You are on page 1of 7

Bioresource Technology 113 (2012) 30–36

Contents lists available at SciVerse ScienceDirect

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

Fermentative biohydrogen production from lactate and acetate


Chao-Wei Wu a, Liang-Ming Whang a,b,c,⇑, Hai-Hsuan Cheng a, Kan-Chi Chan a
a
Department of Environmental Engineering, National Cheng Kung University, No. 1, University Road, Tainan 701, Taiwan
b
Sustainable Environment Research Center (SERC), National Cheng Kung University, No. 1, University Road, Tainan 701, Taiwan
c
Research Center for Energy Technology and Strategy (RCETS), National Cheng Kung University, No. 1, University Road, Tainan 701, Taiwan

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

Article history: In this study, a continuous-flow stirred tank reactor (CSTR) fed with lactate and acetate was operated to
Received 15 October 2011 enrich hydrogen-producing bacteria. By varying the influent substrate concentrations and hydraulic
Received in revised form 25 December 2011 retention times (HRT), the volumetric loading rate (VLR) of 55.64 kg-COD/m3/day seemed to be optimum
Accepted 26 December 2011
for this enriched culture for fermentative hydrogen production from lactate and acetate. The results of
Available online 5 January 2012
batch experiments confirmed that the enriched culture tended to fulfill the e equiv requirement for cell
growth at a lower VLR condition (21.77 kg-COD/m3/day), while it could largely distribute the e equiv for
Keywords:
hydrogen production at a higher VLR condition. However, a maximum lactate/acetate concentration
Bioenergy
Hydrogen production
allowed for enriching this culture existed, especially at a lower HRT condition in which wash-out can
Lactate be an issue for this enriched culture. Finally, the results of cloning and sequencing indicated that Clostrid-
Acetate ium tyrobutyricum was considered the major hydrogen-producing bacteria in the CSTR fed with lactate
Clostridium tyrobutyricum and acetate.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction (Lay et al., 1999), spoiled wheat grains (Kalia et al., 1993), municipal
wastewater and sludge (Kim et al., 2004; Van Ginkel et al., 2005),
The importance of renewable energy sources increases as the is- potato wastes (Zhu et al., 2008), and food wastes (Han and Shin,
sues of fossil fuel exhaustion and global warming become serious. 2004; Kim et al., 2004; Chu et al., 2008; Lee et al., 2008b), but results
Considering our dependence on fossil fuel energy and the global suggest that hydrogen production is more efficient from carbohy-
environment issues, there is an urgent need in developing a clean drates than other materials (Li and Fang, 2007). In fact, fermentative
and renewable energy such as hydrogen. Hydrogen can be pro- hydrogen production from simple sugars, such as sucrose and glu-
duced biologically from potential renewable materials such as car- cose, can be easily achieved at high conversion efficiencies and its
bohydrate-containing biomass and organic wastes (Das and metabolism has been widely studied (Li and Fang, 2007; Lin et al.,
Veziroglu, 2001; Lee et al., 2010). Among these H2-generating pro- 2007). In addition to carbohydrates, fermentative biohydrogen pro-
cesses, the dark fermentation process is considered more favorable duction from lactate/acetate-containing wastes has been discussed
because it offers a potential means to produce H2 without requiring recently (Jo et al., 2008a,b; Lee et al., 2008b; Juang et al., 2011), but
a significant energy input while partially stabilizing a variety of detailed information is rather limited. Although utilization of lac-
complex organic wastes (Davila-Vazquez et al., 2008; Lee et al., tate and acetate for production of butyrate and hydrogen by a num-
2010). ber of microorganisms including Clostridium acetobutylicum,
Hydrogen production from anaerobic waste treatment benefits Clostridium tyrobutyricum, Clostridium beijerinckii, and Butyribacte-
both organic wastes reduction and renewable energy production rium methylotrophicum has been previously reported (Balows
at the same time, but it also creates challenges because the waste et al., 1992; Diez-gonzalez et al., 1995; Hashsham et al., 2000; Jo
materials usually are composed of a variety of substrates that can et al., 2008a,b), this occurrence, however, has been reported only
be used by different species of microorganisms (Whang et al., for batch experiments.
2006; Li and Fang, 2007; Li et al., 2010). Fermentative biohydrogen In this study, we operated a continuous-flow stirred tank reac-
has been studied for the organic fraction of municipal solid wastes tor (CSTR) fed with lactate and acetate in order to enrich lactate/
acetate-utilizing and hydrogen-producing consortia and evaluate
their hydrogen production capacity under different volumetric
loading rate (VLR) conditions. In addition to CSTR operation, batch
⇑ Corresponding author at: Research Center for Energy Technology and Strategy
experiments were conducted to evaluate effects of pH and initial
(RCETS), National Cheng Kung University, No. 1, University Road, Tainan 701,
Taiwan. Tel.: +886 6 2757575x65837; fax: +886 6 2752790. substrate-to-biomass (S0/X0) ratio on fermentative biohydrogen
E-mail address: whang@mail.ncku.edu.tw (L.-M. Whang). production of the enriched culture. Finally, molecular methods

