0% found this document useful (0 votes)
41 views103 pages

Ethanol

Uploaded by

Dana Matei
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
41 views103 pages

Ethanol

Uploaded by

Dana Matei
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

CHAPTER TWO

Synthesis of ethanol and its


catalytic conversion
Jifeng Panga, Mingyuan Zhenga,*, Tao Zhanga,b,*
a
CAS Key Laboratory of Science and Technology on Applied Catalysis, Dalian Institute of Chemical Physics,
Chinese Academy of Sciences, Dalian, Liaoning, China
b
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian,
Liaoning, China
*Corresponding authors: e-mail address: myzheng@[Link]; taozhang@[Link]

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,

Advances in Catalysis, Volume 64 # 2019 Elsevier Inc. 89


ISSN 0360-0564 All rights reserved.
[Link]
90 Jifeng Pang et al.

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. 1 Synthetic methodologies of ethanol and its applications.


94 Jifeng Pang et al.

Bshish et al. summarized the catalysts development for ethanol reforming


(23). Makshina et al. provided a comprehensive summary of the advances
and achievements in chemocatalytic conversion of ethanol to 1,3-BD (24).
Sun and Wang discussed in detail the recent advances in catalytic conversion
of ethanol to a wide range of chemicals such as hydrogen, hydrocarbons,
aromatics, and other oxygenates (8). Recently, Eagan et al. introduced
chemical ways involved in the conversion of ethanol into middle-distillate
fuels (25). These attempts stimulated the ethanol research in large and pro-
vided guidance for its production and utilization. Taking into account of
the great achievements about ethanol production and applications, this
chapter will focus on the research developments in the production of eth-
anol and its transformation to commodity chemicals and advanced fuels.
Different methods for ethanol production will be discussed and compared
based on environmental and economic aspects. Afterwards, its utilization
in producing important chemicals including acetaldehyde, ethyl acetate,
propylene, acetone, 1,3-BD and isobutene, using homogeneous and het-
erogeneous catalysts, will be highlighted. Additionally, the catalytic con-
version of ethanol to advanced fuels such as hydrogen, n-butanol and
hydrocarbons will be summarized. Finally, an outlook about ethanol will
be provided from the perspective of fundamental researches and industrial
applications.

2. Methods for ethanol production


2.1 Production of ethanol via biomass fermentation
At present, ethanol is primarily produced by the fermentation process,
which accounts for more than 80% of global ethanol production, and sub-
stitutes ca. 1% of gasoline used in the world.
Conventional carbohydrates such as sucrose or glucose derived from sug-
arcane, sugar beets or starchy materials, are the first industrialized feedstocks.
Through the fermentation process, the first generation ethanol is produced
(26). A typical fermentation process for ethanol production from corn is
depicted in Fig. 2. The corn grain is first ground into flour, and then mixed
with water to form a mash. After a cooking process to reduce bacteria levels,
yeast is added to initiate the fermentation process. The process generally
takes about 40–50 h at around 308–313 K with a pH of 4.0–5.0 and a sugar
concentration of 30–200 g L1 (27,28). After the fermentation process, the
resulted slurry is transferred to the distillation columns for separation. In the
final stage, the ethanol distillate is dehydrated over molecular sieves or via
Synthesis of ethanol and its catalytic conversion 95

Fig. 2 Process for ethanol production from biomass via the fermentation method.

the azeotropic distillation to obtain anhydrous ethanol. Several factors such


as reaction temperature, pH, fermentation time, agitation rate, and initial
sugar concentration, affect the fermentation process as well as the ethanol
yield. Taking into account of most parameters, the productivity of the fer-
mentation process over commercially applied micro-organisms is typically
between 1 and 10 g L1 h1 (29,30).
Fermentation of biomass to ethanol is a bio catalytic process, typically
occurring in the Saccharomyces cerevisiae. The main metabolic pathway
involved in this process is fermentative glycolysis (31–33). As shown in
Fig. 3, 1 mole of glucose is metabolized to 2 moles of pyruvate via a series
of reactions. The formed pyruvate is then reduced to ethanol with the
release of CO2 under anaerobic conditions. Through these cascade reac-
tions, 2 moles of ethanol are produced from 1 mole of hexose with 2 moles
of CO2 being released. Theoretically, the maximum weight yield of ethanol
in this pathway is 51.1%. During this pathway, ATP is transformed to ADP,
and then regenerated to maintain a constant pH and phosphate con-
centration, which is the sole pathway in yeast that provides energy for
the yeast cellular maintenance (15,34). However, yeast biomass and various
by-products such as glycerol are produced if sufficient ATP is available for
cell growth other than cellular maintenance (35). These glycolytic interme-
diates inevitably decrease the ethanol yield to some extent. Several strategies
have been dedicated to minimize the side-products and increase the ethanol
96 Jifeng Pang et al.

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. 3 Metabolic pathway for ethanol production in the Saccharomyces cerevisiae


(31,32). Reprinted (adapted) with permission from Bai, F. W.; Anderson, W. A.; Moo-Young,
M. Biotechnol. Adv. 2008, 26 (1), 89–105. Copyright (2008) Elsevier.

yield such as free-energy conservation and redox-metabolism engineering


(36). Although great works have been done on the reaction mechanism
study and a high yield of ethanol production has been achieved, the com-
plexity of gene expression networks behind the fermentation process is still
far from being completely understood.
Minimizing the energy consumption is an important way to elevate the
economic viability of the fermentation process, and great efforts have been
dedicated to achieve this goal with mixed results. One approach is to
increase the tolerance of microorganisms during the fermentation process.
The toxicity of ethanol to yeast limits the ethanol concentration to be
around 10–12 wt%. Some ethanol-tolerant strains of yeast, e.g., Red Star
Pasteur Champagne, Lalvin EC-1118 yeasts or “Turbo Yeast,” can elevate
the concentration of ethanol to 20 v% or higher (22). Moreover, research in
yeast physiology has revealed that many strains of Saccharomyces cerevisiae can
potentially tolerate a much higher ethanol concentration than expected
Synthesis of ethanol and its catalytic conversion 97

(over 15% ethanol) (37). Another successful approach is the application of


membrane pervaporation technology for the separation of ethanol from fer-
mentation broths, and some pervaporation membranes have been used in
combination with conventional fermentation processes (38,39). These
advancements play a key role in increasing the industrial efficiency and
reducing the environmental impacts.
Among all fermentation processes, feedstock cost dominates the final
ethanol price, which is estimated to be around 80% of final costs of corn
ethanol. Most of high ethanol production countries have the advantage in
available cheap feedstocks. For instance, owing to the high productivity
of corn, United States has been the world’s largest ethanol producer since
2014. Brazil is another world leading ethanol producer due to its abundant
sugarcane resources (40). In contrast, the ethanol in China is produced
mainly from maize and cassava, and its absolute value is far less than that
of the United States and Brazil (41). On the other side, the feedstock of
the first generation ethanol is competing with food resources, which could
potentially drive food prices higher. Hence, the exploration of more abundant
and cheaper feedstock is crucial for the economic production of ethanol (42).
Lignocellulosic biomass is the most abundant non-edible feedstock, the
majority of which are carbohydrates of hemicellulose and cellulose com-
posed of xylose and glucose. In the past decade, fermentation of lignocellu-
losic biomass to ethanol, i.e., the second generation ethanol, has been
extensively investigated around the globe. Nevertheless, the complex struc-
ture of lignin and (hemi)cellulose evolved by nature makes the plant cell
rather sturdy. It is still a challenge to degrade biomass efficiently under mild
chemical or biochemical conditions. As shown in Fig. 2, many pretreatment
processes are employed to open the rigid structure of lignocellulosic bio-
mass. Preliminary mechanical pretreatment of crushing or milling allows
to reduce the particle size and the degree of polymerization (43). Afterwards,
acid, alkaline or steam explosion pretreatments are conducted to enhance
the sugar solubility and the lignin removal (44–46). The resulted biomass
is then hydrolyzed to mono and/or oligosaccharides using acids or enzymes,
followed by fermentation using yeast or bacterial strains.
The ethanol production from lignocellulosic biomass is highly depen-
dent on the percentage of C6 or C5 sugar recovery after pretreatments
and/or hydrolysis, and the produced inhibitors during the pretreatment
and hydrolysis processes (47). In this whole process, the conversion of cel-
lulosic components into fermentable sugars is the major technological and
economical bottleneck. Other than that, some obstacles such as the high
98 Jifeng Pang et al.

operation cost and the poison effect of inhibitors, must be overcome to


achieve a high ethanol yield. In detail, most pretreatment processes are
cost-ineffective and contribute >20% of the cost for lignocellulosic biomass
ethanol (48). Furthermore, some harmful compounds are inevitably pro-
duced during pretreatment processes, which are detrimental to the sugars
fermentation process (49,50). Meanwhile, the rate for co-fermentation of
C5 and C6 sugars is much slower than that of C6 sugars. All these features
make the cellulosic ethanol relatively expensive although the costs have been
coming down significantly over the past decades. With governments incen-
tive and carbon tax stimulations, lignocellulosic ethanol has been reported to
be produced in the United States and China (22). Even then, lignocellulosic
ethanol is still less economically feasible than the first generation ethanol. For
instance, ethanol production from lignocellulosic biomass was estimated to
be 700–1000 USD/ton in 2016 (the minimum ethanol selling price), which
was much higher than the target price of 300–500 USD/ton (51–53).

2.2 Production of ethanol from ethylene


Hydration of ethylene to ethanol is a proved industrial process, which
was developed by the Shell Chemicals in 1947 (16,54). This reaction is exo-
thermic, and hence the operating temperature is a compromise between
kinetics and thermodynamics (55). In the commercial process, ethylene is
co-fed with water at 573 K under a pressure of 6.0–7.0 MPa, about 5%
of ethylene is converted into ethanol at each pass with diethyl ether as
the main by-product.
Hydration of ethylene to ethanol is an acid catalyzed reaction. Silica gel-
supported phosphoric acid is the first catalyst used for this transformation,
over which the orthophosphoric acid is regarded as the active site. However,
the phosphoric acid is prone to leach via vaporization, causing the catalyst
deactivation and equipment corrosion issues. Although some solid acid cat-
alysts such as WO3/ZrO2 and WO3–TiO2 were developed, the elevated
ethylene price and rapid development of sugar fermentation make this route
unattractive for the large-scale production of ethanol (56,57).

2.3 Direct thermochemical conversion of syngas to ethanol


Syngas, a mixture of carbon monoxide, hydrogen and carbon dioxide in some
cases, is an important building block in the chemical industry. Nearly all
organic resources such as biomass, coal, petroleum and natural gas can be read-
ily converted to syngas by gasification and/or reforming processes (58,59).
Synthesis of ethanol and its catalytic conversion 99

In the commercial operation, carbon-rich materials are rapidly heated to over


973 K in a high-temperature combustion chamber of gasifier and partially
oxidized with controlled air flow. The crude syngas obtained is then purified
to remove contaminants and used as a benchmark material for synthesizing
methanol, oxygenates, gasoline, diesel, and other products (60).
Direct thermochemical conversion of syngas offers a promising route for
ethanol and higher alcohol synthesis. Some excellent works have been ded-
icated to this topic (61–64). Herein, we focus on the recent advances in the
catalyst developments in this area.
In general, direct conversion of syngas to ethanol occurs at a high pres-
sure of ca. 200 bar and temperatures higher than 573 K. This reaction
involves cascade steps. As shown in Fig. 4, H2 first adsorbs dissociatively
to form H atoms on the surface of catalysts. Meanwhile, CO is dissociated
to surface C* and oxygen species. The surface C* species are then hydro-
genated by surface H to generate adsorbed CHx species, followed by the
insertion of non-dissociated CO to an adsorbed acyl species. The acyl
species finally undergo hydrogenation to produce ethanol (61,65,66).

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.

The adsorption state of CO is crucial to the selectivity of ethanol. Without


the dissociation of CO, it is prone to be hydrogenated to methanol. In con-
trast, strongly dissociated CO causes the methane formation. Additionally,
C2+ oxygenates are also co-produced due to the uncontrollable carbon
chain growth. DFT calculations have also been used to predict the product
distribution over certain metal surfaces. Fig. 5 shows the catalyst perfor-
mances for CO hydrogenation as a function of the C and O adsorption
energies on transition-metal (2 1 1) surfaces. Metals like Ru, Co, Ni,
and Rh are identified as the best metals for methanation, and Cu is believed
to be the best metal catalyst for methanol synthesis. Although no good
metal catalyst is predicated yet for ethanol synthesis, the result provides
a fundamental understanding of the transition-metal catalysts behavior
for these reactions. Promising approaches such as modifying the electronic
structure of metals by alloying, using dopants to increase the CdO bond
scission, blocking the active sites for hydrocarbon production and increas-
ing the CHx–CO coupling rates have been proposed to shift the product
distribution to the higher alcohol range (67–69). Up to date, Rh, Mo and
Cu-based catalysts are demonstrated as the most promising catalysts for
ethanol production from syngas.

2.3.1 Rh-based catalysts


The initial attempts for the conversion of syngas to ethanol were conducted
by the Union Carbide Corporation in 1975. Over the optimal 5% Rh/SiO2
catalyst, the CO conversion reached 72% with 37% methane, 7.4% meth-
anol and 33% ethanol at 598 K, a GHSV of 2200 h1 and a pressure of
17.2 MPa (70). Since then, extensive works have been done on studying
the Rh catalyst as it was identified as the most potential element for conver-
ting syngas to ethanol. Still, single Rh component exhibits a poor selectivity
to ethanol and mainly leads to methanol and methane. Various strategies
have been proposed to improve the ethanol selectivity by modifying Rh
with additives and/or elevating the dispersion of Rh on some supports.
Using promoter is the first strategy to increase the ethanol selectivity at
considerable CO conversions. Various elements such as Mn, Fe, Ce, V, La,
Ag and Ti have been screened as additives to modify the Rh catalysts. Typ-
ically, Mn, Li, and Zr are in close interaction with Rh species to form a new

active site of Rh0 x Rh + y  O  Mn + (x  y, 2  n  4), which changes the
CO adsorption from linear, geminal, and bridged forms to tilted form, the
main form to produce ethanol. The new sites also promote the dissociation
Fig. 5 Rates and selectivities for the formation of methane, methanol, and ethanol as a function of carbon and oxygen adsorption energies.
The rates and selectivities are calculated using a microkinetic model. The reaction conditions are as follows: T ¼ 593 K, PCO ¼ 30 bar,
PH2 ¼ 60 bar (67). Reprinted (adapted) with permission from 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. Copyright (2015) Elsevier.
102 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

the highest ethanol selectivity (alcohols/hydrocarbons >50) due to its high


concentration of surface alkyl and CO species (81). Liu et al. loaded Rh on
Ce1–xZrxO2 by a co-precipitation method for CO conversion. At 548 K,
2.4 MPa and a W/F of 10 g h mol1, they obtained around 35.2% ethanol
selectivity at a CO conversion of 27.3% (82). Lopez et al. evaluated the
catalytic performance of the Rh/MCM-41 catalyst and compared it with
a typical Rh/SiO2 catalyst. Although the total selectivity to C2-oxygenated
compounds is similar, the Rh/MCM-41 catalyst showed higher selectivity
to ethanol than the Rh/SiO2 catalyst (77% vs 21%) (83). Chen et al. also
studied the effect of supports on the conversion of syngas to ethanol at
573 K. In contrast with different SBA-15 and silica supports, Rh-Mn/
SBA-15-HPMo catalysts showed the better selectivity to ethanol. The
enhanced redox property of the highly dispersed Rh and Mn oxides
encapsulated within the mesoporous channels should be the main reason
for the improved catalytic performance (84). Considering the highly exo-
thermic and structure sensitive nature for CO hydrogenation, Chen et al.
used β-SiC with superior thermal conductivity for Rh loading. Over
Rh-Mn-Li/SiC catalysts at 578 K, 5.0 MPa and a GHSV of 8000 h1,
the catalytic efficiency reached 97.2 g/g-Rh/h even though the selectivity
was yet to be improved (14.4%) (85). Han et al. supported Rh-Mn-Li on
SiO2-ZrO2, and achieved the highest space–time yields to C2-oxygenates
(i.e., 71.3 g/(kg h)) at a SiO2/ZrO2 of 1:3. However, the selectivity to
ethanol was only 16%, and most of the C2-oxygenates product was acetic
acid (86). Some carbon-based supports were also used for ethanol produc-
tion. Pan et al. deposited Rh-Mn nanoparticles inside or outside of CNTs
for syngas conversion. As shown in Fig. 6, the Rh-Mn particles inside the

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.

CNTs exhibited an enhanced activity with respect to those located on


the outside surface of CNTs. At 603 K and a pressure of 3 MPa, the
C2-oxygenates selectivity reached 38.1% at 18.8% CO conversion (87).
The same group compared CNTs with other carbon materials such as
carbon black, CMK-3 and AC using Rh-Mn-Li-Fe as the active compo-
nent. The CNTs supported catalysts still exhibited the highest activity,
demonstrating that the confinement of the nanochannels and graphitic
structure are crucial in promoting the reaction (88). Kim et al. prepared
Rh-Mn-Li-Fe/OMC catalysts via three different methods (polyol, organ-
ometallic and sonochemical) for alcohol synthesis. Over the sonochemical
method prepared catalysts, the selectivity to C2–4 alcohols reached up to
52.7%, significantly higher than that prepared by the incipient wetness
method (89). They also investigated the influence of Rh particle size on
ethanol production over OMC supports, and found that an average particle
size of 4.6 nm was preferred for high ethanol selectivity (36.0% at 5.5% CO
conversion) (90).

