You are on page 1of 10

Industrial Crops and Products 106 (2017) 21–30

Contents lists available at ScienceDirect

Industrial Crops and Products


journal homepage: www.elsevier.com/locate/indcrop

Exploiting the potential of gas fermentation


Stephanie Redl a,1 , Martijn Diender b,1 , Torbjørn Ølshøj Jensen a , Diana Z. Sousa b ,
Alex Toftgaard Nielsen a,∗
a
Novo Nordisk Foundation Center for Biosustainability, Technical University of Denmark, Hørsholm, Denmark
b
Laboratory of Microbiology, Wageningen University, Wageningen, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: The use of gas fermentation for production of chemicals and fuels with lower environmental impact is a
Received 15 July 2016 technology that is gaining increasing attention. Over 38 Gt of CO2 is annually being emitted from industrial
Received in revised form 20 October 2016 processes, thereby contributing significantly to the concentration of greenhouse gases in the atmosphere.
Accepted 8 November 2016
Together with the gasification of biomass and different waste streams, these gases have the potential for
Available online 19 November 2016
being utilized for production of chemicals through fermentation processes. Acetogens are among the most
studied organisms capable of utilizing waste gases. Although engineering of heterologous production of
Keywords:
higher value compounds has been successful for a number of acetogens, the processes are challenging due
Syngas fermentation
Biomass gasification
to the redox balance and the lack of efficient engineering tools. In this review, we address the availability
Co-cultures of different gaseous feedstock and gasification processes, and we focus on the advantages of alternative
Mixotrophy fermentation scenarios, including thermophilic production strains, multi-stage fermentations, mixed
Acetogens cultures, as well as mixotrophy. Such processes have the potential to significantly broaden the product
Thermophiles portfolio, increase the product concentrations and yields, while enabling the exploitation of alternative
and mixed feedstocks. The reviewed processes also have the potential to address challenges associated
with product inhibition and may contribute to reducing the costs of downstream processing. Given the
widespread availability of gases, such processes will likely significantly impact the transition towards a
more sustainable society.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction energy and carbon source. Acetogenic bacteria are the most stud-
ied and have the greatest industrial potential, making use of the
There is an increasing demand for processes that reduce carbon- Wood-Ljungdahl pathway (WLP) to convert CO2 (and CO). Com-
emissions and ensure carbon neutral and sustainable production pared to the Calvin-Benson-Bassham cycle, which is used by plants,
of energy and commodities for the steadily growing population algae, cyanobacteria, purple bacteria, and some proteobacteria, the
(Pachauri et al., 2014). Previous advances in the production of first WLP is a highly energy efficient CO2 fixation pathway (Hawkins
generation biofuels have raised the feed vs. fuel debate. A promising et al., 2013). Microorganisms employing the WLP are therefore
technology that has gained increasing attention within recent years relevant as biotechnological platforms for the production of biofu-
is gas fermentation, a process in which microorganisms anaerobi- els and biochemicals from one-carbon compounds, and potentially
cally convert a gaseous substrate into biofuels and biochemicals. decrease our dependence on fossil resources.
Several microorganisms have the ability to utilize CO2 and CO as Metabolic traits of gas-fermenting bacteria have been reviewed
recently (Daniell et al., 2016; Dürre and Eikmanns, 2015; Latif et al.,
2014). The scope of this article is to assess the potential of gas fer-
Abbreviations: 3HP, 3-hydroxypropionic acid; 4HB, 4-hydroxybutyrate; ATP,
mentation and ways to exploit this potential. We would like to draw
adenosine triphosphate; BDO, butanediol; GHG, greenhouse gas; MEK, methyl ethyl attention to alternative production scenarios, including multi-stage
ketone; MSW, municipal solid waste; NADH/NAD+, nicotinamide adenine dinucleo- processes, co-cultures, mixotrophy, and thermophilic production
tide; NADPH/NADP+, nicotinamide adenine dinucleotide phosphate; NETL, National strains. To date, pure cultures of mesophilic strains are deployed,
Energy Technology Laboratory; PFOR, pyruvate:ferredoxin oxidoreductase; PHA,
with a focus mainly on ethanol, and on 2,3-BDO production. The
polyhydroxyalkanoates; VFA, volatile fatty acids; VSS, volatile suspended solids;
WLP, Wood-Ljungdahl pathway. aforementioned alternative production scenarios on the contrary
∗ Corresponding author. would broaden the spectrum of products that can be produced from
E-mail address: atn@biosustain.dtu.dk (A.T. Nielsen). CO and CO2 . Additionally, it would be possible to explore combina-
1
These authors contributed equally.

http://dx.doi.org/10.1016/j.indcrop.2016.11.015
0926-6690/© 2016 Elsevier B.V. All rights reserved.
22 S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30

tions of different industrial feedstock streams (gas and sugar) and cement kiln waste gas increases from <50% CO2 (air-fired) to up to
to take advantage of variable process conditions. 100% when being O2 -fired (International Energy Agency, 2014).

2.2. Gasification of low-value carbonaceous materials to syngas


2. Feedstock availability
The availability of substrate for gas fermentation is broadened
The present review focuses on the conversion of CO and CO2
immensely when considering the amount of feedstock that can be
as carbon source by gas fermenting microorganisms. Hereby, CO
converted into carbon- and energy-rich gas streams via gasifica-
serves as carbon source and electron donor. When CO2 serves as
tion. Gasification is defined as the “thermo-chemical conversion
sole carbon source, an additional electron donor is required. Below,
of carbonaceous feedstock to gaseous products through a partial
we describe some of the sources of CO, CO2 , and electron donors,
oxidation process at elevated temperatures” (Mohammadi et al.,
as well as their industrial availability.
2011). Besides fossil fuels, there is a broad range of more sustainable
options: lignocellulosic energy crops such as willow, switchgrass,
2.1. Off-gases from industry, heat and energy generation etc. can serve as feedstock for syngas production. The use of lig-
nocellulosic energy crops has the advantage that their prices are
CO- and CO2 -rich waste gases are an attractive substrate for gas “more stable as they only participate in the energy market” (Daniell
fermentation. Many industrial processes produce large amounts of et al., 2016). Also algae is an abundant biomass that is suitable as
carbon-rich gases that are often left unused, thereby contribut- feedstock for gasification (Azadi et al., 2015).
ing to elevated concentrations of CO2 and CO in the atmosphere. Another option is lignocellulosic biomass waste, which is not
In 2011, about 23% of the total CO2 emissions were derived from suitable for food production or starch- and sugar-based produc-
industrial processes (van der Hoeven, 2013), the second largest tion of chemicals and biofuels. The impact of this option is even
sector contributing to CO2 emissions, after electricity- and heat- more significant when taking into account that the lignin-fraction
generation installations. More than 40% of the CO2 emissions in is not utilized in sugar-based production (both 1st and 2nd gen-
2011 were derived from generation of electricity and heat (van der eration production). In wheat straw the lignin content is around
Hoeven, 2013), which worldwide relies heavily upon coal combus- 20% (Sheldon, 2014), and can be as high as 44.5% in some woody
tion (Gutmann, 2014). Overall, the anthropogenic CO2 emissions biomass (Vassilev et al., 2012). This kind of lignin-rich feedstock,
account to 38 Gt/year (Edenhofer et al., 2014). suitable for gasification, accumulates as agricultural residues or as
Industrial processes produce CO2 emissions through chemical forestry by-products. Crop production (for food, feed, or produc-
reactions that do not involve combustion, of which the following tion of 1st generation biofuels), generates large amounts of residue,
three sub-sectors are the main-contributors: iron and steel (27%), with a residue/crop ratio of 1:1.4 for conventional crops (Kim and
non-metallic minerals (27%), and chemicals and petrochemicals Dale, 2004). Another interesting source for lignocellulosic biofuels
(16%) (International Energy Agency, 2007). For example, 60% of include residues from 2nd generation biofuels production. In the
the CO2 emissions from cement production come from inevitable process of ethanol production from corn stover, for example, 5.9-
chemical reactions in the process (Cement Sustainability Initiative, times (by mass) more lignocellulosic residues are generated than
2014). Those emissions cannot be prevented by heat and energy ethanol (McAloon et al., 2000).
generation with renewables, thus alternative strategies for reduc- Municipal solid waste (MSW) or industrial waste could also
ing GHG emission are required. According to the world steel serve as feedstock for gasification. A total of 251 × 106 t of MSW was
association, 1.7 × 109 t of crude steel were produced worldwide in generated in the US in 2012 after annual increases during the last
2014 (World Steel Association, 2015a) and 1.9 tons CO2 are emit- decades (U. S. Environmental Protection Agency, 2014). Especially
ted per ton crude steel produced (World Steel Association, 2015b), those parts of the complex waste that are of organic origin possess
which accounts to an annual CO2 emission of 3.2 × 109 t. In con- great potential as starting material for gasification. Although their
clusion, large amounts of carbon rich off-gases are available and carbon content is high, with 13% by mass (Staley and Barlaz, 2009),
their conversion into chemicals and fuels has the potential to sig- they are currently left unrecycled. In particular, the biodegrada-
nificantly decrease GHG emissions (Handler et al., 2015; Ou et al., tion of the aforementioned waste fraction has been identified as
2013). one of the main challenges for direct exploitation (Drzyzga et al.,
Off-gases from electricity and heat generation, as well as indus- 2015). Additionally, sludge from waste water treatment could act
trial waste gases are supposedly a substrate that comes free of as potential feedstock for gasification, but is of lesser relevance,
charge. Currently, gas fermentation processes closest to commer- since its carbon-content is relatively low (Drzyzga et al., 2015).
cialization are based on industrial waste gases (LanzaTech, 2016). The feedstock requirements for gasification technologies are
However, not all waste gases are equally suitable for microbial gas generally considered flexible (Daniell et al., 2016). However, there
fermentation, since there are demands with regards to the conti- are certain requirements, for example with regards to the moisture
nuity of the gas stream, the carbon content, as well as the purity of content of the feedstock (Piccolo and Bezzo, 2009). Additionally, the
the gas. technology of large-scale gasifiers restricts the feed rate to around
The CO and CO2 content of the off-gas is dependent on the sec- 2000 t per day (Griffin and Schultz, 2012). There are different gasifi-
tor, but is also heavily dependent on process parameters and can cation technologies available, and fluidized bed gasification is most
therefore vary between production sites. For example, the off-gas suitable for large scale gas production, when taking throughput,
from power plants contains only 3–4% (gas-fired) to 13–14% (coal- costs, complexity, and efficiency into account (Alauddin et al., 2010;
fired) CO2 , but process improvements such as chemical looping Mohammadi et al., 2011). The efficiency of conventional biomass
combustion and oxyfuel-technology (O2 -fired instead of air-fired) gasification, when comparing the lower heating value of the pro-
can increase the power-efficiency and the CO2 -content of the off- duced syngas with that of the gasification feedstock, is around
gas (International Energy Agency, 2014). The off-gases of other 85% for biomass and coal (Ptasinski, 2008). Currently, advances
industrial processes contain high percentage of CO2 , for exam- in biomass gasification technologies are made (Heidenreich and
ple the production of ethylene oxide emitting nearly 100% CO2 Foscolo, 2015), thus more efficient gasification technologies are
(International Energy Agency, 2014). As it is the case for electricity likely to emerge.
and heat generation, the off-gas composition can be greatly influ- The syngas composition depends on the feedstock (Tiquia-
enced by the process parameters: for example, the CO2 content of Arashiro, 2014), but can be greatly influenced by the gasification
S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30 23

