to DMO over the Pd
catalyst (196,197). Reprinted (adapted) with permission from Yue, H.; Ma, X.; Gong, J. Acc.
Chem. Res. 2014, 47 (5), 1483–1492. Copyright (2014) American Chemical Society.
produced NO with methanol and oxygen (Eq. 2). In the overall reaction,
CO reacts with methanol and oxygen to produce DMO with CH3ONO as
an active agent (Eq. 3). The generally accepted mechanism for this reaction
is shown in Fig. 11. In the catalytic cycle, Pd0 is oxidized by CH3ONO to
Pd(OCH3)2, and the adsorbed CO is then inserted into Pd(OCH3)2 to pro-
duce Pd(COOCH3)2. Finally, the Pd(COOCH3)2 undergoes intramolec-
ular coupling elimination reaction to produce DMO, and the Pd2+ is
reduced to Pd0 (196).
To make full use of the Pd sites, Zhao et al. used CNTs to disperse 1 wt%
Pd. The obtained Pd particle size was <5 nm, much smaller than those on
alumina. Correspondingly, the activity of the Pd/CNTs catalyst was much
higher than that of the commercial counterpart (1% Pd/CNTs 80% methyl
nitrite conversion vs 1% Pd/γ-Al2O3 30% methyl nitrite conversion at
393 K) (200). Peng et al. reported that the MgO supported 0.5 wt%
Pd exhibited excellent activity, selectivity and stability for this reaction.
The CO conversion reached 63% with 97% selectivity to DMO at 403 K
in the 120 h TOS (201). The same group supported the homogenous
PdCl2-CuCl2 system on Al2O3 for DMO production. The Cl was found
to have a synergetic effect with Cu2+, which is essential to achieve the high
selectivity toward DMO (99.8%) at 393 K and a GHSV of 3000 h1 (202).
They also synthesized Pd/α-Al2O3 catalysts by a Cu2+-assisted in situ reduc-
tion approach. The Cu2+ ions separated the Pd nanoparticles size to 2.7 nm
on the Al2O3 support, leading to a high activity and stability even at a Pd
loading of 0.1 wt% (97% selectivity at 62% conversion with a space-time yield
of 1332 g L1 h1) (203). The exposed facets and size of Pd are also crucial for
the DMO production. The (111) facet of Pd nanocrystals with 2–3 nm par-
ticle size are active for CO coupling with methanol to DMO, in contrast to
that over the (100) facet. Based on these results, they have successfully synthe-
sized a Pd(111)/α-Al2O3 catalyst through a wet impregnation-solution
chemical reduction method, which showed a STY of >1000 g L1 h1 in
the 100 h TOS (204). Wang et al. loaded Pd onto the surface of the AlOOH
rooted on the Al-fiber, forming a thin-sheet microfibrous-structured
Al-fiber@ns-AlOOH@Pd catalyst. The unique structure endows the cat-
alyst with high activity even at 0.25 wt% Pd loading. Over which, the CO
conversion reached 66% with 94% selectivity to DMO at a GHSV of
3000 L kg1 h1, which was two times higher in intrinsic activity than that
of the traditional Pd/α-Al2O3 catalysts (205). Some researchers introduced
additives to decrease the size of Pd on α-Al2O3 support and thus promoted
the DMO production. Zhao et al. reported that the CeO2 decreased the Pd
size to ca. 13 nm on α-Al2O3, and increased ca. 20% CO conversion. This
promotion effect was confirmed by Fan group’s using the Pd(NO3)2 and
PdCl2 precursors (206,207). An optimized eggshell distribution of Pd
was realized using Pd(NO3)2 precursor, which greatly reduced the Pd loading
to 0.1 wt% even at 90% CO conversion (208). Gao et al. employed Fe
to modify the Pd/α-Al2O3 catalyst over a cordierite monolithic support,
and investigated the influence of calcination temperatures on the reaction.
At calcination temperatures between 473 and 673 K, the monolithic catalyst
122 Jifeng Pang et al.
H3C O O
H3C O
+H2
O
-CH3OH
O CH3 O OH
-H2O
H3C O +H2
-CH3OH
+H2
O
-CH3OH
HO
HO
+H2
-H2O
OH
Ethanol
Ethylene glycol
Fig. 12 Reaction networks for the hydrogenation of DMO to ethanol (196,213).
Synthesis of ethanol and its catalytic conversion 123
treatment, which is much better than that over Cu/SiO2 catalysts (63%
yield) (215). To elevate the long-term stability and activity, they used boron
acid to dope Cu/SiO2 catalysts. The boric acid decreased the size of metallic
Cu, resulting in a strong interaction with the surface cupreous species. Over
the 3%CuB/SiO2 catalyst, the ethanol yield was 85% with low activity loss
during the reaction (216). In a similar way, they also introduced Ni to the
Cu/SiO2 catalysts by an impregnation method. At 1 wt% of Ni, the selec-
tivity to ethanol increased to 90% mainly due to the enhanced H2 adsorption
and dissociation ability of the modified catalyst (217).
Some unique materials have also been tried to disperse Cu species on cat-
alysts for DMO hydrogenation. Zhu et al. dispersed Cu nanoparticles in
mesoporous Al2O3, over which the ethanol yield reached 94.9% during
the 200 h TOS reaction at 543 K and a LHSV of 0.21. The high activity
and stability of the 15Cu85Al catalyst was attributed to the mesoporous
pores structure, which benefited a balanced surface Cu0–Cu+ species
(218). They also tried Cu/ZrO2/Al2O3 catalysts with different structures
(different CuO phases, particle sizes, and CuAl2O4 spinel) for ethanol syn-
thesis. Under optimized calcination temperature of 1023 K, an ethanol yield
up to 97.4% was obtained in the 200 h TOS (219). Recently, Liu et al. used
Mo2C/SiO2 catalysts for DMO hydrogenation. Unlike Cu/SiO2 catalysts,
the hydrogenation of DMO takes place at temperatures as low as 473 K, and
the main intermediate is methyl acetate instead of ethylene glycol. Even
though the yield to ethanol is not comparable to that of Cu/SiO2 catalysts
(70.2% vs >90%), it still shows promising stability (350 h) for ethanol
production (213).
A much greener process for ethanol production was achieved by Xu et al.
using cellulose as the feedstock. In this process, cellulose was first converted
into methyl glycolate with a high yield of 57.7 C % in a one-pot reaction in
methanol at 513 K and 1 MPa O2 over tungsten-based catalysts. The methyl
glycolate was then separated by distillation from the product solutions.
Finally, the methyl glycolate was converted to EG at 473 K and to ethanol
at 553 K with a 50% selectivity over a Cu/SiO2 catalyst. Through the two-
step reactions, methyl glycolate, EG and ethanol can be selectively produced
(Fig. 14A), which provides a versatile way for converting cellulose into
value-added chemicals (221). The same group optimized the last step using
0.1Pt–Cu/SiO2 single-atom alloy catalyst, which enhanced the catalytic
activity and selectivity to ethanol (76.7% at 503 K) in the 700 h TOS (222).
Recently, Yang et al. developed a one-pot approach for the conversion
of cellulosic biomass to ethanol. Over a multifunctional MoOx/Pt/WOx
Synthesis of ethanol and its catalytic conversion 125
Fig. 14 Catalytic conversion of cellulosic biomass to ethanol (A: two steps; B: one-pot)
(220). Reprinted (adapted) with permission from Yang, M.; Qi, H.; Liu, F.; Ren, Y.; Pan, X.;
Zhang, L.; Liu, X.; Wang, H.; Pang, J.; Zheng, M.; Wang, A.; Zhang, T. Joule 2019, 3,
1937–1948. Copyright (2019) Elsevier.
catalyst, 43.2 C% yield of ethanol was obtained with 100% cellulose conver-
sion at 518 K and 6 MPa hydrogen. The reaction involved in the tandem
reactions of converting cellulose to EG (rapid) and hydrogenolysis of EG
to ethanol (slow). Characterizations found that the Pt-W species catalyzed
the conversion of cellulose to EG (223,224), and the isolated and low-
coordinated MoOx bonded with Pt formed 5OMo-Pt-WOx active sites
to promote the CdO bonds cleavage of EG to ethanol (Fig. 14B) (220).
Nearly at the same time, Song et al. combined H2WO4 with Pt/ZrO2
for the direct conversion of cellulose into ethanol. The H2WO4 mainly cat-
alyzed the cleavage of CdC bonds in the glucose unit, whereas Pt/ZrO2 with
Pt0 and Pt2+ species promoted CdO bonds cleavage to yield ethanol. After
5 h reaction at 523 K and hydrogen atmosphere, 32% yield of ethanol was
achieved in an aqueous medium (225). Li et al. developed a multifunctional
Ru-WOx/HZSM-5 catalyst for the one-pot conversion of cellulose to
ethanol. Over the 5Ru-25WOx/HZSM-5 catalyst, 76.8% yield of ethanol
was obtained from the 1 wt% cellulose at 508 K and 3 MPa H2 in the 20 h
reaction. By adjusting the heating ways, high-concentration feedstock
(5 wt%) conversion was achieved with 53.7% yield of ethanol (226).
functional groups of the support (239). Similarly, Morales et al. used ther-
mally treated graphitic oxides, which bear certain functional groups for Cu
catalysts preparation. Over the N doped graphite support, the Cu species
were uniformly dispersed with particle size of 13 nm. It showed 97.8%
acetaldehyde selectivity and a high activity of ca. 13.7 μmolethanol g1 s1,
1 order of magnitude higher than that on Cu/SiO2. The catalyst also
showed superior stability even with co-fed water (240). Additionally,
the reaction conditions also greatly affect the finally products yield. Over
the Cu/ZnO catalyst, acetaldehyde and hydrogen were the main products
at low ethanol conversions, and shifted to ethyl acetate at high ethanol
conversions (241,242). Due to the unique catalytic activity of novel
single-atom catalysts for a variety of hydrogenation and dehydrogenation
reactions (223,243), NiCu single atom alloy nanoparticles were synthe-
sized for the non-oxidative dehydrogenation of ethanol to acetaldehyde
and hydrogen. Different from Cu/SiO2 catalysts remain unreactive up
to 523 K, acetaldehyde was formed at 423 K over the NiCu single atom
alloy catalysts. Based on the Arrhenius-type plots of the reaction rate in
Fig. 15 and characterizations, the authors concluded that the atomically
dispersed Ni in Cu significantly lowered the CdH bond activation
barrier, and then significantly increased the reaction activity (244,245).
Fig. 16 TOF for ethanol dehydrogenation in the absence (■, 523 K) and presence
(▲, 473 K) of oxygen as a function of the gold particle size (246). Reprinted (adapted) with
permission from Guan, Y.; Hensen, E. J. M. Appl. Catal. A Gen. 2009, 361 (1–2), 49–56.
Copyright (2009) Elsevier.
Recently, Au-based catalysts have been widely used for ethanol dehy-
drogenation. Guan et al. synthesized Au/SiO2 catalysts with controlled
Au particle size for the ethanol dehydrogenation reaction. As shown in
Fig. 16, the activity of ethanol dehydrogenation was optimized at an Au par-
ticle size of 6 nm, which was attributed to the surface steps that had a suitable
geometry for the removal of α-H atoms from adsorbed ethoxide. Under
oxygen atmosphere, the intrinsic activity remained unchanged up to about
7 nm and then started to increase with particle size beyond 7 nm. However,
the selectivity to acetaldehyde became lower due to the formation of acids
and carbon dioxide (246). Gazsi et al. screened a variety of supports includ-
ing CeO2, Al2O3, MgO, TiO2, SiO2 and AC to load Au for the dehydro-
genation of ethanol, and found that the product selectivity was highly
dependent on the supports. Over the 1%Au/SiO2 catalyst, >90% selectivity
to acetaldehyde was obtained at 575–775 K due to its low activity in acetal-
dehyde decomposition (247). Wang et al. atomically dispersed Au on nano-
scale ZnZrOx for the low-temperature ethanol dehydrogenation. Owing to
the high dispersion of Au and the unique interaction between Au and sup-
ports (Au-OX), the undesired dehydration reactions were inhibited, and ca.
75% acetaldehyde was obtained at 80% ethanol conversion over the 0.5%Au/
Zn1Zr10Ox catalyst at 598 K and a WHSV of 1.11 gEtOH/(gCat.h) (248).
Ag, Pd-based catalysts and some carbon materials have also been used for
ethanol dehydrogenation. Abu-Zied calcined a mixture of chromic oxide
132 Jifeng Pang et al.
gel and silver nitrate to make silver/chromia catalysts for this reaction.
However, the selectivity to acetaldehyde greatly decreased at high ethanol
conversions due to the co-production of diethyl ether, ethylene and ethyl
acetate (249). Sushkevich et al. studied the activity of Ag/SiO2 catalysts for
this reaction, and achieved a TOF of 3.4 s1 to acetaldehyde. Based on
characterizations, they proposed that the ethanol was first activated by silica
to yield a hydrogen-bonded surface complex. Then, the rate-determining
step of CdH bond cleavage and proton abstraction happened on the close
Ag sites and silica sites, respectively. Finally, H2 and acetaldehyde desorbed
with the regeneration of the active sites (250). Similar to Cu, Pd on ZnO
was also showed high activity for ethanol dehydrogenation at low ethanol
conversions, but the reaction pathway changed to ester productions at high
ethanol conversions. At 493 K and a TOF of 0.087 s1, the selectivity to
acetaldehyde was 100% at 21.7% ethanol conversion. The selectivity
to ethyl acetate, however, reached up to 50% at 60% ethanol conversion
under the similar conditions (251). Interestingly, some N-doped graphene
also exhibited certain activity for ethanol conversion. It catalyzed the dehy-
drogenation of ethanol at 473–623 K, and produced acetaldehyde in 100%
selectivity at <10% ethanol conversions (252).
Ethyl acetate is usually the co-product of acetaldehyde, which is also
another important chemical derived from the ethanol dehydrogenation.