2.3.2 Non-noble metal-based catalysts


Due to the high price as well as the limited availability of noble metals,
extensive studies have been dedicated to explore inexpensive non-noble
metal catalysts for syngas-to-ethanol conversion. In general, the catalysts
can be classified into modified F-T synthesis catalysts, modified methanol
synthesis catalysts and molybdenum (Mo)-based catalysts.

[Link] Modified F-T synthesis catalysts


Although modified F-T catalysts are regarded as promising candidates for the
synthesis of C2+ alcohols, the selectivity to ethanol is still relatively low. The
formed alcohols are basically a mixture of C1–C6 linear alcohols that obeys
the Anderson–Schulz–Flory distribution. In order to improve the ethanol
selectivity, many efforts have been made in catalysts design to depress the
products of methane and methanol, and control the carbon chain growth.
Different metals have been used to modify the electronic properties of
F-T catalysts. Yang et al. loaded Pd onto Fe–Cu–Co-based catalysts, and
obtained 19.5% and 58.7% selectivity to ethanol and alcohols, respectively,
with 81.7% CO conversion at 623 K. The Pd facilitated the reduction
of Fe–Cu–Co-based catalysts, which improved the H2 activation for the
C2+ alcohol production (91). Ning et al. reported a uniformly dispersed
CoGa catalyst with CoGa particles trapped in the ZnAl-LDHs. The produc-
tion of methanol was highly suppressed due to the contiguous distribution
Synthesis of ethanol and its catalytic conversion 105

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).

[Link] Modified methanol synthesis catalysts


The second type of non-noble metal catalysts for ethanol production is
modified methanol synthesis catalysts. The research was initiated due to
the formation of ethanol during methanol synthesis over catalysts containing
a trace amount of alkali. Following this strategy, various promoters were
added to Cu/ZnO/Al2O3 (Cr2O3) catalysts, trying to shift the selectivity
toward ethanol. Although the reaction mechanism is still debating, it is
generally accepted that the CdC formation is the rate-determining step
for C2+ OH synthesis. The presence of some additives can stabilize C1 inter-
mediates (formyl species, formaldehyde, CHx), decrease the energy barrier
for coupling two C1 species to C2-oxygenates, and inhibit the side reactions
(methanol and methane formation) (97–99).
106 Jifeng Pang et al.

The primary elements used to modify the Cu/ZnO/Al2O3 catalyst were


alkali metal ions, which followed the trend of Cs > Rb > K > Na > Li in the
aspect of C2+ OH selectivity and productivity (100). The alkali metal ions
were found to neutralize the surface acidity of catalysts and/or supports,
and facilitate the adsorption of CO molecules on the catalyst surface. The
alkali metal ions also provided base sites for the condensation of C1 species
and thus increased higher alcohols selectivity (101,102). Sun et al. used both
experimental and DFT calculation to investigate the promotion effect of Cs.
Results found that C2+ OH productivity was doubled at 583 K with the
assistance of Cs2O promoters, which should be attributed to the promotion
effect of the initial CdC bond formation. The result was confirmed by DFT
calculations, which disclosed that the stability of key intermediates such as
HCO and H2CO were enhanced in the presence of Cs additives (103).
Other than alkali metals, transition metals have also been used as pro-
moters for boosting ethanol production and/or suppressing CH3OH and
CH4 formation (104). Beiramar et al. studied a series of promoters, i.e.,
Fe, Co, Ru, Zr, Mo, Mg, Mn, and Cr, on Cu/ZnO/Al2O3 catalysts for
converting syngas to ethanol. The reducibility of Cu in Cu/ZnO/Al2O3
catalysts was greatly affected by the promoters, which was easily reduced
in the presence of Mn and Cr species, leading to more active sites for the
CO hydrogenation and the C2+ alcohols production (105). Differently,
Li et al. used B to promote Cu/ZnO/Al2O3 catalysts for ethanol synthesis.
When the B content was 6.0 wt%, the ethanol selectivity reached 37.6%
with 4.9% CO conversion at 523 K and a GHSV of 300 mL g1 h1. Char-
acterization results revealed that a suitable B promoter prevents the growth
of Cu particles and increases the ratio of Cu2O on the catalyst surface, lead-
ing to a balanced Cu+-Cu0 for CO activation and C1 species coupling (106).
The same group also investigated the effect of catalyst preparation methods
on ethanol selectivity and its deactivation mechanisms. By comparison of the
traditional (the precursor was aged, dried, and calcined) and the liquid-phase
prepared (the precursor was aged, re-dispersed, dried, and calcined) catalysts,
the authors found that catalysts prepared by liquid-phase methods showed
superior activity to ethanol. This is attributed to the introduction of Cu+
and the weak acid on the catalyst surface for carbon chain growth (107).
Additionally, the state of Al in the Cu–Zn–Al catalysts also played a key role
in the formation of high carbon alcohols. Liu et al. found that the AlOOH in
catalysts promoted the CO dissociation and chain growth. In contrast, the
Al2O3 had no function of CO dissociation, which caused the formation of
methanol. Hence, to stabilize the AlOOH in catalysts is very crucial to the
production of C2+ alcohols (108).
Synthesis of ethanol and its catalytic conversion 107

[Link] Mo-based catalysts


Mo itself has a relatively low activity for syngas conversion, and yields pri-
marily hydrocarbons. However, the selectivity dramatically shifts toward
alcohols when Mo is bonded with the C, O, S, or P atoms in the presence
of alkali or/and Group VIII metal promoters (109). The incorporation of
heteroatoms into Mo particles weakens the CO adsorption but facilitates
the CO dissociation. On the other hand, the alkali metal ions change
the surface electronic properties and provide the moderate strength of
Mo-CO adsorption for alcohols synthesis. Some 3d transition metals, espe-
cially Ni and Co are found to promote the formation of alkyl group
(CnHx), which is considered as the key intermediate in the formation of
higher alcohols. Over the K, Co modified Mo-based catalysts, the selectiv-
ity to alcohols reached as high as 80%. With the recycle of methanol and
linear alcohols, a high selectivity to ethanol can be achieved. In addition,
the modified catalysts showed less sensitive to CO2 and were less prone to
be deactivated by the coke formation. Hence, this catalysts system appears
to be one of the most promising catalysts at present for the conversion of
syngas to ethanol (110).
Regarding the effect of ligand atoms, most studies focus on MoS2-based
catalysts because MoP- and Mo2C-based catalysts show relatively high
hydrocarbons selectivity. Moreover, MoS2-based catalysts are commercial
catalysts that used for deep hydrodesulfurization and hydrodenitrogenation
of heavy oils, accordingly have great opportunities to be used in ethanol
synthesis.
The catalyst deactivation, regeneration mechanisms and some key factors
affecting the product selectivity were discussed in some excellent works
(61,111,112). For instance, Woo et al. studied the influence of alkali metal
ions on MoS2, and found that the selectivity to alcohols follows the order of
Li < Na < Cs < Rb < K, indicating moderate base additives are desirable
(113). Similar to F-T and methanol synthesis catalysts, Ni, Co, Rh and
Mn have also been used to modify the Mo-based catalysts. Among these,
the Co-promoter demonstrated higher selectivity to ethanol due to the for-
mation of “Co–Mo–S” phases (114–117). The catalyst preparation methods
are crucial to the activity of K-CoMoSx catalysts. Hensley et al. studied the
decomposition behavior of ammonium tetrathiomolybdate at different tem-
peratures for preparing K-CoMoSx catalysts, and found the activity of cat-
alysts was sensitive to the precursor decomposition temperature and the way
in which sulfur is lost in the process. Moderate temperatures like 623 K were
preferable for maintaining the sulfur and the formation of sulfide catalysts
(118). The same group investigated the stability of K-CoMoSx catalysts
108 Jifeng Pang et al.

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).

2.4 Fermentation of syngas to ethanol


Another direct route for ethanol production from syngas is fermentation of
syngas with acetogenic bacteria. In general, C1 compounds such as CO,
CO2 or mixtures thereof are converted with hydrogen via the WL path-
way. Ethanol, acetic acid, and other metabolites such as 2,3-butanediol,
butanol or butyrate can be produced. After downstream processing involv-
ing distillation and dehydration, fuel grade ethanol is obtained (122).
Versatile syngas feedstocks containing CO and variable amounts of H2
and CO2, regardless of industrial off gas or biomass gasification products,
are suitable for the syngas fermentation. This process barely requires any pre-
treatment for the CO-rich syngas feedstock and has a high flexibility in terms
of gas composition and impurity except small amounts of special compounds
such as NO and acetylene (123,124). From the feedstock compatibility
point of view, it is a promising method for the ethanol production.
During the fermentation process, the gas mixture is fed into a bioreactor
along with fresh media mostly consisting of nutrients, salts, trace metals, and
bacteria in water. Owing to the limited interaction between syngas and
bacteria, the bioconversion rate is very low, and syngas is bubbled into
the solution in cycles. Acetogens such as Clostridium ljungdahlii, Clostridium
Synthesis of ethanol and its catalytic conversion 109

thermoaceticum, Clostridium autoethanogenum, Alkalibaculum bacchi, are the


group of bacteria active for fermenting syngas into ethanol (125). Other
key parameters including nutrient levels, pH, temperatures, gas flow rates,
agitation speeds, reactor types, and pressures should be optimized in order
to achieve a high ethanol concentration. In general, this bio process is
performed under the conditions of pH 4–6, 308–315 K and 0–5 bar (126).
As mentioned, the fermentation process follows the WL pathway, also
called the reductive acetyl-CoA pathway, which was first proposed by Wood
and Ljungdahl in 1966 (127,128). As illustrated in Fig. 7, the WL pathway
involves carbonyl and methyl branches with syngas or CO2 and CO2/H2 as
feedstocks. In the methyl branch, CO is oxidized to CO2, and then reduces
to formate by FDH. The formed formate undergoes several reductive steps to
[CH3]-Co-FeS-P and serves as the precursor for methyl group introduction
into acetyl-CoA. For the carbonyl branch, CO can either be used directly, or
generated from CO2, serves as the carbonyl group for acetyl-CoA synthesis
via the CODH (130). After the acetyl-CoA is formed, acetaldehyde can
be produced directly, or through a two-step reaction via acetate. The
acetaldehyde is finally reduced to ethanol by AOR and/or ADHE (122).
More than 100 acetogenic bacteria have been isolated and tested for the
syngas conversion, and the best bacteria belong to the classes of genera
Acetobacterium and Clostridium (131). Singla et al. tried thirteen anaerobic
strains together with enrichment protocols for developing an efficient sys-
tem to produce ethanol. Over the enriched anaerobic mixture TERI

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

including membrane bioreactors, fluidized bed bioreactors, moving bed


bioreactors, bubble column bioreactors, monolith bioreactors and stirred
tank bioreactors were used. The stirred tank bioreactor seems to have
yielded the best results so far, and it has been used in most cases even at dem-
onstration scale (141–144). Some researchers used additives to improve the
mass transfer between gas and liquid. For instance, Kim and Lee used
CoFe2O4@SiO2–CH3 nanoparticles to enhance the mass transfer of syngas.
Double yield of ethanol was obtained after the addition of nanoparticles
using the Clostridium ljungdahlii (0.156 g L1 vs 0.489 g L1; 18 h) (145).

2.5 Production of ethanol from syngas via acetic acid


An alternative indirect process for producing ethanol is the selective hydroge-
nation of syngas derived acetic acid. This process involves several steps such as
conversion of syngas to methanol or dimethyl ether, carbonylation of meth-
anol or dimethyl ether to acetic acid/ester and finally hydrogenation of acetic
acid/ester to ethanol. The atomic efficiency of overall process could be as high
as 71.9%, which makes it a promising method for ethanol production. The
conversion of syngas to methanol or dimethyl ether is a well-developed pro-
cess operated at a commercial scale over Cu/Zn/Al2O3 catalysts (146). Hence,
herein only methanol or dimethyl ether carbonylation and acetic acid/ester
hydrogenation reactions in this process will be introduced.

2.5.1 Methanol carbonylation


Carbonylation of methanol to acetic acid is one of the most important reac-
tions in this process. It was first developed by BASF in 1960 over a Co and
iodide catalyst, operated at 503 K and a pressure of 6.0–8.0 MPa. A 0.1 M
of Co-based homogeneous catalyst was used, and 90% selectivity to acetic
acid was obtained based on the reactant of methanol. Afterwards, Paulik
and Roth in Monsanto discovered an Rh-iodide catalyst system. Compared
with the BASF process, it runs under milder conditions of 423–473 K and a
pressure of 3.0–6.0 MPa. Since then, the Monsanto process has become
the dominant technology for the production of acetic acid (147). Based
on the exclusive licensing rights from Monsanto, BP Chemicals in 1996
announced a new methanol carbonylation process named as Cativa™,
which was a promoted Ir/iodide catalyst with an improved stability, all-
owing operation at low water concentrations. Currently, most industrial
processes utilize Rh (the Monsanto process) and Ir (the Cativa™ process)
catalyzed homogeneous reactions, as illustrated in Fig. 8, which contribute


CH3
CH3OH CH3
I CO
I CO
Rh
Ir
I CO
I CO I
CO I
CH3 I CH3 I
– CH3 –

- 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-

Cativa process Monsanto process


Fig. 8 Proposed catalytic cycles for the methanol carbonylation over the Ir-based (Cativa process) and Rh-based (Monsanto process) catalysts
(148,149).
Synthesis of ethanol and its catalytic conversion 113

ca. 60% of acetic acid production worldwide. The market is shared by


chemical companies such as BASF, Monsanto, BP, Chiyoda, and Celanese.
The present commercialized catalysts are homogeneous complex of Co,
Rh and Ir with iodide salts as co-catalysts. A number of strategies have been
proposed to enhance the stability and activity of catalysts, which are sum-
marized in some excellent papers and reviews (149–151). Since iodide is
employed as the main co-catalyst and acid is the main product, corrosion
problems are inevitable. Additionally, the difficult in separation of the
homogeneous metal complexes from the corrosive product enlarge the
energy consumption. Therefore, it has been strongly recommended to
develop some novel heterogeneous catalytic systems for this reaction, which
will lessen the corrosion problems and ease the recovering of catalysts (152).
Directly immobilize the active metal complexes of Rh or Ir on the surface
of solid supports such as AC, clay, alumina, silica, zeolite or polymers, pro-
vide an excellent way for addressing the catalyst separation issues (153). In
1998, Chiyoda and UOP commercialized a large scale methanol carbonyl-
ation process using a Rh complex immobilized on a poly(vinylpyridine)
resin. The catalyst was proved to be rather stable, and the ionic attachment
strategy was found to be an efficient way for supporting metals (154,155).
Yashima et al. screened different supports such as zeolites, silica, alumina
and ion-exchange resins for anchoring Rh species. Among those materials,
zeolite Y supported Rh catalysts showed highest activity to methyl acetate,
and obtained 76.6% selectivity at 87.7% methanol conversion, 453 K and a
W/F of 1.25 (156). Saikia et al. loaded Rh complex onto the formic acid
activated montmorillonite, and achieved >99% acetic acid yield at >99%
CO conversion after 3 h reactions at 403 K (157).
Besides the prevailing metal of Rh, Ni and Ir complexes were also
immobilized on supports for the carbonylation of methanol to acetic acid
with CH3I as the co-catalyst. Merenov and Abraham supported 3.5%Ni on
AC at different temperatures for this reaction. The catalyst was found to be
relatively active at a reaction temperature of 498 K (25% methanol conver-
sion). Nevertheless, it deactivated irreversibly at higher temperatures
(533 K) due to the formation of inactive nickel carbide phases (152). Mod-
erate amounts of surface oxygen groups were necessary to balance the redox
cycle of nickel and the adsorption of carbonaceous species, as reported by
Pang et al. After adjusting the surface oxygen groups via HNO3 treatments,
the CO conversion lingered around 40% with ca. 80% selectivity to target
products over the Ni/AC catalyst (158). Liu and Chiu modified Ni/AC
with Sn for this reaction. The methanol conversion was significantly
114 Jifeng Pang et al.

increased (>15% conversion improvement) with similar selectivity to


methyl acetate and acetic acid, which was attributed to the increased CO
adsorption in the presence of Sn (159). Kwak et al. synthesized an Ir–La/C
catalyst, which demonstrated high activity, selectivity and stability for
the carbonylation of methanol. Both acetic acid and methyl acetate
were produced with selectivity maintained >99% at a productivity of
1.5 mol acetyl/mol Irs in the >1 month operation. The authors found
that the halocarbonyl structures of Ir–La acted as a Lewis acid to accelerate
the rate-determining step of CO insertion (160).
More environmentally benign catalysts without using homogeneous
CH3I were also developed. Li et al. immobilized Rh and iodide species
on cross-linked copolymers via different functional groups for methanol
conversion. The methanol conversion reached 90% with 50% yield of
methyl acetate and acetic acid in a 2-h reaction at 393 K under 3.0 MPa
CO. Although the activity of the catalyst gradually decreased due to the con-
tinuous iodide leaching (>50 ppm), it can be regenerated over a 15 batch
test (being regenerated twice) (161). Another important option is used mor-
denite (H-MOR) based catalysts, which also showed unique selectivity to
acetic acid. Ellis et al. employed Cu-MOR catalysts for the halide-free car-
bonylation of methanol. After 6 h reaction, the selectivity to acetic acid and
methyl acetate surpassed 70% and maintained for ca. 12 h, after which it
gradually declined with dimethyl ether being produced. Only after the pores
are blocked with non-volatile MTG products, Cu-MOR become active for
carbonylation of methanol (162). Blasco et al. confirmed the reaction mech-
anism for the carbonylation of methanol over Cu-MOR catalysts by using
operando IR spectroscopy and “in situ” magic-angle spinning NMR spec-
troscopy. As shown in Fig. 9, different primary products were produced
over H-MOR and Cu-MOR catalysts. As to H-MOR catalysts, acetic
acid is the main product in contrast to methyl acetate over Cu-MOR cat-
alysts. Cu-MOR catalysts, however, show much higher reaction rate than
H-MOR does. Two neighboring active sites, one bridged hydroxyl and
a neighboring Cu+, were proposed for activating methanol and CO,
respectively (163).
DME is an attractive substrate to replace methanol in the direct carbon-
ylation as it has demonstrated higher selectivity to methyl acetate. Cheung
et al. used DME for carbonylation, and the H-MOR (Si/Al ¼ 10:1) catalyst
showed high selectivity (>99%) to methyl acetate at temperatures between
423 and 463 K. Under anhydrous conditions, DME reacted with protons to
Synthesis of ethanol and its catalytic conversion 115

H-MOR MOR-OCH3 + Dimethy ether + H2O

+ CO

+ H2O
CH3OH MOR-Acylium cation

CH3COOH + Cu-Dimethy ether


Cu-CO +

Methyl acetate

Cu-MOR MOR-OCH3 + Dimethy ether + H2O

Fig. 9 A possible reaction mechanism for the carbonylation of methanol with CO on


H-MOR and Cu-MOR catalysts (163). Reprinted (adapted) with permission from
Blasco, T.; Boronat, M.; Concepción, P.; Corma, A.; Law, D.; Vidal-Moya, J. A. Angew. Chem.
Int. Ed. 2007, 46 (21), 3938–3941. Copyright (2007) Wiley-VCH.

