Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Biofuels

ISSN: 1759-7269 (Print) 1759-7277 (Online) Journal homepage: http://www.tandfonline.com/loi/tbfu20

Ethanol production by syngas fermentation


in a continuous stirred tank bioreactor using
Clostridium ljungdahlii

Bimal Acharya, Animesh Dutta & Prabir Basu

To cite this article: Bimal Acharya, Animesh Dutta & Prabir Basu (2017): Ethanol production by
syngas fermentation in a continuous stirred tank bioreactor using Clostridium ljungdahlii, Biofuels,
DOI: 10.1080/17597269.2017.1316143

To link to this article: http://dx.doi.org/10.1080/17597269.2017.1316143

Published online: 16 May 2017.

Submit your article to this journal

Article views: 8

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tbfu20

Download by: [The UC San Diego Library] Date: 25 May 2017, At: 06:33
BIOFUELS, 2017
https://doi.org/10.1080/17597269.2017.1316143

Ethanol production by syngas fermentation in a continuous stirred tank


bioreactor using Clostridium ljungdahlii
Bimal Acharyaa, Animesh Dutta a
and Prabir Basub
a
School of Engineering, University of Guelph, Guelph, ON, Canada; bFaculty of Engineering, Dalhousie University, Halifax, NS, Canada

ABSTRACT ARTICLE HISTORY


Ontario biomass could be thermochemically processed by dry and wet torrefaction to produce Received 21 November 2016
high quality solid biofuel. These solid fuels or raw biomass could also be gasified to produce Accepted 17 February 2017
syngas. This study analyzes and demonstrates a successful and efficient way of producing KEYWORDS
bioethanol from syngas fermentation using Clostridium ljungdahlii in a laboratory scale Syngas fermentation;
continuous stirred tank bioreactor having an innovative gas supply and effluent extraction bioethanol; mass transfer;
structures. At the beginning of the experiment, a batch process was conducted to grow Clostridium ljungdahlii
microorganisms and allow the growth of the microorganisms to reach to maximum cell density
in a reactor without supplying a gas. Ethanol production was observed by supplying two
different gas compositions which included 100% CO and simulated syngas, mimicking the
composition of syngas extracted from lignocellulosic biomass having 60% CO, 35% H2, and 5%
CO2. CO and syngas were fermented with different gas flow (5–15 mL/min), effluent flow (0.25–
0.75 mL/min), and media flow rates and stirrer speed (300–500 rpm) at atmospheric pressure
and 378C. The gas flow rate, media and effluent flow rate, pH level, and stirrer speed were
controlled during the fermentation process. The exhaust gas was reused for the improvement
of residence time using a loop-back system for improving the gas–liquid mass transfer.
Excessive foam was observed during the fermentation process which was controlled using
diluted antifoam-204. Maximum cell concentration reached 2.4 g/L. The mass transfer
coefficient showed better performance during syngas fermentation than CO fermentation.
More bioethanol production was observed by syngas fermentation than CO fermentation. CO
fermentation produced 0.17–1.33 g/L-effluent ethanol and 8.92–23.67 g/L-effluent acetic acid
whereas syngas fermentation produced 0.85–3.75 g/L-effluent ethanol and 8.89–14.97 g/L-
effluent acetic acid.

Introduction thermochemical conversion process has many advan-


tages over the biochemical conversion process: the
Biofuels produced from lignocellulosic biomass are
possibility of using 100% lignocellulosic biomass
good options for the future replacement of fossil fuels.
including lignin; independence of feedstock composi-
Bioethanol and biodiesel are produced from lignocellu-
tions; elimination of complex pretreatment and high
losic biomass through biochemical and thermochemi-
enzyme costs; and independence of H2:CO:CO2 ratio in
cal transformation processes. Both processes have
syngas fermentation [5]. However, the main constraints
merits and demerits but the biochemical process is a
of syngas fermentation using biocatalysts are low gas–
more matured and dominating process than the ther-
liquid mass transfer, production of inhibitory products
mochemical process in bioethanol production [1]. Eth-
during fermentation, and low productivity associated
anol produced from syngas fermentation using
with cell density [6]. Another likely limiting factor for
industrial off-gas or natural gas or syngas extracted
ethanol production is its high concentration of undis-
from biomass could be a cheaper feedstock solution
sociated acetic acid in the bioreactor [7].
[2]. There are two paths to produce bioethanol from
Syngas, a mixture of carbon monoxide, hydrogen
syngas, using either a metallic catalyst or a biological
and carbon dioxide, is used in the biochemical conver-
catalyst. Metallic catalysts, either metal based or modi-
sion process in the presence of a biocatalyst or microor-
fied methanol catalysts, are expensive and require
ganism to produce biofuel. Syngas fermentation is
higher operating temperatures and pressures [1,3].
defined as the process to convert syngas into various
Bioethanol produced from syngas using a biological
chemical products like ethanol, alcohol, and organic
catalyst is economical and operable at ambient tem-
acid. Commonly used biological catalysts or microorgan-
perature and atmospheric pressure using low grade
isms in syngas fermentation are Clostridium ljungdahlii,
feedstocks including lignocellulosic biomass and
Clostridium ragsdalei, Clostridium carboxidivorans,
municipal wastes [4,5]. Syngas fermentation using the

CONTACT Bimal Acharya bacharya@uoguelph.ca


© 2017 Informa UK Limited, trading as Taylor & Francis Group
2 B. ACHARYA ET AL.

Clostridium autoethanogenum, and Alkalibaculum bacchi [18,22]. Shen et al. and Richter et al. used a hollow fiber
[7–10]. These biocatalysts play an important role to membrane for the gas supply system [7,16]. Trickle
metabolize syngas into bioethanol and acetic acid bed reactors (TBR) are used in syngas fermentation to
through the wood–ljungdahl pathway [11]. There are produce hydrogen using photosynthetic bacterium
number of challenges to economically operate continu- e.g. Rhodospirillum rubrum and Rubrivivax gelatinosus;
ous fermentation by using biomass extracted syngas in CH4 using tri-cultures of R. rubrium; acetate using Pep-
a bioreactor because it needs complex distillation due tostreptococcus productus; and alcohol using Clostrid-
to low ethanol concentration, advanced cleaning of ium ragsdalei [8]. There are no accessible literatures on
toxic contaminants in biomass extracted syngas, and the study of fermentation using a continuous feed gas
cheaper growth medium or media ingredients [12–15]. loop-back supply system with continuous stirred tank
The continuous fermentation process is better than reactor (CSTR). Hence, this study tries to utilize a simpli-
the batch process because continuous flow of nutrition fied continuous stirred tank bioreactor employing a
never allows the decrease of nutrition levels in the novel gas flow with feedback gas supply system and
reactor for the microorganism. This continuous flow of effluent abstraction schemes to improve gas–liquid
nutrition process favors the continuous growth of the mass transfer, ease inhibitory stress, and increase bioe-
biocatalyst at the highest concentration level. How- thanol production and compare the productivity of
ever, a batch culture requires more time to re-grow the bioethanol by using carbon monoxide and syngas. Fur-
fermenting organism after each fermentation cycle. ther, the influence of gas and liquid flow rates on gas
For commercial ethanol production, batch culture may conversion efficiency, bioethanol yield, acetic acid
be unrealistic. Similarly, a two-stage continuous culture yield, pH level, and mass transfer are analyzed.
could be a better option than a single stage-continu-
ous fermentation process because separate treatments
can be made for acetogenesis and solventogenesis Mass transfer and product yield
stages like separate pH level, temperature, working vol- One of the major limitations of syngas fermentation
ume level, dilution rate, nutrition level, growth rate, cell technology is the low mass transfer between substrate
density, and reactor productivity [7]. and microbes. A number of studies were conducted to
The mass transfer in substrate and microbes enhance the mass transfer [5,7,8,11,16]. Researchers
depends on the amount of dissolution of syngas in the observed an improvement in mass transfer rates by
fermentation media. This process brings in more gas– improving the residence time of syngas with biocata-
liquid interfacial area and retention time in the fermen- lyst, gas supply system, gas flow rate, reactor design,
tation media. The mass transfer rates and microbe and gas supply system design. Increasing the agitation
accessibility to syngas can also be improved by creat- speed of the stirrer up to the optimum limit improves
ing finer bubbles or decreased rising velocity in the fer- the accessibility of the syngas to the biocatalyst by
mentation media with an increase in agitation speed forming finer bubbles and reducing the rising velocity
[5]. To improve the mass transfer of gasses into the liq- in the fermentation media. The mass transfer coeffi-
uid medium, different bioreactor designs such as bub- cients, k,CO a/VLand kL,H2 a/VL are the major parameters
ble columns, packed columns, trickle beds, hollow for determining the mass transfer rate which can be
fiber reactors, and continuously stirred tank reactors calculated by using following Equations (1) (general)
with micro-spargers are utilized [6–8,11,16–18]. Richter and (2) (boundary layer theory) [23]:
et al. concluded that continuous culture has more
advantages than batch culture in syngas fermentation h i
kL;i :a  V :
1 dni
dt
technology [7]. Even though there is improvement in ¼2
L
3 (1)
the mass transfer by increasing the agitation speed or VL
4Cout;i 5
 Cin;i
gas flow rate, the overall process may not be commer- Cout;i
ln Cin;i
cially viable due to high energy consumption or high
 sffiffiffiffiffi