Table 1
Theoretical yields for the production of some compounds of interest from CO, CO2 /H2 , and CO/H2 respectively by C. ljungdahlii.

CO CO2 /H2 CO/H2

mol/mol CO g/g CO mol ATP/mol producta mol/mol CO2 g/g CO2 mol ATP/mol producta mol/mol CO g/g CO mol ATP/mol producta

acetate 0.250 0.527 1.13 0.500 0.671 0.125 0.5 1.054 0.625
acetone 0.125 0.259 1.25 0.333 0.440 −0.750 0.25 0.518 0.250
ethanol 0.167 0.274 1.13 0.500 0.523 −0.375 0.5 0.822 0.125
i-propanol 0.111 0.238 1.75 0.333 0.455 0.500 0.25 0.536 0.500
butanol 0.083 0.221 2.75 0.250 0.421 −0.250 0.25 0.662 0.750
a
The estimated energy yield is displayed taking into account the minimal energy yield of the C. ljungdahlii metabolism (Diender et al., 2015; Schuchmann and Müller,
2014).

technology and gasification parameters used (Drzyzga et al., 2015; photosynthesis (Lovley and Nevin, 2013). However, electrosynthe-
McKendry, 2002). Syngas typically contains 40–50% N2 , 15–20% sis technology is still awaiting large-scale demonstration.
H2 , 10–15% CO, 10–15% CO2 and 3–5% CH4 (McKendry, 2002). The Hydrogen can serve as an electron donor to enable fixation
syngas composition subsequently can be shifted towards a cer- of CO2 for many microorganisms. In addition to the natural
tain composition by gas cleaning and reforming, such as (reverse) hydrogen content of syngas, there are several processes that can
water-gas-shift or gas drying. lead to the production of H2 . For instance, H2 can be produced
In mid-2016, 95 active gasification plants are operating world- by hydrogenogenic microorganisms (Das and Veziroǧlu, 2001).
wide (23 in the USA and 72 in the rest of the world) according Another option is production of H2 by electrolysis of water, and
to the NETL gasification plant database of the U.S. Department of technologies of several companies have reached commercial levels
Energy. However, only 16 plants utilize waste or biomass to date, (Bertuccioli et al., 2014). Already about 4% of the global hydrogen
the rest are still petro-fueled. The produced syngas is not yet uti- production is derived from water electrolysis (Gandía et al., 2013).
lized in gas fermentation processes, but is mainly channeled to the The required electricity could be generated by renewables, espe-
chemical industry for production of fuels, fertilizer, etc. (National cially wind power and photovoltaics, to allow energy storage in
Energy Technology Laboratory, 2016). The technology for biomass form of hydrogen during times of excess energy production.
gasification for syngas production exists, but economic incentives, In conclusion, there is a substantial amount of gaseous sub-
as well as the necessary infrastructure are required (Roddy, 2013). strates, electron sources, and feedstock for gasification available.
Until now, only few economic analyses have been published in Making use of the available resources holds a promise for chem-
the scientific literature with respect to fermentation of biomass- ical and fuel production with reduced carbon emission and full
derived syngas (Choi et al., 2010; Piccolo and Bezzo, 2009; Spath exploitation of available resources such as biomass and renewable
and Dayton, 2003). energy. However, the potential of gas fermentation and its role in
the transition towards a bio-based economy can only be exploited if
the processes are also economically attractive. Hence, the gaseous
substrates have to be converted by the microbial production organ-
2.3. CO and H2 as alternative carbon and electron sources in isms to a product of sufficiently high value.
microbial processes

Beyond the utilization of off-gases from industry or syngas gen- 3. Status of current gas fermentation processes
erated by gasification of certain feedstock, there are alternative
ways to supply the microorganisms with carbon and electrons. Due Bacteria fermenting CO2 - and CO-rich gases have a range of
to its high energy content, CO is in general preferred by gas ferment- natural products which are mainly of interest as biocommodities
ing microorganisms over CO2 and H2 . Besides reverse water-gas or biofuels, and some are already on their way to being commer-
shift reaction, some other methods make it possible to generate cialized. Acetate is the typical product of acetogenic bacteria and
CO from CO2 . For example, light-driven processes can be used to advances have been made to improve the production of natural
perform splitting of CO2 into CO (Service, 2009; Woolerton et al., acetate producers by genetic modification (Straub et al., 2014).
2010). Butyrate and 2,3-BDO are other natural products of gas ferment-
In microbial fermentation of C5 and C6 sugars, the substrate ing bacteria that are of interest as biocommodity (Bengelsdorf et al.,
is reduced to CO2 . Therefore microbial processes such as pro- 2013; Cho et al., 2014; Köpke et al., 2011). However, syngas fermen-
duction of 1st and 2nd generation biofuels provide an additional tation processes closer to commercialization generally aim at the
source of CO2 . The fermentation off-gases contain nearly 100% CO2 production of biofuels. Multiple acetogenic bacteria can naturally
(International Energy Agency, 2014). For fermentation of C5/C6 produce ethanol (Abubackar et al., 2011) or butanol (Dürre, 2016).
sugars, the theoretical mass yield of the conversion is 0.51 g ethanol Some other natural products of potential industrial interest have
and 0.49 g CO2 per g sugar. been described, such as 2-oxobutyrate (Daniell et al., 2016), hex-
As mentioned before, fixation of CO2 by microorganisms anol (Daniell et al., 2016), polymers and polymer building blocks
requires an additional electron source. The concept of photosyn- (Drzyzga et al., 2015), PHA (Brown, 2007), 3HP/4HB (Hawkins et al.,
thesis, where electrons are made available by induction with light, 2013), 5-aminovulinic acid, and vitamin B12 (Wiegel, 1994). Table 1
can also be exploited for gas fermentation as cyanobacteria are shows the theoretical production yields for a range of relevant
naturally fixing CO2 through photosynthesis. Recently, Sakimoto compounds, and significant production of certain biochemicals
et al. described a method for artificially photosensitizing the ace- have already been achieved by gas fermentation as summarized
togen Moorella thermoacetica. In this hybrid approach, the cells in Table 2.
were loaded with cadmium sulfide nanoparticles, which enabled Generally, the natural products that can be produced with
acetate production via photosynthesis (Sakimoto et al., 2016). higher yields have relatively low commercial value. This, combined
Some microorganisms can also directly accept electrons from an with relatively high costs of fermentation and gasification reactors,
electrode for the reduction of fixed CO2 , in a process called elec- makes such processes less attractive when compared to conven-
trosynthesis. The process is described to be far more efficient than tional production based on fossil resources or on 1st generation
24 S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30

Table 2
Examples of volumetric productivity, titer, and yield of some important gas fermentation target products.