0960-8524/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2011.12.130
C.-W. Wu et al. / Bioresource Technology 113 (2012) 30–36 31

were applied to investigate microbial ecology of lactate/acetate- equipped with an ICSep COREGEL-87H3 column (7.8  300 mm)
utilizing and hydrogen-producing bacteria enriched in the CSTR. in a 42 °C oven and a RI Range 16 detector. The eluent usedwas
0.008 N sulfuric acid at a flow rate of 0.6 mL/min. The carbohydrate
was analyzed using the phenol–sulfuric acid method (Herbert
2. Methods
et al., 1971). The pH, ORP, NH4+–N and volatile suspended solids
(VSS) were measured according to standard methods (APHA,
2.1. Operation of the hydrogen fermentation bioreactor
1998).
A continuous stirred tank reactor-type (CSTR) bioreactor was
operated in this study in order to enrich lactate/acetate-utilizing 2.4. Calculations of electron-equivalent balance
and hydrogen-producing consortia. The total volume of the biore-
actor was 4 L with a working volume of 2 L and bioreactor was We used Eq. (1) proposed in reference (Lee and Rittmann, 2009;
equipped with a mechanic propeller for completely mixing at an Lee et al., 2008a) for establishing the electron-equivalent (e equiv)
agitating speed of 160 rpm. The influent medium solution, contain- balance in all experiments.
ing predetermined concentrations of lactate and acetate, was eHLa;HAc ¼ eSEP þ eH2 þ ebiomass þ eundet ð1Þ
stored at 4 °C in a refrigerator and continuously fed into the biore-
actors using a peristaltic pump. Each liter of the influent feed was
 
supplemented with the following chemicals (in mg): resazurin, Where e 
HLa;HAc = e equiv of initial substrate, eSEP = e equiv of all

0.175; CaCl26H2O, 32.32; MgCl26H2O, 232.26; KCl, 167.81; soluble end products (SEP), e H2 = e equiv of cumulative hydrogen

MnCl24H2O, 63.87; CoCl26H2O, 3.87; H3BO3, 0.74; CuCl2H2O, gas during the incubation period, e biomass = e equiv of growth bio-

0.35; Na2MoO42H2O, 0.33; ZnCl2, 0.27; FeCl24H2O, 10.62; sodium mass during batch experiment, and e undet = e equiv of not detected
thioglycolate, 217.35; KH2PO4, 119 (Liu et al., 2011). The pH in both in the measured sinks. All values were experimentally measured
bioreactors was controlled at 6 ± 0.1 using a pH controller by the except eundet . The chemical formula for biomass was assumed to
addition of 5% H3PO4and 10% NaOH throughout the experiments. be C5H7O2N (Rittmann and McCarty, 2001). SEP includes formate,
The seeding microorganism was obtained from a lab-scale contin- acetate, lactate, butyrate. The conversion of e equiv values fol-
uous flow bioreactor fed with wasted residues from a bioethanol lows: 1 mol acetate 8 e equiv, 1 mol butyrate 20 e equiv, 1 mol
fermentation process (Juang et al., 2011). In each run, the data lactate 12 e equiv, 1 mol H2 2 e equiv, and 1 mol biomass
were collected from the bioreactors after operation of a period (C5H7O2N) 20 e equiv(Lee et al., 2008a; Rittmann and McCarty,
passing through more than 10 reactor volumes. The amount of 2001).
gases produced in the bioreactor was measured with a wet-gas
flow meter (Shinagawa W-NK-0.5B, Tokyo, Japan). The bioreactors 2.5. DNA extraction, polymerase chain reaction (PCR),cloning,
and the off-gas flow meter were kept in an air-bath incubator and sequencing and phylogenetic analysis of [Fe–Fe] hydrogenase gene
the temperature was maintained at 35 °C.
The genomic DNA samples in biomass taken from the CSTR were
2.2. Fermentative biohydrogen tests extracted using the UltracleanTM Soil DNA Isolation Kit (MoBio Lab-
oratories, Inc., Solana Beach, CA). After DNA extraction, a forward
Fermentative biohydrogen batch tests conducted in this study primer HydH1f (50 -TTIACITSITGYWSYCCIGSHTGG-30 ) and a reverse
were a modified version of biochemical methane potential test primer HydH3r (50 -CAICCIYMIGGRCAISNCAT-30 ) (Schmidt et al.,
originally developed by Owen et al. (1979). The test was carried 2010) was used for amplifying an approximate 500-bp fragment
out in a series of 1 L glass bottles equipped with pH/ORP monitor- of [Fe–Fe]-hydrogenase functional gene for the clone library analy-
ing and gas collection systems. To each bottle with a liquid work- sis. The thermal profile used for the amplification was as follows: a
ing volume of 800 mL, predetermined concentrations of lactate/ hotstart at 95 °C for 10 min, 30 cycles of denaturation (45 s at
acetate and biomass taken from the CSTR were added as fermenta- 94 °C), annealing (45 s at 55 °C) and extension (90 s at 72 °C), and
tion substrate and seeding sludge, respectively. In addition, each a final extension at 72 °C for 3 min. The PCR products of HydH1f-
bottle contained many chemicals as growth nutrients (Liu et al., HydH3r were purified by electrophoresis in a 1.5% (wt/vol) agarose
2011). The bottles were incubated at 35 °C in a water bath tank gel, and used in the construction of the clone library. The purified
equipped with magnetic stirring at a rotational rate of 120 rpm. PCR products were ligated into the pGEM-T Easy vector (Promega,
The pH of the mixed liquor was controlled by feeding with either Madison, WI) and the ligation products were used to transform
NaOH (10%) or H3PO4 (5%). The total volume of gases produced Escherichia coli DH5a competent cells following the manufacture’s
during fermentation was determined using a gas collection system protocol (Invitrogen, Groningen, Netherlands). Plasmids of clones
with water displacement, and the gas composition was analyzed were extracted by Mini-M plasmid DNA extraction system (Vio-
using gas chromatography (GC). Samples were frequently taken gene, Sunnyvale, CA, USA). DNA sequencing reactions were per-
throughout the batch experiments for the determination of organic formed using ABI 3100 and 3730 capillary sequencers (Applied
acids using high-performance liquid chromatography (HPLC). Biosystems, Foster City, CA, USA). BioEdit was used to align cloned
and published sequences in GenBank using the Basic Local Align-
2.3. Analytical methods ment Search Tool program developed by US National Center for Bio-
technology Information (Altschul et al., 1990).
The composition of biogas collected from bioreactors and
batches was analyzed using gas chromatograph (China GC 8900, 3. Results and discussion
Taipei, Taiwan) equipped with a thermal conductivity detector
(TCD). A 2 m stainless column was packed with Hayesep Q (60/ 3.1. Performance of the hydrogen fermentation bioreactor
80 mesh) and installed in a 60 °C oven. The operational tempera-
tures of the injection port, the oven, and the detector were all set The experimental results of the hydrogen fermentation bioreac-
at 60 °C. Nitrogen was used as the carrier gas at a flow rate of tor obtained under five different operational conditions by varying
15 mL/min. Volatile fatty acids (VFAs), alcohols, and sugars were the influent substrate concentrations and HRTs are summarized in
determined using HPLC (Hitachi D2000 system, Tokyo, Japan) Table 1. It was evident that the microorganisms enriched in the
32 C.-W. Wu et al. / Bioresource Technology 113 (2012) 30–36

