Você está na página 1de 9

Bioresource Technology 172 (2014) 41–49

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Enrichment and optimization of anaerobic bacterial mixed culture for


conversion of syngas to ethanol
Ashish Singla a,b, Dipti Verma b, Banwari Lal a,b, Priyangshu M. Sarma a,b,⇑
a
TERI University, 10 Institutional Area, Vasant Kunj, New Delhi 110 070, India
b
TERI, Darbari Seth Block, India Habitat Centre, New Delhi 110 003, India

h i g h l i g h t s

 Anaerobic bacterial mixed culture enriched for conversion of syngas to ethanol.


 Operational parameters optimized for enhancing ethanol production from syngas.
 Semi-continuous fermentation study done for getting increased ethanol production.
 Up-scaling studies done for further enhancing ethanol production from syngas.

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

Article history: The main aim of the present study was to enrich anaerobic mixed bacterial culture capable of producing
Received 13 June 2014 ethanol from synthesis gas fermentation. Screening of thirteen anaerobic strains together with enrich-
Received in revised form 18 August 2014 ment protocol helped to develop an efficient mixed culture capable of utilizing syngas for ethanol pro-
Accepted 19 August 2014
duction. Physiological and operational parameters were optimized for enhanced ethanol production.
Available online 27 August 2014
The optimized value of operational parameters i.e. initial media pH, incubation temperature, initial syn-
gas pressure, and agitation speed were 6.0 ± 0.1, 37 °C, 2 kg cm2 and 100 rpm respectively. Under these
Keywords:
conditions ethanol and acetic acid production by the selected mixed culture were 1.54 g L1 and 0.8 g L1
Syngas
Mixed culture
respectively. Furthermore, up-scaling studies in semi-continuous fermentation mode further enhanced
Optimization ethanol and acetic acid production up to 2.2 g L1 and 0.9 g L1 respectively. Mixed culture TERI SA1
Ethanol was efficient for ethanol production by syngas fermentation.
Up-scaling Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction are also used to convert syngas components into a variety of fuels
and chemicals such as hydrogen, methane, methanol, ethanol, and
Synthesis (syngas) gas, primarily a mixture of CO, CO2 and H2 is acetic acid (Klasson et al., 1992). Biological processes, although rel-
a major feedstock in production of many fuels and chemicals. It can atively slower than chemical reaction have several advantages over
be produced from gasification of several materials such as coal, catalytic process such as elimination of expensive metal catalyst,
wood, municipal solid waste and lignocellulosic biomass (Phillips higher specificity of biocatalyst, lower energy costs, greater resis-
et al., 1993). The essential syngas components CO, CO2 and H2 tance to catalyst poisoning and independence of H2/CO ratio
can be biologically converted into ethanol and other value added (Wolfrum and Watt, 2002). The biological reaction occurs under
compounds such as acetic acid, 2-butanol, n-propanol and polyhy- ambient condition of pH, temperature and pressure with the for-
droxyalkanoate (PHA) (Kundiyana et al., 2010). Catalytic processes mation of specific products. However, direct production of fuels
and chemicals from gasification technology is economically unfa-
vorable and requires very large infrastructure as about 60% of the
Abbreviations: PHA, polyhydroxyalkanoate; PBM, Pfennig’s basal media; TCD, total investment cost in modern methanol plants is accounted for
thermal conductivity detector; FID, flame ionization detector; OD, optical density; syngas generation (Vannby and Winter Madsen, 1992). Therefore,
CO, carbon monoxide; H2, hydrogen; rs, resparged; AA, acetic acid; EtOH, ethanol; the economical advantage of biological processes through develop-
SD, Sludge; CD, cow dung; CF, chicken faeces; VFA, volatile fatty acids.
⇑ Corresponding author at: TERI University, 10 Institutional Area, Vasant Kunj, ment of suitable biocatalyst to ferment gaseous substrate to valu-
New Delhi 110 070, India. Tel.: +91 11 24682100; fax: +91 11 24682144. able products can be considered. Previous study indicates that FT
E-mail address: priyanms@teri.res.in (P.M. Sarma). process has a relative overall energy efficiency of 45%, while gas

http://dx.doi.org/10.1016/j.biortech.2014.08.083
0960-8524/Ó 2014 Elsevier Ltd. All rights reserved.
42 A. Singla et al. / Bioresource Technology 172 (2014) 41–49

