You are on page 1of 10

Bioresource Technology 102 (2011) 8524–8533

Contents lists available at ScienceDirect

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

Bioreactor and process design for biohydrogen production


Kuan-Yeow Show a, Duu-Jong Lee b,⇑, Jo-Shu Chang c
a
Department of Environmental Engineering, Faculty of Engineering and Green Technology, Universiti Tunku Abdul Rahman, Jalan University, Bandar Barat,
31900 Kampar, Perak, Malaysia
b
Department of Chemical Engineering, National Taiwan University of Science and Technology, Taipei 106, Taiwan
c
Department of Chemical Engineering, National Cheng Kung University, Tainan 701, Taiwan

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

Article history: Biohydrogen is regarded as an attractive future clean energy carrier due to its high energy content and
Received 28 November 2010 environmental-friendly conversion. It has the potential for renewable biofuel to replace current hydrogen
Received in revised form 16 April 2011 production which rely heavily on fossil fuels. While biohydrogen production is still in the early stage of
Accepted 18 April 2011
development, there have been a variety of laboratory- and pilot-scale systems developed with promising
Available online 30 April 2011
potential. This work presents a review of advances in bioreactor and bioprocess design for biohydrogen
production. The state-of-the art of biohydrogen production is discussed emphasizing on production path-
Keywords:
ways, factors affecting biohydrogen production, as well as bioreactor configuration and operation. Chal-
Biohydrogen
Biophotolysis
lenges and prospects of biohydrogen production are also outlined.
Bioreactor Ó 2011 Elsevier Ltd. All rights reserved.
Dark fermentation
Photo-fermentation

1. Introduction While attractive features of biohydrogen production have been


demonstrated in laboratory studies, the technology is yet to com-
There have been growing concerns of greenhouse gas emissions pete with commercial hydrogen production processes from fossil
and other pressing environmental issues over the use of fossil fuels. fuels in terms of cost, efficiency and reliability (Lee et al., 2010).
Hydrogen fuel is a promising alternative to conventional fossil Thus, improving the efficiency of biohydrogen production poses a
fuels because it has the potential to eradicate all the environmental major challenge to many researchers. This review provides an
problems that the fossil fuels would create. Many scientists per- overview of the state-of-the-art and perspectives of biohydrogen
ceive hydrogen gas as the future fuel because of its non-polluting production research. The review focuses on biohydrogen produc-
features and its use in highly efficient fuel cells for electricity. tion pathways, major factors affecting biohydrogen production,
However, a major doubt over the use of hydrogen as a clean energy as well as bioreactor configuration and operation. Challenges and
alternative is the way it is produced. The current hydrogen gas pro- prospects of biohydrogen production are also outlined.
duction is unfriendly as it is being generated from fossil fuels
through thermo-chemical processes, such as hydrocarbon reform- 2. Hydrogen production pathways
ing, coal gasification and partial oxidation of heavier hydrocarbons.
Biohydrogen production by microorganisms has attracted Biohydrogen can be generated via different pathways which can
increasing global attention, owing to its potential for inexhaustible, be broadly categorized into two distinct groups, viz. light-depen-
low-cost and renewable source of clean energy. Studies on biohy- dent and dark fermentative processes. Light-depended processes
drogen production have been focusing on biophotolysis of water include photolysis and photo-fermentation, whereas dark fermen-
using algae and cyanobacteria, photo-decomposition of organic tation is the major light independent process.
compounds by photosynthetic bacteria, and dark fermentation
from organic compounds with anaerobes. Anaerobic hydrogen fer- 2.1. Photolysis
mentation appears to be most favorable since hydrogen could be
generated at higher rates. Moreover, the process can be carried Hydrogen can be produced directly via water-splitting process
out on various organic wastes and wastewaters enriched with car- through the photosynthetic capability of algae and cyanobacteria.
bohydrates, thereby achieving sustainable low cost biohydrogen The biohydrogen production occurs via direct absorption of light
production with concomitant waste purification. and transfer of electrons to hydrogenases and/or nitrogenases en-
zymes. Under anaerobic or excessive energy conditions, the excess
⇑ Corresponding author. electrons are released by microorganisms using hydrogenase
E-mail address: djlee@ntu.edu.tw (D.-J. Lee). enzyme which converts the hydrogen ions to hydrogen gas (Turner

0960-8524/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2011.04.055
K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533 8525

et al., 2008). The protons and electrons extracted via the 2.3. Dark fermentation
water-splitting reactions are recombined by chloroplast hydroge-
nase to form molecular hydrogen gas (Hankamer et al., 2007). Dark fermentation under anaerobic conditions seems to be the
In the water-splitting process, oxygen is also produced which in most favorable among the bioproduction processes. The fermenta-
turn suppresses hydrogen production (Kapdan and Kargi, 2006). To tion can be carried out at higher rates and lower cost using various
prevent oxygen buildup leading to hydrogen suppression, research organic substrates and wastewaters (Hallenbeck and Ghosh, 2009).
work has been carried out to identify or engineer less oxygen-sen- Dark fermentation uses primarily anaerobic bacteria, although
sitive algae and bacteria, separate the hydrogen and oxygen se- some algae are also used, on carbohydrate rich substrates grown
quence, or change the ratio of photosynthesis to respiration without the need of light energy (Kapdan and Kargi, 2006).
process (US DOE, 2007). Sulfate has been found effective in sup- Molecular hydrogen formation is generally realized via two
pressing oxygen generation, but the hydrogen production is also pathways in the presence of specific coenzymes, i.e. either by for-
inhibited (Turner et al., 2008). mic acid decomposition route or by the re-oxidization of nicotin-
With the major substrate being water which is abundant, direct amide adenine dinucleotide (NADH) route (Eqs. (1) and (2))
photolysis via water-splitting demonstrates promising future. 2þ þ
However, there are challenges to be dealt with. Current problems NADH þ Hþ þ 2Fd !2Hþ þ NADþ þ 2Fd ð1Þ
such as the need of large surface area to collect light and difficulty
þ 2þ
in achieving continuous hydrogen production under aerobic condi- 2Fd þ 2Hþ ! 2Fd þ H2 ð2Þ
hydrogenase
tions are to be addressed (US DOE, 2007).
Biohydrogen can also be produced through some microorgan- The Embden–Meyerhof or glycolytic pathway is undoubtedly
isms like algae that can directly produce hydrogen under certain the most common route for glucose degradation to pyruvate,
conditions. It has been reported that sulfur-deficient green algae which is found in all major of microorganisms and functions in
whose energy was derived from light under anaerobic conditions the presence or absence of oxygen (Prescott et al., 2002). In this
could induce the hydrogenase reactions to produce hydrogen pathway, glucose is converted into pyruvate associated with the
photo-synthetically (Laurinavichene et al., 2008). Under the sul- conversion of NADH from NAD+ via anaerobic glycolysis, which
fur-deprivation condition, the algae consume large amounts of cel- could be represented by Eq. (3):
lular starch and protein. Apparently, these catabolic reactions help
to sustain indirectly the hydrogen production process. C6 H12 O6 þ 2NADþ ! 2CH3 COCOOH þ 2NADH þ 2Hþ ð3Þ
Photolysis biohydrogen production is deemed feasible if the The disposal of electrons via pyruvate-ferredoxin oxidoreduc-
efficiency of photon conversion can be improved via large-scale al- tase or NADH-ferredoxin oxidoreductase and hydrogenase might
gal bioreactors. Engineering approach to regulate light inputs in be affected by the corresponding NADH and acetyl-CoA levels as
improving the algae photon conversion efficiency can be optimized well as environmental conditions. As a result, the oxidation–reduc-
through the algal bioreactor operation. Substantial increases in tion state has to be balanced through the NADH consumption to
light conversion efficiency from about 5% to 15% had been reported form some reduced compounds, i.e. lactate, ethanol and butanol,
(Laurinavichene et al., 2008). Attaining 10–13% efficiency is feasi- resulting in a lowered hydrogen yield.
ble by using molecular approach to engineer the microorganisms In the recent development, metabolic engineering has received
to better utilize the light energy (Turner et al., 2008). increasing attention in improving biohydrogen production.
Improvements in hydrogen yields by existing pathways has been
2.2. Photo-fermentation attempted by increasing the flux through gene knockouts of com-
peting pathways or increased homologous expression of enzymes
Unlike the photolysis process in which hydrogen is produced by involved in the hydrogen-generating pathways (Hallenbeck and
cyanobacteria and/or green algae directly or indirectly from water, Ghosh, 2009). To further elucidate a better understanding in
purple photosynthetic microbes are able to generate hydrogen improving biohydrogen production, major factors affecting hydro-
from organic substrates through photo-fermentation. Despite the gen production such as operating conditions, feedstock and reactor
relatively lower yields of hydrogen from photosynthetic bacteria, configuration will be systematically reviewed and extensively dis-
the fermentative pathway is a promising process of biohydrogen cussed in the following sections.
production, due mainly to its higher rate of hydrogen evolution
in the absence of any light source (Wang et al., 2010a). The process
can be versatile in terms of types of feed to the microbes (Redwood 3. Factors affecting biohydrogen production
et al., 2009).
In contrast to green algae, the photosynthetic purple bacteria are Inevitably, performance of hydrogen-producing bioreactor sys-
unable to split water by their simple photosynthetic structure. tems and operation are dictated by various factors. Factors associ-
However, these bacteria are able to use simple organic acids or even ated with environmental conditions, process operating conditions,
dihydrogensulfide as electron donor derived from the organic sub- and chemical conditions are reviewed in this section.
strates under anaerobic conditions (Akkerman et al., 2003). The elec-
trons extracted from the organics are pumped through electron 3.1. Culture pH
carriers, during which protons are driven across the membrane.
An electrochemical gradient is developed and the ATP synthase syn- System pH has been recognized to be one of the most important
thesize adenosine triphosphate (ATP) from adenosine diphosphate environmental factors affecting metabolic pathways and the yield
(ADP) by using the energy generated from protons moving down of biohydrogen. Comparative studies with respect to the effect of
the gradient. The surplus ATP energy can be used to transport the pH on continuous hydrogen production revealed that the optimum
electrons further to ferredoxin – an electron carrier in the chloro- pH range to achieve the maximum hydrogen yield or specific
plasts. When molecular nitrogen is deficient, the ATP synthase can, hydrogen production rate was found between 5.2 and 6.0 in most
again with the extra ATP energy, reduce protons into hydrogen gas studies using pure or mixed cultures of bacteria (Oh et al., 2004a;
using the electrons derived from the ferredoxin (Akkerman et al., Zhang et al., 2008a). Fang and Liu (2002) investigated the effect of
2003). In this manner, the organic acids derived from the substrates pH on hydrogen production from glucose in a CSTR over the range
can be ultimately transformed into hydrogen and carbon dioxide. of 4.0–7.0 and found the optimum hydrogen yield occurring at a
8526 K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533