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.

2.5.2 Acetate hydrogenation


Acetic acid is an important carboxylic acid with a broad spectrum of appli-
cations such as in polymers, paints, foods and pharmaceuticals (172). It is also
a good candidate for the synthesis of ethanol owing to the rapid develop-
ment of the methanol or DME carbonylation.
Initially, the hydrogenation of acetic acid was investigated in biomass
pyrolysis oil upgrading in an attempt to remove the corrosive carboxylic
acid in bio oils. A series of metal catalysts have been tried for this reaction.
Rachmady and Vannice loaded Pt on different supports (TiO2, SiO2,
η-Al2O3, and Fe2O3) for the hydrogenation at 423–573 K in a fixed-bed
reactor. Over 2.01%Pt/TiO2 catalysts, the highest selectivity to ethanol
of 70% was achieved at 420 K (173). They also used bimetallic Pt–Fe cata-
lysts for acetic acid hydrogenation, and proposed that the hydrogen dissoci-
ation occurred on Pt sites, and acetic acid adsorption and activation occurred
on FeO to create the surface acyl species (174). Olcay et al. tried different
commercial catalysts including Ru/C, Pt/C, Pd/C, Rh/C, Ir/Al2O3,
Raney Ni and Raney Cu for this reaction at 383–563 K and 5.17 MPa
H2. The Ru/C catalyst showed approximately 80% ethanol selectivity at
ca. 433 K. DFT calculations and kinetic measurements revealed that the
high activity of Ru can be attributed to its intrinsic high reactivity in disso-
ciating acetic acid and acetate to acetyl (CH3CO*) species, which is likely to
be the rate-determining step for acetic acid hydrogenation (175,176). Even
though the selectivity to ethanol was improved over the above mentioned
metals and alloys, the conversion of acetic acid was lower than 20%. This
makes the whole process low efficiency. The addition of Sn to Pt catalysts
has been proved to increase the acetic acid conversion at high ethanol selec-
tivity due to the geometric and electronic effects. The Sn inhibited the
CdC bonds cleavage to form CO, CH4, and C2H6 even at 500–600 K,
and as a result, the acetic acid conversion increased to 80% with ca. 80% eth-
anol selectivity at 528 K, 4 MPa and a LHSV of 0.6 h1 (17,177). Different
supports, additives and preparation methods have also been widely investi-
gated to modify the Pt–Sn catalysts for improving the ethanol yield. Zhang
et al. loaded bimetallic Pt–Sn catalysts on different supports (CNTs, ZrO2,
TiO2, SiO2, and SiC) in the hydrogenation of acetic acid to ethanol. The
Pt–Sn/CNTs catalysts showed the best performance, exhibiting over 97%
acetic acid conversion and 92% ethanol selectivity at 623 K with a LHSV
of 1.5 h1 (178). Zhou et al. further modified Pt–Sn catalysts with K to
block the acid sites and modify the electronic structure of Pt. As the results,
side reactions such as ethanol decomposition, dehydration and esterification
Synthesis of ethanol and its catalytic conversion 117

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).

2.6 Production of ethanol from syngas via dimethyl oxalate


An alternative new approach for ethanol production from syngas has been
proposed via the DMO intermediate. In this approach, the CO was coupled
with methanol to produce DMO, followed by the hydrogenation of
DMO to ethanol with the produced methanol being recycled back to the
feedstock. The production of DMO from methanol and syngas has been
commercialized since 2010 with a capacity of >10,000 tons per year
(194). The product of DMO hydrogenation could be tuned between bulk
ethylene glycol and ethanol, which endows it a versatile process for produc-
ing bulk chemicals from syngas.

2.6.1 Coupling CO and methanol to dimethyl oxalate


Since the discovery of alkyl nitrite as an efficient re-oxidizing agent for
dibutyl oxalate synthesis by UBE, alkyl nitrite has been widely used in
the Pd catalyzed reactions such as the conversion of CO and methanol
to DMO (195). The reaction consists of two separate steps. CO is first
directly coupled with CH3ONO to produce DMO over the coupling cat-
alyst (Eq. 1). Then, the CH3ONO is regenerated by the reaction of
120 Jifeng Pang et al.

(CH3OCO)2
Pd0
DMO

2CH3ONO NO O + CH3OH

O O

CH3O Pd(II) OCH3+NO CH3O C Pd(II) C OCH3


DCPC
Pd(I1) alkoxyl
compound

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).

2CO +2CH 3 ONO ƒ!CH 3 COCOOCH 3 + 2NO (1)


1=2O2 + 2CH 3 OH + 2NO ƒ! 2CH 3 ONO +H2 O (2)
2CO + 2CH 3 OH +1=2O2 ƒ!CH 3 COCOOCH 3 + H2 O (3)

The CO coupling reaction was first developed by Fenton and Steinwand


in 1974. Over the PdCl2-CuCl2 system, 93% yield of carbonate and oxalate
was obtained. However, the reaction suffered from catalysts deactivation
and product separation issues (198). Afterwards, various heterogeneous
Pd-based catalysts were used for the DMO production in gas-solid phase
systems. Currently, DMO has been commercially produced over Pd/α-
Al2O3 catalysts in fixed-bed reactors (195,199). Nevertheless, the Pd load-
ing in this catalyst is 2 wt% with large Pd nanoparticles (10–20 nm), which
greatly increases the catalyst cost and reduces the process economy.
Synthesis of ethanol and its catalytic conversion 121

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.

exhibited a much high Pd efficiency of 733 DMO (g)Pd (g)1h1(194,209).


They added K ions in Pd/NaY catalysts for the CO coupling reaction to
DMO. The K ions were found to enhance the electron density of Pd species,
and facilitate the CO activation, which caused a 25% increase in the STY to
696 g L1 h1, with 86.3% selectivity to DMC at 383 K, 3.6% K loading and a
GHSV of 8000 h1 (210).

2.6.2 Hydrogenation of dimethyl oxalate to ethanol


Besides ethylene glycol, DMO is also an important platform chemical for
ethanol production (211,212). The hydrogenation of DMO to ethanol
follows two main routes. As shown in Fig. 12, the major way involves
the direct hydrogenation of DMO to ethanol over Cu-based catalysts
via the intermediates of methyl glycolate and ethylene glycol. Although
the role of Cu species is still in debate, it is widely accepted that Cu0 sites
activate H2 and adsorb the intermediates through the hydroxyl group,
whereas the Cu+ species stabilize the methoxy and acyl species. The syn-
ergistic effect between Cu0 and Cu+ sites improves the reactivity of the
ester group in DMO and the –OH group of ethylene glycol (191). Differ-
ently, in the minor way over some carbide catalysts, the methyl glycolate
was first dehydrated to methyl acetate, and then hydrogenated to ethanol,
as discovered by conditional experiments (213).

H3C O O
H3C O
+H2

O
-CH3OH
O CH3 O OH

Dimethyl oxalate Methyl glycolate


+H2

-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

Cu sites are found to be highly active for the selective hydrogenation of


CdO bonds and relatively inert for the cleavage of CdC bonds. Hence,
Cu-based catalysts have gained much attention for the selective hydrogena-
tion of DMO to ethanol. Gong et al. developed an ammonia evaporation
hydrothermal method to prepare Cu/SiO2 catalysts for DMO hydrogena-
tion. The ethanol yield reached up to 83% over 20 wt% Cu/SiO2 catalysts at
553 K and a LHSV of 2.0 h1. Based on characterizations, the authors
ascribed the high activity to the coexistence of Cu0 and Cu+. Cu0 is the
key active site for the hydrogenation, while Cu+ facilitates the conversion
of intermediates (i.e., ethylene glycol) to ethanol (214). To balance and sta-
bilize the Cu0/Cu+ active species, the same group prepared a core-sheath
nano-reactor for this reaction. As shown in Fig. 13, the catalyst consist of
a copper-phyllosilicate nanotube (CuPSNT) sheath with Cu NPs either
inside (Cu@CuPSNT-in) or outside (Cu@CuPSNT-out). Compared with
Cu@CuPSNT-out, the Cu@CuPSNT-in nanocatalysts displayed superior
catalytic performance with 92% yield of ethanol at 553 K and a LHSV of
2.0 h1. Additionally, the yield to ethanol still reached 85% after 12 h heat

Fig. 13 Comparison of catalytic performance and stability over Cu@CuPSNT-in, Cu@


CuPSNT-out and Cu/SiO2 catalysts. Reaction conditions: LHSV ¼ 2.0 h1, H2/DMO ¼ 200,
553 K (215). Reprinted (adapted) with permission from Yue, H.; Zhao, Y.; Zhao, S.; Wang, B.;
Ma, X.; Gong, J. Nat. Commun. 2013, 4, 2339. Copyright (2013) Springer Nature.
124 Jifeng Pang et al.

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).

2.7 Comparison of different methods for ethanol production


Due to the great demands of green fuels, a series of methods, as mentioned
above, have been developed for ethanol production. State-of-the-art, pos-
sible intermediates, as well as the advantages/disadvantages of these processes
are summarized and discussed in Table 1.
126 Jifeng Pang et al.

Table 1 Different methods for ethanol production.


Process Intermediates Advantages Problems Status

Sugars Sugar With >90% Elevating feedstock Commercialized


fermentation efficiency cost and conflicting
with food supply
Lignocellulose Abundant Low efficiency Pre-commercial
fermentation feedstock caused high cost
Ethylene Ethylene – Low efficiency and Vanished
hydration high cost
Direct thermal Syngas Abundant Wide products Pre-commercial
conversion of feedstock distribution
syngas
Fermentation Syngas Feedstock Low efficiency Pilot
of syngas compatibility
Methanol Acetic acid High Homogenous Commercialized
carbonylation efficiency catalysts
Dimethyl ether Methyl High – Pilot/
carbonylation acetate efficiency Commercializing
Syngas coupled DMO Multiproducts Low atom –
with methanol economic

Fermentation of sugars derived from crops is the dominant method for


ethanol production, donated as the first generation ethanol. It is a well-
established technology at commercial scale with >90% theoretical yield.
Great efforts have been made for improving the profit margin by increasing
the fermentation efficiency and decreasing the separation cost. Neverthe-
less, the conflict between food and fuel is still in debate. In the past decades,
more efforts have been dedicated to the fermentation of lignocellulosic
biomass to ethanol, which is the second generation ethanol. This process
involves feedstock collection, pretreatment, hydrolysis, fermentation and
separation steps, all of which determines the feasibility of converting ligno-
cellulosic biomass to ethanol. So far, cellulosic ethanol is still not compet-
itive with sugar/starch ethanol. However, the technology development
and engineering integration will greatly reduce the cost of cellulosic etha-
nol, and will promote the transformation in ethanol production from
traditional crops to cellulosic biomass feedstock. At this stage, a full use
of all components in biomass and zero waste discharge should be seriously
considered. In the long run, the development of new microorganisms is
Synthesis of ethanol and its catalytic conversion 127

required to ensure co-fermentation of C5 and C6 sugars efficiently, and the


cellulase enzyme costs should be minimized. This will minimize the cost
gap between the first generation and second generation ethanol (227).
Direct hydration of ethylene to ethanol is unattractive for large-scale pro-
duction in most areas since it is not competitive to fermentation ethanol in
terms of reaction efficiency, cost of raw material and catalysts. Thermal con-
version of syngas to ethanol is a single-step process, and has been extensively
investigated in the past century. Different types of catalysts have been devel-
oped, and various in situ and operando characterization techniques in combi-
nation with theoretical calculations have been dedicated to the mechanism
studies. Nevertheless, it is still a challenge to produce ethanol selectively
due to the complex product distribution and high energy consumption.
Gas fermentation process is rapidly developed, which offers great advantages
in terms of feedstock flexibility, carbon efficiency and production economics.
This process can be operated in a broad range of syngas components without
requiring for an external gas shift operation in the F-T process. Additionally,
the development of gene technology and membrane-based bioreactors has
also improved the reaction efficiency to some extent (142). Therefore, some
pilot or semi-commercial-scale processes have been in operation, (228)
trying to commercialize this process using industrial waste gas streams or
natural gas (229).
Methanol carbonylation process has been commercialized for acetic acid
production using metal-organic complexes and halogen promoters. It has a
high efficiency with ca. 99% selectivity to acetic acid, but has the drawback
of difficulty in catalysts separation. To solve this problem, some efforts have
been made on developing heterogeneous catalysts for a vapor-phase car-
bonylation reaction. The underlining investigation of heterogenized homo-
geneous catalysts showed promising catalytic performance for applications.
Some pilot plants in China have been deliberating to commercialize this
process in the near future (149). Using dimethyl ether as the feedstock
for carbonylation is different from the methanol carbonylation process. It
is a halide-free process with environmentally benign zeolite catalyst in a
high-pressure fixed-bed reactor. After years of exploration, highly stable cat-
alysts have been developed. A 100,000 ton/year pilot is currently running in
China, and several larger scale plants are under construction. To obtain eth-
anol, the resulting acetic acid and methyl acetate are hydrogenated over
Cu-based catalysts. The last methanol conversion process is through the
DMO intermediate, another important chemical in coal chemical industry.
In this process, multiple products including DMO, EG and ethanol can be
128 Jifeng Pang et al.

produced, which makes it a versatile process for producing important


chemicals. Based on the product value, market needs and atom economy
considerations, EG is the primary product. However, ethanol is also an
option owing to the large market requirements. In the past decades, great
efforts have been made for ethanol production. Several processes have been
successful in lab and now operating in pilot or larger scales. For example, the
two-step conversion of methanol to ethanol via the methyl acetate process is
on the stage of pre-commercialization. The chemical way for ethanol pro-
duction from biomass has appeared recently, which provides a potential way
for efficient ethanol production.

3. Catalytic conversion of ethanol to chemicals


and advanced fuels
Up to date, ethanol is one of the largest bulk bio-chemical. As men-
tioned above, the global ethanol production is anticipated to further increase
annually. The increasing availability and decreasing cost of ethanol also stim-
ulate the exploration of ethanol upgrading. Other than being used as a fuel
additive to blend with gasoline, ethanol also serves as an important platform
molecule for synthesizing value-added chemicals and advanced fuels. Spe-
cifically, ethanol is readily converted into other important chemicals such
as acetaldehyde, ethyl acetate, propylene, acetone, 1,3-BD and isobutene
that are difficult to be synthesized from fossil resources. In addition, ethanol
can also be upgraded to advanced fuels including hydrogen, butanol and
hydrocarbons to satisfy the demands of modern engines. Using ethanol as
a fuel additive has been widely evaluated and discussed in some works from
environmental and sustainability point of view (230–232). Therefore, we
herein focus on the chemical conversion of ethanol to important chemicals
and advanced fuels.

3.1 Catalytic conversion of ethanol to chemicals


3.1.1 Conversion of ethanol to acetaldehyde and ethyl acetate
Acetaldehyde is an important intermediate for producing a variety of
chemicals such as pyridine, pentaerythritol, and acetate esters. Currently,
acetaldehyde is mainly produced by the oxidation of ethylene over a homo-
geneous PdCl2 and CuCl2 system named the Wacker process (233). Owing
to the corrosiveness of catalysts and the high cost of purification process, the
Wacker process is gradually replaced by the direct conversion of ethanol
method. Generally, two ways have been developed for the synthesis of
Synthesis of ethanol and its catalytic conversion 129

acetaldehyde from ethanol, i.e., non-oxidative dehydrogenation and oxi-


dative dehydrogenation of ethanol. The former reaction uses ethanol as
the solo feedstock, and generates acetaldehyde with hydrogen as a useful
by-product. Sometimes, ethyl acetate is co-produced even as a main prod-
uct depending on catalysts and reaction conditions. In another method,
oxygen and ethanol are co-fed, producing acetaldehyde with water as
the by-product. To maximize the acetaldehyde yield, the oxidation reac-
tion should be controlled to prevent the excessive oxidation products such
as acetic acid and carbon dioxide.

