review paper on biohydrogen

Upload: coolgk2

Post on 06-Jul-2018

221 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/17/2019 Review Paper on Biohydrogen

    1/22

    Review

    A critical literature review on biohydrogen production by pure

    cultures

    Omneya Elsharnouby a, Hisham Hafez b,*, George Nakhla a,c, M. Hesham El Naggar a

    a Department of Civil and Environmental Engineering, The University of Western Ontario, London, Ontario N6A 5B9, CanadabGreenField Ethanol Inc., Chatham, Ontario N7M 5J4, Canadac

    Department of Chemical and Biochemical Engineering, The University of Western Ontario, London, Ontario N6A 5B9, Canada

    a r t i c l e i n f o

    Article history:

    Received 19 November 2012

    Received in revised form

    21 January 2013

    Accepted 7 February 2013

    Available online 13 March 2013

    Keywords:

    Biohydrogen

    Fermentation

    Pure cultures

    Mesophilic

    Thermophilic

    Anaerobic

    a b s t r a c t

    Global research is moving forward in developing hydrogen as a renewable energy source in

    order to alleviate concerns related to carbon dioxide emissions and depleting fossil fuels

    resources. Biohydrogen has the potential to replace current hydrogen production tech-

    nologies relying heavily on fossil fuels. Batch and continuous systems employing pure

    mesophiles and thermophiles isolates and co-cultures of isolates have been investigated.

    The co-cultures of the isolates achieved better results than mono-cultures of the isolates

    with respect to different parameters. This paper presents a critical review of the literature

    reporting on fermentative biohydrogen production by pure cultures of bacteria in different

    systems. Synergies between different types of bacteria, i.e. strict and facultative, and a

    comparison between mono- and co-cultures, types of feedstocks, and preferred feedstocks

    for mono- and cultures are outlined.

    Copyright  ª  2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights

    reserved.

    1. Introduction

    The challenges of environmental pollution and traditional

    energy reserves depletion are focussing intensive research on

    alternative energy production. Hydrogen is widely regarded as

    one of the most promising energy carriers, because of its high

    efficiency of conversion to usable power, non-polluting 

    oxidation products, and high gravimetric energy   [1]. These

    advantages render hydrogen as an attractive candidate to

    reduce reliance on conventional fossil fuels.

    Biological hydrogen production is suitable for a variety of 

    feedstocks including organic waste material, and is less

    energy intensive compared to other hydrogen processes.

    Biological methods include photosynthetic hydrogen pro-

    duction and dark fermentative hydrogen production.

    Photosynthetic hydrogen production involves transforming 

    solar energy into hydrogen via photosynthetic bacteria.

    However, its application is challenged by the low transfer

    efficiency of light, complexity in reactors design, and low

    hydrogen production rates   [2,3]. On the other hand,

    fermentative hydrogen production facilitates high hydrogen

    production rate through a simple operation, making the

    process an increasingly popular option for hydrogen

    production.

    *   Corresponding author. Tel.:  þ1 519 784 6230; fax:  þ1 519 352 9559.E-mail addresses:   [email protected]   (O. Elsharnouby),   [email protected],   [email protected]   (H. Hafez),

    [email protected] (G. Nakhla), [email protected] (M.H. El Naggar).

     Available online at www.sciencedirect.com

    j o u r n a l h o m e p a g e :   w w w . e l s e v i er . c o m / l o c a t e / he

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 6

    0360-3199/$  e  see front matter Copyright  ª  2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

    mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]://www.sciencedirect.com/science/journal/03603199http://www.elsevier.com/locate/hehttp://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://www.elsevier.com/locate/hehttp://www.sciencedirect.com/science/journal/03603199mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]

  • 8/17/2019 Review Paper on Biohydrogen

    2/22

    Several factors influence the fermentative hydrogen pro-

    duction process, which have to be optimized for enhanced

    performance. Chief among these factors are: inoculum, sub-

    strate, reactor type, and temperature, which seem to impact

    both hydrogen yield and hydrogen production rate, albeit with

    varying importance [4]. Hydrogen yield is significantly influ-

    enced by the inoculum type, as the fermentation end products

    are influenced by the bacterial metabolism. The inoculumused for fermentative hydrogen production include: mixed

    communities of anaerobic bacteria obtained from anaerobic

    sludge digesters [5,6], compost piles [7], and pure cultures of 

    known species of hydrogen-producing bacteria. In pure cul-

    ture systems, metabolic shifts are more easily detected due to

    the reduced diversity of the biomass. Moreover, studies

    employing pure cultures can reveal important information

    regarding conditions that promote high hydrogen yield and

    production rate [8].

    Numerous pure bacterial cultures have been used in recent

    studies to produce hydrogen from various substrates. Never-

    theless, only a few review papers with limitedscope are found

    in the literature addressing fermentative hydrogen produc-tion by pure cultures [9,10]. For example, the review article [9]

    merely presented data on the challenges and prospective of 

    biohydrogen production by pure cultures, without any critical

    analysis, and provided minimal insight on the potential

    applications.

    In this paper, a critical review of 195 studies employing 

    pure cultures was conducted considering the most important

    parameters [1e103]. The relative effectiveness of co-cultures

    of pure isolates and mono-cultures of these isolates is dis-

    cussed. In addition, comparative studies between employing 

    thermophilic and mesophilic cultures, batch and continuous

    systems, and the different types of feedstocks, are evaluated.

    Table 1   summarizes the data of considered studies withrespect to operational and performance parameters. Sixteen

    different types of pure cultures were employed in fermenta-

    tive hydrogen production processes solely or co-cultured in

    the studies listed in Table 1. It should be noted that for ease of 

    comparison, the hydrogen production rate and hydrogen yield

    from these studies were normalized to L H2 /L/day and mol H2 /

    mol hexose equivalent, respectively.

    2. Effect of synergies between co-cultures ontechnical and economic efficiencies

    Ten independent studies considered in this review havecompared the effectiveness of co-cultures of pure isolates

    with their mono-cultures in fermentative hydrogen produc-

    tion.  Table 2  summarizes the operational and performance

    parameters of the mono-cultures and co-cultures studied. In

    allten studies, the co-cultures achievedbetterresults than the

    mono-cultures with respect to different parameters. Exam-

    ining the available literature, it is evident that the motivation

    behind employing co-cultures, rather than mono-cultures,

    was either economical or technical. From economy view

    point, co-cultures can help maintain anaerobic conditions for

    strict high hydrogen producers and eliminate the need for an

    expensive reducing agent. From the technical view point, co-

    cultures can improve the hydrolysis of complex sugars and

    plant biomass, and can provide a wider range of pH for bac-

    teria to fermenthydrogen.In moststudies, both economic and

    technical reasons were important considerations and inter-

    dependant. It was also found that there are primarily three

    different types of co-culture of pure isolates. An explanation

    of the synergistic effect in the co-culture processes for each

    type is provided below.