fermentation has an overall energy efficiency of 57%, in terms of gas. In addition, semi-continuous and scale-up fermentation stud-
energy in feedstock converted to final product (Griffin and ies from 130 mL to 1 L volumes were also carried out to enhance
Schultz, 2012). the ethanol production.
Microorganisms act as biocatalysts to convert syngas into
chemicals and fuels. Anaerobic bacteria such as Clostridium
2. Methods
ljungdahlii, Clostridium ragsdalei, Clostridium carboxidivorans,
Clostridium aceticum, Acetobacterium woodii and Clostridium ther-
2.1. Source of inoculum and fermentation medium
moaceticum convert biomass generated syngas composed of CO,
CO2 and H2 to ethanol and acetic acid (Kundiyana et al., 2010).
Micro-organisms used for the present study were taken from
Bioethanol production through acetogenic fermentation of the
TERI’s culture collection. TERI is a referral centre and culture repos-
gaseous substrates (CO, CO2 and H2) follows the Wood-ljungdahl
itory for bacterial strains in India (Sarma et al., 2004; Agrawal et al.,
and acetyl-CoA pathways. Syngas fermentation for ethanol pro-
2010; Kaur et al., 2009; Singh et al., 2014). An additional approach
duction based on the acetyl-CoA metabolic pathway is an emerg-
of enrichment was also employed using sludge (SD), cow dung
ing area and the process require significant research
(CD) and chicken faeces (CF) samples collected from Gwal pahari
interventions. Several research groups have explored the use of
(Gurgaon, Haryana, 28°280 N 77°1.90 E) India. Pfennig’s basal media
anaerobic bacteria to convert syngas to ethanol (Gaddy and
(PBM) medium containing (per liter) 50 mL mineral stock solution,
Clausen, 1992; Younesi et al., 2005). However, these studies have
10 mL trace metal stock solution, 10 mL vitamin stock solution, 1 g
yet to define a methodology for generating high ethanol produc-
yeast extract, 1 mL resazurin (0.1%) and 20 mL L-cysteine HCl
tion levels with a stable culture. The selection of appropriate
(2.5%) was used for the enrichment of microbial strains. The com-
microbes for efficient syngas fermentation is often a challenging
position of the mineral, vitamin and trace metal stock solutions has
task. Therefore, the isolation and engineering of new microbial
been previously reported (Gaddy and Clausen, 1992). Unless spec-
species, which are more productive and robust, need to be
ified, all the experiments were performed in 130 mL Wheaton
employed. Results from published literature show that tempera-
serum bottles containing 40 mL of the liquid medium. Anaerobic
ture, pH, reducing agents, redox potential, agitation speed, gas
condition during media preparation was maintained as described
partial pressure, gas compositions and media components have
by (Singh et al., 2014). The bottles were sealed with butyl rubber
significant effects on cell growth and ethanol production (Hurst
stopper and aluminum cap and then sterilized at 121 °C and
and Lewis, 2010). Optimization of these parameters is critical
15 psi for 15 min. Vitamin solution was added after sterilization.
for assessing the potential sustainability of ethanol from syngas.
The initial pH of the medium was adjusted at 6.0 ± 0.1 with 2 N
In the recent years, a general understanding regarding cell
KOH and 2 N HCl solutions.
growth and metabolism leading to higher conversion of syngas
to liquid fuel has been made. Syngas fermentation using Clostrid-
ium P7 showed that increase in CO partial pressure from 0.35 to 2.2. Screening and enrichment of syngas utilizing microbial mixed
2.0 atm. resulted in 440% improvement in cell growth and etha- culture for ethanol production
nol production shifted from non-growth associated to growth
associated phase (Hurst and Lewis, 2010). Studies on agitation In the present study initially 13 microbial cultures were col-
speed showed 50% decrease in ethanol production when agitation lected from TERI’s culture collection (Sarma et al., 2004; Agrawal
speed was increased to 250 rpm from 150 rpm (Ramachandriya, et al., 2010; Kaur et al., 2009; Singh et al., 2014) and screened for
2006). Similar studies on bioreactor design and up-scaling syngas fermentation to ethanol. An additional approach of enrich-
parameters showed six fold increase in ethanol production when ment was also designed. Sludge, cow dung and chicken faeces sam-
syngas fermenting strain Clostridium P11 was scaled up from 7.5 L ples were used for the enrichment of syngas utilizing microbial
to 100 L pilot scale fermentor (Kundiyana et al., 2010). Further- mixed culture. All samples were subjected to multiple enrichment
more, studies on the aspects of process optimization have also cycles by adding (10% w/v) sample inoculum into 125 mL of Whea-
enhanced the conversion rates and ethanol production ton serum bottle with 40 mL of freshly prepared medium. The cul-
(Kundiyana et al., 2011). ture bottles were sparged and headspace pressurized up to
Based on the points highlighted above, the present study was 1 kg cm2 with commercial syngas (CO 20%, CO2 15%, H2 20%,
aimed at developing anaerobic mixed bacterial culture that have CH4 3% balance N2). Incubation was done at 37 °C in a rotary shaker
the capability to utilize syngas for ethanol production. However at 150 rpm. After 5 days (10% v/v) enriched culture was subse-
there is a considerable number of studies on syngas fermentation quently transferred into fresh medium and this process was
using mesophilic pure cultures (Tanner, 2005; Maddipati et al., repeated for 6 cycles. During each sub-culturing cycle the culture
2011). Limited studies using mixed-cultures have been reported, bottles were analyzed for syngas utilization, Volatile fatty acids
although mixed culture fermentation is more suited for industrial (VFA) and solvent production (as described in Section 2.6). The
applications, when compared to pure culture fermentation. Some enrichment cycles were done in duplicates.
of the advantages are: (i) no need for highly sterile cultivation,
(ii) presence of high microbial diversity, which offers increased 2.3. Optimization of operational parameters for enhanced ethanol
adaptation capacity, (iii) possibility of mixed substrates co-fermen- production by the selected mixed culture
tation, and (iv) higher capacity for continuous processing. The
mixed culture presents an opportunity for higher alcohols produc- Based on enrichment results, selected syngas fermenting mixed
tion from syngas. Semi-continuous fermentations in a 3 L fermen- culture was taken up for further study. Different operational
tor with the mixed culture and CSL medium resulted in a twofold parameters were investigated for enhanced ethanol production.
more total alcohol production than in the YE medium. The synergy To investigate the effect of temperature on growth and ethanol
between strain CP15 and Clostridium propionicum in the mixed cul- production by the selected mixed culture, the experimental set
ture in bottle fermentations resulted in 50% higher efficiency in was incubated at temperatures 30 °C, 37 °C and 50 °C at pH 6.0,
converting propionic acid, butyric acid and hexanoic acid to their agitation speed 150 rpm and syngas pressure in the reactor bottle
respective alcohol (Liu et al., 2013). Further, different operational headspace of 1 kg cm2. Effect of initial medium pH on ethanol
parameters such as pH, temperature, syngas pressure and agitation production by selected mixed culture TERI SA1 was studied with
speed were optimized for enhanced ethanol production from syn- the pH 5.0, 6.0, 7.0 and 8.0. The pH was adjusted with 2 N HCl
A. Singla et al. / Bioresource Technology 172 (2014) 41–49 43