Volumetric Titer Yield References Comments


productivity

Acetate 479 mmol/l/day 833 mM n.a. (Straub et al., 2014) A. woodii (ME), controlled batch
7.92 mmol/l/day 10 mM n.a. (Kim et al., 2016) Multistage process of
carboxydotrophic, hydrogenogenic
culture converting CO into H2 /CO2 ,
off-gas in second reactor converted
mainly to acetate by the acetogen T.
kivui
11.92 mmol/l/day 97.7 mM n.a. (Xu et al., 2015) Acetogenic enrichment culture grown
on H2 /CO2

Ethanol 193 mmol/l/day 450 mM 0.14 mol/mol (Richter et al., 2013) C. ljungdahlii ERI-2 Yield calculated
CO from 28% of introduced CO present in
carbon of ethanol
0.25 mmol/l/day 5.5 mM n.a. (Xu et al., 2015) Acetogen enrichment culture grown on
H2 /CO2

Butanol n.a. 25.6 mM n.a. (Köpke and Liew, 2012) C. autoethanogenum (GM)
3.5 mmol/l/day 6 mM 0.074 mol/mol (Diender et al., 2016) Co-culture of C. kluyveri and C.
CO autoethanogenum
2.7 mmol/l/daya 8 mM 0.0080 mol/mol (Ramió-Pujol et al., C. carboxydivorans P7 at 25 ◦ C
CO 2015)

Hexanol 2 mmol/l/day 4 mM 0.05 mol/mol (Diender et al., 2016) Co-culture of C. kluyveri and C.
CO autoethanogenum
2.1 mmol/l/daya 5 mM 0.0038 mol/mol (Ramió-Pujol et al., C. carboxydivorans P7 at 25 ◦ C
CO 2015)

Acetone 10.9 mmol/l/day 52 mM n.a. (Hoffmeister et al., A.woodii (GM)


2016)

Butyrate 227 mmol/l/day 220 mM n.a. (Vasudevan et al., Chain-elongating open mixed culture
2014) fed with effluent of reactor syngas
fermenting C. ljungdahlii ERI-2
8.5 mmol/l/day 11 mM 0.136 mol/mol (Diender et al., 2016) Co-culture of C. kluyveri and C.
CO autoethanogenum
6.3 mmol/l/daya 6 mM 0.0135 mol/mol (Ramió-Pujol et al., C. carboxydivorans P7 at 25 ◦ C
CO 2015)

Caproate 14.7 mmol/l/day 8.6 mM n.a. (Vasudevan et al., Chain-elongating open mixed culture
2014) fed with effluent from reactor with
syngas fermenting C. ljungdahlii ERI-2
2.5 mmol/l/day 6 mM 0.074 mol/mol (Diender et al., 2016) Co-culture of C. kluyveri and C.
CO autoethanogenum
4.6 mmol/l/daya 6 mM 0.0073 mol/mol (Ramió-Pujol et al., C. carboxydivorans P7 at 25 ◦ C
CO 2015)
a
Converted from mmol/g protein/h to volumetric production rates using protein concentrations and maximal production rates reported.

bio-production (Piccolo and Bezzo, 2009; Wei et al., 2009). There- been reported for acetone (Becker et al., 2009; Hoffmeister et al.,
fore, attempts of producing higher value compounds from CO2 and 2016), n-Butanol (Dürre, 2016; Köpke et al., 2010), butyrate (Ueki
CO gases with genetically engineered strains are being pursued. et al., 2014), isoprene (Beck et al., 2013; Chen et al., 2013), iso-
Two main approaches can be taken: engineering of an autotrophic propanol (Köpke et al., 2012), and MEK (Mueller et al., 2013). The
strain to produce a heterologous compound, or engineering of a recently published review by Liew et al. gives an excellent overview
heterotrophic producer to utilize gaseous substrates. Initial suc- of the advances in genetic engineering of syngas-fermenting bac-
cess to the latter approach, conferring autotrophy on heterotrophic teria (Liew et al., 2016). Cueto-Rojas et al. gives an overview of the
host organisms, has been achieved within the recent years. Engi- potential products which can be formed from gaseous (and non-
neering of CO2 fixation in a heterotroph was reported for the first gaseous) substrates when taking the Gibbs free energy content per
time when the 3HP/4HB pathway was engineered into a hyperther- electron into account (Cueto-Rojas et al., 2015).
mophilic host, Pyrococcus furiosus, for the production of n-butanol
from H2 /CO2 (Keller et al., 2013). Efforts have also been made to
4. Alternative gas fermentation scenarios
engineer the WLP for CO into E. coli (Burk et al., 2009), however, only
low rates have been reported (Daniell et al., 2016). Antonovsky et al.
Progress is being made in the field of genetic modification
(2016) recently published a hybrid approach of rational metabolic
of gas fermenting microorganisms. However, pathways (natural
engineering and laboratory evolution. The evolved E. coli strain
or heterologous) have to make it possible to harvest the energy
employs a fully functional Calvin-Benson-Bassham cycle to fix CO2 ,
that can be generated by product formation. The metabolism of
while energy and reduction equivalents are generated by another
syngas-fermenting bacteria operates at a thermodynamic limit
supplied compound. The developed strain has been the first report
(Schuchmann and Müller, 2014) and the ATP yield for a given
of a fully functional carbon fixation cycle in a heterologous host
product is dependent on the production strain (Bengelsdorf et al.,
(Antonovsky et al., 2016).
2013; Bertsch and Müller, 2015). The expression of heterolo-
More research is focused on the engineering of gas-fermenting
gous pathways therefore often disturbs the energy balance of the
microorganisms. Successful heterologous production in Clostridium
cell. To fully exploit the potential of gas fermentation, alternative
autoethanogenum, C. ljungdahlii, and Acetobacterium woodii has
approaches such as multi-stage fermentation, mixed fermentation,
S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30 25