    The first type of co-cultures involves strict and facultativeanaerobes. Obligate anaerobes are extremely sensitive to O2,

    and their H2-producing abilities are inhibited by a slight

    amount of O2, which requires the addition of a reducing agent

    such as   L-cysteine to stabilize H2   production. In order to

    eliminate the cost of the expensive reducing agent, facultative

    anaerobes are used to consume O2 in a medium, so anaerobic

    conditions are readily attained without the need for a

    reducing agent. Therefore, a strict anaerobe such as   Clos-

    tridium sp., and a facultative anaerobe such as  Enterobacter sp.

    are co-cultured in the same reactor under optimum culture

    conditions for H2 production, to achieve stable and high-yield

    H2 production without a reducing agent.

     Jenni et al.   [11]   investigated the applicability of mixedculture of Clostridium butyricum and Escherichia coli for stable H2production without any reducing agents. They utilized

    glucose as substrate, at a temperature of 37   C, and an opti-

    mum pH of 6.5 in a batch reactor. The authors noted that the

    gas production pattern of batch fermentations by E. coli and C.

    butyricum   differed, i.e.   E. coli  continued gas production long 

    after the exponential growth. Mono-cultures of  E. coli and  C.

    butyricum   achieved H2   yields of 1.45 mol-H2 /mol-glucose

    consumed, and 2.09 mol- H2 /mol-glucose consumed, while

    the co-cultures achieved 1.65 mol-H2 /mol-glucose consumed.

    It should be emphesized, however, that even though   C.

    butyricum achieved a higher molar hydrogen yield i.e. specific

    hydrogen production, the co-culture facilitated a greaterglucose conversion efficiency resulting in a higher overall

    volumetric hydrogen production. These findings indicated

    that employing co-cultures was more economical by elimi-

    nating the expensive reducing agent, and technically more

    effective with higher hydrogen production.

    Haruhiko et al.   [12]   examined the O2   tolerance and H2-

    producing stability of the mono- and mixed cultures of   C.

    butyricum and  Enterobacter aerogenes  in batch and continuous

    flowstudies using starch as a substrate at temperature of 37 C

    and pHs of 6.5 and 5.5, respectively. In the batch study,   E.

    aerogenes   hardly produced H2   from starch since it had no

    ability to utilize starch. H2 production by C. butyricum without

    a reducing agent occurred after a long lag time of 12 h. In caseof  C. butyricum with 0.1% L-cysteine as reducing agent, H2 was

    evolved after a short lag time of 5 h. On the other hand, H 2production by the mixed culture of  C. butyricum  and  E. aero-

    genes occurred after even a shorter lag time of less than 2 h,

    and the amount of H2 evolved was the largest at 175% of that

    produced by C. butyricum with a reducing agent. This confirms

    that the mixed culture could produce H2 without a reducing 

    agent, since   E. aerogenes   consumed O2   rapidly, and thus

    maintaining the anaerobic conditions conducive for bio-

    hydrogen production. In the continuous-flow study, the case

    of  C. butyricum with a reducing agent, exhibited H2 production

    after 15 h due to removal of O2 in the reactor by the reducing 

    agent. The hydrogen production by  C. butyricum  without the

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 64946

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    3/22

    Table 1 e Operational and performance parameters of the reviewed 194 experiments.

    Culture(s) Reactortype

    TC

    Substratetype

    Substrateconcentration

    (g/L)

    pH Hydrogen yie(mol H2 /mol glucose

    equivalent)

    1-Caldicellulosiruptor saccharolyticus

    Batch 70 Pretreated wheat

    straw

    10 7.2 3.8

    Batch 70 Pretreated barelystraw 20l

    7 ND

    Batch 70 Glucose 20 7 3.4

    Batch 70 Carrot pulp

    hydrolysate

    10 7 2.8

    Batch 72 Glucose 31 7 2.8

    Batch 72 Glucose 10 7 3.4

    Batch 72 PSPa 10 7 3.5

    Batch 72 PSP-H2b 10 7 3.4

    Batch 72 GXSc 10 6.8m 3.2

    Batch 72 SSBd 20l 6.8m 2.8

    Batch 70 Sucrose 10 7m 2.96

    Batch 72 Glu/Xyle 10 7 3.4

    Batch 72 Glu/Xyle 14 7 3.3

    Batch 72 Glu/Xyle 28 7 2.4

    Batch 72 Miscanthushydrolysate

    10 7 3.4

    Batch 72 Miscanthus

    hydrolysate

    14 7 3.3

    Batch 72 Miscanthus

    hydrolysate

    28 7 2.4

    2-Thermotoga neapolitana

    DSM 4359   CSABR 75 Xylose 5 7 3.36

    DSM 4359   Batch 75 Xylose 5 7.5 1.31

    Batch 75 Glucose 27 7 3

    Batch 75 Glucose 10 7 2.9

    Batch 75 PSP-H2b 10 7 3.3

    Batch 75 PSPa 10 7 3.8

    Batch 75 Glucose 10 7 3.5

    Batch 75 Glucose 20 7 3.4

    Batch 75 Fructose 10 7 3.4 Batch 75 fructose 20 7 3.2

    Batch 75 Glucose  þ

    Fructose

    10 7 3.3

    Batch 75 Glucose  þ

    Fructose

    20 7 3

    Batch 75 Carrot pulp

    hydrolysate

    10 7 2.7

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    4/22

    Table 1 e ( continued )

    Culture(s) Reactortype

    TC

    Substratetype

    Substrateconcentration

    (g/L)

    pH Hydrogen yie(mol H2 /mol glucose

    equivalent)