Ethyl acetate has a wide range of applications, and has been industrially pro-
duced using Cu/Cr2O3 catalysts by the Davy Process Technology Limited
(253). Very similar to the acetaldehyde production, Cu-based catalysts are
primarily investigated for the reaction. For instance, Santacesaria et al. devel-
oped a commercial Cu/copper chromite catalyst for ethanol dehydrogena-
tion at 493–513 K, 2 MPa and a contact time of 98 g h mol1. The ethanol
conversion was 65% with 98–99% selectivity to ethyl acetate (254). After
performing a kinetic study over the same catalyst, the same group employed
the Langmuir–Hinshelwood–Hougen–Watson kinetic model to interpret
the experimental data and the reaction mechanism. Acetaldehyde was iden-
tified as the key intermediate for this reaction, which reacted with another
molecule of ethanol or an ethoxide specie to form a hemiacetal. The formed
hemiacetal was then dehydrogenated to ethyl acetate (255,256). Yu et al.
modified the Cu/SiO2 catalyst with ZrO2, and found that the yield to ethyl
acetate was improved due to the synergistic effect between Cu0 and acid sites
(257). Freitas et al. studied Cu/ZrO2 catalysts in this reaction, and attributed
to the high activity of catalysts to the high density of base sites of O2– and a
heterogeneous distribution of Cu0/Cu+ species (258). They also compared
Synthesis of ethanol and its catalytic conversion 133
different ZrO2 phases, i.e., amorphous (am-), monoclinic (m-), and tetrag-
onal (t-) phases, in this reaction. The highest selectivity to ethyl acetate
was achieved on Cu/m-ZrO2 catalysts (259). Apart from Cu-based cata-
lysts, Pd-based catalysts are also used for ethyl acetate synthesis from etha-
nol. Sánchez et al. explored different supports to load Pd particles (1 wt%),
and found that the reducible supports such as ZnO and SnO2 favored the
formation of alloyed Pd phases, which improved the ethyl acetate
selectivity (260).
and an O2/ethanol ratio of 2. The yield to ethyl acetate and acetic acid
increased significantly at a higher temperature or O2/ethanol ratio (263).
To improve the activity of Au, different supports were tried to disperse Au
for this reaction. Takei et al. screened 23 supports for loading Au nanoparticles
and used in oxidative ethanol dehydrogenation. Strongly acidic MoO3 or
weakly basic La2O3 supported Au catalysts showed >95% selectivity to acet-
aldehyde at temperatures above 473 K, whereas p-type and n-type semi-
conductive metal oxides supported catalysts resulted in the dominant
formation of complete oxidation products (i.e., CO2) or deep oxidation
product of acetic acid. The different performance of Au on supports was
attributed to the varying stability of surface metal ethoxide species, which
was controlled by the amount of surface oxygen species (264). Sobolev
et al. used TiO2 to support Au for the oxidative dehydrogenation of eth-
anol. Due to the direct participation of active oxygen species, the reaction
temperature started at as low as 398 K with a “double-peak” profile (265).
Mielby et al. used mesoporous silicalite-1 to encapsulate Au nanoparticles,
and obtained Au/Recryst-S1 catalysts with Au particles of 2–3 nm, over
which 50% ethanol conversion and 98% selectivity toward acetaldehyde
were achieved at 473 K and a O2/ethanol ratio of 1 (266).
Synthesis of ethanol and its catalytic conversion 135
Since the electronic effects may exist between different metals, some bime-
tallic catalysts were synthesized and applied for the reaction. Bauer et al. pre-
pared Au-Cu/SiO2 catalysts for ethanol conversion. After oxidization of
the catalyst at 573 K, Au core with a thin CuOx shell structure was formed.
The ethanol conversion reached 90% with >80% selectivity to acetalde-
hyde over the corn-shell bimetallic catalysts at 473 K in a 50 h TOS
(267). A highly efficient and robust Au/MgCuCr2O4 catalyst was synthe-
sized by Liu and Hensen, giving ca. 100% ethanol conversion and ca. 95%
acetaldehyde selectivity during the >500 h of reaction at 523 K and a
GHSV of 100,000 mL g cat1 h1. The Cu species in the chromite-spinel
phase have a synergistic effect with Au in activating O2 and stabilizing the
Au nanoparticles during the reaction (18). The same group also developed
Au–Ir catalysts on SiO2, which showed an enhanced activity as compared
with the pure Au nanoparticles (271). Similarly, a synergistic interaction
between Au and Cu species was observed at a ratio of 1:1, and whole eth-
anol conversion with 100% selectivity to acetaldehyde was achieved at
523 K albeit a low catalyst stability (268). Additionally, Au nanoparticles
supported on copper silicate nanotubes showed a higher activity (98% eth-
anol conversion and 93% acetaldehyde selectivity at 523 K) than that
over other supports like MgSiO3 due to the synergistic effect between
Au and Cu (269).
Some other metals including V, Pd and carbon materials were used for
the oxidative dehydrogenation of ethanol. Chimentão et al. employed Na to
increase the dispersion of V on MCM-41 and TiO2 for ethanol conversion.
The acetaldehyde selectivity reached >95% at <20% ethanol conversion
over 0.5%Na-VOX catalysts. However, the ethanol conversion gradually
decreased with TOS (272). Other supports including TiO2/SiO2, ZrO2/SiO2,
HT derived oxides, Al2O3, CeO2, ZrO2, TiO2, and SBA-15 were also
used to load the VOx for ethanol oxidation. The supports showed a
strong influence on the nature of V species, evidenced by changing
the activation energy and oxidative dehydrogenation rate in ethanol con-
version (273). Consequently, the optimized VOX/ZrO2/SiO2 catalysts
were found to give the highest activity due to the suitable ionicity in sup-
ports (274,275). Recently, CNTs were directly used as catalysts to con-
vert ethanol into acetaldehyde. Approximately 60% ethanol conversion
and 93% acetaldehyde selectivity were achieved over the optimized
CNT1000 catalysts at 543 K. Characterizations showed that the C]O
groups generated on the CNT surface were the active sites for ethanol
conversion (276).
136 Jifeng Pang et al.
CH3CH2OH
[1] + CH3CH2OH [6] – H2
– H2O CH3CHO
CH3CH2OCH2CH3 [7] + CH3CHO [12] + H2O [13] + CH3CHO
[2] – H2O CH3C(O)OCH2CH3 – H2 CH3CH(OH)CH2CHO
2CH2=CH2 – CH2=CH2 [14] – H2 [16] – H2O
[18]
[3] [8] CH3C(O)OH CH3C(O)CH2CHO
CH2=CHCH2CH3 [9] + CH3C(O)OH
[4] – H2O [15] – CO
CH3CH=CHCH3 CH3C(O)CH2C(O)OH
[5] + CH2=CH2 [10] – CO2
2CH2=CHCH3 CH3C(O)CH3 CH2=CHCH2CHO
[11] + H2 (or + CH3CH3OH, –CH3CHO) [17] – CO
– H2O
CH2=CHCH3
Fig. 17 Proposed reaction pathways for the conversion of ethanol to propylene (278).
Reprinted (adapted) with permission from Iwamoto, M.; Tanaka, M.; Hirakawa, S.;
Mizuno, S.; Kurosawa, M. ACS Catal. 2014, 4 (10), 3463–3469. Copyright (2014) American
Chemical Society.
Synthesis of ethanol and its catalytic conversion 137
best results of 37% propylene yield in the presence of water. Their results
demonstrated that this reaction mainly proceeds via the acetone interme-
diate (302). Recently, Xia et al. synthesized Y-modified ZrO2 catalysts
by a co-precipitation method for ethanol to propylene conversion. The
addition of Y increased the catalyst surface area, changed the acid and base
sites, and modified the crystalline structure of ZrO2, which improved the
propylene yield up to 44.0% (303).
Fig. 18 Proposed reaction mechanisms for the conversion of ethanol to 1,3-BD (Black:
Toussaint’s generally accepted mechanism, Red: Fripiat’s Prins mechanism, Blue: Ostro-
mislensky’s hemiacetal rearrangement) (315).
140 Jifeng Pang et al.
DFT calculations over MgO surface (315). The energy barrier for ethanol
dehydration to ethylene was found to be lower than that for ethanol dehy-
drogenation (33.5 kcal/mol vs 39.6 kcal/mol), suggesting ethylene is the
main by-product. The energy barrier of the aldol condensation is about
16 kcal/mol, slightly lower than the direct reaction of ethylene and acetal-
dehyde (29 kcal/mol), suggesting that Prins condensation is also an alterna-
tive route (highlighted in red). Moreover, the hemiacetal rearrangement
mechanism was also proposed, which involves reaction between ethanol
and acetaldehyde to yield 1-ethoxyethanol and then to 1.3-BD (highlighted
in blue). Recently, Chieregato et al. investigated the reaction mechanism
over pure MgO catalysts. Crotyl alcohol and 3-buten-1-ol, rather than
acetaldol and crotonaldehyde, are identified as key intermediates to 1,3-
BD in the Lebedev process (316). In spite of the different reaction mecha-
nisms, multifunctional catalysts with a balanced amounts and strengths of
acid/base and redox sites are essential parameters that need to be considered
in catalysts design. Herein, we focus on different multifunctional catalysts for
1,3-BD production via one-step process developed by Lebedev and two-
step process known as the Ostromislensky process. In general, the catalysts
could be classified into mixed oxides, metal enhanced catalysts and others.
catalyst was sensitive to the precursors used in the catalyst preparation pro-
cess. For instance, Ohnishi et al. reported that a wet-kneading catalyst made
with MgCl2 as the precursor had a worse performance than that made from
Mg(NO3)2 due to the chloride ion residual (320).
The ratio between SiO2 and MgO is another key parameter that affects
the acidic and basic properties of catalysts. Ohnishi et al. found that the best
142 Jifeng Pang et al.
Co, Ni, Cu, Zn, Ag) on MgO2–SiO2 catalysts, and found that the addition
of metallic Ag and Cu greatly improved the ethanol conversion and 1,3-BD
selectivity. For instance, over the Ag modified MgO2–SiO2 catalysts, the
1.3-BD yield reached 55% at full ethanol conversion (37% ethanol conver-
sion and 39% 1,3-BD selectivity over the MgO–SiO2 catalyst) at 623 K
(332). In the following research, they found the addition of Ag enhanced
the dehydrogenation of ethanol without affecting the weak acid and strong
base sites for the aldol condensation. Hence, the highest 1,3-BD productiv-
ity of 0.29 gBDg1Cat h
1
was obtained over 1.0%Ag/MgO–SiO2 catalysts at a
1
WHSV of 1.2 h (337). Dagle et al. screened different silica supports in the
conversion of ethanol 1,3-BD. It was found that the silica support changed
the dispersion of Zr and Ag species, and then affected the Lewis acid sites of
catalysts. Over the optimized 1Ag/4ZrO2/SiO2-SBA-16 catalyst, 70%
butadiene yield was obtained at 99% ethanol conversion (598 K) (346).
Ivanova’s group did a systematic study on converting ethanol to 1,3-BD
over Ag/Zeolites catalysts. After screening a variety of metals such as Ag,
Cu and Ni, they found Ag was the most active, selective, and stable metal.
Among the oxides including MgO, ZrO2, Nb2O5, TiO2 and Al2O3, ZrO2
was found to be the best choice in the condensation of acetaldehyde and the
MPV reaction of crotonaldehyde. The optimized catalyst contained 1 wt%
Ag and 10 wt% Zr on silica, provided 74 mol% selectivity to butadiene at an
ethanol conversion of 88% at 593 K with a WHSV of 0.04 h1 (338). Later,
different Zr containing zeolites were also tested to support metallic Ag for
this reaction. The BEA(200) zeolites incorporated with isolated Zr showed
the better activity, giving 56% selectivity to 1,3-BD at an ethanol conver-
sion of 48% (339). FT-IR and DFT studies revealed that the isolated Zr
atoms connected to three –O–Si linkages and one OH group in the zeolite
crystalline structure afforded open Lewis acid sites for efficient conversion
of ethanol to 1,3-BD (347). To increase the amount of Zr open sites, they
prepared the ZrBEA using post-synthesis modification method by
dealumination of the parent zeolite followed by treatment with ZrOCl2
in a DMSO solution, and the highest 1,3-BD formation rate of 3
mmolg1 s1 at ca. 60% selectivity was achieved over the 1%Ag/Zr(3.5)-
BEA(75) catalyst (340). Finally, they investigated the role of different zeo-
lites on CdC condensation reactions, and different mechanisms were
proposed among the (Sn, Zr, Ti) BEA zeolites. For the most active support
of ZrBEA, two aldehyde molecules were co-adsorbed at the open M(IV)
Lewis acid site, and the protons were transferred between the adsorbents
via the M-OH group of the open site during the condensation reactions
(348,349).
146 Jifeng Pang et al.
which was then reduced by an MPV reaction with ethanol to form crotyl
alcohol. Finally, 1,3-BD was formed after the dehydration of the crotyl alco-
hol (Fig. 18, highlighted in bold black). The 1,3-BD productivity is highly
sensitive to the ethanol/acetaldehyde ratio due to a competitive adsorption
between the reactants on Lewis acid sites (355).
When using ethanol as the sole feedstock, dual fixed-bed reactors are
employed for coupling ethanol dehydrogenation and condensation reac-
tions. Nevertheless, the catalysts should be carefully selected to balance
these steps. Klein et al. screened different metals and zeolites for 1,3-BD
production. The Cu/SiO2 catalysts were highly active and stable for eth-
anol dehydrogenation. The acetaldehyde selectivity reached 100% at 20%
ethanol conversion during a 90 h TOS in the first stage. In the second stage,
the resulted mixture of ethanol and acetaldehyde was converted over doped
zeolite catalysts with different structures. Specifically, over the MgO10\K-β
280 catalyst, the highest 1,3-BD selectivity of 72% was obtained (356).
Cheong et al. used the mesocellular siliceous foam to load Cu and Zr sites.
Given the inhibited ethanol dehydration reaction, 1,3-BD selectivity was
improved to 73% at 96% ethanol conversion with an unprecedented
productivity of 1.4 gBD/gcatalyst h1 (357).
Since the first step of acetone production from ethanol was discussed in
the former section, herein we only summarize the progresses in converting
acetone to isobutene. Currently, acidic zeolites and mixed oxides are two
typical kinds of catalyst used for this reaction. In 1993, Hutchings et al. used
zeolite β and ZSM-5 for the conversion of acetone to isobutene, and
obtained >80% selectivity to isobutene at 65% acetone conversion over
the zeolite β catalysts (358). Nevertheless, the zeolite β was prone to be
Synthesis of ethanol and its catalytic conversion 149
deactivated due to the strong Brønsted acid induced coke formation. Hence,
Tago et al. used alkali metals to depress the acidity of β zeolites by an
ion-exchange method. Over the K-β zeolite catalyst, the strong acid sites
were mostly eliminated, which depressed the undesired reactions, leading
to an increased isobutylene yield (55%) with a higher stability (359).
Recently, Herrmann and Iglesia studied the elementary steps of isobutene
(C4) and acetic acid formation from acetone over acidic zeolites catalysts
(typically, MFI and β). It was found that there are two distinct reaction
pathways, i.e., anhydrous (MO: mesityl oxide) and H2O-mediated (DA:
diacetone alcohol) routes, for the conversion of C6 species (from acetone
coupling) to C4 species. In contrast to the minor way of H2O-mediated
route involving β-scission of C6 ketols by protons, the predominant anhy-
drous pathway involves the β-scission of C6 ketols via van der Waals contacts
within vicinal microporous voids (360).