power to volume ratio and stress to the biocatalysts VL kL;i :aVL Di
[19,20]. If the amount of cells in the bioreactor is not  i ¼ (2)
VL kL;i :aVL j Dj
sufficient to consume the gas provided, then the
increase in syngas flow rate means an eventual
Similarly, the overall mass transfer coefficient value
increase in the loss of syngas at the exhaust system,
can be determined for a hollow fiber membrane by
resulting in lower gas conversion efficiency. The pro-  
C
ductivity of a biocatalyst decreases with an increase in Equation (3) [24]. The slope of ln Cout;iout;i
Cin;i vs: time
stress either by mechanical means or an increase in can be used to determine the mass transfer coefficient:
cell density [18,21]. In orthodox syngas fermentation
   

technology, a single orifice or multiple orifices are used Cout; i Q L


ln ¼ exp kL;i a t (3)
to release the gas into the fermentation broth in both Cout; i  Cin; i VL VL
the batch and continuous fermentation process
BIOFUELS 3

where outlined by Liu et al. [10]:

dni total moles of ethanol produced


¼ instantaneous molar rate of gas componets i Ethanol Yield ¼ total moles of CO consumed
 100% (5)
dt transfered into the medium; mmol 1 mole of ethanol produced
h 6 moles of CO consumed
Maximum cell mass  Initial cell mass
VL ¼ dynamic liquid holdup volume in the Cell Mass Yield ¼
moles of CO consumed
bioreactor; L
(6)
kL; i orj :a VL ¼ volumetric mass transfer coefficient
Total moles of ½Gasi consumed
of supplied gas; h1 ½Gasi utilization % ¼  100%
Total moles of ½Gasi supplied

i ¼ components of composition of gas e:g: CO; H2 ; (7)


CO2 ; CH4
Cin; i ¼ liquid phase concentration of gas reactant i Methodology
in equilibrium by Henry’ s law with the gas at
the inlet of bioreactor, mmol CSTBR fabrication
h
Cout; i ¼ liquid phase concentration of gas reactant i The continuous stirred tank bioreactor (CSTBR) was
0
in equilibrium by Henry s law With the gas designed and fabricated at Bio-Renewable Innovative
at the outlet of bioreactor, mmol
h Lab (BRIL), University of Guelph, ON. The fabricated 3L
Di, Dj = diffusivities for i and j gases e.g. at 37  C, CO laboratory scale low cost reactor (made with transpar-
(3.26 £ 10¡9) and H2 (6.48 £ 10¡9) m2/s, ent PVC pipe type PVC-9002-86-20 for tank and Plexi-
Q = liquid recirculation rate glass sheet type VH-100 Acrylic resin for base and lid)
L = length of gas supply pipe line, m with working volume of 2L broth media is handy, sim-
t = sampling time, s ple, robust, and easy for maintenance, data reading,
a = specific surface area of gas supply system, m2. and cleaning. A low cost innovative gas diffuser was
positioned at 1 cm and on the fringe, and an in situ cell
Equation (1) indicates that an increase in the inlet
retention filter at 15 cm height of the reactor. The reac-
syngas increases the volumetric mass transfer coeffi-
tor was connected to a temperature and pH meter
cient. However, Devarapalli et al. [8] observed that kL,
(PHE-1411, Omega Environmental Inc., Laval, QC, Can-
CO a/VL increased by 24% when the syngas flow rate
ada) and digital pressure gauge (PHH-222, Omega
was doubled while kL,H2 a/VL decreased by 46% which
Environmental Inc., Laval, QC, Canada). A gas flow
may be due to limitations of concentration of hydro-
meter (PMR1-0106018, Cole Parmer Inc., USA) and a
gen by large concentration of liquid phase carbon
micro-pump (GE-F155001, Gilson Inc, USA) of the
monoxide. Orgill et al. concluded that a decrease in
desired rate by using F117938 tubes were connected
the mass transfer coefficient kL,i a/VL for both carbon
to the reactor to monitor the gas flow amount and
monoxide and hydrogen is due to high liquid holdup
insertion/extraction of the media/product. Membrane
volume, VL, a result of the increased liquid recirculation support was positioned at two-thirds level of the biore-
rate [17]. The hydrogenase activity of CO limits the mix-
actor to install the membrane on the higher level of
ing ability of hydrogen gas into the media [25]. Simi-
the fermentation broth. For a continuous supply of
larly, the mass transfer coefficient declines with the
media and extraction of effluent, modified 2L anaero-
rise in media flow rate in a pure CO environment [17].
bic glass jars were used.
Klasson et al. used Equation (4) to determine the
mass transfer coefficient (kL) for a slightly soluble gas-
eous substrate [26]: Microorganism and media preparation
An anaerobic bacterium biocatalyst, Clostridium ljung-
1 dNsG kL a G
¼ Ps  PsL (4) dahlii, an American Type Culture Collection
VL dt H
(ATCC#55380) was collected from Cedarlane, Burling-
where ton, Ontario. Cedarlane supplied the broth media
agents and preparation manual. About 1L of broth
NsG ¼ molar substrate transfered from gas phase media was prepared by adding 10 g of peptone, 10 g
ðmolÞ of beef extract, 3 g of yeast extract, 5 g of salt, 1 g of
VL ¼ volume of the bioreactor (L) soluble starch, 3 g of sodium acetate, 4 ml of 0.0025%
PsG and PsL = partial pressure of the gaseous substrate resazurin as an oxygen indicator, and 1000 ml of deion-
in gas and liquid phase (atm) ized water. After all ingredients dissolved, the medium
H ¼ Henry's laws constant (L.atm/mol). was boiled for 10 min to drive off oxygen. The medium
was cooled by bubbling with oxygen free gas, 0.5 g of
During CO and syngas fermentation, the product L-cysteine HCl as a reducing agent was added and the
yield can also be expressed by using Equations (5–7) as pH adjusted to 6.8. This was dispensed under an
4 B. ACHARYA ET AL.