co-cultivation, mixotrophy, and thermophilic production strains pH 7 and lowering the pH significantly decreases their efficiency
therefore need to be considered. (Ganigué et al., 2016). Vasudevan and co-workers have shown that
effluent from a syngas reactor, containing acetate and ethanol,
4.1. Multi-stage fermentation could be used to feed a reactor performing chain elongation, indi-
rectly obtaining elongated acids from syngas (Vasudevan et al.,
As the formation of products from gas fermentation is influenced 2014). A pure culture of C. ljungdahlii strain ERI-2 was grown in
by environmental factors and interesting products are mainly the first stage in order to ferment the syngas, while a mixed cul-
produced under conditions non-optimal for growth, high concen- ture obtained from a chain-elongating reactor was used in the
trations of desired products can often not be obtained in a single second part of the system. The separation of these two processes
processing step. Therefore, applying multiple processing stages allows both systems to function at optimal conditions, obtaining
could broaden the portfolio of products obtained from syngas fer- production rates for butyrate and caproate of 20 and 1.7 g/l/day,
mentation together with higher end-product concentrations. One respectively.
example is the combination of fermentation of easily accessible For the production of longer chain lipids (C16-C18) from syn-
C5- and C6-compounds of the lignocellulosic biomass with syn- gas, Hu and coworkers constructed a two-stage system (Hu et al.,
gas fermentation. The residual lignin-rich biomass, not converted 2016). This was done by growing M. thermoacetica in the first stage
via sugar fermentation, can be gasified, making previously inac- on syngas, mainly producing acetate as the final product. The out-
cessible carbon available for syngas fermenters. Additionally, as going broth was fed into an aerobic reactor containing a genetically
mentioned above, CO2 off-gas from the sugar fermentation might engineered Yarrowia lipolytica, a widely used oleaginous yeast.
be introduced to the gas fermentation, reducing waste emission The system accumulated up to 18 g/l C16-C18 lipids at a rate of
and increasing the overall yield of the process. This process can be 0.19 g/l/h. This system nicely shows how carbon from syngas can be
seen as an upgraded version of the sugar platform accessing the full fixated into more complex products by using a multi-stage system.
range of carbon present in the initial material. In theory, such a two-stage system would allow for production of
Syngas fermentation by acetogens generally results in a mix- other complex products as the second reactor can contain different
ture of both alcohols and acids under optimal growth conditions genetically altered strains.
(Bertsch and Müller, 2015). Lowering pH and decreasing the con- Not only biological processes can be coupled for production
centration of yeast extract has shown to move the spectrum of bio-based chemicals. The integration of bio-based processes
towards alcohol production (Abubackar et al., 2012). In addition, with chemical processes enables the production of chemicals
many microorganisms cannot tolerate the presence of CO at high rarely obtained from biological systems. A collaboration of Lan-
concentrations, requiring partly removal of the toxic gas in order zaTech with Invista and SK Innovation proposes production of
to operate efficiently. To produce interesting amounts of one type 1,3-butadiene via a two-step process involving a biological and
of product, and to do this efficiently, the fermentation process can chemical step (Green Chemicals Blog, 2014; Invista, 2015). Butadi-
be performed in multiple stages. ene precursors, 2,3-BDO and 1,3-BDO are generated biologically via
In a study by Kim et al. (2016), acetate was produced as syngas fermentation. Subsequently these products can be thermo-
an end-product in a multistage fermentation process using steel catalytically dehydrated to form butadiene. Additionally, there are
mill off-gas as feedstock. In the first part of the process the attempts to make production of butadiene a one-step process,
hydrogenogenic Thermococcus onnurineus converted a large part of involving an engineered biological catalyst. Both the one- and two-
the CO fraction in the gas to H2 and CO2 . The resulting H2 /CO/CO2 step process are still in the development stage, but these initiatives
mixture was subsequently fed into a reactor containing the homo- show the possibilities for combining different processes in order to
acetogenic Thermoanaerobacter kivui, allowing for generation of expand the scope of syngas derived products.
acetate as the main end-product at a maximal volumetric pro-
duction rate of 0.33 mmol/l/h. Absence of the initial biological 4.2. Mixed cultures and defined co-cultures
gas-treatment step caused T. kivui not to grow in the mentioned
study due to CO inhibition (Kim et al., 2016). Recently, T. kivui The utilization of open or defined mixed cultures has some
was shown to be capable of adapting to growth on CO up to advantages over pure culture approaches (Drzyzga et al., 2015):
100% headspace composition, but this took several transfers with (i) microbial interactions can result in the production of different
increasing CO pressures (Weghoff and Müller, 2016). final products, (ii) some strains may detoxify the environment from
A multistage approach can also be used for the production of a toxic compound or reduce the concentration of a certain prod-
ethanol, generating acetate in the first step, followed by solven- uct allowing for a more efficient conversion of the gas or increased
togenesis in the second system (Grethlein and Jain, 1992). Richter product yield, (iii) cultures are more robust and resilient to changes
et al. (2013) designed a two stage system for producing ethanol in environmental conditions, (iv) less prone to contamination by
from syngas using C. ljungdahlii. The first reactor was employed foreign bacteria or collapse due to phage infection. Trials for syngas
using acetogenic conditions, producing acetate as the main end- conversion by open mixed cultures resulted mainly in the produc-
product. The broth was subsequently fed to a second fermenter, tion of methane, hydrogen, or acetate.
which employed a lower pH causing acetate to be converted to In general, methanogens are poor CO utilizers (Daniels et al.,
ethanol. An ethanol production rate of 0.37 g/l/h was achieved with 1977; Diender et al., 2015), and grow slowly on the substrate.
this system with maximal concentrations up to 450 mM (Richter In mixed cultures it was observed that methane could be pro-
et al., 2013). Carbon and hydrogen recovery in the ethanol end- duced indirectly from CO. Under thermophilic conditions, CO was
product were 28% and 74%, respectively. Performance of different first converted to hydrogen, subsequently acting as substrate for
clostridial strains were tested in this system showing C. ljungdahlii methanogenesis. Under mesophilic conditions it was observed that
PETC as the most optimal producer in this case (Martin et al., 2016). CO was mainly converted to acetate, which subsequently acted as
In addition to production of relatively simple products such as substrate for methanogenesis. (Sipma et al., 2003, 2004). Grow-
acetate and ethanol, longer chain VFAs could be produced via a sim- ing methanogens in a mixed culture containing carboxydotrophs,
ilar multistage system. These VFAs are produced via the process such as acetogens and hydrogenogens, does not only enable pro-
of chain elongation, in which shorter chain VFAs are converted to duction of intermediate substrates, but also for the removal of CO
longer chain VFAs via reverse ␤-oxidation (Angenent et al., 2016). as a toxic compound. Open mixed cultures not adapted to growth
Strains performing this metabolism often grow optimally around on CO were shown to be able to convert syngas to biogas (Guiot
26 S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30