Due to the similar operation conditions and catalysts for conversion of
ethanol to acetone and acetone to isobutene, it is very attractive to couple
these two steps for the direct conversion of ethanol to isobutene. Sun et al.
developed a multifunctional ZnxZryOz mixed oxide catalyst to fulfill the
one step process. By adjusting the Zn/Zr ratio, catalyst with passivated
strong Lewis acid and weakened Brønsted acid sites were obtained. Over
then, the undesirable dehydration reactions were suppressed, the polymer-
ization/coking formation were mitigated, and the yield of isobutene was
elevated to 83% (21). The authors also investigated the influence of reaction
parameters on the isobutene yield over different ZnxZryOz catalysts, trying
to disclose the role of different acid sites. It was found that weak Brønsted
acid sites were active for the conversion of acetone to isobutene, whereas
strong Brønsted acid sites were responsible for the coke formation. Over
an improved Zn1Zr8O17 catalyst, an isobutene yield of 79% was achieved
with an ethanol molar fraction of 8.3% at 748 K (361). The Lewis acid–base
pairs over ZnxZryOz catalysts played the key role in theoretically conversion
of acetone to isobutene (88.9%). The absence of Brønsted acid sites elim-
inated the isobutene isomerization and undesired polymerization/coke
reactions, and increased the catalyst stability to more than 200 h TOS with
<2% activity loss (362).
C2 to C1(Ir-CeO2 interface)
CO
CH4 H2 C2 chemistry
CO2 (CeO2 surface)
–
CH3COO
CH3CH2O–
+H
Had –OH
COad
CO32– CHx Ir OH
CeO2
Fig. 19 Schematic view of the bifunctional mechanism of ethanol steam reforming over
Ir/CeO2 catalysts (381). Reprinted (adapted) with permission from Wang, F.; Cai, W.;
Descorme, C.; Provendier, H.; Shen, W.; Mirodatos, C.; Schuurman, Y. Int. J. Hydrogen.
Energ. 2014, 39 (31), 18005–18015. Copyright (2014) Elsevier.
oxidized into acetate species on the ceria support, and then they were trans-
formed to CHx and carbonyls at the noble metal/ceria interface. The
obtained CHx and carbonyls migrated over the Ir particles to hydrogen,
CO and methane, as illustrated in Fig. 19 (381). Recently, Dai et al. encap-
sulated nano Pt particles in the hollow Beta microreactor for ethanol steam
reforming. In contrast to Pt-SiO2 and Pt-Beta, the resulted catalyst spatially
inhibited the particles aggregation, and unreacted ethanol and acetaldehyde
diffusion, which led to the higher stability and activity for the hydrogen
production (100% ethanol conversion with >60 H2 selectivity at 623 K
for 15 h) (382).
[Link].2 Non-noble metal catalysts Although precious metals are
very active for this reaction, their high cost still limits the practical appli-
cation at large scales. Inexpensive non-noble metal catalysts such as Ni and
Co have been extensively investigated for the ethanol steam reforming
reaction.
Ni is the primary candidate because of its excellent CdC breaking prop-
erties. According to the thermodynamic equilibrium of steam reforming,
Gong’s group conducted this reaction under mild conditions to produce
CO-free hydrogen. They found that skeletal Ni-based catalysts and co-
precipitation method prepared catalysts with high-loading of Ni on Al2O3
exhibited considerable H2 selectivity (35%) with less than 100 ppm CO in
the products. However, coking and particle sintering were still observed dur-
ing the reaction (383,384). Montero et al. studied the coke deactivation
154 Jifeng Pang et al.
0h 1.5 h 4h 8h 20 h
Fresh catalyst
0
Support Ni particle Filamentous coke Ni-carbide Non filamentous coke
increased butanol production to 12% at ca. 725 K (351). The catalyst prep-
aration method also greatly affected their activity. Dı́az et al. prepared mixed
oxide catalysts under different conditions to create different distributions of
acid and base sites. It was found that catalysts prepared at super saturation
conditions have large amounts of strong base sites, which showed the best
performances (ca. 20% selectivity at 473 K) (430,431). Appel et al. com-
pared MgO, Al2O3 and Mg-Al mixed oxides in ethanol conversion, and
concluded that adjacent weak acid and medium base sites are needed for
producing the C4 compound. Over the MgAl 3:1 catalyst, the butanol
selectivity reached ca. 37% with 35% ethanol converted at 623 K (432).
The cations and anions of HAP can be replaced by guest species, which
in turn affects the acid–base properties of catalysts, and thus changes the
butanol selectivity. Onda et al. substituted Ca-HAPs with Sr, which
exhibited a high butanol selectivity of ca. 80% at an ethanol conversion
between 1% and 24% at 573 K in a fixed-bed reactor. Conditional experi-
ments revealed that the Sr-HAP catalyst promoted the aldol condensation
and inhibited the coking by favoring the hydrogen transfer, resulting in
the high selectivity to butanol (440). The same group synthesized
Sr-HAP catalysts with different Sr/P molar ratios for ethanol conversion.
Although there was an increase in the amount of strong acid sites, the density
of base sites significantly higher than that of the acid sites density, which cau-
sed the high butanol selectivity of 86.4% over the Sr-HAP (1.70) catalyst
under the conditions of 573 K and a W/F of 130 h g/mol (441). Dumeignil
et al. investigated the influence of HAP anions on ethanol conversion, and
found that the anions greatly influenced the acidity/basicity ratio of HAP.
An optimized acidity/basicity ratio of 5 was achieved on the HAP-CO3 cat-
alyst, which led to the 40% ethanol conversion and 30% yield of heavier
alcohols at 673 K and a GHSV of 5000 mL h1 g1 (442).
The reaction pathway over HAP was revisited by Meunier et al. over a
commercial HAP catalyst. Through an analysis of the thermodynamic and
kinetic data, it was suggested that two reaction pathways take place simul-
taneously. The dominant one was the direct condensation of two ethanol
molecules (the direct route), (443) whereas the minor one went through
the condensation of ethanol with acetaldehyde (the Guerbet route). Com-
paring metal and HAP catalysts, it was proposed that butanol was formed
mainly by the direct route on metal-free basic oxides at high temperatures,
however the pathway changed to the Guerbet route over the metal con-
taining system under mild conditions (444).
Pd-Mg-Al gave the best performances in terms of stability and butanol selec-
tivity (2.8% ethanol conversion and 72.7% butanol selectivity at 473 K for
5 h). In addition, water was found to be detrimental for butanol production
(453). Recently, Pang et al. reported the conversion of ethanol to butanol
over HT derived NiMgAlO catalysts. At 523 K in a fixed bed reactor, the
n-butanol and C4–C8 alcohols selectivity reached 55.2% and 85%, respec-
tively, at 18.7% ethanol conversion. The high activity of NiMgAlO catalysts
was attributed to metallic Ni sites that promoted the hydrogen spillover, and
strong base and moderate acid sites that enhanced the aldol condensation
reactions (Fig. 21) (451). Sun et al. used the HT (Mg-Al) precursors to
prepared Cu or Ni doped porous metal oxides, over which obtained 22%
butanol yield at 56% ethanol conversion and 593 K in a batch reactor.
Fig. 21 Proposed reaction network for ethanol conversion (red indicates main
by-products). Reprinted (adapted) with permission from Pang, J.; Zheng, M.; He, L.; Li, L.;
Pan, X.; Wang, A.; Wang, X.; Zhang, T. J. Catal. 2016, 344, 184–193. Copyright (2016)
Elsevier.
166 Jifeng Pang et al.
Ethanol
DEHYDRATION
Diethyl ether
C2H4
Ethylene
Ethylene
C2H4 surface species
C2H4
C3+HYDROCARBONS
Butene I
surface species C4H8 C4H8 Butene
xC2H4
dimerization of ethylene to butene
C5+ II
Cali C3H6 ,C4H8
aliphatic formation of propene and butene
xC2H4
surface species
III
Aromatics Caro C3H6 Propene
aromatic surface species formation of propene
and there’s still a long way for the controllable synthesis 1,3-BD from ethanol.
The two steps method first partly dehydrogenated ethanol to acetaldehyde,
then converted the unreacted ethanol and acetaldehyde to 1,3-BD. By this
method, robust catalysts are easier to be coupled for this complex reaction,
and the 1,3-BD yield has been improved to some extent. Additionally, eth-
anol is also a key platform chemical for the production of high energy fuels
such as hydrogen, 1-butanol and hydrocarbons. Ethanol is a non-toxic
hydrogen carrier for on-station hydrogen generation. Numerous catalysts
have been developed to improve hydrogen yield by the steaming reforming
reaction, but the present technology still has the problems like catalysts deac-
tivation and high energy consumption. Some novel technologies such as
membrane reactor and chemical looping have also been applied in ethanol
steam reforming for the production of high purity hydrogen. Metal com-
plexes, mixed oxides, HAP and metal enhanced oxide catalysts with strong
base sites and weak acid sites have been developed for converting ethanol to
butanol. Although the activity is still limited under present situations, it pro-
vides an alternative method for ethanol upgrading. Meanwhile, ethanol has
been directly upgraded to hydrocarbons over metal or metal oxide modified
zeolite catalysts. Nevertheless, extensive works should be done on the prod-
uct distributions control and the catalyst stability improvement.
After years of endeavor, the ethanol production has increased dramati-
cally, and several novel routes exhibit potential for commercial applications.
The increased availability of ethanol also stimulates its applications. Ethanol
has been used as a platform chemical for producing important chemicals and
advanced fuels. Moreover, it also triggers fundamental research interests in
both biorefinery and catalysis:
(1) Enzymatic processes vs chemical conversion technologies. Biocatalysis
has been proved to be the most active way for the first generation eth-
anol production. Especially with the development of genetic engineer-
ing, the mild enzymatic process has been demonstrated to be a potential
way for the second generation ethanol production. However, the bio-
catalysis process has a lower efficiency as compared with chemical ones.
In this regards, chemical processes are also promising for ethanol pro-
duction in spite of its multiple step reactions. For instance, the conver-
sion of syngas to ethanol via DME carbonylation process has been
applied at a pilot scale in China. Even though these chemical processes
are still in a cradle, it exhibits high potential for future ethanol produc-
tion. Another example is the catalytic conversion of biomass to ethanol,
which emerges recently but shows high potential for ethanol
172 Jifeng Pang et al.
Acknowledgments
This work was supported by the National Science Foundation of China (21690081,
21690084, 21721004 and 21776268), “Transformational Technologies for Clean Energy
and Demonstration,” Strategic Priority Research Program of the Chinese Academy of
Sciences, Grant No. XDA 21060200. The authors would like to thank Prof. Junming
Sun and Yong Wang from Washington State University, Dr. Joby Sebastian and Prof.
Aiqin Wang from Dalian Institute of Chemical Physics, for their suggestions and fruitful
discussions. Dedicated to the 70th anniversary of Dalian Institute of Chemical Physics, CAS.
References
1. Asif, M.; Muneer, T. Renew. Sust. Energ. Rev. 2007, 11 (7), 1388–1413.
2. Dresselhaus, M. S.; Thomas, I. L. Nature 2001, 414 (6861), 332–337.
3. Abdmouleh, Z.; Alammari, R. A. M.; Gastli, A. Renew. Sust. Energ. Rev. 2015, 45,
249–262.
4. Himmel, M. E.; Ding, S. Y.; Johnson, D. K.; Adney, W. S.; Nimlos, M. R.;
Brady, J. W.; Foust, T. D. Science 2007, 315 (5813), 804–807.
5. Wu, C. Z.; Yin, X. L.; Yuan, Z. H.; Zhou, Z. Q.; Zhuang, X. S. Energy 2010, 35 (11),
4445–4450.
6. Goldemberg, J. Science 2007, 315 (5813), 808–810.
7. Gupta, A.; Verma, J. P. Renew. Sust. Energ. Rev. 2015, 41, 550–567.
8. Sun, J.; Wang, Y. ACS Catal. 2014, 4 (4), 1078–1090.
9. Kujawska, A.; Kujawski, J.; Bryjak, M.; Kujawski, W. Renew. Sust. Energ. Rev. 2015,
48, 648–661.
10. Soleas, G. J.; Diamandis, E. P.; Goldberg, D. M. J. Clin. Lab. Anal. 1997, 11 (5),
287–313.
11. Rosillo-Calle, F.; Walter, A. Energy Sustain. Dev. 2006, 10 (1), 20–32.
12. Mussatto, S. I.; Dragone, G.; Guimarães, P. M. R.; Silva, J. P. A.; Carneiro, L. M.;
Roberto, I. C.; Vicente, A.; Domingues, L.; Teixeira, J. A. Biotechnol. Adv. 2010,
28 (6), 817–830.
13. Augusto Horta Nogueira, L.; Silva Capaz, R. Glob. Food Sec. 2013, 2 (2), 117–125.
14. Li, J.; Kazakov, A.; Dryer, F. L. J. Phys. Chem. A 2004, 108 (38), 7671–7680.
15. Stevenson, B. J.; Liu, J.-W.; Kuchel, P. W.; Ollis, D. L. J. Biotechnol. 2012, 157 (1),
113–123.
16. Maki, Y.; Sato, K.; Isobe, A.; Iwasa, N.; Fujita, S.; Shimokawabe, M.; Takezawa, N.
Appl. Catal. A Gen. 1998, 170 (2), 269–275.
17. Zhang, K.; Zhang, H. T.; Ma, H. F.; Ying, W. Y.; Fang, D. Y. Catal. Lett. 2014, 144
(4), 691–701.
18. Liu, P.; Hensen, E. J. M. J. Am. Chem. Soc. 2013, 135 (38), 14032–14035.
19. Angelici, C.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ChemSusChem 2013, 6 (9),
1595–1614.
20. Ramasamy, K. K.; Wang, Y. J. Energy. Chem. 2013, 22 (1), 65–71.
21. Sun, J.; Zhu, K.; Gao, F.; Wang, C.; Liu, J.; Peden, C. H. F.; Wang, Y. J. Am. Chem.
Soc. 2011, 133 (29), 11096–11099.
22. Baeyens, J.; Kang, Q.; Appels, L.; Dewil, R.; Lv, Y.; Tan, T. Prog. Energ. Combust. Sci.
2015, 47, 60–88.
23. Bshish, A.; Yakoob, Z.; Narayanan, B.; Ramakrishnan, R.; Ebshish, A. Chem. Pap.
2011, 65 (3), 251–266.
24. Makshina, E. V.; Dusselier, M.; Janssens, W.; Degreve, J.; Jacobs, P. A.; Sels, B. F.
Chem. Soc. Rev. 2014, 43 (22), 7917–7953.