[Link] Non-oxidative dehydrogenation of ethanol


Direct dehydrogenation of ethanol to acetaldehyde is an attractive method as
it avoids the over oxidation of substrates, and produces another useful prod-
uct, hydrogen, rather than water. Various catalysts have been exploited for
this reaction. Generally, the active sites of catalysts are metals from group IB,
Pd and some special carbon materials.
Cu-based catalysts have demonstrated high activity for the ethanol dehy-
drogenation (234). However, the selectivity and stability of catalysts are
highly dependent on the structure of Cu. Bulk Cu foam has been used
for the ethanol dehydrogenation, and shows inferior performance to that
of supported Cu catalysts due to the large Cu particles and low accessible
of Cu sites (235). Hence, various supports were used to disperse Cu species
for this reaction. Chang et al. loaded Cu species on rice husk ashes for eth-
anol dehydrogenation at 473–573 K. Owing to the high dispersion of Cu,
the selectivity to acetaldehyde was 100% at <80% ethanol conversion, supe-
rior to the benchmark catalyst of Cu/SiO2 (236). They also modified the Cu
with Cr on silica and explored other catalyst synthesis method (i.e., ion
exchange) to promote the catalyst activity and mitigate the catalyst deac-
tivation. As the result, the dispersion of Cu was greatly enhanced from 7%
to >60%, which led to a strong interaction with Cr, and thus resulted in
the enhanced activity (ca. 80% ethanol conversion at 548 K) and stability
for acetaldehyde formation (237,238). Recently, Wang et al. employed
the MC to support Cu species for ethanol conversion. The inert inner sur-
face of MC enriched the ethanol and the scarce surface functional groups
(some –OH and –COOH groups) inhibited the secondary reactions of
acetaldehyde. Hence, the selectivity to acetaldehyde reached 95.1% with
83% ethanol conversion at 553 K and a GHSV of 26,000 h1, which is
much higher than that of Cu/SBA-15 catalysts. Kinetic measurements
indicated that the selectivity to acetaldehyde is influenced by the surface
130 Jifeng Pang et al.

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. 15 Arrhenius-type plots of the reaction rate in ethanol dehydrogenation, normal-


ized by the surface area of Cu over Cu, Pd0.01Cu, Pt0.01Cu, and Ni0.01Cu nanoparticles
(244). Reprinted (adapted) with permission from Shan, J.; Liu, J.; Li, M.; Lustig, S.; Lee, S.;
Flytzani-Stephanopoulos, M. Appl. Catal. B Environ. 2018, 226, 534–543. Copyright
(2018) Elsevier.
Synthesis of ethanol and its catalytic conversion 131

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).

[Link] Oxidative dehydrogenation of ethanol


Oxidative dehydrogenation of ethanol employs oxygen to activate ethanol.
This process has a lower standard enthalpy compared to the non-oxidative
dehydrogenation reaction (Eqs. 4 and 5). Hence, the reaction could be con-
ducted under relatively mild conditions. After years of explorations, several
effective catalysts including Au-based catalysts, Pd-based catalysts and some
non-noble metal catalysts have been developed for this reaction.
Cat:
C2 H5 OH ƒƒƒ! CH 3 CHO + H2 ΔH ¼ 68 kJ=mol (4)
Cat:
C2 H5 OH +1=2O2 ƒƒƒ! CH 3 CHO +H2 O ΔH ¼ 179 kJ=mol (5)
Since Masatake Haruta disclosed the high activity of Au nano particles in
CO oxidation, interests in Au-based catalysis have increased dramatically
(261). Various Au catalysts with different particle sizes, morphology, and
on different supports have been synthesized and tested in oxidative dehydro-
genation of ethanol, as summarized in Table 2. The particle size and state
of Au are crucial to oxidation reactions. Zheng and Stucky synthesized
monodispersed Au nanoparticles on different supports including zeolites,
α-Fe2O3, TiO2, HAP, Al2O3, ZnO, fumed SiO2 and SiO2, and investigated
the size effect for ethanol conversion. The acetaldehyde selectivity was opti-
mized to ca. 90% with 20–24% ethanol conversion at 473 K over Au/SiO2
catalysts with Au particle sizes of 3.5 and 8.2 nm, much higher than that of
catalysts with 6.3 nm Au particles (ethyl acetate as the by-product) (262).
Gong and Mullins studied the reaction on Au(111) surface by using
temperature-programmed desorption and molecular beam reactive scatter-
ing. Results showed that ethanol first underwent OdH bond cleavage to
form a surface ethoxide. Then, βdCdH bond was selectively activated
to produce acetaldehyde and water over oxygen covered Au(111) (270).
Nanoporous gold was active for oxidative ethanol dehydrogenation, giv-
ing ca. 90% selectivity to acetaldehyde at 62% ethanol conversion, 473 K
134 Jifeng Pang et al.

Table 2 Catalytic conversion of ethanol to acetaldehyde over different catalysts.


Catalyst
(particle size nm) Conditions Con. Sel. Ref.
0.5%Au/SiO2(3.5) 473 K; ethanol(l): O2 ¼ 1:1000; 1 g catalyst 22 90 (262)
Nanoporous Au 473 K; ethanol(g): O2 ¼ 1:2 22 90 (263)
1.0%Au/MoO3 513 K; ethanol(g)/O2 ¼ 1/3; 99 94 (264)
SV ¼ 20,000 mL h1 gcat1
2%Au/TiO2(2.1) 398 K; ethanol(g)/O2 ¼ 1/9; 58 76 (265)
GHSV ¼ 3600 h1
1%Au/Recryst-S1 473 K; ethanol(g)/O2 ¼ 1/1 50 98 (266)
Au–CuOx/SiO2 473 K; ethanol/O2 ¼ 0.61:1 90 >80 (267)
0.9%Au/ 523 K, ethanol(g)/O2 ¼ 1/3, >99 95 (18)
MgCuCr2O4 GHSV ¼ 100,000 mL g cat1 h1
0.2%Au/0.2%Cu 523 K, ethanol/O2 ¼ 1:1.2; 100 100 (268)
GHSV ¼ 4800 h1
Au/CuSiO3 523 K, ethanol/O2 ¼ 1:3; 98 93 (269)
GHSV ¼ 100,000 mL g cat1 h1

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.

3.1.2 Conversion of ethanol to propylene and acetone


Propylene is an important starting material for preparing polymers, and also a
key chemical intermediate for synthesizing propylene oxide, polypropylene
and acrylonitrile. The demand for propylene is growing rapidly with an
annual rate of 5–6% (277). Propylene is generally produced by the fluid cat-
alytic cracking and steam thermal cracking of naphtha. Owing to the limited
storage of petroleum resources and deteriorating environmental issues, con-
version of renewable ethanol to propylene provided an alternate option for
sustainable production of propylene.
Catalytic conversion of ethanol to propylene is a rather complex reac-
tion involving in CdC bond formations and cleavages. The reaction path-
ways are highly dependent on catalysts, as summarized by Iwamoto et al. in
2014 (278–280). As shown in Fig. 17, over Ni-MCM-41 catalysts (1–5
steps in Fig. 17), ethylene is the key intermediate for propylene formation,
which is first dimerized to 1-butene and then isomerized to butenes. The
2-butenes finally reacts with another ethylene to produce propylene. Dif-
ferently, over Sc/In2O3 catalysts (6-12-9-10-11, steps in Fig. 17), ethanol is
first dehydrogenated to acetaldehyde, and then oxidized to acetic acid with
water or a surface hydroxyl group. The obtained acetic acid is further
converted into acetone and carbon dioxide via the ketonization reaction.
Finally, the formed acetone is hydrogenated and dehydrated to propylene.

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

To get a high yield of propylene, different reaction steps should be bal-


anced, and multiple active sites need to be coupled in one catalyst.
H-ZSM-5 is the most investigated zeolite for the conversion of ethanol
to propylene. Fujitani’s group screened ZSM-5 catalysts with different
Si/Al ratios and additives for the reaction. The Zr-ZSM-5 catalyst with
a Si/Al ratio of 80 afforded the highest propylene yield of 32% at 773 K
with ethylene as the main by-product (281). Afterwards, they investigated
the P modified ZSM-5 zeolites (282) and MFI-type zeolites (283), and
found that the acid sites of catalysts greatly affect the propylene yield.
The reaction follows the route of dehydration of ethanol to ethylene
and then forms propylene via carbine species (284). Sano’s group found that
the acid strength of catalyst is a crucial factor for the propylene production.
They modified the H-ZSM-5 with a variety of alkali metals and P additives
to improve the propylene yield (285–287). The same strategy was
employed by several other researchers (288–291). Nevertheless, the yield
of propylene was still around 30%. The crystal size of the zeolite is another
factor affecting the propylene selectivity. Small-sized HZSM-5 showed
higher propylene selectivity and better stability than the larger ones, which
was attributed to the abundant secondary pores and short channels that
reduced the diffusion path lengths for coke formation (292–294). Given
the high activity of HZSM-5 and SAPO-34 catalysts for the direct conver-
sion of ethylene to propylene, (295,296) HZSM-5/SAPO-34 composites
catalysts were synthesized and investigated in ethanol conversion. A higher
propylene yield of ca. 34.5% was achieved at 775–823 K over a HZSM-5/
SAPO-34 weight ratio of 4 (297,298). According to the thermodynamic
evaluation, pressures lower than 1 bar and temperatures higher than
523 K are favorable for high propylene production. The maximum propyl-
ene yield is roughly 42% irrespective of the reaction conditions (299).
Inspired by the formation of ketones from alcohols over mixed oxide
catalysts, transition metal oxides have been explored for the conversion
of ethanol to propylene. Iwamoto’s group found that In2O3-based oxides
were notably active for the reaction. Various metals were loaded on In2O3
to improve the stability and activity of catalysts. Over the Sc-modified
In2O3 catalysts, the yield to propylene reached 60% in the presence of water
and hydrogen. The Sc addition prevented the reduction of In2O3, and the
co-feeding of water and hydrogen inhibited the coke formation and ace-
tone decomposition, respectively. Hence, the stability of catalysts was
improved significantly (300,301). They also investigated Ce-based catalysts
for this reaction with 31 metals as additives. Yttrium (20%) on Ce gave the
138 Jifeng Pang et al.

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).

2CH 3 CH 2 OH +H2 O ƒ!CH 3 COCH 3 +CO2 +4H2 (6)

Apart from the propylene production, acetone is also selectively pro-


duced over a variety of oxide catalysts (Eq. 6). Since 1930s, the catalytic
conversion of ethanol to acetone has been investigated. At first, various
basic oxide catalysts such as Fe2O3-ZnO/CaO and ZnO-CaO were used,
giving 50–90% yields of acetone with ca. 60–100% ethanol conversion at
673–723 K. They suggested that basicity of catalysts enhanced the acetate
species (CH3COO–) formation, which was the rate-determining step for
this process (oxidization then condensation) (304–306). Later, Nishiguchi
et al. observed that acetone is the main product over CuO/CeO2 catalysts
during their investigation on the steam reforming of ethanol at 653 K. They
proposed that acetone was produced via the aldol condensation of acetal-
dehyde, which was oxidized by the O of the support and then decomposed
to produce carbon dioxide, acetone and hydrogen (condensation then
oxidization) (307). Recently, Appel’s group did a series of work on the
mechanism study. They used physical mixture catalysts of Cu/ZnO/
Al2O3 and m-ZrO2 to study the process of acetone synthesis. It was found
that ethanol was dehydrogenated to acetaldehyde over Cu species, and the
acetaldehyde was then oxidized to acetate by the superficial O of ZrO2.
The acetate species were condensed to acetone and CO2 (oxidization then
condensation). During this process, water was dissociated on the Cu surface
to produce H2 and O species that migrated to ZrO2 (308). In the subsequent
work, they found that the strong base and acid sites were crucial to the dehy-
drogenation and condensation reactions, and Zn2+ species were responsible
to the water dissociation for closing the catalytic cycle over the ZnxZr1–
xO2–y-based catalysts (309). Based on the reaction mechanism, they used
Ag to dope CeO2 for this reaction, and achieved elevated acetone selectivity
(60%) with depressed ethylene selectivity. The improved activity of
Ag/CeO2 lies in the excellent catalytic activity of CeO2 in H2O dissociation
and the synergistic effect between Ag and CeO2 (310).
Synthesis of ethanol and its catalytic conversion 139

3.1.3 Conversion of ethanol to 1,3-butadiene and isobutene


1,3-BD is an important monomer in the synthesis of rubbers, elastomers and
resins with a consumption of >10 million metric tons (311). The production
of 1,3-BD from ethanol has a long history, and some excellent works have
been dedicated to this subject (19,24). In 1920–1930s, the catalytic conver-
sion of ethanol to 1,3-BD was extensively investigated, and plant scale pro-
duction of 1,3-BD was achieved in Russia and German. In the past decade,
this process has been revisited because of the low availability of oil feedstocks
for cracking, the exploration of shale gas and the increasing interest in green
chemistry. One pot conversion of ethanol to 1,3-BD has been regarded as a
valuable method for the sustainable production of 1,3-BD (312,313).
Although it has been proved to be an industrially practiced technology, it
still needs a thorough investigation to find new catalysts for more successful
industrial applications and to get a better understanding of the reaction
mechanism.
The reaction mechanism for converting ethanol to 1,3-BD is rather com-
plex, and many pathways have been suggested and debated (24,314). It
is generally accepted that it is a multi-step reaction, as shown in Fig. 18. Eth-
anol first undergoes dehydrogenation to acetaldehyde, followed by a self-
condensation to yield acetaldol. Then, dehydration of acetaldol produces
crotonaldehyde, which is further reduced to crotyl alcohol via the MPV
reaction with ethanol. Finally, crotyl alcohol is dehydrated to 1,3-BD
(highlighted in bold black). Taifan et al. studied different reaction steps by

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.

[Link] The one-step reaction process for conversion of ethanol to 1,3-BD


[Link].1 Mixed oxide catalysts Mixed oxide catalysts, with adjust-
able basic and acidic properties, have long been proved to be active in con-
verting ethanol to 1,3-BD. One of the typical and most widely investigated
is mixed Mg and Si oxides, in which the Mg oxide provides the base sites,
and the interface between Mg oxide and silica offers the acid sites. Specif-
ically, SiO2 increases the structural defects in MgO, forming stepped MgO
surface to promote the dehydrogenation of ethanol, as proved by Zhang
et al. in recent study (317). To get a high yield of 1,3-BD, the basic and
acidic properties of catalysts were subtly adjusted by changing preparation
methods or Si/Mg ratios.
Different preparation methods have been used to improve the 1,3-BD
selectivity. As shown in Table 3, Natta et al. used the wet-kneading
method, i.e., a wet method by mechanically or magnetically mixed the
solid precursor materials, to prepare the SiO2–MgO materials, which
showed super performance to the mixture of MgO and SiO2 in terms of
activity and stability (318,329). These conclusions were consistent with
Weckhuysen group’s results that the wet-kneaded SiO2–MgO catalysts
were more basic, and the smaller SiO2 spheres with MgO provided the best
balance of base and acid sites (319,330,331). Additionally, the SiO2–MgO
Synthesis of ethanol and its catalytic conversion 141

Table 3 Conversion of ethanol to 1,3-BD over different mixed oxides catalysts.


WHSV S1,3-BD
Catalystsa Preparing method T (K) (h21) Con. (%) (%) Ref.
MgO/SiO2(0.83) Wet-kneading 623 0.03 53.3 30 (318)
SiO2–MgO(1:1) Physical mixture 698 – 18.3 1.6 (319)
SiO2–MgO(1:1) Wet-kneading 51.5 31.6
SiO2–MgO(1:1) co-precipitation 46.1 13.9
MgO-SiO2(1:1) Wet-kneading 623 0.15 50 84 (320)
[Mg(OH)2]
MgO––SiO2(1:1) Wet-kneading 0.15 71 7
[MgCl2]
NiO–MgO– Impregnated 553 12.5 59 90 (321)
SiO2(0.2:1.5:1)
4%ZnO/MgO– Impregnated 648 1 56.0 62.1 (322)
SiO2(1:1)
ZnO-talc(1:2)b Preparation method 673 – 41.8 50.1 (323)
ZrO2–ZnO/ Impregnated 598 0.3 32 68 (324)
MgO:SiO2(95:5)
ZrZn/MgO– Co-precipitation 648 0.62 40 35.9 (325)
SiO2(1:1)
K/ZrZn/MgO– Co-precipitation 44 57.8
SiO2(1:1)
Na–ZnxZryOz Hard template 623 6.2 97 47 (326)
Zn–La–Zr–SiOx Impregnated 598 1 17.7 70.2 (327)
673 2 100 60.2
Zn–Y/beta Two-step metallation 603 0.3 90 81 (328)
procedure
673 7.9 82 63
a
The number in bracket is the mass ratio of Mg/Si.
b
The contact time (W/F) is 5.49 g h mol1.