Table 1
Summary of operation and performance of the fermentative biohydrogen bioreactor.

Operational parameter Unit Run 1 Run 2 Run 3 Run 4 Run 5


HRT h 23 18 18 12 12
Lactate/acetate conc. mg/L 15000/4500 15000/4500 30000/9000 30000/9000 15000/4500
VLR kg-COD/m3/day 21.77 27.82 55.64 83.46 41.73
Biogas production L-biogas/day 3.25 ± 0.60 5.94 ± 1.92 19.40 ± 0.50 27.90 ± 2.20 11.61 ± 1.17
H2 content % 20 ± 3 25 ± 10 40±1 40 ± 2 29 ± 2
Specific H2 production rate mmol/g-VSS/h 0.48 ± 0.11 1.59 ± 1.05 4.27 ± 0.20 4.72 ± 0.73 3.08 ± 0.74
Volumetric H2 production rate mmol/L/h 0.56 ± 0.56 1.40 ± 0.90 6.60 ± 0.36 9.48 ± 0.93 2.86 ± 0.47
H2 yield mmol/mmol-HLa 0.09 ± 0.02 0.16 ± 0.02 0.42 ± 0.02 0.39 ± 0.04 0.21 ± 0.03

bioreactor were capable of producing hydrogen by utilizing lactate Run 4 ðVLR ¼ 83:46 kg  COD=m3 =dayÞ
and acetate. In addition, the hydrogen production performance of HLa þ 0:38 HAc þ 0:02 NH3 ! 0:08 C5 H7 O2 N0:45 H2
the bioreactor seemed to be a function of the calculated volumetric
þ 0:67 CO2 þ 0:68 HBu ðEq: 4Þ
loading rate (VLR). As the VLR increased from 21.77 to 83.46 kg-
COD/m3/day, the specific hydrogen production rate increased from
0.48 to 4.72 mmol/g-VSS/hr. The trend was similar for the hydro- Run 5 ðVLR ¼ 41:73 kg  COD=m3 =dayÞ
gen content and the hydrogen yield, but there was no increase of HLa þ 0:16 HAc þ 0:02 NH3 ! 0:05 C5 H7 O2 N þ 0:21 H2
these two parameters as the VLR increased from 55.64 to
þ 0:51 CO2 þ 0:66 HBu ðEq: 5Þ
83.46 kg-COD/m3/day. It is likely that the maximum VLR for the
microorganisms enriched in the bioreactor has been reached, since
an incomplete utilization of lactate was occurred at the VLR of where HLa is lactate, HAc is acetate, and HBu is butyrate. In
55.64 kg-COD/m3/day. view of these stoichiometric equations, the enriched culture fer-
Table 2 summarizes the yields (based on per mol of lactate) of mented lactate and acetate to produce biomass, butyrate, isobuty-
lactate/acetate fermentation end products in the bioreactor under rate, CO2 and H2, although the yield for hydrogen production
five different VLR conditions. In general, the enriched microorgan- varied with the VLR condition.
isms utilized lactate and acetate for biomass production, with The calculated fractions of e equiv distributions (in percent-
butyrate as the major SEP. Production of hydrogen and carbon age) for biomass and hydrogen at different VLR and food-to-bio-
dioxide that accompanies the formation of butyrate from lactate/ mass (F/M) conditions are presented in Fig. 1. As the VLR
acetate fermentation has been reported previously (Diez-gonzalez increased from 21.77 to 55.64 kg-COD/m3/day, the e equiv distri-
et al., 1995; Matsumoto and Nishimura, 2007). Our results, with a bution toward to hydrogen increased linearly, while the e equiv to
satisfactory recovery in carbon and electron balance, confirmed the biomass decreased correspondingly. Similar trends were also
proposed metabolism of lactate and acetate for hydrogen produc- found for biomass and hydrogen when the F/M ratio increased
tion that was previously discussed in several hydrogen fermenta- from 17.1 to 32.4 kg-COD/kg-VSS/day. It is hypothesized that at a
tion bioreactors (Jo et al., 2008a; Lee et al., 2008b; Juang et al., lower VLR or F/M condition with limited e equiv available the en-
2011). Furthermore, the stoichiometric equations obtained at dif- riched culture distributes a higher percentage of e equiv for cell
ferent VLR conditions were summarized as follows: synthesis, while more excess e equiv is available for hydrogen
production at a higher VLR or F/M condition in which the e equiv
Run 1 ðVLR ¼ 21:77 kg  COD=m3 =dayÞ
requirement for cell synthesis can be easily satisfied.
HLa þ 0:27 HAc þ 0:03 NH3 ! 0:07 C5 H7 O2 N þ 0:09 H2 Indeed, at the VLR of 55.64 kg-COD/m3/day or the F/M ratio of
þ 0:36 CO2 þ 0:65 HBu ðEq: 1Þ 32.4 kg-COD/kg-VSS/day in the Run 3, there was about 2500 mg/L
of residual lactate remained in the effluent, due presumably to that
the maximum VLR or the maximum F/M ratio for the enriched cul-
Run 2 ðVLR ¼ 27:82 kg  COD=m3 =dayÞ ture has been reached. Compared to the operational condition of
HLa þ 0:28 HAc þ 0:02 NH3 ! 0:06 C5 H7 O2 N þ 0:16 H2 the Run 3, similar influent concentrations of lactate (30000 mg/L)
and acetate (9000 mg/L) but a lower HRT (12 h) were applied in
þ 0:43 CO2 þ 0:69 HBu ðEq: 2Þ
the Run 4 with the VLR of 83.46 kg-COD/m3/day or the F/M ratio
of 36.2 kg-COD/kg-VSS/day, resulting in a higher e equiv demand
Run 3 ðVLR ¼ 55:64 kg  COD=m3 =dayÞ for cell synthesis. Although in the Run 4 about 3500 mg/L of resid-
HLa þ 0:42 HAc þ 0:02 NH3 ! 0:056 C5 H7 O2 N þ 0:47 H2 ual lactate still remained available for hydrogen production, no fur-
þ 0:7 CO2 þ 0:71 HBu ðEq: 3Þ ther increase of e equiv toward to hydrogen production was
observed as expected. This is likely that either the maximum VLR

Table 2
Summary of fermentation end product yields observed under different operational HRT conditions.

HRT Unit Run 1 Run 2 Run 3 Run 4 Run 5


h 23 18 18 12 12
Operation VLR kg-COD/m3/day 21.77 27.82 55.64 83.46 41.73
Substrate Lactate mol/mol 1.00 1.00 0.90 0.87 0.99
Consumption Acetate mol/mol 0.27 0.28 0.38 0.33 0.16
Gas Hydrogen mol/mol 0.09 ± 0.02 0.16 ± 0.10 0.42 ± 0.02 0.39 ± 0.04 0.21 ± 0.03
Carbon dioxide mol/mol 0.36 ± 0.07 0.43 ± 0.10 0.63 ± 0.01 0.58 ± 0.05 0.51 ± 0.04
Biomass Biomass mol/mol 0.07 ± 0.02 0.06 ± 0.01 0.05 ± 0.00 0.07 ± 0.01 0.05 ± 0.01
Organic acids Isobutyrate mol/mol 0.04 ± 0.00 0.03 ± 0.01 0.01 ± 0.01 0.02 ± 0.00 0.03±
Butyrate mol/mol 0.61 ± 0.02 0.66 ± 0.05 0.63 ± 0.02 0.57 ± 0.03 0.63±0.03
De equiv % –3.5 –6.9 –5.0 –6.1 –10.3
C recovery % 93.2% 96.0% 99.9% 99.1% 96.8%
C.-W. Wu et al. / Bioresource Technology 113 (2012) 30–36 33

60

Accumulated gas (mmol/L)


50
H2
40

30

20 CO2
10

0
0 5 10 15
Time (h)

120

Concentration (mmol/L)
100
Lactate
80
Butyrate
60

40
Acetate
20

0
0 5 10 15
Time (h)

Fig. 1. The calculated fractions of e equiv distributions (in percentage) for biomass 1400
and hydrogen at different VLR and food-to-biomass (F/M) conditions. h: Biomass,
j: hydrogen. 1200
Biomass (mg-VSS/L)

