You are on page 1of 19

catalysts 

Review 
Review
Olefins from Biomass Intermediates: A Review 
Olefins from Biomass Intermediates: A Review
Vasiliki Zacharopoulou 1 and Angeliki A. Lemonidou 1,2,* 
Vasiliki Zacharopoulou 1 and Angeliki A. Lemonidou 1,2, *
1  Department of Chemical Engineering, Aristotle University of Thessaloniki, University Campus, 
1
Department of Chemical Engineering, Aristotle University of Thessaloniki, University Campus,
Thessaloniki 54124, Greece; vzacharopoulou@auth.gr 
54124 Thessaloniki, Greece; vzacharopoulou@auth.gr
2   Chemical Process Engineering Research Institute (CERTH/CPERI), P.O. Box 60361 Thermi,   
2
Chemical Process Engineering Research Institute (CERTH/CPERI), P.O. Box 60361 Thermi,
Thessaloniki 57001, Greece 
57001 Thessaloniki, Greece
*  Correspondence: alemonidou@cheng.auth.gr; Tel.: +30‐2310‐996‐273 
* Correspondence: alemonidou@cheng.auth.gr; Tel.: +30-2310-996-273
Received: 16 November 2017; Accepted: 19 December 2017; Published:   
Received: 16 November 2017; Accepted: 19 December 2017; Published: 23 December 2017

Abstract:  Over
Abstract: Over  the
the  last
last  decade,
decade,  increasing
increasing  demand
demand  forfor  olefins
olefins  and
and  their
their  valuable
valuable  products
products  has
has 
prompted research on novel processes and technologies for their selective production. As olefins 
prompted research on novel processes and technologies for their selective production. As olefins
are predominately dependent on fossil resources, their production is limited by the finite reserves 
are predominately dependent on fossil resources, their production is limited by the finite reserves
and the associated economic and environmental concerns. The need for alternative routes for olefin 
and the associated economic and environmental concerns. The need for alternative routes for olefin
production is
production is imperative
imperative in in order
order to
to meet
meet the
the exceedingly
exceedingly highhigh demand,
demand, worldwide.
worldwide. Biomass
Biomass isis 
considered a promising alternative feedstock that can be converted into the valuable olefins, among 
considered a promising alternative feedstock that can be converted into the valuable olefins, among
other  chemicals 
other chemicals and and fuels.
fuels.  Through 
Through processes  such  as
processes such as  fermentation,
fermentation,  gasification,
gasification,  cracking
cracking  and
and 
deoxygenation,  biomass  derivatives  can  be  effectively  converted  into  C 2–C4  olefins.  This  short 
deoxygenation, biomass derivatives can be effectively converted into C2 –C4 olefins. This short review
review focuses on the conversion of biomass‐derived oxygenates into the most valuable olefins, e.g., 
focuses on the conversion of biomass-derived oxygenates into the most valuable olefins, e.g., ethylene,
ethylene, propylene, and butadiene. 
propylene, and butadiene.

Keywords: olefins; biomass; ethylene; propylene; butadiene; catalysis 
Keywords: olefins; biomass; ethylene; propylene; butadiene; catalysis
 

1. Introduction 
1. Introduction
The importance
The importance of –C44 olefins
of CC22–C olefins (i.e.,
(i.e., CC22H
H44,, C
C33H66, , butenes,  and  C
butenes, and C44H66) ) has  been highlighted
has been highlighted 
because of their numerous applications as key building blocks in the chemical industry, linked with 
because of their numerous applications as key building blocks in the chemical industry, linked with
the increasing
the increasing  needs 
needs of of 
thethe  expanding 
expanding global global  population 
population [1].  lower
[1]. These These olefins
lower are olefins  are  prevalent
the most the  most 
prevalent organic compounds, with the highest production volumes, worldwide, highly dependent 
organic compounds, with the highest production volumes, worldwide, highly dependent on crude oil
on crude oil and natural gas products [2]. It is estimated that 400 million tons of olefins are annually 
and natural gas products [2]. It is estimated that 400 million tons of olefins are annually produced,
produced, 
using using tons
one billion one as
billion  tons  as  feedstock,
hydrocarbon hydrocarbon  via feedstock, 
processes suchvia  processes  such  as 
as fluid-catalytic fluid‐catalytic 
cracking, steam
cracking, steam cracking, and dehydrogenation [3]. Almost 60% of the global feedstocks are used in 
cracking, and dehydrogenation [3]. Almost 60% of the global feedstocks are used in FCC units, and
FCC units, and approximately 40% in steam cracking processes. (Figure 1) Produced olefins can be 
approximately 40% in steam cracking processes. (Figure 1) Produced olefins can be used in a wide
used in a wide spectrum of high‐end applications such as packaging, construction, solvents, coatings, 
spectrum of high-end applications such as packaging, construction, solvents, coatings, and synthetic
and synthetic fibers [4]. 
fibers [4].

 
Figure 1. Olefin production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014, 
Figure 1. Olefin production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014,
WILEY-VCH Verlag GmbH & Co.WILEY‐VCH Verlag GmbH & Co.   

Catalysts 2018, 8, 2; doi: 10.3390/catal8010002    www.mdpi.com/journal/catalysts 
Catalysts 2018, 8, 2; doi:10.3390/catal8010002 www.mdpi.com/journal/catalysts
Catalysts 2018, 8, 2  2 of 18 

C2H4 constitutes the most predominant olefin in the global market and is primarily produced 
Catalysts 2018, 8, 2 2 of 19
via  naphtha  steam  cracking,  among  various  hydrocarbon  feedstocks,  as  well  as  through  ethane 
thermal cracking. Globally, 57% (Figure 2) of the C2H4 volume is produced via naphtha and gas oil 
C2 H4 constitutes
steam  cracking  and  38%  thethrough 
most predominant
ethane  and olefin
LPG in the global market
(Liquefied  and is primarily
Petroleum  produced
Gas)  steam  cracking. 
via naphtha steam cracking, among various hydrocarbon feedstocks, as well as through ethane
Naphtha is a liquid fraction obtained from petroleum refining processes, such as catalytic cracking 
thermal cracking. Globally, 57% (Figure 2) of the C2 H4 volume is produced via naphtha and gas
and hydrocracking. Depending on its origin, it contains variable amounts of paraffins, aromatic, and 
oil steam cracking and 38% through ethane and LPG (Liquefied Petroleum Gas) steam cracking.
olefinic compounds. The ratio of these components can indicate the process that the specific fraction 
Naphtha is a liquid fraction obtained from petroleum refining processes, such as catalytic cracking
can be used for the optimum results. At high temperatures (i.e., 650–750 °C), naphtha and gas oil can 
and hydrocracking. Depending on its origin, it contains variable amounts of paraffins, aromatic, and
yield 30 and 25 wt. % of C
olefinic compounds. The2ratio H4, respectively. In the case of ethane, added along with naphtha in the 
of these components can indicate the process that the specific fraction
feed can
stream,  yields  of  C H   can 
be used for the optimum results.
2 4 reach 
At80 
highwt.  %  [5].  Its  (i.e.,
temperatures eminent 
650–750 ◦ C), naphtha
industrial  uses and
cause  the 
gas oil world 
can
demand for C
yield 30 and 2H 4 to incessantly increase, as it can be used for significant applications, such as the 
25 wt. % of C2 H4 , respectively. In the case of ethane, added along with naphtha in the
feed stream, yields of C2 H4 can
production  of  intermediate  reach 80 wt.
chemicals,  % [5]. Its
mainly  in eminent industrial
the  industry  of uses cause the
plastics.  world
i.e.,  demand(e.g., 
polymers 
for C2 H4 to incessantly
poly‐ethylene),  increase, as it canhydroformylation, 
propionaldehyde—via  be used for significant applications,
vinyl  such as halogenation 
chloride—via  the production and 
of intermediate chemicals, mainly in the industry of plastics. i.e., polymers
de‐hydrohalogenation,  alpha‐olefins—via  oligomerization,  and  C2H4  oxide  and  acetaldehyde—via  (e.g., poly-ethylene),
propionaldehyde—via hydroformylation, vinyl chloride—via halogenation and de-hydrohalogenation,
oxidation [4,6,7]. In recent years, bio‐ethanol has been extensively studied as an alternate feedstock 
alpha-olefins—via oligomerization, and C2 H4 oxide and acetaldehyde—via oxidation [4,6,7]. In recent
for  C2H4  production [8]. Other  bio‐derived  compounds  such  as methanol and  dimethyl‐ether,  can 
years, bio-ethanol has been extensively studied as an alternate feedstock for C2 H4 production [8].
also be used as a feedstock for C 2H4, via Methanol to Olefins (MTO) and Dimethyl‐ether to Olefins 
Other bio-derived compounds such as methanol and dimethyl-ether, can also be used as a feedstock
(DMTO) processes [9,10]. Bio‐ethylene can also be produced via bio‐synthesis from various enzymes 
for C2 H4 , via Methanol to Olefins (MTO) and Dimethyl-ether to Olefins (DMTO) processes [9,10].
or microorganisms [11]. 
Bio-ethylene can also be produced via bio-synthesis from various enzymes or microorganisms [11].

 
Figure 2. C
Figure 2. C2H2H 4 production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014,
4 production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014, 
WILEY-VCH Verlag GmbH & Co. WILEY‐VCH Verlag GmbH & Co. 

C3HC 6 3is 
H6the  second 
is the secondmost 
most significant 
significantolefin,  conventionally produced
olefin, conventionally produced  viavia  steam 
steam cracking, 
cracking, as a as  a 
co-product, or through fluid catalytic cracking (FCC). (Figure
co‐product,  or  through  fluid  catalytic  cracking  (FCC).  (Figure  3) C H and gasoline production severely
2 4 3)  C2H4  and  gasoline  production 
affect C 3 H 6 production; recently, steam crackers switch to ethane feedstocks, suppressing concurrent
severely affect C3H6 production; recently, steam crackers switch to ethane feedstocks, suppressing 
production of
concurrent  production  C H
3 6 , while its demand outpaces existing steam and
of  C3H6,  while  its  demand  outpaces  existing fluid catalytic cracking
steam  and capacity [12].
fluid  catalytic 
C3 H6 is mainly used for the production of polypropylene, as well as for the synthesis of numerous
cracking capacity [12]. C3H6 is mainly used for the production of polypropylene, as well as for the 
platform chemicals (e.g., cumene, acrylonitrile, propylene-oxide) [13]. The importance of C3 H6 in
synthesis  of  numerous  platform  chemicals  (e.g.,  cumene,  acrylonitrile,  propylene‐oxide)  [13].  The 
the C3 value chain addresses the need for alternative processes; using the conventional technology,
importance 
C3 H6 canof  alsoC3Hbe6 produced
in  the  C3through
  value  chain  addresses 
C4 -C8 olefin the  or
cracking, need  for  alternative 
through processes; 
FCC under severe using  the 
conditions
conventional technology, C
in order to increase produced 3H6 can also be produced through C4‐C8 olefin cracking, or through FCC 
volume [12]. On-purpose production methods that include propane
under  severe  conditions 
dehydrogenation in  order 
(PDH) [14], to  increase 
olefin metathesis [15],produced 
or methanolvolume 
to olefins[12]. 
(MTO)On‐purpose  production 
[16,17], have recently
methods that include propane dehydrogenation (PDH) [14], olefin metathesis [15], or methanol to 
been implemented, as well as using other unconventional feedstocks [18], such as bio-alcohols and
olefins  (MTO) 
vegetable [16,17],  have  recently  been  implemented,  as  well  as  using  other  unconventional 
oils.
feedstocks [18], such as bio‐alcohols and vegetable oils. 

 
importance  of  C3H6  in  the  C3  value  chain  addresses  the  need  for  alternative  processes;  using  the 
conventional technology, C3H6 can also be produced through C4‐C8 olefin cracking, or through FCC 
under  severe  conditions  in  order  to  increase  produced  volume  [12].  On‐purpose  production 
methods that include propane dehydrogenation (PDH) [14], olefin metathesis [15], or methanol to 
olefins  (MTO)  [16,17],  have  recently  been  implemented,  as  well  as  using  other  unconventional 
Catalysts 2018, 8, 2 3 of 19
feedstocks [18], such as bio‐alcohols and vegetable oils. 

Catalysts 2018, 8, 2  3 of 18 
 
Figure 3. C
3. C3 H3H 6 production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014, 
Figure 6 production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014,
WILEY‐VCH Verlag GmbH & Co. (*XTP refers to propylene production from any kind of feedstock) 
WILEY-VCH Verlag GmbH & Co. (*XTP refers to propylene production from any kind of feedstock).

C4 olefins are mostly produced via fluid catalytic cracking, as well as through steam cracking 
C4 olefins are mostly produced via fluid catalytic cracking, as well as through steam cracking
(Figure  4).  Most  common  C4  olefins  are  C4H6,  isobutylene,  and  butenes;  C4H6  is  a  key  chemical, 
(Figure 4). Most common C4 olefins are C4 H6 , isobutylene, and butenes; C4 H6 is a key chemical,
currently  used  for  the  production  of  polymers  (i.e.,  rubbers),  butylenes  are  expended  in  the  fuel 
currently used for the production of polymers (i.e., rubbers), butylenes are expended in the fuel
industry for the production of blending components and octane enhancers, and n‐butenes are used 
industry for the production of blending components and octane enhancers, and n-butenes are used
as  co‐monomers  of  polyethylene  and  for  the  synthesis  of  higher  olefins.  C4H6  is  the  prevalent  C4 
as co-monomers of polyethylene and for the synthesis of higher olefins. C4 H6 is the prevalent C4
olefin,  conventionally  co‐produced  via  naphtha  and  gas  oil  cracking,  along  with  C3H6  and  C2H4, 
olefin, conventionally co-produced via naphtha and gas oil cracking, along with C3 H6 and C2 H4 ,
among others; C4H6 yield via cracking is substantially low, highlighting the need for more targeted 
among others; C4 H6 yield via cracking is substantially low, highlighting the need for more targeted
production methods [4]. Demand for C4H6 is expected to increase, as it can be used as a bio‐based 
production methods [4]. Demand for C4 H6 is expected to increase, as it can be used as a bio-based
feedstock for greener rubbers. Several bio‐based routes have based proposed for C4H6‐ production; 
feedstock for greener rubbers. Several bio-based routes have based proposed for C4 H6 - production;
bio‐ethanol constitutes the most promising feedstock for this purpose. 
bio-ethanol constitutes the most promising feedstock for this purpose.

 
Figure 4. C4 olefin production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014,
Figure 4. C4 olefin production methods using hydrocarbon feedstocks. Reproduced from [3]. 2014, 
WILEY-VCH Verlag GmbH & Co.
WILEY‐VCH Verlag GmbH & Co.   