oxygen free environment and autoclaved at 121 C. The experiment, an anaerobic environment was created at
freeze dried 0.5 ml of biocatalyst was transferred asep- atmospheric pressure. The pH level was set to 4.5 § 0.5
tically into the broth media in test tubes, cultured for by using base-NaOH or acid-HCl whenever necessary
72 h and disseminated anaerobically in the broth and the stirrer speed was adjusted from 300 to
media jar at 37 C. A Bunson burner was used to create 500 rpm [7,18]. The medium and effluent extraction
an aseptic environment, and an incubator was used to flow rates were set at 0.5 § 0.25 mL/min. Pressure was
maintain a constant temperature environment. Finally, controlled by using a back-pressure regulator in the
the cultured microorganism was transferred anaerobi- reactor. To capture the unwanted exhaust products, a
cally into the CSTBR for the syngas fermentation bubbler was used.
process [6]. The main aim of syngas fermentation here is to
extract bioethanol from the gasification of biomass
products. The composition of syngas released from the
Experimental setup and procedure
gasification of all types of biomass consists of CO, H2,
A schematic diagram of the experimental setup for CO2, CH4 and few other gases with tar and ashes [18].
syngas fermentation is presented in Figure 1. All appa- A similar composition of syngas from a gas supplier
ratus for the experiment including CSTBR and accesso- was used during the experiment. For the investigation
ries were sterilized in an autoclave at 121 C for 20 min. of products, gas and liquid samples were collected
Nitrogen gas was used inside the reactor to create an intermittently. The nutrient level and activities of the
anaerobic environment. Initially, the prepared and ster- microorganism was maintained by continuously sup-
ilized 1.9 L fresh broth media was transferred to the 3 L plying broth media using a micro-pump. The cell mass,
CSTBR in a nitrogen environment. An inoculum of excluding the suspended cells in the bioreactor, could
100 mL was added in the medium in the same environ- be captured in a droplet, form a biofilm or be sus-
ment. The medium plus inoculum was kept airtight. To pended in the medium. Gas flow rates of 5, 10, and 15
drive off the oxygen, nitrogen was continuously mL/min; media flow rates of 0.25, 0.5, and 0.75 mL/
pumped into the reactor with medium for 2 h to create min; and stirrer speeds of 300 and 500 rpm were
a completely anaerobic environment. The reactor was tested. Different batches of testing were conducted for
placed in a temperature controlled environment for 48 different combinations: Batch 1 and 2 (B11 and B21)
h to allow propagation of microorganisms in the biore- for 5 mL/min gas, 0.25 mL/min media and effluent flow
actor. The batch was then ready to run in the reactor rate, and 300 rpm of stirrer; Batch 3 and 4 (B12 and
with different fermenting gases, carbon monoxide, B22) for 10 mL/min gas, all other parameters
and syngas in a continuous fermentation process. Car- unchanged. Batch numbers are numbered as B1x for
bon monoxide was run in the first batch at the begin- the fermentation of carbon monoxide and B2x for the
ning then switched to syngas in the second batch fermentation of syngas (x = 1–18 signifying each batch
fermentation processes. After 48 h of each batch, the having fresh broth medium with varying gas flow rate,
reactor was operated continuously for 45 days. media/affluent flow rate, and stirrer speed). Cell density
Pure carbon monoxide was supplied to the reactor increased to a maximum level with the addition of
in the first cycle and then switched to syngas with 60% fresh medium in each batch. Foam formation was con-
CO, 35% H2, and 5% CO2 at rates of 5–15 mL/min, and trolled by using diluted antifoam-204 (1/100 mL dilu-
bioethanol production was recorded. Throughout the tion with deionized water).

Figure 1. CSTBR semi-continuous fermentation setup.


BIOFUELS 5

The pH level of the fresh medium was set to 6.8. sieve. There were two detectors: a flame ionization
After the addition of inoculum, the pH level decreased detector and a thermal conductivity detector. A stan-
rapidly and stabilized at 4.4 § 0.4. This may be due to dard gas mixture was utilized for the standardization
formation of number of acidic compounds like propa- of the analyzer readings. A Hamilton gas tight syringe
nol-C3H8O, propionic acids-C3H6O2, nonanoic acid- (Hamilton Co., Revo, NV, USA) was used to inject
C9H18O2, benzaldehyde including bioethanol-C2H6O, or 100 mL in a helium gas environment. The gas samples
acetic acid-CH3COOH [4,6]. During the culturing of the were collected from the gas inlet and outlet of the bio-
microorganism in the fresh medium in the first 48 h, reactor. Helium gas at 2 mL/min was supplied automat-
the pH level changed from 5.8 to 5.0. This may be due ically during the experiment from the gas cylinder. Gas
to rapid formation of acetic acid at the beginning. The analysis data were collected from the experiment.
pH level lowered from 5.0 to 4.4 during the continuous
process due to formation of bioethanol and acetic
acid. The addition of base-NaOH was made when the Results and discussion
pH level dropped below 4.0 as shown in Figure 1. The Operation of the bioreactor
variation of pH level could be due to formation of ace-
tic compounds, and flow of media, affluent, and fer- The bioreactor was operated continuously at 37 C after
menting gas. 48 h of bacterial growth. Fresh broth medium was
pumped at different liquid flow rates into the bioreac-
tor after achieving an appropriate cell population dur-
Analytical methods ing the first 48 h. The stirrer rate was set at 300 rpm for
Cell growth one set of continuous process and 500 rpm for another
Cell growth of Clostridium ljungdahlii was observed by set of continuous process. During the continuous pro-
quantifying the cell dry weight of it. Three samples cess, liquid flow rates were chosen from 0.25, 0.5, and
were taken from the bioreactor in each batch and the 0.75 mL/min; and gas flow rates were chosen from 5,
average was plotted with standard deviation. Optical 10, and 15 mL/min. The pH of the bioreactor was kept
cell density was analyzed using a spectrophotometer at 4.0–4.8. During the experiment, in the first week of
at 600 nm and the cell concentration or cell dry weight the first run, excessive foam was observed in the biore-
was determined using a standard calibration curve. actor. Diluted antifoam liquid was used to control the
foam, but all biocatalyst were dead after the addition
Solvent analysis of 10 drops of 1/100 diluted antifoam liquid. This may
The amounts of bioethanol and acetic acid were ana- be due to an excessive amount of antifoam liquid. The
lyzed using a GC-MS system, equipped with a Bruker antifoam liquid was 1/100 diluted with deionized water
BR-SWAX column (30 m x 0.25 mm internal diameter and added 1–2 drops at a time into the bioreactor
with a 0.25 mm phase thickness), gas chromatograph whenever excessive foam was observed. The experi-
(7890A) and mass spectrometer (59756MS) of Agilent ment was repeated without considering any of the pre-
Technologies, USA. Oven temperature was set at 44 C vious data. During the second continuous run, i.e. after
and held for 3.5 min. Intermittent temperature was set 45 days, the experiment had to be repeated due to
at 200 C with ramp of 5 C/min. Helium carrier gas was loss of microorganism caused by air leakage in the
supplied at a rate of 1 mL/min and the sample was experimental setup. With an increase in the liquid flow
inoculated in an un-fragmented mode at 280 C. The rate, the increased dilution reduced cell population
0.5 mL solvent samples were transferred to 15 mL vials and productivity. Out of the three liquid flow rates, the
and incubated at 75 C for 5 min which was then equili- optimum was 0.5–0.75 mL/min. The optimum condi-
brated with a 75 mm carboxen-polydimethyle-siloxane tion for bioethanol production was observed at a stirrer
(CAR-PDMS) fiber dipped in a headspace for 20 min. rate of 300 rpm, gas flow rate of 10–15 mL/min, and
The volatile compounds were thermally desorbed in liquid flow rate of 0.25–0.5 mL/min.
the injector port by manually injecting and exposing
the fiber for 8 min. The mass spectrometer scanned
Cell concentration and pH
10–150 m/z at an interval of 1 s. Ionization was formed
by an electronic impact at 200 C, while the transfer Measurements of cell optical density or cell dry weight
line was kept at 250 C. The +ve ion mode data were (cell concentration) were taken in each batch when the
collected and products were extracted physically by condition changed. Cell concentration in the first batch
using a SPME holder (Supelco, USA), a hotplate, and a (B11 and B21) of each run reached a maximum level.
metal support with clamps. After the first 48 h, the cell concentration never crossed
the cell concentration of first batch (2.4 g/L). This may
Gas analysis be due to the lowering of pH by the formation of acidic
Gas samples were analyzed using 8610C GC (SRI Instru- components, stirring, bubbling of gas, and depleting of
ments, Torrance, CA) having two columns molecular nutrition levels of the medium. Nutrition levels of the
6 B. ACHARYA ET AL.