pH of 5.5. Chang and Lin (2004) observed hydrogen production 3.4. Nutrients
from glucose using continuous cultures under identical conditions
except pH values, and concluded that higher hydrogen yield and In hydrogen fermentation processes, nitrogen, phosphate and
production rate are attained at a pH of 5.7 compared with pH other inorganic trace minerals are necessary supplements for car-
6.4. Ren et al. (2009) emphasized the maximum hydrogen yield bohydrate-based feedstocks in order to obtain optimal cell cultiva-
was achieved only when microbial reactions followed an ethanol tion and hydrogen production. Previous study indicated that
fermentation type that occurred at a pH of about 4.5. organic nitrogen seems to be more favorable for hydrogen evolu-
Venkata Mohan (2009) reported that the initial pH values of tion compared with inorganic one (Yokoi et al., 2001). The highest
5.5–7.5 may represent the optimum and acceptable ranges of pH level of H2 production from sweet potato starch residue by a re-
for hydrogen production, whereas the hydrogen yield may be shar- peated batch culture containing C. butyricum IFO13949 when
ply dropped at pH out of the optimum range. While system pH be- 1.0% polypepton was added as nitrogen source was reported (Yokoi
low 6 had shown negative influence on the substrate degradation et al., 2001). In contrast, addition of urea, (NH4)2SO4 or NH4Cl re-
efficiency due to the inhibition of methanogens, increase in feed sulted in the absence of hydrogen evolution by the same culture.
pH has resulted in suppressed hydrogen production. pH control Phosphate is one of the important inorganic nutrients required
could stimulate the microorganisms to achieve maximum hydro- for optimal hydrogen production (Lin and Lay, 2004a). Excess
gen production ability because the activity of hydrogenase was phosphate may favor VFAs and hydrogen production over solvent
inhibited at low or high pH during fermentation (Venkata Mohan, production, so phosphate supplementation may be needed with
2009). carbohydrate-rich feeds.
Lin and Lay (2004b) examined the effect of trace elements
3.2. Hydraulic retention time including Mg, Na, Zn, Fe, K, I, Co, NHþ 4 , Mn, Ni, Cu, Mo and Ca on
hydrogen production by a C. pasteurianum-predominant culture.
Hydraulic retention time (HRT) could be used as a tool to se- The authors found elements Mg, Na, Zn and Fe being important
lect microbial populations whose growth rates are able to catch supplements and proposed an optimum nutrient formulation con-
up with mechanical dilution created by continuous volumetric taining (mg l1) MgCl26H2O 120, NaCl 1000, ZnCl2 0.5 and FeS-
flow. Typical specific growth rate of methane-producing bacteria O47H2O 3 for the optimal hydrogen yield. However, less effect of
is about 0.0167–0.02 h1, which is much lower than that of NHþ 4 and a lower value of optimum iron concentration indicated
hydrogen-producing bacteria of about 0.172 h1 (Lo et al., here are inconsistent with those values recommended in other
2009a). Therefore, hydrogen-producing bacteria can be retained studies (Lin and Lay, 2004c; Wang and Wan, 2009).
while washing out methane-producing bacteria by regulating Iron is a key component in the enzymatic activity of hydrogen
HRT. Lo et al. (2009a) found that the methane concentration ran- production. Biological formation and consumption of molecular
ged between 0.0011 and 0.0058 mol l1 at low dilution rates hydrogen (H2) are catalyzed by hydrogenases, of which three phy-
(D = 0.002–0.0167 h1) was hardly measureable at higher dilu- logenetically unrelated types are known: [NiFe]-hydrogenases,
tions (D > 0.075 h1), indicating negligible methanogenic activity [FeFe]-hydrogenases, and [Fe]-hydrogenase (Shima et al., 2008).
at high dilution rates. This means that HRT is capable of inhibit- Among these classes, [FeFe]- and [NiFe]-hydrogenases are the ma-
ing or terminating methanogenesis in hydrogen production via jor enzymes; most of hydrogenases found in microorganisms are
anaerobic fermentation. belong to one of these two enzymes except [Fe]-hydrogenase that
Zhang et al. (2006a) found that shortening the HRT from 8 h to were discovered in some methanogens. The mostly monomeric
6 h would reduce microbial diversity associated with inhibition of [FeFe]-hydrogenases are more involved in H2 evolution, and dis-
propionate production without affecting the existence of dominant play high sensitivity to oxygen (O2) and carbon monoxide (CO)
species, leading to an increase in the hydrogen yield. Similar find- (Bleijlevens et al. 2004; Frey, 2002). [NiFe]-hydrogenases, com-
ing was also reported by other researchers (Hussy et al., 2003). posed of two subunits, are involved in H2 oxidation, but also cata-
These results provided evidence that hydrogen yield, a function lyze reversible reactions (Adams and Hall, 1979; Nishihara et al.,
of the microbial populations, can be affected by HRT (Zhang 1997). These enzymes are less active than [FeFe]-hydrogenases
et al., 2006a). This might explain the view of HRT-dependent by 10–102-fold, but much more tolerant to O2 or CO in a reversible
hydrogen yield obtained by some other researchers in the similar manner (Frey, 2002). This property is a significant advantage for
systems (Fan et al., 2006; Venkata Mohan, 2009). the biotechnological application of hydrogenase in biohydrogen
production, since it is not necessary to protect the production pro-
3.3. Hydrogen partial pressure cess from O2.