Overall,
Overall, existing
existing  olefin
olefin production
production  isis highly dependent on
highly dependent  fossil resources that
on fossil resources  that are
are expended
expended 
into
into energy-consuming
energy‐consuming  processes, heavily
processes,  contributing
heavily  to the environmental
contributing  affliction. Continuously
to  the  environmental  affliction. 
increasing
Continuously demand has turned
increasing  the petrochemical
demand  industry
has  turned  the  towards optimization
petrochemical  of existing
industry  towards  processes in
optimization  of 
order to meet challenging capacities, while lowering the vast production costs. Recently,
existing processes in order to meet challenging capacities, while lowering the vast production costs.  investigation
of alternative feedstocks has been considerably studied in order to offer an attractive solution
Recently, investigation of alternative feedstocks has been considerably studied in order to offer an 
untrammeled from the limited crude oil reserves; coal, natural gas, and biomass are some of the
attractive solution untrammeled from the limited crude oil reserves; coal, natural gas, and biomass 
examples [2,4]. Coal has been used in coal-rich countries as a feedstock for chemicals’ production;
are some of the examples [2,4]. Coal has been used in coal‐rich countries as a feedstock for chemicals’ 
even though it reduces dependence on fossil resources, the resulting CO2 emissions limit
production; even though it reduces dependence on fossil resources, the resulting CO the extent of
2 emissions limit 
its applications [4]. Methane prices have recently dropped due to technological advances,
the  extent  of  its  applications  [4].  Methane  prices  have  recently  dropped  due  to  technological  enabling
the use of shale gas as an attractive, economical feedstock. Thus, cost-effective olefin production,
advances, enabling the use of shale gas as an attractive, economical feedstock. Thus, cost‐effective 
via steam cracking, has been enabled, already implemented in the olefin market, primarily for the
olefin production, via steam cracking, has been enabled, already implemented in the olefin market, 
production of C3 H6 and higher olefins
primarily for the production of C [12,19,20]. Renewable feedstocks offer great advantages in terms
3H6 and higher olefins [12,19,20]. Renewable feedstocks offer great 
of sustainability, energy consumption, environmental pollution, CO2 emissions, and cost;2 emissions, 
advantages in terms of sustainability, energy consumption, environmental pollution, CO biomass is
an
and abundant carbon-source
cost;  biomass  with potential
is  an  abundant  to replacewith 
carbon‐source  fossilpotential 
resourcesto [21]. Through
replace  processes
fossil  such
resources  as
[21]. 
fermentation, hydro-deoxygenation or gasification, light olefins can be produced
Through processes such as fermentation, hydro‐deoxygenation or gasification, light olefins can be from bio-feedstocks,
produced from bio‐feedstocks, or from bio‐intermediates (e.g., ethanol, butanol, naphtha, methanol, 
and propane) via dehydration, metathesis, and steam‐cracking, among others [22]. 
In  this  context,  the  alternative  production  of  olefins  from  biomass  intermediates  is  reviewed 
and  compared  to  the  conventional  processes,  with  emphasis  on  the  most  promising  chemical 
technologies for future applications that progress biomass valorization. 
Catalysts 2018, 8, 2 4 of 19

or from bio-intermediates (e.g., ethanol, butanol, naphtha, methanol, and propane) via dehydration,
metathesis, and steam-cracking, among others [22].
In this context, the alternative production of olefins from biomass intermediates is reviewed and
compared to the conventional processes, with emphasis on the most promising chemical technologies
for future applications that progress biomass valorization.

2. Olefins from Biomass


Catalysts 2018, 8, 2  4 of 18 
Following the incessantly increasing demand worldwide, technological advances have enabled
Catalysts 2018, 8, 2  4 of 18 
discussed  in  this 
the production ofchapter.  Overall, 
light olefins fromolefins 
various from  biomass  are  predominantly 
biomass-derived produced 
feedstocks, obtaining from mixtures
mainly biomass 
intermediates, and more specifically, from alcohols, diols, and other oxygenates. These intermediates 
discussed 
of in  this Specific
C2 –C4 olefins. chapter. processes
Overall,  olefins 
can be from  biomass 
selective to theare  predominantly 
production produced 
of a certain from 
product, asbiomass 
will be
are, in most cases, formed via fermentation, hydro‐deoxygenation, or gasification processes (Figure 5). 
intermediates, and more specifically, from alcohols, diols, and other oxygenates. These intermediates 
discussed in this chapter. Overall, olefins from biomass are predominantly produced from biomass
are, in most cases, formed via fermentation, hydro‐deoxygenation, or gasification processes (Figure 5). 
intermediates, and more specifically, from alcohols, diols, and other oxygenates. These intermediates
are, in most cases, formed via fermentation, hydro-deoxygenation, or gasification processes (Figure 5).

 
 
Figure 5. Biomass to olefins primary routes. 
Figure 5. Biomass to olefins primary routes.
Figure 5. Biomass to olefins primary routes. 
2.1. Ethylene (C
2.1. Ethylene (C22H H44) )
2.1. Ethylene (C
As mentioned 2H4) 
mentioned  above,  steam  cracking  (mainly 
As above, steam cracking (mainly usingusing 
naphthanaphtha  as  a  feedstock) 
as a feedstock) is the mostis extensively
the  most 
extensively 
As  used 
mentioned  process 
above,  for  C
steam 2Hcracking 
4  production  from 
(mainly  hydrocarbons. 
using  naphtha 
used process for C2 H4 production from hydrocarbons. Steam cracking operating conditions Steam 
as  a  cracking 
feedstock)  is  operating 
the  most 
require
conditions 
extensively 
vast require 
amountsused  vast 
process 
of energy, amounts 
thus for  of  energy, 
C2H4  production 
increasing thus 
the production increasing 
from  the 
cost,hydrocarbons. production  cost, 
Steam  cracking 
and heavily contributing and  heavily 
operating 
to environmental
contributing to environmental issues. Even though the installed technology for these processes has 
conditions  require  vast  amounts  of  energy,  thus  increasing  the 
issues. Even though the installed technology for these processes has already been modifiedproduction  cost,  and  heavily 
for
already  been  modified  for  increased  efficiency,  novel  production  methods 
contributing to environmental issues. Even though the installed technology for these processes has 
increased efficiency, novel production methods have been explored in order to reduce production have  been  explored  in 
order 
already 
cost to been 
and reduce 
to production 
modified 
substitute the cost 
for finite
increased and 
fossil to  substitute 
efficiency, 
resources. novel 
Apartthe  finite 
C2fossil 
production 
from resources. 
H4methods  have Apart 
conventional been  from  from
C2H
explored 
production 4 
in 
conventional 
order  to  production 
reduce  production  from  petrochemicals, 
cost  and  to  C 2H4  the 
substitute  can  be  effectively 
finite  fossil 
petrochemicals, C2 H4 can be effectively produced from renewable feedstocks, such as plants, produced 
resources.  from 
Apart  renewable 
from  C2H4 

feedstocks, such as plants, microorganisms, and bio‐alcohols (Figure 6). 
conventional  production 
microorganisms, from  petrochemicals, 
and bio-alcohols (Figure 6). C2H4  can  be  effectively  produced  from  renewable 
feedstocks, such as plants, microorganisms, and bio‐alcohols (Figure 6). 

 
Figure 6. Schematic chart of C22H
Figure 6. Schematic chart of C H44 production methods. 
production methods.  
Figure 6. Schematic chart of C2H4 production methods. 
C2H4  biosynthesis  is  a  natural  pathway,  as  it  constitutes  an  important  hormone  that  plants 
recognize 
C2H4  and  produce is 
biosynthesis  [11].  ACC  synthase 
a  natural  pathway, and 
as  oxidase  are  two 
it  constitutes  an  vital  enzymes 
important  that  enable 
hormone  C2H4 
that  plants 
production from ACC (1‐aminocyclopropane‐1carboxylic acid) and SAM (S‐adenosyl methionine), 
recognize  and  produce  [11].  ACC  synthase  and  oxidase  are  two  vital  enzymes  that  enable  C2H4 
along  with  carbon  dioxide  and  HCN,  via  the  Yang  cycle,  starting  from  methionine  via  three 
production from ACC (1‐aminocyclopropane‐1carboxylic acid) and SAM (S‐adenosyl methionine), 
consecutive  reaction  steps:  (a)  methionine  is  converted  into  SAM  by  SAM  synthetase;  (b)  ACC 
Catalysts 2018, 8, 2 5 of 19

C2 H4 biosynthesis is a natural pathway, as it constitutes an important hormone that plants


recognize and produce [11]. ACC synthase and oxidase are two vital enzymes that enable C2 H4
production from ACC (1-aminocyclopropane-1carboxylic acid) and SAM (S-adenosyl methionine),
along with carbon dioxide and HCN, via the Yang cycle, starting from methionine via three consecutive
reaction steps: (a) methionine is converted into SAM by SAM synthetase; (b) ACC synthetase
converts SAM to ACC; and finally, (c) ACC is converted into C2 H4 by ACC oxidase [23]. (Scheme 1)
Microorganisms, such as bacteria and fungi have also been reported to produce C2 H4 starting
from methionine via the KMBA (2-keto-4-methylthiobutyric acid) formation pathway or through
Catalystsȱ2018,ȱ8,ȱ2ȱ 5ȱofȱ18ȱ
2-oxoglutarate conversion [24–26]. Ethylene production rate can reach 2859.2 µmol/gCDW/h
ΐmol/gCDW/hȱ
(CDW-Cell (CDWȬCellȱ
Dry Weight) fromDryȱ Weight)ȱputida
Pseudomonas fromȱ [27].
Pseudomonasȱ putidaȱ
Despite the [27].ȱ
fact that Despiteȱ theȱtechnology
bio-synthesis factȱ thatȱ
bioȬsynthesisȱ technologyȱ isȱ atȱ itsȱ earlyȱ stages,ȱ recentȱ technoȬeconomicȱ analysesȱ
is at its early stages, recent techno-economic analyses highlight the potential of these processes highlightȱ theȱ
potentialȱofȱtheseȱprocessesȱforȱonȬpurposeȱC
for on-purpose C2 H4 production. However,2H 4ȱproduction.ȱHowever,ȱfurtherȱstudiesȱareȱrequiredȱ
further studies are required in order to reduce cost
inȱ
andorderȱ toȱ reduceȱ
overcome costȱ andȱ
or improve overcomeȱ
aspects such asorȱ improveȱ aspectsȱ
productivity suchȱ asȱ
and product productivityȱ
separation. andȱ productȱ
Advances in the
separation.ȱAdvancesȱinȱtheȱbiotechnologicalȱprocessesȱcouldȱimproveȱproductivityȱandȱreduceȱcost,ȱ
biotechnological processes could improve productivity and reduce cost, enabling future development
enablingȱfutureȱdevelopmentȱofȱC
of C2 H4 biosynthesis methods [28].2H4ȱbiosynthesisȱmethodsȱ[28].ȱ

ȱ
Schemeȱ1.ȱStepsȱforȱC
Scheme 1. Steps for C22H44ȱbioȬsynthesis.ȱ
bio-synthesis.

Ethanol,ȱ
Ethanol, derivedȱ
derived fromȱfrom biomass,ȱ
biomass, i.e.,ȱ
i.e., cellulose,ȱ
cellulose, corn,ȱ
corn, andȱ
and sugarcaneȱ
sugarcane [8]ȱ [8] isȱ
is theȱ mostȱ common
the most commonȱ
bioȬfeedstockȱforȱtheȱproductionȱofȱC
bio-feedstock for the production of 2H4.ȱSeveralȱcompaniesȱworldwide,ȱsuchȱasȱBraskemȱ(SaoȱPaulo,ȱ
C2 H4 . Several companies worldwide, such as Braskem
Brazil),ȱ Axensȱ (RueilȬMalmaison,ȱ
(Sao Paulo, Brazil), Axens (Rueil-Malmaison, France),ȱ Solvayȱ
France),(Brussels,ȱ Belgium),ȱBelgium),
Solvay (Brussels, andȱ BPȱ (London,ȱ Unitedȱ
and BP (London,
Kingdom),ȱfocusȱonȱresearchȱandȱplantȱoperationȱforȱethanolȱdehydrationȱintoȱC
United Kingdom), focus on research and plant operation for ethanol dehydration into C2 H4 [13–15], 2H4ȱ[13–15],ȱasȱtheȱ

globalȱ demandȱ
as the global demandforȱ Cfor
2H4ȱ isȱ continuouslyȱ increasingȱ [29,30].ȱ Evenȱ thoughȱ eachȱ companyȱ hasȱ
C2 H4 is continuously increasing [29,30]. Even though each company has
developedȱ andȱ installedȱ
developed and installed specific specificȱ technologiesȱ
technologies forȱreaction,
for this thisȱ reaction,ȱ
ethanolethanolȱ dehydrationȱ
dehydration typically typicallyȱ
includes
includesȱ twoȱ steps;ȱ reactionȱ ofȱ ethanolȱ dehydrationȱ andȱ purificationȱ ofȱ
two steps; reaction of ethanol dehydration and purification of products [29]. Bio-ethanol can be productsȱ [29].ȱ BioȬethanolȱ
canȱbeȱproducedȱviaȱfermentationȱprocessesȱcausingȱsugarsȱtoȱselectivelyȱbreakȱdownȱintoȱethanol,ȱ
produced via fermentation processes causing sugars to selectively break down into ethanol, at high
atȱ highȱ
rates, rates,ȱ
with withȱby-product
limited limitedȱ byȬproductȱ
formationformationȱ [31].ȱ However,ȱ
[31]. However, the strong theȱ strongȱ dependenceȱ
dependence onȱ sugarȱ
on sugar production
productionȱ costsȱ
costs limit their limitȱ theirȱ[30,32].
application applicationȱ
More [30,32].ȱ
complexMoreȱ complexȱ
feedstocks withfeedstocksȱ
higher marketwithȱavailability
higherȱ marketȱ
and
availabilityȱandȱthusȱlowerȱcostȱ(i.e.,ȱcellulose,ȱhemicellulose,ȱandȱlignocellulose)ȱareȱmoreȱattractiveȱ
thus lower cost (i.e., cellulose, hemicellulose, and lignocellulose) are more attractive but their direct
butȱtheirȱdirectȱconversionȱintoȱethanolȱisȱchallengingȱandȱhasȱnotȱbeenȱreportedȱtoȱbeȱsatisfactorilyȱ
conversion into ethanol is challenging and has not been reported to be satisfactorily viable for
viableȱforȱapplicationsȱinȱtheȱindustryȱ[33].ȱ
applications in the industry [33].
BioȬethanolȱ dehydrationȱ
Bio-ethanol dehydration is anisȱendothermic
anȱ endothermicȱ
reaction,reaction,ȱ requiringȱmoderate
requiring relatively relativelyȱ moderateȱ
temperatures
temperaturesȱ (i.e.,ȱ 180–500ȱ °C)ȱ andȱ theȱ presenceȱ ofȱ aȱ catalyst.ȱ Theȱ mechanismȱ
(i.e., 180–500 ◦ C) and the presence of a catalyst. The mechanism of bio-ethanol dehydration consists of ofȱ bioȬethanolȱ
dehydrationȱconsistsȱofȱtheȱfollowingȱsteps:ȱ(a)ȱprotonationȱofȱtheȱhydroxylȱgroupȱbyȱanȱacidȱcatalyst,ȱ
the following steps: (a) protonation of the hydroxyl group by an acid catalyst, (b) deprotonation of the
(b)ȱdeprotonationȱofȱtheȱmethylȱgroupȱbyȱtheȱconjugateȱbaseȱofȱtheȱcatalyst,ȱandȱ(c)ȱrearrangementȱtoȱ
methyl group by the conjugate base of the catalyst, and (c) rearrangement to form C2 H4 (Scheme 2) [30].
formȱC 2H4ȱ(Schemeȱ2)ȱ[30].ȱTheȱselectionȱofȱtheȱmostȱsuitableȱcatalystȱforȱbioȬethanolȱdehydrationȱ
The selection of the most suitable catalyst for bio-ethanol dehydration into C2 H4 is of key importance,
intoȱC 2H4ȱisȱofȱkeyȱimportance,ȱinȱorderȱtoȱlowerȱreactionȱtemperatureȱandȱovercomeȱcommonȱissues,ȱ
in order to lower reaction temperature and overcome common issues, such as catalyst deactivation
suchȱ
due toasȱcoke
catalystȱ deactivationȱ
formation, particledueȱ toȱ cokeȱ
collision, formation,ȱ
and particleȱ
agglomeration collision,ȱ
[29,30]. Acidandȱ agglomerationȱ
catalysts, [29,30].ȱ
such as zeolites
Acidȱcatalysts,ȱsuchȱasȱzeolitesȱandȱsilicoaluminophosphatesȱ(SAPO),ȱhaveȱbeenȱwidelyȱusedȱforȱthisȱ
and silicoaluminophosphates (SAPO), have been widely used for this reaction, over the last decades,
reaction,ȱoverȱtheȱlastȱdecades,ȱasȱtheyȱpromoteȱselectiveȱconversionȱintoȱtheȱdesiredȱproduct,ȱwithȱ
as they promote selective conversion into the desired product, with extremely high conversion and
extremelyȱhighȱconversionȱandȱselectivityȱvalues,ȱdespiteȱtheȱfactȱthatȱcatalystsȱneedȱtoȱbeȱfrequentlyȱ
selectivity values, despite the fact that catalysts need to be frequently regenerated [30,34]. Actually,
regeneratedȱ[30,34].ȱActually,ȱSAPOȱcatalystsȱareȱsignificantlyȱactiveȱinȱthisȱreaction,ȱreachingȱ98.0%ȱ
SAPO catalysts are significantly active in this reaction, reaching 98.0% selectivity to ethylene, at 250 ◦ C,
selectivityȱtoȱethylene,ȱatȱ250ȱ°C,ȱoverȱSAPOȬ11Ȭ4ȱandȱ98.4%,ȱatȱ340ȱ°C,ȱoverȱMnȬSAPOȬ34ȱ[35,36].ȱ
over SAPO-11-4 and 98.4%, at 340 ◦ C, over Mn-SAPO-34 [35,36]. However, over modified HZSM-5
However,ȱ overȱ modifiedȱ HZSMȬ5ȱ andȱ MCMȬ41ȱ catalysts,ȱ selectivityȱ toȱ C2H4ȱ isȱ evenȱ higherȱ (i.e.,ȱ
99.0%)ȱ[37].ȱMoreover,ȱaluminaȬbasedȱcatalysts,ȱsuchȱasȱAl2O3ȬMgO/SiO2,ȱexhibitedȱhighȱconversionȱ
andȱselectivityȱvaluesȱ(i.e.,ȱ99.0ȱandȱ97.0%,ȱrespectively)ȱrequiringȱregenerationȱoverȱlongerȱperiodsȱ
ofȱ operationȱ [29].ȱ TungstenȬbasedȱ heteropolyacids,ȱ supportedȱ onȱ variousȱ substancesȱ areȱ highlyȱ
selectiveȱ atȱ relativelyȱ lowȱ reactionȱ temperaturesȱ (i.e.,ȱ 180–250ȱ °C)ȱ [38].ȱ Itȱ hasȱ beenȱ provenȱ thatȱ
Catalysts 2018, 8, 2 6 of 19