1000 Biomass
or the maximum F/M ratio for the enriched culture has been 800
reached, especially at such a lower HRT in which wash-out can be 600
an issue for this enriched culture, forcing this culture to distribute
more e equiv for growth. In comparison with the operational con- 400
dition of the Run 4, the same HRT but a lower influent concentra- 200
tions of lactate (15000 mg/L) and acetate (4500 mg/L) were
0
applied in the Run 5, resulting in a high F/M ratio of 42.9 kg-COD/ 0 5 10 15
kg-VSS/day and a drop in e equiv for hydrogen production. Under Time (h)
such a low HRT and insufficient substrate (lactate) condition, wash-
Fig. 2. Fermentative biohydrogen performance of the enriched culture at pH 6.5.
out is a serious issue for this enriched culture, forcing this culture to
distribute more e equiv for growth although the substrate (lactate)
was obviously insufficient. In summary, the operational condition bolic reaction described in this study (Hashsham et al., 2000; Jo
with a VLR of 55.64 kg-COD/m3/day or the F/M ratio of 32.4 kg- et al., 2008a).
COD/kg-VSS/day seemed to be optimum for this enriched culture Table 3 summarizes the H2 yield and specific H2 production rate
for fermentative hydrogen production from lactate and acetate. (SHPR) in fermentative biohydrogen tests conducted at different
pH conditions. In the batch1 experiment, the results indicated that
3.2. Effect of pH on fermentative biohydrogen among examined pH conditions (pH 5–6.5) pH 5.5 was favored for
achieving anoptimum hydrogen production potential with the
Batch experiments were conducted to evaluate fermentative highest H2 yieldof 0.54 mol-H2/mol-HLa and a second highest
biohydrogen production potential at different pH conditions. SHPR of 215 mL-H2/g-VSS/hr. In batches 2–4, the results showed
Fig. 2 shows the results of fermentative biohydrogen test for the en- that pH 6 was favored over pH 7 for attaining higher H2 yield
riched culture at pH 6.5, including the concentrations of lactate, and SHPR in the three initial substrate-to-biomass ratios examined
acetate, biomass, SEPs, and the cumulative gas production. The re- (S0/X0 = 5, 15, 50), with a higher initial biomass concentration the
sults confirmed the utilization of lactate and acetate by the en- more significant in the difference. The calculated e equiv distribu-
riched culture for biomass growth, formation of butyrate, and tions (in percentage) for hydrogen, biomass and butyrate obtained
production of hydrogen. Once the lactate was depleted, biomass in the batch 1 experiment are summarized in Fig. 3. As the pH in-
growth, butyrate formation, and gas production also ceased, despite creased from pH 5 to pH 6.5, the e equiv toward to biomass in-
that acetate remained available. Production of hydrogen that creased from 2.4% to 7.2%, due presumably to a preferred growth
accompanies the formation of butyrate from glucose fermentation condition at a neutral pH. As a result, a lower e equiv was distrib-
is a typical metabolic pathway by many clostridial species (Ljung- uted toward to hydrogen production (Liu et al., 2011). These find-
dahl et al., 1989). Only a few clostridial species such as C. acetobu- ings agreed with previous study that higher hydrogen production
tylicum and C. tyrobutyricum as well as other species like and SHPR were attained at pH 5.5 and 6 when using ethanol fer-
Butyribacterium methylotrophicum have been related to the meta- mentation residues as the substrate in which utilization of lactate
34 C.-W. Wu et al. / Bioresource Technology 113 (2012) 30–36

Table 3 Table 4
Hydrogen performancein fermentative biohydrogen tests conducted with enriched Fermentative biohydrogen performance under different X0 and S0/X0.
culture under different pH conditions.
X0 (mg- S0 (HLa/HAc) S0/ Yield (mol-H2/ SHPR(mL-H2/ Lag T
pH S0 (HLa/ X0 Yield SHPR* VSS/L) (mg-HLa/L) X0 mol-HLa) g-VSS/h) (h)
HAc)
300 7500/2250 25 0.51 153 3.3
mg/L mg-VSS/ mol-H2/mol- mL-H2/g-VSS/
300 15000/4500 50 0.53 188 4.89
L HLa h
300 30000/9000 100 0.49 128 9.7
Batch 1 5 10000/3000 600 0.48 142 1000 7500/2250 7.5 0.57 119 1.53
5.5 10000/3000 600 0.54 215 1000 15000/4500 15 0.59 151 2.91
6 10000/3000 600 0.42 207 1000 30000/9000 30 0.45 148 8.27
6.5 10000/3000 600 0.47 227 1700 7500/2250 4.5 0.76 181 0.3
Batch 2 6 15000/4500 300 0.53 188 1700 15000/4500 9 0.79 265 0.77
7 15000/4500 300 0.51 153 1700 30000/9000 18 0.58 235 1.96
Batch 3 6 15000/4500 1000 0.59 151
7 15000/4500 1000 0.45 129
Batch 4 6 15000/4500 1700 0.79 265
7 15000/4500 1700 0.61 169
tion at a higher VLR condition (55.64 kg-COD/m3/day). S0, on the
S0: initial substrate, X0: initial biomass. other hand, slightly affect the e equiv distributions of biomass
*
SHPR: Specific hydrogen production rate.
and hydrogen, but the S0 (HLa/HAc) at 30000/9000 mg/L did show
inhibition on biological activities in growth and hydrogen produc-
tion. This observation suggests that a maximum lactate/acetate
and acetate was found to be the mechanism for hydrogen produc-
concentration allowed for enriching this culture in a CSTR is ex-
tion (Juang et al., 2011).
isted, especially at a lower HRT condition in which wash-out can
be an issue for this enriched culture.
3.3. Fermentative biohydrogen tests at different S0/X0 ratio and X0