and 2 N KOH solutions. Similarly the effect of different initial syn- 3. Results and discussion
gas pressure in the reactor bottle headspace of 0.5, 1.0, 1.5, 2.0, 2.5
and 3.0 kg cm2 and different rate of agitation (100, 150 and 3.1. Screening and enrichment of syngas utilizing microbial mixed
200 rpm) were studied for analyzing the effect on ethanol produc- culture for ethanol production
tion. All the experiments were performed in duplicate and the data
points are average of the duplicate ± standard deviation (less than The aim of the present study was to develop anaerobic mixed
5% of average). bacterial culture capable of utilizing syngas as carbon source to
produce ethanol. In this study, natural anaerobes obtained from
sludge, cow dung and chicken faeces and thirteen cultures
2.4. Semi-continuous fermentation studies for enhanced ethanol obtained from TERI’s culture collection were examined for maxi-
production by mixed culture TERI SA1 mum ethanol production by utilizing syngas as carbon source. In
most natural environments, anaerobes are predominant and have
The experiment was conducted in semi-continuous mode significant ability to transform organics to alcohols (Fan et al.,
(sparged with fresh syngas after every 24 h and fermentation in 2003). In addition, thirteen bacterial strains from TERI’s culture
batch mode) for enhanced ethanol production by the selected collection were selected because this centre is a culture repository
mixed culture. The change in pH, cell biomass, acetic acid and eth- of anaerobic microbes such as Clostridium sps., isolated from vari-
anol production was analyzed at regular time interval of 24 h for a ous sources (Agrawal et al., 2010; Kaur et al., 2009; Singh et al.,
period of 216 h. The culture bottles were resparged after every 24 h 2014). Moreover, Chicken faeces are a good source for enrichment
with fresh syngas and maintained with the syngas pressure of of anaerobic mixed bacterial culture for conversion of syngas to
2 kg cm2 in the headspace. The experiment was performed in ethanol. Gaddy and Clausen (1992) have been previously reported
duplicates. a mesophilic syngas fermenting bacterial strain (C. ljungdahlii)
belonging to a novel species from the similar source. Out of the
thirteen strains of TERI culture collection and three sample inocu-
lums, the mixed culture enriched from chicken faeces utilized syn-
2.5. Scale up studies using mixed culture TERI SA1 for increased
gas as carbon source to produce maximum ethanol and acetic acid,
ethanol production
and was selected for subsequent studies (Fig. 1). Other enriched
strains did not utilize syngas efficiently to produce ethanol as they
Scale up studies for the selected mixed culture was performed
might not have adapted for utilizing syngas as carbon source. After
in 1 L wheaton serum bottles with 300 mL of medium in semi-
4th enrichment cycle, the selected mixed culture TERI SA1enriched
continuous mode at pH 6.0, temperature 37 °C and agitation of
from chicken faeces was capable of utilizing 100% of the initial sup-
100 rpm. The syngas pressure in the reactor bottle headspace
plied CO. Each enrichment cycle was of 120 h. The maximum
was maintained at 1 kg cm2. Sample was taken for analyzing
amount of ethanol and acetic acid produced by mixed culture TERI
change in pH, cell growth, acetic acid and ethanol production
SA1 after 5th enrichment was 0.15 g L1 and 1.21 g L1 respec-
at regular time interval of 24 h for a period of 216 h. The reactor
tively (Fig. 1). Thus, TERI SA1 being most efficient was taken up
bottle headspace gas was replaced with fresh syngas and main-
for further optimization studies. Although similar study has been
tained with same syngas pressure after every 24 h. The experi-
done previously but there are crucial differences between the
ment was performed in duplicate and the data points are
strain reported in the previous study (Gaddy and Clausen, 1992)
average of the duplicate ± standard deviation (less than 5% of
and the microbial mixed culture enriched in the current study, (i)
average).
The sample was collected from similar source but different geo-
graphical location, (ii) The bacterial diversity present in the sample
varies according to the climate conditions as well as eating habit of
2.6. Analytical methods the animal living in the particular region, (iii) The current study
was related to the enrichment and optimization of a mixed bacte-
Gas phase concentrations in the reactor bottle headspace was rial culture rather than previous pure culture study by (Gaddy and
analyzed by using Agilent Technologies 7890A series gas chro- Clausen, 1992), (iv) Moreover, mixed culture fermentation is cost
matograph (GC) equipped with thermal conductivity detector effective approach than pure culture.
(TCD). GC was fitted with 4 m long Carbosieve-II (80/100 mesh
size) 11.80  2 mm SS column (Nucon, India). Argon was used
as carrier gas with the flow rate of 11.4 mL/min. The oven, injec- 3.2. Optimizing operational parameters for maximizing ethanol
tor and detector temperature were 100 °C, 120 °C and 150 °C, production
respectively. For VFAs and solvent analysis, liquid samples were
centrifuged at 10,000 rpm for 10 min and supernatant was fil- 3.2.1. Effect of pH on cell growth and ethanol production
tered with 0.22 lm millipore membrane filters. Ethanol and pH is a critical parameter to obtain optimal microbial activity in
acetic acid concentration was analyzed using flame ionization the culture media. The extracellular pH directly influences the
detector (FID) equipped with DB-WAXETR 30 m  0.53 mm  intracellular pH, membrane potential, proton motive force, and
1.0 micron packed column (Agilent 7890, USA). Argon was used consequently substrate utilization and product profile (Devi
as the carrier gas at the flow rate of 0.947 mL/min. The injector et al., 2010). There was an increase in ethanol production and cell
and detector temperature were 220 °C and 250 °C, respectively. biomass as pH increased from 5.0 to 6.0 (Fig. 2a). The maximum
The oven temperature was set at 60 °C for 1 min and increased cell biomass (0.14 g L1), ethanol (1.48 g L1) and acetic acid and
to 170 °C and 200 °C at ramping rate of 20 °C/min and (0.6 g L1) production by mixed culture TERI SA1 was observed
10 °C/min respectively. The calibration curve obtained for all when initial pH of the medium was 6.0. There is a narrow range
the standards showed R2 value more than 0.998. The bacterial of pH for every organism, in which the cells are metabolically
growth was monitored by measuring the optical density (OD) active. A change in pH of environment can lead to death of the cells
at 600 nm using UV visible spectrophotometer (UV-2450, and consequently results in loss of biological activity. Therefore,
Shimadzu, Japan). Initial and final pH was checked by pH meter the metabolite production decreased gradually as the pH increased
(Mettler Toledo, India). beyond 6.0, confirming it as the optimum pH value for the selected
44 A. Singla et al. / Bioresource Technology 172 (2014) 41–49