medium were maintained in the reactor by continu- fermentation. A similar observation was found with cell
ously supplying fresh medium to the reactor at a speci- concentration. This may be due to higher formation of
fied rate for each cycle (or batch). Cell growth rate acetic acid during CO fermentation than syngas
decreased with a decrease in pH value and an increase fermentation.
in temperature, so continuous monitoring of the pH
level and temperature was carried out and the pH level
Product profiles
was maintained by addition of base. The pH of fermen-
tation is a significant parameter in controlling the sub- The acetyl-CoA biochemical or wood–ljungdahl path-
strate metabolism and changing the physiological way is followed by Clostridium ljungdahlii for cell
state. Cell growth, selectivity of product, and metabolic growth, acetic acid formation, and adenosine triphos-
by-products discharge are affected by pH level [18]. phate (ATP) during the growth stage of syngas fermen-
The pH level decreased during the first cycle of the fer- tation at pH 4–7 [27]. The conversion of acetyl-CoA to
mentation process due to formation of weak organic acetate needs two major steps, where the first step fol-
acids which may permeate through the cell wall, accu- lows conversion from acetyl-CoA to acetate then acet-
mulate inside the cell, and reduce the inside pH by H+ aldehyde with reduced ferredoxin and finally to
ions. This phenomenon helps to convert the acetogen- ethanol via alcohol dehydrogenase (ADH), and the sec-
esis phase to the solventogenesis phase. ond step follows direct conversion of acetyl-CoA to
During CO and syngas fermentation, pH decreased acetaldehyde then to ethanol by reducing acetalde-
rapidly during the first 48 h from 5.80 to 5.0 during the hyde [28]. Adjusting the pH level by adding base to the
batch process at a gas flow rate of 5 mL/min, media/ bioreactor causes ethanol production to be improved
affluent circulation rate of 0.25 mL/min, and stirrer but the growth of unwanted acetic acid was also
speed of 300 rpm. The pH level further decreased and observed. Acetic acid production during CO fermenta-
reached an almost stable condition at 4.4 § 0.4 for CO tion is more than acetic acid formation during syngas
and syngas fermentation during the continuous pro- fermentation. More acetic acid formation during CO
cess as shown in Figure 2. The rapid change in pH level fermentation may be due to the less conversion of ace-
at the beginning indicates that acid formation is high tic acid in to the ethanol caused by thin layer formation
which may be due to an increase in formation of acetic or by low reactivity with CO or high CO flow rate.
acid. A continuous variation in pH level was observed In each batch of continuous process, the products
throughout the experiment which may be due to a var- are bioethanol and acetic acid. Production of bioetha-
iation in acid formation during the fermentation pro- nol and acetic acid increased with cell concentration,
cess and disturbances caused to the biocatalyst by CO and syngas utilization in a linear pattern, which
variation of media and affluent in/out process, stirrer increases or decreases according to pH level, cell con-
speed, and gas flow rate. Lowering the pH (indicating centration level, gas flow, and liquid flow. This signifies
the formation of acetic acid) triggered a transformation that continuous production of bioethanol and acetic
in cell metabolism from acetogenesis to solventogene- acid occurs in each cycle. Each batch of cycle was
sis. Cell concentration and ethanol yield are directly refreshed after 120 h by introducing fresh broth media.
proportional to pH variation. The pH level is always at The production of bioethanol varied from 0.30 § 0.07
the higher level during syngas fermentation than CO to 1.3 § 0.05 g/L and acetic acid varied from 8.60 §

Figure 2. pH profiles and cell concentration during CO and syngas fermentation.


BIOFUELS 7

0.40 to 23.67 § 0.80 g/L during CO fermentation t = time duration of continuous fermentation in bio-
whereas production of bioethanol varied from 0.80 § reactor (48–1088 h)
0.02 to 3.75 § 0.08 g/L and acetic acid varied from 5.20 R = R-squared value or coefficient of determination
§ 0.80 to 14.20 § 0.50 g/L during syngas fermentation. Suffix = production from CO or syngas e.g. YCO
In the continuous extraction process, when the stirrer means bioethanol from carbon monoxide.
speed is set at 300 rpm (B11–B19 and B21–B29), the
production of ethanol and acetic acid batch was from R-squared value indicates that the proposed model
0.41 to 1.33 g/L in B11–B19 and from 0.92 to 3.75 g/L is valid or near to the experimental results. Figures 3a
in B21–B29. In each gas flow with the 300 rpm stirrer, and 3b show that bioethanol production increased
ethanol production increased with the increase in liq- with an increase in the gas flow rate up to an optimal
uid flow (media in and affluent out) rate. When the gas flow rate. The optimum gas flow and liquid flow/
flow doubled from 5 to 10 mL/min, ethanol production extraction rate for the production of bioethanol was
increased by 58% during CO fermentation, whereas found to be 10–15 mL/min and 0.25–0.5 mL/min. How-
49% during syngas fermentation. ever, liquid flow rate and stirrer speed has less impact
The acetic acid formation is more at the beginning on bioethanol production than gas flow rate. Acetic
of the experiment. Such acetic acid formation contin- acid formation was almost proportional to bioethanol
ued during the continuous processes. This may be due formation. However, the production of acetic acid was
to biofilm formation and acetogenesis process. Liquid higher with higher gas flow rate during CO fermenta-
flow rate varied from 0.25 to 0.75 mL/min; there was tion and lower during syngas fermentation. This may
11–29% variation in bioethanol and acetic acid forma- be due to more effective solventogenesis during syn-
tion with variation of the liquid flow. This may be due gas fermentation than CO fermentation. The maximum
to an increase in nutrition to the cells and an increase bioethanol during CO fermentation was 1.33 g/L at 10
in gas–liquid interaction. Similarly, an increase in the mL/min gas flow rate, 0.25 mL/min liquid flow rate,
gas flow rate increased production of bioethanol by and 300 rpm. The maximum bioethanol during syngas
30–60% and increased production of acetic acid by 3% fermentation was 3.75 at 15 mL/min gas flow rate, 0.25
to 40%. mL/min liquid flow rate, and 300 rpm. Acetic acid for-
From the plot of the above experimental results and mation reached 23.67 g/L during 15 mL/min gas flow
data, the following models were extracted to represent rate, 0.75 mL/min liquid flow rate, and 300 rpm during
bioethanol and acetic acid production: CO fermentation whereas acetic acid formation was
14.97 mL/min during syngas fermentation at 10 mL/
YCO ¼ 8  1012 t4  2  108 t3 þ 2  105 t 2 min gas flow rate, 0.25 mL/min liquid flow rate, and
300 rpm.
 0:0044 t þ 0:73 (8a)

R2CO ¼ 0:924 (8b) Product formation and gas conversions


16 6 12 5 9 4
ZCO ¼ 5  10 t  2  10 t þ 3  10 t During the fermentation process, carbon monoxide is
6 3
 3  10 t þ 0:0011 t  0:187 t
2 converted into ethanol, acetate and carbon dioxide
through the wood–ljungdahl pathway using biocatalyst
þ 19:614 (9a) Clostridium ljungdahlii. Supplied gas carbon monoxide
and hydrogen are useful for the cell growth of the
R2CO ¼ 0:9945 (9b)
microorganism. The concentration of carbon monoxide
11 4 8 3 6
YSyngas ¼  1  10 t þ 2  10 t  9  10 t 2
continuously decreases with the increase in concentra-
þ 0:0036 t þ 0:615 tion of carbon dioxide [16,18]. Carbon dioxide may
(10a) react with the hydrogen as an electron donor to yield
R2Syngas ¼ 0:987 (10b) ethanol and acetate.