The dissolved hydrogen concentration in the liquid phase, con- 3.5. Temperature
tributing to the hydrogen partial pressure, is one of the key factors
affecting microbial pathways in fermentative hydrogen produc- Microbes are capable of producing hydrogen in a temperature
tion. Hydrogen production is less favorable as the hydrogen partial range of 15–85 °C (Kanai et al., 2005), but laboratory-scale studies
pressure rises (Tiwari et al., 2006). from a review indicated that about 73% studies were carried out
Thus, it is important to remove excess hydrogen from the sys- with mesophilic cultures (Li and Fang, 2007). Chang and Lin
tem to maintain hydrogen production. Several studies have re- (2004) studied the hydrogen production capability of a mixed cul-
vealed that reducing hydrogen partial pressure can considerably ture under varying temperatures from 15 to 34 °C and found that
enhance hydrogen production. Many strategies of removal or sep- hydrogen yield and specific hydrogen production rate increased
arating excess hydrogen gas have been developed to avoid the neg- with temperature, achieving respective maximum values of
ative effect of the hydrogen accumulation in the gas phase. Mizuno 359 mmol l1 d1 and 1.42 mol H2 mol1 glucose at 30–34 °C and
et al. (2000) found an increase in the hydrogen yield from 0.85 to 28–32 °C, respectively. However, both values obtained at the ambi-
1.43 mol mol1 hexose when the reactor is sparged with nitrogen ent temperatures are much lower compared with the correspond-
at 15 times hydrogen produced rate. Hussy et al. (2003) also re- ing values of 574 mmol l1 d1 and 1.70 mol H2 mol1 glucose at a
ported a 1.5 times enhanced hydrogen yield from 1.3 to consistent temperature of 35 ± 1 °C. Minnan et al. (2005) also re-
1.9 mol mol1 hexose by sparging the reactor with nitrogen to re- ported that increasing temperatures from 25 °C to 35–36 °C im-
duce hydrogen in the off-gas from 50% to 7%. proved the hydrogen production activity.
K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533 8527

Studies using mesophilic cultures (Lee et al., 2006) seemed to able to support a wide variety of metabolite activities, and they are
indicate that, although hydrogen-producing bacteria are able to potentially more resilient to changes in environmental conditions
perform at ambient temperature conditions, increasing tempera- (Hallenbeck and Ghosh, 2009).
ture in the mesophilic regime always improves the hydrogen pro-
duction, while further increasing culture temperature beyond
3.8. Feedstock
mesophilic range may cause a decrease in hydrogen production,
presumably dependent on physiological characteristics of micro-
Simple sugars such as glucose, sucrose and lactose are readily
bial cultures.
biodegradable and are thus preferred as model substrates for
hydrogen production. However, the costs for pure carbohydrate
3.6. Substrate concentration
sources are high for practical-scale hydrogen production, which
can only be viable when based on renewable and low cost sources.
The effect of substrate concentration on hydrogen production,
As reported by numerous studies of biohydrogen fermentative pro-
however, has been a point of debate. In studies concerning the ef-
cesses, carbohydrates are the main source of hydrogen. Thus,
fect of substrate concentration, Kim et al. (2006) found hydrogen
wastes and biomass rich in sugars and/or complex carbohydrates
yield to increase with increasing glucose concentration from 10
turn out to be the most suitable feedstocks for biohydrogen gener-
to 35 g l1 at a HRT of 12 h, whereas Kyazze et al. (2006) examined
ation (Lo et al., 2008, 2009a; Luo et al., 2011; Ntaikou et al., 2010).
continuous hydrogen production at 12 h HRT on 10–50 g l1 su-
Some complex substrates are not ideal for fermentative hydro-
crose and found that the hydrogen yield decreased from
gen production due to their complex structures. The main technical
1.7 ± 0.2 mol H2 mol1 hexose added at 10 g l1 sucrose to
difficulty for practical hydrogen production is the process stability
0.8 ± 0.1 mol H2 mol1 hexose added at 50 g l1. A decrease in
and the short duration of the hydrogen production from wastes
hydrogen yield could occur at HRTs of 2.5 h and 10 h when the
and complex substrates. Studies have indicated that it might be
CSTR feeding strength increased from 10 g glucose l1 to 40 g glu-
uneconomical to recover hydrogen from waste such as wastewater
cose l1 (Van Ginkel and Logan, 2005). These studies indicate that
sludge (Okamoto et al., 2000; Van Ginkel et al., 2005). The hydro-
besides substrate concentration, other operating conditions such
gen yields obtained were rather low.
as HRT and composition of microbial cultures also affect continu-
Complex substrates, however, after being pretreated by some
ous hydrogen production. As iron is a key component in the enzy-
methods, they can be easily used by hydrogen-producing bacteria
matic activity of biohydrogen production, the observed decrease of
(Luo et al., 2010). For example, Zhang et al. (2007b) reported that
the hydrogen yield in continuous cultures when increasing the or-
the hydrogen yield from cornstalk wastes after acidification pre-
ganic loading rate could also be due to limitation (starvation) in
treatment was much larger than that from cornstalk wastes with-
iron compound. At high glucose inlet concentrations, the limiting
out any pretreatment. Complex solid wastes, such as organic
factor may not be glucose but iron in the substrate.
residues, food processing, mixed wastes, digester sludge, and mu-
A wide range of substrate concentration ranging from 2.5 to
nicipal wastes have also been tested as feedstocks for fermentative
120 g hexose l1 had been reported (Zhang et al., 2007a). With a
hydrogen production (Juang et al., 2011; Kim and Lee, 2010; Ntai-
high feeding sucrose concentration of up to 30–40 g COD l1 the
kou et al., 2010). Apart from carbohydrates, such wastes usually
granular sludge-based continuously stirred anaerobic bioreactor
have quite high contents of proteins and fats, and thus their bio-
produced hydrogen more efficiently at a short HRT of 0.5 h, with
transformation to hydrogen is comparatively lower than those ob-
the highest volumetric rate of 15 l h1 l1 and an optimal yield of
tained from carbohydrate based wastewaters. Biotransformation of
ca. 3.5 mol H2 mol1 sucrose (Wu et al., 2006). Although the opti-
wastes and wastewater towards hydrogen can be considered quite
mum hydrogen yield was obtained at a glucose concentration of
appealing from both the environmental and the economical stand-
10 g l1 in granule sludge or biofilm sludge reactors, Zhang et al.
point (Ntaikou et al., 2010).
(2007a) noted that hydrogen production was almost kept consis-
tent at a substrate concentration ranging from 10 to 30 g l1, but
sharp decrease in hydrogen yield took place at glucose concentra- 4. Bioreactor configuration and operation
tions larger than 40 g l1.
4.1. Photo-bioreactors
3.7. Seed culture
Design of a photo-bioreactor depends on microbiological pro-
Clostridium and Enterobacter are most widely used as inocu- cesses associated with microalgae, diatoms or cyanobacteria.
lums for fermentative hydrogen production (Wang and Wan While these photo-heterotrophic bacteria differ in photochemical
2009; Zhang et al., 2006b, 2008b). Most of the studies using pure efficiency, absorption coefficient and size, the light regime includ-
cultures of bacteria for fermentative hydrogen production were ing light and dark cycles is assumed to be much more determining
conducted in batch mode and used glucose as substrate (Wang than biological factors (Akkerman et al., 2003). Hence the produc-
and Wan, 2009). Mixed cultures of bacteria from anaerobic sludge, tivity of a photo-bioreactor is light-dependent, and a large surface
municipal sewage sludge, compost and soil had been widely used to volume ratio is a prerequisite for a productive photo-bioreactor
as inoculums for fermentative hydrogen production (Li and Fang, for optimal light exposure of the algae. Provisions for thermal con-
2007). Fermentative hydrogen production processes using mixed trol and monitoring flow rates, pH, and dissolved oxygen, sulfur,
cultures are more practical than those using pure cultures, as the and hydrogen are essential. Technical development is now moving
former are simpler to operate and easier to control, and may have towards devising gas-tight systems, microalgae culturing as well as
a broader source of feedstock (Li and Fang, 2007). In addition, there computer-controlled system for monitoring and automatic nutri-
are also several other potential merits of using microbial consortia ent delivery and culture dilution.
instead of pure cultures. Practical hydrogen fermentations will Photo-bioreactors have been designed to achieve an economi-
have to be carried out under non-sterile conditions using readily cal, rapid multiplication and high density of the microalgae culture
available feedstocks with minimal pretreatment. Mixed consortia (Evens et al., 2000). Different photo-bioreactor designs including
would be able to address these issues as they have been selected flat plate, tubular, pond or pool-type, have been investigated (Akk-
for growth and are robust under non-sterile conditions (Zhang erman et al., 2003). Detailed cost analyses must be conducted on
et al., 2007b, 2008b). As microbial enriched consortia, they are also minimizing the materials and operating costs, and also for optimiz-
8528 K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533