and MCM-41 catalysts, selectivity to C2 H4 is even higher (i.e., 99.0%) [37]. Moreover, alumina-based
catalysts, such as Al2 O3 -MgO/SiO2 , exhibited high conversion and selectivity values (i.e., 99.0 and
97.0%, respectively) requiring regeneration over longer periods of operation [29]. Tungsten-based
heteropolyacids, supported on various substances are highly selective at relatively low reaction
temperatures (i.e., 180–250 ◦ C) [38]. It has been proven that acid-base sites on the catalyst surface are
linked with the product distribution of this reaction; ethanol dehydration most probably proceeds via
the formation of intermediate species and highly depends on temperature, ethanol partial pressure, and
the nature of acid-base sites [30,39–43]. In fact, as ethanol dehydration is an endothermic reaction, the
reaction temperature strongly affects C2 H4 yield. Basic catalysts, such as MgO or CaO have also been
used with results that resembled those, over acidic catalysts. Phosphoric acid, oxides, molecular sieves,
Catalystsȱ2018,ȱ8,ȱ2ȱ 6ȱofȱ18ȱ
and other heteropoly acid catalysts can also be used for this purpose [29]. Syndol is a commercial
canȱalsoȱbeȱusedȱforȱthisȱpurposeȱ[29].ȱSyndolȱisȱaȱcommercialȱcatalyst,ȱusedȱbyȱHalconȱSDȱ(USA),ȱtoȱ
catalyst, used by Halcon SD (USA), to obtain high ethanol conversion and selectivity to ethylene, at
obtainȱ highȱ
400–500 ethanolȱ conversionȱ
◦ C [29,44–46]. andȱ selectivityȱ
Overall, a selection toȱ ethylene,ȱ
of the most promising atȱcatalysts
400–500ȱfor °Cȱethanol
[29,ȱ 44–46].ȱ Overall,ȱ
conversion intoaȱ
selectionȱofȱtheȱmostȱpromisingȱcatalystsȱforȱethanolȱconversionȱintoȱC
C2 H4 are presented in Table 1. 2 H 4 ȱareȱpresentedȱinȱTableȱ1.ȱ

ȱ
Schemeȱ2.ȱMechanismȱofȱbioȬethanolȱdehydrationȱtoȱC
Scheme H44.ȱ.
2. Mechanism of bio-ethanol dehydration to C22H

Tableȱ1.ȱSelectedȱcatalystsȱforȱbioȬethanolȱdehydrationȱtoȱC
Table H44..ȱ
1. Selected catalysts for bio-ethanol dehydration to C22H
Ethanolȱ Selectivityȱtoȱ
Catalystȱ Ethanol Selectivity Temperatureȱ(°C) WHSVȱa/LHSVȱbȱ(hƺ1)ȱ Referenceȱ
Catalyst Conversionȱ(%)ȱ C2H4ȱ(%)ȱ to Temperature (◦ C) WHSV a /LHSV b (h−1 ) Reference
Conversion (%) C2 H4 (%)
MnȬSAPOȬ34ȱ 99.4ȱ 98.4ȱ 340ȱ 2.0ȱaȱ [35]ȱ
Mn-SAPO-34
0.5%LaȬ2%PȬHZSMȬ5ȱ 99.4
100.0ȱ 98.4
99.9ȱ 340
240–280ȱ 2.0aȱa
2.0ȱ [35]
[47]ȱ
0.5%La-2%P-HZSM-5 100.0 99.9 240–280 2.0 a [47]
TPAȬMCMȬ41ȱ*ȱ 98.0ȱ 99.9ȱ 300ȱ 2.9ȱaȱa [48]ȱ
TPA-MCM-41 * 98.0 99.9 300 2.9 b [48]
SynDolȱ**ȱ 99.0ȱ 96.8ȱ 450ȱ 26–234ȱ bȱ [29,ȱ44–46]ȱ
SynDol ** 99.0 96.8 450 26–234 [29,44–46]
STAȬMCMȬ41ȱ***ȱ 99.0ȱ 99.9ȱ 250ȱ 2.9ȱaȱa [49]ȱ
STA-MCM-41 *** 99.0 99.9 250 2.9 [49]
*ȱtungstophosphoricȱacidȱ(TPA),ȱ**ȱMgOȬAl2O3/SiO2,ȱ***ȱsilicotungsticȱacidȱ(STA).ȱ
* tungstophosphoric acid (TPA), ** MgO-Al2 O3 /SiO2 , *** silicotungstic acid (STA). a Weight hourly space velocity
(WHSV), b Liquid hourly space velocity (LHSV).
aȱWeightȱhourlyȱspaceȱvelocityȱ(WHSV),ȱ bȱLiquidȱhourlyȱspaceȱvelocityȱ(LHSV)ȱ

Methanolȱ isȱ anotherȱ alcoholȱ thatȱ canȱ beȱ convertedȱ intoȱ olefins,ȱ throughȱ theȱ wellȬknownȱ
Methanol is another alcohol that can be converted into olefins, through the well-known
methanolȬtoȬolefinsȱ (MTO)ȱ process.ȱ Theȱ MTOȱ reactionȱ isȱ oneȱ ofȱ theȱ mostȱ importantȱ processesȱ forȱ
methanol-to-olefins (MTO) process. The MTO reaction is one of the most important processes
producingȱ olefinsȱ fromȱ aȱ C1ȱ feedstockȱ [50].ȱ MTOȱ wasȱ initiallyȱ proposed,ȱ inȱ 1977,ȱ byȱ Mobilȱ
for producing olefins from a C1 feedstock [50]. MTO was initially proposed, in 1977, by Mobil
Corporationȱ [51],ȱ followedȱ byȱ numerousȱ researchȱ studies,ȱ focusingȱ onȱ theȱ developmentȱ ofȱ aȱ
Corporation [51], followed by numerous research studies, focusing on the development of a
commerciallyȱavailableȱtechnologyȱ[52,17].ȱWithinȱthisȱcontext,ȱtheȱfirstȱMTOȱplantȱwasȱinstalledȱinȱ
commercially available technology [17,52]. Within this context, the first MTO plant was installed
2010,ȱinȱChina,ȱforȱtheȱproductionȱofȱlightȱolefinsȱfromȱcoalȱ[53].ȱ
in 2010, in China, for the production of light olefins from coal [53].
BioȬmethanolȱ canȱ beȱ producedȱ viaȱ severalȱ processesȱ includingȱ pyrolysis,ȱ bioȬsynthesis,ȱ
Bio-methanol can be produced via several processes including pyrolysis, bio-synthesis,
gasification,ȱandȱelectrolysis,ȱusingȱaȱwideȱrangeȱofȱbiomassȱwaste,ȱsuchȱasȱagricultural,ȱforest,ȱandȱ
gasification, and electrolysis, using a wide range of biomass waste, such as agricultural, forest,
municipalȱ [17,ȱ 53–55].ȱ Althoughȱ mostȱ ofȱ theseȱ processesȱ areȱ currentlyȱ underȱ development,ȱ theȱ
and municipal [17,53–55]. Although most of these processes are currently under development, the
installationȱ ofȱ bioȬmethanolȱ productionȱ unitsȱ requiresȱ furtherȱ studiesȱ onȱ designȱ andȱ energyȱ
installation of bio-methanol production units requires further studies on design and energy efficiency
efficiencyȱinȱorderȱtoȱensureȱfeasibilityȱofȱlargeȬscaleȱapplicationȱ[56].ȱ
in order to ensure feasibility of large-scale application [56].
MTOȱisȱanȱautocatalyticȱreactionȱthatȱconventionallyȱtakesȱplaceȱatȱmoderateȱtemperatureȱ(i.e.,ȱ
MTO is an autocatalytic reaction that conventionally takes place at moderate temperature
300–450ȱ °C),ȱ◦overȱ acidicȱ catalysts.ȱ Aȱ numberȱ ofȱ suggestionsȱ onȱ theȱ mechanismȱ ofȱ MTOȱ hasȱ beenȱ
(i.e., 300–450 C), over acidic catalysts. A number of suggestions on the mechanism of MTO has
proposed,ȱshowingȱthatȱmostȱconclusionsȱagreeȱthatȱtheȱreactionȱnetworkȱconsistsȱofȱatȱleastȱthreeȱ
been proposed, showing that most conclusions agree that the reaction network consists of at least
mainȱ pathways:ȱ (a)ȱ directȱ methanolȱ conversion;ȱ (b)ȱ directȱ conversionȱ ofȱ ethylene;ȱ andȱ (c)ȱ ethaneȱ
three main pathways: (a) direct methanol conversion; (b) direct conversion of ethylene; and (c) ethane
methylationȱbyȱ methanolȱ[52].ȱ Dependingȱ onȱ theȱ catalystȱ used,ȱ MTOȱ canȱ selectivelyȱyieldȱ mainlyȱ
methylation by methanol [52]. Depending on the catalyst used, MTO can selectively yield mainly C2 H4
C2H4ȱ andȱ C3H6;ȱ mostȱ studiesȱ focusȱ onȱ methanolȱ conversionȱ intoȱ ethylene,ȱ asȱ theȱ mainȱ product.ȱ
and C3 H6 ; most studies focus on methanol conversion into ethylene, as the main product. (Scheme 3)
(Schemeȱ3)ȱOverȱSAPOȱcatalystsȱandȱzeolites,ȱbioȬmethanolȱcanȱbeȱselectivelyȱconvertedȱintoȱlightȱ
olefins;ȱoverȱSAPOȬ34,ȱselectivityȱtoȱC2H4ȱandȱC3H6ȱisȱ60.0%,ȱatȱ350–425ȱ°Cȱ[57].ȱModificationȱonȱtheȱ
catalystȱsynthesisȱprocedureȱstronglyȱaffectsȱproductȱdistribution.ȱDimethylȱetherȬtoȬolefinsȱ(DMTO)ȱ
isȱ aȱ similarȱ processȱ thatȱ yieldsȱ theȱ sameȱ products.ȱ BioȬdimethylȱ etherȱ canȱ beȱ producedȱ fromȱ
lignocellulosicȱbiomass,ȱviaȱpyrolysisȱandȱgasification,ȱresultingȱinȱtheȱformationȱofȱolefinsȱ(i.e.,ȱC2H4ȱ
andȱC3H6)ȱandȱsyngas,ȱfollowingȱtheȱbioliq®ȱconceptȱdevelopedȱinȱKarlsruheȱInstituteȱofȱTechnology,ȱ
Catalysts 2018, 8, 2 7 of 19

Over SAPO catalysts and zeolites, bio-methanol can be selectively converted into light olefins; over
SAPO-34, selectivity to C2 H4 and C3 H6 is 60.0%, at 350–425 ◦ C [57]. Modification on the catalyst
synthesis procedure strongly affects product distribution. Dimethyl ether-to-olefins (DMTO) is a
similar process that yields the same products. Bio-dimethyl ether can be produced from lignocellulosic
biomass, via pyrolysis and gasification, resulting in the formation of olefins (i.e., C2 H4 and C3 H6 ) and
syngas, following the bioliq® concept developed in Karlsruhe Institute of Technology, which is already
implemented in a large scale unit in Germany [58–60]. DMTO can take place at high temperature
(i.e., 723 ◦ C) and low pressure (i.e., 4 bar), fully converting dimethyl ether (DME) into C2 H4 (45.0%),
C3 H6 (39.0%), butenes (8.0%), and other light gases [10]. DME to olefins conversion is driven by a
substantially complex reaction mechanism based on methylation, oligomerization, and hydrocarbon
formation and cracking reactions, over zeolite catalysts [61].
Catalystsȱ2018,ȱ8,ȱ2ȱ 7ȱofȱ18ȱ

ȱ
Schemeȱ3.ȱBioȬmethanolȱconversionȱintoȱolefinsȱ(MTO).ȱ
Scheme 3. Bio-methanol conversion into olefins (MTO).