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.

catalyst for 1,3-BD production have a MgO:SiO2 ratio of 1:1 (320).


Makshina et al. dry milled different ratios of Mg(OH)2 and silica for this
reaction, and found a MgO:SiO2 ratio of two to be the best (332). Recently,
Cavani’s group used the MgO–SiO2 catalysts prepared by the sol-gel tech-
nique for ethanol conversion. The catalysts with a low Si-content, i.e.,
Mg/Si atomic ratio in the range between 9 and 15, gave better 1,3-BD
yields. Over these catalysts, the base and Lewis acid sites co-existed, which
seems critical to the ethanol activation and the dehydration of intermediately
formed alkenols to 1,3-BD (333). Nevertheless, the best MgO:SiO2 ratio
differs greatly in literatures, which should be attributed to the different
reaction conditions and catalyst preparation methods.
Although the SiO2–MgO-based catalysts have been extensively inves-
tigated, the yield of 1,3-BD still needs to be improved. Some metal oxides
have been employed to modify mixed oxide catalysts. Kitayama et al. syn-
thesized large surface area ternary oxide catalysts of nickel magnesium
silicate for ethanol conversion. By adjusting the ratio of Ni, Mg and Si
(Si/Mg ¼ 1.5), the acidic and basic properties of catalysts were optimized
to achieve 90% selectivity to 1,3-BD at ca. 59% ethanol conversion
(321). Larina et al. used ZnO to increase the activity of SiO2–MgO catalysts.
The introduction of ZnO caused two types of Lewis acid sites (the contact
zones of MgO–SiO2 and ZnO–SiO2 phases), which enhanced the ethanol
dehydrogenation step, and eventually improved the 1,3-BD selectivity
from 27% to 62.7% at 623 K (322). These results were consist with Baba’s
conclusions, suggesting that the introduction of Zn is crucial to improve the
acid and base sites of catalysts, and correspondingly increase the 1,3-BD
yield and catalysts stability (323).
Some researchers used multiple metal oxides to improve the activity of
MgO–SiO2 catalysts. Jones’s group supported Zr(IV) and Zn(II) on MgO–
SiO2 at a weight ratio of 95:5 to elevate the 1,3-BD selectivity to 68.7% at
598 K. A synergistic effect between ZrO2 and ZnO was observed. ZnO
promoted the ethanol dehydrogenation step and ZrO2 was expected to assist
the aldol condensation and crotonaldehyde reduction steps (324). They also
added alkali metals to reduce the acid sites of catalysts in order to suppress
the dehydration of ethanol and increase the 1,3-BD selectivity (72% selec-
tivity to 1,3-BD and acetaldehyde) (325). Gao et al. investigated the
promotion effect of ZrO2 through the experimental study and DFT cal-
culations. Among different facets of ZrO2, the ZrO2 (111) is the best one
in the adsorption of reaction species with unique acidic and basic proper-
ties, which showed high reactivity and selectivity (major by-product is
ethylene) (334).
Synthesis of ethanol and its catalytic conversion 143

Apart from MgO–SiO2-based catalysts, many other oxides have also


been used for ethanol conversion to give considerable 1,3-BD yields.
ZnxZryOz is a versatile material with abundant Lewis acid-base pairs. Baylon
et al. used Na to dope the ZnxZryOz-H for decreasing the surface strong
Brønsted acid sites that are responsible for the side reaction like ethanol
dehydration. Over 2000 ppm Na doped Zn1Zr10Oz, the base and weak
Brønsted acid sites remained to give 47% selectivity to 1,3-BD at 97% eth-
anol conversion (326). Larina et al. synthesized different Zn–La–Zr–Si oxide
catalysts for this reaction. The Zn-containing sites enhanced the ethanol
dehydrogenation rate, La-containing base sites and Zr-containing acid sites
increased the rates of the aldol condensation and MPV reductions. There-
fore, the selectivity to 1,3-BD was comparable to the industrial catalysts at a
high productivity of 0.7 gBD gcat 1 h1 (327). Dai et al. employed Beta con-
fined bicomponent Zn–Y clusters for this reaction. Due to the pore confine-
ment and suitable acid base sites of Zn-Y/Beta catalysts, high productivity of
2.33 gBD/gcat/h was achieved with a 1,3-BD selectivity of 63% at 673 K
and a WHSV of 7.9 h1 (328). Afterwards, they investigated the deactiva-
tion mechanism of Zn-Y/Beta catalyst in ethanol to 1,3-BD conversion. By
different characterizations, they found that the self and cross-condensation of
acetaldehyde and acetone produced the long chain unsaturated aldehydes/
ketones, which covered the active sites of catalysts, and leading to the catalyst
deactivation (335). Recently, Wang et al. used the MFI zeolite nanosheet
to support ZnHf species for the conversion of ethanol to 1,3-BD, which
demonstrated enhanced 1,3-BD selectivity and ethanol conversion than
the ZnHf-MFI (Micropore) catalyst (ca. 10% enhancement in ethanol con-
version and 1.3-BD selectivity). After doping with 1.7% Li, the 1.3-BD
selectivity reached 73% with 64.6% ethanol conversion at 593 K and a
WHSV of 0.47 h1 (336).

[Link].2 Metal enhanced mixed oxide catalysts The dehydrogena-


tion of ethanol to acetaldehyde is one of the key step for 1,3-BD production.
Metals are always introduced to metal oxides for activating the ethanol,
which promote the ethanol conversion even under mild conditions. Some
metals such as Ni and Ru not only enhance the ethanol dehydrogenation but
also trigger the CdC bonds cleavage and induce serious side reactions.
Hence, moderately active metals including Ag, Cu, Au are used to get a high
1,3-BD yield from ethanol. They can promote the ethanol conversion with-
out causing side reactions under the same reaction conditions.
Table 4 lists typical metal enhanced oxides catalysts for the conversion of
ethanol to 1,3-BD. Makshina et al. screened a variety of metals (Cr, Mn, Fe,
Table 4 Conversion of ethanol to 1,3-BD over different metal enhanced mixed oxides catalysts.
Catalysts Preparation T (K) WHSV (h21) Con. (%) S1,3-BD (%) Ref.
0.05%Ag/MgO–SiO2(2:1) Incipient wetness impregnation 623 0.014a 97.7 63 (332)
1.0Ag/MgO–SiO2 Incipient wetness impregnation 753 1.2 84 50 (337)
1Ag/10ZrO2/SiO2 Incipient wetness impregnation 593 0.31 45.6 73.0 (338)
1Ag/10ZrO2/SiO2 Incipient wetness impregnation 593 0.04 88 73.9
Ag/ZrBEA200 Incipient wetness impregnation 593 0.32 48 56 (339)
Ag/Zr3.5BEA75 Incipient wetness impregnation (post-synthesis method) 593 3 14.9 58.7 (340)
CuTaSiBEA Two-step postsynthesis method 598 0.5 88 73 (341)
CuO/SiO2–MgO(1:1)b Wet-kneading 698 – 74 49 (330)
Cu–Ag/MgO–SiO2 Impregnated (Benzoxaxine reduced) 573 – Ca. 62 75 (342)
Cu:Zr:Zn(1:1.5:0.5)/SiO2 Impregnated 648 1.5 44.6 67.4 (343)
Hf2.5Zn16/SiO2 Impregnation 633 0.64 99.2 71 (344)
Cu1.0Hf3.0Zn0.5/SiO2 Impregnation 633 0.64 99.4 72
Au/MgO–SiO2 Deposition–precipitation method 573 1.1 60 78.3 (345)
a
The data is LHSV.
b
Adding tetraethyl orthosilicate to ethanol/NH3.
Synthesis of ethanol and its catalytic conversion 145

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.

Cu is another important active metal in the conversion of ethanol to 1,3-


BD. Kyriienko et al. investigated the activity of different metal (Ag, Cu, Zn)
containing TaSiBEA for ethanol conversion. The metals modified the acid
and base sites of TaSiBEA catalysts and added dehydrogenation sites, which
accelerated the ethanol dehydrogenation and subsequent steps of the ethanol
to 1,3-BD process. Over the CuTaSiBEA catalyst, the selectivity to 1,3-BD
reached 73% at 88% ethanol conversion, 598 K and a WHSV of 0.5 h1
(341). Angelici et al. introduced CuO to the MgO–SiO2, which boosted
the production of the intermediate of acetaldehyde and changed the rate-
determining step of the process. As a result, the 1,3-BD yield and selectivity
increased to 40% and 53%, respectively, at 698 K (330). Through the
operando measurements, the same group identified that the majority of
Cu2+ sites in the catalyst were rapidly reduced to metallic Cu under the reac-
tion conditions, which was responsible for the increased dehydrogenation
activity and the BD yield. In addition, the deactivation of that catalyst was
ascribed to carbonaceous deposition on the active sites rather than sintering
of the metal particles (350). Tripathi et al. introduced Cu to Ag/MgO–SiO2
catalysts by a benzoxaxine reduction method. This approach enhanced the
stability of the catalyst and improved the 1,3-BD selectivity to some extent
(46.4% 1,3-BD yield after 6 h TOS) (342).
Multi-metallic catalysts have been synthesized for the conversion of eth-
anol to 1,3-BD. Jones et al. utilized a variety of silica impregnated bi- and
trimetallic catalysts for the conversion of ethanol into 1,3-BD. Both Zn(II)
and Zr(IV) sites showed suitable Lewis acid, which were crucial to the
desirable acid sites catalyzed condensation but not for ethanol dehydration
to by-products (i.e., ethylene and diethyl ether). The highest selectivity of
67% was obtained over the Cu:Zr:Zn(1:1.5:0.5)/SiO2 catalyst (343).
Recently, De Baerdemaeker et al. studied silica supported Cu(II) or
Hf(IV) and Zn(II) for this reaction. The addition of Cu(II) or Hf(IV) and
Zn(II) improved the 1,3-BD selectivity to 72% at nearly 100% ethanol con-
version. The Cu(II) acted as the dehydrogenation catalyst, likewise the
Zn(II) over Hf(IV) and Zn(II) catalysts. Moreover, the Zn(II) component
effectively suppressed the dehydration activity of Hf(IV), which mainly cat-
alyzed the condensation of acetaldehyde. The usage of hemimorphite as
the Zn(II) precursor was proved to be successful. It suppressed the ethylene
formation and promoted the production of 1,3-BD to 0.264 gBD gcat 1 h1
at 70.2% yield and 99.2% ethanol conversion under the reaction conditions
of 633 K using a feed rate of 0.64 gEtOH gcat 1 h1 (344).
Synthesis of ethanol and its catalytic conversion 147

The metallic Au supported on MgO–SiO2 was also tried for converting


ethanol to 1,3-BD. Shylesh et al. used Au nanoparticles to promote the eth-
anol redox processes over Au/MgO–SiO2 catalysts, leading to the enhanced
ethanol dehydrogenation. A high activity and selectivity to 1,3-BD (60%)
was achieved at a low temperature of 573 K and a high partial pressure of
ethanol (20 kPa) via a subtle balance of acid–base and redox sites. The
authors also found that this process reduced the greenhouse gas emissions
by as much as 155% as compared to the petroleum-based process (345).

[Link] The two-step process for conversion of ethanol to 1,3-BD


Besides the above-discussed one-step process, a two-step process, named
after the Ostromyslensky’s process, is also used for 1,3-BD production. It
was commercialized by Carbide and Carbon Chemicals Corporation in
the United States (314). Usually, the two-step process includes dehydroge-
nation of ethanol to a mixture of ethanol and acetaldehyde, followed by the
transformation of ethanol and acetaldehyde into 1,3-BD. In detail, metal
catalysts such as Cu were used for ethanol dehydrogenation in the first step,
then ethanol and acetaldehyde were co-converted to 1,3-BD over different
condensation catalysts.
To simplify the two-step process, ethanol and acetaldehyde mixture was
directly used as the feedstock in the second step reaction over different cat-
alysts including HTs, sepiolites, aluminum oxides, and silica-based mate-
rials (351). Some other new materials were also investigated for this
process. For instance, Chae et al. screened a series of Ta oxide supported
on ordered mesoporous silica with different pores and particle sizes. Over
the Ta/SBA-15 catalyst, 47.1% of ethanol and acetaldehyde conversion
with 79.0% selectivity to 1,3-BD was achieved at an ethanol/acetaldehyde
molar ratio of 2.5 (352). Kyriienko et al. incorporated Ta into the vacant
sites of BEA zeolite by a two-step post-synthesis method. This special struc-
ture improved the 1,3-BD selectivity to 90% at 30.7% feedstock conversion
with an ethanol/acetaldehyde molar ratio of 2.2 (353). Han et al. synthe-
sized a series of ZrO2–SiO2 catalysts for ethanol and acetaldehyde conver-
sion. Owing to the suitable density of Lewis acid sites from highly dispersed
ZrO2, 69.7% 1,3-BD selectivity was obtained at 2 wt% ZrO2 loading with
an ethanol/acetaldehyde molar ratio of 3.5 and a WHSV of 1.8 h1 (354).
To better understand the reaction mechanism for the mixed ethanol/
acetaldehyde to 1,3-BD reaction, M€ uller et al. employed DRIFTS-MS for
in situ investigation of Ta-BEA catalysts. According to their studies, two
acetaldehyde molecules underwent condensation to form crotonaldehyde,
148 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).

[Link] Conversion of ethanol to isobutene


Besides 1,3-BD, another C4 olefin of isobutene has also been selectively syn-
thesized from ethanol via the intermediate of acetone. As shown in
Eqs. (6)–(8), ethanol is first converted to acetone, carbon dioxide and hydro-
gen in the presence of water, and the formed acetone is further converted to
isobutene with the release of another carbon dioxide.

3CH 3 COCH 3 ƒƒ! 2i  C4 H8 +CO + H2 O + H2 (7)

3CH 3 CH 2 OH +H2 O ƒƒ! i  C4 H8 + 2CO + 6H2 (8)

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).

3.2 Catalytic conversion of ethanol to advanced fuels


Ethanol is being widely used as a green fuel additive, however, some issues
including low energy density, miscibility with water still inhibited it for large
150 Jifeng Pang et al.

portion mixed in gasoline. Therefore, some strategies have been developed


to convert ethanol into some advanced fuels such as hydrogen, butanol and
hydrocarbons.

3.2.1 Ethanol as a hydrogen carrier


Hydrogen is a clean energy carrier with energy density as high as 120.7 kJ/g.
Nevertheless, as a flammable gas, hydrogen is a highly dangerous compound
to be stored and transported. Finding an economic and safe way for hydro-
gen storage and transportation is still challenging. An attractive solution of
this subject is to use small oxygenates as hydrogen carriers. Among these
compounds, ethanol has been identified as a promising hydrogen source
due to its high hydrogen to carbon (H/C) atomic ratio, easy handling,
and renewable nature (363).
There are two main processes for hydrogen production from ethanol,
i.e., partial oxidation and steam reforming. Partial oxidation process uses
oxygen to activate ethanol, and has a low reaction temperature but suffering
from low hydrogen production (364,365). In contrast, steam reforming is a
cost-effective process with higher amounts of hydrogen production, which
has been extensively investigated due to the potential application in fuel cells
(23,366). Steam reforming is a strongly endothermic reaction, which pro-
duces only H2 and CO2 with an increase of the whole moles number.
According to the thermodynamic and kinetic equilibriums, the reaction is
preferable to be happened at high temperatures, low pressures and at high
water-to-ethanol ratios. In general, the reaction is suggested to be operated
at above 650 K, atmospheric pressure and a water-to-ethanol molar ratio
higher than 10 to maximize the hydrogen production and also minimize
the CO, CH4 and carbon deposition formation (367). At temperatures
higher than 1073 K, the H2 selectivity is nearly 100%. However, the high
water-to-ethanol ratio and high reaction temperatures cause high energy
consumptions for this process.
The production of H2 from ethanol involves several different reaction
networks. Ideal reaction pathway includes ethanol dehydrogenation to acet-
aldehyde, followed by the CdC bond cleavage to produce CO and CH4,
which is then reformed to hydrogen. Apart from the main reaction, numbers
of reactions occur under reaction conditions, as shown in Table 5, which
causes the production of undesirable CO, CH4 and coke. In recent years,
great efforts have been dedicated to surpass the side reactions, develop
coke-resistant catalysts and minimize the energy consumption of this pro-
cess. Due to the rapid development in this area, we cannot cover all the
Synthesis of ethanol and its catalytic conversion 151

Table 5 Typical reaction pathways of ethanol steam reforming (368–371).


ΔH298
Reactions Equation (kJ/mol) Remarks
Sufficient steam C2H5OH + 174 Highest hydrogen
supply 3H2O ! 2CO2 + 6H2 production
Insufficient steam C2H5OH + H2O ! 2CO + 4H2 256 Undesirable
supply products of CO
C2H5OH + 2H2 ! 2CH4 + H2O 157
Dehydrogenation C2H5OH ! CH3CHO + H2 68 Reaction step for
hydrogen
Dehydration C2H5OH ! C2H4 + H2O 45 One reason for
coking
Decomposition C2H5OH ! 1/2CO2 + 3/2CH4 74 Undesirable
products of CO
C2H5OH ! CO + CH4 + H2 49
and CH4
Decomposition CH3CHO ! C3H6O + CH4 – Undesirable
products of CH4
Reforming C3H6O + H2O ! CO2 + H2 – Process for
improve hydrogen
CH3CHO + H2O ! CO2 + H2 –
productivity
CH4 + H2O ! CO + 3H2 206.2
CO + H2O ! CO2 + H2 41.2
C + H2O ! CO2 + H2 –
Decarbonylation CH3CHO ! CO + CH4 – Less hydrogen
production
Methanation COX + H2 ! CH4 –
Coking 2CO ! CO2 + C 175.1 Possible way for
coking
C2H4 ! polymers ! coke

catalysts or engineering aspects, which had extensively addressed in some


excellent works (363,367,372). Herein, we mainly focus on the strategies
in improving catalyst activity and stability, and ways to minimize the energy
consumption.