et al., 2011). After enrichment and adaptation of anaerobic sludge, gives an overview of achieved key production data of some relevant
a CO conversion efficiency of 75% was obtained. The overall sys- compounds by gas fermentation processes.
tem was mass transfer-limited, but a specific activity of 20 mmol The use of mixed and co-cultures shows that microbial inter-
CH4 /gVSS/day is expected to be achieved under thermophilic con- actions allow for the production of different end-products. Due
ditions. Complementation of traditional anaerobic digestion with to competition for substrate and complexity of mixed communi-
gasification of the non-degraded fraction and subsequent fermen- ties, co-cultures are in theory more efficient and easier to control.
tation of syngas could result in a significant increase in biogas yield, However, it remains to be assessed whether co-cultures are robust
which for municipal waste can be up to six times higher than in enough to maintain over long periods in continuous systems. An
traditional anaerobic digestion (Guiot et al., 2011). However, the alternative use for co-cultures would be to act as bio-augmentation
additional energy investment for the gasification may reduce the in open non-sterile systems. However, further study is required to
energy efficiency of the overall process. assess feasibility of such an approach.
In the work of Alves et al. (2013), a thermophilic enrich-
ment produced mainly H2 and subsequently acetate. Predominant 4.3. Mixotrophy
microorganisms in this culture included Desulfotomaculum, Ther-
mincola, and Thermoanaerobacter (Alves et al., 2013). Another study Some organisms are able to use two or more carbon/electron
on CO conversion to H2 by mixed cultures was performed by sources simultaneously, as observed in some phototrophs using
Liu et al. (2016). Here methanogens were inhibited with chloro- both CO2 and organic substrates at the same time for growth. Dur-
form in order to obtain production of mainly hydrogen (Liu et al., ing mixotrophic growth with sugars, electrons, and CO2 released
2016). CO conversion to hydrogen was 91% in a continuous reac- from the fermentation can be fixed via the WLP. Fast et al. (2015)
tor operated with gas recycling, with an average production rate of described an anaerobic, non-photosynthetic, mixotrophic fermen-
6.1 mmol/gVSS/day. tation in which 2 CO2 and 8 electrons from glycolysis are used in
Acetate production from syngas by mixed cultures has also been the WLP (pathway coupling), generating 3 acetyl-CoA instead of
pursued, obtaining concentrations of up to 23 g/l at a production the 2 acetyl-CoA normally generated. This causes the fermentation
rate of 0.72 g/l/day, and a maximal yield of 0.663 gacetate /gCO (Nam to have a net CO2 exhaust of zero. Despite that the pathway from
et al., 2016). Also mixtures of acetate and ethanol could be obtained glycolysis to acetyl-CoA yields 8 electrons (the amount required in
by mixed cultures (Xu et al., 2015); however, their final concentra- the WLP), an additional electron donor can be required as the gly-
tions being lower compared to other approaches. In both studies, colysis mainly yields low redox potential electrons (mainly in the
pH was found to be of major influence on the production capacity of form of NADH, −340 mV). Only when a pyruvate ferredoxin oxi-
the mixed cultures, being optimal at pH 7. Recently, Ganigué et al. doreductase (PFOR), generating reduced ferredoxin, is present, all
(2016) studied formation of higher alcohols by open mixed cultures 8 electrons can be stoichiometrically used in the WLP (Table 3). In
fed with syngas. Chain elongated products were generated in these other cases the more reduced electrons have to be derived from H2
mixed cultures at higher pH while higher alcohols were formed at (−414 mV) via bifurcation processes or directly from CO (−520 mV)
lower pH, showing that pH is an essential parameter for produc- in order to complete the WLP (Table 3). Mixotrophic growth with
tion of acids or alcohols in a mixed culture (Ganigué et al., 2016). CO as additional substrate could also result in higher yields of
Higher alcohols were also produced via mixed cultures by feeding more reduced products. Therefore, depending on the degree of
syngas and longer chain VFAs (Liu et al., 2014). Carboxydotrophic reduction of the final end-product, the respective electron donor
bacteria, such as C. ljungdahlii, are able to convert carboxylic acids is either generated or has to be supplied (Fast et al., 2015). One
into their corresponding alcohols, using syngas as electron donor. of the challenges in mixotrophic cultivation is to prevent catabo-
Propionic acid, n-butyric acid, isobutyric acid, n-valeric acid, and n- lite repression, causing one of the substrates not to be used in
caproic acid are converted into n-propanol, n-butanol, isobutanol, the presence of the other (Görke and Stülke, 2008). For the use of
n-pentanol, and n-hexanol (Perez et al., 2013). sugars and H2 /CO2 or CO, both the WLP and the glycolysis need
Mixed cultures are more robust than pure cultures, but their to be active to have the desired effect. Toxic properties of CO
complexity and potential competition within the system often might stimulate mixotrophy as its oxidation by the microorgan-
results in a poor understanding of these systems and in a low yield isms can prevent inhibition, and might be necessary for preserving
of desired products. An alternative is to use defined synthetic co- other catabolic routes active. An indirect application of the con-
cultures, which are less complex while still having the benefits from cept of mixotrophy is the co-utilization of lower value products
microbial interactions. Defined co-cultures are currently largely from syngas-fermenting bacteria with sugars. Chotani et al. (2012)
applied in food industrial fermentation processes, such as produc- described for example a method to enhance isoprene production in
tion of dairy products and alcoholic beverages (Bader et al., 2010). E. coli by co-utilization of acetate and glucose (Chotani et al., 2012).
In the case of syngas fermentation, Diender et al. (2016) described
a co-culture of C. autoethanogenum and C. kluyveri, capable of pro- 4.4. Overview of products from alternative gas fermentation
ducing butyrate, caproate, and their respective alcohols from CO as scenarios
substrate. Butyrate, caproate, butanol, and hexanol were produced
at rates of 8.5, 2.5, 3.5, and 2.0 mmol/l/day, respectively. Other Fig. 1 gives an overview of the product range that can
configurations of defined co-cultures and bio-augmentation of be achieved with multistage fermentation, mixed cultures, and
open-culture systems might stimulate production of different bio- defined co-cultures in comparison to the products of pure cul-
chemicals from syngas. Defined co-cultures have advantages over tures of natural and engineered hosts. Considering the approaches
open cultures, which have to deal with competition for substrate involving non-engineered syngas fermenters, multistage fermenta-
and potential removal of interesting products. Such competition tion is currently the most suitable option for generating a relatively
was observed in a sulfate reducing system where methanogens, broad scope of products from gas waste streams. Especially options
sulfate reducers, and acetogens competed for utilization of H2 , combining the conversion of syngas by native organisms with the
resulting in lower efficiency of the overall system (Sipma et al., generation of products by genetically engineered strains using the
2006). Next to applications in food industry, patents have been products of the syngas fermenters are interesting. These systems
filed on the use of co-cultures in applications used for butanol pro- do however lack microbial interaction, which may be necessary
duction from syngas (Datta and Reeves, 2013) and production of for the production of different products like higher alcohols. An
biofuels using photosynthetic co-cultures (Contag, 2015). Table 2 avenue to explore is the utilization of mixed and co-cultures for
S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30 27