    Batch 75 Carrot pulp

    hydrolysate

    20 7 2.4

    DSM 4359   Fed

    batch-CSABR

    75 Xylose 5 7.5m 2.66

    Fed batch-

    CSABR

    75 Glucose 5 7.5m 3.2

    Fed batch-

    CSABR

    75 Sucrose 5 7.5m 2.5

    Batch 85 Glucose 2.5 7.5 3.75

    Batch 77 Glucose 2.5 7.5 3.85

    Batch 70 Glucose 2.5 7.5 3.18

    Batch 65 Glucose 2.5 7.5 3.09

    Batch 60 Glucose 2.5 7.5 2.04

    Batch 80 Glucose 7.5 7.5 1.84

    Batch 80 Glu/Xyle 10 7 3.3

    Batch 80 Glu/Xyle 14 7 3.2

    Batch 80 Glu/Xyle 28 7 2.5

    Batch 80 Miscanthushydrolysate

    10 7 2.9

    Batch 80 Miscanthus

    hydrolysate

    14 7 3.2

    Batch 80 Miscanthus

    hydrolysate

    28 7 2

    Batch-

    Serum

    bottels

    80 Glucose 5 7.5 3.85

    3-Clostridium DMHC-10

    Batch 37 Glucose 10 5 3.35

    4-Enterobacter Cloacae

    IIT-BT08   Batch 36 Sucrose 10 6 3.014

    Batch 36 Glucose 10 6 2.2

    Batch 36 Cellobiose 10 6 1.42

    F.P01   Batch 36 Maltose 10 5 0.729 DM11   Batch 37 Glucose 10 6.5 3.31

    Batch 37 Glucose 10 6.5 2.2

    Batch 37 Glucose 10 6.5 3.1

    5-Clostridium beijerinckii

    L9   Batch 35 Glucose 3 7.2 2.81

    Fanp 3   Batch 35 Glucose 10 6.47e6.98 2.52

    AM21B   Batch 36 Glucose 10 6.5 2

    AM21B   Batch 36 Starch 10 6.5 1.8

    RZF-1108   Batch 35 Glucose 9 7 1.97

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    5/22

    Batch 30 Glucose 2.34 6.3 ND

    Batch 36 Starch 10 6.8 ND

    Batch 41 Starch 10 6.8 ND

    Batch 36 Starch 10 6m ND

    Batch 36 Starch 10 7m ND

    L9   Batch 35 Glucose 6 6.4 1.72

    RZF-1108   Batch 35 Glucose 10 6.5 1.96

    6-Clostridium butyricum

    ATCC19398   Batch 36 Glucose 3 7.2 2.29

    CGS5   Batch 37 Sucrose 17.8 5.5 1.39 W5   Batch 39 Glucose 10 6.5 0.81

    EB6   Batch 37 POMEf  ND 5.5 0.22

    EB6   Batch 37 Glucose 10 6 0.6

    Batch 37 SCB hemicellulose

    hydrolysateh20k 5.5 1.73g 

    Batch 37 Glucose (and 200e

    400 mg/l phenol))

    5 6.5 1.46

    CGS5   Batch 37 Xylan hydrolysate 14.2 7.5 0.84

    CGS5   Batch 37 Pretreated straw

    hydrolysate

    9.2 7.5 0.91

    CWBI1009   Batch 30 Glucose 5 5.2m 1.7

    EB6   Batch 37 Glucose 15.7 5.6 2.2

    TISTR 1032   Batch 37 Sugarcane juice 22.3

    (sucrose)

    6.5 1.33

    TISTR 1032   Batch 37 Sucrose 22.3 6.5 1.34 TISTR 1032

    (immobilized)

    Repeated

    batch

    37 Sugarcane juice 22.3

    (sucrose)

    6.5 1.52

    CGS5   Batch 37 Chlorella vulgaris ESP6

    (microalgal hydrolysate)

    9 5.5m ND

    W5   Batch 37 Molass 100 7 1.63

    TM-9A   Batch 37 Glucose 10 8 3.1

    TM-9A   Batch 37 Arabinose 10 8 0.06

    TM-9A   Batch 37 Raffinose 10 8 2.7

    TM-9A   Batch 37 Sucrose 10 8 1.49

    TM-9A   Batch 37 Trehalose 10 8 1.61

    TM-9A   Batch 37 Xylose 10 8 0.59

    TM-9A   Batch 37 Cellobiose 10 8 0.94

    TM-9A   Batch 37 Cellulose 10 8 0.06

    TM-9A   Batch 37 Ribose 10 8 0.84

    TM-9A   Batch 37 Galactose 10 8 0.86

    TM-9A   Batch 37 Fractose 10 8 0.84

    TM-9A   Batch 37 Mannose 10 8 0.67

    W5   Batch 39 Molasses 100 6.5 1.85

    Batch 37 Glucose 3 6.5 2.09

    C. butyricum

    and Escherichia

    coli

    Batch 37 Glucose 3 6.5 1.65

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    6/22

    Table 1 e ( continued )

    Culture(s) Reactortype

    TC

    Substratetype

    Substrateconcentration

    (g/L)

    pH Hydrogen yie(mol H2 /mol glucose

    equivalent)

    C. butyricum and

    Enterobacter

    aerogenes HO-39

    Repeated

    batch

    37 Sw eet potato

    starch residue

    ND 5.25m 2.7

    7-Thermoanaerobactermathranii A3N

    A3N   Batch 70 Sucrose 10 8 2.69

    A3N   Batch 70 Glucose 10 8 2.64

    A3N   Batch 70 Xylose 5 8 2.5

    A3N   Batch 70 Starch 5 8 ND

    8-Thermoanaerobacterium thermosaccharolyticum

    W16   Batch 60 Xylose 10 6.5 2.62

    PSU-2   Batch 60 Sucrose 20 6.25 2.53

    W16   Batch 60 Xylose  þ

    Glucose

    10 6.5 2.45

    W16   Batch 60 Glucose 10 6.5 2.42

    W16   Batch 60 Hydrolysed

    corn stover

    ND 7 2.24g 

    W16   Batch 60 Glucose 5 6.7 2.07

    W16   Batch 60 Glucose 10 7 ND W16   Batch 60 Xylose 10 7 ND

    W16   Batch 60 Glucose  þ

    Xylose  þ

    Arabinose

    7.5,2.2,0.3 7 ND

    W16   Batch 60 Hydrolysed

    corn stover

    ND 7 ND

    PSU-2   Cont. UAi 60 Sucrose 20 5.5 1.11

    PSU-2   Cont.UASB j 60 Sucrose 20 5.5 1.77

    W16   Batch 60 Xylose 12.24 6.8 2.84

    Thermoanaerobacterm

    thermosaccharolyticum

    GD17 and C. thermocellum JN4

    Batch 60 Cellulose 5 4.4 1.8

    9-Ethanoligenens harbinese

    YUAN-3   Batch 35 Glucose 10 5 1.91

    YUAN-3   CSTR 35 Glucose 10 5 1.93 B49   Batch 37 Glucose 9 7 1.83

    B49   Batch 37 Glucose 12 7 1.71

    B49   Batch 37 Glucose 6 7 1.36

    B49   Batch 35 Glucose 10 6 1.67

    B49   Batch 37 Glucose 10 7 2.2

    B49   Batch 35 Glucose 14.5 6 2.2

    B49   Batch 35 Glucose 10 6 2.26

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    7/22

    Ethanoligenens

    harbinese B49

    and C. acetobutylicum X9

    Batch 37 Microcrystalline

    cellulose

    10 5 1.32

    10-Klebsiella pneumoniae

    ECU-15   Batch 37 Glucose 10 6 2.07

    DSM2026   Batch 37 Glycerol 20 6.5 0.53

    11-Pantoea agglomerans

    Batch 37 Glucose 10 7.2 3.8

    Batch 37 Glucose 20 7.2 4.2

    Batch 37 Glucose 20(saline

    conditions)

    7.2 3.3

    12-Clostridium tyrobutyricum

     JM1   Batch 37 Glucose 20 6.3 3.24

     JM1   CSTR 37 Glucose 5 6.7 1.81

    FYa102   Batch 35 Glucosen 3 7.2 1.47

    FYa102   CSTR 35 Glucoseo 12 6 1.06

    FYa102   CSTR 35 Glucosep 12 6 1.42

    ATCC 25755   Fed batch 37 Glucose 50 5.7 2.33

    13-Clostridium acetobutylicum

    M121   Batch 37 Glucose 3 7.2 1.8

    X9   Batch 37 Microcrystalline

    cellulose

    10 5 0.59

    ATCC 824   Cont.