In Table 3, in addition to pH, the S0/X0 ratio seemed to affect the 3.4. Microbial community in continuous cultures fed with lactate and
hydrogen production potential as well. Therefore, fermentative acetate medium
biohydrogen tests were conducted in order to evaluate the effect
of initial biomass (X0) and substrate (S0) on hydrogen production Microbial community of the enriched culture fed with lactate
and the results are presented in Table 4. In general, increase in and acetate was evaluated based on the results of cloning and
H2 yield and SHPR but decrease in lag time were associated with sequencing for the [Fe–Fe] hydrogenase functional gene. The re-
an increase in X0. However, increase of S0 (HLa/HAc) from 7500/ sults of cloning and sequencing indicated that one dominant oper-
2250 to 15000/4500 slightly increased H2 yield and SHPR, but a ational taxonomic unit (OTU), representing 93% of the 130 clones
noticeable decrease in H2 yield occurred at the S0 (HLa/HAc) of retrieved, was identified to be closely related to C. tyrobutyricum
30000/9000. In association with the S0/X0 ratio, the trend was less JM1 (Jo et al., 2008b), with a similarity of 99%. Therefore, C. tyrobu-
clear but the hydrogen production performance generally de- tyricum was considered the major hydrogen-producing bacteria in
creased with an increased S0/X0 ratio. These observations agreed the bioreactor fed with lactate and acetate.
with previous findings that the increase of the S0/X0 ratio in batch C. tyrobutyricum has been widely reported for its capability of
experiments led to a decrease in the hydrogen production (Juang producing hydrogen from utilizing different sugars such as glucose
et al., 2011; Van Ginkel et al., 2001). and xylose (Ljungdahl et al., 1989; Balows et al., 1992; Jo et al.,
Fig. 4 presents the dependences of the calculated e equiv dis- 2008ab; Zhu and Yang, 2004; Lin et al., 2007; Liu et al., 2011;
tributions (in percentage) for hydrogen and biomass on S0 and Whang et al., 2011).Only a few reports have mentioned that C. tyro-
X0, and the corresponding fitted surfaces by regression. The en- butyricum is also capable of utilizing lactate as the major substrate
riched culture, at a higher X0 value, tended to distribute more e for growth but acetate is required for hydrogen production (Balows
equiv toward to hydrogen than biomass, while more e equiv et al., 1992; Diez-Gonzalez et al., 1995). In fact, it hasbeen shown
was distributed toward to biomass at a lower X0 value. These re- that the conversion of lactate and acetate to butyrate and hydrogen
sults supported the findings observed in the bioreactor that the en- is thermodynamically favorable (Hashsham et al., 2000; Matsum-
riched culture tended to fulfill the e equiv requirement for cell oto and Nishimura, 2007; Jo et al., 2008a). Indeed, by examining
growth at a lower VLR condition (21.77 kg-COD/m3/day), while it several isolated clostrdial species including C. acetobutylicum, C.
could start to largely distribute the e equiv for hydrogen produc- diolis, C. butyricum, and C. beijerinckii, previous studies (Diez-Gonz-
alez et al., 1995; Matsumoto and Nishimura, 2007) have confirmed
fermentative biohydrogen ability of these species from utilizing
lactate and acetate, although the growth rate of C. acetobutylicum
on lactate and acetate (0.05 h1) is much lower than that on glu-
cose (0.67 h1) (Diez-Gonzalez et al., 1995). In addition, Diez-
Gonzalez et al. (1995) has proposed the metabolic pathway of C.
acetobutylicum strain P262 for the utilization of lactate and con-
firmed that acetate was required as an electron acceptor.
Table 5 summarizes stoichiometric equations for hydrogen and
butyrate production from lactate and acetate utilization proposed
in previous studies. These equations suggest that approximately
1 mol of lactate and 0.43–0.5 mol of acetate utilization formed
0.5–0.6 mol of hydrogen and 0.7–0.83 mol of butyrate. In compar-
ison with stoichiometric equations obtained in this study
(Eqs (1)–5), relatively less acetate was consumed (0.16–0.42 mol)
Fig. 3. The calculated e equiv distributions (in percentage) for hydrogen, biomass and less hydrogen was produced (0.09–0.47 mol), especially at
and butyrate obtained in the batch 1 experiment. lower VLR conditions (VLR < 41.73 kg-COD/m3/day). As discussed
C.-W. Wu et al. / Bioresource Technology 113 (2012) 30–36 35

Fig. 4. The dependences of the calculated e equiv distributions (in percentage) for hydrogen and biomass on S0 and X0 (upper), and the corresponding fitted surfaces by
regression (bottom).