Table 3
Stoichiometric conversion of different combinations of electron donors is displayed for the formation of acetyl-CoA or acetate.

Pathway Electron donor CO2 generated Acetyl-CoA formation Acetate formation

Number of molecules ATP yielda Number of molecules ATP yielda

WLP 4 CO 2 1 0.125 1 1.125


4 H2 −2 1 −0.875c 1 0.125c
2 H2 + 2 COb 0 1 −0.75c 1 0.625c
Glycolysisd C6 sugar 2 2 2 2 4
Glycolysis + WLP (in presence of PFOR) C6 sugar 0 3 1.125 3 4.125
Glycolysis + WLPe (H2 ) C6 sugar + 28 H2 −14 10 −6 10 4
Glycolysis + WLPe (CO) C6 sugar + 4 CO 2 4 0.25 4 4.25
a
For the WLP, energy yields can vary strongly per organism, depending on the type of enzymes present and final end-product. Here the estimated energy yield is displayed
taking into account the minimal energy yield of the C. ljungdahlii metabolism (Diender et al., 2015; Schuchmann and Müller, 2014).
b
Data is displayed under the assumption that CO and H2 are co-utilized.
c
A negative ATP yield is obtained for the WLP up to formation of acetyl-CoA from H2 /CO2 or a syngas mixture rich in hydrogen, this yield becomes positive when acetate
or acetate derived products (e.g. ethanol) are formed.
d
This reaction yields 8 electrons that need to be re-oxidized via other pathways (not included).
e
It is assumed that the reaction from pyruvate to acetyl-CoA is not catalysed by a PFOR enzyme and that the electrons derived from glycolysis and acetyl-CoA formation
are all in the form of NADH and are reinvested in the WLP.

Fig. 1. Product portfolio of native and engineered acetogens utilizing CO, CO2 and H2 (grey arrows). Multi-stage fermentation (orange arrows) and mixed co-cultures (green
arrows) broaden the range of potential products. TAGs: triacylglycerides. MEK: methyl ethyl ketone.

syngas conversion. An added value of mixed and co-cultures is the Several anaerobic, carboxydotrophic thermophiles have been
potential to utilize non-CO tolerant microbes, like methanogens, characterized to date (Henstra et al., 2007; Tiquia-Arashiro, 2014).
to grow on syngas by-products. Finally, mixotrophy is interesting The product spectrum of those thermophilic carboxydotrophs com-
for increasing the yield of already existing platforms or decreasing prises acetate, formate, H2 S, H2 , and CH4 (Henstra et al., 2007;
CO2 exhaust from these systems. However, suitable strains need to Tiquia-Arashiro, 2014). The first thermophilic acetogen to naturally
be selected or engineered to have multiple pathways active at the produce ethanol was the strain Moorella sp. HUC22-1, published in
same time. 2004 (Sakai et al., 2005, 2004). Efforts are being made to develop the
genetic toolbox of acetogenic thermophiles in order to broaden the
range of products. Although production of industrially relevant het-
4.5. Thermophilic gas fermenting microbes erologous products have not been reported so far, most progress is
made with M. thermoacetica (Iwasaki et al., 2013; Kita et al., 2013a,
Thermophilic production strains are known to have certain 2013b). Recently, we analyzed a hypothetical process in which ace-
advantages over mesophiles and have attracted significant atten- tone is produced in a bubble column fermentation at 60 ◦ C from
tion also for gas fermentation. The risk of contamination is reduced, corn stover-derived syngas with a M. thermoacetica strain with
since most contaminants are mesophiles (Bosma et al., 2013), and regard to thermodynamic and economic aspects (manuscript in
the oxygen solubility (although also the CO and H2 solubility) is preparation). We found that the process has the potential to be
reduced at elevated temperatures (Tiquia-Arashiro, 2014). Fur- economically interesting, if the high costs for biomass gasification
thermore, the use of a thermophilic production host offers the are reduced.
opportunity of low cost downstream processing via condensation Fig. 2 shows compounds with relatively low boiling point, which
and distillation, since the product enters the gas phase and can may be interesting product candidates for thermophilic production
be condensed from the gas stream leaving the reactor. Since the strains. In M. thermoacetica, the generation of 1 acetyl-CoA requires
product does not accumulate in the fermentation broth, it solves 0.5 ATP (Eq. (1)); net ATP is only generated in the subsequent step
problems with product toxicity and inhibition of production path- of acetate formation. Acetyl-CoA formation from CO leads to a net
ways.
28 S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30

Fig. 2. Compounds with relatively low boiling point are interesting products for gas fermentation processes using thermophiles. The temperature range and optimum of the
thermophilic acetogen M. thermoacetica (Drake and Daniel, 2004) is highlighted in orange.

Fig. 3. The ATP yield and the mol of CO2 consumed/produced based on the energy generation mechanism in M. thermoacetica proposed by Schuchmann and Müller, 2014.
MEK: methyl ethyl ketone.

ATP balance of +0.5 ATP (Eq. (2)). 5. Conclusion

Acetyl-CoA from H2 /CO2 : 4 H2 + 2CO2 − > 1acetyl-CoA


Vast amounts of industrial waste gases are currently emit-
+ 2 H2 O − 0.5ATP (1) ted to the atmosphere. Together with gasification of other waste
streams and underutilized fractions of biomass, these gases have
the potential to serve as substrates for gas fermentation processes
for the production of biochemicals. Such processes are attract-
Acetyl-CoA from CO : 4CO + 2 H2 O − > 1acetyl-CoA ing increasing attention since they may help to significantly curb
the emission of greenhouse gases, while at the same time having
+ 0.5ATP + 2CO2 (2)
the potential of decreasing the production cost of certain chem-
icals. A significant amount of research is currently focused on
For the production of 1 mol acetone, 2 acetyl-CoA are required. developing tools for metabolic engineering of novel acetogenic
Fig. 3 shows the ATP yield per 1 mol of product. The ATP yield calcu- production organisms, and novel products are constantly being
lations are based on the energy conservation mechanism proposed pursued. Heterologous production of higher value compounds has
by Schuchmann and Müller, 2014. For the production of ethanol, 1 been successful for some strains, but the reactions are often chal-
acetyl-CoA is reduced to acetaldehyde and further to ethanol, with lenging due to the redox balance, and mainly ethanol and BDO
NADH as electron donor in each step. The 2 NAD+ can generate an production is reaching commercial levels. Currently, there are only
additional 0.5 ATP. Acetone can be further reduced with NADH to few examples of biomass-based gasification plants operating at a
isopropanol, which generates an additional 0.25 ATP. The produc- commercial scale, and none of these are reported to utilize syn-
tion of isoprene via the mevalonate pathway requires 3 acetyl-CoA, gas for gas fermentation processes. To investigate the economic
2 NADH, and 3 ATP. The energy requirement for isoprene produc- aspect of these processes, it is therefore important to also conduct
tion is relatively high, and 4 ATP are required per mol product when techno-economic studies. Alternative fermentation processes have
H2 /CO2 serves as substrate. For the production of methyl ethyl the potential to address many of the challenges associated with pro-
ketone (MEK), 2 acetyl-CoA are converted via pyruvate into ace- duction of biochemicals through gas fermentation. In this review
tolactate, which is decarboxylated into acetoin. Acetoin is further we have focused on the recent development of processes utiliz-
reduced with NADH into (R,S)-2,3-BDO, which undergoes conver- ing for example multi-stage fermentation, mixed communities,
sion to MEK (Mueller et al., 2013). In comparison, the production defined co-cultures, mixotrophic processes, as well as thermophilic
of the native product of M. thermoacetica yields 1.5 ATP for the production organisms. Such production scenarios have the poten-
growth on CO, and 0.5 ATP for the growth on H2 /CO2 . Assuming that tial for significantly broadening the product portfolio, increasing
0.5 ATP is the lower limit to maintain and gain the biomass, only production titers and yields, exploiting different and also mixed
the production of the more reduced compounds acetone, ethanol, feedstock streams, removing inhibitors and decreasing the cost of
isopropanol, and MEK from CO generates enough energy. For the downstream processing. The future will undoubtedly show further
other production scenarios, by-product formation would be neces- development of industrial processes from this rapidly emerging
sary. field of research.
S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30 29