ZSyngas ¼  3  1013 t5 þ 1  109 t 4  1  106 t3


(11a) i. 6CO þ 3H2 O ! CH3 CH2 OH þ 4CO2
þ 5  103 t2  0:0814 t þ 13:18 ii. 4CO þ 2H2 O ! CH3 COOH þ 2CO2
R2Syngas ¼ 0:9862 (11b) iii. 4CO þ 6H2 ! 2CH3 CH2 OH þ 2CO2
iv. 2CO2 þ6H2 ! CH3 CH2 OH þ 3H2 O
where v. 2CO2 þ4H2 ! CH3 COOH þ 2H2 O

Y = bioethanol yield in g/L either by CO or syngas Hydrogen intake increased after a couple of days of
fermentation operation of the continuous fermentation process,
Z = acetic acid yield in g/L either by CO or syngas which confirmed the utilization of homoacetic fermen-
fermentation tation pathway to produce anaerobic oxidation of
hydrogen and carbon dioxide by Clostridium ljungdahlii
8 B. ACHARYA ET AL.

Figure 3. (a) Bioethanol yield during CO and syngas fermentation; (b) acetic acid yield during CO and syngas fermentation.

[18]. Cell activity was governed by the fermentation Ethanol production increased with an increase in
pH, concentration of enzymes, substrates, cofactors, the consumption of CO and H2. However, acetic acid
effectors and inhibitors. Richter et al. stated that etha- production decreased with an increase in ethanol pro-
nol production can be improved by using the alde- duction as shown in Figure 4a. Maximum CO consump-
hyde:ferredoxin oxidoreductase and ADH route, which tion occurs when maximum ethanol production is
consist of the requisite manufacturing of acetate and a reached. The production of ethanol and acetic acid not
further reduction of un-dissociated acetic acid to etha- only depends on syngas conversion but also on pH, liq-
nol [29]. The provision of more electrons into the cul- uid flow, and gas flow. The variability observed may be
ture in the presence of 5.07 mM of cysteine-HCl in the due to the variation of other factors. Similarly, during
medium having a pH of 5.9 increases ethanol produc- the CO fermentation, bioethanol production increased
tion by 48% [30]. with an increase in CO conversion and achieved
BIOFUELS 9

Figure 4. (a) Bioethanol, acetic acid, CO and H2 conversion during syngas fermentation; (b) Bioethanol, acetic acid, and CO conver-
sion during CO fermentation.

highest levels at 75% CO conversion. However, acetic production from CO is more suitable than H2 without
acid production increased with CO conversion which dependency on gas partial pressure and pH value [31].
may be due to thin film formation at the top of the bio- Younesi et al. found that a higher amount of hydrogen
reactor or less activities of the microorganism during is used after depletion of carbon monoxide during syn-
the CO fermentation process (see Figure 4b). During gas fermentation using Clostridium ljungdahlii [32].
CO fermenation, the production of ethanol is lesser However, such a situation was never experienced here
than the acetic acid. Pressure buildup during the fer- due to continuous supply of syngas from the same
mentation process is due to an increase in the gas sup- tank.
ply rate which also retards cell activities. Experimental Due to the higher concentration of cells at the
results show that the utilization of hydrogen is less beginning, hydrogen consumption was more at the
than the amount of CO during syngas fermentation. beginning than in later stages/cycles. However, con-
This matches thermodynamically because electron sumption declined slightly because of the decrease in
10 B. ACHARYA ET AL.

cell concentration. This change in cell concentration production varies upto 22% with the liquid flow and
was due to the change in the environment including 43% with the gas supply.
pH level. However, CO conversion increased slightly
with ethanol production during syngas fermentation.
Product yields and comparisons
Cells consume CO as it starts to flow in the bioreactor
and reduces the inhibition effect of CO on hydroge- The cell density reaches at maximum during the cell
nase. CO conversion decreased with an increase in ace- growth during first 48 h of experiment. After 48 h, the
tic acid production during CO fermentation. This may CO and syngas supplied then the ethanol and acetic
be due to an increase in CO flow. The experiment was acid formation starts in the bioreactor. The measure-
conducted with a loop-back system which increased ments are carried out at different stirrer speed of
the consumption of CO and H2 due to double interac- 300 rpm and 500 rpm to observe the effect of liquid
tion of the gas with the cell-liquid of the bioreactor. flow and gas flow. Gas flow rate has more impact on
The liquid flow rate was changed every 120 h, but ethanol and acetic acid production than liquid flow
keeping same stirrer and gas flow. The ethanol rate (Figures 5a–6b). Ethanol production increased by

Figure 5. (a) Comparative study of ethanol yield-CO vs. syngas fermentation (Batch 1–9); (b) Comparative study of ethanol yield-CO
vs. syngas fermentation (Batch 10–18).
BIOFUELS 11

Figure 6. (a) Comparative study of acetic acid yield-CO vs. syngas fermentation (Batch 1–9); (b) Comparative study of acetic acid
yield-CO vs. syngas fermentation (Batch 10–18).

29–41% when liquid flow rate was doubled or tripled. 0.5 mL/min liquid flow at 300 rpm; and 1.19 g/L during
Ethanol production increased by 85–151% when gas batch 10–18 at 10 mL/min gas flow, 0.25 mL/min liquid
flow rate doubled. However, ethanol production flow at 500 rpm. During syngas fermentation, maxi-
declined after an increase in gas rate more than 100%. mum ethanol production was 3.75 g/L during batch 1–
Acetic acid production increased 18–22% by increasing 9 CO fermentation at 15 mL/min gas flow, 0.25 mL/min
liquid flow rate, and increased by 82% when gas flow liquid flow at 300 rpm; and 3.05 g/L during batch 10–
rate increased by 100–200%. This may be due to the 18 at 10 mL/min gas flow, 0.75 mL/min liquid flow at
disturbance of excessive flow of gas or formation of 500 rpm. The optimum flow rate for syngas fermenta-
acetic acid due to the continuous supply of fresh tion was 10–15 mL/min and liquid flow at 0.5–0.75mL/
media or formation of biofilm on the bioreactor. Acetic min and stirrer speed above 300 rpm but below
acid formation declined during syngas fermentation 500 rpm. CO fermentation produced less ethanol and
whereas it continued to grow during CO fermentation. more acetic acid so is not recommended to produce
During CO fermentation, maximum ethanol production ethanol. However, it could be a good option for acetic
was 1.33 g/L during batch 1–9 at 10 mL/min gas flow, acid production.
12 B. ACHARYA ET AL.

Table 1. Ethanol, acetic acid concentration andgas conversion efficiency in syngas fermentation.
CO fermentation with feedback loop (B1X) Syngas fermentation with feedback loop (B2X)
Stirrer CO flow Liquid flow CO CO Syngas Syngas CO H2
Batch speed rate mL/ rate mL/ bioethanol CO acetic conversion Syngas flow bioethanol g/ acetic acid conversion conversion
X Hours rpm min min g/L acid g/L % rate mL/min L g/L % %
1 0–120 300 5 0.25 0.41 8.92 60 5 0.92 9.14 57 59
2 121–240 300 5 0.5 0.53 10.11 67 5 1.34 11.23 63 61
3 241–360 300 5 0.75 0.75 12.15 71 5 1.55 13.44 66 45
4 36W80 300 10 0.25 0.76 14.44 60 10 1.89 14.97 70 43
5 481–600 300 10 0.5 1.33 13.88 75 10 2.84 13.88 68 42
6 601–720 300 10 0.75 1.28 14.57 68 10 3.11 12.09 71 47
7 721–840 300 15 0.25 1.32 16.27 55 15 3.75 11.64 79 38
8 841–960 300 15 0.5 1.13 18.22 57 15 3.61 10.99 75 41
9 961–1080 300 15 0.75 1.07 21.53 56 15 3.05 10.41 74 42
10 0–120 500 5 0.25 0.49 9.12 65 5 1.09 8.89 68 55
11 121–240 500 5 0.5 0.57 11.99 71 5 1.12 10.43 66 39
12 241–360 500 5 0.75 0.64 12.61 65 5 1.47 12.91 68 43
13 361–480 500 10 0.25 1.19 15.87 64 10 2.34 14.08 71 42
14 481–600 500 10 0.5 1.02 19.34 66 10 2.47 13.72 72 42
15 601–720 500 10 0.75 0.97 20.09 68 10 3.05 11.75 75 41
16 721–840 500 15 0.25 0.95 18.45 62 15 2.87 11.01 69 37
17 841–960 500 15 0.5 0.86 21.21 55 15 2.54 11.23 67 39
18 961–1080 500 15 0.75 0.81 23.67 58 15 2.12 11.34 65 45