ing the yield and gas collection. Critical parameters in the cost that self-flocculated granular sludge occurred at a HRT of 0.5 h
analyses include the light environment, the climate and land space, and reached a concentration of up to 35.4 g VSS l1, which in turns
reactor construction materials, mechanism of culture mixing, reac- resulted in a significant increase in hydrogen production rate up to
tor maintenance and long-term operational stability with maximal 15 l h1 l1. In another CSTR study with granular sludge fermenting
gas production. Adequate light supply is vital and light limitations glucose wastewater (10 g l1) at a pH of 5.5 and 37 °C, Zhang et al.
must be kept to a minimum. Light conversion efficiencies are low (2007c) reported a maximum hydrogen yield of 1.81 mol H2 mol1
(theoretically not more than 10%) and tend to decrease at higher glucose and a maximum hydrogen production rate of 3.20 l h1 l1
light intensities because of light saturation effect (Akkerman at a HRT of 0.5 h. Show et al. (2007) and Zhang et al. (2007d) found
et al., 2003). Hence for efficient biohydrogen production, it is crit- that formation of granular sludge significantly increased overall
ical to dilute the light and distribute it over the reactor volume, and reactor biomass to as much as 16.0 g VSS l1, which enabled CSTR
mix the culture at high rates, so that cells are exposed to the light to operate at an OLR of up to 20 g glucose h1 l1 and hence en-
only for a short time. hanced performance in H2 production.
The exact configuration of the bioreactor also needs to be estab-
lished for the most effective use of light and surface area. Biomass 4.2.2. Anaerobic sequencing batch reactor
mixing is hence significant to ensure uniform dispersion of nutri- The influence of substrate loading rate on fermentative hydro-
ents and illumination of the culture, and to prevent sedimentation gen production was studied in biofilm configured sequencing
of culture. Modular design of experimental system should be al- batch reactor using chemical wastewater as substrate (Vijaya
lowed for possible scale-up. Such commercial scale should achieve Bhaskar et al., 2008). Reactor was operated with selectively en-
sustainable gas output and high hydrogen yields with compact riched anaerobic mixed microflora at different organic loading
configuration. Trapping and removal of hydrogen gas in the system rates after adjusting the feed to a pH of 6.0 to provide suitable envi-
are also important design consideration for photo-bioreactors. Gi- ronment for acidogenic bacterial function. Variation in hydrogen
ven the current advancement in photo-biohydrogen production, production yield from 6.064 to 13.44 mol H2 kg1 COD d1 was ob-
technical and economic strategies for cycling of the microalgae served with change in organic loading rate from 6.3 to
between sulfur-deprivation and supply must be developed 7.9 kg COD m3 d1 (Vijaya Bhaskar et al., 2008). Increase in the or-
(Laurinavichene et al., 2008). ganic loading rate showed marked reduction in COD removal effi-
ciency from 22.6% to 17.2%.
4.2. Dark fermentation bioreactors In another study using a sequencing batch reactor (Buitron and
Carvajal, 2010), the effect of the temperature, the hydraulic reten-
Studies on batch, semi-continuous and continuous hydrogen- tion time (HRT), and initial substrate concentration on hydrogen
producing bioreactors have been conducted. Batch hydrogen fer- production from Tequila vinasse was examined. When 25 °C and
mentation normally brings about lower hydrogen production rates 12 h HRT were applied, insignificant quantity of biogas was pro-
in comparison with its semi-continuous or continuous counterpart. duced. Using a longer HRT of 24 h and temperatures of 25–35 °C,
From the engineering’s standpoint, continuous hydrogen produc- biogas containing hydrogen was produced. A maximum volumetric
tion is preferred. Besides the extensively studied Continuous Stir- hydrogen production rate of 50.5 ml H2 h1 l1 and an average
red Tank Reactor (CSTR), numerous biohydrogen bioreactor hydrogen content in the biogas of 29.2% were obtained when the
processes such as Anaerobic Sequencing Batch Reactor (ASBR), reactor was fed with 3 g COD l1 at 35 °C and 12 h HRT.
Membrane Bioreactor (MBR), Fixed-bed Bioreactor, Fluidized-bed
Bioreactor and Upflow Anaerobic Sludge Blanket (UASB) Bioreactor 4.2.3. Membrane bioreactor
have been developed with high production yields and output. Use of a membrane bioreactor (MBR) to control biomass con-
centration was reported (Oh et al., 2004b). Operated with a HRT
4.2.1. Continuous stirred tank reactor of 3.3 h, the biomass concentration increased from 2.2 g l1 in a
Continuous stirred tank reactor (CSTR) are commonly used for control reactor (no membrane chemostat) to 5.8 g l1 in an anaer-
continuous hydrogen production (Younesi et al., 2008; Ding obic MBR by controlling sludge retention time (SRT) at 12 h. Under
et al., 2010). In a CSTR, hydrogen-producing microbes are com- such operating conditions, the hydrogen production rate increased
pletely-mixed and suspended in the reactor liquor from the mixing from 0.50 to 0.64 l h1 l1. Increasing the SRT would further en-
pattern. Under such hydrodynamics, a good substrate-microbes hance biomass retention which favors substrate utilization, but re-
contact and mass transfer can be accomplished. On the other hand, sults in a decrease in hydrogen production rate. In a review by Li
the CSTR is unable to maintain high levels of fermentative biomass and Fang (2007) on studies of hydrogen production by MBR, hydro-
because of the rapidly mixed operating pattern. Biomass washout gen production rates in a range of 0.25–0.69 l h1 l1 was reported.
may occur at short hydraulic retention times (HRTs), thus the It was concluded, however, that MBR did not exhibit superiority
hydrogen production rates are considerably restricted. Depending other high-rate hydrogen production systems. Membrane fouling
on the HRT, volatile suspended solids retained in the bioreactor and high operating cost also limit the use of MBR process in biohy-
normally ranged between 1 and 4 g VSS l1 (Show et al., 2007, drogen fermentation.
2010; Zhang et al., 2006a). The highest CSTR hydrogen production
rate of 1.12 l h1 l1 was reported in fermentation of sucrose with a 4.2.4. Fixed-bed bioreactor
mixed hydrogen-producing culture. A fixed-bed bioreactor is operated with support carriers packed
Fang et al. (2002) demonstrated that hydrogen-producing aci- within the tank. The hydraulic mixing regime is less turbulent
dogenic sludge could agglutinate into granules in a well-mixed comparing with the CSTR, this results in higher mass transfer resis-
CSTR treating a synthetic sucrose-base wastewater at 26 °C and a tance along with lowered rates of substrate conversion and hydro-
pH of 5.5. Operated at a HRT of 6 h, formation of the granular gen production. High hydrogen yield could not be maintained
sludge significantly enhanced retention of biomass (up to consistently in a fixed-bed reactor, because pH gradient distribu-
20 g l1) with hydrogen production rate of up to 0.54 l h1 l1. tion along the reactor column caused a heterogeneous distribution
The CSTR system examined by Wu et al. (2006) was initially seeded of microbial activity. To overcome the mass transfer resistance and
with silicone-immobilized sludge at 40 °C and pH 6.6 ± 0.2, and pH heterogeneous distribution, recirculation flow was recom-
reactor performance was examined at a HRT of 0.5–6 h and an mended. Fixed-bed reactors with tubular, tapered or rhomboid
influent sucrose concentration of 10–40 g COD l1. It was noted shape configurations operated at a HRT of 1.08 h were reported.
K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533 8529