    Trickling bed reactor

    30 Glucose 10 6.2 0.9

    ATCC 824   Batch 36 Cassava

    wastewater

    5k 7 2.41

    14-Escherichia coli

    Batch 37 Glucose 3 6.5 1.45

    S3   Batch 30 Glucose 5 6.8 0.84

    S6   Batch 30 Glucose 5 6.8 0.49

    WDHL   Batch 37 Glucose 15 6 0.3

    WDHL   Batch 37 Galactose 15 6 1.12

    WDHL   Batch 37 Lactose 15 6 1.02

    WDHL   Batch 37 Glucose  þ

    Galactose

    7.5, 7.5 6 1.02

    DJT135   Batch 35 Arabinose 10 6.5 1.2

    DJT135   Batch 35 Fractose 10 6.5 1.27

    DJT135   Batch 35 Galactose 10 6.5 0.69

    DJT135   Batch 35 Glucose 10 6.5 1.51

    DJT135   Batch 35 Lactose 10 6.5 0.37

    DJT135   Batch 35 Maltose 10 6.5 0.2

    DJT135   Batch 35 Mannitol 10 6.5 0.88

    DJT135   Batch 35 Sorbitol 10 6.5 1.36

    DJT135   Batch 35 Sucrose 10 6.5 0.35

    DJT135   Batch 35 Terhalose 10 6.5 0.52

    DJT135   Batch 35 Xylose 10 6.5 0.68

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    8/22

    Table 1 e ( continued )

    Culture(s) Reactortype

    TC

    Substratetype

    Substrateconcentration

    (g/L)

    pH Hydrogen yie(mol H2 /mol glucose

    equivalent)

    SH5   Fed-Batch 37 Soduim

    formate

    16.1 6.5 ND

    15-Clostridium thermocellum

     JN4   Batch 60 Cellulose 5 4.4 0.8

    7072   Batch 55 Cellulose 5 7.4 1.2

    7072   CSTR, 100 L Cron stalk 30 7.4 0.45

    7072   CSTR, 10 L 55 Cron stalk 30 7.4 0.43

    7072   Batch 55 Cron stalk 5 7.4 ND

    ATCC 27405   CSTR 60 Cellulose 4 7 1.29

    ATCC 27405   CSTR 60 Cellulose 3 7 1.53

    ATCC 27405   CSTR 60 Cellulose 2 7 1.65

    ATCC 27405   CSTR 60 Cellulose 1.5 7 0.98

    ATCC27405   Batch 55 Delignefied

    wood fibres

    0.1 6.5 1.6

    ATCC 27405   Batch 60 Cellulose 1 6.8 1.9

    C. thermocellum

    and C. thermosaccharolyticm

    Batch 55 Corn stalk

    waste

    10 7.2 ND

    C. thermocellum and

    C.thermosaccharolyticm

    CSTR 55 Corn stalk

    waste

    10 7.2 ND

    C. thermocellum DSM1237

    and C.thermopalmarium

    DSM 5974

    Batch 55 Cellulose 9 7 1.36

    16- Enterobacter aerogenes

    HO-39   Batch 38 Glucose 10 6.5 1

    HO-39   Fed batch 37 Glucose 10 6.5 0.8

    HO-39   Batch 38 Maltose 10 6.5 0.7

    ATCC29007   Batch 38 Glucose 21.25 6.13 ND

    Batch 38 Glucose 0.2 7 ND

    Batch 37 Glycerol 20 7 0.2

    Batch 37 Glucose 10 5.8 0.89

    E 82005   Batch 38 Glucose 10 5.8 1

    IAM 1183   Batch 37 Xylose 5 6.3 2.64

    IAM 1183   Batch 37 Galactose 10 6.3 2.82 IAM 1183   Batch 37 Mannose 25 6.3 0.96

    IAM 1183   Batch 37 Rahamnose 5 6.3 0.48

    IAM 1183   Batch 37 Arabinose 10 6.3 1.56

    ATCC35029   Batch 37 Glycorel 21 ND 1.22

    NCIMB 10102   Continuous

    packed col.

    40 Corm starch

    hydrolysate

    ND 5.5 2.55

    NCIMB 10102   Batch 40 Starch

    hydrolysate

    20 6.5 1.09g 

    W23   Batch 35 Glucose 5 6.5 1.87

    http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    9/22

    Enterobacter aerogenes

    W23 and Candida

    maltosa HY-35

    Batch 35 Glucose 5 6.5 2.19

    Enterobacter aerogenes

    and C. butyricum

    Batch 36 Starch ND 6.5 2

    Enterobacter aerogenes

    and C. butyricum

    (immobiolized)

    Repeated

    batch

    36 Starch ND 5.5 2.6

    ND: Not defined.a Untreated potato steam peels, Molar yields were based on the amount of starch in untreated PSP assuming 100% starch consumption.

    b The starch in the PSP was liquefied with alpha-amylase, and then the liquefied starch was further hydrolysed to glucose by amyloglucosidase.

    c Mixture of glucose, xylose and sucrose.

    d Sweet sorgham bagasse.

    e Glucose: Xylose  ¼  7:3.

    f Palm oil mill effluent.

    g Mol H2 /mol total sugar.

    h Sugarcane bagasse hemicellulose hydrolysate.

    i Up flow carrier free anaerobic system.

     j Up flow anaerobic sludge blanket.

    k g COD/L.

    l g sugars/L.

    m Controlled pH.

    n 2 g/L peptone was added with the substrate.

    o 8 g/L peptone was added with the substrate.p 0.36 g/L, 1.4 g/L peptone and ammonium chloride respectively, were added with the substrate.

  • 8/17/2019 Review Paper on Biohydrogen

    10/22

    reducing agent was inhibited by oxygen, and could not be

    recovered at all even after 50 h, indicating complete damage

    for the bacterial cells. In contrast, H2 production by the mixed

    culture, which was initially inhibited by oxygen, was recov-

    ered immediately within 0.5 h. These results confirm that the

    mixed culture can remove O2   in the reactor and recover H2production immediately.

    Similarly, Yokoi et al.   [13]  investigated the synergies be-tween strict and facultative anaerobes. A sustained high bio-

    hydrogen yield of 2.7 mol H2 /mol glucose was attained by a

    mixed culture of   C. butyricum   and   E. aerogenes HO-39.   The

    mixed bacteria utilized starch waste consisting of sweet po-

    tato starch residue asa carbon source, and corn steepliquor as

    a nitrogen source. The experiment was conducted in a fed

    batchculture, at a temperature of 37 C, and a controlled pH of 

    5.25. The results proved that a mixed culture of  C. butyricum

    and E. aerogenes could produce hydrogen from starch at a high

    yield of more than 2 mol of hydrogen per 1 mol of glucose

    without any reducing agents, since  E. aerogenes, a facultative

    anaerobe, removed oxygen and generated anaerobic condi-

    tions in the reactor.

    The second type of co-cultures reported in the literature

    was between cellulose degrading anaerobes and high

    hydrogen producers via fermenting simple sugars. The most

    common dark fermentation procedure employed to generate

    hydrogen from cellulose materials involved expensive pre-

    treatment processes, such as delignification, and hydrolysis

    [14]. Since pre-treatment processes are expensive, fermenta-

    tive hydrogen production from cellulosic materials is desir-able. Therefore, many studies investigated employing co-

    cultures of two bacterial strains: one with the capability of 

    hydrolysing cellulose, and the other is a high hydrogen pro-

    ducer utilizing simple sugars.