The present results as shown in Table 1 match with Devarapalli et al. reported 5.7 g/L of ethanol using a
other publications. Shen et al. reported that ethanol TBR with the same biocatalyst [8]. The highest produc-
production increases with an increase in gas flow rate tion of ethanol reported was 23.93 g/L by Shen et al.
up to a point, then decreases with a further increase in during syngas fermentation using a hollow fiber mem-
gas flow rate [16]. Shen et al. also observed a reduction brane reactor and the same biocatalyst [16]. Aghbashlo
of bioethanol concentration with an increase in liquid et al. observed the maximum exergy efficiency of a
flow rate due to an increase in cell dilution. They sug- continuous bioreactor at 8 exergetic productive index
gested 10–15 mL/min could be the optimum limit of at media flow rate of 0.55 mL/min [36].
gas flow rate for CO and syngas fermentation for a bio- Mass transfer efficiency, which depends on the reac-
reactor. Obviously, gas flow rate varies with the design tor configuration, agitation speed, gas flow, and media
parameters of the bioreactor. Bredwell et al. and Klas- flow rate, is another important technical parameter to
son et al. observed that gas conversion is adversely improve ethanol productivity [5,7,16]. Hollow fiber
affected by high gas flow during the fermentation pro- membrane is an effective option to improve gas–liquid
cess [20,33]. Alsaker et al. and Ungerman and Heindel mass transfer [7,16]. The bubbling size and retention
observed lower ethanol concentrations by increasing time of syngas in the fermentation media can also
the gas flow rate and stirrer speed due to an increase improve gas–liquid mass transfer. These factors are
in stress on the biocatalyst [19,34]. The ethanol produc- dependent on the bioreactor design. Hence, bioreactor
tivity of the fermentation was improved by lowering design is a primary factor to improve gas–liquid mass
the cell dilution, i.e. by lowering the liquid flow rate transfer. The position of the gas supply system in the
[7,35]. However, acetic acid production increased even bioreactor may also have an impact on gas retention
with an increase in media flow rate and gas flow rate time or gas–liquid mass transfer. Another limiting factor
during the CO fermentation process after the optimum of ethanol production during syngas fermentation is
limit of flow rates [16]. the characteristics of the biofilm [16]. Double gas supply
The results of the present investigation are com- throughout the fringe of the reactor enhances the gas
pared with results of others in Table 2. Acetic acid for- retention time compared to the gas supply system
mation in the present research is higher than other installed at the center of the reactor. This may be due
findings during the CO fermentation process. In almost to the greater velocity of the gas bubbles at the fringe
all observation, syngas fermentation shifts from aceto- than at the center. Prior to the rupturing of the bubble
genic phase to solventogenic phase with the decreas- on the surface layer of the media on the bioreactor, the
ing pH level of the culture [9]. peripheral gas bubble travels horizontally and voyages
From this study (Table 2), the maximum concentra- longer time and distances. A gas loop-back system
tions of ethanol and acetic acid were 1.33 and 23.67 g/ improves gas–liquid mass transfer by improving reten-
L during the CO fermentation and 3.75 and 14.97 g/L tion time (due to improving the travel time of the bub-
during syngas fermentation. Maximum ethanol and bles) better than a single gas supply system. Sudiyo and
acetate were 7.52 and 3.43 g/L [32] during CO fermen- Andersson observed the lateral movement of bubbles
tation and 9.6 and 6.1 g/L [38] during syngas fermenta- towards the center due to stirrer vortex [37].
tion in a CSTR using Clostridium ljungdahlii as Cell retention and growth of the cell depends on the
biocatalyst. Richter et al. reported the ethanol produc- nutrition of the media. A continuous flow of fresh
tion as 2.2 g/L and Younessi et al. reported 0.55 g/L media also helps with cell density. Maintaining cell
using C. ljungdahlii in a stirred tank reactor [32]. density improves bioethanol production. The higher
Table 2 . Product yield with different reactors for CO fermentation and syngas fermentation.
CO fermentation Syngas fermentation
Reactor Biocatalyst Operation Yield (g/L) Ref Reactor Biocatalyst Operation Yield (g/L) Ref
Continuous stirred tank Clostridium ljungdahlii Yeast medium Eth:1.33 This study Continuous stirred tank Clostridium ljungdahlii Yeast medium Eth:3.77 This
pH: 4.6 AA:23.67 pH: 4.6 AA:14.97 study
Continuous stirred tank Clostridium ljungdahlii pH: 4 Eth: 7.52g/L [32] Two stage continuous still Clostridium ljungdahlii pH: 4.75 Eth:2.2 [7]
Temp: 37 Ac: 3.43g/L tank Temp: 37 AA:1.9
Cell density: 10g/L
Continuous stirred tank Butyribacterium pH: 6–6.8 Acetate: 7.96 [32] Continuous stirred tank Clostridium ljungdahlii pH: 4.5 Eth+ AA:11 Cell density: [43]
methylotrophicum Temp: 37 mmol/L Temp:37 2g/L
Butyrate: 5.95 g/L
Gas lift reactor Eubacterium limosum pH: 6.8 Acetate: 0.1mmol/ [38] Continuous stirred tank Clostridium ljungdahlii - Eth:0.55 AA:1.3 [32]
Temp: 37 L Cell density: 10g/L
Eth: 38mmol/L
Continuous gas feed Clostridium pH:4.75 Eth: 0.91 [9] Stirred tank Clostridium ljungdahlii Yeast medium Eth:0.2 [22]
autoethanogenum Tungstem AA: 0.91 pH: 5.5 AA:2.2
Continuous CO feed Clostridium pH: 5.75 Eth: 5.55 [39] Trickle bed co-current Clostridium ragsdalei Yeast Extract, pH:46 Eth:5.7 [8]
carboxidivorans Butanol: 2.66 AA:12.3
Optimized medium and one single Clostridium pH: 4.75–5.75 Eth: 4.26 [40] Trickle bed counter-current Clostridium ragsdalei Yeast Extract, pH:46 Eth:3.0
gas-fed autoethanogenum AA: N/A AA:5.3
Gas lift reactor Eubacterium limosum pH: 6.8 Eth:0.2 [41] Stirred tank Clostridium ragsdalei Corn Steep Liquor, pH: no Eth:9.6 [38]
Temp: 37 AA:2.2 control AA:6.1
Cell density:
0.75g/L
Continuous stirred tank Carboxydothermus 6.8-7.0 Eth: 0.09mM [42] Stirred tank Clostridium Yeast Medium, pH: No Eth:2.2 [44]
hydrogenoformans Ac:0.51mM carboxidivorans control AA:1.9
Hollow fiber membrane Clostridium Yeast Medium, pH: 4.5–5.5 Eth: 23.93 [16]
stirred tank carboxidivorans AA: 4.79
Stirred tank Clostridium Modified Mineral Medium Eth:2.2 [9]
autoethanogenum pH: 4.75 AA:1.9
BIOFUELS
13
14 B. ACHARYA ET AL.