Table 5
improvement, this study was the first one, using a continuous flow
Stoichiometric equations for hydrogen and butyrate production from lactate and bioreactor, producing hydrogen from lactate and acetate
acetate utilization proposed in previous studies. utilization.
Equation Reference
1 Lactate + 0.5 acetate ? 0.75 Matsumoto and
butyrate + CO2 + 0.5 H2 + 0.5 H2O Nishimura (2007) 4. Conclusions
2 Lactate + 0.43 acetate ? 0.7 Diez-Gonzalez et al.
butyrate + CO2 + 0.57 H2 + 0.7 H2O (1995)
This study demonstrates fermentative hydrogen production
3 Lactate + 0.4 acetate + 0.7 H+ ? 0.7 Hashsham et al. (2000)
butyrate + CO2 + 0.6 H2 + 0.4 H2O from lactate and acetate utilization in a continuous flow bioreactor.
4 Lactate + 0.5 acetate ? 0.83 butyrate + 0.75 Juang et al. (2011) The VLR of 55.64 kg-COD/m3/day or the F/M ratio of 32.4 kg-COD/
CO2 + 0.5 H2 + 0.83 H2O kg-VSS/day seems to be optimum for the enriched culture for fer-
mentative hydrogen production from lactate and acetate. However,
a maximum lactate/acetate concentration allowed for enriching
this culture in a CSTR exists at a lower HRT condition, in which
previously, it is likely that the VLR condition or the S0/X0 ratio has wash-out can be an issue for this enriched culture. Finally, the re-
an effect on the e equiv distribution into biomass and hydrogen, sults of cloning and sequencing indicate that C. tyrobutyricum is
resulting in the difference in stoichiometry. Despite the fact that considered the major hydrogen-producing bacteria in the bioreac-
the hydrogen yield was slightly lower and required further tor fed with lactate and acetate.
36 C.-W. Wu et al. / Bioresource Technology 113 (2012) 30–36