The study reviewed that the rhomboid bioreactor with conver- model is used to formulate prediction of the flow field and the
gent–divergent configuration has a maximum hydrogen produc- reaction kinetics model then portrays the reaction conversion pro-
tion rate of 1.60 l h1 l1 as compared with tapered reactor cess. The coupled model is verified and used to simulate the behav-
(1.46 l h1 l1) and tubular reactor (1.40 l h1 l1). The superiority ior of the EGSB reactor for biohydrogen production. The coupled
in the production rate could be attributed to higher turbulent mix- model also demonstrates a qualitative relationship between
ing favoring mass transfer and low gas hold-up. It was observed hydrodynamics and biohydrogen production.
that both hydrogen production and substrate conversion rates in-
creased with slurry recycling ratio. 4.3. Microbial electrolysis cells
Support carriers play an important role in biomass retention
and hydrogen production in fixed-bed reactors. Loofah sponge, ex- Studies of microbial electrolysis cells or microbial electro-
panded clay, and activated carbon had been used to support hydrogenesis cells (MECs) for biohydrogen production through
growth of sewage sludge culture (Lee et al., 2006; Lo, 2009b). It electro-hydrogenesis have been reported recently (Call and Logan,
was found that loofah sponge is inefficient for biomass immobili- 2008; Wang et al., 2011). The MEC is a modified microbial fuel cell,
zation, while other carriers exhibit better biomass yields. Activated in which organic substrates are converted into biohydrogen. In
carbon is a better choice for support carriers, which resulted in the contrast to microbial fuel cells, a MEC operates with a supply of
maximum hydrogen production rate of 1.32 l h1 l1 by the fixed- an external voltage rather than generated by it. The energy supply
bed bioreactor operated at a HRT of 1 h and a sucrose concentra- is needed since the substrate conversion is not spontaneous under
tion of 20 g l1. standard conditions (Call and Logan, 2008). Anaerobic microorgan-
isms in a MEC decompose organic substances and transfer elec-
4.2.5. Fluidized-bed bioreactor trons to the anode. After traveling through an external load, the
In an examination on immobilized sewage sludge, Lin et al. electrons combine at the cathode, with protons and oxygen form-
(2009) found the microbial culture could produce hydrogen effi- ing water. Hydrogen gas production occurs at the cathode via reac-
ciently in a 3-phase fluidized-bed reactor at a HRT ranging from tion of hydrogen ion with electrons. MECs can be operated with
2 to 6 h with a maximal steady-state rate of 1.82 l h1 l1 and an versatility of microbial community and substrates. In addition to
optimal yield of 4.26 mol H2 mol1 sucrose. A much higher hydro- decomposing fermentation metabolites as substrate, glucose or
gen production rate of 2.36 l h1 l1 was recorded by Zhang et al. glucose-containing substrates such as cellulose can also be con-
(2007d) who examined biofilm culture growing on granulated acti- verted into hydrogen (Call and Logan, 2008).
vated carbon in an anaerobic fluidized-bed reactor. Operated at Initial design of most MEC systems used flat electrode. Flat elec-
HRTs of 0.5–4 h and influent glucose concentrations of 10– trode designs, however, have limited surface area for the exoelec-
30 g l1, a culture pH of 4.0 was found to be most favorable for trogens causing the membranes to increase the ohmic resistances.
hydrogen production. At that operating pH, biofilm sludge concen- Alternative design uses a graphite brush for the exoelectrogen sub-
tration up to 21.5 g VSS l1 was retained. The findings suggested strate (anode) without the need for membrane separator. This de-
superiority of fluidized-bed in biohydrogen production in compar- sign managed to decrease Vapp from 1.0 V using a gas diffusion
ison with the fixed-bed counterpart. membrane and 0.5 V with a Nafion membrane to 0.4 V in the mem-
braneless design. It was reported the efficiency is a function of the
4.2.6. Upflow anaerobic sludge blanket bioreactor lower heating value of the hydrogen divided by the lower heating
Upflow Anaerobic Sludge Blanket (UASB) bioreactor has been value of the organic material plus the electrical energy provided
used in biohydrogen research because of its good treatment effi- (Call and Logan, 2008). The efficiency was raised from 23% with a
ciency and capability in retaining high biomass concentration. It gas diffusion membrane and 53% with a Nafion membrane to
was found that hydrogen yield was HRT-dependent and stabilized 76% in the membraneless reactor. Under these conditions, a hydro-
at 1.5 mol H2 mol1 sucrose at HRT of 8–20 h (Chang and Lin, gen production rate of 3.12 m3 m3 reactor per day can be
2004). The yields decreased drastically at a HRT of 4 or 24 h. At a achieved (Call and Logan, 2008). However, the methane production
HRT of 8 h, the maximum hydrogen production and specific hydro- rates also increased to an average of 3.5% methane in the gaseous
gen production rates were recorded at 0.25 l h1 l1 and product. To control the methanogenesis, strategies involving inter-
53.5 mmol H2 g1 VSS d1, respectively. Biomass retention reached mittent draining and air exposure or in situ air-sparging have been
the maximum level of 7.2 g VSS l1 at a HRT of 24 h, but decreased proposed (Call and Logan, 2008). However, these strategies will
to 5.0 g l1 at the optimum HRT of 8 h (Table 1). likely increase operation and maintenance requirements leading
Anaerobic granular sludge bed bioreactors were supplemented to more costly systems. In addition to methane suppression, con-
with activated carbon carriers and combined with distributors in- tinuous operation, decreasing the pH, operating under carbon lim-
stalled at different locations to investigate the effect of distributor/ ited conditions, increasing the microorganisms tolerance to
carrier on biohydrogen production efficiency (Lo et al., 2009b). The impurities, and examining other feedstocks are all issues to be
results show that plastic net stimulated the substrate/microorgan- addressed.
isms contact and sludge granulation, thereby leading to a much MEC can be incorporated into the second stage of a two-stage
better hydrogen production performance when compared with hybrid bioreactor system which will be discussed in the following
those obtained from traditional CSTR. The highest hydrogen pro- section. In the second stage reaction, electricity applied to a MEC
duction rate (7.89 l h1 l1) and yield (3.42 mol H2 mol1 sucrose) provides the energy needed in converting organic acids to hydro-
were obtained when two pieces of plastic nets were installed at gen (Call and Logan, 2008; Wang et al., 2011). In a MEC, the micro-
both 4 cm and 16 cm from the bottom of bioreactor without carrier bial consortia accomplish dark fermentation in situ, and the
addition operated at a HRT of 0.5 h. Addition of carriers led to sig- resulting fermentation metabolites are used by the electrogenic
nificant improvement on the hydrogen production efficiency at a microorganisms. The electrogenic community oxidizes their sub-
longer HRT (1–4 h) when compared with the carrier-absent strate using the anode as terminal electron acceptor.
system. Achieving satisfactory efficiency of MEC with nominal electrical
A coupled hydrodynamics-reaction kinetics model was formu- inputs, however, is still a major challenge. Hydrogen production
lated from computational fluid dynamics code to simulate a gas–li- rates are still substantially lower than those derived from dark fer-
quid–solid three-phase expanded granular sludge bed (EGSB) mentations at relatively high voltages (800 mV). There remain sub-
reactor for the first time (Wang et al., 2010b). The hydrodynamics stantial challenges to the practical implementation of this
8530 K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533

Table 1
Biohydrogen production performance of some dark fermentation systems.

Process Optimal HRTa (h) HPRb (L/L h) SHPRc (mmol/g-VSS h) Biomass (g-VSS/L) Ref.
Fixed-bed 1 1.32 3.73 15.8 Lee et al. (2006)
Tricking biofilter 4 1.07 1.99 24 Oh et al. (2004a)
AFBR 2 1.82 – – Lin et al. (2009)
AFBR 1 2.36 4.9 21.5 Zhang et al. (2007d)
Packed-bed 0.5 7.41 12.82 25.8 Lee et al. (2006)
Packed-bed 10 0.25 3.70 3.02 Palazzi et al. (2000)
UASB 8 0.28 2.47 5.06 Chang and Lin (2004)
UASB 26.7 0.05 – – Kotsopoulos et al. (2006)
CSTR 6 0.54 1.30 20 Fang et al. (2002)
CSTR 0.5 15.09 19.27 35.84 Wu et al. (2006)
CSTR 0.5 3.20 8.31 16.03 Zhang et al. (2007c)
Column reactor 0.25 7.49 – 35 Zhang et al. (2008c)
a
On the basis of hydrogen production rate.
b
Hydrogen production rate.
c
Specific hydrogen production rate.

technology. These include the replacement of expensive platinum philic dark fermentation. In second stage photobioreactor, acetate
electrode, the high current densities and the reduction of the elec- is converted to hydrogen and carbon dioxide (Nath and Das,
trical input requirement. 2006). The combination could be expected to reach as close to
the theoretical maximum production of 12 mol of hydrogen per
mol glucose equivalent as possible. While the research so far have
4.4. Hybrid bioreactors indicated encouraging findings, the main technical barrier to the
practical application of photo-fermentation lies with the low pho-
A basic principle of the hybrid two-stage bioreactors is that bio- tosynthetic efficiencies with concomitant low hydrogen yields un-
fermentation of substrate to hydrogen and organic acids takes der moderate to high light intensities.
place in a conventional reactor in a first stage of process, and addi- In another approach, MECs are incorporated into the second
tional gaseous energy in the form of methane or hydrogen is ex- stage, in which electricity applied to a microbial fuel cell provides
tracted from in the second stage (Koutrouli et al., 2009). In the energy needed in converting organic acids to hydrogen (Call
optimizing gas production, a different reactor for the second stage and Logan, 2008; Wang et al., 2011). In principle, a second stage
operated under different conditions (such as higher pH and longer MEC could produce 12 H2/glucose with only a small input of
HRT) is applied (Hwang et al., 2011). energy.
An attempt for maximizing the overall energy extraction in the It has been reported that the hybrid two-stage system has al-
second stage is the utilization of photo-fermentation (Kapdan and ready been scaled up to pilot plant stage (Ueno et al., 2007). Such
Kargi, 2006) or fuel cell (Garcia-Pena et al., 2009). This can be a hybrid system offers several merits over conventional methano-
achieved by recovering additional hydrogen from the metabolites genesis, including an effective solubilization of substrates and in-
of the dark fermentation. Non-sulfur photo-synthetic microbes creased tolerance to high OLRs. Successful operation of such
capable of converting organic acids to hydrogen were cultivated systems using actual wastes has also been documented, including
in the system (Redwood et al., 2009). Theoretically, complete con- a pilot-scale fermentation of kitchen waste generating relatively
version of organics into hydrogen via photo-fermentations is pos- high production rates of 5.4 m3 H2 m3 d1 and 6.1 m3 CH4
sible owing to the fact that hydrogen production is driven by m3 d1 while achieving COD removal of 80% (Ueno et al., 2007).
ATP-dependent nitrogenase. The required ATP is formed with the The superior performance of this hybrid system is illustrated from
energy derived from the sunlight. the methane yields which were two times better than a single-
The feasibility of hybrid two-stage photo-fermentation system stage process (Ueno et al., 2007).
has been demonstrated in which the second stage was actually
fed with effluent from a hydrogen-producing reactor, although
hydrogen yields were well below the hypothetical values (Nath 4.5. Multi-stage bioreactors
et al., 2008). On the other hand, a combination of dark and
photo-fermentation in a hybrid two-stage system can improve Multi-stage bioreactors involving three or even four stages
the overall yield of hydrogen (Nath and Das, 2006). The synergy (Fig. 1) have been proposed to maximize the hydrogen production
of the hybrid system lies in the maximal utilization of the substrate from the substrate (US DOE, 2007; Wang et al., 2011). Sunlight is
due to improved thermodynamics. In the first stage, the biomass is first filtered through the first stage direct photolysis reactor where
fermented to acetate, carbon dioxide and hydrogen in a thermo- the visible light is utilized by blue green algae, and the unfiltered