Enrichment study
0.16 1.6

Acetic acid (g L-1)


Ethanol (g L-1)
0.12 1.2

0.08 0.8

0.04 0.4

0 0

S3
HS
S10

S25

T01

T08A
T08
T03
T05
SD

T08b
CD

IT2
TM9A

MT55
TERI SA1
Enrichment cultures
Ethanol Acec acid
Fig. 1. Acetic acid and ethanol production pattern of the enrichment cultures. The experiment was performed in duplicate and the data points are average of the
duplicate ± standard deviation (less than 5% of average).

mixed culture (Fig. 2a). Similar pH range has also been reported for and metabolite production decreased gradually as the agitation
syngas fermenting bacterial strains Clostridium autoethanogenum, speed was further increased to 150 and 200 rpm (Fig. 2c). Cell lysis
Butyribacterium methylotropphicum (Abrini et al., 1994) and (Shen of the mixed culture TERI SA1 was observed at high agitation rates.
et al., 1999). In addition, Rajagopalan et al. (2002) observed maxi- The drop in ethanol and acetic acid formation at 200 rpm followed
mum ethanol concentration of 0.56 g L1 with C. carboxidivorans P7 by 150 rpm could possibly be due to the shear sensitivity and CO
at similar pH range. pH optimization studies for the selected mixed toxicity of mixed culture TERI SA1 to higher agitation rates at high
culture led to increased ethanol production in comparison to the syngas pressure (Gaddy et al., 2003; Ramachandriya, 2006).
previously reported in literature by (Rajagopalan et al., 2002). Enzymes such as CODH and hydrogenase involved in the acetyl-
CoA pathway could be affected due to the shear and the presence
3.2.2. Effect of temperature on cell growth and ethanol production of excess CO also inhibits growth by inhibiting the CO-dehydroge-
Temperature changes have significant effect upon microbial nase. Moreover, presence of excess CO unfortunately results in
growth and activity. In most cases, the temperature of the culture poor H2 conversion which eventually causes cell lysis (Gaddy
media is decided based on the specific microorganism. The fermen- et al., 2003; Ramachandriya, 2006). Ramachandriya (2006)
tation temperature not only affects substrate utilization, growth observed that ethanol concentration decreased by 50% when P11
rate and membrane lipid composition of the acetogens, but also cells were agitated at 250 rpm (0.9 g L1) as compared to
gas substrate availability because gas solubility increases with 150 rpm (1.8 g L1). Similarly, the maximum acetic acid concentra-
decreasing temperature (Munasinghe and Khanal, 2010; tions were 52% lower when cells were agitated at 250 rpm (2 g L1)
Kundiyana et al., 2011). Acetic acid and ethanol was desirable than 150 rpm (4.2 g L1).
product from syngas fermentation (Fig. 2b). As shown in Fig. 2b
the maximum amount of cell biomass (0.17 g L1), ethanol 3.2.4. Effect of syngas pressure on cell growth and ethanol production
(1.48 g L1) and acetic acid (0.6 g L1) produced by mixed culture One potential bottleneck of syngas bioconversion is the mass
TERI SA1 was at 37 °C. High temperature did not favor syngas uti- transfer limitation due to the sparingly soluble nature of the sub-
lization, as there was no cell growth, ethanol and acetic acid pro- strate. Hence, one way to overcome this limitation is by increasing
duction at 50 °C, confirming the mesophilic nature of selected the pressure. In batch fermentation, different CO pressures mean
mixed culture TERI SA1. In case of mesophilic syngas fermenting different gaseous substrate concentrations which are directly pro-
bacterial strains, higher temperature above 37 °C affects the portional to the product formation and cell growth (Abubackar
growth, as carbon monoxide (CO) and hydrogen (H2) components et al., 2012). As the initial syngas pressure in the reactor bottle
of syngas have a decreased solubility with increase in temperature headspace was increased there was a linear increase in cell growth,
(Munasinghe and Khanal, 2010). There was very less difference ethanol and acetic acid production up to 2 kg cm2 (Fig. 2d). How-
between the cell growth as well as product formation efficiency ever, as the pressure was increased beyond 2 kg cm2 (up to 2.5
of the mixed culture at temperature 30 °C and 37 °C (Fig. 2b). As and 3.0 kg cm2) both ethanol and acetic acid production gradually
the optimum temperature for mesophilic acetogens are between decreased (Fig. 2d). This can be interpreted from this study that
30 °C and 40 °C, while thermophilic acetogens grow best between high syngas pressure (above 2 kg cm2) in the reactor bottle head-
55 °C and 60 °C (Daniell et al., 2012). These results are efficient space probably inhibited CO consumption, thus inhibiting the
than the previous findings of (Abrini et al., 1994) for C. autoethano- growth and product formation efficiency of the selected mixed cul-
genum which produced up to 0.32 g L1 ethanol at similar temper- ture TERI SA1. Maximum amount of ethanol and acetic acid pro-
ature range. duction achieved by the mixed culture TERI SA1 at pressure
2 kg cm2 was 1.54 g L1 and 0.9 g L1 respectively which was
3.2.3. Effect of agitation speed on cell growth and ethanol production comparatively higher than achieved at 0.5 kg cm2 (Fig. 2d). In a
Agitation plays a very important role in anaerobic fermenta- similar study by (Hurst and Lewis, 2010) increase of CO partial
tions utilizing gaseous substrates because gases such as CO, CO2 pressure from 0.35 to 2.0 atm. increased the cell concentration
and H2 are less soluble in water and mass transfer of these sub- from 0.20 to 1.08 g L1, which represented a 440% improvement.
strates to the cells are an important factor in reactor performance Klasson et al. (1991), examined that there was a linear relation
(Kapic et al., 2006). The maximum amount of ethanol (1.47 g L1) between the reaction rate and CO partial pressures up to 1.6 atm.
and acetic acid (0.85 g L1) production achieved by selected mixed At CO partial pressure of 2.5 atm. there was a short period of CO
culture TERI SA1 was at agitation of 100 rpm (Fig. 2c). Cell growth uptake, thereafter the culture failed to utilize the CO. This was
A. Singla et al. / Bioresource Technology 172 (2014) 41–49 45