Synthesis of ethanol and its catalytic conversion 175
25. Eagan, N. M.; Kumbhalkar, M. D.; Buchanan, J. S.; Dumesic, J. A.; Huber, G. W. Nat.
Rev. Chem. 2019, 3 (4), 223–249.
26. Gerbens-Leenes, W.; Hoekstra, A. Y. Environ. Int. 2012, 40, 202–211.
27. Torija, M. J.; Rozès, N.; Poblet, M.; Guillamón, J. M.; Mas, A. Int. J. Food Microbiol.
2003, 80 (1), 47–53.
28. Lin, Y.; Zhang, W.; Li, C. J.; Sakakibara, K.; Tanaka, S.; Kong, H. N. Biomass Bioenergy
2012, 47, 395–401.
29. Zhang, N.; Steven Green, V.; Ge, X.; Savary, B. J.; Xu, J. Bioresour. Technol. 2014, 155,
189–197.
30. Inal, M.; Yiğitoğlu, M. J. Chem. Technol. Biot. 2011, 86 (12), 1548–1554.
31. Murphy, M.; Carey, C., Eds.; Avian Energetics and Nutritional Ecology; Springer US,
1996; pp 31–60.
32. Bai, F. W.; Anderson, W. A.; Moo-Young, M. Biotechnol. Adv. 2008, 26 (1),
89–105.
33. Koutinas, A. A.; Vlysidis, A.; Pleissner, D.; Kopsahelis, N.; Lopez Garcia, I.;
Kookos, I. K.; Papanikolaou, S.; Kwan, T. H.; Lin, C. S. K. Chem. Soc. Rev. 2014,
43 (8), 2587–2627.
34. Semkiv, M. V.; Dmytruk, K. V.; Abbas, C. A.; Sibirny, A. A. BMC Biotechnol. 2014, 14
(1), 1–9.
35. Della-Bianca, B.; Basso, T.; Stambuk, B.; Basso, L.; Gombert, A. Appl. Microbiol. Biot.
2013, 97 (3), 979–991.
36. Gombert, A. K.; van Maris, A. J. A. Curr. Opin. Biotech. 2015, 33, 81–86.
37. Devantier, R.; Scheithauer, B.; Villas-B^ oas, S. G.; Pedersen, S.; Olsson, L. Biotechnol.
Bioeng. 2005, 90 (6), 703–714.
38. Naik, P. V.; Bernstein, R.; Vankelecom, I. F. J. J. Appl. Polym. Sci. 2016, 133 (28)
43670.
39. Abels, C.; Carstensen, F.; Wessling, M. J. Membrane Sci. 2013, 444, 285–317.
40. Jonker, J. G. G.; Junginger, H. M.; Verstegen, J. A.; Lin, T.; Rodrı́guez, L. F.;
Ting, K. C.; Faaij, A. P. C.; van der Hilst, F. Appl. Energy 2016, 173, 494–510.
41. Chen, W.; Wu, F.; Zhang, J. Renew. Energy 2016, 85, 939–944.
42. Naik, S. N.; Goud, V. V.; Rout, P. K.; Dalai, A. K. Renew. Sust. Energ. Rev. 2010, 14
(2), 578–597.
43. Pang, J.; Zheng, M.; Li, X.; Sebastian, J.; Jiang, Y.; Zhao, Y.; Wang, A.; Zhang, T.
ACS Sust. Chem. Eng. 2019, 7 (1), 679–687.
44. Pang, J.; Zheng, M.; Wang, A.; Sun, R.; Wang, H.; Jiang, Y.; Zhang, T. AIChE J.
2014, 60 (6), 2254–2262.
45. Lynd, L. R.; Liang, X.; Biddy, M. J.; Allee, A.; Cai, H.; Foust, T.; Himmel, M. E.;
Laser, M. S.; Wang, M.; Wyman, C. E. Curr. Opin. Biotech. 2017, 45, 202–211.
46. Pandiyan, K.; Singh, A.; Singh, S.; Saxena, A. K.; Nain, L. Renew. Energy 2019, 132,
723–741.
47. Singh, R.; Shukla, A.; Tiwari, S.; Srivastava, M. Renew. Sust. Energ. Rev. 2014, 32,
713–728.
48. Yang, B.; Wyman, C. E. Biofuel. Bioprod. Bior. 2008, 2 (1), 26–40.
49. Klein-Marcuschamer, D.; Oleskowicz-Popiel, P.; Simmons, B. A.; Blanch, H. W. Bio-
technol. Bioeng. 2012, 109 (4), 1083–1087.
50. Koppram, R.; Tomas-Pejo, E.; Xiros, C.; Olsson, L. Trends Biotechnol. 2014, 32 (1),
46–53.
51. [Link]
52. Joelsson, E.; Erdei, B.; Galbe, M.; Wallberg, O. Biotechnol. Biofuels 2016, 9 (1), 1–16.
53. Mika, L. T.; Csefalvay, E.; Nemeth, Á. Chem. Rev. 2018, 118 (2), 505–613.
54. Momose, H.; Kusumoto, K.; Izumi, Y.; Mizutani, Y. J. Catal. 1982, 77 (1), 23–31.
55. Llano-Restrepo, M.; Muñoz-Muñoz, Y. M. Fluid Phase Equilib. 2011, 307 (1), 45–57.
176 Jifeng Pang et al.
56. Katada, N.; Iseki, Y.; Shichi, A.; Fujita, N.; Ishino, I.; Osaki, K.; Torikai, T.; Niwa, M.
Appl. Catal. A Gen. 2008, 349 (1–2), 55–61.
57. Chu, W.; Echizen, T.; Kamiya, Y.; Okuhara, T. Appl. Catal. A Gen. 2004, 259 (2),
199–205.
58. Jin, E.; Zhang, Y.; He, L.; Harris, H. G.; Teng, B.; Fan, M. Appl. Catal. A Gen. 2014,
476, 158–174.
59. Molino, A.; Chianese, S.; Musmarra, D. J. Energy. Chem. 2016, 25 (1), 10–25.
60. Chuang, S. S. C.; Zhang, L. In Handbook of Climate Change Mitigation and Adaptation;
Chen, W.-Y., Suzuki, T., Lackner, M., Eds.; Springer New York: New York, NY,
2016; pp 1–18.
61. Luk, H. T.; Mondelli, C.; Ferre, D. C.; Stewart, J. A.; Perez-Ramirez, J. Chem. Soc.
Rev. 2017, 46 (5), 1358–1426.
62. Spivey, J. J.; Egbebi, A. Chem. Soc. Rev. 2007, 36 (9), 1514–1528.
63. Subramani, V.; Gangwal, S. K. Energy Fuel. 2008, 22 (2), 814–839.
64. Schumann, J.; Medford, A. J.; Yoo, J. S.; Zhao, Z.-J.; Bothra, P.; Cao, A.; Studt, F.;
Abild-Pedersen, F.; Nørskov, J. K. ACS Catal. 2018, 8 (4), 3447–3453.
65. Ichikawa, M.; Fukushima, T. J. Chem. Soc. Chem. Commun. 1985, 6, 321–323.
66. Haider, M. A.; Gogate, M. R.; Davis, R. J. J. Catal. 2009, 261 (1), 9–16.
67. Medford, A. J.; Vojvodic, A.; Hummelshøj, J. S.; Voss, J.; Abild-Pedersen, F.; Studt, F.;
Bligaard, T.; Nilsson, A.; Nørskov, J. K. J. Catal. 2015, 328, 36–42.
68. Vojvodic, A.; Norskov, J. K. Natl. Sci. Rev. 2015, 2 (2), 140–143.
69. Cao, A.; Schumann, J.; Wang, T.; Zhang, L.; Xiao, J.; Bothra, P.; Liu, Y.; Abild-
Pedersen, F.; Nørskov, J. K. ACS Catal. 2018, 8 (11), 10148–10155.
70. Bhasin, M. M.; O’Connor, G. L. Belgian Patent 824,822, 1975, to Union Carbide.
71. Wang, Y.; Luo, H.; Liang, D.; Bao, X. J. Catal. 2000, 196 (1), 46–55.
72. Yu, J.; Mao, D.; Ding, D.; Guo, X.; Lu, G. J. Mol. Catal. A Chem. 2016, 423, 151–159.
73. Lin, P. Z.; Liang, D. B.; Luo, H. Y.; Xu, C. H.; Zhou, H. W.; Huang, S. Y.; Lin, L. W.
Appl. Catal. A Gen. 1995, 131 (2), 207–214.
74. Hu, J.; Wang, Y.; Cao, C.; Elliott, D. C.; Stevens, D. J.; White, J. F. Catal. Today 2007,
120 (1), 90–95.
75. Palomino, R. M.; Magee, J. W.; Llorca, J.; Senanayake, S. D.; White, M. G. J. Catal.
2015, 329, 87–94.
76. Wang, J.; Zhang, Q.; Wang, Y. Catal. Today 2011, 171 (1), 257–265.
77. Xue, F.; Chen, W.; Song, X.; Cheng, X.; Ding, Y. RSC Adv. 2016, 6 (42),
35348–35353.
78. Yang, N.; Medford, A. J.; Liu, X.; Studt, F.; Bligaard, T.; Bent, S. F.; Nørskov, J. K.
J. Am. Chem. Soc. 2016, 138 (11), 3705–3714.
79. Huang, X.; Teschner, D.; Dimitrakopoulou, M.; Fedorov, A.; Frank, B.;
Kraehnert, R.; Rosowski, F.; Kaiser, H.; Schunk, S.; Kuretschka, C.; Schl€ ogl, R.;
Willinger, M. G.; Trunschke, A. Angew. Chem. Int. Ed. 2019, 58 (26), 8709–8713.
80. Li, C.; Liu, J.; Gao, W.; Zhao, Y.; Wei, M. Catal. Lett. 2013, 143 (11), 1247–1254.
81. Katzer, J. R.; Sleight, A. W.; Gajardo, P.; Michel, J. B.; Gleason, E. F.; McMillan, S.
Discuss. Faraday Soc. 1981, 72 (0), 121–133.
82. Liu, Y.; Murata, K.; Inaba, M.; Takahara, I.; Okabe, K. Catal. Today 2011, 164 (1),
308–314.
83. Lopez, L.; Velasco, J.; Montes, V.; Marinas, A.; Cabrera, S.; Boutonnet, M.; J€arås, S.
Catalysts 2015, 5 (4), 1737.
84. Chen, G.; Guo, C.-Y.; Zhang, X.; Huang, Z.; Yuan, G. Fuel Process. Technol. 2011, 92
(3), 456–461.
85. Chen, W.; Ding, Y.; Xue, F.; Song, X.; Ning, L. Catal. Commun. 2016, 85, 44–47.
86. Han, L. P.; Mao, D. S.; Yu, J.; Guo, Q. S.; Lu, G. Z. Appl. Catal. A Gen. 2013, 454,
81–87.
Synthesis of ethanol and its catalytic conversion 177
87. Pan, X. L.; Fan, Z. L.; Chen, W.; Ding, Y. J.; Luo, H. Y.; Bao, X. H. Nat. Mater. 2007,
6 (7), 507–511.
88. Fan, Z. L.; Chen, W.; Pan, X. L.; Bao, X. H. Catal. Today 2009, 147 (2), 86–93.
89. Kim, T. W.; Kim, M. J.; Chae, H. J.; Ha, K. S.; Kim, C. U. Fuel 2015, 160, 393–403.
90. Kim, M. J.; Kim, T. W.; Chae, H. J.; Kim, C. U.; Jeong, S. Y.; Kim, J. R.; Ha, K. S.
J. Nanosci. Nanotechnol. 2016, 16 (2), 2004–2009.
91. Yang, X.; Zhu, X.; Hou, R.; Zhou, L.; Su, Y. Fuel Process. Technol. 2011, 92 (10),
1876–1880.
92. Ning, X.; An, Z.; He, J. J. Catal. 2016, 340, 236–247.
93. Ding, M.; Tu, J.; Qiu, M.; Wang, T.; Ma, L.; Li, Y. Appl. Energy 2015, 138, 584–589.
94. Feng, W.; Wang, Q.; Jiang, B.; Ji, P. Ind. Eng. Chem. Res. 2011, 50 (19), 11067–11072.
95. Wang, P.; Bai, Y.; Xiao, H.; Tian, S.; Zhang, Z.; Wu, Y.; Xie, H.; Yang, G.; Han, Y.;
Tan, Y. Catal. Commun. 2016, 75, 92–97.
96. Gao, W.; Zhao, Y.; Liu, J.; Huang, Q.; He, S.; Li, C.; Zhao, J.; Wei, M. Catal. Sci.
Technol. 2013, 3 (5), 1324–1332.
97. Hilmen, A. M.; Xu, M.; Gines, M. J. L.; Iglesia, E. Appl. Catal. A Gen. 1998, 169 (2),
355–372.
98. Nunan, J. G.; Bogdan, C. E.; Klier, K.; Smith, K. J.; Young, C.-W.; Herman, R. G.
J. Catal. 1988, 113 (2), 410–433.
99. Zhang, L.; Bai, B.; Bai, H.; Huang, W.; Gao, Z.-H.; Zuo, Z.-J.; Lv, Y.-K. Phys. Chem.
Chem. Phys. 2017, 19 (29), 19300–19307.
100. Klier, K.; Herman, R. G.; Nunan, J. G.; Smith, K. J.; Bogdan, C. E.; Young, C.-W.;
Santiesteban, J. G. In Methane Conversion; Bibby, D. M., Chang, C. D., Howe, R. F.,
Yurchak, S., Eds.; Elsevier: Amsterdam, 1988; p. 109.
101. Gupta, M.; Smith, M. L.; Spivey, J. J. ACS Catal. 2011, 1 (6), 641–656.
102. Herman, R. G. Catal. Today 2000, 55 (3), 233–245.
103. Sun, J.; Cai, Q.; Wan, Y.; Wan, S.; Wang, L.; Lin, J.; Mei, D.; Wang, Y. ACS Catal.
2016, 6 (9), 5771–5785.
104. Zhang, R.; Sun, X. Wang, B. J. Phys. Chem. C 2013, 117 (13), 6594–6606.
105. Beiramar, J. M.; Griboval-Constant, A.; Khodakov, A. Y. ChemCatChem 2014, 6 (6),
1788–1793.
106. Li, J.; Gao, Z. H.; Li, S. J.; Zuo, Z. J.; Huang, W. Energy Sources Part A 2016, 38 (16),
2383–2389.
107. Liu, Y.-J.; Zuo, Z.-J.; Li, C.; Deng, X.; Huang, W. Appl. Surf. Sci. 2015, 356,
124–127.
108. Liu, Y.; Liu, C.; Deng, X.; Huang, W. RSC Adv. 2015, 5 (120), 99023–99027.
109. Zaman, S.; Smith, K. J. Catal. Rev. 2012, 54 (1), 41–132.
110. Portillo, M. A.; Perales, A. L. V.; Vidal-Barrero, F.; Campoy, M. Fuel Process. Technol.
2016, 151, 19–30.