[Link] Catalyst development


[Link].1 Noble metal catalysts Noble metals have been extensively
studied over oxide supports including Al2O3, MgO, ZrO2 and CeO2 for
ethanol steam reforming. Rh was found to have higher activity and
152 Jifeng Pang et al.

hydrogen selectivity in comparison to Pd, Pt, Ni and Ru under the identical


conditions (373). It has a unique role within the 4d transition metals for
CdC and CdH bonds dissociation. Nevertheless, these catalysts still suffer
from deactivations via either metal sintering or coking. Attempts have been
made to reduce these issues by improving the dispersion of metals and
changing the electronic property or structure of catalysts. For instance,
Campos-Skrobot et al. used K to modify the Rh/NaY zeolites (high surface
area, 440 m2/g), over which the ethanol conversion increased from 62% to
97% with 68% selectivity to H2 at 573 K and a water-to-ethanol molar ratio
above five (374). To minimize the coke formation, Coronel et al. loaded Rh
nanoparticles on La2O3–SiO2 supports with different contents of La2O3. The
Rh/La2O3-SiO2 catalyst with 15 wt% of La2O3 exhibited higher stability
albeit localized carbon depositions (25 h, 100% ethanol conversion, 60%
H2 selectivity) (375). González-Gil et al. developed an RhCeNi/Al2O3
multimetallic catalyst that was active and stable in the hydrogen production
at both bench and pilot-plant scales. The H2 yield was 95% with a steam to
ethanol ratio of 7 at 973 K. During the 78 h operation, a constant hydrogen
yield of 0.23 Nm3 h1 was reported (376).
Ru-based catalysts at relatively high Ru loading (2 wt% or larger) also
exhibit high hydrogen selectivity, comparable to that of Rh (373). But, high
levels of ethylene and polymerized coke are produced (377). In order to solve
this problem, Carbajal-Ramos et al. used the Ce0.8Zr0.2O2 support to load
Ru by a co-precipitation method. The Ce0.8Zr0.2O2 support improved the
dispersion of Ru, which had better metal-support interactions. The catalyst
showed a high activity for the water gas shift reaction and an inhibition effect
on the ethylene formation, achieving the high H2 selectivity (90%) and neg-
ligible carbon deposition on the catalyst (378,379).
Recently, some noble metals were selected as model catalysts for the
mechanistic or kinetic study. Choong et al. investigated the ethanol steam
reforming reaction using in situ DRIFTS over Pt (Rh, Pd)/Ca-Al2O3 cat-
alysts. Formate species were exclusively observed, which was regarded as an
intermediate in this reaction (380). Wang et al. carried out kinetic study of
ethanol steam reforming over an Ir/CeO2 catalyst. The apparent energy for
this reaction was calculated to be 58 kJ/mol and the reaction orders with
respect to ethanol, H2O and H2 were 0.6, 0.5 and 0.9, respectively.
Accordingly, a bi-functional reaction mechanism was proposed, i.e., the
ethanol and water were activated on the ceria surface and the CdC bond
breaking of the C2 intermediates happened on the noble metal/ceria inter-
face. In detail, the ethoxy species derived from ethanol were activated and
Synthesis of ethanol and its catalytic conversion 153

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

C atom O atom H atom

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

Stage 1 Stage 2 Stage 3

0
Support Ni particle Filamentous coke Ni-carbide Non filamentous coke

Fig. 20 Deactivation stages of the Ni/La2O3–Al2O3 catalyst during ethanol steam


reforming (385). Reprinted (adapted) with permission from Montero, C.; Ochoa, A.;
Castaño, P.; Bilbao, J.; Gayubo, A. G. J. Catal. 2015, 331, 181–192. Copyright (2015) Elsevier.

mechanisms of Ni on the support of La2O3-α-Al2O3 in a fluidized bed reac-


tor (385). As shown in Fig. 20, there are three reaction stages in the deacti-
vation process. At the first reaction stage, the filamentous coke is preferable
produced, which drags Ni particles to the end of the fibers. Under this con-
dition, the performance is relatively steady due to the accessible of Ni par-
ticles. Then, the nonfilamentous coke becomes significant with the
decreasing yield of filamentous coke at the second reaction stage, which
blocks the Ni sites and causes the severe drop in ethanol conversion. Finally,
the nonfilamentous coke continuously blocks the Ni sites at the third reac-
tion stage. However, the deactivation rate decreases due to the protection of
the unreacted ethanol in the stream.
Co is another metal that exhibited high activity for CdC bonds scission,
and it has been investigated in ethanol steam reforming, as summarized by
Ozkan’s group (386). The Co particle size played an important role in deter-
mining the activity of ethanol steam reforming over Co/CNTs catalysts. As
the Co particle sizes decreased to be <3 nm, both the surface-specific activ-
ity and stability of catalyst were improved, which was attributed to the
increased unsaturated cobalt surface atoms and the lower fraction of terrace
sites (387). Similar conclusions were also obtained over the Ce modified
Co/MgAl2O4 catalysts. The addition of 7.8% Ce increased the dispersion
and reducibility of Co species, which resulted in a high H2 production
(100% ethanol conversion with 5.2 mol H2/mol ethanol) (388). G€ und€uz
and Dogu incorporated Mg and/or Co into the framework of mesoporous
alumina. The obtained Co@Mg-mesoporous alumina demonstrated the
higher hydrogen yield of 5.2 per 1 mole of ethanol due to the suitable
Lewis acid sites and balanced Co0 and CoO phases within the framework
(389). The activity and stability of Co catalysts were also highly dependent
on the supported materials. Among the oxides of ZrO2, TiO2, CeO2,
Synthesis of ethanol and its catalytic conversion 155

MgAl2O4 and γ-Al2O3, CeO2 was found to be an excellent support for


this reaction due to its abundant oxygen vacancies that facilitate oxygen
mobility and storage (390). Turczyniak et al. used the high-surface-area
Co/CeO2 catalysts to study the reaction mechanism by synchrotron-based
XPS, absorption spectroscopy and on-line mass spectrometry at different
pressures. Metallic Co and partially reduced cerium oxide were detected
at low pressures, which promoted the CdC bonds cleavage, and favored
the CO and H2 production (391). Different from the previous reaction
mechanism via the CdC bonds cleavage of acetaldehyde, Wang’s group
found that acetaldehyde was first converted to acetone over CoO and base
sites (CoII–ZrO2 catalysts), then the acetone was transformed to hydrogen
over the metallic Co sites (392). Recently, the mechanisms of ethanol
steam reforming over the Co13/CeO2x model catalysts was revealed by
DFT calculations. Different Co species and Ce oxides played special roles
in this reaction. Specifically, ethanol was first dehydrogenated on the Cox+
site to CH3CHO, which reacted with the hydroxyl that derived from the
water dissociation on CeO2 x, and formed acetic acid on the interface.
Then, the acetic acid moved to the Co0 site for CdC bonds cleavages
to CO2. Finally, the H2 was produced on the Cox+ site (393).
Due to the high activity of Ni and Co, they are used together for eth-
anol steam reforming, and a remarkable synergistic effect was observed in
most cases. Szijjártó et al. co-loaded Ni and Co on the MgAl2O4 support,
and obtained a higher hydrogen selectivity. They proposed that the defect
sites of the electron-rich support and the alloy particles promoted the initial
dehydrogenation of ethanol to acetaldehyde, and the left Ni species cata-
lyzed the following steam reforming of acetaldehyde (394). Moretti et al.
also compared the catalytic performance of Ni, Co and NiCo alloy over
mixed Ce-Zr oxides, and a significant synergetic effect was identified
between Co and Ni species. The best performance in terms of ethanol con-
version (95%) and hydrogen selectivity (>80%) was achieved over the
bimetallic NiCo/Ce-Zr(9:1) catalysts at 873 K (395).
In spite of these progresses, the stability of non-noble metals still needs to
be improved (372). The first and mostly used method is alloying the non-
noble metals. The introduction of very small amounts of noble metal such as
Rh, Pt and Ru increases the stability and activity of the non-noble metals
because of its strong interactions and synergistic effects. For instance,
González-Gil et al. incorporated Rh to the Ni/Al2O3 catalyst for ethanol
stream reforming. The Rh inhibited the carbon filament formation with
negligible intermediate products, and resulted in the improved stability of
156 Jifeng Pang et al.

Ni particles in the 78 h TOS (376,396). Similarly, a secondary metal such as


Fe, Cu, and Sn, also improved the catalytic activity and stability of Ni via the
formation of nanoalloys (397). The other way to stabilize the metals is
through the strong interaction between metal and supports. Researchers
used the HT-like compounds to prepare the Co-based catalysts, which
had excellent anti-sintering properties because of the strong metal–support
interactions (398). Recently, spatially confined methods have been used to
stabilize metals under fierce reaction conditions, which restrict metal parti-
cles by a well-defined cavities or channels, and provide a special micro-
environment to avoid the particles sintering (399). By studying chemical
structure of Co-HT catalysts, Huck-Iriart et al. found that the high stability
of Co-HT catalysts should be attributed to the presence of Coδ+ active spe-
cies, which inhibited the formation of metallic Co, and avoided the catalyst
deactivation by carbon deposition (400).

[Link] New technologies used for hydrogen production


In the products of steam reforming, CO, CH4 and CO2 always coexist with
hydrogen, making the gas inappropriate to be used in fuel cells. Water gas
shift reaction, pressure swing adsorption and/or Pd membrane separation are
required to further improve the hydrogen purity for fuel-cell application.
Recently, some new technologies have been used for the steam reforming
of ethanol to produce pure hydrogen. The most typical technologies are the
membrane reactor and chemical looping.

[Link].1 Membrane reactor Membrane reactor technology inte-


grates the chemical reaction and the product separation, which not only
reduces the capital costs but also shifts the reaction equilibrium toward
desirable reactions. In the case of ethanol steam reforming reaction, in a
membrane reactor, both the hydrogen yield and hydrogen selectivity
are improved with a high purity hydrogen stream (401). Murmura et al.
employed a Pt/NiCeO2 catalyst with Pd membrane reactor for steam
reforming of ethanol. Due to the equilibrium shift effect, a high throughput
of 4.5 mol H2 per mol ethanol was produced at 753 K and a low steam to
ethanol ratio (402). Similarly, Hedayati et al. achieved higher hydrogen pro-
ductivity over Pd-Rh/CeO2 catalysts in a high pressure reactor containing
Pd–Ag separation membranes. It was found that the energy efficiency was
also improved at high pressures (i.e., 4–12 bar) with pure hydrogen being
produced for fuel cells (403). For the first time, Ma et al. conducted an inves-
tigation on a large-scale catalytic membrane reactor for hydrogen generation
Synthesis of ethanol and its catalytic conversion 157

over a Pd/Au/Pd/Au composite asymmetric membrane. The reactor dis-


played a H2 yield of 11 and 21 Nm3 m2 h1 bar0.5 at 623 and 723 K,
respectively. The simulation models were consistent with the experimental
results, which give good guidance to pilot-scale hydrogen production from
ethanol steam reforming by using membrane reactors (404).

[Link].2 Chemical looping Steam reforming via chemical looping is


also a highly efficient method for hydrogen production from ethanol. This
process occurs in a dual fluidized bed with air and/or stream fed to the reac-
tor containing oxygen carrier catalysts. Unlike traditional steam reforming
process, the coke can be oxidized to generate the heat, which will be stored
in the ceramic matrix. The endothermic reforming reaction coupled with
the exothermic oxidation reactions minimizes the energy consumption
(405,406). The oxygen carrier is the key issue for chemical looping, and
Ni-based carriers have gained great attention for ethanol reforming (407).
For instance, Wang et al. encapsulated CeO2 and NiO nanoparticles in
SBA-15, denoted by xCeNi/SBA-15, as the oxygen carrier. The catalyst
exhibited the highest ethanol conversion of >90.0% and >84.7% hydrogen
selectivity with superior anti-coking capacity (408). The same performance
in activity and stability were also achieved by Jiang et al. although they used a
different oxygen carrier, NiO/montmorillonite (409).

3.2.2 Conversion of ethanol to butanol


Table 6 depicts the typical properties of fuels additives including methanol,
ethanol, butanol and gasoline. Compared to methanol and ethanol, butanol
has the advantages of high energy content, low volatility, high boiling point,
similar octane number, air/fuel ratio and adiabatic flame temperature similar
to that of gasoline. Moreover, butanol can be blended with gasoline at any
ratios. It is noncorrosive and immiscible with water at most conditions. All
these features make butanol a potential fuel additive for partially replacement
of gasoline. The performance of butanol blended fuels in different engines
was studied. Interestingly, it was found that butanol can be used safely and
favorably at high blending ratios with both gasoline and diesel, from the
viewpoints of thermal efficiency and exhaust emissions (411–413).
Although butanol is an advanced fuel additive, its production from
renewable biomass is still a challenge. Butanol can be produced from biomass
via the fermentation process, which is developed by a French microbiologist
named Louis Pasteur, and called as the ABE process (414). The concentra-
tion of butanol produced by fermentation is limited to 1–3 v/v% due to the
158 Jifeng Pang et al.

Table 6 Typical properties of methanol, ethanol, butanol and gasoline (9,410).


Property fuels Methanol Ethanol Butanol Gasoline
Chemical formula CH3OH C2H5OH C4H9OH C6.75H12.99
(C4–C12)
Molecular weight (g/mol) 32 46.07 74.12 100–105
Energy density (MJ/L) 16 19.6 29.2 39–46
Density (g/m at 20 °C)
3
0.79 0.79 0.81 0.72–0.76
Solubility in water (g/L at 20 °C) Miscible 79 Immiscible Immiscible
Octane number (R + M)/2 98.6 99.1 89 86–94
Boiling temperature (°C) 64.5 78.3 118 25–275
Latent heat of vaporization 1178 904 716 380–500
(25 kJ/kg)
Auto-ignition temperature (°C) 465 422 343 257
Stoichiometric air/fuel ratio 6.4 9.0 11.2 14.7
Adiabatic flame temperature (K) 1890 2310 2340 2370
Limit of added volume (v%) 5–30 ca. 10 >30 –

self-inhibition of butanol, which seriously decreases the process efficiency


but increases the operation scale. Up to date, it is impossible to meet the fuel
market by the fermentation process. Given the increased availability, limited
blending volume in gasoline, and other inevitable drawbacks of ethanol
(415), it is highly desired to explore chemical routes to upgrade ethanol into
butanol. After decades of exploration, several kinds of catalysts have been
developed for the transformation. Generally, it can be classified to homoge-
neous catalysts, metal oxide catalysts, HAP catalysts and metal-based
catalysts.

[Link] Homogeneous catalysts


In 2009, Ishii et al. first developed the catalytic conversion of ethanol to
butanol over homogenous [Ir(acac)(cod)] (cod: 1,5-cyclooctadiene, acac:
acetylacetonate) catalysts. The selectivity to butanol reached 67% at 18%
conversion in the presence of a 1,4-bis(diphenylphosphino) butane ligand
as well as NaOEt and 1,7-octadiene additives. Longer-chain alcohols were
formed as side products because of the uncontrollable aldol condensations
(416). Afterwards, Wass et al. employed a new catalyst composed of Ru
Synthesis of ethanol and its catalytic conversion 159

complexes with bis(diphenylphosphanyl)methane and NaOEt for ethanol


conversion. After 4 h reaction, 90.0% selectivity to butanol was achieved
at 20.4% ethanol conversion and 423 K. The results demonstrated that
the high selectivity to butanol was due to the small bite angle ligands, which
confined the uncontrolled base-catalyzed aldol condensation of acetalde-
hyde (417). The same group also investigated a variety of ligands to couple
the Ru-based catalysts, and found that mixed donor phosphine-amine
ligands were highly active in the presence of NaOEt, which showed
90% selectivity to butanol at 31% ethanol conversion (418). Since water
is always involved in the ethanol production and storage processes, it is
highly attractive to develop water tolerant catalysts for this reaction. Xu
et al. developed an Ir-based catalyst with water soluble ligand L10 (bat-
hophenanthroline disulfonic acid, disodium salt) and KOH/NaOAc for eth-
anol conversion in water. The ethanol conversion reached 52% with 81%
butanol selectivity at 423 K after the 16 h reaction (419).
For most catalytic systems, high carbon alcohols (C6, C8, C10) are always
observed during ethanol conversion process due to the uncontrollable aldol
condensations. To improve the product selectivity, Jones et al. developed a
bifunctional Ir complex catalyst coupled with bulky Ni or Cu hydroxides for
ethanol conversion. It was confirmed that the Ir complex facilitated the
dehydrogenation of ethanol to acetaldehyde, and the sterically crowded
Ni and Cu hydroxides catalyzed the aldol coupling reaction of acetaldehyde
exclusively to crotonaldehyde, which prevented the formation of high car-
bon alcohols. Under optimized conditions of 423 K and 24 h, 37% ethanol
was converted with >99% selectivity to butanol (420). Recently, Fu et al.
developed a homogeneous non-noble-metal catalyst, i.e., a Mn-ligand cat-
alyst (a well-defined PNP and NNP pincer Mn catalyst), for upgrading eth-
anol to butanol. The catalysts exhibited the highest TON (114120), TOF
(3078 h1) with 92% selectivity to butanol at 11.2% ethanol conversion
under the conditions of 433 K for 12 h. Results demonstrated that the high
selectivity was attributed to the “N–H moiety” and the PNP pincer struc-
ture of the catalyst (421).
Besides ethanol self-coupling, conversion of ethanol with methanol to
isobutanol has also been attempted. Wingad et al. screened a variety of
Ru complex catalysts for methanol/ethanol mixtures conversion. The
ruthenium diphosphine complexes with small bite angle disphosphines
showed the best performance. It demonstrated >99% selectivity to
isobutanol at >75% conversion via a Guerbet-type mechanism (422). Liu
et al. used Ir-based catalysts immobilized on N functionalized carbon
160 Jifeng Pang et al.

materials in water for upgrading ethanol with methanol to isobutanol. The


isobutanol selectivity surpassed 90% with 49% ethanol conversion at 433 K
for 12 h, even in the presence of some typical biogenic impurities such as
glycerol, acetic acid, unreacted carbohydrates and other organics (423).