AnȱoverallȱcomparisonȱofȱtheȱmainȱC2H4ȱproductionȱprocessesȱisȱshownȱinȱTableȱ2.ȱDespiteȱofȱ
An overall comparison of the main C2 H4 production processes is shown in Table 2. Despite of the
theȱ technologicalȱ advancesȱ ofȱ theȱ lastȱ decades,ȱ bioȬethyleneȱ productionȱ processesȱ cannotȱ replaceȱ
technological advances of the last decades, bio-ethylene production processes cannot replace those
thoseȱdependentȱonȱfossilȱresources.ȱAsȱbioȬsynthesisȱprocessesȱareȱaȱrecentȱresearchȱsubject,ȱfurtherȱ
dependent on fossil resources. As bio-synthesis processes are a recent research subject, further studies
studiesȱ onȱ theȱ reductionȱ ofȱ theȱ costȱ andȱ onȱ increasingȱ theȱ productivityȱ areȱ essentialȱ priorȱ toȱ
on the reduction of the cost and on increasing the productivity are essential prior to considering their
consideringȱtheirȱindustrialȱapplication.ȱBioȬethanolȱdehydrationȱisȱtheȱmostȱpromisingȱalternativeȱ
industrial application. Bio-ethanol dehydration is the most promising alternative for the production
forȱtheȱproductionȱofȱ“green”ȱC2H4,ȱasȱnumerousȱstudiesȱhaveȱfocusedȱonȱtheȱselectionȱofȱtheȱmostȱ
of “green” C2 H4 , as numerous studies have focused on the selection of the most suitable catalyst,
suitableȱcatalyst,ȱloweringȱtheȱreactionȱtemperature,ȱwhileȱincreasingȱtheȱyieldȱtoȱC2H4.ȱAsȱethanolȱ
lowering the reaction temperature, while increasing the yield to C2 H4 . As ethanol dehydration has
dehydrationȱhasȱalreadyȱbeenȱimplemented,ȱtheȱonlyȱsetȬbackȱforȱindustrialȱbioȬethanolȱdehydrationȱ
already been implemented, the only set-back for industrial bio-ethanol dehydration to C2 H4 is the
toȱC2H4ȱisȱtheȱproductionȱandȱavailabilityȱofȱbioȬethanol.ȱMTOȱandȱDMTOȱprocessesȱhaveȱalsoȱbeenȱ
production and availability of bio-ethanol. MTO and DMTO processes have also been extensively
extensivelyȱ studiedȱ withȱ encouragingȱ resultsȱ regardingȱ theirȱ commercialȱ applications.ȱ Likewise,ȱ
studied with encouraging results regarding their commercial applications. Likewise, bio-methanol
bioȬmethanolȱ andȱ bioȬDMEȱ productionȱ isȱ alsoȱ limited,ȱ loweringȱ prospectȱ productivity.ȱ However,ȱ
and bio-DME production is also limited, lowering prospect productivity. However, future increases
futureȱincreasesȱofȱtheȱproducedȱbioȬfeedstocksȱcouldȱeliminateȱthisȱissue,ȱachievingȱhighȱyields,ȱinȱ
of the produced bio-feedstocks could eliminate this issue, achieving high yields, in cost-competitive
costȬcompetitiveȱprocesses,ȱasȱsteamȬcrackingȱunits.ȱFutureȱstudiesȱshouldȱfocusȱonȱcostȱreductionȱ
processes, as steam-cracking units. Future studies should focus on cost reduction linked with the
linkedȱwithȱtheȱimplementationȱofȱtheȱbioȬbasedȱmethods.ȱ
implementation of the bio-based methods.
Tableȱ2.ȱComparisonȱofȱC2H4ȱproductionȱprocesses.ȱ
Table 2. Comparison of C2 H4 production processes.
BioȬEthanolȱ
Processȱ SteamȱCrackingȱ BioȬSynthesisȱ MTOȱ DMTOȱ
Dehydrationȱ
Bio-Ethanol
Process Steam Cracking Bio-Synthesis MTO DMTO
Feedstockȱ HCȱ(Naphtha)ȱ ACC/SAMȱ Dehydration
BioȬethanolȱ BioȬmethanolȱ BioȬDMEȱ
Operatingȱ
Feedstock 675–700ȱ°Cȱ
HC (Naphtha) Ambient,ȱaerobicȱ
ACC/SAM Bio-ethanol 300–500ȱ°C,ȱ
Bio-methanol 675–750ȱ°C,ȱ
Bio-DME
180–500ȱ°Cȱ
Conditionsȱ
Operating atmosphericȱpressureȱ
675–700 ◦ C atmospheric conditionsȱ
Ambient, aerobic lowȱpressureȱ
300–500 ◦ C, lowȱpressure
675–750 ◦ C,
180–500 ◦ C
Conditions Alreadyȱinstalledȱ
pressure Selectiveȱsustainableȱ
conditions Commercialȱ low pressure low pressure
Advantagesȱ Closeȱtoȱcommercialȱapplication
technologyȱ
Already installed productionȱ
Selective sustainable applicationȱ
Commercial
Advantages Energyȱintense,ȱ Close to commercial application
technology production application
Limitedȱ
environmentalȱ
Energy intense, Increasedȱcost,ȱlowȱ Limited Limitedȱproductionȱofȱ
Disadvantagesȱ Increased cost, low bioȬethanolȱ Limited production of
Disadvantages concerns,ȱfiniteȱ
environmental concerns, productivityȱ bio-ethanol bioȬfeedstocksȱ
productivity supplyȱ bio-feedstocks
resourcesȱ
finite resources supply
YieldȱtoȱC
Yield to C22H
H44ȱ 31.3%ȱ
31.3% Ȭȱ- 99.9%ȱ
99.9% 41.5%ȱ
41.5% 45.0%ȱ
45.0%

2.2.ȱPropyleneȱ(C
2.2. Propylene (C33H
H66)ȱ)
AsȱtheȱdemandȱforȱC
As the demand for C33HH6ȱincreases,ȱresearchȱfocusesȱonȱonȬpurposeȱproductionȱofȱC 3H6,ȱbasedȱ
6 increases, research focuses on on-purpose production of C3 H6 , based
onȱbiomassȱresources,ȱaimingȱatȱsubstitutingȱoilȬbasedȱfeedstocks,ȱinȱmoreȱenvironmentallyȱfriendlyȱ
on biomass resources, aiming at substituting oil-based feedstocks, in more environmentally friendly
processes.ȱCorn,ȱvegetableȱoilsȱandȱotherȱbiomassȱproductsȱhaveȱbeenȱeffectivelyȱusedȱasȱfeedstocksȱ
processes. Corn, vegetable oils and other biomass products have been effectively used as feedstocks
forȱtheȱproductionȱofȱbioȬpropylene,ȱviaȱprocessesȱsuchȱasȱgasification,ȱmetathesis,ȱdehydrogenation,ȱ
for the production of bio-propylene, via processes such as gasification, metathesis, dehydrogenation,
fermentation,ȱandȱcrackingȱ[22].ȱ
fermentation, and cracking [22].
cost‐competitive processes, as steam‐cracking units. Future studies should focus on cost reduction 
linked with the implementation of the bio‐based methods. 

Table 2. Comparison of C2H4 production processes. 
Catalysts 2018, 8, 2 Bio‐Ethanol  8 of 19
Process  Steam Cracking  Bio‐Synthesis  MTO  DMTO 
Dehydration 
Feedstock  HC (Naphtha)  ACC/SAM  Bio‐ethanol  Bio‐methanol  Bio‐DME 
Conventionally, 675–700 °C 
Operating  C3 H6 is a by-product of C2 H4 production via steam cracking
Ambient, aerobic  of hydrocarbons
300–500 °C,  675–750 °C, 
180–500 °C 
andConditions 
FCC of gas oil.atmospheric pressure  conditions  various hydrocarbons (e.g.,
(Figure 7) At high temperatures, low pressure  low pressure
naphtha, ethane, and
Already installed  Selective sustainable  Commercial 
propane) are co-fed, under the most suitable operating conditions, to selectively
Advantages  yield C3 H6 [62]. FCC,
Close to commercial application
technology  production  application 
also uses hydrocarbons to produce
Energy intense, 
C H
3 6 , at moderate pressure and high temperatures, over zeolites,
Limited 
such as, ZSM-5, often modified
environmental  with metals to increase
Increased cost, low selectivity to C H
3 6 [63]. It is considered greener
Limited production of 
Disadvantages  bio‐ethanol 
than steam cracking concerns, finite 
due to lower energy productivity 
demand and decreased CO2 emissions. bio‐feedstocks 
Apart from the
supply 
resources 
well-known steam cracking and FCC, olefin metathesis and methanol to C3 H6 are also considered
Yield to C2H4  31.3%  ‐  99.9%  41.5%  45.0% 
alternatives for industrial application. Moreover, alkanes can be converted into alkenes via catalytic
dehydrogenation; propane can be used as a feedstock, at high temperature and atmospheric pressure.
2.2. Propylene (C3H6) 
Propane dehydrogenation plants have already been installed worldwide by several companies;
As the demand for C
Linde/BASF 3H6 increases, research focuses on on‐purpose production of C
(Alabama, USA), Lummus Technology (New Jersey, USA), Snamprogetti/Yarsintez 3H6, based 

on biomass resources, aiming at substituting oil‐based feedstocks, in more environmentally friendly 
(Jubail, Saudi Arabia), UOP (Illinois, USA), etc., mostly using chromium, and Pt–Sn catalysts [64].
processes. Corn, vegetable oils and other biomass products have been effectively used as feedstocks 
Methanol to C3 H6 is actually included in methanol to olefins reactions, described in the previous
for the production of bio‐propylene, via processes such as gasification, metathesis, dehydrogenation, 
chapter; zeolites (i.e., ZSM-5) are the most active catalysts in order to selectively produce C3 H6 from
fermentation, and cracking [22]. 
methanol [65].

 
Figure 7. Schematic chart of C3 H6 production methods.

As mentioned above, bio-ethanol can be produced through various methods. Apart from
constituting a valuable feedstock for C2 H4 production, bio-ethanol can also be used for C3 H6
production; C2 H4 from bio-ethanol can undergo dimerization followed by the well-known metathesis
reaction, along with butenes, thus producing C3 H6 [66]. For this process, not only bio-ethylene can
be derived from biomass (i.e., bio-ethanol), but also bio-butylene from bio-butanol dehydration [67].
In fact, butanol can be produced from bio-ethanol, via the Guerbet process, where higher alcohols
can be formed, upon condensation of two primary alcohols [68]. For this purpose, basic oxides,
such as MgO and hydrotalcites have been active in ethanol conversion into butanol, reaching 85%
selectivity [69]. N-butanol can also be produced through fermentation; several companies (e.g., Versalis
-San Donato Milanese, Italy, Global Bioenergies - Evry, France) have already developed biochemical
processes for the production of C4 alcohols. Through the ABE process (Acetone-Butanol-Ethanol), C4
alcohols can be produced via carbohydrate fermentation by genetically modified micro-organisms [39].
The metathesis reaction can yield C3 H6 using C2 H4 and butenes as a feedstock, via two different
approaches: (a) dimerization of bio-ethylene and then reaction with remaining bio-ethylene and
(b) direct reaction of bio-ethylene and bio-butene. (Scheme 4) C2 H4 dimerization processes have
already been implemented in the industry (e.g., AlphaButol by Axens- Rueil-Malmaison, France),
operating under relatively mild conditions (i.e., 0–100 ◦ C) and over metal catalysts, such as Ti and
Ni [70]. In direct reaction processes, (e.g., Olefin Conversion Technology-OCT ABB Lummus Global),
both homogeneous and heterogeneous catalysts have been employed. Heterogeneous (e.g., tungsten,
TheȱmetathesisȱreactionȱcanȱyieldȱC3H6ȱusingȱC2H4ȱandȱbutenesȱasȱaȱfeedstock,ȱviaȱtwoȱdifferentȱ
approaches:ȱ (a)ȱ dimerizationȱ ofȱ bioȬethyleneȱ andȱ thenȱ reactionȱ withȱ remainingȱ bioȬethyleneȱ andȱ (b)ȱ
directȱreactionȱofȱbioȬethyleneȱandȱbioȬbutene.ȱ(Schemeȱ4)ȱC2H4ȱdimerizationȱprocessesȱhaveȱalreadyȱ
beenȱimplementedȱinȱtheȱindustryȱ(e.g.,ȱAlphaButolȱbyȱAxensȬȱRueilȬMalmaison,ȱFrance),ȱoperatingȱ
underȱ relativelyȱ mildȱ conditionsȱ (i.e.,ȱ 0–100ȱ °C)ȱ andȱoverȱ metalȱ catalysts,ȱ suchȱ asȱ Tiȱ andȱ Niȱ[70].ȱ Inȱ
Catalysts 2018, 8, 2 9 of 19
directȱ reactionȱ processes,ȱ (e.g.,ȱ Olefinȱ Conversionȱ TechnologyȬOCTȱ ̄̅̅ȱ Lummusȱ Global),ȱ bothȱ
homogeneousȱ andȱ heterogeneousȱ catalystsȱ haveȱ beenȱ employed.ȱ Heterogeneousȱ (e.g.,ȱ tungsten,ȱ
molybdenum,ȱorȱRheniumȱoxides,ȱonȱaluminaȱorȱsilica)ȱareȱusuallyȱpreferredȱoverȱhomogeneous,ȱsuchȱ
molybdenum, or Rhenium oxides, on alumina or silica) are usually preferred over homogeneous,
asȱ organometallicȱ
such as organometallic complexes.ȱ Tungstenȱ
complexes. Tungstenoxides,ȱ supportedȱ
oxides, supportedonȱon
silicaȱ haveȱ
silica havebeenȱ
beenusedȱ
usedforȱ
for theȱ
the OCTȱ
OCT

process,ȱatȱtemperaturesȱaboveȱ260ȱ°Cȱandȱ30–35ȱbar,ȱreachingȱmoreȱthanȱ90.0%ȱselectivityȱtoȱC
process, at temperatures above 260 C and 30–35 bar, reaching more than 90.0% selectivity to C63,ȱforȱ 3H H6 ,
60.0%ȱ
for buteneȱ
60.0% buteneconversionȱ [71].ȱ[71].
conversion PreȬreductionȱ
Pre-reductiontreatment,ȱ asȱ wellȱ
treatment, as asȱ
wellincreasedȱ acidityȱ
as increased obtainedȱ
acidity usingȱ
obtained
moreȱ acidicȱ
using supports,ȱ
more acidic enhancesȱ
supports, catalyticȱ
enhances activityȱ
catalytic atȱ lowerȱ
activity temperatureȱ
at lower [72].ȱRheniumȱ
temperature catalystsȱ
[72]. Rhenium haveȱ
catalysts
have exhibited selectivity ~100%, however fast deactivation requires continuous regeneration [73].ȱ
exhibitedȱ selectivityȱ ~100%,ȱ howeverȱ fastȱ deactivationȱ requiresȱ continuousȱ regenerationȱ [73].
MolybdenaȱcatalystsȱhaveȱalsoȱbeenȱusedȱinȱtheȱShellȱHighȱOlefinȱProcessȱ(SHOP),ȱproducingȱ΅Ȭolefinsȱ
Molybdena catalysts have also been used in the Shell High Olefin Process (SHOP), producing α-olefins
throughȱtheȱoligomerizationȱofȱC
through the oligomerization of C22HH4,ȱfollowedȱbyȱolefinȱmetathesisȱ[74].ȱ
4 , followed by olefin metathesis [74].

ȱ
Scheme 4. C3 H6 formation via metathesis.