111. Dorokhov, V. S.; Kamorin, M. A.; Rozhdestvenskaya, N. N.; Kogan, V. M. C. R.
Chim. 2016, 19 (10), 1184–1193.
112. Xiao, K.; Bao, Z.; Qi, X.; Wang, X.; Zhong, L.; Fang, K.; Lin, M.; Sun, Y. Chin. J.
Catal. 2013, 34 (1), 116–129.
113. Woo, H. C.; Park, T. Y.; Kim, Y. G.; Nam, I.-S.; Lee, J. S.; Chung, J. S. In: Studies in
Surface Science and Catalysis; Guczi, L., Solymosi, F., Tetenyi, P., Eds.; Vol. 75 ; Elsevier,
1993; pp. 2749–2752.
114. Li, Z.; Fu, Y.; Jiang, M. Appl. Catal. A Gen. 1999, 187 (2), 187–198.
115. Li, Z.; Fu, Y.; Bao, J.; Jiang, M.; Hu, T.; Liu, T.; Xie, Y. Appl. Catal. A Gen. 2001, 220
(1–2), 21–30.
116. Andersson, R.; Boutonnet, M.; J€arås, S. Fuel 2013, 107, 715–723.
117. Konarova, M.; Tang, F.; Chen, J.; Wang, G.; Rudolph, V.; Beltramini, J.
ChemCatChem 2014, 6 (8), 2394–2402.
178 Jifeng Pang et al.
118. Menart, M. J.; Hensley, J. E.; Costelow, K. E. Appl. Catal. A Gen. 2012, 437–438, 36–43.
119. Hensley, J. E.; Pylypenko, S.; Ruddy, D. A. J. Catal. 2014, 309, 199–208.
120. Yang, Y.; Wang, Y.; Liu, S.; Song, Q.; Xie, Z.; Gao, Z. Catal. Lett. 2009, 127 (3),
448–455.
121. Xie, W.; Zhou, J.; Ji, L.; Sun, S.; Pan, H.; Zhu, J.; Gao, C.; Bao, J. RSC Adv. 2016, 6
(45), 38741–38745.
122. Molitor, B.; Richter, H.; Martin, M. E.; Jensen, R. O.; Juminaga, A.; Mihalcea, C.;
Angenent, L. T. Bioresour. Technol. 2016, 215, 386–396.
123. Drake, H. L.; Gossner, A. S.; Daniel, S. L. In: Incredible Anaerobes: From Physiology to
Genomics to Fuels; Wiegel, J., Maier, R. J., Adams, M. W. W., Eds.; Vol. 1125 ;
Wiley-Blackwell, 2008; pp. 100–128.
124. Mohammadi, M.; Najafpour, G. D.; Younesi, H.; Lahijani, P.; Uzir, M. H.;
Mohamed, A. R. Renew. Sust. Energ. Rev. 2011, 15 (9), 4255–4273.
125. Koepke, M.; Held, C.; Hujer, S.; Liesegang, H.; Wiezer, A.; Wollherr, A.;
Ehrenreich, A.; Liebl, W.; Gottschalk, G.; Duerre, P. PNAS 2010, 107 (29),
13087–13092.
126. Griffin, D. W.; Schultz, M. A. Environ. Prog. Sustain. 2012, 31 (2), 219–224.
127. Ljungdah, L. G.; Wood, H. G. Annu. Rev. Microbiol. 1969, 23, 515-&.
128. Wood, H. G. FASEB J. 1991, 5 (2), 156–163.
129. Martin, M. E.; Richter, H.; Saha, S.; Angenent, L. T. Biotechnol. Bioeng. 2016, 113 (3),
531–539.
130. Ragsdale, S. W.; Pierce, E. BBA Proteins Proteom. 2008, 1784 (12), 1873–1898.
131. Drake, H. L.; G€ oßner, A. S.; Daniel, S. L. Ann. N. Y. Acad. Sci. 2008, 1125 (1),
100–128.
132. Singla, A.; Verma, D.; Lal, B.; Sarma, P. M. Bioresour. Technol. 2014, 172, 41–49.
133. Liu, K.; Atiyeh, H. K.; Tanner, R. S.; Wilkins, M. R.; Huhnke, R. L. Bioresour.
Technol. 2012, 104, 336–341.
134. Mohammadi, M.; Younesi, H.; Najafpour, G.; Mohamed, A. R. J. Chem. Technol. Biot.
2012, 87 (6), 837–843.
135. Richter, H.; Martin, M. E.; Angenent, L. T. Energies 2013, 6 (8), 3987–4000.
136. Phillips, J. R.; Klasson, K. T.; Clausen, E. C.; Gaddy, J. L. Appl. Biochem. Biotech. 1993,
39 (1), 559–571.
137. Munasinghe, P. C.; Khanal, S. K. Bioresour. Technol. 2010, 101 (13), 5013–5022.
138. Abubackar, H. N.; Veiga, M. C.; Kennes, C. Bioresour. Technol. 2015, 186, 122–127.
139. Abubackar, H. N.; Bengelsdorf, F. R.; D€ urre, P.; Veiga, M. C.; Kennes, C. Appl.
Energy 2016, 169, 210–217.
140. Gao, J.; Atiyeh, H. K.; Phillips, J. R.; Wilkins, M. R.; Huhnke, R. L. Bioresour. Technol.
2013, 147, 508–515.
141. Munasinghe, P. C.; Khanal, S. K. Biotechnol. Progr. 2010, 26 (6), 1616–1621.
142. Abubackar, H. N.; Veiga, M. C.; Kennes, C. Biofuel Bioprod. Bior. 2011, 5 (1), 93–114.
143. Ungerman, A. J.; Heindel, T. J. Biotechnol. Progr. 2007, 23 (3), 613–620.
144. Devarapalli, M.; Atiyeh, H. K.; Phillips, J. R.; Lewis, R. S.; Huhnke, R. L. Bioresour.
Technol. 2016, 209, 56–65.
145. Kim, Y.-K.; Lee, H. Bioresour. Technol. 2016, 204, 139–144.
146. Bozzano, G.; Manenti, F. Prog. Energ. Combust. 2016, 56, 71–105.
147. Paulik, F. E.; Roth, J. F. Chem. Commun. 1968, (24), 1578.
148. Thomas, C. M.; S€ uss-Fink, G. Coord. Chem. Rev. 2003, 243 (1–2), 125–142.
149. Budiman, A. W.; Nam, J. S.; Park, J. H.; Mukti, R. I.; Chang, T. S.; Bae, J. W.;
Choi, M. J. Catal. Surv. Asia 2016, 20 (3), 173–193.
150. Haynes, A. In Catalytic Carbonylation Reactions; Beller, M. Ed.; Vol. 18, Springer, 2006;
pp. 179–205.
Synthesis of ethanol and its catalytic conversion 179
151. Haynes, A. In: Advances in Catalysis; Gates, B. C., Knozinger, H., Jentoft, F. C., Eds.;
Vol. 53 ; Academic Press Inc, 2010; pp. 1–45.
152. Merenov, A. S.; Abraham, M. A. Catal. Today 1998, 40 (4), 397–404.
153. Corma, A.; Garcia, H. Adv. Synth. Catal. 2006, 348 (12–13), 1391–1412.
154. De Blasio, N.; Tempesti, E.; Kaddouri, A.; Mazzocchia, C.; Cole-Hamilton, D. J.
J. Catal. 1998, 176 (1), 253–259.
155. Jones, C. W. Top. Catal. 2010, 53 (13), 942–952.
156. Yashima, T.; Orikasa, Y.; Takahashi, N.; Hara, N. J. Catal. 1979, 59 (1), 53–60.
157. Saikia, P. K.; Sarmah, P. P.; Borah, B. J.; Saikia, L.; Dutta, D. K. J. Mol. Catal. A Chem.
2016, 412, 27–33.
158. Pang, F.; Song, F.; Zhang, Q.; Tan, Y.; Han, Y. Chem. Eng. J. 2016, 293, 129–138.
159. Liu, T. C.; Chiu, S. J. Ind. Eng. Chem. Res. 1994, 33 (3), 488–492.
160. Kwak, J. H.; Dagle, R.; Tustin, G. C.; Zoeller, J. R.; Allard, L. F.; Wang, Y. J. Phys.
Chem. Lett. 2014, 5 (3), 566–572.
161. Li, F.; Chen, B.; Huang, Z.; Lu, T.; Yuan, Y.; Yuan, G. Green Chem. 2013, 15 (6),
1600–1607.
162. Ellis, B.; Howard, M. J.; Joyner, R. W.; Reddy, K. N.; Padley, M. B.; Smith, W. J.
In Studies in Surface Science and Catalysis; Hightower, J. W., Delgass, W. N.,
Iglesia, E., Alexis, T. B., Eds.; Vol. 101 ; Elsevier, 1996; pp. 771–779.
163. Blasco, T.; Boronat, M.; Concepción, P.; Corma, A.; Law, D.; Vidal-Moya, J. A.
Angew. Chem. Int. Ed. 2007, 46 (21), 3938–3941.
164. Cheung, P.; Bhan, A.; Sunley, G. J.; Iglesia, E. Angew. Chem. Int. Ed. 2006, 45 (10),
1617–1620.
165. Cheung, P.; Bhan, A.; Sunley, G. J.; Law, D. J.; Iglesia, E. J. Catal. 2007, 245 (1),
110–123.
166. Bhan, A.; Allian, A. D.; Sunley, G. J.; Law, D. J.; Iglesia, E. J. Am. Chem. Soc. 2007, 129
(16), 4919–4924.
167. Liu, J.; Xue, H.; Huang, X.; Wu, P. H.; Huang, S.-J.; Liu, S. B.; Shen, W. Chin. J.
Catal. 2010, 31 (7), 729–738.
168. Liu, J.; Xue, H.; Huang, X.; Li, Y.; Shen, W. Catal. Lett. 2010, 139 (1), 33–37.
169. Xue, H.; Huang, X.; Zhan, E.; Ma, M.; Shen, W. Catal. Commun. 2013, 37, 75–79.
170. Rasmussen, D. B.; Christensen, J. M.; Temel, B.; Studt, F.; Moses, P. G.; Rossmeisl, J.;
Riisager, A.; Jensen, A. D. Angew. Chem. Int. Ed. 2015, 54 (25), 7261–7264.
171. Lu, X. China Creates World’s First Coal-to-Ethanol Production Line; 2017, http://
[Link]/ns_17179/ue/201703/t20170320_175086.html.
172. Vidra, A.; Nemeth, A. Period. Polytech. Chem. Eng. 2018, 62 (3), 245–256.
173. Rachmady, W.; Vannice, M. A. J. Catal. 2000, 192 (2), 322–334.
174. Rachmady, W.; Vannice, M. A. J. Catal. 2002, 209 (1), 87–98.
175. Olcay, H.; Xu, L.; Xu, Y.; Huber, G. W. ChemCatChem 2010, 2 (11), 1420–1424.
176. Shangguan, J.; Olarte, M. V.; Chin, Y. H. J. Catal. 2016, 340, 107–121.
177. Alcala, R.; Shabaker, J. W.; Huber, G. W.; Sanchez-Castillo, M. A.; Dumesic, J. A.
J. Phys. Chem. B 2005, 109 (6), 2074–2085.
178. Zhang, S.; Duan, X.; Ye, L.; Lin, H.; Xie, Z.; Yuan, Y. Catal. Today 2013, 215,
260–266.
179. Zhou, M.; Zhang, H.; Ma, H.; Ying, W. Fuel Process. Technol. 2016, 144, 115–123.
180. Xu, G.; Zhang, J.; Wang, S.; Zhao, Y.; Ma, X. Front. Chem. Sci. Eng. 2016, 10 (3),
417–424.
181. Dong, X.; Lei, J.; Chen, Y.; Jiang, H.; Zhang, M. Appl. Catal. B Environ. 2019, 244,
448–458.
182. Adkins, H.; Folkers, K. J. Am. Chem. Soc. 1931, 53 (3), 1095–1097.
183. Shih, Y. S.; Jen, C. M. J. Chin. Chem. Soc. 1984, 31 (3), 301–305.
180 Jifeng Pang et al.
216. Zhao, S.; Yue, H.; Zhao, Y.; Wang, B.; Geng, Y.; Lv, J.; Wang, S.; Gong, J.; Ma, X.
J. Catal. 2013, 297, 142–150.
217. Zhao, Y.; Zhao, S.; Geng, Y.; Shen, Y.; Yue, H.; Lv, J.; Wang, S.; Ma, X. Catal. Today
2016, 276, 28–35.
218. Zhu, Y.; Kong, X.; Li, X.; Ding, G.; Zhu, Y.; Li, Y. W. ACS Catal. 2014, 4 (10),
3612–3620.
219. Zhu, Y.; Kong, X.; Zhu, S.; Dong, F.; Zheng, H.; Zhu, Y.; Li, Y. W. Appl. Catal.
B Environ. 2015, 166–167, 551–559.
220. Yang, M.; Qi, H.; Liu, F.; Ren, Y.; Pan, X.; Zhang, L.; Liu, X.; Wang, H.; Pang, J.;
Zheng, M.; Wang, A.; Zhang, T. Joule 2019, 3, 1937–1948.
221. Xu, G.; Wang, A.; Pang, J.; Zhao, X.; Xu, J.; Lei, N.; Wang, J.; Zheng, M.; Yin, J.;
Zhang, T. ChemSusChem 2017, 10 (7), 1390–1394.
222. Yang, C.; Miao, Z.; Zhang, F.; Li, L.; Liu, Y.; Wang, A.; Zhang, T. Green Chem. 2018,
20 (9), 2142–2150.
223. Wang, A.; Zhang, T. Acc. Chem. Res. 2013, 46 (7), 1377–1386.
224. Zheng, M.; Pang, J.; Sun, R.; Wang, A.; Zhang, T. ACS Catal. 2017, 7 (3),
1939–1954.
225. Song, H.; Wang, P.; Li, S.; Deng, W.; Li, Y.; Zhang, Q.; Wang, Y. Chem. Commun.
2019, 55 (30), 4303–4306.
226. Li, C.; Xu, G.; Wang, C.; Ma, L.; Qiao, Y.; Zhang, Y.; Fu, Y. Green Chem. 2019, 21
(9), 2234–2239.
227. Hamelinck, C. N.; Hooijdonk, G. V.; Faaij, A. P. C. Biomass Bioenergy 2005, 28 (4),
384–410.
228. Kundiyana, D. K.; Huhnke, R. L.; Wilkins, M. R. J. Biosci. Bioeng. 2010, 109 (5),
492–498.