    For example, Yan et al. [15] investigated biohydrogen pro-

    duction from cellulose using the thermophilic anaerobic

    bacterium Clostridium thermocellum JN4, and the co-cultures of 

    the aforementioned bacterium with   Thermoanaerobacterium

    thermosaccharolyticum GD17. The   C. thermocellum JN4   can

    decompose cellulose but cannot completely utilize the cello-

    biose and glucose produced by the cellulose degradation,

    while the   T. thermosaccharolyticum GD17   can utilize mono-

    sugars. The experiment was conducted at pH of 4.4, and

    Table 2 e Operational and performance parameters for studies employing mono and co-cultures.

    Culture(s) Reactortype

    TC

    Substratetype

    Substrateconcentration

    (g/L)

    pH Hydrogenyield (mol H2 /mol/glucose,

    hexose equivalent)

    H2productionrate (L/L/d)

    Ref. no.

    Clostridium butyricum   Batch 37 Glucose 3 6.5 2.09 0.41   [11]

    Escherichia coli   Batch 37 Glucose 3 6.5 1.45 0.33   [11]C. butyricum and

    Escherichia coli

    Batch 37 Glucose 3 6.5 1.65 0.52   [11]

    Enterobacter aerogenes

    and C. butyricum

    Batch 37 Starch ND 6.5 2 ND   [12]

    Enterobacter aerogenes

    and C. butyricum

    (immobiolized)

    Repeated

    batch

    37 Starch ND 5.5 2.6 ND   [12]

    Enterobacter aerogenes

    HO-39 and C. butyricum

    Repeated

    batch

    37 Sweet

    potato

    starch

    residue

    ND 5.25a 2.7 0.977   [13]

    C. thermocellum JN4   Batch 60 Cellulose 5 4.4 0.8 0.01   [15]

    Thermoanaerobacterium

    thermosaccharolyticum

    GD17 and C. thermocellum JN4

    Batch 60 Cellulose 5 4.4 1.8 0.33   [15]

    C. acetobutylicum X9   Batch 37 Microcrystall ine

    cellulose

    10 5 0.59 21.33   [14]

    C. acetobutylicum X9 and

    Ethanoligenens harbinese

    Batch 37 Microcrystall ine

    cellulose

    10 5 1.32 11.08   [14]

    Clostridium thermocellum

    and C. thermosaccharolyticum

    Batch 55 Corn stalk

    waste

    10 7.2 ND 0.34   [16]

    Clostridium thermocellum

    and C. thermosaccharolyticum

    CSTR 55 Corn stalk

    waste

    10 7.2 ND 0.44   [16]

    Clostridium thermocellum

    DSM1237 and C.

    thermopalmarium

    DSM5974

    Batch 55 Cellulose 9 7 1.36 0.42   [17]

    Enterobacter aerogenes W23   Batch 35 Glucose 5 6.5 1.87 5.8   [18]

    Enterobacter aerogenesW23 and Candida

    maltosa HY-35

    Batch 35 Glucose 5 6.5 2.19 6.27   [18]

    ND: Not defined.

    a Controlled pH.

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 64954

    http://-/?-http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    11/22

    temperature of 60 C in a batch reactor.The C. thermocellum JN4

    resulted in hydrogen yield of about 0.8 mol H2 /mol glucose,

    with lactate as the main product. When  C. thermocellum JN4

    was co-cultured with T. thermosaccharolyticum GD17, hydrogen

    production was doubled and H2 yield increased to a high level

    of 1.8 mol H2 /mol glucose. Butyrate was the most abundant

    byproduct and lactate was not detected at the end of the co-

    cultures process.Aijie et al.  [14] employed dark fermentation of microcrys-

    talline cellulose to produce biohydrogen using mono and co-

    cultures. Clostridium acetobutylicum ATCC 824(X),   a high

    hydrogen producer from microcrystalline cellulose was uti-

    lized to produce hydrogen at temperature of 37   C, and pH of 

    5.0. The mono-culture of  X9 yielded hydrogen after a 5-h time

    lag. The corresponding hydrogen yield, maximum hydrogen

    production rate, and cellulose hydrolysis ratio reached

    755 mL/L medium, 6.4 mmol H2 /h/g dry cell, and 68.3%,

    respectively. The co-cultures of  C. acetobutylicum X9 and strain

    Ethanoligenens harbinense B49, which can produce hydrogen

    efficiently from both monosaccharides and microcrystalline

    cellulose, yielded hydrogen immediately following initiationof fermentation. The hydrogen yield, maximum hydrogen

    production rate, and cellulose hydrolysis ratio of 1810 mL/L

    medium, 55.4 mmol H2 /h/g dry cell, and 77.6%, respectively,

    were achieved. The strain B49 rapidly removed reducedsugars

    produced by cellulose hydrolysis by   X9, hence improving 

    cellulose hydrolysis and subsequent hydrogen production.

    Another example of technical and economical efficiencies

    attained by the synergies between co-cultures in fermentative

    hydrogen production process is provided by Qian et al.   [16].

    The authors utilized a combination of cellulose-hydrolysing 

    bacteria and highly efficient hydrogen producing bacteria to

    optimize hydrogen production from cellulosic waste. They

    employed a mix of  C. thermocellum   and   Clostridium thermo-saccharolyticum utilizing corn stalk waste. C. thermocellum is a

    cellulose-degrading bacterium, which has the potential for

    direct hydrogen production from lignocellulosic waste, hence

    eliminating the need for an extensive hydrolysing process, but

    with low biohydrogen yield. The C. thermosaccharolyticum is  a

    non-cellulolytic high hydrogen-producing stain. The experi-

    ments were conducted in both batch and continuous-flow

    modes at temperature of 55   C, and pH of 7.2. At the end of 

    the   C. thermocellum   mono-culture experiment, cellobiose,

    glucose, and xylose contents were found in the fermentation

    broth. However, cellobiose and xylose were not detected at the

    end of the   C. thermosaccharolyticum   and   C. thermocellum   co-

    cultures experiments because   C. thermosaccharolyticum   cul-tures produced hydrogen, organic acids (acetate), and other

    components. The hydrogen yield in the co-culture batch

    fermentation reached 68.2 mL/g-cornstalk, which was 94.1%

    higher than that in the mono-culture, and the rate of 

    hydrogen production reached 14.1 mL H2 /L/h. A hydrogen

    yield of 74.9 mL/g cornstalk, as well as production rate of 

    18.5 mL H2 /L/h were achieved using the optimized co-culture

    method in the scaled-up reactor.

    Alei et al.   [17]   investigated employing a cellulolytic

    hydrogen-producing bacterium and a non-cellulolytic high

    hydrogen-producing bacterium. In their study, C. thermocellum

    DSM 1237, a cellulolytic hydrogen-producing bacterium, was

    co-cultured with   Clostridium thermopalmarium DSM 5974,   a

    non-cellulolytic high hydrogen-producing bacterium. The

    bacteria utilized cellulose as the sole substrate, at tempera-

    ture of 55   C, and pH of 7.0. The co-culture produced nearly

    double the amount of hydrogen produced in  C. thermocellum

    monocultures. Ethanol and acetate were the main metabolites

    in   C. thermocellum   monocultures, whereas the co-cultures

    produced butyrate as the main metabolite. These results

    support the synergies between cellulolytic anaerobes andhigh-yield biohydrogen producers. It is thought that using co-

    cultures of various types of complex sugars degrading anaer-

    obes in general and high hydrogen producers would give the

    same results as using co-cultures of cellulose degrading an-

    aerobes and high hydrogen producers.