H2 Biomass feed

Water Photo- Dark Microbial


Photolysis Voltage
fermentation fermentation electrolysis supply

Organic acids

Fig. 1. Multi-stage bioreactor system. Adapted from US DOE (2007).


K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533 8531

infrared light is used by photo-synthetic microbes in the second Technical challenges in achieving practical applications of bio-
stage photo-fermentative reactor. The effluent from the second hydrogen include lowering the cost of hydrogen production, deliv-
stage photo-fermentation together with the biomass feedstock is ery, storage, conversion, and end use applications. Economics
fed into a third stage dark fermentation reactor where the bacteria would strongly favor large-scale hydrogen production systems.
convert the substrate into hydrogen and organic acids. As the efflu- The problem with the dark hydrogen fermentation is that in the
ent is enriched with organic acids, a supply of external organic dark, fermentative bacteria produce only relatively small amounts
acids for the photo-fermentative process can be eliminated. The of hydrogen, typically less than 30% stoichiometrically. Economic
forth stage is the use of a MEC to produce hydrogen. The MEC uses feasibility will not be sustainable at these yield level. Currently,
the organic acids generated from the dark fermentation under less than 10% of the algae photosynthetic capacity was utilized
light-independent process. It thus can be operated during the night for biohydrogen production. Research is underway in improving
or low light condition (US DOE, 2007). further biohydrogen production capacity using molecular engi-
Integration of multiple processes produces significant chal- neering approach. With technology advancement, biohydrogen
lenges for the reactor engineering, system design, process control, production may offer a sustainable alternative energy resource in
and operation and maintenance. The major challenges with the future.
coproduction of hydrogen and oxygen from photolytic hydrogen
production include photosynthetic and respiration capacity ratio, 6. Conclusions
co-culture balance, and concentration and processing of cell bio-
mass (Holladay et al., 2009). Extensive research in the past two decades have reviewed
The increasing interest on hythane, a highly efficient and ultra promising prospect of biohydrogen production. There have been
clean burning alternative fuel (a mixture of hydrogen and meth- substantial improvement and development in both the yield and
ane) which is probably the most promising biogas for industrial volumetric production rates of hydrogen fermentations. For realis-
applications, has led many researchers to attempt hydrogen pro- tic applications that make economic sense, hydrogen yields and
duction by anaerobic digestion of biomass in hybrid or multi-stage production rates must at least surpass considerably the present
bioreactors. An experimental work optimizing a two phase ther- achievements. Technological breakthrough must be sought after
mophilic anaerobic digestion process for bio-hythane production to extract most of hydrogen from substrate, if not all. Investigation
treating source collected organic fraction of municipal solid waste addressing this challenge should be one of focuses of future re-
was carried out (Cavinato et al., 2009). An hydrogen rich biogas search. Scaling up of the process to full-scale systems will further
production of 0.06 m3 gas/kgTVS was reached in the fermentative refine the technology for future commercialization.
reactor, showing a 65% CO2 content. The methanogenic phase
shown constant stability parameters (pH value 8, VFA References
255 mg COD/l, ammonia 1150 mg N/l) and a specific gas produc-
tion of 0.73 m3/kgTVS. Adams, M.W., Hall, D.O., 1979. Purification of the membrane-bound hydrogenase of
Escherichia coli. J. Biol. Chem. 183 (1), 11–22.
Akkerman, I., Janssen, M., Rocha, J.M.S., Reith, J.H., Wijffels, R.H., 2003.
Photobiological hydrogen production: photochemical efficiency and
5. Challenges and prospects bioreactor design. In: Reith, J.H., Wijffels, R.H., Barten, H. (Eds.), Biomethane
and Biohydrogen: Status and Perspectives of Biological Methane and Hydrogen
Production. Dutch Biological Hydrogen Foundation, Hague, pp. 124–145.
The relatively low hydrogen yield and production rate are two Bleijlevens, B., Buhrke, T., van der Linden, E., Friedrich, B., Albracht, S.P., 2004. The
common challenges for the biological hydrogen-producing sys- auxiliary protein HypX provides oxygen tolerance to the soluble [NiFe]-
hydrogenase of ralstonia eutropha H16 by way of a cyanide ligand to nickel.
tems. Enhancement in hydrogen yield may be possible by using J. Biol. Chem. 279 (45), 46686–46691.
suitable microbial strain, process modification, efficient bioreactor Buitron, G., Carvajal, C., 2010. Biohydrogen production from Tequila vinasses in an
design and also genetic and molecular engineering technique, to anaerobic sequencing batch reactor: effect of initial substrate concentration,
temperature and hydraulic retention time. Bioresour. Technol. 101 (23), 9071–
redirect metabolic pathway. Some integrated strategies are now 9077.
being developed such as the two-step fermentation process (acido- Call, D., Logan, B.E., 2008. Hydrogen production in a single chamber microbial
genic + photobiological or acidogenic + methanogenic processes) electrolysis cell lacking a membrane. Environ. Sci. Technol. 42, 3401–3406.
Cavinato, C., Bolzonella, D., Eusebi, A.L., Pavan, P., 2009. Bio-hythane production by
or the use of modified microbial fuel cells (de Vrije and Claasen,
thermophilic two-phase anaerobic digestion of organic fraction of municipal
2003; Ueno et al., 2007). Applying genetic and metabolic engineer- solid waste: preliminary results. AIDIC Conference Series, 09, 61-66.
ing techniques to improve the hydrogen yields of those microbial doi:10.3303/ACOS0909008.
cultures with higher hydrogen production rates might also be an- Chang, F.Y., Lin, C.Y., 2004. Biohydrogen production using an up-flow anaerobic
sludge blanket reactor. Int. J. Hydrogen Energy 29 (1), 33–39.
other feasible option (Chittibabu et al., 2006; Nath and Das, 2006). Chittibabu, G., Nath, K., Das, D., 2006. Feasibility studies on the fermentative
Apart from that more easily degradable substance, green wastes hydrogen production by recombinant Escherichia coli BL-21. Process Biochem.
will, most probably, to a large extent be the targeted feedstocks for 41 (3), 682–688.
de Vrije, T., Claasen, P.A.M., 2003. Dark hydrogen fermentation. In: Reith, J.H.,
hydrogen fermentation because of their abundance. Lignocellulose Wijffels, R.H., Barten, H. (Eds.), Bio-methane and Bio-hydrogen. Dutch Biological
is the most abundant renewable natural resource and substrate Hydrogen Foundation, Hague, pp. 103–123.
available for conversion to fuels. There is great potential in the Ding, J., Wang, X., Zhou, X.F., Ren, N.Q., Guo, W.Q., 2010. CFD optimization of
continuous stirred-tank (CSTR) reactor for biohydrogen production. Bioresour.
use of green waste biomass as a renewable source of energy via Technol. 101 (18), 7005–7013.
microbial breakdown to sugars that can then be converted to bio- Evens, T.J., Chapman, D.J., Robbins, R.A., D’Asaro, E.A., 2000. An analytical pat-plate
hydrogen (Lo et al., 2008, 2009a). photobioreactor with a spectrally attenuated light source for the incubation of
phytoplankton under dynamic light regimes. Hydrobiologia 434, 55–62.
There has been a growing resistance against the use of energy Fan, K.S., Kan, N.R., Lay, J.J., 2006. Effect of hydraulic retention time on anaerobic
crops as feedstocks for biofuels generation. This backlash was cen- hydrogenesis in CSTR. Bioresour. Technol. 97 (1), 84–89.
tered on the food-vs-fuel debate, with the main arguments that Fang, H.H.P., Liu, H., 2002. Effect of pH on hydrogen production from glucose by a
mixed culture. Bioresour. Technol. 82 (1), 87–93.
crops that could support human dietary needs are diverted to the
Fang, H.H.P., Liu, H., Zhang, T., 2002. Characterization of a hydrogen-producing
production of biofuels thus inflates food prices. As an answer to granular sludge. Biotechnol. Bioeng. 78 (1), 44–52.
those issues, the production of second generation biofuels is pro- Frey, M., 2002. Hydrogenases: hydrogen-activating enzymes. ChemBiochem 3 (2–
posed, i.e. biofuels produced by feedstocks that are not competitive 3), 153–160.
Garcia-Pena, E.I., Guerrero-Barajas, C., Ramirez, D., Arriaga-Hurtado, L.G., 2009.
to edible crops such as wastes and residues (Juang et al., 2011; Kim Semi-continuous biohydrogen production as an approach to generate
and Lee, 2010; Ntaikou et al., 2010). electricity. Bioresour. Technol. 100 (24), 6369–6377.
8532 K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533