229. Daniell, J.; K€ opke, M.; Simpson, S. Energies 2012, 5 (12), 5372.
230. Niven, R. K. Renew. Sust. Energ. Rev. 2005, 9 (6), 535–555.
231. Kohse-H€ oinghaus, K.; Oßwald, P.; Cool, T. A.; Kasper, T.; Hansen, N.; Qi, F.;
Westbrook, C. K.; Westmoreland, P. R. Angew. Chem. Int. Ed. 2010, 49 (21),
3572–3597.
232. Bergthorson, J. M.; Thomson, M. J. Renew. Sust. Energ. Rev. 2015, 42, 1393–1417.
233. Takei, T.; Iguchi, N.; Haruta, M. Catal. Surv. Asia 2011, 15 (2), 80–88.
234. Cunningham, J.; Al-Sayyed, G. H.; Cronin, J. A.; Fierro, J. L. G.; Healy, C.;
Hirschwald, W.; Ilyas, M.; Tobin, J. P. J. Catal. 1986, 102 (1), 160–171.
235. Chladek, P.; Croiset, E.; Epling, W.; Hudgins, R. R. Can. J. Chem. Eng. 2007, 85 (6),
917–924.
236. Chang, F. W.; Kuo, W. Y.; Lee, K. C. Appl. Catal. A Gen. 2003, 246 (2), 253–264.
237. Chang, F. W.; Kuo, W. Y.; Yang, H. C. Appl. Catal. A Gen. 2005, 288 (1–2),
53–61.
238. Chang, F. W.; Yang, H. C.; Roselin, L. S.; Kuo, W. Y. Appl. Catal. A Gen. 2006, 304,
30–39.
239. Wang, Q. N.; Shi, L.; Lu, A. H. ChemCatChem 2015, 7 (18), 2846–2852.
240. Morales, M. V.; Asedegbega-Nieto, E.; Bachiller-Baeza, B.; Guerrero-Ruiz, A. Carbon
2016, 102, 426–436.
241. Nobuhiro, I.; Nobutsune, T. Bull. Chem. Soc. Jpn 1991, 64 (9), 2619–2623.
242. Fujita, S. I.; Iwasa, N.; Tani, H.; Nomura, W.; Arai, M.; Takezawa, N. React. Kinet.
Catal. Lett. 2001, 73 (2), 367–372.
243. Qiao, B.; Wang, A.; Yang, X.; Allard, L. F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T.
Nat. Chem. 2011, 3, 634.
244. Shan, J.; Liu, J.; Li, M.; Lustig, S.; Lee, S.; Flytzani-Stephanopoulos, M. Appl. Catal.
B Environ. 2018, 226, 534–543.
182 Jifeng Pang et al.
245. Shan, J.; Janvelyan, N.; Li, H.; Liu, J.; Egle, T. M.; Ye, J.; Biener, M. M.; Biener, J.;
Friend, C. M.; Flytzani-Stephanopoulos, M. Appl. Catal. B Environ. 2017, 205,
541–550.
246. Guan, Y.; Hensen, E. J. M. Appl. Catal. A Gen. 2009, 361 (1–2), 49–56.
247. Gazsi, A.; Koós, A.; Bánsági, T.; Solymosi, F. Catal. Today 2011, 160 (1), 70–78.
248. Wang, C.; Garbarino, G.; Allard, L. F.; Wilson, F.; Busca, G.; Flytzani-
Stephanopoulos, M. ACS Catal. 2016, 6 (1), 210–218.
249. Abu-Zied, B. M. Appl. Catal. A Gen. 2000, 198 (1–2), 139–153.
250. Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. ChemCatChem 2013, 5 (8), 2367–2373.
251. Iwasa, N.; Yamamoto, O.; Tamura, R.; Nishikubo, M.; Takezawa, N. Catal. Lett.
1999, 62 (2), 179–184.
252. Li, S.; Wang, W.; Liu, X.; Zeng, X.; Li, W.; Tsubaki, N.; Yu, S. RSC Adv. 2016, 6
(16), 13450–13455.
253. Colley, S. W.; Tabatabaei, J.; Waugh, K. C.; Wood, M. A. J. Catal. 2005, 236 (1),
21–33.
254. Santacesaria, E.; Carotenuto, G.; Tesser, R.; Di Serio, M. Chem. Eng. J. 2012, 179,
209–220.
255. Inui, K.; Kurabayashi, T.; Sato, S.; Ichikawa, N. J. Mol. Catal. A Chem. 2004, 216 (1),
147–156.
256. Carotenuto, G.; Tesser, R.; Di Serio, M.; Santacesaria, E. Catal. Today 2013, 203,
202–210.
257. Yu, X.; Zhu, W.; Gao, S.; Chen, L.; Yuan, H.; Luo, J.; Wang, Z.; Zhang, W. Chem.
Res. Chin. Univ. 2013, 29 (5), 986–990.
258. Freitas, I. C.; Damyanova, S.; Oliveira, D. C.; Marques, C. M. P.; Bueno, J. M. C.
J. Mol. Catal. A Chem. 2014, 381, 26–37.
259. Sato, A. G.; Volanti, D. P.; Meira, D. M.; Damyanova, S.; Longo, E.; Bueno, J. M. C.
J. Catal. 2013, 307, 1–17.
260. Sánchez, A. B.; Homs, N.; Fierro, J. L. G.; Piscina, P. R. D. L. Catal. Today 2005,
107–108, 431–435.
261. Masatake, H.; Tetsuhiko, K.; Hiroshi, S.; Nobumasa, Y. Chem. Lett. 1987, 16 (2),
405–408.
262. Zheng, N.; Stucky, G. D. J. Am. Chem. Soc. 2006, 128 (44), 14278–14280.
263. Kong, X.-M.; Shen, L.-L. Catal. Commun. 2012, 24, 34–37.
264. Takei, T.; Iguchi, N.; Haruta, M. New J. Chem. 2011, 35 (10), 2227–2233.
265. Sobolev, V. I.; Koltunov, K. Y.; Simakova, O. A.; Leino, A.-R.; Murzin, D. Y. Appl.
Catal. A Gen. 2012, 433–434, 88–95.
266. Mielby, J.; Abildstrøm, J. O.; Wang, F.; Kasama, T.; Weidenthaler, C.; Kegnæs, S.
Angew. Chem. 2014, 126 (46), 12721–12724.
267. Bauer, J. C.; Veith, G. M.; Allard, L. F.; Oyola, Y.; Overbury, S. H.; Dai, S. ACS Catal.
2012, 2 (12), 2537–2546.
268. Redina, E. A.; Greish, A. A.; Mishin, I. V.; Kapustin, G. I.; Tkachenko, O. P.;
Kirichenko, O. A.; Kustov, L. M. Catal. Today 2015, 241 (Pt. B), 246–254.
269. Du, X.; Fu, N.; Zhang, S.; Chen, C.; Wang, D.; Li, Y. Nano Res. 2016, 9 (9),
2681–2686.
270. Gong, J.; Mullins, C. B. J. Am. Chem. Soc. 2008, 130 (49), 16458.
271. Guan, Y.; Hensen, E. J. M. J. Catal. 2013, 305, 135–145.
272. Chimentão, R. J.; Herrera, J. E.; Kwak, J. H.; Medina, F.; Wang, Y.; Peden, C. H. F.
Appl. Catal. A Gen. 2007, 332 (2), 263–272.
273. Beck, B.; Harth, M.; Hamilton, N. G.; Carrero, C.; Uhlrich, J. J.; Trunschke, A.;
Shaikhutdinov, S.; Schubert, H.; Freund, H.-J.; Schl€ ogl, R.; Sauer, J.;
Schom€acker, R. J. Catal. 2012, 296, 120–131.
274. Lin, Y.-C.; Chang, C.-H.; Chen, C.-C.; Jehng, J.-M.; Shyu, S.-G. Catal. Commun.
2008, 9 (5), 675–679.
Synthesis of ethanol and its catalytic conversion 183
275. Hidalgo, J. M.; Tišler, Z.; Kubicka, D.; Raabova, K.; Bulanek, R. J. Mol. Catal.
A Chem. 2016, 420, 178–189.
276. Wang, J.; Huang, R.; Feng, Z.; Liu, H.; Su, D. ChemSusChem 2016, 9 (14),
1820–1826.
277. Khanmohammadi, M.; Amani, S.; Garmarudi, A. B.; Niaei, A. Chin. J. Catal. 2016, 37
(3), 325–339.
278. Iwamoto, M.; Tanaka, M.; Hirakawa, S.; Mizuno, S.; Kurosawa, M. ACS Catal. 2014,
4 (10), 3463–3469.
279. Iwamoto, M.; Kasai, K.; Haishi, T. ChemSusChem 2011, 4 (8), 1055–1058.
280. Iwamoto, M. Catal. Today 2015, 242 (Pt. B), 243–248.
281. Song, Z.; Takahashi, A.; Mimura, N.; Fujitani, T. Catal. Lett. 2009, 131 (3), 364–369.
282. Song, Z.; Takahashi, A.; Nakamura, I.; Fujitani, T. Appl. Catal. A Gen. 2010, 384
(1–2), 201–205.
283. Xia, W.; Takahashi, A.; Nakamura, I.; Shimada, H.; Fujitani, T. J. Mol. Catal. A Chem.
2010, 328 (1–2), 114–118.
284. Takahashi, A.; Xia, W.; Wu, Q.; Furukawa, T.; Nakamura, I.; Shimada, H.;
Fujitani, T. Appl. Catal. A Gen. 2013, 467, 380–385.
285. Goto, D.; Harada, Y.; Furumoto, Y.; Takahashi, A.; Fujitani, T.; Oumi, Y.;
Sadakane, M.; Sano, T. Appl. Catal. A Gen. 2010, 383 (1–2), 89–95.
286. Furumoto, Y.; Harada, Y.; Tsunoji, N.; Takahashi, A.; Fujitani, T.; Ide, Y.;
Sadakane, M.; Sano, T. Appl. Catal. A Gen. 2011, 399 (1–2), 262–267.
287. Furumoto, Y.; Tsunoji, N.; Ide, Y.; Sadakane, M.; Sano, T. Appl. Catal. A Gen. 2012,
417–418, 137–144.
288. Inoue, K.; Okabe, K.; Inaba, M.; Takahara, I.; Murata, K. React. Kinet. Mech. Cat.
2010, 101 (2), 477–489.
289. Song, Z.; Liu, W.; Chen, C.; Takahashi, A.; Fujitani, T. React. Kinet. Mech. Cat. 2013,
109 (1), 221–231.
290. Huangfu, J.; Mao, D.; Zhai, X.; Guo, Q. Appl. Catal. A Gen. 2016, 520, 99–104.
291. Xia, W.; Wang, F.; Mu, X.; Chen, K.; Takahashi, A.; Nakamura, I.; Fujitani, T. Catal.
Commun. 2017, 91, 62–66.
292. Meng, T.; Mao, D.; Guo, Q.; Lu, G. Catal. Commun. 2012, 21, 52–57.
293. Takamitsu, Y.; Yamamoto, K.; Yoshida, S.; Ogawa, H.; Sano, T. J. Porous. Mater.
2014, 21 (4), 433–440.
294. Xia, W.; Chen, K.; Takahashi, A.; Li, X.; Mu, X.; Han, C.; Liu, L.; Nakamura, I.;
Fujitani, T. Catal. Commun. 2016, 73, 27–33.
295. Oikawa, H.; Shibata, Y.; Inazu, K.; Iwase, Y.; Murai, K.; Hyodo, S.; Kobayashi, G.;
Baba, T. Appl. Catal. A Gen. 2006, 312, 181–185.
296. Lin, B.; Zhang, Q.; Wang, Y. Ind. Eng. Chem. Res. 2009, 48 (24), 10788–10795.
297. Duan, C.; Zhang, X.; Zhou, R.; Hua, Y.; Zhang, L.; Chen, J. Fuel Process. Technol.
2013, 108, 31–40.
298. Bai, T.; Zhang, X.; Wang, F.; Qu, W.; Liu, X.; Duan, C. J. Energy. Chem. 2016, 25 (3),
545–552.
299. Lehmann, T.; Seidel-Morgenstern, A. Chem. Eng. J. 2014, 242, 422–432.
300. Shouta, M.; Mika, K.; Masashi, T.; Masakazu, I. Chem. Lett. 2012, 41 (9), 892–894.
301. Iwamoto, M.; Mizuno, S.; Tanaka, M. Chem. Eur. J. 2013, 19 (22), 7214–7220.
302. Hayashi, F.; Tanaka, M.; Lin, D.; Iwamoto, M. J. Catal. 2014, 316, 112–120.
303. Xia, W.; Wang, F.; Mu, X.; Chen, K.; Takahashi, A.; Nakamura, I.; Fujitani, T. Catal.
Commun. 2017, 90, 10–13.
304. Nakajima, T.; Tanabe, K.; Yamaguchi, T.; Matsuzaki, I.; Mishima, S. Appl. Catal.
1989, 52 (3), 237–248.
305. Hino, M.; Arata, K. J. Chem. Soc. Chem. Commun. 1988, 17, 1168–1169.
306. Nakajima, T.; Yamaguchi, T.; Tanabe, K. J. Chem. Soc. Chem. Commun. 1987, 6,
394–395.
184 Jifeng Pang et al.
307. Nishiguchi, T.; Matsumoto, T.; Kanai, H.; Utani, K.; Matsumura, Y.; Shen, W.-J.;
Imamura, S. Appl. Catal. A Gen. 2005, 279 (1), 273–277.
308. Rodrigues, C. P.; Zonetti, P. C.; Silva, C. G.; Gaspar, A. B.; Appel, L. G. Appl. Catal.
A Gen. 2013, 458, 111–118.
309. Silva-Calpa, L. D. R.; Zonetti, P. C.; de Oliveira, D. C.; de Avillez, R. R.;
Appel, L. G. Catal. Today 2017, 289, 264–272.
310. de Lima, A. F. F.; Zonetti, P. C.; Rodrigues, C. P.; Appel, L. G. Catal. Today 2017,
279, 252–259.
311. White, W. C. Chem. Biol. Interact. 2007, 166 (1–3), 10–14.
312. Patel, A. D.; Meesters, K.; den Uil, H.; de Jong, E.; Blok, K.; Patel, M. K. Energy. Envi-
ron. Sci. 2012, 5 (9), 8430–8444.
313. Cespi, D.; Passarini, F.; Vassura, I.; Cavani, F. Green Chem. 2016, 18 (6), 1625–1638.
314. Quattlebaum, W. M.; Toussaint, W. J.; Dunn, J. T. J. Am. Chem. Soc. 1947, 69 (3),
593–599.