pH optimization
(a)

Cell biomass (g L-1)


1.8 0.16

AA & EtOH. (g L-1)


1.5
0.12
1.2
0.9 0.08
0.6
0.04
0.3
0 0
5 6 7 8
Initial media pH
Ethanol Acec acid Cell biomass

Temperature optimization
(b)
1.8 0.2

L-1)
AA & EtOH. ( g L-1)

1.5 0.16

Cell biomass (g
1.2
0.12
0.9
0.08
0.6
0.3 0.04
0 0
30 37 50
Incubation temperature (°C)
Ethanol Acec acid Cell biomass

(c) 1.8
Agitation speed optimization
0.2

Cell biomass (g L-1)


AA & EtOH. (g L-1)

1.5 0.16
1.2
0.12
0.9
0.08
0.6
0.3 0.04

0 0
100 150 200
Agitation speed (rpm)
Ethanol Acec acid Cell biomass

Syngas pressure optimization


(d)
1.8 0.2
Cell biomass (g L-1)

1.5 0.16
AA & EtOH. (g L-1)

1.2
0.12
0.9
0.08
0.6
0.3 0.04
0 0
0.5 1 1.5 2 2.5 3
Syngas Pressure (kg cm-2)
Ethanol Acec acid Cell biomass
Fig. 2. (a) Optimization of initial media pH, (b) Incubation temperature, (c) Agitation speed and (d) Initial syngas pressure in the reactor headspace for cell biomass, acetic
acid and ethanol production by syngas fermenting mixed culture TERI SA1. The experiment was performed in duplicate and the data points are average of the
duplicate ± standard deviation (less than 5% of average).
46 A. Singla et al. / Bioresource Technology 172 (2014) 41–49

probably due to the insufficient cell concentration which could not and gradually increased till the cells entered the stationary phase.
keep the reaction at the mass transfer limited stage and caused A maximum acetic acid concentration of 1.67 g L1 was reached at
toxicity for the microbe. It was suggested that high CO partial 192 h (Fig. 3b). Acetic acid concentration then decreased to
pressure could be employed after a sufficient cell concentration 1.3 g L1 at the end of the fermentation. Acetic acid concentrations
was achieved (Klasson et al., 1991). Younesi et al. (2005) also were comparatively lower between 120 and 168 h. Consumption of
studied the effect of various initial total pressures of syngas acetic acid by mixed culture TERI SA1 was observed at later stages
(0.8–1.8 atm.) in the batch culture of C. ljungdahlii. of fermentation. Acetic acid consumption was associated with pro-
duction of ethanol after 192 h of fermentation. This was also
observed in other syngas fermentation studies using Clostridium
3.3. Semi-continuous fermentation studies for enhanced ethanol
strain P11 (Panneerselvam et al., 2010) and C. carboxidivorans P7
production by the selected mixed culture TERI SA1
(Hurst and Lewis, 2010). Ethanol production started in the expo-
nential phase (Fig. 3b). The rate of ethanol production increased
3.3.1. Cell growth and pH profiles
in the early exponential phase and then slightly decreased till
Cell mass concentrations of mixed culture TERI SA1 increased
the cell entered in stationary phase. During stationary phase the
during the first 144 h (Fig. 3a). However, cells were in the lag phase
ethanol concentration again start increasing and decreased with
in the first 24 h. After 24 h cells entered the exponential phase and
there was increase in cell mass concentration until 144 h. In addi- the decline in cell concentration. This trend in ethanol production
was due to the effect of syngas pressure in reactor headspace
tion, cells entered the stationary phase after 144 h. The maximum
cell mass concentration of 0.2 g L1 was observed at 144 h. Minor (Hurst and Lewis, 2010). The maximum ethanol concentration
was 2.3 g L1 at 72 h, which was 49% higher than achieved during
changes in cell mass concentrations were observed during the sta-
tionary phase until 216 h. The pH of the culture media decreased batch fermentation conditions (Fig. 3b). Higher ethanol production
in semi-continuous fermentation mode could possibly be due to
during cell growth due to the production of acetic acid (Fig. 3a).
The pH decreased from 6.0 to 4.6 until 192 h, at which time the the sparging of fresh syngas after every 24 h as compared to batch
fermentation (sparged and pressurized with syngas just in the
acetic acid concentrations in medium reached a maximum
(Fig. 3a). Subsequently, a minor change in pH was observed after beginning of fermentation).
192 h.
3.4. Scale up studies using mixed culture TERI SA1 for increased
3.3.2. Product profiles ethanol production
Acetic acid and ethanol were the major products formed during
syngas fermentation in culture media (Fig. 3b). Acetic acid is a pri- 3.4.1. Cell growth and pH profiles
mary metabolite, and its production is associated with cell growth. The growth and pH profiles of mixed culture TERI SA1 is shown
The production of acetic acid was higher during the growth phase in Fig. 4a. There was a short lag Phase (24 h) followed by an