    The third type of co-cultures reported in the literature in-

    volves aciduric hydrogen producing microorganisms and high

    hydrogen producers. Due to the inhibitory impact of organic

    acids produced during fermentation on biohydrogen pro-

    ducers, near neutral or weak acidic conditions are mandatory

    to attain high hydrogen yields. However, pH control is un-

    economical due to the large quantity of chemicals needed.

    Aciduric microorganisms can produce hydrogen at low pHs,hence reducing or even eliminating buffering requirements.

    Thus, using a hydrogen producing bacterium co-cultured with

    aciduric microorganisms can achieve stable and high-

    hydrogen production at low pHs   [18]. In this study, a batch

    experiment was carried out to measure the hydrogen-

    producing ability of a mixed culture of  Candida maltosa HY-

    35, an aciduric microorganism, which can produce hydrogen

    at pH as low as 1.3, and a facultative anaerobe  E. aerogenes W-

    23,  which has aciduric hydrogen producing properties (can

    produce hydrogen at pH of 4.0 ). These mixed cultures at 35  C

    attained a hydrogen yield of 1735 mL/L, representing 17.15%

    and 119.90% higher yield than the monocultures  E. aerogenes

    W-23   and   C. maltosa HY-35, respectively. Meanwhile, theaverage hydrogen production rate of the mixed culture was

    261.1 mL/h/L, which was 7.85% and 146.23% higher than those

    of the monoculture of  E. aerogenes W-23 and  C. maltosa HY-35,

    respectively. In this case, the co-cultures of hydrogen pro-

    ducing aciduric microorganism and high hydrogen producing 

    anaerobe allowed for dispensing or reducing amount of buff-

    ering agents by providing a wider range of pH for bacteria to

    ferment within. In addition,   Candida  consumed the lactate,

    succinic and citric acids produced by   Enterobacter   during 

    fermentation and slowed the shift in pH. These results

    confirm the synergies between the mixed cultures.

    The success of co-cultures of high hydrogen producing 

    bacteria and hydrogen producing aciduric bacteria empha-sizes the potential for employing co-cultures of multi species

    high-hydrogen producing-bacteria with different optimum pH

    ranges to naturally realize hydrogen production over a wide

    pH range. For instance, it was found that the optimum pH for

    hydrogen production and growth rate for Clostridium DMHV-10

    is 5.0 [19]. Thebacterium fermented 10 g/L glucose with a yield

    of 3.35 mol H2 /mol glucose, at temperature of 37   C, and an

    initial pH of 5.0. The bacterium is capable of growth and

    hydrogen production within a pH range of 5.0e7.0. On the

    other hand, it was reported thatthe optimum pH for hydrogen

    production and growth rate for the hydrogen producer  Enter-

    obacter Cloacae was found to be 6.5 [20]. The bacterium yielded

    3.31 mol H2 /mol glucose from fermenting 10 g/L glucose at a

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 6   4955

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    12/22

    temperature of 37   C, and an initial pH of 6.5. This bacterium

    was capable of growing and producing hydrogen within a pH

    range of 4.5e8.0. Thus, these two strains could be co-cultured

    at an initial pH of 6.5 eliminating the need for a buffer and a

    reducing agent to maintain the anaerobic conditions for the

    Clostridium bacteria.

    3. Comparative study between thermophilesand mesophiles

    Temperature is one of the most important operational pa-

    rameters in fermentative H2 production. Temperature affects

    the growth rate metabolic pathways of microorganism, sub-

    strate hydrolysis rate and hydrogen production rate.

    Fermentative reactions can be operated at mesophilic

    (25e40   C), thermophilic (40e65   C) or hyperthermophilic

    (>80   C) temperatures   [21]. It has been demonstrated that

    within a specific temperature range, increasing the tempera-

    ture accelerates hydrogen production, with sharply dropping 

    activity of hydrogen producers outside the optimum temper-ature range   [8]. The approximately 200 investigations

    reviewed in this study can be classified into two groups: 116

    studies were conducted within mesophilic range; and 78

    studies were carried out within thermophilic range. Per-

    forming experiments employing mesophilic cultures is

    generally less expensive. However, it was reported that ther-

    mophilic and hyperthermophilic cultures seem to exhibit su-

    perior performance in hydrogen production. The highest

    reported hydrogen yields in the literature, which were close to

    the theoretical maximum of 4.0 mol-H2 /mol-glucose, were

    achieved by using extreme thermophiles [22,23].

    In general, thermophiles are thought to be robust micro-

    organisms that produce stable enzymes. It is widely acceptedthat more hydrogen can be produced under thermophilic

    conditions than under mesophilic conditions  [24]. However,

    the data available in the literature does not always support

    this hypothesis, and seem to be substrate dependant. This is

    because some mesophilic bacteria have better bacterial ki-

    netics than thermophilic ones utilizing the same substrate,

    despite operating at much lower temperatures. For instance,

    the hyperthermophilic bacterium,  Thermotoga neapolitana in a

    batch experiment at a temperature of 77   C, and a pH of 7.5,

    was capable of producing 3.85 mol H2 /mol glucose, from 2.5 g/

    L glucose, with a hydrogen production rate of 0.56 L/L/d  [22].

    Giuliana et al.   [23]   reported that   T. neapolitana   achieved a

    maximum hydrogen yield of 3.85 mol H2 /mol and a maximumhydrogen production rate of 1.2 L/L/d, utilizing 5 g/L of glucose

    in serum bottles at a temperature of 80   C, pH of 7.5. On the

    other hand, the maximum hydrogen yield of 3.8 mol H2 /mol

    glucose and hydrogen production rate of 1.82 L/L/d were

    attained by the mesophilic bacterium   Pantoea agglomerans

    utilizing 10 g/L glucose as substrate, at a temperature of 37  C,

    and a pH of 7.2 [25]. Although the latter bacterium was oper-

    ating at mesophilic tempratures, it produced hydrogen at a

    higher rate than the thermophilic one from glucose (mono-

    saccharide), and with almost the same yield.

    The maximum hydrogen yield reported in the literature by

    a thermophile utilizing fructose (another type of mono-

    saccharides) was 3.4 mol H2 /mol hexose equivalent, with a

    maximum hydrogen production rate of 2.4 L/L/d, by the bac-

    teria T. neapolitana [26]. The bacteria utilized 10 g/L fructose at

    a temperature of 75   C, and a pH of 7.0. Nevertheless, the

    maximum hydrogen yield reported in the literature by a

    mesophile utilizing 10 g/L fructose in a batch at a temperature

    of 35   C, and a pH of 6.5 was 1.27 mol H 2 /mol hexose equiva-

    lent by the bacteria E. coli [27].