[Link] Metal oxide catalysts


One of the key step for butanol synthesis is the condensation reaction, which
is preferred to take place at basic conditions. Strong liquid bases are always
involved in the homogenous systems. Due to the environmental and corro-
sion problems, the liquid bases are gradually replaced by solid oxide catalysts
(424). In the past decades, various solid base and acid-base oxide catalysts
have been exploited for the reaction.
Coville et al. screened most of the conventional solid base catalysts
including MgO, CaO and BaO, and found that MgO was highly active
and selectivity for this reaction. Over the MgO catalyst, the butanol yield
reached 18.4% at 56.1% ethanol conversion. A great deal of by-products
(gaseous products, esters and ethers) were also observed, which was attrib-
uted to the harsh reaction conditions (>573 K) and thus uncontrolled con-
densation and decomposition reactions (425). Kozlowski and Davis
employed Na modified zirconia catalysts for ethanol conversion. Introduc-
tion of Na species decreased the density of acid sites and correspondingly
increased the density of base sites, which suppressed the side reactions of
ethanol dehydration. However, it also inhibited the condensation reaction
and produced large amounts of acetaldehyde (426). Isotopic transient analysis
suggests that the intrinsic TOF of this reaction is only 0.040 s1 on MgO, due
to the low ethanol dehydrogenation activity (427). To elucidate key inter-
mediates for ethanol conversion, Cavani et al. used MgO as a model catalyst
for the mechanism study, and drew a unifying picture of different routes that
lead to butanol, butadiene and ethylene, which provides a solid foundation
for the products selectivity control and mechanism understanding (316,428).
Mixed oxides of MgO and Al2O3 that derived from HT are the typical
oxide catalysts used for ethanol conversion. Iglesia et al. prepared a series of
MgyAlOx catalyst from HT for the condensation reaction. It was found that
the structure and surface properties of catalysts greatly affect the product
selectivity. Condensation of ethanol was favored on Mg-rich samples with
Lewis acid–strong Brønsted base pairs (429). Ordóñez et al. prepared three
different HT-derived mixed oxides with incorporated framework Fe3+ for
ethanol conversion. The substitution of Al3+ by Fe3+ largely decreased the
acid sites, which in turn suppressed ethanol dehydration and resulted in an
Synthesis of ethanol and its catalytic conversion 161

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).

[Link] HAP catalysts


HAP materials, typically calcium-HAP, are well-known minerals used in
the biological field. Recently, the interest in this material was extended
to the field of catalysis, particularly as bifunctional catalysts or supports.
The stoichiometry of calcium-HAP shows a molar Ca/P ratio of 1.67,
i.e., Ca10(PO4)6(OH)2. Interestingly, the number of acid and base sites in
HAP can be tuned through the substitution of calcium or phosphate ions,
which shows high potential for alcohols conversion (433,434).
Besides used in different condensation reactions, HAP has been tried in
ethanol upgrading (435,436). Tsuchida et al. tried various oxide catalysts,
and found nonstoichiometric HAP with a Ca/P molar ratio of 1.64 selec-
tively catalyzed ethanol to butanol at atmospheric pressure. At 573 K and
a contact time of 1.78 s, ethanol conversion and butanol selectivity were
14.7% and 76%, respectively, with high carbon alcohols as co-products
(437). By changing reaction conditions (>673 K) and feedstocks (different
alcohols including ethanol, 1-butanol, 1-hexanol, 2-ethyl-1-butanol,
1-octanol and 2-ethyl-1-hexanol), gasoline with a research octane number
of 99 was also obtained. The Guerbet reaction and the Lebedev reaction
were demonstrated to be responsible for the CdC formation, followed
by dehydrogenation and dehydration to aldehydes and olefins (438).
The same group also investigated the influence of Ca/P molar ratios on
butanol selectivity. It was found that the Ca/P ratio greatly affected the
distribution of acid and base sites, which subsequently influenced the buta-
nol selectivity. Meanwhile, the distribution of high carbon alcohols is
closely related to parallel reactions of ethanol dehydration, dehydrogena-
tion and the Lebedev reactions, which were surpassed by the medium base
sites of HAP (Ca/P ¼ 1.67), and resulted in the C4 alcohols selectivity to
70.7% with 19% ethanol conversion at 545 K (439).
162 Jifeng Pang et al.

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).

[Link] Metal-based catalysts


For the conversion of ethanol to butanol, ethanol activation is regarded as
the rate determining step because ethanol is particularly difficult to be
dehydrogenated under mild conditions. Hence, metals have been intro-
duced to promote the hydrogen transfer reaction, as shown in Table 7.
The initial oxide used to load metal is Al oxides due to its suitable acidic
and basic properties. Jiang et al. reported the coupling of ethanol to butanol
over Ni supported Al2O3 catalysts. The ethanol conversion and butanol
selectivity were 19.1% and 64.3%, respectively (445). Riittonen et al.
screened several metals supported on Al2O3 for the one-pot conversion
Table 7 Conversion of ethanol to butanol over different metal enhanced catalysts.
Ethanol Butanol Main
Catalyst Reactor Conditions conversion (%) selectivity (%) by-product Ref.
1
8%Ni/Al2O3 Fixed bed 673 K, LHSV: 0.67 h 19.1 64.3 Acetaldehyde (445)
20% Ni/Al2O3 Autoclave 523 K, 72 h 25 80 Diethyl ether (446)
1
1.8% Cu/Al2O3 Fixed bed 513 K, LHSV: 4.3 h 16 64 Ethyl acetate (447)
1
HTC-500 Fixed bed 513 K, LHSV: 4.3 h 14 69 GAS (447)
8% Ni/Al2O3 Fixed bed 523 K, WHSV: 6.4 h1 35 61.7 1-Hexanol (448)
1
3Cu1Ce/AC Fixed bed/Autoclave 523 K, LHSV: 4 mL (h g cat.) 46.2 41–55 1-Hexanol (449)
Cu5MgAl5O Autoclave 533 K, 5 h 9 80 Acetal C6 (450)
NiMgAlO Fixed bed 533 K WHSV: 3.2 h1 18.7 55.2 1-Hexanol (451)
NiCu–PMO Autoclave 593 K, 6 h 56 39.3% 1-Hexanol (452)
Pd5MgAlO Autoclave 473 K, 5 h 3.8 72.7 Acetaldehyde (453)
Co + NaHCO3 Autoclave 473 K, 3d <5 69 2-Ethyl (454)
butanol
8%Ni/Al2O3 + 5% Fixed bed 603 K, 120 atm – 50–60 Hexanol (455)
Pd-8%Fe/Al2O3
Cu/CeO2, scCO2 Fixed bed 463 K, LHSV: 1.97 h1 16 60 – (456)
UiO-66-encapsulated Fixed bed 523 K, 2 MPa 49.8 48.6 >C4 products (457)
nanopalladium
164 Jifeng Pang et al.

of ethanol to butanol. The selectivity to butanol were shown in the follow-


ing sequence: Ni > Pt > Au Rh > Ru ≫ Ag, where a 25% ethanol conver-
sion with 80% butanol selectivity was reached over the 20.7% Ni/Al2O3
catalyst at 523 K for 72 h (446). In 2015, they revisited the reaction in a fixed
bed reactor at a LHSV of 4.3 h1. In comparison with Cu/Al2O3 and
Co/Al2O3 catalysts, the Ni/Al2O3 demonstrated similar activity to previous
study, and achieved ca. 30% ethanol conversion and 60% selectivity to
butanol at 513 K and a LHSV of 4.3 h1. They proposed that the adjacent
acid-base sites over inverse spinel metal aluminate structure containing octa-
hedrally coordinated metal cations were crucial to butanol selectivity (447).
Ghaziaskar and Xu conducted the reaction under sub-/supercritical condi-
tions. The ethanol conversion was 35% with approximately 62% butanol
selectivity at 523 K under the pressure of 176 bar over 8% Ni/Al2O3 catalysts.
Meanwhile, about 20% selectivity to hexanol was obtained due to the sec-
ondary reaction of butanol (448). Dias et al. compared the catalytic conver-
sion of ethanol to butanol with the fermented butanol under different
sugarcane distillery. Over the Ni/Al2O3 catalyst, vapor-phase catalysis pres-
ented higher potential for industrial implementation than liquid-phase
catalysis, which was an economic performance similar to that of the current
ABE process (458,459). Additives were used to improve the catalyst prop-
erties and thus reactions. Miller et al. reported that La2O3 modified
Ni/Al2O3 catalysts reduced the formation of both ethyl acetate and diethyl
ether, which gave a total alcohol yield of 38% and a total high carbon alco-
hol selectivity exceeding 85%, much higher than that over the Ni/Al2O3
catalyst (<50% high carbon alcohol selectivity) (460). The mathematical
modeling of butanol production from ethanol was established by Riittonen
et al. over a Cu/Al2O3 catalyst. Five parallel reactions were identified when
acetaldehyde, ethyl acetate, diethyl ether, and diethoxyethane were consid-
ered as potential intermediates. A mathematical model was then constructed
to interpret the reaction results and predict the trends of products under
different reaction conditions (461).
Taking advantages of the unique properties of HT precursor, mixed
oxides of MgO and Al2O3 were used to load metals for this reaction.
Tanchoux et al. prepared a series of Cu–Mg–Al mixed oxide catalysts from
LDHs. At 533 K for 5 h, the ethanol conversion and butanol selectivity
reached 9% and 80%, respectively, with 1,1-dimethoxymethane as the only
by-product (450). They also synthesized various M–Mg–Al (M ¼ Pd, Ag,
Mn, Fe, Cu, Sm, Yb) mixed oxide catalysts. Medium and high strength base
sites were found to determine the catalytic performances of the catalysts, and
Synthesis of ethanol and its catalytic conversion 165

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.

The characterization results revel that metal nanoparticles or alloys as well as


the existence of balanced acid–base sites played the key role in the excellent
catalytic performance (452).
Other strategies have also been proposed for ethanol conversion with
considerable butanol selectivity. Zhang et al. used commercial Co metal
powder for ethanol conversion at 473 K in the presence of NaHCO3.
The butanol selectivity reached 69%, but it took 3 days to achieve ca. 4%
ethanol conversion (454). Dziugan et al. designed a two-zone reactor system
for ethanol conversion. Over 8%Ni/Al2O3 and 5%Pd-8%Fe/Al2O3 cata-
lysts, 13% (v/v) of the butanol was obtained even under moderate conditions
(80 atm and 563 K), and this number increased to 31.4% (v/v) by distillation
of the reaction mixture and recycling of the low-boiling fraction back into
the reactor (455). Earley et al. conducted the ethanol reaction under scCO2
conditions over different Cu/CeO2 catalysts. The presence of CO2 has a
positive effect on ethanol conversion due to the redox cycle of Ce2O3
and CeO2 (Ce2O3 + CO2 ! CeO2 + CO). Meanwhile, Cu particles on high
surface area CeO2 showed the best activity for butanol formation as com-
pared with other oxide supports including Al2O3, ZSM-5, TiO2, and Si/Al
oxides (456). Large surface area supports were also used to improve metal
and oxide dispersions for this reaction. Jiang et al. used AC to load Cu
and CeO2 species. This resulted in a highly dispersed Cu and Ce sites
and thus benefited ethanol conversion and aldol condensation reactions,
leading to a high butanol yield of ca. 20% at 523 K in batch and fixed-
bed reactors (449). Afterwards, they developed the UiO-66-encapsulated
nanopalladium catalysts for ethanol conversion. Due to the suitable Lewis
acid that derived from the coordinatively unsaturated Zr sites and the
enhanced dehydrogenation activity of Pd, 49.8% of ethanol conversion
and 48.6% of selectivity toward butanol was achieved at 523 K and
2 MPa pressuer in the 200 h TOS (457).

3.2.3 Conversion of ethanol to hydrocarbons


Another important alternative approach for the upgrading of ethanol is the
conversion of ethanol into fuel range hydrocarbons, i.e., C5+ hydrocarbons
that have similar properties to the fossil fuels. The process is very similar to
the MTG process, which generally operates at temperature >623 K and
co-produces C1–C4 products (462).
Zeolites are the mostly used catalysts for the conversion of ethanol to
hydrocarbons. In the past years, great efforts have been dedicated on catalysts
improvement, modification and mechanistic studies (463). For H-ZSM-5
Synthesis of ethanol and its catalytic conversion 167

catalysts, it has a similar reaction mechanism to form hydrocarbons from


methanol and ethanol, but it is more pronounced to be deactivated by coke
deposition for ethanol conversion than the MTG process (20,464,465). The
kinetics study by Gayubo et al. revealed that the coke forming precursors are
ethylene and olefins, which undergo oligomerization and condensation to
coke. By treated HZSM-5 with the alkali and/or increased the water con-
tent in ethanol, the concentration of intermediates decreased to attenuate
the catalysts deactivation (466). Van der Borght et al. used Ni, Ga and Fe
to modify ZSM-5 catalysts. A positive effect for C3+ hydrocarbons was
observed when the metal content was <1 wt% due to an increase in the
amount of acid sites (467). Narula et al. developed the bimetallic InV-
ZSM-5 catalyst for ethanol conversion, which produced more C3+ products
than H-ZSM-5 catalysts did in the absence of hydrogen (468). Differently,
Hamieh et al. introduced hydroquinone to scavenge radicals during the coke
formation, which improved the stability of ZSM-5 to some extent (469).
Researchers also tried other zeolites with different pore structures for ethanol
conversion. Kittikarnchanaporn and Jitkarnka studied large pore zeolites of
HY and HBeta for this reaction, which produced heavier oils, especially
C6–C8 hydrocarbons and aromatic compounds (470). Chistyakov et al.
supported Pd-Zn on pilot MFI/alumina system for converting ethanol to
hydrocarbons. The catalyst was very stable with high selectivity to hydrocar-
bons because of the robust activity of the Pd-Zn alloy (471).
To understand the mechanism of catalyst deactivation, conditional exper-
iments and some novel characterizations were employed for this reaction.
Ramasamy et al. co-fed different oxygenates such as ethyl acetate, acetic acid
and acetaldehyde with ethanol. The presence of reactive acetaldehyde
enhanced the formation of high molecular weight aromatic compounds,
which deactivated the catalysts via the pore-blocking mechanism. In contrast,
acetic acid deactivated the catalysts via the active-site poisoning mechanism
(strong adsorption of acetate ion) (472). Based on experimental and compu-
tational investigations, D€aumer et al. proposed that aromatic compounds
with long linear/branched side-chains blocked the pore, and changed the
selectivity from hydrocarbons to C2 products during the reaction processes
(473). Differently, Pinard et al. demonstrated that coke played different role
in the catalyst deactivation. Some coke in the zeolite channels blocked the
pores and others formed radicals promoting ethylene polymerizations
(474). Recently, Van der Borght et al. proposed a two consecutive stage reac-
tion mechanism for the conversion of ethanol to hydrocarbons. As shown in
Fig. 22, in the first stage, ethanol was dehydrated to ethylene directly or
168 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

Fig. 22 Proposed reaction mechanisms for the conversion of ethanol to hydrocarbons


(475). Reprinted (adapted) with permission from 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. Copyright (2016) Wiley-VCH.

through a bimolecular pathway with diethyl ether as the intermediate. In the


second stage, there were three different consecutive pathways for hydrocar-
bons production, i.e., butane production from ethylene dimerization, pro-
pylene, butene and aromatics production via aliphatic and aromatic surface
intermediates, respectively (475).
Although there have been significant achievements in conversion of
ethanol to hydrocarbons, it is still a challenge to control the product distri-
butions. Hydrocarbons ranging from C1 to C5+, including olefins, paraffins
and aromatics are generally co-produced. Therefore, high-selectivity cata-
lysts need to be explored, so that the energy consumption for products
separation can be saved. Additionally, the deactivation mechanism of
zeolites and the strategies to lessen the deactivation process still need to
be further investigated (476).