Acknowledgements Lay, J.J., Lee, Y.J., Noike, T., 1999. Feasibility of biological hydrogen production from
organic fraction of municipal solid waste. Water Res. 33 (11), 2579–2586.
Lee, H.S., Rittmann, B.E., 2009. Evaluation of metabolism using stoichiometry in
The authors would like to acknowledge the financial support fermentative biohydrogen. Biotechnol. Bioeng. 102 (3), 749–758.
from the National Science Council of Taiwan under Grant NSC Lee, H.S., Salerno, M.B., Rittmann, B.E., 2008a. Thermodynamic evaluation on H2
production in glucose fermentation. Enviro. Sci. Technol. 42 (7), 2401–2407.
98-3114-E-006-013 and NSC100-3113-E-006-017, and NSC 101-
Lee, Z.K., Li, S.L., Lin, J.S., Wang, Y.H., Kuo, P.C., Cheng, S.S., 2008b. Effect of pH in
3113-E-006-016. fermentation of vegetable kitchen wastes on hydrogen production under a
thermophilic condition. Int. J. Hydrogen Energy 33, 5234–5241.
Lee, H.S., Vermaas, W.F.J., Rittmann, B.E., 2010. Biological hydrogen production:
prospects and challenges. Trends Biotechnol. 28 (5), 262–271.
References
Li, C.L., Fang, H.H.P., 2007. Fermentative hydrogen production from wastewater and
solid wastes by mixed cultures. Crit. Rev. Environ. Sci. Technol. 37 (1), 1–39.
Altschul, S.F., Gish, W., Miller, W., Myers, E.W., Lipman, D.J., 1990. Basic local Li, S.L., Whang, L.M., Chao, Y.C., Wang, Y.H., Wang, Y.F., Hsiao, C.J., Tseng, I.C., Bai,
alignment search tool. J. Mol. Biol. 215 (3), 403–410. M.D., Cheng, S.S., 2010. Effects of hydraulic retention time on anaerobic
APHA, 1998. Standard methods for the examination of water and wastewater. hydrogenation performance and microbial ecology of bioreactors fed with
American Public Health Association, Washington, DC. glucose-peptone and starch-peptone. Int. J. Hydrogen Energy 35, 61–70.
Balows, A., Truper, H., Dworkin, M., Harder, W., Schleifer, K.H., 1992. The Lin, P.Y., Whang, L.M., Wu, Y.R., Ren, W.J., Hsiao, C.J., Li, S.L., Chang, J.S., 2007.
prokaryotes. A handbook on the biology of bacteria: ecophysiology, isolation, Biological hydrogen production of the genus Clostridium: metabolic study and
identification, applications. Springer-Verlag GmbH, New York. mathematical model simulation. Int. J. Hydrogen Energy 32 (12), 1728–1735.
Chu, C.F., Li, Y.Y., Xu, K.Q., Ebie, Y., Inamori, Y., Kong, H.N., 2008. A pH- and Liu, I.C., Whang, L.M., Ren, W.J., Lin, P.Y., 2011. The effect of pH on the production of
temperature-phased two-stage process for hydrogen and methane production biohydrogen by clostridia: thermodynamic and metabolic considerations. Int. J.
from food waste. Int. J. Hydrogen Energy 33, 4739–4746. Hydrogen Energy 36 (1), 439–449.
Das, D., Veziroglu, T.N., 2001. Hydrogen production by biological processes: a Ljungdahl, L.G., Hugenholtz, J., Wiegel, J., 1989. Acetogenic and acid-producing
survey of literature. Int. J. Hydrogen Energy 26 (1), 13–28. bacteria. In: Minton, N.P., Clarke, D.J. (Eds.), Biotechnology Handbooks:
Davila-Vazquez, G., Arriaga, S., Alatriste-Mondragon, F., de Leon-Rodriguez, A., Clostridia. Plenum Press, New York.
Razo-Flores, E., 2008. Fermentative biohydrogen production: trends and Matsumoto, M., Nishimura, Y., 2007. Hydrogen production by fermentation using
perspectives. Sci. Biotechnol. Rev. Environ. 7 (1), 27–45. acetic acid and lactic acid. J. Biosci. Bioeng. 103 (3), 236–241.