Hallenbeck, P.C., Ghosh, D., 2009. Advances in fermentative biohydrogen Nishihara, H., Miyashita, Y., Aoyama, K., Kodama, T., Igarashi, Y., Takamura, Y., 1997.
production: The way forward? Trends Biotechnol. 27 (5), 287–297. Characterization of an extremely thermophilic and oxygen-stable membrane-
Hankamer, B., Lehr, F., Rupprecht, J., Mussgnug, J.H., Posten, C., Kruse, O., 2007. bound hydrogenase from a marine hydrogen-oxidizing bacterium
Photosynthetic biomass and H2 production by green algae: from bioengineering Hydrogenovibrio marinus. Biochem. Biophy. Res. Comm. 232, 766–770.
to bioreactor scale-up. Phys. Plant 131, 10–21. Ntaikou, I., Antonopoulou, G., Lyberatos, G., 2010. Biohydrogen production from
Holladay, J.D., Hu, J., King, D.L., Wang, Y., 2009. An overview of hydrogen production biomass and wastes via dark fermentation. Waste Biomass Valoriz. 1 (1), 21–39.
technologies. Catal. Today 139, 244–260. Oh, Y.K., Kim, S.H., Kim, M.S., Park, S., 2004a. Thermophilic biohydrogen production
Hussy, I., Hawkes, F.R., Dinsdale, R., Hawkes, D.L., 2003. Continuous fermentative from glucose with trickling biofilter. Biotechnol. Bioeng. 88 (6), 690–698.
hydrogen production from a wheat starch co-product by mixed microflora. Oh, S.E., Lyer, P., Bruns, M.A., Logan, B.E., 2004b. Biological hydrogen production
Biotechnol. Bioeng. 84 (6), 619–626. using a membrane bioreactor. Biotechnol. Bioeng. 87 (1), 119–127.
Hwang, J.H., Choi, J.A., Abou-Shanab, R.A.I., Min, B., Song, H., Kim, Y., Lee, E.S., Jeon, Okamoto, M., Miyahara, T., Mizuno, O., Noike, T., 2000. Biological hydrogen
B.H., 2011. Feasibility of hydrogen production from ripened fruits by a potential of materials characteristic of the organic fraction of municipal solid
combined two-stage (dark/dark) fermentation system. Bioresour. Technol. 102 wastes. Water Sci. Technol. 41 (3), 25–32.
(2), 1051–1058. Palazzi, E., Fabiano, B., Perego, P., 2000. Process development of continuous
Juang, C.P., Whang, L.M., Cheng, H.H., 2011. Evaluation of bioenergy recovery hydrogen production by Enterobacter aerogenes in a packed column reactor.
processes treating organic residues from ethanol fermentation process. Bioproc. Eng. 22 (3), 205–213.
Bioresour. Technol. 102 (9), 5394–5399. Prescott, L.M., Klein, D.A., Harley, J.P., 2002. Microbiology. McGraw-Hill, New York.
Kanai, T., Imanaka, H., Nakajima, A., Uwamori, K., Omori, Y., Fukui, T., Atomi, H., Redwood, M.D., Paterson-Beedle, M., Macaskie, L.E., 2009. Integrating dark and light
Imanaka, T., 2005. Continuous hydrogen production by the hyperthermophilic bio-hydrogen production strategies: towards the hydrogen economy. Rev.
archaeon, Thermococcus kodakaraensis KOD1. J. Biotechnol. 116 (3), 271–282. Environ. Sci. Biotechnol. 8 (2), 149–185.
Kapdan, I.K., Kargi, F., 2006. Bio-hydrogen production from waste materials. Enzym. Ren, N., Wang, A., Cao, G., Xu, J., Gao, L., 2009. Bioconversion of lignocellulosic biomass
Microb. Technol. 38 (5), 569–582. to hydrogen: potential and challenges. Biotechnol. Adv. 27 (6), 1051–1060.
Kim, S.H., Han, S.K., Shin, H.S., 2006. Effect of substrate concentration on hydrogen Shima, S., Pilak, O., Vogt, S., Schick, M., Stagni, M.S., Meyer-Klaucke, W., Warkentin,
production and 16S rDNA-based analysis of the microbial community in a E., Thauer, R.K., Ermler, U., 2008. The crystal structure of [Fe]-hydrogenase
continuous fermenter. Process Biochem. 41 (1), 199–207. reveals the geometry of the active site. Science 321 (5888), 572–575.
Kim, M.S., Lee, D.Y., 2010. Fermentative hydrogen production from tofu-processing Show, K.Y., Zhang, Z.P., Tay, J.H., Liang, T.D., Lee, D.J., Jiang, W.J., 2007. Production of
waste and anaerobic digester sludge using microbial consortium. Bioresour. hydrogen in a granular sludge-based anaerobic continuous stirred tank reactor.
Technol. 101 (1), S48–S52. Int. J. Hydrogen Energy 32 (18), 4744–4753.
Kotsopoulos, A., Zeng, J., Angelidaki, I., 2006. Biohydrogen production in granular Show, K.Y., Zhang, Z.P., Tay, J.H., Liang, T.D., Lee, D.J., Ren, N., Wang, A., 2010. Critical
up-flow anaerobic sludge blanket (UASB) reactors with mixed cultures under assessment of anaerobic processes for continuous biohydrogen production from
hyper-thermophilic temperature (70 °C). Biotechnol. Bioeng. 94 (2), 296–302. organic wastewater. Int. J. Hydrogen Energy 35 (24), 13350–13355.
Koutrouli, E.C., Kalfas, H., Gavala, H.N., Skiadas, I.V., Stamatelatou, K., Lyberatos, G., Tiwari, M.K., Guha, S., Harendranath, C.S., Tripathi, S., 2006. Influence of extrinsic
2009. Hydrogen and methane production through two-stage mesophilic factors on granulation in UASB reactor. Appl. Microbiol. Biotechnol. 71 (2), 145–
anaerobic digestion of olive pulp. Bioresour. Technol. 100 (15), 3718–3723. 154.
Kyazze, G., Martinez-Perez, N., Dinsdale, R., Premier, G.C., Hawkes, F.R., Guwy, A.J., Turner, J., Sverdrup, G., Mann, M.K., Maness, P.C., Kroposki, B., Ghirardi, M., Evans,
Hawkes, D.L., 2006. Influence of substrate concentration on the stability and yield R.J., Blake, D., 2008. Renewable hydrogen production. Int. J. Energy Res 32 (5),
of continuous biohydrogen production. Biotechnol. Bioeng. 93 (5), 971–979. 379–407.
Laurinavichene, T.V., Kosourov, S.N., Ghirardi, M.L., Seibert, M., Tsygankov, A.A., Ueno, Y., Tatara, M., Fukui, H., Makiuchi, T., Goto, M., Sode, K., 2007. Production of
2008. Prolongation of H2 photoproduction by immobilized, sulfur-limited hydrogen and methane from organic solid wastes by phase-separation of
Chlamydomonas reinhardtii cultures. J. Biotechnol. 134, 275–277. anaerobic process. Bioresour. Technol. 98 (9), 1861–1865.
Lee, K.S., Lin, P.J., Chang, J.S., 2006. Temperature effects on biohydrogen production US Department of Energy DOE, 2007. Hydrogen, Fuel Cells and Infrastructure
in a granular sludge bed induced by activated carbon carriers. Int. J. Hydrogen Technologies Program, Multi-Year Research, Development and Demonstration
Energy 31 (4), 465–472. Plan, US Department of Energy.
Lee, H.S., Vermaas, W.F.J., Rittmann, B.E., 2010. Biological hydrogen production: Van Ginkel, S., Logan, B.E., 2005. Inhibition of biohydrogen production by
prospects and challenges. Trend Biotechnol. 28 (5), 262–271. undissociated acetic and butyric acids. Environ. Sci. Technol. 39 (23), 9351–9356.
Li, C.L., Fang, H.H.P., 2007. Fermentative hydrogen production from wastewater and Van Ginkel, S.W., Oh, S.E., Logan, B.E., 2005. Biohydrogen gas production from food
solid wastes by mixed cultures. Crit. Rev. Environ. Sci. Technol. 37 (1), 1–39. processing and domestic wastewaters. Int. J. Hydrogen Energy 30 (15), 1535–
Lin, C.Y., Lay, C.H., 2004a. Effects of carbonate and phosphate concentrations on 1542.
hydrogen production using anaerobic sewage sludge microflora. Int. J. Venkata Mohan, S., 2009. Harnessing of biohydrogen from wastewater treatment
Hydrogen Energy 29 (3), 275–281. using mixed fermentative consortia: process evaluation towards optimization.
Lin, C.Y., Lay, C.H., 2004b. A nutrient formulation for fermentative hydrogen Int. J. Hydrogen Energy 34 (17), 7460–7474.
production using anaerobic sewage sludge microflora. Int. J. Hydrogen Energy Vijaya Bhaskar, Y., Venkata Mohan, S., Sarma, P.N., 2008. Effect of substrate loading
30 (3), 285–292. rate of chemical wastewater on fermentative biohydrogen production in
Lin, C.Y., Lay, C.H., 2004c. Carbon/nitrogen-ratio effect on fermentative biofilm configured sequencing batch reactor. Bioresour. Technol. 99 (15),
hydrogen production by mixed microflora. Int. J. Hydrogen Energy 29 (1), 41–45. 6941–6948.
Lin, C.N., Wu, S.Y., Chang, J.S., Chang, J.S., 2009. Biohydrogen production in a three- Wang, J., Wan, W., 2009. Factors influencing fermentative hydrogen production: a
phase fluidized bed bioreactor using sewage sludge immobilized by ethylene- review. Int. J. Hydrogen Energy 34 (2), 799–811.
vinyl acetate copolymer. Bioresour. Technol. 100 (13), 3298–3301. Wang, Y.Z., Liao, Q., Zhu, X., Tian, X., Zhang, C., 2010a. Characteristics of hydrogen
Lo, Y.C., Bai, M.D., Chen, W.M., Chang, J.S., 2008. Cellulosic hydrogen production production and substrate consumption of Rhodopseudomonas palustris CQK01 in
with a sequencing bacterial hydrolysis and dark fermentation strategy. an immobilized-cell photobioreactor. Bioresour. Technol. 101 (11), 4034–4041.
Bioresour. Technol. 99 (17), 8299–8303. Wang, X., Ding, J., Guo, W.Q., Ren, N.Q., 2010b. A hydrodynamics-reaction kinetics
Lo, Y.C., Su, Y.C., Chen, C.Y., Chen, W.M., Lee, K.S., Chang, J.S., 2009a. Biohydrogen coupled model for evaluating bioreactors derived from CFD simulation.
production from cellulosic hydrolysate produced via temperature-shift- Bioresour. Technol. 101 (24), 9749–9757.
enhanced bacterial cellulose hydrolysis. Bioresour. Technol. 100 (23), 5802– Wang, A., Sun, D., Cao, G., Wang, H., Ren, N.Q., Wu, W.M., Logan, B.E., 2011.
5807. Integrated hydrogen production process from cellulose by combining dark
Lo, Y.C., Lee, K.S., Lin, P.J., Chang, J.S., 2009b. Bioreactors configured with distributors fermentation, microbial fuel cells, and a microbial electrolysis cell. Bioresour.
and carriers enhance the performance of continuous dark hydrogen Technol. 102 (5), 4137–4143.
fermentation. Bioresour. Technol. 100 (19), 4381–4387. Wu, S.Y., Hung, C.H., Lin, C.N., Chen, H.W., Lee, A.S., Chang, J.S., 2006. Fermentative
Luo, G., Talebnia, F., Karakashev, D., Xie, L., Zhou, Q., Angelidaki, I., 2011. Enhanced hydrogen production and bacterial community structure in high-rate anaerobic
bioenergy recovery from rapeseed plant in a biorefinery concept. Bioresour. bioreactors containing silicone-immobilized and self-flocculated sludge.
Technol. 102 (2), 1433–1439. Biotechnol. Bioeng. 93 (5), 934–946.
Luo, G., Xie, L., Zou, Z.H., Wang, W., Zhou, Q., 2010. Evaluation of pretreatment Yokoi, H., Saitsu, A., Uchida, H., Hirose, J., Hayashi, S., Takasaki, Y., 2001. Microbial
methods on mixed inoculum for both batch and continuous thermophilic hydrogen production from sweet potato starch residue. J. Biosci. Bioeng. 91 (1),
biohydrogen production from cassava stillage. Bioresour. Technol. 101 (3), 959– 58–63.
964. Younesi, H., Najafpour, G., Ismail, K.S.K., Mohamed, A.R., Kamaruddin, A.H., 2008.
Minnan, L., Jinli, H., Xiaobin, W., Huijuan, X., Jinzao, C., Chuannan, L., Fengzhang, Z., Biohydrogen production in a continuous stirred tank bioreactor from synthesis
Liangshu, X., 2005. Isolation and characterization of a high H2-producing strain gas by anaerobic photosynthetic bacterium: Rhodopirillum rubrum. Bioresour.
Klebsiella oxytoca HP1 from a hot spring. Res. Microbiol. 156 (1), 76–81. Technol. 99 (7), 2612–2619.
Mizuno, O., Ohara, T., Shinya, M., Noike, T., 2000. Characteristics of hydrogen Zhang, Z.P., Show, K.Y., Tay, J.H., Liang, D.T., Lee, D.J., Jiang, W.J., 2006a. Effect of
production from bean curd manufacturing waste by anaerobic microflora. hydraulic retention time on biohydrogen production and anaerobic microbial
Water Sci. Technol. 42 (3–4), 345–350. community. Process Biochem. 41 (10), 2118–2123.
Nath, K., Das, D., 2006. Amelioration of biohydrogen production by a two-stage Zhang, H.S., Bruns, M.A., Logan, B.E., 2006b. Biological hydrogen production by
fermentation process. Ind. Biotechnol. 2, 44–47. Clostridium acetobutylicum in an unsaturated flow reactor. Water Res. 40 (4),
Nath, K., Muthukumar, M., Kumar, A., Das, D., 2008. Kinetics of two-stage 728–734.
fermentation process for the production of hydrogen. Int. J. Hydrogen Energy Zhang, Z.P., Show, K.Y., Tay, J.H., Liang, D.T., Lee, D.J., 2007a. Biohydrogen
33 (4), 1195–1203. production with anaerobic fluidized bed reactors – a comparison of
K.-Y. Show et al. / Bioresource Technology 102 (2011) 8524–8533 8533