(a)
6.5 0.25
Cell biomass (g L-1)
6 0.2

5.5 0.15
pH

5 0.1

4.5 0.05

4 0
24 48 72 96 120 144 168 192 216
Time (h)
pH Cell biomass

(b) 3 1.8
L-1)

2.5 1.5
Ethanol (g L-1)

Acetic acid (g

2 1.2
1.5 0.9
1 0.6
0.5 0.3
0 0
24 48 72 96 120 144 168 192 216
Time (h)
EtOH. AA
Fig. 3. (a) Cell growth and pH profiles and (b) ethanol and acetic acid production at different time intervals by mixed culture TERI SA1 under semi-continuous fermentation
conditions. The experiment was performed in duplicate and the data points are average of the duplicate ± standard deviation (less than 5% of average).
A. Singla et al. / Bioresource Technology 172 (2014) 41–49 47

(a)
6.5 0.2

Cell biomass (g L-1)


6 0.16

5.5 0.12

pH 5 0.08

4.5 0.04

4 0
24 48 72 96 120 144 168 192 216
Time (h)
pH cell biomass

EtOH. & AA Production (g L-1)


(b)
12 2.5
CO & H2 utilization (mM)

10 2
8
1.5
6
1
4
2 0.5
0 0
0
24

48

72

96

120

144

168

192

216
24, rs

48, rs

72, rs

96, rs

120, rs

144, rs

168, rs

192, rs
Time (h)
EtOH. AA CO H₂
Fig. 4. (a) Change in pH and cell biomass concentration, (b) Ethanol and acetic acid production and CO and H2 consumption at different time intervals by mixed culture TERI
SA1 under semi-continuous fermentation conditions in 1 L reactor bottles. The experiment was performed in duplicate and the data points are average of the
duplicate ± standard deviation (less than 5% of average).

exponential phase. The cell concentration increased exponentially because 1 L reactor bottle could not tolerate high syngas pressure
during the first 72 h of fermentation. The maximum cell biomass of 2 kg cm2. It was reported that ethanol production by C. carbox-
concentration of 0.17 g L1 was achieved at 144 h. The pH of idivorans P7 was non-growth related when the partial pressure of
culture media decreased from 6 to 5.1 during cell growth in the CO in the headspace was below 106 kPa (Hurst and Lewis, 2010).
first 72 h due to the production of acetic acid (Fig. 4a). However, However, ethanol formation was growth associated with C. carbox-
the pH of the medium decreased from 6 to 4.7 after 72 h, during idivorans P7 when the partial pressure of CO in the headspace was
which cells were still growing. Then, the pH of culture media above 106 kPa. In the present study, the rate of ethanol production
slightly decreased after 120 h due to stationary phase of bacterial was high in the exponential phase and decreased with time espe-
cells (Fig. 4a). cially with the increase in cell biomass concentration. After 216 h
of fermentation, ethanol in 1 L reactor bottle was 2.0 g L1
3.4.2. Gas consumption and product profiles (Fig. 4b). Both CO and H2 were utilized by mixed culture TERI
Acetic acid and ethanol production with mixed culture TERI SA1 SA1 for growth, acetic acid and ethanol production. The total moles
was observed during the acidogenic and solventogenic phases of CO and H2 consumed by mixed culture TERI SA1 during syngas
respectively. Acetic acid concentration increased in the first 72 h fermentation is shown in Fig. 4b. About 28% and 3% of the total
and further increased till 192 h (Fig. 4b). Acetic acid formation by CO and H2 moles were consumed by mixed culture TERI SA1 during
mixed culture TERI SA1 was growth related, which resulted a the first 144 h of fermentation for mostly growth and acetic acid
decrease in the pH of the fermentation medium (Fig. 4b). The max- production respectively. In the acetyl-CoA pathway, both H2 and
imum acetic acid concentration was 1.9 g L1 at 192 h. After the CO serve as an energy and electron source for cell growth and
concentration of acetic acid reached a maximum, mixed culture product formation (Ragsdale, 2004). Significant amounts of carbon
TERI SA1 started to consume it. About 8.4% of the acetic acid was from CO can be converted to cell biomass and ethanol if H2 is uti-
consumed between 192 and 216 h (Fig. 4b). The ethanol produc- lized as an electron source. In contrast, high concentration of CO
tion was observed during exponential phase which decreased after was found to inhibit hydrogenase enzyme and thus reduce the
the stationary phase. The slight increase in ethanol production was uptake of H2 in non-CO fermenting organisms (Bennett et al.,
observed during stationary phase when the pH was around 4.7 2000). This could explain the decrease in H2 consumption during
after 120 h (Fig. 4a and b). Ethanol formation by mixed culture TERI the stationary phase with mixed culture TERI SA1 (Fig. 4b). Trends
SA1 in 1L reactor bottle fermentation was growth as well as non- in CO consumption profile were similar to ethanol profiles (Fig. 4b).
growth associated depending on the actual syngas pressure in Ethanol production started after 24 h of the fermentation. For eth-
the reactor bottle headspace. The syngas pressure during 1 L reac- anol production and cell maintenance between 24 and 216 h,
tor bottles fermentation study was maintained at 1.0 kg cm2 approximately 34% and 2% of the total moles of CO were consumed
48 A. Singla et al. / Bioresource Technology 172 (2014) 41–49

Table 1
Comparison of ethanol yield by different mesophilic syngas fermenting bacterial strains.