Diez-Gonzalez, F., Russell, J.B., Hunter, J.B., 1995. The role of an nad-independent Owen, W.F., Stuckey, D.C., Healy Jr., J.B., Young, L.Y., McCarty, P.L., 1979. Bioassay for
lactate-dehydrogenase and acetate in the utilization of lactate by clostridium- monitoring biochemical methane potential and anaerobic toxicity. Water Res.
acetobutylicum strain p262. Arch. Microbio. 164 (1), 36–42. 13 (6), 485–492.
Han, S.K., Shin, H.S., 2004. Biohydrogen production by anaerobic fermentation of Rittmann, B.E., McCarty, P.L., 2001. Environmental biotechnology: principles and
food waste. Int. J. Hydrogen Energy 29, 569–577. applications: McGraw-Hill, New York.
Hashsham, S.A., Fernandez, A.S., Dollhopf, S.L., Dazzo, F.B., Hickey, R.F., Tiedje, J.M., Schmidt, O., Drake, H.L., Horn, M.A., 2010. Hitherto Unknown Fe–Fe – hydrogenase
Criddle, C.S., 2000. Parallel processing of substrate correlates with greater gene diversity in anaerobes and anoxic enrichments from a moderately acidic
functional stability in methanogenic bioreactor communities perturbed by fen. Appl. Environ. Microbiol. 76 (6), 2027–2031.
glucose. Appl. Environ. Microbiol. 66 (9), 4050–4057. Van Ginkel, S.W., Oh, S.E., Logan, B.E., 2005. Biohydrogen gas production from food
Herbert, D., Philipps, P., Strange, R., 1971. Carbohydrate analysis. Methods Enzymol. processing and domestic wastewaters. Int. J. Hydrogen Energy 30, 1535–1542.
5, 265–277. Van Ginkel, S., Sung, S., Lay, J.J., 2001. Biohydrogen production as a function of pH
Jo, J.H., Lee, D.S., Park, D., Park, J.M., 2008a. Biological hydrogen production by and substrate concentration. Enviro. Sci. Technol. 35 (24), 4726–4730.
immobilized cells of Clostridium tyrobutyricum JM1 isolated from a food waste Whang, L.M., Hsiao, C.J., Cheng, S.S., 2006. A dual-substrate steady-state model for
treatment process. Bioresour. Technol. 99 (14), 6666–6672. biological hydrogen production in an anaerobic hydrogen fermentation process.
Jo, J.H., Lee, D.S., Park, J.M., 2008b. The effects of pH on carbon material and energy Biotechnol. Bioeng. 95 (3), 492–500.
balances in hydrogen-producing Clostridium tyrobutyricum JM1. Bioresour. Whang, L.M., Lin, C.A., Liu, I.C., Wu, C.W., Cheng, H.H., 2011. Metabolic and energetic
Technol. 99 (17), 8485–8491. aspects of biohydrogen production of Clostridium tyrobutyricum: the effects of
Juang, C.P., Whang, L.M., Cheng, H.H., 2011. Evaluation of bioenergy recovery hydraulic retention time and peptone addition. Bioresour. Technol. 102 (18),
processes treating organic residues from ethanol fermentation process. 8378–8383.
Bioresour. Technol. 102 (9), 5394–5399. Zhu, H., Stadnyk, A., Béland, M., Seto, P., 2008. Co-production of hydrogen and
Kalia, V.C., Jain, S.R., Kumar, A., Joshi, A.P., 1993. Fermentation of biowaste to H2 by methane from potato waste using a two-stage anaerobic digestion process.
Bacillus licheniformis. World J. Microbiol. Biotechnol. 10, 224–227. Bioresour. Technol. 99, 5078–5084.
Kim, S.H., Han, S.K., Shin, H.S., 2004. Feasibility of biohydrogen production by Zhu, Y., Yang, S.T., 2004. Effect of pH on metabolic pathway shift in fermentation of
anaerobic co-digestion of food waste and sewage sludge. Int. J. Hydrogen xylose by Clostridium tyrobutyricum. J. Biotechnol. 110 (2), 143–157.
Energy 29, 1607–1616.

You might also like