Solely 1-butene, produced via bio-butanol dehydration, can also be used as a feedstock for C3 H6
formation, as 1-butene is isomerized to 2-butene, and then both react via metathesis to form C3 H6
and 2-pentene [75]. Moreover, oligomerization/cracking of C2 H4 from bio-ethanol can result in C3 H6
production [76–78]. The direct ethanol to C3 H6 process (ETP) has recently been explored, over catalysts
such as zeolites [79–81] and metal oxide catalysts [82–84] in order to increase yield to C3 H6 against the
most common by-products (i.e., C2 H4 , butenes and aromatic hydrocarbons), improve stability and
suppress coke formation. Thus, as the yield to C3 H6 rarely exceeds 40.0%, the need for novel catalytic
materials is imperative [85].
Other biomass-derived oxygenates, such as polyols, aldehydes, and ketones can also be converted
into hydrocarbons through oxygen removal catalytic reactions (hydrogenation/ dehydrogenation, and
hydro-deoxygenation). In this context, over transition metal oxides, glycerol and other C3 oxygenated
compounds can be converted into C3 H6 . Glycerol is a low cost molecule that can be produced via
fermentation, transesterification and hydrogenolysis reactions, from biomass feedstocks. It can be
upgraded into valuable compounds through a number of chemical or biological routes [33]. Glycerol to
olefins (GTO) methods require the complete removal of oxygen, which is an intricate task. Its catalytic
conversion into C3 H6 is a novel research subject that has recently attracted attention; Hultenberg and
Brandin recently filed a patent introducing the production of lower hydrocarbons (e.g., ethane, propane,
and propene) from glycerol, over WO3 on ZrO2 and Pt on CeO2 catalysts. [86] Fadigas et al. initially
explored glycerol conversion, in a continuous flow process, over Ni, and Fe–Mo metal catalysts
supported on activated carbon, obtaining high selectivity values to C3 H6 [87]. Schmidt’s group
reported glycerol cracking into propanal, acrolein, C3 H6 , and C2 H4 , in three steps (i.e., dehydration,
hydrogenation, and upgrading), over HZSM-5 zeolites, Pd/α-Al2 O3 , and HBEA zeolites, respectively
for each step/reaction [88].
Yu et al used Ir/ZrO2 and H-ZSM5 catalysts, in two steps, in order to selectively produce
C3 H6 via hydro-deoxygenation in a fixed-dual-bed reactor. Over the optimized reaction conditions
(i.e., 250 ◦ C and 1 bar hydrogen pressure), selectivity to C3 H6 reached 85.0%, for complete glycerol
conversion [89]. (Table 3) Sun et al. used WO3 -Cu/Al2 O3 catalysts, at 250 ◦ C under hydrogen
flow (atmospheric pressure) to obtain 47.4% selectivity to C3 H6 for 100% glycerol conversion [91].
Combining WO3 -Cu/Al2 O3 and SiO2 -Al2 O3 in a dual-bed reactor, selectivity to C3 H6 reached 84.8%.
Mota et al, used Fe/Mo catalysts supported on activated carbon to obtain 100 glycerol conversion and
90.0% selectivity to C3 H6 at 300 ◦ C [92]. In all the above studies, a number of by-products has been
detected in the gas phase (e.g., carbon dioxide, propane, and ethylene). Further studies by our group,
in a batch reactor, over molybdena-based catalysts supported on carbon, report glycerol production
Catalysts 2018, 8, 2 10 of 19

into C3 H6 with 100% selectivity in the gas phase (88.0% glycerol conversion and 76.0% selectivity to
C3 H6 ) at 300 ◦ C and under hydrogen atmosphere (80 bar) (Figure 8) [90]. Our group has also proven
that reducible molybdenum oxides selectively drive this reaction into C3 H6 , most probably via a
reverse Mars—van Krevelen mechanism; formation of Mo4+ and Mo5+ species most likely drives the
reaction to the formation of the desired product. C3 H6 is most probably formed via two consecutive
cycles, in a one-step reaction: (a) due to the presence of oxygen vacancies, the adsorption of glycerol
proceeds with the two adjacent hydroxyls forming an unstable cyclic intermediate which in turn is
released as 2-propenol and (b) the latter after re-adsorption with the remaining hydroxyl is further
deoxygenated to C3 H6 [93]. Dow Global Technologies have patented a GTO process in a batch reactor,
reaching 96% selectivity to C3 H6 for 24% glycerol conversion. Hydroiodic acid acts as a catalyst in
subsequent reduction-oxidation cycles, over reductive atmosphere [92].

Table 3. Selected catalysts for one-step glycerol conversion into C3 H6 .


Catalystsȱ2018,ȱ8,ȱ2ȱ 10ȱofȱ18ȱ
Glycerol Selectivity to
Catalyst Temperature (◦ C) WHSV (h−1 ) Reference
Conversion (%) C3 H6 (%)
GTOȱ Ir/ZrO
processȱ inȱ aȱ batchȱ reactor,ȱ100.0
reachingȱ 96%ȱ85.0
selectivityȱ toȱ C 3H6ȱ forȱ 24%ȱ glycerolȱ conversion.ȱ
2 & HZSM-5-30 250 1.0 [89]
Hydroiodicȱ acidȱCarbon
Fe–Mo/Black actsȱ asȱ aȱ catalystȱ
88.0 inȱ subsequentȱ
76.0 reductionȬoxidationȱ
300 cycles,ȱ
- overȱ reductiveȱ
[90]
WO3 -Cu/Al2 O3 & SiO2 -Al2 O3 100.0 84.8 250 - [91]
atmosphereȱ[92].ȱ
Fe/Mo 100.0 90.0 300 5.4 [92]

ȱ
Figure 8. Effect of reaction time on selectivity over the Fe–Mo/BC_A catalyst (H2 pressure: 8.0 MPa,
Figureȱ8.ȱEffectȱofȱreactionȱtimeȱonȱselectivityȱoverȱtheȱFe–Mo/BC_Aȱcatalystȱ(H2ȱpressure:ȱ8.0ȱMPa,ȱ
temperature: 300 ◦ C). Reproduced from [90]. Copyright 2015, Royal Society of Chemistry.
temperature:ȱ300ȱ°C).ȱReproducedȱfromȱ[90].ȱCopyrightȱ2015,ȱRoyalȱSocietyȱofȱChemistry.ȱ

Bio-oil,
BioȬoil,ȱproduced
producedȱfrom fromȱcatalytic pyrolysis
catalyticȱ pyrolysisȱof ofȱ
fats, oilsoilsȱ
fats,ȱ andandȱ
other low-values
otherȱ compounds,
lowȬvaluesȱ can also
compounds,ȱ canȱ
be used as a feedstock for the production of olefins, through processes that are
alsoȱbeȱusedȱasȱaȱfeedstockȱforȱtheȱproductionȱofȱolefins,ȱthroughȱprocessesȱthatȱareȱalreadyȱusedȱinȱ already used in the
petrochemicals’
theȱ petrochemicals’ȱindustry [94]. Syntroleum
industryȱ Corporation
[94].ȱ Syntroleumȱ has already
Corporationȱ hasȱimplemented similar processed
alreadyȱ implementedȱ similarȱ
(i.e., Bio-Synfining), where vegetable oils and fats can be converted into fuels
processedȱ(i.e.,ȱBioȬSynfining),ȱwhereȱvegetableȱoilsȱandȱfatsȱcanȱbeȱconvertedȱintoȱfuelsȱandȱpropaneȱ and propane [95].
Neste Oil is also using the NExBTL (Next Generation Biomass to Liquid) process
[95].ȱNesteȱOilȱisȱalsoȱusingȱtheȱNExBTLȱ(NextȱGenerationȱBiomassȱtoȱLiquid)ȱprocessȱinȱorderȱtoȱ in order to produce
liquid
produceȱfuels and fuelsȱ
liquidȱ olefins
andȱ[96]. Through
olefinsȱ [96].ȱsteam
Throughȱcracking
steamȱand fluid catalytic
crackingȱ andȱ fluidȱcracking,
catalyticȱliquid fuelsliquidȱ
cracking,ȱ and
C H can be formed, while via the first route olefins are primarily produced
fuelsȱandȱC3H6ȱcanȱbeȱformed,ȱwhileȱviaȱtheȱfirstȱrouteȱolefinsȱareȱprimarilyȱproducedȱ[97].ȱTheȱfirstȱ
3 6 [97]. The first step
of the steam cracking process is hydro-deoxygenation of fatty acids and triglycerides, resulting in
stepȱofȱtheȱsteamȱcrackingȱprocessȱisȱhydroȬdeoxygenationȱofȱfattyȱacidsȱandȱtriglycerides,ȱresultingȱ
green hydrocarbons
inȱ greenȱ and naphtha.
hydrocarbonsȱ Hydro-deoxygenation
andȱ naphtha.ȱ proceeds over
HydroȬdeoxygenationȱ conventional
proceedsȱ overȱ hydrotreating
conventionalȱ
catalysts, under high hydrogen pressure and moderate temperatures (i.e., 280–400 ◦ C) [98]. Noble metal
hydrotreatingȱcatalysts,ȱunderȱhighȱhydrogenȱpressureȱandȱmoderateȱtemperaturesȱ(i.e.,ȱ280–400ȱ°C)ȱ
catalysts have also been used, supported on carbon, silica, alumina, or zeolites [99]. The second step
[98].ȱNobleȱmetalȱcatalystsȱhaveȱalsoȱbeenȱused,ȱsupportedȱonȱcarbon,ȱsilica,ȱalumina,ȱorȱzeolitesȱ[99].ȱ
involves
Theȱ secondȱ stepȱand
gasoline C3 H6 gasolineȱ
involvesȱ production viaCcracking
andȱ at atmospheric
3H6ȱ productionȱ pressure,
viaȱ crackingȱ at 800 ◦ C. Milder
atȱ atmosphericȱ reaction
pressure,ȱ atȱ
conditions enhance C H formation,
800ȱ°C.ȱMilderȱreactionȱconditionsȱenhanceȱC
3 6 while suppressing C H and aromatics
3H6ȱformation,ȱwhileȱsuppressingȱC
2 4 production.
2H4ȱandȱaromaticsȱ

production.ȱ
Tableȱ4ȱsummarizesȱtheȱmainȱcharacteristicsȱofȱtheȱmostȱimportantȱC3H6ȱproductionȱmethods.ȱ
Throughȱ steamȱ crackingȱ C3H6ȱ canȱ beȱ formedȱ asȱ aȱ byȬproductȱ ofȱ C2H4ȱ production,ȱ atȱ highȱ
temperatures,ȱinȱaȱmarkedlyȱenergyȱintenseȱprocess.ȱLowerȱtemperaturesȱofȱFCCȱmakeȱthisȱprocessȱ
moreȱ environmentallyȱ friendly,ȱ comparedȱ toȱ steamȱ cracking.ȱ However,ȱ yieldȱ toȱ C3H6ȱ isȱ highlyȱ
dependentȱ onȱ theȱ feedstockȱ andȱ onȱ theȱ selectedȱ condition.ȱ Catalystȱ deactivationȱ andȱ productȱ
Catalysts 2018, 8, 2 11 of 19

Table 4 summarizes the main characteristics of the most important C3 H6 production methods.
Through steam cracking C3 H6 can be formed as a by-product of C2 H4 production, at high temperatures,
in a markedly energy intense process. Lower temperatures of FCC make this process more
environmentally friendly, compared to steam cracking. However, yield to C3 H6 is highly dependent on
the feedstock and on the selected condition. Catalyst deactivation and product recovery are among the
disadvantages of this rather popular method. Propane dehydrogenation is another C3 H6 production
process from finite resources that also requires high temperatures and atmospheric pressure. Recently
installed units highlight the potential of this application for on-purpose C3 H6 production at high
yields. Via metathesis reactions olefins can be converted into C3 H6 , at relatively low temperatures
and moderate pressure. Bio-olefins are the most suitable feedstock but their availability limits the
productivity of this method. Glycerol to olefins is a novel research subject that aims at the production
of “green” C3 H6 through catalytic reactions, at moderate temperature. Nonetheless, more in-depth
research is essential on the most suitable catalysts that will selectively drive the reaction. Even though
C3 H6 production from glycerol is still a lab-scale application, future studies focusing on scale-up and
Catalystsȱ2018,ȱ8,ȱ2ȱ analyses will enable industrial application of these biomass-valorization processes.
techno-economic 11ȱofȱ18ȱ

Tableȱ4.ȱComparisonȱofȱC H6ȱproductionȱprocesses.ȱ
4. Comparison of C33H
Table 6 production processes.
Processȱ SteamȱCrackingȱ FCC Dehydrogenation Metathesisȱ GTO
Process
Feedstockȱ Steam Cracking
HCȱ FCC
HCȱ Dehydrogenation
Propaneȱ Metathesis
BioȬolefinsȱ GTO
Glycerolȱ
Feedstock HC HC Propane
500–700ȱ°C,ȱ Bio-olefins
0–260ȱ°C,ȱ Glycerol
Operatingȱ 750–900ȱ°C,ȱ 500–550ȱ°C,ȱ 250–400ȱ°C,ȱ
Operating 750–900 ◦ C, 500–550 ◦ C, 500–700 ◦ C,
atmosphericȱ ◦ C,
moderateȱ
0–260 250–400 ◦ C,
Conditionsȱ moderateȱpressureȱ moderateȱpressure hydrogenȱpressure
Conditions moderate pressure moderate pressure atmospheric pressureȱ
pressure moderatepressureȱ
pressure hydrogen pressure
Greenerȱthanȱ
Greener than
Alreadyȱinstalledȱ
Already installed SteamȱCracking,ȱ
Steam Cracking, Alreadyȱinstalledȱ
Already installed Alreadyȱ
Already installed Sustainableȱ
Sustainable
Advantagesȱ
Advantages
technology
technologyȱ Flexibility of
Flexibilityȱofȱ units
unitsȱ units
installedȱunitsȱ production
productionȱ
operation
operationȱ
Energyȱintense,ȱ
Energy intense, Yieldȱdependsȱonȱ
Yield depends on Catalystȱ
Limitedȱ Require hydrogen
Requireȱhydrogenȱ
environmental
environmentalȱ feedstock, catalyst
feedstock,ȱcatalystȱ Catalyst deactivation, Limited production
deactivation,ȱ
Disadvantages
Disadvantagesȱ productionȱofȱ atmosphere,
atmosphere,ȱ
concerns, finite
concerns,ȱfiniteȱ deactivation,ȱ endothermic
deactivation, reaction
endothermicȱ of bio-feedstocks
bioȬfeedstocksȱ lab-scale
labȬscaleȱ
resources product recovery
resourcesȱ productȱrecoveryȱ reactionȱ
Yield to C33H
YieldȱtoȱC H66ȱ 18.0%
18.0%ȱ 25.0%
25.0%ȱ 85.0%
85.0%ȱ 90.0%
90.0%ȱ 90.0%
90.0%ȱ

2.3. Butadiene (C44H


2.3.ȱButadieneȱ(C H66)ȱ)
AlternativeȱfeedstocksȱhaveȱbeenȱproposedȱinȱorderȱtoȱeffectivelyȱproduceȱbioȬbutadieneȱfromȱ
Alternative feedstocks have been proposed in order to effectively produce bio-butadiene
renewableȱ
from resources,ȱ
renewable thusȱ thus
resources, substitutingȱ theȱ energyȱ
substituting consumingȱ
the energy processesȱ
consuming ofȱ naphthaȱ
processes cracking.ȱ
of naphtha Inȱ
cracking.
Figureȱ9,ȱtheȱvariousȱproductionȱprocessesȱforȱC
In Figure 9, the various production processes for4H C46ȱareȱsummarized.ȱ
H6 are summarized.

ȱ
Figure 9. Schematic chart of C44H
Figureȱ9.ȱSchematicȱchartȱofȱC H66ȱproductionȱmethods.ȱ
production methods.