315. Taifan, W. E.; Bucko, T.; Baltrusaitis, J. J. Catal. 2017, 346, 78–91.
316. Chieregato, A.; Ochoa, J. V.; Bandinelli, C.; Fornasari, G.; Cavani, F.; Mella, M.
ChemSusChem 2015, 8 (2), 377–388.
317. Zhang, M.; Gao, M.; Chen, J.; Yu, Y. RSC Adv. 2015, 5 (33), 25959–25966.
318. Kvisle, S.; Aguero, A.; Sneeden, R. P. A. Appl. Catal. 1988, 43 (1), 117–131.
319. Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A. Catal. Sci.
Technol. 2015, 5 (5), 2869–2879.
320. Ohnishi, R.; Akimoto, T.; Tanabe, K. J. Chem. Soc. Chem. Commun. 1985, 22,
1613–1614.
321. Kitayama, Y.; Satoh, M.; Kodama, T. Catal. Lett. 1996, 36 (1), 95–97.
322. Larina, O. V.; Kyriienko, P. I.; Soloviev, S. O. Catal. Lett. 2015, 145 (5), 1162–1168.
323. Sekiguchi, Y.; Akiyama, S.; Urakawa, W.; Koyama, T.-R.; Miyaji, A.; Motokura, K.;
Baba, T. Catal. Commun. 2015, 68, 20–24.
324. Lewandowski, M.; Babu, G. S.; Vezzoli, M.; Jones, M. D.; Owen, R. E.; Mattia, D.;
Plucinski, P.; Mikolajska, E.; Ochenduszko, A.; Apperley, D. C. Catal. Commun. 2014,
49, 25–28.
325. Da Ros, S.; Jones, M. D.; Mattia, D.; Pinto, J. C.; Schwaab, M.; Noronha, F. B.;
Kondrat, S. A.; Clarke, T. C.; Taylor, S. H. ChemCatChem 2016, 8 (14), 2376–2386.
326. Baylon, R. A. L.; Sun, J.; Wang, Y. Catal. Today 2016, 259 (Part 2), 446–452.
327. Larina, O. V.; Kyriienko, P. I.; Soloviev, S. O. Theor. Exp. Chem. 2016, 52 (1), 51–56.
328. Dai, W.; Zhang, S.; Yu, Z.; Yan, T.; Wu, G.; Guan, N.; Li, L. ACS Catal. 2017, 7 (5),
3703–3706.
329. Natta, G.; Rigamonti, R. Chim. Ind. 1947, 29, 239–243.
330. Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.
ChemSusChem 2014, 7 (9), 2505–2515.
331. Chung, S. H.; Angelici, C.; Hinterding, S. O. M.; Weingarth, M.; Baldus, M.;
Houben, K.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS Catal. 2016, 6 (6),
4034–4045.
332. Makshina, E. V.; Janssens, W.; Sels, B. F.; Jacobs, P. A. Catal. Today 2012, 198 (1),
338–344.
333. Ochoa, J. V.; Bandinelli, C.; Vozniuk, O.; Chieregato, A.; Malmusi, A.; Recchi, C.;
Cavani, F. Green Chem. 2016, 18 (6), 1653–1663.
334. Gao, M.; Zhang, M.; Yu, Y. Catal. Lett. 2016, 146 (12), 2450–2457.
335. Yan, T.; Yang, L.; Dai, W.; Wang, C.; Wu, G.; Guan, N.; Hunger, M.; Li, L. J. Catal.
2018, 367, 7–15.
336. Wang, C.; Zheng, M.; Li, X.; Li, X.; Zhang, T. Green Chem. 2019, 21 (5), 1006–1010.
337. Janssens, W.; Makshina, E. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;
Kerkhofs, S.; Martens, J. A.; Jacobs, P. A.; Sels, B. F. ChemSusChem 2015, 8 (6),
994–1008.
Synthesis of ethanol and its catalytic conversion 185
338. Sushkevich, V. L.; Ivanova, I. I.; Ordomsky, V. V.; Taarning, E. ChemSusChem 2014, 7
(9), 2527–2536.
339. Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. Green Chem. 2015, 17 (4), 2552–2559.
340. Sushkevich, V. L.; Ivanova, I. I. ChemSusChem 2016, 9 (16), 2216–2225.
341. Kyriienko, P. I.; Larina, O. V.; Soloviev, S. O.; Orlyk, S. M.; Calers, C.; Dzwigaj, S.
ACS Sust. Chem. Eng. 2017, 5 (3), 2075–2083.
342. Tripathi, A.; Faungnawakij, K.; Laobuthee, A.; Assabumrungrat, S.; Laosiripojna, N.
Int. J. Chem. React. Eng. 2016, 14 (5), 945–954.
343. Jones, M. D.; Keir, C. G.; Iulio, C. D.; Robertson, R. A. M.; Williams, C. V.;
Apperley, D. C. Catal. Sci. Technol. 2011, 1 (2), 267–272.
344. De Baerdemaeker, T.; Feyen, M.; M€ uller, U.; Yilmaz, B.; Xiao, F. S.; Zhang, W.;
Yokoi, T.; Bao, X.; Gies, H.; De Vos, D. E. ACS Catal. 2015, 5 (6), 3393–3397.
345. Shylesh, S.; Gokhale, A. A.; Scown, C. D.; Kim, D.; Ho, C. R.; Bell, A. T.
ChemSusChem 2016, 9 (12), 1462–1472.
346. Dagle, V. L.; Flake, M. D.; Lemmon, T. L.; Lopez, J. S.; Kovarik, L.; Dagle, R. A.
Appl. Catal. B Environ. 2018, 236, 576–587.
347. Sushkevich, V. L.; Palagin, D.; Ivanova, I. I. ACS Catal. 2015, 5 (8), 4833–4836.
348. Palagin, D.; Sushkevich, V. L.; Ivanova, I. I. J. Phys. Chem. C 2016, 120 (41),
23566–23575.
349. Zhang, M.; Zhuang, J.; Yu, Y. Appl. Surf. Sci. 2018, 458, 1026–1034.
350. Angelici, C.; Meirer, F.; van der Eerden, A. M. J.; Schaink, H. L.; Goryachev, A.;
Hofmann, J. P.; Hensen, E. J. M.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS
Catal. 2015, 5 (10), 6005–6015.
351. León, M.; Dı́az, E.; Vega, A.; Ordóñez, S.; Auroux, A. Appl. Catal. B Environ. 2011,
102 (3–4), 590–599.
352. Chae, H. J.; Kim, T. W.; Moon, Y. K.; Kim, H. K.; Jeong, K. E.; Kim, C. U.;
Jeong, S. Y. Appl. Catal. B Environ. 2014, 150–151, 596–604.
353. Kyriienko, P. I.; Larina, O. V.; Soloviev, S. O.; Orlyk, S. M.; Dzwigaj, S. Catal.
Commun. 2016, 77, 123–126.
354. Han, Z.; Li, X.; Zhang, M.; Liu, Z.; Gao, M. RSC Adv. 2015, 5 (126),
103982–103988.
355. M€uller, P.; Burt, S. P.; Love, A. M.; McDermott, W. P.; Wolf, P.; Hermans, I. ACS
Catal. 2016, 6 (10), 6823–6832.
356. Klein, A.; Keisers, K.; Palkovits, R. Appl. Catal. A Gen. 2016, 514, 192–202.
357. Cheong, J. L.; Shao, Y.; Tan, S. J. R.; Li, X.; Zhang, Y.; Lee, S. S. ACS Sust. Chem.
Eng. 2016, 4 (9), 4887–4894.
358. Hutchings, G. J.; Johnston, P.; Lee, D. F.; Williams, C. D. Catal. Lett. 1993, 21 (1),
49–53.
359. Tago, T.; Konno, H.; Ikeda, S.; Yamazaki, S.; Ninomiya, W.; Nakasaka, Y.;
Masuda, T. Catal. Today 2011, 164 (1), 158–162.
360. Herrmann, S.; Iglesia, E. J. Catal. 2018, 360, 66–80.
361. Liu, C.; Sun, J.; Smith, C.; Wang, Y. Appl. Catal. A Gen. 2013, 467, 91–97.
362. Sun, J.; Baylon, R. A. L.; Liu, C.; Mei, D.; Martin, K. J.; Venkitasubramanian, P.;
Wang, Y. J. Am. Chem. Soc. 2016, 138 (2), 507–517.
363. Navarro, R. M.; Peña, M. A.; Fierro, J. L. G. Chem. Rev. 2007, 107 (10), 3952–3991.
364. Hohn, K. L.; Lin, Y. C. ChemSusChem 2009, 2 (10), 927–940.
365. Deluga, G. A.; Salge, J. R.; Schmidt, L. D.; Verykios, X. E. Science 2004, 303 (5660),
993–997.
366. Nabgan, W.; Tuan Abdullah, T. A.; Mat, R.; Nabgan, B.; Gambo, Y.; Ibrahim, M.;
Ahmad, A.; Jalil, A. A.; Triwahyono, S.; Saeh, I. Renew. Sust. Energ. Rev. 2017, 79,
347–357.
367. Hou, T.; Zhang, S.; Chen, Y.; Wang, D.; Cai, W. Renew. Sust. Energ. Rev. 2015, 44,
132–148.
186 Jifeng Pang et al.
368. Ni, M.; Leung, D. Y. C.; Leung, M. K. H. Int. J. Hydrog. Energy 2007, 32 (15),
3238–3247.
369. Vaidya, P. D.; Rodrigues, A. E. Chem. Eng. J. 2006, 117 (1), 39–49.
370. Nanda, S.; Rana, R.; Zheng, Y.; Kozinski, J. A.; Dalai, A. K. Sustain. Energ. Fuels 2017,
1 (6), 1232–1245.
371. Zanchet, D.; Santos, J. B. O.; Damyanova, S.; Gallo, J. M. R.; Bueno, J. M. C. ACS
Catal. 2015, 5 (6), 3841–3863.
372. Li, S.; Gong, J. Chem. Soc. Rev. 2014, 43 (21), 7245–7256.
373. Liguras, D. K.; Kondarides, D. I.; Verykios, X. E. Appl. Catal. B Environ. 2003, 43 (4),
345–354.
374. Campos-Skrobot, F. C.; Rizzo-Domingues, R. C. P.; Fernandes-Machado, N. R. C.;
Cantão, M. P. J. Power Sources 2008, 183 (2), 713–716.
375. Coronel, L.; Múnera, J. F.; Tarditi, A. M.; Moreno, M. S.; Cornaglia, L. M. Appl.
Catal. B Environ. 2014, 160–161, 254–266.
376. González-Gil, R.; Herrera, C.; Larrubia, M. A.; Mariño, F.; Laborde, M.;
Alemany, L. J. Int. J. Hydrog. Energy 2016, 41 (38), 16786–16796.
377. Bilal, M.; Jackson, S. D. Catal. Sci. Technol. 2012, 2 (10), 2043–2051.
378. Carbajal-Ramos, I. A.; Gomez, M. F.; Condó, A. M.; Bengió, S.;
Andrade-Gamboa, J. J.; Abello, M. C.; Gennari, F. C. Appl. Catal. B Environ. 2016,
181, 58–70.
379. Mudiyanselage, K.; Al-Shankiti, I.; Foulis, A.; Llorca, J.; Idriss, H. Appl. Catal.
B Environ. 2016, 197, 198–205.
380. Choong, C.; Zhong, Z.; Huang, L.; Borgna, A.; Hong, L.; Chen, L.; Lin, J. ACS Catal.
2014, 4 (7), 2359–2363.
381. Wang, F.; Cai, W.; Descorme, C.; Provendier, H.; Shen, W.; Mirodatos, C.;
Schuurman, Y. Int. J. Hydrog. Energy 2014, 39 (31), 18005–18015.
382. Dai, R.; Zheng, Z.; Sun, C.; Li, X.; Wang, S.; Wu, X.; An, X.; Xie, X. Fuel 2018, 214,
88–97.
383. Zhang, C.; Zhang, P.; Li, S.; Wu, G.; Ma, X.; Gong, J. Phys. Chem. Chem. Phys. 2012,
14 (10), 3295–3298.
384. Wang, T.; Ma, H.; Zeng, L.; Li, D.; Tian, H.; Xiao, S.; Gong, J. Nanoscale 2016, 8 (19),
10177–10187.
385. Montero, C.; Ochoa, A.; Castaño, P.; Bilbao, J.; Gayubo, A. G. J. Catal. 2015, 331,
181–192.
386. Sohn, H.; Ozkan, U. S. Energy Fuels 2016, 30 (7), 5309–5322.
387. da Silva, A. L. M.; den Breejen, J. P.; Mattos, L. V.; Bitter, J. H.; de Jong, K. P.;
Noronha, F. B. J. Catal. 2014, 318, 67–74.
388. Barroso, M. N.; Gomez, M. F.; Arrúa, L. A.; Abello, M. C. Int. J. Hydrog. Energy 2014,
39 (16), 8712–8719.
389. G€und€ uz, S.; Dogu, T. Appl. Catal. B Environ. 2015, 168–169, 497–508.
390. Sun, C.; Li, H.; Chen, L. Energy. Environ. Sci. 2012, 5 (9), 8475–8505.
391. Turczyniak, S.; Teschner, D.; Machocki, A.; Zafeiratos, S. J. Catal. 2016, 340,
321–330.
392. Sun, J.; Karim, A. M.; Mei, D.; Engelhard, M.; Bao, X.; Wang, Y. Appl. Catal.
B Environ. 2015, 162, 141–148.
393. Li, M.-R.; Song, Y. Y.; Wang, G. C. ACS Catal. 2019, 9 (3), 2355–2367.
394. Szijjártó, G. P.; Pászti, Z.; Sajó, I.; Erdőhelyi, A.; Radnóczi, G.; Tompos, A. J. Catal.
2013, 305, 290–306.
395. Moretti, E.; Storaro, L.; Talon, A.; Chitsazan, S.; Garbarino, G.; Busca, G.;
Finocchio, E. Fuel 2015, 153, 166–175.
396. González-Gil, R.; Chamorro-Burgos, I.; Herrera, C.; Larrubia, M. A.; Laborde, M.;
Mariño, F.; Alemany, L. J. Int. J. Hydrog. Energy 2015, 40 (34), 11217–11227.
Synthesis of ethanol and its catalytic conversion 187
397. Bussi, J.; Musso, M.; Quevedo, A.; Faccio, R.; Romero, M. Catal. Today 2017, 296,
154–162.
398. Espinal, R.; Taboada, E.; Molins, E.; Chimentao, R. J.; Medina, F.; Llorca, J. RSC
Adv. 2012, 2 (7), 2946–2956.
399. Pan, X.; Bao, X. Acc. Chem. Res. 2011, 44 (8), 553–562.
400. Huck-Iriart, C.; Soler, L.; Casanovas, A.; Marini, C.; Prat, J.; Llorca, J.; Escudero, C.
ACS Catal. 2018, 8 (10), 9625–9636.