Acknowledgements Dürre, P., Eikmanns, B.J., 2015. C1-carbon sources for chemical and fuel production
by microbial gas fermentation. Curr. Opin. Biotechnol. 35, 63–72.
Dürre, P., 2016. Butanol formation from gaseous substrates. FEMS Microbiol. Lett.
This work was supported by The Novo Nordisk Foundation and 363 (6), 1–7.
a Ph.D. grant from the People Programme (Marie Curie Actions) Daniell, J., Nagaraju, S., Burton, F., Köpke, M., Simpson, S.D., 2016. Low-carbon fuel
of the European Union Seventh Framework Programme FP7- and chemical production by anaerobic gas fermentation. Adv. Biochem. Eng.
Biotechnol., 1–29.
People-2012-ITN, under grant agreement No. 317058, “BACTORY”. Daniels, L., Fuchs, G., Thauer, R.K., Zeikus, J.G., 1977. Carbon monoxide oxidation by
Research of M. Diender and D.Z. Sousa is supported by a ERC grant methanogenic bacteria. J. Bacteriol. 132 (1), 118–126.
(project 323009) of the European Union Seventh Framework Pro- Das, D., Veziroǧlu, T.N., 2001. Hydrogen production by biological processes: a
survey of literature. Int. J. Hydrogen Energy 26 (1), 13–28.
gramme FP7 and a Gravitation grant (project 024.002.002) of the
Datta, R., Reeves, A., 2013. Syntrophic co-culture of anaerobic microorganism for
Netherlands Ministry of Education, Culture and Science and the production of n-butanol from syngas. US20140206066 A1.
Netherlands Science Foundation(NWO). Diender, M., Stams, A.J.M., Sousa, D.Z., 2015. Pathways and bioenergetics of
anaerobic carbon monoxide fermentation. Front. Microbiol. 6, 1725.
Diender, M., Stams, A.J.M., Sousa, D.Z., 2016. Production of medium-chain fatty
acids and higher alcohols by a synthetic co-culture grown on carbon monoxide
References or syngas. Biotechnol. Biofuels 9 (1), 1.
Drake, H.L., Daniel, S.L., 2004. Physiology of the thermophilic acetogen Moorella
Abubackar, H.N., Veiga, M.C., Kennes, C., 2011. Biological conversion of carbon thermoacetica. Res. Microbiol. 155 (10), 869–883.
monoxide: rich syngas or waste gases to bioethanol. Biofuels Bioprod. Biorefin. Drzyzga, O., Revelles, O., Durante-Rodríguez, G., Díaz, E., García, J.L., Prieto, A.,
5 (1), 93–114. 2015. New challenges for syngas fermentation: towards production of
Abubackar, H.N., Veiga, M.C., Kennes, C., 2012. Biological conversion of carbon biopolymers. J. Chem. Technol. Biotechnol. 90 (10), 1735–1751.
monoxide to ethanol: effect of pH, gas pressure, reducing agent and yeast Edenhofer, O., Pichs-Madruga, R., Sokona, Y., Farahani, E., Kadner, S., Seyboth, K.,
extract. Bioresour. Technol. 114, 518–522. Adler, A., Baum, I., Brunner, S., Eickemeier, P., 2014. Climate change 2014:
Alauddin, Z.A.B.Z., Lahijani, P., Mohammadi, M., Mohamed, A.R., 2010. Gasification Mitigation of climate change. Working group III contribution to the fifth
of lignocellulosic biomass in fluidized beds for renewable energy assessment report of the Intergovernmental Panel on Climate Change. UK and
development: a review. Renew. Sustain. Energy Rev. 14 (9), 2852–2862. New York.
Alves, J.I., van Gelder, A.H., Alves, M.M., Sousa, D.Z., Plugge, C.M., 2013. Moorella U. S. Environmental Protection Agency, 2014. Municipal Solid Waste Generation,
stamsii sp. nov., a new anaerobic thermophilic hydrogenogenic carboxydotroph Recycling, and Disposal in the United States: Facts and Figures for 2012. https://
isolated from digester sludge. Int. J. Syst. Evol. Microbiol. 63 (11), 4072–4076. www.epa.gov/sites/production/files/2015-09/documents/2012 msw fs.pdf.
Angenent, L.T., Richter, H., Buckel, W., Spirito, C.M., Steinbusch, K.J.J., Plugge, C.M., Fast, A.G., Schmidt, E.D., Jones, S.W., Tracy, B.P., 2015. Acetogenic mixotrophy:
Strik, D.P., Im Grootscholten, T., Buisman, C.J.N., Hamelers, H.V.M., 2016. Chain novel options for yield improvement in biofuels and biochemicals production.
elongation with reactor microbiomes: open-culture biotechnology to produce Curr. Opin. Biotechnol. 33, 60–72.
biochemicals. Environ. Sci. Technol. 50 (6), 2796–2810. Görke, B., Stülke, J., 2008. Carbon catabolite repression in bacteria: many ways to
Antonovsky, N., Gleizer, S., Noor, E., Zohar, Y., Herz, E., Barenholz, U., Zelcbuch, L., make the most out of nutrients. Nat. Rev. Microbiol. 6 (8), 613–624.
Amram, S., Wides, A., Tepper, N., 2016. Sugar synthesis from CO2 in Gandía, L.M., Arzamendi, G., Diéguez, P.M., 2013. Renewable hydrogen energy: an
Escherichia coli. Cell 166 (1), 115–125. overview. Renew. Hydrogen Technol., 1–17.
Azadi, P., Brownbridge, G., Mosbach, S., Inderwildi, O., Kraft, M., 2015. Simulation Ganigué, R., Sánchez-Paredes, P., Bañeras, L., Colprim, J., 2016. Low fermentation
and life cycle assessment of algae gasification process in dual fluidized bed pH is a trigger to alcohol production, but a killer to chain elongation. Front.
gasifiers. Green Chem. 17 (3), 1793–1801. Microbiol. 7 (702), 1–11.
Bader, J., Mast-Gerlach, E., Popović, M.K., Bajpai, R., Stahl, U., 2010. Relevance of Green Chemicals Blog, 2014. INVISTA and LanzaTech Expand Collaboration
microbial coculture fermentations in biotechnology. J. Appl. Microbiol. 109 (2), (accessed June 2016) http://greenchemicalsblog.com/2014/06/05/invista-and-
371–387. lanzatech-expand-collaboration/.
Beck, Z.Q., Cervin, M.A., Chotani, G.K., Diner, B.A., Fan, J., Peres, C.M., Sanford, K.J., Grethlein, A.J., Jain, M.K., 1992. Bioprocessing of coal-derived synthesis gases by
Scotcher, M.C., Wells, D.H., Whited, G.M., 2013. Recombinant anaerobic anaerobic bacteria. Trends Biotechnol. 10, 418–423.
acetogenic bacteria for production of isoprene and/or industrial bio-products Griffin, D.W., Schultz, M.A., 2012. Fuel and chemical products from biomass
using synthesis gas. US20140234926 A1. syngas: a comparison of gas fermentation to thermochemical conversion
Becker, U., Grund, G., Orschel, M., Doderer, K., Löhden, G., Brand, G., Dürre, P., routes. Environ. Prog. Sustain. Energy 31 (2), 219–224.
Lederle, S., Bahl, H.J., Fischer, R.J., 2009. Zellen und Verfahren zur Herstellung Guiot, S.R., Cimpoia, R., Carayon, G., 2011. Potential of wastewater-treating
von Aceton 10. EP2421960 A1. anaerobic granules for biomethanation of synthesis gas. Environ. Sci. Technol.
Bengelsdorf, F.R., Straub, M., Dürre, P., 2013. Bacterial synthesis gas (syngas) 45 (5), 2006–2012.
fermentation. Environ. Technol. 34 (13–14), 1639–1651. Gutmann, K., 2014. Europe’s Dirty 30: How the EU’s Coal-fired Power Plants Are
Bertsch, J., Müller, V., 2015. Bioenergetic constraints for conversion of syngas to Undermining Its Climate Efforts. CAN Europe, WWF European Policy Office,
biofuels in acetogenic bacteria. Biotechnol. Biofuels 8 (1), 1. HEAL, the EEB and Climate. https://europeanclimate.org/wp-content/uploads/
Bertuccioli, L., Chan, A., Hart, D., Lehner, F., Madden, B., Standen, E., 2014. 2014/07/Dirty-30-report-finale2.pdf.
Development of Water Electrolysis in the European Union: Report for Fuel Handler, R.M., Shonnard, D.R., Griffing, E.M., Lai, A., Palou-Rivera, I., 2015. Life Cycle
Cells and Hydrogen Joint Undertaking. http://www.fch.europa.eu/sites/ Assessments of ethanol production via gas fermentation: anticipated
default/files/study%20electrolyser 0-Logos 0.pdf. greenhouse gas emissions for cellulosic and waste gas feedstocks. Ind. Eng.
Bosma, F., van der Oost, E., de Vos, John M., Willem van Kranenburg, R., 2013. Chem. Res. 55 (12), 3253–3261.
Sustainable production of bio-based chemicals by extremophiles. Curr. Hawkins, A.S., McTernan, P.M., Lian, H., Kelly, R.M., Adams, M.W.W., 2013.
Biotechnol. 2 (4), 360–379. Biological conversion of carbon dioxide and hydrogen into liquid fuels and
Brown, R.C., 2007. Hybrid thermochemical/biological processing. Appl. Biochem. industrial chemicals. Curr. Opin. Biotechnol. 24 (3), 376–384.
Biotechnol. 137 (1–12), 947–956. Heidenreich, S., Foscolo, P.U., 2015. New concepts in biomass gasification. Prog.
Schilling, Christophe, H., Burgard, A., Trawick, John, D., 2009. Methods and Energy Combust. Sci. 46, 72–95.
organisms for utilizing synthesis gas or other gaseous carbon sources and Henstra, A.M., Sipma, J., Rinzema, A., Stams, A.J.M., 2007. Microbiology of synthesis
methanol. WO/2009/094485. gas fermentation for biofuel production. Curr. Opin. Biotechnol. 18 (3),
Cement Sustainability Initiative, 2014. Cement Sector View on Industrial 200–206.
Innovation (accessed June 2016) https://www.iea.org/media/workshops/2014/ Hoffmeister, S., Gerdom, M., Bengelsdorf, F.R., Linder, S., Flüchter, S., Öztürk, H.,
industryreviewworkshopoct/2 Session1 A CSICement 231014.pdf. Blümke, W., May, A., Fischer, R.-J., Bahl, H., 2016. Acetone production with
Chen, W.Y., Liew, F., Koepke, M., 2013. Recombinant microorganisms and uses metabolically engineered strains of Acetobacterium woodii. Metab. Eng. 36,
therefor. US 20130323820 A1. 37–47.
Cho, C., Jang, Y., Moon, H.G., Lee, J., Lee, S.Y., 2014. Metabolic engineering of Hu, P., Chakraborty, S., Kumar, A., Woolston, B., Liu, H., Emerson, D.,
clostridia for the production of chemicals. Biofuels Bioprod. Biorefin. 9 (2), Stephanopoulos, G., 2016. Integrated bioprocess for conversion of gaseous
211–225. substrates to liquids. P. Natl. Acad. Sci. U. S. A. 113 (14), 3773–3778.
Choi, D., Chipman, D.C., Bents, S.C., Brown, R.C., 2010. A techno-economic analysis International Energy Agency, 2007. Tracking Industrial Energy Efficiency and CO2
of polyhydroxyalkanoate and hydrogen production from syngas fermentation Emissions, https://www.iea.org/publications/freepublications/publication/
of gasified biomass. Appl. Biochem. Biotechnol. 160 (4), 1032–1046. tracking emissions.pdf.
Chotani, G.K., Nielsen, A.T., Vaviline, D.V., 2012. Methods for increasing microbial International Energy Agency, 2014. IEA Energy Training Week 2014: Carbon
production of isoprene, isoprenoids, and isoprenoid precursor molecules using Capture and Storage. https://www.iea.org/media/training/presentations/
glucose and acetate co-metabolism. WO2013052914 A3. etw2014/course4and5/Day 2 Session 2 CCS.pdf.
Contag, P.R., 2015. Organism co-culture in the production of biofuels. US8986962 Invista, 2015. INVISTA and LanzaTech Make Breakthrough for Bio-derived
B2. Butadiene Production (accessed June 2016) http://www.invista.com/en/news/
Cueto-Rojas, H.F., van Maris, A.J., Wahl, S.A., Heijnen, J.J., 2015. pr-invista-and-lanzatech-make-breakthrough-for-bio-derived-butadiene-
Thermodynamics-based design of microbial cell factories for anaerobic production.html.
product formation. Trends Biotechnol. 33 (9), 534–546.
30 S. Redl et al. / Industrial Crops and Products 106 (2017) 21–30