C4H6ȱ isȱ alsoȱ aȱ byȬproductȱ ofȱ C2H4ȱ productionȱ throughȱ crackingȱ ofȱ hydrocarbons.ȱ C4H6ȱ
productionȱfromȱbioȬethanolȱisȱaȱpromisingȱalternativeȱthatȱhasȱbeenȱindustriallyȱimplementedȱpriorȱ
toȱtheȱinstallationȱofȱnaphthaȱcrackingȱtechnologies;ȱethanolȱtoȱC4H6ȱprocessesȱhaveȱbeenȱusedȱsinceȱ
1920,ȱconstitutingȱtheȱmainȱproductionȱpracticeȱuntilȱtheȱendȱofȱWorldȱWarȱIIȱ[22,100].ȱMainȱroutesȱ
ofȱ thisȱ processȱ includeȱ dehydrogenation,ȱ dehydration,ȱ andȱ condensationȱ overȱ suitableȱ catalysts,ȱ
eitherȱ inȱ oneȬstepȱ (i.e.,ȱ Lebedevȱ approach)ȱ (Schemeȱ 5)ȱ orȱ twoȬstepȱ processesȱ (i.e.,ȱ Ostromisslenskiȱ
ȱ
Catalysts 2018, 8, 2 12 of 19
Figureȱ9.ȱSchematicȱchartȱofȱC4H6ȱproductionȱmethods.ȱ

C
C44HH66ȱ isȱ
is alsoȱ
also aȱ byȬproductȱ of
a by-product ofȱ CC22H
H44ȱ production
productionȱ through
throughȱcracking
crackingȱofofȱhydrocarbons.
hydrocarbons.ȱ C C44H66ȱ
productionȱfromȱbioȬethanolȱisȱaȱpromisingȱalternativeȱthatȱhasȱbeenȱindustriallyȱimplementedȱpriorȱ
production from bio-ethanol is a promising alternative that has been industrially implemented prior
toȱtheȱinstallationȱofȱnaphthaȱcrackingȱtechnologies;ȱethanolȱtoȱC
to the installation of naphtha cracking technologies; ethanol to C44H H66ȱprocessesȱhaveȱbeenȱusedȱsinceȱ
processes have been used since
1920,ȱconstitutingȱtheȱmainȱproductionȱpracticeȱuntilȱtheȱendȱofȱWorldȱWarȱIIȱ[22,100].ȱMainȱroutesȱ
1920, constituting the main production practice until the end of World War II [22,100]. Main routes of
ofȱ
thisthisȱ processȱ
process includeȱ
include dehydrogenation,ȱ
dehydrogenation, dehydration,ȱ
dehydration, andȱ condensationȱ
and condensation overȱ suitableȱ
over suitable catalysts,catalysts,ȱ
either in
eitherȱ
one-stepinȱ oneȬstepȱ (i.e.,ȱ approach)
(i.e., Lebedev Lebedevȱ approach)ȱ
(Scheme 5) (Schemeȱ 5)ȱ orȱprocesses
or two-step twoȬstepȱ (i.e.,
processesȱ (i.e.,ȱ Ostromisslenskiȱ
Ostromisslenski approach)
approach)ȱ(Schemeȱ6)ȱ[100].ȱInȱtheȱfirstȱcase,ȱmultifunctionalȱcatalystsȱhaveȱbeenȱemployed,ȱmainlyȱ
(Scheme 6) [100]. In the first case, multifunctional catalysts have been employed, mainly alumina
aluminaȱandȱmagnesiaȬsilicaȱ
and magnesia-silica catalysts,catalysts,ȱviaȱ C2H4ȱ dimerizationȱ
via C2 H4 dimerization andȱ subsequentȱmetathesis,ȱwhileȱ
and subsequent metathesis, while at the latter, atȱ
theȱlatter,ȱethanolȱwasȱinitiallyȱdehydrogenatedȱintoȱacetaldehyde,ȱoverȱcopperȬbasedȱcatalysts,ȱatȱ
ethanol was initially dehydrogenated into acetaldehyde, over copper-based catalysts, at moderate
moderateȱ
temperatures, temperatures,ȱ andȱ thenȱ
and then ethanol ethanolȱ
reacted with reactedȱ withȱ to
acetaldehyde acetaldehydeȱ toȱ formȱ
form C4 H6 , over Ta2 OC54H 6,ȱ overȱas
catalyst, Tawell
2O5ȱ

catalyst,ȱasȱwellȱasȱotherȱoxides,ȱsupportedȱonȱsilicaȱ[101,102].ȱSeveralȱresearchȱprojectsȱsuggestȱthatȱ
as other oxides, supported on silica [101,102]. Several research projects suggest that aldol condensation
aldolȱcondensationȱofȱacetaldehydeȱisȱaȱkeyȱstepȱinȱethanolȱconversionȱintoȱC
of acetaldehyde is a key step in ethanol conversion into C4 H6 [103]. On the other 4H6ȱ[103].ȱOnȱtheȱotherȱ
hand, latest works,
hand,ȱlatestȱworks,ȱincludingȱDRIFTSȱ(DiffuseȱReflectionȱInfraredȱSpectroscopy)ȱanalysesȱandȱDFTȱ
including DRIFTS (Diffuse Reflection Infrared Spectroscopy) analyses and DFT (Density Functional
(DensityȱFunctionalȱTheory)ȱcalculations,ȱproposeȱthatȱethanolȱcarbanionicȱintermediatesȱareȱofȱkeyȱ
Theory) calculations, propose that ethanol carbanionic intermediates are of key importance for the
importanceȱforȱtheȱreactionȱmechanismȱ[104].ȱ
reaction mechanism [104].

ȱ
Catalystsȱ2018,ȱ8,ȱ2ȱ Schemeȱ5.ȱC
Scheme 5. C44H
H66ȱproductionȱthroughȱtheȱLebedevȱprocess.ȱ
production through the Lebedev process. 12ȱofȱ18ȱ

ȱ
Schemeȱ6.ȱC
Scheme 6. C44H
H66ȱproductionȱthroughȱtheȱOstromisslenskiȱprocess.ȱ
production through the Ostromisslenski process.

Apartȱ fromȱ bioȬethanol,ȱ C4H6ȱ canȱ beȱ formedȱ fromȱ otherȱ biomassȱ derivedȱ oxygenatesȱ suchȱ asȱ
Apart from bio-ethanol, C4 H6 can be formed from other biomass derived oxygenates such as
butanolsȱandȱbutanediols,ȱproducedȱthroughȱbiomassȱfermentationȱorȱgasificationȱ[105,106].ȱInȱfact,ȱ
butanols and butanediols, produced through biomass fermentation or gasification [105,106]. In fact,
bioȬ1,4Ȭbutanediolȱ isȱ currentlyȱ producedȱ byȱ Genomatica,ȱ atȱ aȱ smallȱ scaleȱ plant,ȱ planningȱ onȱ
bio-1,4-butanediol is currently produced by Genomatica, at a small scale plant, planning on installing
installingȱanȱindustrialȱscaleȱoneȱwithȱNovamontȱ[107].ȱAcidȱcatalyzedȱdehydrationȱofȱbioȬbutanolsȱ
an industrial scale one with Novamont [107]. Acid catalyzed dehydration of bio-butanols yields a
yieldsȱ aȱ mixtureȱ ofȱ nȬbutenes.ȱ Subsequentȱ dehydrogenationȱ resultsȱ inȱ theȱ productionȱ ofȱ C4H6.ȱ
mixture of n-butenes. Subsequent dehydrogenation results in the production of C4 H6 . N-butanol
NȬbutanolȱ canȱ beȱ convertedȱ intoȱ 1Ȭbuteneȱ orȱ nonlinearȱ C4ȱ olefins,ȱ viaȱ dehydration,ȱ overȱ catalystsȱ
can be converted into 1-butene or nonlinear C4 olefins, via dehydration, over catalysts with mild
withȱ mildȱ orȱhighȱacidity,ȱrespectively;ȱoverȱ zeolitesȱaȱ~60%ȱyieldȱtoȱ isobuteneȱhasȱ beenȱ obtainedȱ
or high acidity, respectively; over zeolites a ~60% yield to isobutene has been obtained [108].
[108].ȱ Theȱ overwhelmingȱ costȱ ofȱ theȱ buteneȱ dehydrogenationȱ stepȱ hasȱ ledȱ toȱ researchȱ onȱ directȱ
The overwhelming cost of the butene dehydrogenation step has led to research on direct dehydration
dehydrationȱ ofȱ butanediolsȱ toȱ C4H6,ȱ overȱ suitableȱ catalysts;ȱ overȱ sodiumȱ phosphateȱ catalysts,ȱ
of butanediols to C4 H6 , over suitable catalysts; over sodium phosphate catalysts, 1,4-butanediol can be
1,4Ȭbutanediolȱ canȱ beȱ convertedȱ intoȱ C4H6ȱ atȱ 280ȱ °C,ȱ whileȱ conversionȱ ofȱ 1,3Ȭbutanediolȱ requiresȱ
converted into C4 H6 at 280 ◦ C, while conversion of 1,3-butanediol requires higher temperatures, over
higherȱ temperatures,ȱ overȱ theȱ sameȱ catalystsȱ [109].ȱ C4H6ȱ yieldȱ upȱ toȱ 95%ȱ hasȱ beenȱ reportedȱ fromȱ
the same catalysts [109]. C4 H6 yield up to 95% has been reported from 1,4-butanediol [110] and 90%
1,4Ȭbutanediolȱ [110]ȱ andȱ 90%ȱ fromȱ 1,3Ȭbutanediolȱ [103].ȱ Inȱ bothȱ cases,ȱ increasedȱ byȬproductȱ
from 1,3-butanediol [103]. In both cases, increased by-product formation is a critical issue that further
formationȱisȱaȱcriticalȱissueȱthatȱfurtherȱstudiesȱonȱcatalyticȱapproachesȱcouldȱaddress.ȱDehydrationȱ
studies on catalytic approaches could address. Dehydration of other C4 diols (e.g., 2,3-butanediol) is
ofȱ otherȱ C4ȱ diolsȱ (e.g.,ȱ 2,3Ȭbutanediol)ȱ isȱ muchȱ moreȱ challenging,ȱ requiringȱ multipleȱ andȱ moreȱ
much more challenging, requiring multiple and more complex reaction steps [111]. 2,3-butanediol,
complexȱreactionȱstepsȱ[111].ȱ2,3Ȭbutanediol,ȱproducedȱviaȱglucoseȱfermentation,ȱcanȱbeȱconvertedȱ
produced via glucose fermentation, can be converted into C4 H6 , over scandium oxide catalysts, at
intoȱ C4H6,ȱ overȱ scandiumȱ oxideȱ catalysts,ȱ atȱ highȱ temperaturesȱ (i.e.,ȱ 411ȱ °C),ȱ reachingȱ 88%ȱ yield.ȱ
high temperatures (i.e., 411 ◦ C), reaching 88% yield. Using two catalytic beds, with scandium oxide
Usingȱtwoȱcatalyticȱbeds,ȱwithȱscandiumȱoxideȱandȱalumina,ȱselectivityȱtoȱC4H6ȱwasȱ94%,ȱprovingȱ
and alumina, selectivity to C4 H6 was 94%, proving the feasibility of direct double bond dehydration
theȱ feasibilityȱ ofȱ directȱ doubleȱ bondȱ dehydrationȱ ofȱ 2,3Ȭbutanediolȱ [91].ȱ Alternatively,ȱ twoȬstepȱ
of 2,3-butanediol [91]. Alternatively, two-step processes, using two different catalysts (e.g., silica
processes,ȱusingȱtwoȱdifferentȱcatalystsȱ(e.g.,ȱsilicaȱorȱaluminaȬbased)ȱforȱsubsequentȱdehydrations,ȱ
or alumina-based) for subsequent dehydrations, have also been proposed, yielding C4 H6 via the
haveȱ alsoȱ beenȱ proposed,ȱ yieldingȱ C4H6ȱ viaȱ theȱ formationȱ ofȱ unsaturatedȱ alcoholsȱ [112].ȱ Novelȱ
formation of unsaturated alcohols [112]. Novel chemical and biochemical technologies enable the
chemicalȱandȱbiochemicalȱtechnologiesȱenableȱtheȱproductionȱofȱC4H6ȱfromȱsyngasȱoriginatingȱfromȱ
production of C4 H6 from syngas originating from the gasification of biomass or waste gases from
theȱgasificationȱofȱbiomassȱorȱwasteȱgasesȱfromȱtheȱsteelȱindustry;ȱbutanediolsȱcanȱbeȱproducedȱfromȱ
the steel industry; butanediols can be produced from syngas via fermentation [113]. As syngas can
syngasȱviaȱfermentationȱ[113].ȱAsȱsyngasȱcanȱbeȱproducedȱfromȱvariousȱorganicȱmaterials,ȱsuchȱasȱ
be produced from various organic materials, such as biomass, it is rather inexpensive. Thus, it is
biomass,ȱ itȱ isȱ ratherȱ inexpensive.ȱ Thus,ȱ itȱ isȱ anȱ excellentȱ resourceȱ forȱ theȱ productionȱ ofȱ valuableȱ
an excellent resource for the production of valuable bio-chemicals and bio-fuels [114]. Bio-catalytic
bioȬchemicalsȱ andȱ bioȬfuelsȱ [114].ȱ BioȬcatalyticȱ processesȱ enableȱ syngasȱ fermentationȱ inȱ moderateȱ
conditions,ȱincreasingȱenergyȱsaving,ȱimprovingȱproductȱyieldȱandȱinvolvingȱlessȱtoxicȱcompoundsȱ
orȱ productsȱ [115,116].ȱ Despiteȱ theȱ factȱ thatȱ syngasȱ fermentationȱ isȱ stillȱ anȱ immatureȱ approach,ȱ itsȱ
futureȱpotentialȱcannotȱbeȱignored.ȱ
Catalysts 2018, 8, 2 13 of 19

processes enable syngas fermentation in moderate conditions, increasing energy saving, improving
product yield and involving less toxic compounds or products [115,116]. Despite the fact that syngas
fermentation is still an immature approach, its future potential cannot be ignored.
As lighter olefins are most preferably produced via steam cracking, C4 H6 production is rather
limited through this process. Via dehydrogenation, butane—or butenes via oxidative dehydrogenation,
can be selectively converted into C4 H6 but the high temperatures and catalyst deactivation are key
disadvantages impeding commercialization. Bio-ethanol conversion into C4 H6 is the most promising
alternative; numerous studies have focused on the selection of the most suitable catalyst. However,
catalytic deactivation and increased by-product formation are still critical issues that further research
could resolve. Via dehydration/hydrogenation steps, bio-butanol can be effectively converted into
C4 H6 , but the limited availability of the feedstock, along with the significantly high cost due to high
temperature (up to 700 ◦ C) and to the hydrogenation step, hinder industrial application. (Table 5) On
the other hand, relatively lower temperatures of highly selective bio-butanediol dehydration could
enable its applicability, even though production of the bio-feedstock is quite limited and by-product
formation is not negligible. To conclude, all bio-based methods for C4 H6 production are still in lab-scale.
Nonetheless, in the future, novel catalysts can be synthesized for the selective C4 H6 production.
Moreover, further studies on production of the biomass-derived feedstocks could reduce the cost and
increase their availability in order to facilitate the implementation of these processes.

Table 5. Comparison of C4 H6 production processes.