    The maximum hydrogen yield attained by a thermophileutilizing sucrose (i.e. di-saccharide) was 2.96 mol H2 /mol

    hexose equivalent with a hydrogen production rate of 4.5 L/L/

    d, which was achieved using  Caldicellulosiruptor saccharolyticus,

    at a pH of 7, a temperature of 70   C, and an initial concen-

    tration of 10 g/L, in a batch reactor   [28]. The maximum

    hydrogen yield of a mesophile utilizing sucrose was reported

    by Narendra et al. [29] for E. cloacae. They achieved a hydrogen

    yield and a hydrogen production rate of 3.1 mol H2 /mol hexose

    equivalent and 15.84 L/L/d, respectively, at a temperature of 

    36  C, a pH of 6.0, and initial sucrose concentration of 10 g/L.

    It has beenrecently reported [30] that mesophilic anaerobic

    bacteria cannot utilize cellulose (i.e. complex sugars) effec-

    tively. The addition of exogenous cellulose enzymes isnecessary for hydrolysis of cellulose to generate H2 by meso-

    philic anaerobic bacteria. On the other hand, thermophilic

    anaerobic bacteria can effectively utilize cellulose   [31], and

    therefore, they have a great potential for H2 production from

    cellulose without the addition of exogenous cellulose  [15]. In

    addition, the high operating temperature of the thermophiles

    enhances the hydrolysis rate. For example, Rumana et al. [32]

    reported that the bacterium C. thermocellum utilized 1 g/L cel-

    lulose, and produced a hydrogen yield of 1.9 mol H2 /mol

    hexose equivalent. On the other hand, the maximum attained

    hydrogen yield reported in the literature from mesophiles (C.

    acetobutylicum  and   E. harbinense) utilizing cellulose at a con-

    centration of 10 g/L was only 1.32 mol H2 /mol hexose equiv-alent [14]. These results confirm that thermophiles can more

    effectively utilize complex sugars for hydrogen production

    than mesophiles.

    4. Bioreactor configuration

    Fermentative hydrogen production was applied in both batch

    reactors and continuous systems. Batch-mode reactors are

    easily and flexibly operated. This has resulted in the wide

    utilization of batch reactors for determining the biohydrogen

    potential of organic substrates. In industrial context, however,

    continuous bioprocesses are recommended for practicalconsiderations such as waste stock management, economic

    feasibility, and practical engineering design [33].

    Seven cases of batch studies are reported in the literature

    to be successfully scaled up or applied to continuous-flow

    systems. The performance of scaled up continuous systems

    varied from one study to another. In some cases, the same or

    even better performance was achieved compared to their

    batch counterparts. In other cases, however, the continuous

    systems displayed lesser but stable performance. Table 3 il-

    lustrates the differences in performance and operational pa-

    rameters between the batches and continuous-flow systems

    employing the same bacteria, and utilizing the same sub-

    strates. Scrutinizing the results in Table 3, it may be concluded

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 64956

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    13/22

    Table 3 e Operational and performance parameters of the batch and their continuous counterparts reviewed studies.

    Culture Reactortype

    TC Substratetype

    Substrateconcentration

    (g/L)

    pH Hydrogen yield(mol H2 /mol

    glucose, hexoseequivalent)

    1-Clostridium thermocellum

    7072   Batch 55 Corn stalk 5 7.4 ND

    7072   CSTR (10 L) 55 Corn stalk 30 7.4 0.43

    7072   CSTR (100 L) 55 Corn stalk 30 7.4 0.45 2-Clostridium tyrobutyricum

    FYa102   CSTR 35 Glucoseb 12, 8 6 1.06

    FYa102   CSTR 35 Glucosec 12, 0.35, 1.4 6 1.42

    FYa102   Batch 35 Glucosed 3, 2 7.2 1.47

     JM1   CSTR 37 Glucose 5 6.7 1.81

     JM1   Batch 37 Glucose 20 6.3 3.24

    3-Ethanoligenens harbinese

    YUAN-3   Batch 35 Glucose 10 5 1.91

    YUAN-3   CSTR 35 Glucose 10 5 1.93

    4-Thermotoga neapolitana

    DSM 4359   Batch 75 Xylose 5 7.5 1.31

    DSM 4359   CSTR 75 Xylose 5 7 3.36

    5-Thermoanaerobacterium thermosaccharolyticum

    PSU-2   Cont. UASBa 60 Sucrose 20 5.5 1.77

    PSU-2   Batch 60 Sucrose 20 6.25 2.53

    ND: Not defined.

    a UASB  ¼  up flow anaerobic sludge blanket.

    b 2 g/L peptone was added with the substrate.

    c 8 g/L peptone was added with the substrate.

    d 0.36 g/L, 1.4 g/L peptone and ammonium chloride respectively, were added with the substrate.

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    14/22

    that optimizing the operational parameters is the key for

    achieving a successful continuous-flow system.

    Continuous-flow system performance was superior to the

    batch in thermophilic fermentation of cornstalk by   C. ther-

    mocellum 7072 [1]. In the batch test involving a substrate of 5 g/

    L cornstalk at a temperature of 55  C and a pH of 7.4, hydrogen

    yield and hydrogen production rate of 38.8 mL/g and 201.4mL/

    L/h were achieved. The continuous stirred tank reactor (CSTR)was fed with 30 g/L of cornstalk, and operated at the same

    temperature and pH as the batch test. The maximum

    hydrogen production rate in the 10 L CSTR, and the 100 L CSTR

    was 767.5, and 739.9 mL/L/h, respectively. The authors

    attributed the higher hydrogen production rates and shorter

    lag-phase period observed in the two CSTRs to the improved

    mixing conditions in the two reactors, compared to the

    anaerobic bottles. The hydrogen yield in the 10 and 100 L

    CSTRs reached 58.3 and 61.4 mL/g of cornstalk. Acetate and

    ethanol were the major end products of fermentation by  C.

    thermocellum   for cornstalk, and the ratio of ethanol/acetate

    was lower in both CSTRs than in the 125 mL anaerobic bottles.

    The substrate concentration was the only different opera-tional parameter between the CSTRs and batch tests. Thus, it

    was presumed that it also contributed to the higher produc-

    tion rate and yield, as higher substrate concentration in-

    creases fermentation rate. The shift from the ethanol to the

    acetate pathway in the CSTRs explained the higher attained

    hydrogen yields in the CSTRs.

    Cheng and Liu   [1]   used  C. thermocellum   and achieved a

    hydrogen yield of 1.2 mol H2 /mol hexose equivalent, at a pH of 

    7.0, and a temperature of 60   C. These results are consistent

    with the 0.98e1.65 mol H2 /mol hexose equivalent observed by

    Lauren et al. [34] employing a CSTR at neutral pH, and influent

    of cellulose concentration in the range of 1.5e4 g/L.