Iwasaki, Y., Kita, A., Sakai, S., Takaoka, K., Yano, S., Tajima, T., Kato, J., Nishio, N., Piccolo, C., Bezzo, F., 2009. A techno-economic comparison between two
Murakami, K., Nakashimada, Y., 2013. Engineering of a functional thermostable technologies for bioethanol production from lignocellulose. Biomass Bioenergy
kanamycin resistance marker for use in Moorella thermoacetica ATCC 39073. 33 (3), 478–491.
FEMS Microbiol. Lett., 8–12. Ptasinski, K.J., 2008. Thermodynamic efficiency of biomass gasification and
Köpke, M., Liew, F., 2012. Production of butanol from carbon monoxide by a biofuels conversion. Biofuels Bioprod. Biorefin. 2 (3), 239–253.
recombinant microorganism. Patent WO2012/053905. Ramió-Pujol, S., Ganigué, R., Bañeras, L., Colprim, J., 2015. Incubation at 25C
Köpke, M., Held, C., Hujer, S., Liesegang, H., Wiezer, A., Wollherr, A., Ehrenreich, A., prevents acid crash and enhances alcohol production in Clostridium
Liebl, W., Gottschalk, G., Dürre, P., 2010. Clostridium ljungdahlii represents a carboxidivorans P7. Bioresour. Technol. 192, 296–303.
microbial production platform based on syngas. P. Natl. Acad. Sci. U. S. A. 107 Richter, H., Martin, M.E., Angenent, L.T., 2013. A two-stage continuous fermentation
(29), 13087–13092. system for conversion of syngas into ethanol. Energies 6 (8), 3987–4000.
Köpke, M., Mihalcea, C., Bromley, J.C., Simpson, S.D., 2011. Fermentative production Roddy, D.J., 2013. A syngas network for reducing industrial carbon footprint and
of ethanol from carbon monoxide. Curr. Opin. Biotechnol. 22 (3), 320–325. energy use. Appl. Therm. Eng. 53 (2), 299–304.
Köpke, M., Simpson, S., Liew, F., Chen, W., 2012. Fermentation process for Sakai, S., Nakashimada, Y., Yoshimoto, H., Watanabe, S., Okada, H., Nishio, N., 2004.
producing isopropanol using a recombinant microorganism. US20120252083 Ethanol production from H2 and CO2 by a newly isolated thermophilic
A1. bacterium, Moorella sp. HUC22-1. Biotechnol. Lett. 26 (20), 1607–1612.
Keller, M.W., Schut, G.J., Lipscomb, G.L., Menon, A.L., Iwuchukwu, I.J., Leuko, T.T., Sakai, S., Nakashimada, Y., Inokuma, K., Kita, M., Okada, H., Nishio, N., 2005. Acetate
Thorgersen, M.P., Nixon, W.J., Hawkins, A.S., Kelly, R.M., 2013. Exploiting and ethanol production from H2 and CO2 by Moorella sp. using a repeated
microbial hyperthermophilicity to produce an industrial chemical, using batch culture. J. Biosci. Bioeng. 99 (3), 252–258.
hydrogen and carbon dioxide. P. Natl. Acad. Sci. U. S. A. 110 (15), 5840–5845. Sakimoto, K.K., Wong, A.B., Yang, P., 2016. Self-photosensitization of
Kim, S., Dale, B.E., 2004. Global potential bioethanol production from wasted crops nonphotosynthetic bacteria for solar-to-chemical production. Science 351
and crop residues. Biomass Bioenergy 26 (4), 361–375. (6268), 74–77.
Kim, T.W., Bae, S.S., Lee, J.W., Lee, S.-M., Lee, J.-H., Lee, H.S., Kang, S.G., 2016. A Schuchmann, K., Müller, V., 2014. Autotrophy at the thermodynamic limit of life: a
biological process effective for the conversion of CO-containing industrial model for energy conservation in acetogenic bacteria. Nat. Rev. Microbiol.,
waste gas to acetate. Bioresour. Technol. 211, 792–796. 809–821.
Kita, A., Iwasaki, Y., Sakai, S., Okuto, S., Takaoka, K., Suzuki, T., Yano, S., Sawayama, Service, R.F., 2009. Solar fuels. Sunlight in your tank. Science (New York, NY) 326
S., Tajima, T., Kato, J., 2013a. Development of genetic transformation and (5959), 1472.
heterologous expression system in carboxydotrophic thermophilic acetogen Sheldon, R.A., 2014. Green and sustainable manufacture of chemicals from
Moorella thermoacetica. J. Biosci. Bioeng. 115 (4), 347–352. biomass: state of the art. Green Chem. 16 (3), 950–963.
Kita, A., Iwasaki, Y., Yano, S., Nakashimada, Y., Hoshino, T., Murakami, K., 2013b. Sipma, J., Lens, P.N.L., Stams, A.J.M., Lettinga, G., 2003. Carbon monoxide conversion
Isolation of thermophilic acetogens and transformation of them with the pyrF by anaerobic bioreactor sludges. FEMS Microbiol. Ecol. 44 (2), 271–277.
and kanr genes. Biosci. Biotechnol. Biochem. Sipma, J., Meulepas, R.J., Parshina, S.N., Stams, A.J., Lettinga, G., Lens, P.N., 2004.
LanzaTech, 2016. ArcelorMittal, LanzaTech and Primetals Technologies Announce Effect of carbon monoxide, hydrogen and sulfate on thermophilic (55 ◦ C)
Partnership to Construct Breakthrough D 87 m Biofuel Production Facility. hydrogenogenic carbon monoxide conversion in two anaerobic bioreactor
http://www.lanzatech.com/arcelormittal-lanzatech-primetals-technologies- sludges. Appl. Microbiol. Biotechnol. 64 (3), 421–428.
announce-partnership-construct-breakthrough-e87m-biofuel-production- Sipma, J., Lettinga, G., Stams, A.J.M., Lens, P.N.L., 2006. Hydrogenogenic CO
facility/. Accessed June 2016. conversion in a moderately thermophilic (55 ◦ C) sulfate-fed gas lift reactor:
Latif, H., Zeidan, A.A., Nielsen, A.T., Zengler, K., 2014. Trash to treasure: production competition for CO-Derived H2 . Biotechnol. Progr. 22 (5), 1327–1334.
of biofuels and commodity chemicals via syngas fermenting microorganisms. Spath, P.L., Dayton, D.C., 2003. Preliminary screening-technical and economic
Curr. Opin. Biotechnol. 27, 79–87. assessment of synthesis gas to fuels and chemicals with emphasis on the
Liew, F., Martin, M.E., Tappel, R.C., Heijstra, B.D., Mihalcea, C., Köpke, M., 2016. Gas potential for biomass-derived syngas: (NREL/TP-510-34929).
fermentation—a flexible platform for commercial scale production of Staley, B.F., Barlaz, M.A., 2009. Composition of municipal solid waste in the United
low-carbon-fuels and chemicals from waste and renewable feedstocks. Front. States and implications for carbon sequestration and methane yield. J. Environ.
Microbiol. 7, 694. Eng. 135 (10), 901–909.
Liu, K., Atiyeh, H.K., Stevenson, B.S., Tanner, R.S., Wilkins, M.R., Huhnke, R.L., 2014. Straub, M., Demler, M., Weuster-Botz, D., Dürre, P., 2014. Selective enhancement of
Mixed culture syngas fermentation and conversion of carboxylic acids into autotrophic acetate production with genetically modified Acetobacterium
alcohols. Bioresour. Technol. 152, 337–346. woodii. J. Biotechnol. 178, 67–72.
Liu, Y., Wan, J., Han, S., Zhang, S., Luo, G., 2016. Selective conversion of carbon Tiquia-Arashiro, S.M., 2014. Thermophilic Carboxydotrophs and their Applications
monoxide to hydrogen by anaerobic mixed culture. Bioresour. Technol. 202, in Biotechnology. Springer, New York City, NY, pp. 3–978 (doi 10).
1–7. Ueki, T., Nevin, K.P., Woodard, T.L., Lovley, D.R., 2014. Converting carbon dioxide to
Lovley, D.R., Nevin, K.P., 2013. Electrobiocommodities: powering microbial butyrate with an engineered strain of Clostridium ljungdahlii. mBio 5 (5),
production of fuels and commodity chemicals from carbon dioxide with e01636–14.
electricity. Curr. Opin. Biotechnol. 24 (3), 385–390. Vassilev, S.V., Baxter, D., Andersen, L.K., Vassileva, C.G., Morgan, T.J., 2012. An
Martin, M.E., Richter, H., Saha, S., Angenent, L.T., 2016. Traits of selected Clostridium overview of the organic and inorganic phase composition of biomass. Fuel 94,
strains for syngas fermentation to ethanol. Biotechnol. Bioeng. 113 (3), 1–33.
531–539. Vasudevan, D., Richter, H., Angenent, L.T., 2014. Upgrading dilute ethanol from
McAloon, A., Taylor, F., Yee, W., Ibsen, K., Wooley, R., 2000. Determining the cost of syngas fermentation to n-caproate with reactor microbiomes. Bioresour.
producing ethanol from corn starch and lignocellulosic feedstocks. National Technol. 151, 378–382.
Renewable Energy Laboratory Report. Weghoff, M.C., Müller, V., 2016. CO metabolism in the thermophilic acetogen
McKendry, P., 2002. Energy production from biomass (part 3): gasification Thermoanaerobacter kivui. Appl. Environ. Microbiol. 82 (8), 2312–2319.
technologies. Bioresour. Technol. 83 (1), 55–63. Wei, L., Pordesimo, L.O., Igathinathane, C., Batchelor, W.D., 2009. Process
Mohammadi, M., Najafpour, G.D., Younesi, H., Lahijani, P., Uzir, M.H., Mohamed, engineering evaluation of ethanol production from wood through
A.R., 2011. Bioconversion of synthesis gas to second generation biofuels: a bioprocessing and chemical catalysis. Biomass Bioenergy 33 (2), 255–266.
review. Renew. Sustain. Energy Rev. 15 (9), 4255–4273. Wiegel, J., 1994. Acetate and the potential of homoacetogenic bacteria for
Mueller, A.P., Köpke, M., Nagaraju, S., 2013. Recombinant microorganisms and uses industrial alications. In: Acetogenesis. Springer, pp. 484–504.
therefor. US 2013/0330809 A1. Woolerton, T.W., Sheard, S., Reisner, E., Pierce, E., Ragsdale, S.W., Armstrong, F.A.,
Nam, C.W., Jung, K.A., Park, J.M., 2016. Biological carbon monoxide conversion to 2010. Efficient and clean photoreduction of CO2 to CO by enzyme-modified
acetate production by mixed culture. Bioresour. Technol. 211, 478–485. TiO2 nanoparticles using visible light. J. Am. Chem. Soc. 132 (7), 2132–2133.
National Energy Technology Laboratory, 2016. Gasification Plant Databases World Steel Association, 2015a. Steel Statistical Yearbook 2015. https://www.
(accessed June 2016) http://www.netl.doe.gov/research/coal/energy-systems/ worldsteel.org/dms/internetDocumentList/bookshop/2015/Steel-Statistical-
gasification/gasification-plant-databases. Yearbook-2015/document/Steel%20Statistical%20Yearbook%202015.pdf.
Ou, X., Zhang, X., Zhang, Q., Zhang, X., 2013. Life-cycle analysis of energy use and World Steel Association, 2015b. Sustainability Indicators (accessed June 2016)
greenhouse gas emissions of gas-to-liquid fuel pathway from steel mill off-gas https://www.worldsteel.org/statistics/Sustainability-indicators.html.
in China by the LanzaTech process. Front Energy 7 (3), 263–270. Xu, S., Fu, B., Zhang, L., Liu, H., 2015. Bioconversion of H2 /CO2 by acetogen enriched
Pachauri, R.K., Allen, M.R., Barros, V.R., Broome, J., Cramer, W., Christ, R., Church, cultures for acetate and ethanol production: the impact of pH. World J.
J.A., Clarke, L., Dahe, Q., Dasgupta, P., 2014. Climate Change 2014: Synthesis Microbiol. Biotechnol. 31 (6), 941–950.
Report. Contribution of Working Groups I, II and III to the Fifth Assessment van der Hoeven, M., 2013. CO2 Emissions From Fuel Combustion: IEA Statistics.
Report of the Intergovernmental Panel on Climate Change. IPCC. https://www.iea.org/publications/freepublications/publication/
Perez, J.M., Richter, H., Loftus, S.E., Angenent, L.T., 2013. Biocatalytic reduction of CO2EmissionsFromFuelCombustionHighlights2015.pdf.
short-chain carboxylic acids into their corresponding alcohols with syngas
fermentation. Biotechnol. Bioeng. 110 (4), 1066–1077.

You might also like