Process Steam Cracking Dehydrogenation Lebedev/Ostromisslenski Dehydration


Feedstock Naphtha Butane/Butenes Bio-ethanol Bio-butanediols
Operating 750–900 ◦ C,
600–700/400–500 ◦ C 400–650 ◦ C 250–350 ◦ C
Conditions moderate pressure
Well-established
Bio-based, on-purpose Bio-based, on-purpose
Advantages Installed technology technology,
production production
on-purpose production
Energy demanding,
Limited production of
environmental concerns, High endothermicity, Catalyst deactivation,
Disadvantages bio-feedstock, various
finite resources, catalyst deactivation various by-products
by-products
limited production
Yield to C4 H6 4.5% 70.0%/71.8% 72.0%/56.5% up to 95.0%

3. Concluding Remarks
Production of bio-olefins is a broad research field that is continuously expanding, as the demand
is incessantly increasing, worldwide. Biomass derived intermediates offer numerous opportunities for
alternative reaction pathways, yielding ethylene, propylene, or butadiene, in less energy demanding
and cost effective processes that do not exploit the finite fossil resources. Bio-olefins can be produced
via numerous processes, some of which have already been implemented in industrial applications.
Ethanol dehydration is the most promising bio-based process for bio-ethylene production that
has already been installed, operating at relatively moderate temperatures. MTO and DMTO are close
to commercial application, using bio-methanol and bio-DME as feedstocks, for selective ethylene
production. In all cases, limited bio-ethanol supply is a key drawback that affects possibility of
industrial applications. Bio-synthesis is an interesting research subject on selective sustainable
production, but more in-depth studies are essential in order to increase productivity and significantly
lower the cost.
Bio-propylene can be effectively produced through bio-olefin metathesis and glycerol to olefins
methods, operating at relatively moderate conditions with increased bio-propylene yields. In fact,
olefin metathesis technology has already been implemented; however, limited bio-feedstock availability
strongly affects productivity and viability of this process. GTO methods include a wide range of
processes that could yield bio-propylene in lab-scale applications. Nevertheless, studies on the most
Catalysts 2018, 8, 2 14 of 19

efficient and stable catalyst that will lower hydrogen demand are expected to make these processes
applicable in the near future.
Bio-butadiene can be primarily produced from bio-ethanol via the Lebedev/Ostromisslenski
methods that have been extensively studied in the past decades. Dehydration of bio-butanediols
is the most promising approach, reaching 95% yield to butadiene, at lower temperature than
the other alternatives. Bio-butanol can also be used as a feedstock for bio-butadiene production,
but the high operation cost due to hydrogenation step and high temperature limits feasibility of
implementation. Additionally, C4 bio-feedstock production is also limited in order to ensure future
viable industrial applications.
Overall, several bio-based processes have been proposed with high potential in bio-olefin
yields. Even though the majority of them are still in laboratory scale, a few have already been
implemented around the world. The main drawback in the scale-up of these processes is the availability
of the bio-feedstocks which can be produced from various biomass derivatives via fermentation,
bio-synthesis, cracking, and deoxygenation among others. Future studies should mainly focus on
increasing the productivity of these methods, along with reducing the cost, in order to facilitate
their implementation in bio-olefin production units. In most catalytic approaches, novel low-cost
catalytic systems with improved properties regarding selectivity and reaction conditions should also
be researched to advance future applications.

Acknowledgments: This research has been financed by the State Scholarships Foundation (IKY) through the
program “RESEARCH PROJECTS FOR EXCELLENCE IKY/SIEMENS”.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. BP. Statistical Review of World Energy; BP: London, UK, 2017.
2. Amghizar, I.; Vandewalle, L.A.; Van Geem, K.M.; Marin, G.B. New Trends in Olefin Production. Engineering
2017, 3, 171–178. [CrossRef]
3. Bender, M. An Overview of Industrial Processes for the Production of Olefins—C4 Hydrocarbons.
ChemBioEng Rev. 2014, 1, 136–147. [CrossRef]
4. Torres Galvis, H.M.; de Jong, K.P. Catalysts for Production of Lower Olefins from Synthesis Gas: A Review.
ACS Catal. 2013, 3, 2130–2149. [CrossRef]
5. Matar, S.; Hatch, L.F. Chemistry of Petrochemical Processes; Gulf Professional Publishing: Houston, TX, USA, 2009.
6. Weissermel, K.; Arpe, H. Oxidation Products Ethylene. In Industrial Organic Chemistry; Wiley-VCH Verlag
GmbH: Weinheim, Germany, 1978; pp. 145–192; ISBN 9783527619191.
7. Zimmermann, H.; Walzl, R. Ethylene. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag
GmbH & Co. KGaA: Weinheim, Germany, 2009; ISBN 9783527306732.
8. Mohsenzadeh, A.; Zamani, A.; Taherzadeh, M.J. Bioethylene Production from Ethanol: A Review and
Techno-economical Evaluation. ChemBioEng Rev. 2017, 4, 75–91. [CrossRef]
9. Chang, C.D. Methanol Conversion to Light Olefins. Catal. Rev. 1984, 26, 323–345. [CrossRef]
10. Haro, P.; Trippe, F.; Stahl, R.; Henrich, E. Bio-syngas to gasoline and olefins via DME—A comprehensive
techno-economic assessment. Appl. Energy 2013, 108, 54–65. [CrossRef]
11. McManus, M.T. The Plant Hormone Ethylene; Wiley-Blackwell: Hoboken, NJ, USA, 2012; ISBN 1444330039.
12. Ding, J.; Hua, W. Game Changers of the C3 Value Chain: Gas, Coal, and Biotechnologies. Chem. Eng. Technol.
2013, 36, 83–90. [CrossRef]
13. Eisele, P.; Killpack, R. Propene. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH &
Co. KGaA: Weinheim, Germany, 2011; ISBN 9783527306732.
14. Sahebdelfar, S.; Zangeneh, F.T. Dehydrogenation of Propane to Propylene Over Pt-Sn/Al2 O3 Catalysts:
The influence of operating conditions on product selectivity. Iran. J. Chem. Eng. 2010, 7, 51–57.
15. Lwin, S.; Wachs, I.E. Olefin Metathesis by Supported Metal Oxide Catalysts. ACS Catal. 2014, 4, 2505–2520.
[CrossRef]
Catalysts 2018, 8, 2 15 of 19

16. Chen, D.; Moljord, K.; Holmen, A. A methanol to olefins review: Diffusion, coke formation and deactivation
on SAPO type catalysts. Microporous Mesoporous Mater. 2012, 164, 239–250. [CrossRef]
17. Tian, P.; Wei, Y.; Ye, M.; Liu, Z. Methanol to Olefins (MTO): From Fundamentals to Commercialization.
ACS Catal. 2015, 5, 1922–1938. [CrossRef]
18. Werpy, T.; Petersen, G.; Aden, A.; Bozell, J. Top Value Added Chemicals from Biomass. Volume 1-Results of
Screening for Potential Candidates from Sugars and Synthesis Gas; U.S. Department of Energy: Washington, DC,
USA, 2004.
19. Bruijnincx, P.C.A.; Weckhuysen, B.M. Shale Gas Revolution: An Opportunity for the Production of Biobased
Chemicals? Angew. Chem. Int. Ed. 2013, 52, 11980–11987. [CrossRef] [PubMed]
20. Siirola, J.J. The impact of shale gas in the chemical industry. AIChE J. 2014, 60, 810–819. [CrossRef]
21. Fiorentino, G.; Ripa, M.; Ulgiati, S. Chemicals from biomass: Technological versus environmental feasibility.
A review. Biofuels Bioprod. Biorefin. 2017, 11, 195–214. [CrossRef]
22. Chieregato, A.; Ochoa, J.V.; Cavani, F. Olefins from Biomass. In Chemicals and Fuels from Bio-Based Building
Blocks; Cavani, F., Albonetti, S., Basile, F., Gandini, A., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA:
Weinheim, Germany, 2016. [CrossRef]
23. Yang, S.F.; Hoffman, N.E. Ethylene Biosynthesis and its Regulation in Higher Plants. Annu. Rev. Plant Physiol.
1984, 35, 155–189. [CrossRef]
24. Lynch, S.; Eckert, C.; Yu, J.; Gill, R.; Maness, P.-C. Overcoming substrate limitations for improved production
of ethylene in E. coli. Biotechnol. Biofuels 2016, 9, 3. [CrossRef] [PubMed]
25. Nickerson, W.J. Ethylene as a metabolic product of the pathogenic fungus, Blastomyces dermatitidis.
Arch Biochem. 1948, 17, 225–233. [PubMed]
26. Eckert, C.; Xu, W.; Xiong, W.; Lynch, S.; Ungerer, J.; Tao, L.; Gill, R.; Maness, P.-C.; Yu, J. Ethylene-forming
enzyme and bioethylene production. Biotechnol. Biofuels 2014, 7, 33. [CrossRef] [PubMed]
27. Wang, J.P.; Wu, L.X.; Xu, F.; Lv, J.; Jin, H.J.; Chen, S.F. Metabolic engineering for ethylene production by
inserting the ethylene-forming enzyme gene (efe) at the 16S rDNA sites of Pseudomonas putida KT2440.
Bioresour. Technol. 2010, 101, 6404–6409. [CrossRef] [PubMed]
28. Markham, J.N.; Tao, L.; Davis, R.; Voulis, N.; Angenent, L.T.; Ungerer, J.; Yu, J. Techno-economic analysis
of a conceptual biofuel production process from bioethylene produced by photosynthetic recombinant
cyanobacteria. Green Chem. 2016, 18, 6266–6281. [CrossRef]
29. Zhang, M.; Yu, Y. Dehydration of ethanol to ethylene. Ind. Eng. Chem. Res. 2013, 52, 9505–9514. [CrossRef]
30. Fan, D.; Dai, D.J.; Wu, H.S. Ethylene formation by catalytic dehydration of ethanol with industrial
considerations. Materials 2013, 6, 101–115. [CrossRef] [PubMed]
31. Kosaric, N.; Duvnjak, Z.; Farkas, A.; Sahm, H.; Bringer-Meyer, S.; Goebel, O.; Mayer, D.; Kosaric, N.;
Duvnjak, Z.; Farkas, A.; et al. Ethanol. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag
GmbH & Co. KGaA: Weinheim, Germany, 2011; pp. 1–72; ISBN 9783527306732.
32. Althoff, J.; Biesheuvel, K.; De Kok, A.; Pelt, H.; Ruitenbeek, M.; Spork, G.; Tange, J.; Wevers, R. Economic
Feasibility of the Sugar Beet-to-Ethylene Value Chain. ChemSusChem 2013, 6, 1625–1630. [CrossRef] [PubMed]
33. Sheldon, R.A. Green and sustainable manufacture of chemicals from biomass: State of the art. Green Chem.
2014, 16, 950–963. [CrossRef]
34. Fukumoto, M.; Kimura, A. Process for the Manufacture of Ethylene by Dehydration of Ethanol. EP 2594546
A1, 17 November 2011.
35. Chen, Y.; Wu, Y.; Tao, L.; Dai, B.; Yang, M.; Chen, Z.; Zhu, X. Dehydration reaction of bio-ethanol to ethylene
over modified SAPO catalysts. J. Ind. Eng. Chem. 2010, 16, 717–722. [CrossRef]
36. Wu, L.; Shi, X.; Cui, Q.; Wang, H.; Huang, H. Effects of the SAPO-11 synthetic process on dehydration Of
ethanol to ethylene. Front. Chem. Sci. Eng. 2011, 5, 60–66. [CrossRef]
37. Zhang, D.; Wang, R.; Yang, X. Effect of P content on the catalytic performance of P-modified HZSM-5
catalysts in dehydration of ethanol to ethylene. Catal. Lett. 2008, 124, 384–391. [CrossRef]
38. Patrick, G.B.; Partington, S.R. Process for Preparing Ethene. U.S. Patent 8822748 B2, 8 October 2008.
39. Lanzafame, P.; Centi, G.; Perathoner, S. Evolving scenarios for biorefineries and the impact on catalysis.
Catal. Today 2014, 234, 2–12. [CrossRef]
40. Kang, M.; DeWilde, J.F.; Bhan, A. Kinetics and Mechanism of Alcohol Dehydration on γ-Al2 O3 : Effects of
Carbon Chain Length and Substitution. ACS Catal. 2015, 5, 602–612. [CrossRef]
Catalysts 2018, 8, 2 16 of 19

41. Christiansen, M.A.; Mpourmpakis, G.; Vlachos, D.G. DFT-driven multi-site microkinetic modeling of ethanol
conversion to ethylene and diethyl ether on γ-Al2 O3 (1 1 1). J. Catal. 2015, 323, 121–131. [CrossRef]
42. Potter, M.E.; Cholerton, M.E.; Kezina, J.; Bounds, R.; Carravetta, M.; Manzoli, M.; Gianotti, E.; Lefenfeld, M.;
Raja, R. Role of Isolated Acid Sites and Influence of Pore Diameter in the Low-Temperature Dehydration of
Ethanol. ACS Catal. 2014, 4, 4161–4169. [CrossRef]
43. DeWilde, J.F.; Czopinski, C.J.; Bhan, A. Ethanol Dehydration and Dehydrogenation on γ-Al2 O3 : Mechanism
of Acetaldehyde Formation. ACS Catal. 2014, 4, 4425–4433. [CrossRef]
44. Kochar, N.K.; Merims, R.; Padia, A.S. Ethylene from Ethanol. Chem. Eng. Prog. 1981, 6, 66–70.
45. Ondrey, G. The Launch of a New Bioethylene-Production Process; Chemical Engineering: New York, NY, USA, 2014.
46. Chen, G.; Li, S.; Jiao, F.; Yuan, Q. Catalytic dehydration of bioethanol to ethylene over TiO2 /γ-Al2 O3 catalysts
in microchannel reactors. Catal. Today 2007, 125, 111–119. [CrossRef]
47. Zhan, N.; Hu, Y.; Li, H.; Yu, D.; Han, Y.; Huang, H. Lanthanum-phosphorous modified HZSM-5 catalysts in
dehydration of ethanol to ethylene: A comparative analysis. Catal. Commun. 2010, 11, 633–637. [CrossRef]
48. Ciftci, A.; Varisli, D.; Tokay, K.C.; Sezgi, N.A.; Dogu, T. Dimethyl ether, diethyl ether & ethylene from alcohols
over tungstophosphoric acid based mesoporous catalysts. Chem. Eng. J. 2012, 207–208, 85–93.
49. Varisli, D.; Dogu, T.; Dogu, G. Silicotungstic acid impregnated MCM-41-like mesoporous solid acid catalysts
for dehydration of ethanol. Ind. Eng. Chem. Res. 2008, 47, 4071–4076. [CrossRef]
50. Chang, C.D.; Silvestri, A.J. The conversion of methanol and other O-compounds to hydrocarbons over zeolite
catalysts. J. Catal. 1977, 47, 249–259. [CrossRef]
51. Olsbye, U.; Svelle, S.; Bjorgen, M.; Beato, P.; Janssens, T.V.W.; Joensen, F.; Bordiga, S.; Lillerud, K.P. Conversion
of methanol to hydrocarbons: How zeolite cavity and pore size controls product selectivity. Angew. Chem.
Int. Ed. 2012, 51, 5810–5831. [CrossRef] [PubMed]
52. Güllü, D.; Demirbas, A. Biomass to methanol via pyrolysis process. Energy Convers. Manag. 2001, 42,
1349–1356. [CrossRef]
53. Liu, Z.; Liang, J. Methanol to olefins conversion catalysts Curr. Opin. Solid State Mater. Sci. 1999, 4, 80–84.
[CrossRef]
54. Galindo, C.P.; Badr, O. Renewable hydrogen utilisation for the production of methanol. Energy Convers.
Manag. 2007, 48, 519–527. [CrossRef]
55. Taylor, C.E.; Noceti, R.P.; Joseph, R.D. New developments in the photocatalytic conversion of methane to
methanol. Catal. Today 2000, 55, 259–267. [CrossRef]
56. Shamsul, N.S.; Kamarudin, S.K.; Rahman, N.A.; Kofli, N.T. An overview on the production of bio-methanol
as potential renewable energy. Renew. Sustain. Energy Rev. 2014, 33, 578–588. [CrossRef]
57. Liang, J.; Li, H.; Zhao, S.; Guo, W.; Wang, R.; Ying, M. Characteristics and performance of SAPO-34 catalyst
for methanol-to-olefin conversion. Appl. Catal. 1990, 64, 31–40. [CrossRef]
58. Raffelt, K.; Henrich, E.; Koegel, A.; Stahl, R.; Steinhardt, J.; Weirich, F. The BTL2 Process of Biomass Utilization
Entrained-Flow Gasification of Pyrolyzed Biomass Slurries. Appl. Biochem. Biotechnol. 2006, 129, 153–164.
[CrossRef]
59. Henrich, E.; Dahmen, N.; Dinjus, E. Cost estimate for biosynfuel production via biosyncrude gasification.
Biofuels Bioprod. Biorefin. 2009, 3, 28–41. [CrossRef]
60. Wright, M.M.; Brown, R.C.; Boateng, A.A. Distributed processing of biomass to bio-oil for subsequent
production of Fischer-Tropsch liquids. Biofuels Bioprod. Biorefin. 2008, 2, 229–238. [CrossRef]
61. Kvisle, S.; Fuglerud, T.; Kolboe, S.; Olsbye, U.; Lillerud, K.P.; Vora, B.V.; Kvisle, S.; Fuglerud, T.; Kolboe, S.;
Olsbye, U.; et al. Methanol-to-Hydrocarbons. In Handbook of Heterogeneous Catalysis; Wiley-VCH Verlag
GmbH & Co. KGaA: Weinheim, Germany, 2008; ISBN 9783527610044.
62. Ren, T.; Patel, M.; Blok, K. Olefins from conventional and heavy feedstocks: Energy use in steam cracking
and alternative processes. Energy 2006, 31, 425–451. [CrossRef]
63. Akah, A.; Al-Ghrami, M. Maximizing propylene production via FCC technology. Appl. Petrochem. Res. 2015,
5, 377–392. [CrossRef]
64. Sattler, J.J.H.B.; Ruiz-Martinez, J.; Santillan-Jimenez, E.; Weckhuysen, B.M. Catalytic Dehydrogenation of
Light Alkanes on Metals and Metal Oxides. Chem. Rev. 2014, 114, 10613–10653. [CrossRef] [PubMed]
65. Zhang, S.; Gong, Y.; Zhang, L.; Liu, Y.; Dou, T.; Xu, J.; Deng, F. Hydrothermal treatment on ZSM-5 extrudates
catalyst for methanol to propylene reaction: Finely tuning the acidic property. Fuel Process. Technol. 2015,
129, 130–138. [CrossRef]
Catalysts 2018, 8, 2 17 of 19