401. Iulianelli, A.; Basile, A. Catal. Sci. Technol. 2011, 1 (3), 366–379.
402. Murmura, M. A.; Patrascu, M.; Annesini, M. C.; Palma, V.; Ruocco, C.;
Sheintuch, M. Int. J. Hydrog. Energy 2015, 40 (17), 5837–5848.
403. Hedayati, A.; Le Corre, O.; Lacarrière, B.; Llorca, J. Catal. Today 2016, 268, 68–78.
404. Ma, R.; Castro-Dominguez, B.; Mardilovich, I. P.; Dixon, A. G.; Ma, Y. H. Chem.
Eng. J. 2016, 303, 302–313.
405. Fan, L. S.; Zeng, L.; Luo, S. W. AIChE J. 2015, 61 (1), 2–22.
406. Zhao, X.; Zhou, H.; Sikarwar, V. S.; Zhao, M.; Park, A. H. A.; Fennell, P. S.;
Shen, L. H.; Fan, L. S. Energy Environ. Sci. 2017, 10 (9), 1885–1910.
407. Luo, M.; Yi, Y.; Wang, S.; Wang, Z.; Du, M.; Pan, J.; Wang, Q. Renew. Sust. Energ.
Rev. 2018, 81, 3186–3214.
408. Wang, K.; Dou, B.; Jiang, B.; Song, Y.; Zhang, C.; Zhang, Q.; Chen, H.; Xu, Y. Int. J.
Hydrog. Energy 2016, 41 (30), 12899–12909.
409. Jiang, B.; Dou, B.; Wang, K.; Zhang, C.; Song, Y.; Chen, H.; Xu, Y. Chem. Eng. J.
2016, 298, 96–106.
410. Jin, C.; Yao, M.; Liu, H.; Lee, C. F.; Ji, J. Renew. Sust. Energ. Rev. 2011, 15 (8),
4080–4106.
411. Rakopoulos, D. C.; Rakopoulos, C. D.; Giakoumis, E. G.; Dimaratos, A. M.;
Kyritsis, D. C. Energy Convers. Manag. 2010, 51 (10), 1989–1997.
412. Elfasakhany, A. Energy Convers. Manage 2015, 95, 398–405.
413. Elfasakhany, A. Fuel 2016, 163, 166–174.
414. Uyttebroek, M.; Van Hecke, W.; Vanbroekhoven, K. Catal. Today 2015, 239, 7–10.
415. Puligundla, P.; Smogrovicova, D.; Obulam, V. S. R.; Ko, S. J. Ind. Microbiol. Biotechnol.
2011, 38 (9), 1133–1144.
416. Koda, K.; Matsu-ura, T.; Obora, Y.; Ishii, Y. Chem. Lett. 2009, 38 (8), 838–839.
417. Dowson, G. R. M.; Haddow, M. F.; Lee, J.; Wingad, R. L.; Wass, D. F. Angew. Chem.
Int. Ed. 2013, 52 (34), 9005–9008.
418. Wingad, R. L.; Gates, P. J.; Street, S. T. G.; Wass, D. F. ACS Catal. 2015, 5 (10),
5822–5826.
419. Xu, G. Q.; Lammens, T.; Liu, Q.; Wang, X. C.; Dong, L. L.; Caiazzo, A.; Ashraf, N.;
Guan, J.; Mu, X. D. Green Chem. 2014, 16 (8), 3971–3977.
420. Chakraborty, S.; Piszel, P. E.; Hayes, C. E.; Baker, R. T.; Jones, W. D. J. Am. Chem.
Soc. 2015, 137 (45), 14264–14267.
421. Fu, S.; Shao, Z.; Wang, Y.; Liu, Q. J. Am. Chem. Soc. 2017, 139 (34), 11941–11948.
422. Wingad, R. L.; Bergstroem, E. J. E.; Everett, M.; Pellow, K. J.; Wass, D. F. Chem.
Commun. 2016, 52 (29), 5202–5204.
423. Liu, Q.; Xu, G.; Wang, X.; Mu, X. Green Chem. 2016, 18 (9), 2811–2818.
424. Hattori, H. Chem. Rev. 1995, 95 (3), 537–558.
425. Ndou, A. S.; Plint, N.; Coville, N. J. Appl. Catal. A Gen. 2003, 251 (2), 337–345.
426. Kozlowski, J. T.; Davis, R. J. J. Energy. Chem. 2013, 22 (1), 58–64.
427. Birky, T. W.; Kozlowski, J. T.; Davis, R. J. J. Catal. 2013, 298, 130–137.
428. Cavani, F.; Trifiro, F.; Vaccari, A. Catal. Today 1991, 11 (2), 173–301.
429. Di Cosimo, J. I.; Apesteguı́a, C. R.; Gines, A. M. J. L.; Iglesia, E. J. Catal. 2000, 190
(2), 261–275.
430. León, M.; Dı́az, E.; Ordóñez, S. Catal. Today 2011, 164 (1), 436–442.
188 Jifeng Pang et al.
431. Ordóñez, S.; Dı́az, E.; León, M.; Faba, L. Catal. Today 2011, 167 (1), 71–76.
432. Carvalho, D. L.; de Avillez, R. R.; Rodrigues, M. T.; Borges, L. E. P.; Appel, L. G.
Appl. Catal. A Gen. 2012, 415–416, 96–100.
433. Rodrigues, E. G.; Keller, T. C.; Mitchell, S.; Perez-Ramirez, J. Green Chem. 2014, 16
(12), 4870–4874.
434. Diallo-Garcia, S.; Laurencin, D.; Krafft, J.-M.; Casale, S.; Smith, M. E.; Lauron-
Pernot, H.; Costentin, G. J. Phys. Chem. C 2011, 115 (49), 24317–24327.
435. Sun, H.; Su, F. Z.; Ni, J.; Cao, Y.; He, H.-Y.; Fan, K. N. Angew. Chem. Int. Ed. 2009,
48 (24), 4390–4393.
436. Ghantani, V. C.; Lomate, S. T.; Dongare, M. K.; Umbarkar, S. B. Green Chem. 2013,
15 (5), 1211–1217.
437. Tsuchida, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Ind. Eng. Chem. Res. 2006, 45
(25), 8634–8642.
438. Tsuchida, T.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Ind. Eng. Chem. Res.
2008, 47 (5), 1443–1452.
439. Tsuchida, T.; Kubo, J.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. J. Catal.
2008, 259 (2), 183–189.
440. Ogo, S.; Onda, A.; Yanagisawa, K. Appl. Catal. A Gen. 2011, 402 (1–2), 188–195.
441. Ogo, S.; Onda, A.; Iwasa, Y.; Hara, K.; Fukuoka, A.; Yanagisawa, K. J. Catal. 2012,
296, 24–30.
442. Silvester, L.; Lamonier, J. F.; Faye, J.; Capron, M.; Vannier, R. N.; Lamonier, C.;
Dubois, J. L.; Couturier, J. L.; Calais, C.; Dumeignil, F. Catal. Sci. Technol. 2015, 5
(5), 2994–3006.
443. Scalbert, J.; Thibault-Starzyk, F.; Jacquot, R.; Morvan, D.; Meunier, F. J. Catal. 2014,
311, 28–32.
444. Meunier, F. C.; Scalbert, J.; Thibault-Starzyk, F. C. R. Chim. 2015, 18 (3), 345–350.
445. Yang, K.-W.; Jiang, X. Z.; Zhang, W.-C. Chin. Chem. Lett. 2004, 15, 1497–1500.
446. Riittonen, T.; Toukoniitty, E.; Madnani, D. K.; Leino, A. R.; Kordas, K.; Szabo, M.;
Sapi, A.; Arve, K.; Warna, J.; Mikkola, J. P. Catalysts 2012, 2 (1), 68–84.
447. Riittonen, T.; Eranen, K.; Maki-Arvela, P.; Shchukarev, A.; Rautio, A.-R.;
Kordas, K.; Kumar, N.; Salmi, T.; Mikkola, J. P. Renew. Energy 2015, 74, 369–378.
448. Ghaziaskar, H. S.; Xu, C. RSC Adv. 2013, 3 (13), 4271–4280.
449. Jiang, D.; Wu, X.; Mao, J.; Ni, J.; Li, X. Chem. Commun. 2016, 52 (95), 13749–13752.
450. Marcu, I. C.; Tichit, D.; Fajula, F.; Tanchoux, N. Catal. Today 2009, 147 (3–4),
231–238.
451. Pang, J.; Zheng, M.; He, L.; Li, L.; Pan, X.; Wang, A.; Wang, X.; Zhang, T. J. Catal.
2016, 344, 184–193.
452. Sun, Z.; Vasconcelos, A. C.; Bottari, G.; Stuart, M. C. A.; Bonura, G.; Cannilla, C.;
Frusteri, F.; Barta, K. ACS Sust. Chem. Eng. 2017, 5 (2), 1738–1746.
453. Marcu, I. C.; Tanchoux, N.; Fajula, F.; Tichit, D. Catal. Lett. 2013, 143 (1), 23–30.
454. Zhang, X.; Liu, Z.; Xu, X.; Yue, H.; Tian, G.; Feng, S. ACS Sust. Chem. Eng. 2013, 1
(12), 1493–1497.
455. Dziugan, P.; Jastrzabek, K. G.; Binczarski, M.; Karski, S.; Witonska, I. A.;
Kolesinska, B.; Kaminski, Z. J. Fuel 2015, 158, 81–90.
456. Earley, J. H.; Bourne, R. A.; Watson, M. J.; Poliakoff, M. Green Chem. 2015, 17 (5),
3018–3025.
457. Jiang, D.; Fang, G.; Tong, Y.; Wu, X.; Wang, Y.; Hong, D.; Leng, W.; Liang, Z.;
Tu, P.; Liu, L.; Xu, K.; Ni, J.; Li, X. ACS Catal. 2018, 8 (12), 11973–11978.
458. Pereira, L. G.; Dias, M. O. S.; Junqueira, T. L.; Pavanello, L. G.; Chagas, M. F.;
Cavalett, O.; Maciel Filho, R.; Bonomi, A. Chem. Eng. Res. Des. 2014, 92 (8),
1452–1462.
Synthesis of ethanol and its catalytic conversion 189
459. Dias, M. O. S.; Pereira, L. G.; Junqueira, T. L.; Pavanello, L. G.; Chagas, M. F.;
Cavalett, O.; Maciel Filho, R.; Bonomi, A. Chem. Eng. Res. Des. 2014, 92 (8),
1441–1451.
460. Jordison, T. L.; Lira, C. T.; Miller, D. J. Ind. Eng. Chem. Res. 2015, 54 (44),
10991–11000.
461. Riittonen, T.; Salmi, T.; Mikkola, J. P.; Warna, J. Top. Catal. 2014, 57 (17–20),
1425–1429.
462. Derouane, E. G.; Nagy, J. B.; Dejaifve, P.; van Hooff, J. H. C.; Spekman, B. P.;
Vedrine, J. C.; Naccache, C. J. Catal. 1978, 53 (1), 40–55.
463. Galadima, A.; Muraza, O. J. Ind. Eng. Chem. 2015, 31, 1–14.
464. Johansson, R.; Hruby, S. L.; Rass-Hansen, J.; Christensen, C. H. Catal. Lett. 2008, 127
(1), 1.
465. Ramasamy, K. K.; Wang, Y. Catal. Today 2014, 237, 89–99.
466. Gayubo, A. G.; Alonso, A.; Valle, B.; Aguayo, A. T.; Bilbao, J. AIChE J. 2012, 58 (2),
526–537.
467. Van der Borght, K.; Galvita, V. V.; Marin, G. B. Appl. Catal. A Gen. 2015, 492,
117–126.
468. Narula, C. K.; Li, Z.; Casbeer, E. M.; Geiger, R. A.; Moses-Debusk, M.; Keller, M.;
Buchanan, M. V.; Davison, B. H. Sci. Rep. 2015, 5, 16039.
469. Hamieh, S.; Canaff, C.; Tayeb, K. B.; Tarighi, M.; Maury, S.; Vezin, H.; Pouilloux, Y.;
Pinard, L. Eur. Phys. J. Spec. Top. 2015, 224 (9), 1817–1830.
470. Kittikarnchanaporn, J.; Jitkarnka, S. Clean Technol. Environ. Policy 2015, 17 (5),
1127–1137.
471. Chistyakov, A. V.; Gubanov, M. A.; Murzin, V. Y.; Zharova, P. A.; Tsodikov, M. V.;
Kriventsov, V. V.; Gekhman, A. E.; Moiseevd, I. I. Russ. Chem. Bull. 2014, 63, 88–93.
472. Ramasamy, K. K.; Gerber, M. A.; Flake, M.; Zhang, H.; Wang, Y. Green Chem. 2014,
16 (2), 748–760.
473. D€aumer, D.; R€auchle, K.; Reschetilowski, W. ChemCatChem 2012, 4 (6), 802–814.
474. Pinard, L.; Tayeb, K. B.; Hamieh, S.; Vezin, H.; Canaff, C.; Maury, S.; Delpoux, O.;
Pouilloux, Y. Catal. Today 2013, 218, 57–64.
475. Van der Borght, K.; Batchu, R.; Galvita, V. V.; Alexopoulos, K.; Reyniers, M.-F.;
Thybaut, J. W.; Marin, G. B. Angew. Chem. Int. Ed. 2016, 55 (41), 12817–12821.
476. Chowdhury, A. D.; Lucini Paioni, A.; Whiting, G. T.; Fu, D.; Baldus, M.;
Weckhuysen, B. M. Angew. Chem. Int. Ed. 2019, 58 (12), 3908–3912.
477. Bai, S.; Shao, Q.; Wang, P.; Dai, Q.; Wang, X.; Huang, X. J. Am. Chem. Soc. 2017,
139 (20), 6827–6830.
478. Wang, L.; Wang, L.; Zhang, J.; Liu, X.; Wang, H.; Zhang, W.; Yang, Q.; Ma, J.;
Dong, X.; Yoo, S. J.; Kim, J.-G.; Meng, X.; Xiao, F.-S. Angew. Chem. Int. Ed.
2018, 57 (21), 6104–6108.
479. Singh, A.; Rangaiah, G. P. Ind. Eng. Chem. Res. 2017, 56 (18), 5147–5163.
480. Pang, J.; Sun, J.; Zheng, M.; Li, H.; Wang, Y.; Zhang, T. Appl. Catal. B Environ. 2019,
254, 510–522.
481. Dou, J.; Sun, Z.; Opalade, A. A.; Wang, N.; Fu, W.; Tao, F. Chem. Soc. Rev. 2017, 46
(7), 2001–2027.
190 Jifeng Pang et al.