Figure 7. (a) Batch comparison of ethanol yield-CO vs. syngas fermentation (Batch 1–18); (b) Batch comparison of acetic acid yield-
CO vs. syngas fermentation (Batch 1–18).

ethanol concentration observed may be due to an ljungdahlii resulting in less cell productivity. Ethanol
increase in the gas–liquid mass transfer by improving production at either stirrer speed increased with an
the retention time due to the feedback gas flow system increase in liquid flow rate and gas flow rate. Liquid
and continuous extraction system. flow rate has less impact on ethanol and acetic acid
The concentration profile of ethanol and acetic acid production than gas flow rate in both fermentation
throughout the experiment from batch B101–118 and processes. Ethanol production started declining after
B201–218 are presented in Figures 7a and b. The con- Batch #107 and #115 during syngas fermentation and
centration of acetic acid increased almost linearly for after Batch #105 and #112 during CO fermentation.
stirrer speeds 300 and 500 rpm during the CO fermen- This may be due to declining acetic acid in the syngas
tation; concentration of acetic acid is maximum for fermentation and increasing acetic acid formation in
batch B205 and B214. Ethanol production was slightly the CO fermentation process. Acetogenesis is more
better at 300 rpm than 500 rpm (Figures 7a,b). This pronounced in CO fermentation, whereas solventogen-
may be due to disturbance of the Clostridium esis is more pronounced in syngas fermentation.
BIOFUELS 15

Mass transfer analysis To enhance the mass transfer, the presence of gas
has been increased by the gas loop-back system and
The mass transfer coefficient of CO, syngas-CO, and
the cell density remains almost at the steady state due
syngas-H2 was plotted and compared as in Figure 8.
to a continuous supply of nutrition with replacement
Mass transfer is always higher in syngas-CO than CO
of fresh media. Gas–liquid mass transfer can be
alone. The mass transfer coefficient increased with an
improved by using a proper size of gas bubble diame-
increase in run time and reached a maximum at B27
ter because it enhances the specific surface area avail-
with 34.02 (1/h) during syngas fermentation, and B15
able for mass transfer. The present gas orifice system
with 24.74 (1/h) during CO fermentation. Klasson et al.
generates small bubbles travelling slowly in a circular
and Bredwell et al. observed mass transfer coefficients
fashion which increases the gas residence time in the
up to 35.5 (1/h). The mass transfer coefficient declined
bioreactor and improves gas–liquid mass transfer
during later batches. This may be due to formation of
which then improves the mass transfer coefficient.
more acetic acid in the CO fermentation than syngas
Orgill et al. noted a decrease in gas–liquid mass trans-
fermentation. The mass transfer coefficient increased
fer coefficient of CO and H2 with an increase in liquid
up to 38% with syngas-CO and 30% with CO with an
recirculation rate or liquid holdup volume VL [17]. Simi-
increase in the gas flow rates; and decreased up to
lar observations were found at higher liquid flow rates
20% with syngas-CO and up to 11% with 100% CO in
during this experiment.
tripling the liquid flow rate. Cell activity and gas uptake
also affected the coefficient. This may be due to an
increase in volume VL. The depleted nutrition level
decreased the cell density which consequently Conclusions
decreased the mass transfer coefficient.
Continuous production of bioethanol in addition to
The major limiting factor in the fermentation pro-
acetic acid from carbon monoxide fermentation and
cess is the potential bottleneck of the mass transfer of
syngas fermentation using a biocatalyst Clostridium
gas into liquid. This is more severe for carbon monox-
ljungdahlii in a designed and developed economical
ide, hydrogen, and its mixture in comparison with the
continuous stirred tank bioreactor has been proved
aerobic process because the solubility of CO and H2 on
successfully. We compared bioethanol production
mass basis with respect to oxygen is only 60% and 4%
using two different gases (CO and syngas) using an
respectively [20]. Mass transfer limitation may arise
innovative gas loop-back system. In both fermentation
due to gas–liquid interface, transport of gas to media,
processes, bioethanol production was closely moni-
diffusion of gas into the liquid layer, and diffusion of
tored by varying pH level, broth media flow rates, gas
the gaseous substrate to the intracellular reaction site
flow rates, and stirrer speed. The pH level impacts bioe-
[45]. Improvements in impeller designs, liquid flow pat-
thanol production, gas consumption, and cell growth
terns, mixing residence time, baffle design, aerated
environment. Maximum bioethanol production was
power efficiency, and application of bubble dispersers
1.33 g/L during CO fermentation at 10 mL/min gas
are but a few parameters to enhance the mass transfer
flow, 300 rpm, 0.5 mL/min liquid flow and 3.75 g/L dur-
capability [5].
ing syngas fermentation at 15 mL/min gas flow,
300 rpm, 0.25 mL/min liquid flow. Similarly, the maxi-
mum acetic acid production was 23.67 g/L during CO
fermentation at 15 mL/min gas flow, 500 rpm, 0.75 mL/
min liquid flow and 14.97 g/L during syngas fermenta-
tion at 10 mL/min gas flow, 300 rpm, 0.25 mL/min liq-
uid flow. The higher amount of acetic acid is due to the
continuous process where the fresh broth media (addi-
tion of nutrients) is supplied continuously in the biore-
actor. The addition of nutrients favors the growth of
the microorganism. The maximum gas consumption
observed was 75% during CO fermentation and 79%
during syngas (65% CO, 30% H2, and 5% CO2) fermen-
tation. The syngas fermentation process is better than
the CO fermentation process for bioethanol production
because this process produced more ethanol. Further
research is recommended for the conversion of acetic
acid into bioethanol either by replacing the gas supply
system or by improving the engineering design param-
Figure 8. Mass transfer coefficient of CO, syngas (CO) and syn- eters of the bioreactor to improve the mass transfer
gas (H2). coefficient.
16 B. ACHARYA ET AL.

Disclosure statement and ethanol production performance. Biochem Engg J.