66. Banks, R.L.; Kukes, S.G. New developments and concepts in enhancing activities of heterogeneous metathesis
catalysts. J. Mol. Catal. 1985, 28, 117–131. [CrossRef]
67. Knifton, J.F.; Sanderson, J.R.; Stockton, M.E. Tert-butanol dehydration to isobutylene via reactive distillation.
Catal. Lett. 2001, 73, 55–57. [CrossRef]
68. Kozlowski, J.T.; Davis, R.J. Heterogeneous catalysts for the guerbet coupling of alcohols. ACS Catal. 2013, 3,
1588–1600. [CrossRef]
69. Dowson, G.R.M.; Haddow, M.F.; Lee, J.; Wingad, R.L.; Wass, D.F. Catalytic conversion of ethanol into an
advanced biofuel: Unprecedented selectivity for n-butanol. Angew. Chem. Int. Ed. 2013, 52, 9005–9008.
[CrossRef] [PubMed]
70. HOOD, A.D., Jr.; Bridges, R.S. Staged Propylene Production Process. WO 2014110125 A1, 17 July 2014.
71. Mol, J.C. Industrial applications of olefin metathesis. J. Mol. Catal. A Chem. 2004, 213, 39–45. [CrossRef]
72. Huang, S.; Liu, S.; Xin, W.; Bai, J.; Xie, S.; Wang, Q.; Xu, L. Metathesis of ethene and 2-butene to propene on
W/Al2 O3 -HY catalysts with different HY contents. J. Mol. Catal. A Chem. 2005, 226, 61–68. [CrossRef]
73. Bouchmella, K.; Hubert Mutin, P.; Stoyanova, M.; Poleunis, C.; Eloy, P.; Rodemerck, U.; Gaigneaux, E.M.;
Debecker, D.P. Olefin metathesis with mesoporous rhenium-silicium-aluminum mixed oxides obtained via a
one-step non-hydrolytic sol-gel route. J. Catal. 2013, 301, 233–241. [CrossRef]
74. Busca, G. Heterogeneous Catalytic Materials: Solid State Chemistry, Surface Chemistry and Catalytic Behaviour;
Elsevier: Warsaw, Poland, 2014.
75. Popoff, N.; Mazoyer, E.; Pelletier, J.; Gauvin, R.M.; Taoufik, M. Expanding the scope of metathesis: A survey
of polyfunctional, single-site supported tungsten systems for hydrocarbon valorization. Chem. Soc. Rev.
2013, 42, 9035–9054. [CrossRef] [PubMed]
76. Inoue, K.; Inaba, M.; Takahara, I.; Murata, K. Conversion of ethanol to propylene by H-ZSM-5 with Si/Al2
ratio of 280. Catal. Lett. 2010, 136, 14–19. [CrossRef]
77. Kazuhisa, M.; Takahara, I.; Inaba, M. Propane Formation by Aqueous-Phase Reforming of Glycerol over
Pt/H-ZSM5 Catalysts. React. Kinet. Catal. Lett. 2008, 93, 59–66.
78. Lin, B.; Zhang, Q.; Wang, Y. Catalytic conversion of ethylene to propylene and butenes over H-ZSM-5.
Ind. Eng. Chem. Res. 2009, 48, 10788–10795. [CrossRef]
79. Huangfu, J.; Mao, D.; Zhai, X.; Guo, Q. Remarkably enhanced stability of HZSM-5 zeolite co-modified with
alkaline and phosphorous for the selective conversion of bio-ethanol to propylene. Appl. Catal. A Gen. 2016,
520, 99–104. [CrossRef]
80. Zhang, N.; Mao, D.; Zhai, X. Selective conversion of bio-ethanol to propene over nano-HZSM-5 zeolite:
Remarkably enhanced catalytic performance by fluorine modification. Fuel Process. Technol. 2017, 167, 50–60.
[CrossRef]
81. Hayashi, F.; Iwamoto, M. Yttrium-modified ceria as a highly durable catalyst for the selective conversion of
ethanol to propene and ethene. ACS Catal. 2013, 3, 14–17. [CrossRef]
82. Hayashi, F.; Tanaka, M.; Lin, D.; Iwamoto, M. Surface structure of yttrium-modified ceria catalysts and
reaction pathways from ethanol to propene. J. Catal. 2014, 316, 112–120. [CrossRef]
83. Iwamoto, M.; Mizuno, S.; Tanaka, M. Direct and selective production of propene from bio-ethanol on
Sc-loaded In2 O3 catalysts. Chem. A Eur. J. 2013, 19, 7214–7220. [CrossRef] [PubMed]
84. Xia, W.; Wang, F.; Mu, X.; Chen, K.; Takahashi, A.; Nakamura, I.; Fujitani, T. Catalytic performance of
H-ZSM-5 zeolites for conversion of ethanol or ethylene to propylene: Effect of reaction pressure and
SiO2 /Al2 O3 ratio. Catal. Commun. 2017, 91, 62–66. [CrossRef]
85. Xue, F.; Miao, C.; Yue, Y.; Hua, W.; Gao, Z. Direct conversion of bio-ethanol to propylene in high yield over
the composite of In2 O3 and zeolite beta. Green Chem. 2017. [CrossRef]
86. Hulteberg, C.; Brandin, J. Process for Preparing Lower Hydrocarbons from Glycerol. U.S. Patent 20110224470
A1, 15 September 2011.
87. Souza, F.J.C.; Gambetta, R.; de Araujo Mota, C.J.; da Conceicao Goncalves, V.L. Preparation of Heterogeneous
Catalysts Used in Selective Hydrogenation of Glycerin to Propene, and a Process for the Selective
Hydrogenation of Glycerin to Propene. U.S. Patent 8841497 B2, 24 June 2009.
88. Blass, S.D.; Hermann, R.J.; Persson, N.E.; Bhan, A.; Schmidt, L.D. Conversion of glycerol to light olefins and
gasoline precursors. Appl. Catal. A Gen. 2014, 475, 10–15. [CrossRef]
89. Yu, L.; Yuan, J.; Zhang, Q.; Liu, Y.M.; He, H.Y.; Fan, K.N.; Cao, Y. Propylene from renewable resources:
Catalytic conversion of glycerol into propylene. ChemSusChem 2014, 7, 743–747. [CrossRef] [PubMed]
Catalysts 2018, 8, 2 18 of 19

90. Zacharopoulou, V.; Vasiliadou, E.S.; Lemonidou, A.A. One-step propylene formation from bio-glycerol over
molybdena-based catalysts. Green Chem. 2015, 17, 903–912. [CrossRef]
91. Sun, D.; Yamada, Y.; Sato, S. Efficient production of propylene in the catalytic conversion of glycerol.
Appl. Catal. B Environ. 2015, 174, 13–20. [CrossRef]
92. Deshpande, R.; Davis, P.; Pandey, V.; Kore, N. Dehydroxylation of Crude Alcohol Streams Using a
Halogen-Based Catalyst. WO 2013090076 A1, 20 June 2013.
93. Zacharopoulou, V.; Vasiliadou, E.; Lemonidou, A.A. Exploring the reaction pathways of bio-glycerol
hydro-deoxygenation to propene over Molybdena-based catalysts. ChemSusChem 2017. [CrossRef] [PubMed]
94. Pyl, S.P.; Schietekat, C.M.; Reyniers, M.F.; Abhari, R.; Marin, G.B.; Van Geem, K.M. Biomass to olefins:
Cracking of renewable naphtha. Chem. Eng. J. 2011, 176–177, 178–187. [CrossRef]
95. Ramin, A.H.; Lynn Tomlinson, G.R. Biorenewable Naphtha Composition and Methods of Making Same. U.S.
Patent 8581013 B2, 12 November 2013.
96. Vermeiren, W.; Van Gyseghem, N. A process for the Production of Bio-Naphtha from Complex Mixtures of
Natural Occurring Fats & Oils. WO 2011012439 A1, 3 February 2011.
97. Bielansky, P.; Weinert, A.; Schönberger, C.; Reichhold, A. Catalytic conversion of vegetable oils in a continuous
FCC pilot plant. Fuel Process. Technol. 2011, 92, 2305–2311. [CrossRef]
98. Mortensen, P.M.; Grunwaldt, J.D.; Jensen, P.A.; Knudsen, K.G.; Jensen, A.D. A review of catalytic upgrading
of bio-oil to engine fuels. Appl. Catal. A Gen. 2011, 407, 1–19. [CrossRef]
99. Ruddy, D.A.; Schaidle, J.A.; Ferrell, J.R., III; Wang, J.; Moens, L.; Hensley, J.E. Recent advances in
heterogeneous catalysts for bio-oil upgrading via “ex situ catalytic fast pyrolysis”: Catalyst development
through the study of model compounds. Green Chem. 2014, 16, 454–490. [CrossRef]
100. Angelici, C.; Weckhuysen, B.M.; Bruijnincx, P.C.A. Chemocatalytic conversion of ethanol into butadiene and
other bulk chemicals. ChemSusChem 2013, 6, 1595–1614. [CrossRef] [PubMed]
101. Quattlebaum, W.M., Jr.; Toussaint, W.J. Process of Making Olefins. U.S. Patent 2407291 A, 10 September 1946.
102. Toussant, W.J.; Dunn, J.T.; Jackson, D.R. Production of Butadiene from Alcohol. Ind. Eng. Chem. 1947, 39,
120–125. [CrossRef]
103. Makshina, E.V.; Dusselier, M.; Janssens, W.; Degrève, J.; Jacobs, P.A.; Sels, B.F. Review of old chemistry
and new catalytic advances in the on-purpose synthesis of butadiene. Chem. Soc. Rev. 2014, 43, 7917–7953.
[CrossRef] [PubMed]
104. Chieregato, A.; Ochoa, J.V.; Bandinelli, C.; Fornasari, G.; Cavani, F.; Mella, M. On the chemistry of ethanol on
basic oxides: Revising mechanisms and intermediates in the lebedev and guerbet reactions. ChemSusChem
2015, 8, 377–388. [CrossRef] [PubMed]
105. Xiu, Z.L.; Zeng, A.P. Present state and perspective of downstream processing of biologically produced
1,3-propanediol and 2,3-butanediol. Appl. Microbiol. Biotechnol. 2008, 78, 917–926. [CrossRef] [PubMed]
106. Celińska, E.; Grajek, W. Biotechnological production of 2,3-butanediol-Current state and prospects.
Biotechnol. Adv. 2009, 27, 715–725. [CrossRef] [PubMed]
107. Mark, J.B.; Anthony, P.B.; Robin, E.O.; Jun, S.P.P. Microorganisms for Producing Butadiene and Methods
Related Thereto. WO 2012177710 A1, 27 December 2012.
108. Zhang, D.; Barri, S.A.I.; Chadwick, D. N-Butanol to iso-butene in one-step over zeolite catalysts. Appl. Catal.
A Gen. 2011, 403, 1–11. [CrossRef]
109. Arpe, H.-J.; Weissermel, K. Industrial Organic Chemistry; Wiley-VCH: Weinheim, Germany, 2010;
ISBN 3527320024.
110. Reppe, W.; Steinhofer, A.; Daumiller, G. Verfahren zur Herstellung von Diolefinen. DE 899350, 10 August 1953.
111. Kopke, M.; Mihalcea, C.; Liew, F.; Tizard, J.H.; Ali, M.S.; Conolly, J.J.; Al-Sinawi, B.; Simpson, S.D.
2,3-Butanediol Production by Acetogenic Bacteria, an Alternative Route to Chemical Synthesis, Using
Industrial Waste Gas. Appl. Environ. Microbiol. 2011, 77, 5467–5475. [CrossRef] [PubMed]
112. Duan, H.; Yamada, Y.; Sato, S. Selective dehydration of 2,3-butanediol to 3-buten-2-ol over ZrO2 modified
with CaO. Appl. Catal. A Gen. 2014, 487, 226–233. [CrossRef]
113. Mohammadi, M.; Najafpour, G.D.; Younesi, H.; Lahijani, P.; Uzir, M.H.; Mohamed, A.R. Bioconversion of
synthesis gas to second generation biofuels: A review. Renew. Sustain. Energy Rev. 2011, 15, 4255–4273.
[CrossRef]
Catalysts 2018, 8, 2 19 of 19

114. Choi, D.W.; Chipman, D.C.; Bents, S.C.; Brown, R.C. A techno-economic analysis of polyhydroxyalkanoate
and hydrogen production from syngas fermentation of gasified biomass. Appl. Biochem. Biotechnol. 2010, 160,
1032–1046. [CrossRef] [PubMed]
115. Bredwell, M.D.; Srivastava, P.; Worden, R.M. Reactor Design Issues for Synthesis-Gas Fermentations.
Biotechnol. Prog. 1999, 15, 834–844. [CrossRef] [PubMed]
116. Madhukar, G.R.; Elmore, B.B.; Huckabay, H.K. Microbial conversion of synthesis gas components to
useful fuels and chemicals, Microbial Conversion of Synthesis Gas Components to Useful Fuels and
Chemicals Symposium. In Seventeenth Symposium on Biotechnology for Fuels and Chemicals; Wyman, C.E.,
Davison, B.H., Eds.; ABAB Symposium, Volume 57/58; Humana Press: Totowa, NJ, USA, 1996.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like