biofilm-based and granule-based systems. Int. J. Hydrogen Energy 33 (5), 1559– Zhang, Z.P., Adav, S.S., Show, K.Y., Tay, J.H., Liang, D.T., Lee, D.J., 2008a.
1564. Characteristics of rapidly formed hydrogen-producing granules and biofilms.
Zhang, M.L., Fan, Y.T., Xing, Y., Pan, C.M., Zhang, G.S., Lay, J.J., 2007b. Enhanced Biotechnol. Bioeng. 101 (5), 926–936.
biohydrogen production from cornstalk wastes with acidification pretreatment Zhang, Z.P., Show, K.Y., Tay, J.H., Liang, T.D., Lee, D.J., 2008b. Enhanced continuous
by mixed anaerobic cultures. Biomass Bioenergy 31 (4), 250–254. biohydrogen production by immobilized anaerobic microflora. Energy Fuels 22
Zhang, Z.P., Show, K.Y., Tay, J.H., Liang, D.T., Lee, D.J., Jiang, W.J., 2007c. Rapid (1), 87–92.
formation of hydrogen-producing granules in an anaerobic continuous stirred Zhang, Z.P., Show, K.Y., Tay, J.H., Liang, T.D., Lee, D.J., Wang, J.Y., 2008c. The role of
tank reactor induced by acid incubation. Biotechnol. Bioeng. 96 (6), 1040–1050. acid incubation in rapid immobilization of hydrogen-producing culture in
Zhang, Z.P., Tay, J.H., Show, K.Y., Yan, R., Liang, D.T., Lee, D.J., Jiang, W.J., 2007d. anaerobic upflow column reactors. Int. J. Hydrogen Energy 33 (19), 5151–5160.
Biohydrogen production in a granular activated carbon anaerobic fluidized bed
reactor. Int. J. Hydrogen Energy 32 (2), 185–191.

You might also like