2014;85:21–29.
No potential conflict of interest was reported by the authors. [17] Orgill JJ, Atiyeh HK, Devarapalli M, et al. A comparison of
mass transfer coefficients between trickle-bed, hollow
fiber membrane and stirred tank reactors. Bioresour
ORCID Technol. 2013;133:340–346.
[18] Mohammadi M, Younesi H, Najafpour G, et al. Sustain-
Animesh Dutta http://orcid.org/0000-0002-9995-807X
able ethanol fermentation from synthesis gas by Clos-
tridium ljungdahlii in a continuous stirred tank
bioreactor. J Chem Technol Biotechnol. 2012;87:837–
References
843.
[1] Subramani V, Gangwal SK. A review of recent literature [19] Ungerman AJ, Heindel TJ. Carbon monoxide mass trans-
to search for an efficient catalytic process for the con- fer for syngas fermentation in a stirred tank reactor with
version of syngas to ethanol. Energ Fuel. 2008;22:814– dual impeller configurations. Biotechnol Prog.
839. 2007;23:613–620.
[2] Jiang Y, Liu J, Jiang W, et al. Current status and pros- [20] Bredwell MD, Srivastava P, Worden RM. Reactor design
pects of industrial bio-production of n-butanol in China. issues for synthesis-gas fermentations. Biotechnol Prog.
Biotechnol Adv. 2015;33:1493–1501. 1999;15:834–844.
[3] Datta R, Maher MA, Jones C, et al. Ethanol-the primary [21] Alsaker KV, Paredes C, Papoutsakis ET. Metabolite stress
renewable liquid fuel. J Chem Technol Biotechnol. and tolerance in the production of biofuels and chemi-
2011;86:473–480. cals: gene-expression-based systems analysis of buta-
[4] Acharya B, Roy P, Dutta A. Review of syngas fermenta- nol, butyrate, and acetate stresses in the anaerobe
tion processes for bioethanol. Biofuels. 2014;5(5):551– Clostridium acetobutylicum. Biotechnol Bioeng.
564. 2010;105:1131–1147.
[5] Munasinghe PC, Khanal SK. Syngas fermentation to bio- [22] Cotter JL, Chinn MS, Grunden AM. Influence of process
fuel: evaluation of carbon monoxide mass transfer coef- parameters on growth of Clostridium ljungdahlii and
ficient (kLa) in different reactor configurations. Clostridium autoethanogenum on synthesis gas.
Biotechnol Progr. 2010;26(6):1616–1621. Enzyme Microb Technol. 2009;44:281–288.
[6] Roy P, Dutta A, Chang S. Development and evaluation of [23] Sherwood TK, Pigford RL, Wilke CR. Mass transfer. New
a functional bioreactor for CO fermentation into etha- York: McGraw-Hill Inc; 1975.
nol. Bioresour Bioproc. 2016;3(1):1–8. [24] Ahmed T, Semmens MJ. Use of sealed and hollow fibers
[7] Richter H, Martin ME, Angenent LT. A two-stage continu- for bubble less membrane aeration: experimental stud-
ous fermentation system for conversion of syngas into ies. J Membr Sci. 1992;60(1-2):1–10.
ethanol. Energies. 2013;6:3987–4000. [25] Skidmore BE. Syngas fermentation: quantification of
[8] Devarapalli M, Atiyeh HK, Phillips JR, et al. Ethanol pro- assay techniques, reaction kinetics and pressure depen-
duction during semi-continuous syngas fermentation in dencies of the Clostridium P11 hydrogenase [Master's
a trickle bed reactor using Clostridium ragsdalei. Biore- thesis]. Provo, UT: Bringham Young University; 2010.
sour Technol. 2016;209:56–65. [26] Klasson KT, Ackerson CMD, Clausen EC, et al. Biological
[9] Abubackar HN, Veiga MC, Kennes C. Biological conver- conversion of synthesis gas into fuels. Int J Hydrogen
sion of carbon monoxide: rich syngas or waste gases to Energ. 1992;17(4):281–288.
bioethanol. Biofuels Bioprod Bioref. 2011;5:93–114. [27] Abubackar HN, Bengelsdorf FR, Durre P, et al. Improved
[10] Liu K, Atiyeh HK, Tanner RS, et al. Fermentative produc- operating strategy for continuous fermentation of car-
tion of ethanol from syngas using novel moderately bon monoxide to fuel-ethanol by clostridia. Appl
alkaliphilic strains of Alkalibaculum bacchi. Bioresour Energy. 2016;169:210–217.
Technol. 2012;104:336–341. [28] Kopke M, Held C, Hujer S, et al. Clostridium ljungdahlii
[11] Huhnke RL, Lewis RS, Tanner RS. Isolation and character- represents a microbial production platform based on
ization of novel Clostridial Species. US Patent No. syngas. Proc Natl Acad Sci. 2010;107(29):13087–13092.
7704,723. 2010. [29] Richter H, Molitor B, Wei H, et al. Ethanol production in
[12] Saxena J, Tanner R. Optimization of a corn steep syngas-fermenting Clostridium ljungdahlii is controlled
medium for production of ethanol from synthesis gas by thermodynamics rather than by enzyme expression.
fermentation by Clostridium ragsdalei. World J Microbiol Energy Environ Sci. 2016;9:2392–2399.
Biotechnol. 2012;28:1553–1561. [30] Mohammadi M, Mohamed AR, Najafpour G, et al. Clos-
[13] Phillips JR, Remondet NM, Atiyeh HK, et al. Designing tridium ljungdahlii for production of biofuel from syn-
Syngas Fermentation Medium for Fuels and Bulk Chemi- thesis gas. Energy Sources, Part A: Recovery, Utilization,
cals Production. Proceedings of 2011 ASABE Annual and Environmental Effects. 2016;38(3):427–434.
International Meeting; August 7-10; Louisville, KY, USA, [31] Hu P, Bowen SH, Lewis RS. A thermodynamic analysis of
Paper # 1111052; p. 1–12; 2011. electron production during syngas fermentation. Biore-
[14] Xu D, Tree DR, Lewis RS. The effects of syngas impurities sour Technol. 2011;102:8071–8076.
on syngas fermentation to liquid fuels. Biomass Bioen- [32] Younesi H, Najafpour G, Mohamed AR. Ethanol and ace-
ergy. 2011;35:2690–2696. tate production from synthesis gas via fermentation
[15] Vane LM, Alvarez FR, Huang Y, et al. Experimental valida- processes using anaerobic bacterium, Clostridium ljung-
tion of hybrid distillation-vapor permeation process for dahlii. Biochem Eng J. 2005;27:110–119.
energy efficient ethanol-water separation. J Chem Tech- [33] Klasson KT, Ackerson CMD, Clausen EC, et al. Biological
nol Biotechnol. 2010;85:502–511. conversion of coal and coal-derived synthesis gas. Fuel.
[16] Shen Y, Brown R, Wen Z. Syngas fermentation of Clos- 1993;72:1673–1678.
tridium carboxidivorans P7 in a hollow fiber membrane [34] Alsaker KV, Paredes C, Papoutsakis ET. Metabolite stress
biofilm reactor: evaluating the mass transfer coefficient and tolerance in the production of biofuels and
BIOFUELS 17

chemicals: gene-expression-based systems analysis of [40] Abubackar HN, Veiga MC, Kennes C. Carbon monoxide
butanol, butyrate, and acetate stresses in the anaerobe fermentation to ethanol by Clostridium autoethanoge-
Clostridium acetobutylicum. Biotechnol Bioeng. num in a bioreactor with no accumulation of acetic
2010;105:1131–1147. acid. Bioresour Technol. 2015;186:122–127.
[35] Kundiyana DK, Huhnke RL, Wilkins MR. Syngas fermenta- [41] Chang IS, Kim DH, Kim BH, et al. CO fermentation of
tion in a 100-L pilot scale fermentor: design and process Eubacterium limosum KIST612. J Microbiol Biotechnol.
considerations. J Biosci Bioeng. 2010;109(5):492–498. 1998;8:134–140.
[36] Aghbashlo M, Tabatabaei M, Dadak A, et al. Exergy- [42] Haddad M, Cimpoia R, Guiot SR. Performance of Carbox-
based performance analysis of a continuous stirred bio- ydothermus hydrogenoformans in a gas-lift reactor for
reactor for ethanol and acetate fermentation from syn- syngas upgrading into hydrogen. Int J Hydrogen Energ.
gas via Wood–Ljungdahl pathway. Chem Eng Sci. 2014;39:2543–2548.
2016;143:36–46. [43] Younesi H, Najafpour G, Mohamed AR. Liquid fuel pro-
[37] Sudiyo R, Andersson B. Bubble trapping and coales- duction from synthesis gas via fermentation process in
cence at the baffles in stirred tank reactors. AIChE J. a continuous tank bioreactor (CSTBR) using Clostridium
2007;53:2232–2239. ljungdahlii. Iranian J Biotechnol. 2006;4:45–53.
[38] Maddipati P, Atiyeh HK, Bellmer DD, et al. Ethanol pro- [44] Ukpong MN, Atiyeh HK, De Lorme MJ, et al. Physiologi-
duction from syngas by Clostridium strain P11 using cal response of Clostridium carboxidivorans during con-
corn steep liquor as a nutrient replacement to yeast version of synthesis gas to solvents in a gas-fed
extract. Bioresour Technol. 2011;102:6494–6501. bioreactor. Biotechnol. Bioeng. 2012;109:2720–2728.
[39] Fernandez-Naveira A, Abubackar HN, Veiga MC, et al. [45] Mohammadi M, Najafpour GD, Younesi H, et al. Biocon-
Efficient butanol-ethanol (BE) production from carbon version of synthesis gas to second generation biofuels:
monoxide fermentation by Clostridium carboxidivorans. a review. Renew Sustain Energy Rev. 2011;15:4255–
Appl Microbiol Biotechnol. 2010;100(7):3361–3370. 4273.

You might also like