Microbial species Topt (°C) pHopt Yield (g/l) EtOH. Products References
Clostridium ljungdahlii 37 N.A 0.6 Acetate, ethanol Younesi et al. (2005)
Clostridium autoethanogenum 37 5.8–6.0 0.32 Acetate, ethanol Abrini et al. (1994)
Clostridium ljungdahlii 37 4.0–5.0 1.0 Acetate, ethanol Gaddy and Clausen (1992)
Clostridium carboxidivorans P7T 37 5.8–5.9 0.56 Acetate, ethanol, butanol Rajagopalan et al. (2002)
Clostridium ragsdalei 32 6.0 1.89 Acetate, ethanol Kundiyana et al. (2011)
Butyribacterium methylotropphicum 37 6.0 N.A Acetate, ethanol, butanol Shen et al. (1999)
Eubacterium limosum 38–39 7.0–7.2 N.A Acetate, ethanol Chang et al. (2001)
Peptostreptococcus productus 37 7.0 N.A Acetate, ethanol Lorowitz and Bryant (1984)
A. bacchi 37 5.8–7.0 1.7 Acetate, ethanol Liu et al. (2012)
Mixed culture TERI SA1 37 6.0 2.3 Acetate, ethanol This study
Mixed culture 37 8.0 8.0 Acetate, ethanol Liu et al. (2013)