4. Conclusions and outlook


In the past decades, great attention has been paid to the ethanol pro-
duction and applications due to its unique roles in bridging the biomass con-
version and modern fuel and chemical markets. The production of ethanol
Synthesis of ethanol and its catalytic conversion 169

increased dramatically with more economic and environmentally benign


ways. Ethanol has been regarded as one of the most important bulk chemical
derived from biomass, and the “ethanol economy” also boosts its down-
stream applications including directly used as the fuel additive or upgrading
to value added chemicals or fuels.
After thousands years’ evolution and the recent government stimula-
tions, the fermentation process has been commercialized in a large scale,
and continues to be the major method for ethanol production. The first gen-
eration ethanol employs the low polymerized sugars as feedstocks, which
causes the food versus fuel debates. Therefore, more abundant non-food lig-
nocellulosic biomass is being used to produce the second generation ethanol.
Although challenges like pretreatment, impurities, inhibitor influences, and
co-fermentation of C5–C6 sugars exist, the lignocellulosic biomass is still
being considered as the most promising natural resource for ethanol produc-
tion. Different from the biomass-based ethanol, the ethanol derived from
ethylene had been vanished for years due to its poor economic competitive-
ness in most areas. On the other hand, several other routes to ethanol have
been developed owing to the rapid development of C1 chemistry. The first
one is the direct conversion of syngas to high carbon alcohols over hetero-
geneous catalysts. Numerous catalysts, including Rh-based catalysts, modi-
fied F-T and methanol synthesis catalysts, and Mo-based catalysts, have been
exploited for ethanol production. So far, the activity and selectivity of the
catalysts to ethanol is still limited due to the uncontrollable formation of a
wide range of alcohols. To fulfill the commercial application, more robust
catalysts with high selectivity to ethanol production should be developed
based on the molecular-level understanding of the reaction mechanism.
Syngas fermentation process has the advantage of flexible CO-rich gas
resources. Pilot plant scale operations have been tested to evaluate the fea-
sibility of this process. The main challenge lies in the low reaction efficiency
because of the limited contact interface between gas and enzymes. In the past
several years, conversion of syngas to ethanol via methanol carbonylation
and acetate hydrogenation has made great progresses. On the basis of com-
mercialized methanol carbonylation, some efforts have been dedicated to
replace the homogenous catalysts with the heterogeneous ones given the
similar reaction mechanisms. Some catalysts exhibited comparable activity
to commercial catalysts, but more works such as the catalysts stability
improvement, low price catalysts development need to be done before its
applications. Meanwhile, a breakthrough in the conversion of DME to eth-
anol has been achieved using halide-free heterogeneous catalysts. By using a
170 Jifeng Pang et al.

consecutive reaction over catalysts of zeolites and Cu-based catalysts, etha-


nol was obtained with high selectivity during the long time run process. It
offers a new commercially potential process for ethanol production. During
the investigation of converting syngas to EG, a new way for ethanol produc-
tion was proposed by employing dimethyl oxalate as the intermediate.
DMO, EG and ethanol could be selectively produced depending on the
choice of catalysts and the reaction conditions. It provides a versatile way
for multiple products production. However, the ethanol production has a
lower atom economy as compared to that of EG. Besides, conversion of car-
bon dioxide into ethanol has been hotly investigated, which will provide a
good way for the energy transformation (477,478). Recently, catalytic con-
version of biomass to ethanol has been emerged, providing a high potential
way for efficient carbon neutral ethanol production.
Ethanol is primarily used as a fuel additive, which accounts for more than
80% of ethanol production. Additionally, it could also serve as a platform
molecule for producing valuable chemicals and advanced fuels. Ethanol
can be selectively dehydrogenated to acetaldehyde, which is a key interme-
diate in the production of butanol, propylene and 1,3-BD. For ethanol to
acetaldehyde conversion, non-oxidative dehydrogenation and oxidative
dehydrogenation processes have been developed. The former one is atom-
ically more economic since hydrogen is produced as a useful product.
Cu-based catalysts exhibited high activity but suffering from quick deac-
tivation. The latter operated at lower temperatures but sacrificed the
hydrogen, and Au-based catalysts demonstrated the high potential for acet-
aldehyde production via the control of Au particles. Conversion of ethanol
to propylene is a rather complex reaction involving both CdC bonds for-
mation and cleavage, it is a challenge to get high yield of propylene. Up to
now, a maximum yield of ca. 60% to propylene was reported in the presence
of water and hydrogen. Moreover, the stability of catalysts is yet to be
improved. Both one-step and two-step methods were used for the synthesis
of 1,3-BD from ethanol. For the one step method, multiple active sites
should be coupled in one catalyst to accomplish the complicated cascade
reaction steps. For the mixed oxide catalysts, ethanol was activated at rela-
tively high temperatures due to the relatively low dehydrogenation activity,
which caused side reactions such as decomposition and dehydrogenations.
Introduction of metals significantly enhanced the ethanol dehydrogenation
reaction and decreased the reaction temperatures. However, the kinetic
parameters of these elementary reaction steps are difficult to be balanced.
Moreover, the acid and base sites of catalysts are hard to be controlled,
Synthesis of ethanol and its catalytic conversion 171

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.

production. Currently, the enzymatic and catalytic processes co-exist to


flourish the ethanol production, and their developments are highly
dependent on the technology evaluations and the availability of feed-
stock at the locations.
(2) Integrate novel technologies into the ethanol chain. Besides the new
routes and catalysts for ethanol production, some other novel technol-
ogies should be used in the ethanol chain to augment the economic fea-
sibility. Taking ethanol separation as an example, the prevailing method
for ethanol separation is distillation, which is an energy intensive process
because of the low ethanol concentration. On the other hand, some
membrane technologies provide competitive advantages over the con-
ventional separation techniques, which have been incorporated into the
ethanol production process. Even though currently the cost of mem-
branes makes the separation expensive, it still provides high potential
to drive down the capital and operation costs by improving the mem-
brane materials and combing the distillation-membrane processes (479).
Then, analytical grade ethanol will be obtained with improved eco-
nomic feasibility.
(3) Full conversion of the “plant cell.” In lignocellulosic biomass, hemicel-
lulose, cellulose and lignin co-exist to form the rigid plant cell. At pre-
sent, the second generation ethanol only liberates the C6 sugars from
biomass for ethanol fermentation, which enlarges the cost of sugars
and also causes inhibitors from hemicellulose. To make the process
more economic feasible, all the components in the biomass should
be used in the biorefinery process. For instance, the hemicellulose is
hydrolyzed into C5 sugars for xylitol production, while the lignin is
used as the resin additive or monomers, which will decrease the inhib-
itors and margin the price of ethanol production. Additionally, the full
utilization of biomass with “zero” waste will make the process more
eco-friendly (480).
(4) Technology and engineering. The rapidly development of fermenta-
tion and chemical technologies promote the ethanol production and
applications. In contrast, engineering is also indispensable to the current
biorefinery process, especially to the new developed technologies. For
instance, for the fermentation of syngas to ethanol, great achievements
have been made on enzyme selections. Nevertheless, for the large scale
application, more efforts should be done on the interaction between gas
and the enzymes, which is the determinant factor for this process.
Hence, chemical engineering is recommended to be involved to solve
Synthesis of ethanol and its catalytic conversion 173

the mess transfer problems. Moreover, most processes such as pretreat-


ments, separations and reactor designs are highly dependent on the
chemical engineering, which is a fantastic way to make the whole pro-
cess more energy saving and economic feasible.
(5) Study reaction mechanisms by in situ technologies. To understand the
reactions mechanisms in molecular level, key intermediates of these cat-
alytic cycles should be identified and characterized, which will guide
the catalysts design. During the last decades, novel techniques for cat-
alysts characterization have been rapidly developed. Some instruments
could even be used under reaction conditions including high pressures,
temperatures and different phases, which shed light on deeper insight
into the reaction mechanisms and catalysts designs. To date, transmis-
sion and attenuated total reflection infrared spectroscopy, X-ray absorp-
tion spectroscopy as well as modified Raman, UV–vis, EPR and
M€ ossbauer have been used in in situ/operando studies. Besides, great
efforts have been devoted to narrow the gaps (i.e., pressure gap)
between the ultrahigh vacuum environment and catalysis under work-
ing conditions in the aspects of XPS, STEM, which will uncover the
authentic structures of catalysts (481). Additionally, the blossom of
supercomputing stimulates DFT calculations in clusters or big mole-
cules level simulation, which will bridge the gap between reactions
and theories. Moreover, the achievements in DFT calculations will
guide the catalysts design in experiments, and vice versa.
(6) Selectively activation of certain bonds under complex reaction net-
works. Ethanol is the simplest alcohol with CdC, CdO, CdH
and OdH bonds, which can be used as a model molecule to study
the activation of different functional groups. Currently, it is still a chal-
lenge to control the product selectivity because of the complex reaction
networks. Different model catalysts should be designed for certain ele-
mentary step, which will simplify the reaction and provide a close rela-
tionship between catalysts and reaction results. For instance, during the
conversion of ethanol to 1,3-BD process, dehydrogenation, aldol con-
densation and MPV reactions are occurred separately over different cat-
alysts to minimize the uncontrollable side reactions. Then, the reaction
steps, reaction mechanisms and the interaction between structure of cat-
alysts and product distributions can be disclosed more clearly. By con-
trolling the kinetic of different steps, 1,3-BD productivity could be easily
controlled over binary or trinary bed reactors. According to these results,
multiple active sites catalysts can be then designed more efficiently.
174 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.

184. Santiago, M. A. N.; Sánchez-Castillo, M. A.; Cortright, R. D.; Dumesic, J. A. J. Catal.


2000, 193 (1), 16–28.
185. Wang, S.; Guo, W.; Wang, H.; Zhu, L.; Yin, S.; Qiu, K. New J. Chem. 2014, 38 (7),
2792–2800.
186. Huang, X.; Ma, M.; Miao, S.; Zheng, Y.; Chen, M.; Shen, W. Appl. Catal. A Gen.
2017, 531, 79–88.
187. Qin, H.; Guo, C.; Sun, C.; Zhang, J. J. Mol. Catal. A Chem. 2015, 409, 79–84.
188. Ye, C. L.; Guo, C. L.; Zhang, J. L. Fuel Process. Technol. 2016, 143, 219–224.
189. Zhu, Y. M.; Shi, X. W. L. Bull. Kor. Chem. Soc. 2014, 35 (1), 141–146.
190. Zhu, Y. M.; Shi, L. J. Ind. Eng. Chem. 2014, 20 (4), 2341–2347.
191. Di, W.; Cheng, J.; Tian, S.; Li, J.; Chen, J.; Sun, Q. Appl. Catal. A Gen. 2016, 510,
244–259.
192. Lu, Z.; Yin, H.; Wang, A.; Hu, J.; Xue, W.; Yin, H.; Liu, S. J. Ind. Eng. Chem. 2016,
37, 208–215.
193. Zhou, W.; Kang, J.; Cheng, K.; He, S.; Shi, J.; Zhou, C.; Zhang, Q.; Chen, J.;
Peng, L.; Chen, M.; Wang, Y. Angew. Chem. Int. Ed. 2018, 57 (37), 12012–12016.
194. Gao, X.; Zhao, Y.; Wang, S.; Yin, Y.; Wang, B.; Ma, X. Chem. Eng. Sci. 2011, 66 (15),
3513–3522.
195. Yamamoto, Y. Catal. Surv. Asia 2010, 14 (3), 103–110.
196. Yue, H.; Ma, X.; Gong, J. Acc. Chem. Res. 2014, 47 (5), 1483–1492.
197. Xu, G. H.; Ma, X.; He, F.; Chen, H. F. Ind. Eng. Chem. Res. 1995, 34 (7), 2379–2382.
198. Fenton, D. M.; Steinwand, P. J. J. Org. Chem. 1974, 39 (5), 701–704.
199. Uchiumi, S. I.; Ataka, K.; Matsuzaki, T. J. Organomet. Chem. 1999, 576 (1–2), 279–289.
200. Zhao, T. J.; Chen, D.; Dai, Y. C.; Yuan, W. K.; Holmen, A. Ind. Eng. Chem. Res.
2004, 43 (16), 4595–4601.
201. Peng, S. Y.; Xu, Z. N.; Chen, Q. S.; Wang, Z. Q.; Chen, Y.; Lv, D. M.; Lu, G.;
Guo, G. C. Catal. Sci. Technol. 2014, 4 (7), 1925–1930.
202. Lv, D. M.; Xu, Z. N.; Peng, S. Y.; Wang, Z. Q.; Chen, Q. S.; Chen, Y.; Guo, G. C.
Catal. Sci. Technol. 2015, 5 (6), 3333–3339.
203. Peng, S. Y.; Xu, Z. N.; Chen, Q.-S.; Chen, Y. M.; Sun, J.; Wang, Z. Q.; Wang, M. S.;
Guo, G. C. Chem. Commun. 2013, 49 (51), 5718–5720.
204. Xu, Z. N.; Sun, J.; Lin, C. S.; Jiang, X.-M.; Chen, Q. S.; Peng, S. Y.; Wang, M. S.;
Guo, G. C. ACS Catal. 2013, 3 (2), 118–122.
205. Wang, C.; Han, L.; Chen, P.; Zhao, G.; Liu, Y.; Lu, Y. J. Catal. 2016, 337, 145–156.
206. Zhao, X. G.; Lin, Q.; Xiao, W.-D. Appl. Catal. A Gen. 2005, 284 (1–2), 253–257.
207. Jin, E.; He, L.; Zhang, Y.; Richard, A. R.; Fan, M. RSC Adv. 2014, 4 (90),
48901–48904.
208. Lin, Q.; Ji, Y.; Jiang, Z. D.; Xiao, W. D. Ind. Eng. Chem. Res. 2007, 46 (24),
7950–7954.
209. Wang, S.; Zhang, X.; Zhao, Y.; Ge, Y.; Lv, J.; Wang, B.; Ma, X. Front. Chem. Sci. Eng.
2012, 6 (3), 259–269.
210. Dong, Y.; Shen, Y.; Zhao, Y.; Wang, S.; Ma, X. ChemCatChem 2015, 7 (16),
2460–2466.
211. Yue, H.; Zhao, Y.; Ma, X.; Gong, J. Chem. Soc. Rev. 2012, 41 (11), 4218–4244.
212. Pang, J.; Zheng, M.; Sun, R.; Wang, A.; Wang, X.; Zhang, T. Green Chem. 2016, 18,
342–359.
213. Liu, Y.; Ding, J.; Sun, J.; Zhang, J.; Bi, J.; Liu, K.; Kong, F.; Xiao, H.; Sun, Y.; Chen, J.
Chem. Commun. 2016, 52 (28), 5030–5032.
214. Gong, J.; Yue, H.; Zhao, Y.; Zhao, S.; Zhao, L.; Lv, J.; Wang, S.; Ma, X. J. Am. Chem.
Soc. 2012, 134 (34), 13922–13925.
215. Yue, H.; Zhao, Y.; Zhao, S.; Wang, B.; Ma, X.; Gong, J. Nat. Commun. 2013, 4, 2339.
Synthesis of ethanol and its catalytic conversion 181

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.

About the authors


Jifeng Pang received his PhD in 2012
from Dalian Institute of Chemical Physics,
Chinese Academy of Sciences, under the
supervision of Prof. Tao Zhang. Afterwards,
he joined Prof. Tao Zhang’s group as an
Assistant Professor. In 2017, he moved to
Washington State University as a visiting
scholar with Prof. Yong Wang for 6 months.
He has published >30 scientific papers, and
applied >20 Chinese and PCT patents. Cur-
rently he is an Associate Professor in Dalian
Institute of Chemical Physics. His research
mainly focuses on the catalytic conversion
of lignocellulosic biomass into glycols and
upgrading ethanol to value-added chemicals.

Mingyuan Zheng received his PhD in 2005


from Dalian Institute of Chemical Physics,
Chinese Academy of Sciences, under the
supervision of Prof. Tao Zhang. Afterwards,
he worked therein up to now, mainly focus-
ing on catalytic conversion of biomass to
polyols. He has published >70 papers in
peer-reviewed scientific journals, and filed
>100 Chinese and PCT patents. He was pro-
moted to full professor in 2018. Currently, he
is engaging in the study of biomass conver-
sion to chemicals, the renewable chemistry,
as well as the industrial applications.
Synthesis of ethanol and its catalytic conversion 191

Tao Zhang received his PhD in Physical


Chemistry from Dalian Institute of Chemical
Physics, Chinese Academy of Sciences, in
1989. He was promoted to full professor of
DICP in 1995, and appointed as the director
of DICP in the period of 2007–2017. He is
Academician of the Chinese Academy of
Sciences since 2013. In 2017, he was pro-
moted as the vice president of the Chinese
Academy of Sciences. Over the past decades,
Prof. Zhang has developed a great number of
nano and subnano metallic catalysts for
applications in energy conversion and envi-
ronmental control. He authored and coauthored >400 publications in
peer-reviewed scientific journals, and applied >200 patents. He is editor-
in-chief of Chinese Journal of Catalysis (since 2015) and is serving on the
advisory boards of Applied Catalysis B (since 2008), ChemPhysChem (since
2010), ACS Sustainable Chemistry &Engineering (since 2013), Industrial &
Engineering Chemistry Research (2014). Prof. Zhang has received many
research awards, including the Science and Technology Progress Award
of HLHL Foundation (2016), Distinguished Award of Chinese Academy
of Sciences (2010), Zhou Guang Zhao Foundation Award for Applied
Science (2009), Excellent Young Scientist Award of Chinese Catalysis
Society (2008), and National Award of Technology Invention (Second
Grade, 2008, 2006, 2005).

You might also like