by mixed culture TERI SA1 respectively. This shows that most of Devi, M.P., Mohan, S.V., Mohanakrishna, G., Sarma, P.N., 2010. Regulatory influence
of CO2 supplementation on fermentative hydrogen production process. Int. J.
the CO consumed by mixed culture TERI SA1 was utilized for eth-
Hydrogen Energy 35, 10701–10709.
anol production. The detailed comparison for ethanol yield of Fan, Y.T., Liao, X.Ch., Lu, H.J., Hou, H.W., 2003. Study on biohydrogen production by
mixed culture TERI SA1 with different syngas fermenting strains anaerobic biological fermentation of organic wastes. Environ. Sci. 24, 152–155.
reported from mesophilic origin is also provided in Table 1. Gaddy, J.L., Clausen, E.C., 1992. Clostridium ljungdahlii, an anaerobic ethanol and
acetate producing microorganism. U.S. Patent 5173429.
Gaddy, J.L., Arora, D.K., Ko, C.W., Phillips, J.R., Basu, R., Wikstrom, C.V., Clausen, E.C.,
2003. Methods for increasing the production of ethanol from microbial
4. Conclusion fermentation. U.S. Patent 0211585 A1.
Griffin, D.W., Schultz, M.A., 2012. Fuel and chemical products from biomass syngas:
a comparison of gas fermentation to thermochemical conversion routes.
Enriched anaerobic mixed culture TERI SA1 was capable of uti- Environ. Prog. Sustain. Energy 31, 219–224.
lizing syngas as carbon source for ethanol production. Optimiza- Hurst, K.M., Lewis, R.S., 2010. Carbon monoxide partial pressure effects on the
metabolic process of syngas fermentation. Biochem. Eng. J. 48, 159–165.
tion of operational parameters enhanced ethanol production
Kapic, A., Jones, S., Heindel, T., 2006. Carbon monoxide mass transfer in a syngas
from 0.15 g L1 to 1.54 g L1 indicating 927% improvement. Subse- mixture. Ind. Eng. Chem. Res. 45, 9150–9155.
quent study of semi-continuous and up-scaling parameters further Kaur, G., Mandal, A.K., Nihlani, M.C., Lal, B., 2009. Control of sulfidogenic bacteria in
produced water from the Kathloni oilfield in northeast India. Int. Biodeterior.
enhanced ethanol production up to 2.3 g L1 which was consider-
Biodegradation 63, 151–155.
ably higher than 1.89 g L1 with Clostridium ragsdalei (Kundiyana Klasson, K.T., Ackerson, M.D., Clausen, E.C., Gaddy, J.L., 1991. Bioreactor design for
et al., 2011) but lower than 8 g L1 with reported mixed culture synthesis fermentations. Fuel 70, 605–614.
(Liu et al., 2013), therefore higher volume bio-reactor studies with Klasson, K.T., Ackerson, M.D., Clausen, E.C., 1992. Bioconversion of syngas into liquid
or gaseous fuels. Enzyme Microb. Technol. 14, 602–608.
selected mixed culture TERI SA1 needs to be undertaken to provide Kundiyana, D.K., Huhnke, R.L., Wilkins, M.R., 2010. Syngas fermentation in a 100-L
further insights. pilot scale fermentor: design and process considerations. J. Biosci. Bioeng. 109,
492–498.
Kundiyana, D.K., Wilkins, M.R., Maddipati, P.B., Huhnke, R.L., 2011. Effect of
temperature, pH and buffer on syngas fermentation using Clostridium strain P11.
Acknowledgements Bioresour. Technol. 102, 5794–5799.
Liu, K., Atiyeh, H.K., Tanner, R.S., Wilkins, M.R., Huhnke, R.L., 2012. Fermentative
The financial support for this research was provided by Depart- production of ethanol from syngas using novel moderately alkaliphilic strains of
Alkalibaculum bacchi. Bioresour. Technol. 104, 336–341.
ment of Biotechnology, Ministry of Science and Technology, India. Liu, K., Atiyeh, H.K., Stevenson, B.S., Tanner, R.S., Wilkins, M.R., Huhnke, R.L., 2013.
The authors are thankful to Dr. R.K. Pachauri Director-General, Mixed culture syngas fermentation and conversion of carboxylic acids into
TERI, for providing us excellent infrastructure support for the alcohols. Bioresour. Technol. 152, 337–346.
Lorowitz, W.H., Bryant, M.P., 1984. Peptostreptococcus productus strain that grows
study. The authors also thank Mr. Samarjeet (TERI, India) and Mr.
rapidly with CO as the energy source. Appl. Environ. Microbiol. 47, 961–964.
Shivaji (TERI, India) for their technical assistance and Ms. Sneha Maddipati, P., Atiyeh, H.K., Bellmer, D.D., Huhnke, R.L., 2011. Ethanol production
Singh (TERI University, India), Ms. Mohita Sharma (TERI University, from syngas by Clostridium strain P11 using corn steep liquor as a nutrient
India) and Ms. Anchal priya (TERI University, India) for helpful replacement to yeast extract. Bioresour. Technol. 102, 6494–6501.
Munasinghe, P.C., Khanal, S.K., 2010. Biomass-derived syngas fermentation into
discussions. biofuels: opportunities and challenges. Bioresour. Technol. 101, 5013–5022.
Panneerselvam, A., Wilkins, M.R., DeLorme, M.J.M., Atiyeh, H.K., Huhnke, R.L., 2010.
Effects of various reducing agents on syngas fermentation by Clostridium
References ragsdalei. Biol. Eng. 2, 135–144.
Phillips, J., Klasson, K., Clausen, E., Gaddy, J., 1993. Biological production of ethanol
from coal synthesis gas. Appl. Biochem. Biotechnol. 39–40, 559–571.
Abrini, J., Naceau, H., Nyns, E.J., 1994. Clostridium autoethanogenum, sp. nov., an
Ragsdale, S.W., 2004. Life with carbon monoxide. Crit. Rev. Biochem. Mol. Biol. 39,
anaerobic bacterium that produced ethanol from carbon monoxide. Arch.
165–195.
Microbiol. 161, 345–351.
Rajagopalan, S., Datar, P., Lewis, R.S., 2002. Formation of ethanol from carbon
Abubackar, H.N., Veiga, M.C., Kennes, C., 2012. Biological conversion of carbon
monoxide via a new microbial catalyst. Biomass Bioenergy 23, 487–493.
monoxide to ethanol: effect of pH, gas pressure, reducing agent and yeast
Ramachandriya, K.D., 2006. Effect of biomass generated producer gas, methane and
extract. Bioresour. Technol. 114, 518–522.
physical parameters on producer gas fermentations by Clostridium strain p11
Agrawal, A., Vanbroekhoven, Karolien, Lal, Banwari, 2010. Diversity of culturable
(Master’s thesis). OkState. 140–151.
sulfidogenic bacteria in two oil–water separation tanks in the north-eastern oil
Sarma, P.M., Bhattacharya, D., Krishnan, S., Lal, B., 2004. Degradation of polycyclic
fields of India. Anaerobe 16, 12–18.
aromatic hydrocarbon by a newly discovered enteric bacterium, Leclercia
Bennett, B., Lemon, B.J., Peters, J.W., 2000. Reversible carbon monoxide binding and
adecarboxylata. Appl. Environ. Microbiol. 70, 3163–3166.
inhibition at the active site of the Fe-only hydrogenase. Biochemistry 39, 7455–
Shen, G.J., Shieh, J.S., Grethlein, A.J., Jain, M.K., Zeikus, J.G., 1999. Biochemical basis
7460.
for carbon monoxide tolerance and butanol production by Butyribacterium
Chang, I.S., Kim, B.H., Lovitt, R.W., Bang, J.S., 2001. Effect of CO partial pressure on
methylotrophicum. Appl. Microbiol. Biotechnol. 51, 827–832.
cell-recycled continuous CO fermentations by Eubacterium limosum KIST612.
Singh, S., Sarma, P.M., Lal, B., 2014. Biohydrogen production by
Process Biochem. 37, 411–421.
Thermoanaerobacterium thermosaccharolyticum TERI S7 from oil reservoir flow
Daniell, J. et al., 2012. Commercial biomass syngas fermentation. Energies 5, 5372–
pipeline. Int. J. Hydrogen Energy 39, 4206–4214.
5417.
A. Singla et al. / Bioresource Technology 172 (2014) 41–49 49

Tanner, R.S., 2005. Clostridium carboxidivorans sp. nov., a solvent-producing Wolfrum, E.J., Watt, A.S., 2002. Bioreactor design studies for a hydrogen-producing
Clostridium isolated from an agricultural settling lagoon, and reclassification bacterium. Appl. Biochem. Biotechnol. 98, 611–625.
of the acetogen Clostridium scatologenes strain SL1 as Clostridium drakei sp. nov. Younesi, H., Najafpour, G., Mohamed, A.R., 2005. Ethanol and acetate production
Int. J. Syst. Evol. Microbiol. 55, 2085–2091. from synthesis gas via fermentation processes using anaerobic bacterium
Vannby, R., Winter Madsen, S.E.L., 1992. New developments in synthesis gas Clostridium ljungdahlii. Biochem. Eng. J. 27, 110–119.
production. Hydrocarbon Technol. Int. 105.

Você também pode gostar