    In another study, Liang-Ming et al.   [35]  investigated thefermentative biohydrogen production in CSTRs using   Clos-

    tridium tyrobutyricum FYa102. Two CSTRs were employed in

    this study: one, denoted (GP), was fed with 12 g/L of glucose

    and 8 g/L of peptone, while the other, denoted (GA), was fed

    with 12 g/L of glucose, 1.4 g/L of ammonium chloride, and

    0.360 g/L of peptone. The experiments were carried out at a

    temperature of 35  C, and a pH of 6.0. The hydrogen yield and

    hydrogen production rate achieved for the (GA) and (GP) re-

    actors were 1.42 mol H2 /mol glucose, and 3.1 L/L/d, and

    1.06 mol H2 /mol glucose, and 10.3 L/L/d, respectively. On the

    other hand, Pei-Ying et al.   [31], conducted biochemical

    hydrogen potential (BHP) tests to investigate the metabolism

    of glucose fermentation and hydrogen production perfor-mance of   C. tyrobutyricum FYa102   in batches. Glucose and

    peptone were used in the fermentation medium at initial

    concentrations of 3 g/L, and 2 g/L, respectively. The experi-

    ment was conducted at a temperature of 35  C and a pH of 7.2.

    The attained hydrogen yield and hydrogen production rates

    were 1.47 mol H2 /mol glucose, and 1.6 L/L/d, respectively. In

    glucose re-feeding experiments, the  C. tyrobutyricum FYa102

    fermented additional glucose during re-feeding tests, pro-

    ducing a substantial quantity of hydrogen. The higher

    hydrogen production rates attained in the CSTRs were

    attributed to the higher rate of fermentation resulting from

    the higher concentrations of glucose and peptone in the me-

    dium, and the better mixing conditions in the CSTRs.

    Although the CSTRs were fed with a much higher concentra-

    tion of glucose, the hydrogen yields were almost the same or a

    bit less than those attained in the batch studies. This differ-

    ence in hydrogen yields was attributed to the different

    ambient pHs in the batches and the CSTRs experiments. It is

    believed that if the CSTRs experiments were conducted at a

    pH of 7.2, higher hydrogen yields would have been attained. It

    must be noted, however, that in the CSTRs study   [35], thehydrogen production rate in the (GP) reactor was 3.5 times

    higher than the (GA) due to a much higher organic loading 

    rate, but the yield of the (GP) was only 75% of that of the (GA)

    due to a lower glucose fraction.

     Ji et al. [36] immobilized the hydrogen producing anaerobe,

    C. tyrobutyricum JM1   in a packed-bed reactor using poly-

    urethane foam as support media. The hydraulic retention

    time (HRT) condition for maximum hydrogen production rate

    in this system was 2 h, where the main metabolite was

    butyrate with low lactate concentration, andhydrogen yield of 

    1.81 mol H2 /mol glucose was attained at a pH of 6.7, temper-

    ature of 37   C, and a feed glucose concentration of 5 g/L.

    Therefore, the immobilized system was an effective and sta-ble approach for continuous hydrogen production for efficient

    utilization of carbon substrates with good hydrogen-

    producing performance. However, in a later study by the

    same group [37], the effects of pH on hydrogen fermentation

    of glucose by the same bacterium were investigated in batch

    cultivations. The batcheswere conducted at different pHs (6.0,

    6.3, 6.7), temperature of 37   C, and a glucose concentration of 

    20 g/L. The initial low glucose concentration (such as the 5 g/L

    of glucose used in the previous study  [36]) resulted in a low

    fermentation rate, and consequently a low hydrogen yield. It

    was proven that a pH of 6.3 was optimum for hydrogen pro-

    duction with a high concentration of butyrate, and a hydrogen

    yield of 3.24 mol H2 /mol glucose. The lower hydrogen yieldachieved in the continuous-flow system was attributed to the

    un-optimized pH and substrate concentration.

    Defeng et al.   [38]   studied hydrogen production of auto-

    aggregative (self-flocculating granular)  E. harbinense YUAN-3

    in a batch reactor and a continuous stirred-tank reactor

    (CSTR), with glucose as substrate under non-sterile condi-

    tions. In the batch reactor, the optimized operational condi-

    tions constituteda pH of 5.0, temperature of 35 C, and glucose

    concentration of 10 g/L. The maximum hydrogen yield and

    hydrogen production rate under the optimum operational

    conditions were 1.91 mol H2 /mol glucose and 1.66 L/L/d,

    respectively. In the CSTR, hydrogen gas yield reached a

    maximum of 1.93 mol H2 /mol glucose, and H2 production ratereached a maximum of 19.6 L/L/d.The strain YUAN-3 was well

    retained in thereactor. The overflow rate of cells was less than

    0.1% in the continuous flow reactor, at a dilution rate of 0.5/h.

    However, after 7 days of continuous operation some other

    hydrogen-producing bacterial species appeared and formed a

    stable community with   YUAN-3. The hydrogen yield

    decreased from 0.93 mol H2 /mol glucose to 1.5 mol H2 /mol

    glucose and stabilized thereafter. The dominant populations

    in the continuous-flow reactor were affiliated with  M. hominis,

    and  M. sueciensis,  and the majority of dominant populations

    belonged to E. harbinense, which were enriched during opera-

    tion of the reactor. These results indicate that a successful

    continuous operation was achieved. It is evident from the

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 64958

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    15/22

    above results that optimizing the operational parameters for

    the auto-aggregative bacteria achieved continuous stable

    hydrogen production, despite the occurrence of microbial

    shift.

    Tien et al.  [30]  investigated biohydrogen production from

    xylose by   T. neapolitana in batch culture using serum bottles

    and a continuously stirred anaerobic bioreactor (CSABR). A

    maximum hydrogen production rate of 0.44 L H2 /L/d and amaximum hydrogen yield of 1.31 mol H2 /mol hexose equiva-

    lent were obtained in the serum bottles test, at an initial

    xylose concentration of 5.0 g/L, and a pH of 7.5. The CSABR

    was run at uncontrolled and controlled pH conditions. In the

    uncontrolled pH experiments, the fermentation process

    ceased before the complete consumption of substrate due to

    the drastic decrease in pH. In pH-controlled cultures, much

    higher H2  production and xylose utilization rates were ach-

    ieved as evident from the levels of acetic acid varying from 2.5

    to 3.5 g/L compared to 2.5 g/L in the uncontrolled batches. In

    contrast to acetic acid production, lactic acid production was

    the lowest under pH-controlled conditions. Subsequently, the

    H2   production rate increased exponentially reaching themaximum level. The maximum H2  yield, and hydrogen pro-

    ductionrateof 2.8 molH2 /mol xylose consumed, and 2.66 L H2 /

    L/d were measured while the pH was maintained at 7.0. It

    appears that controlling pH at neutral limit instead of an

    initial pH of 7.5 was the key for a stable continuous system.

    Another example of the effect of un-optimized operational

    parameters on the performance of continuous-flow system

    was provided by Sompong et al.   [39]. They investigated the

    fermentation of 20 g/L sucrose by the bacterium  T. thermo-

    saccharolyticum strain PSU-2 in an UASB bioreactor. The system

    was stable, and the hydrogen yield and production rate of 

    1.77 mol H2 /mol hexose and 5.9 L H2 /L/d were achieved at a

    temperature of 60   C and a pH of 5.5. However, the same au-thors investigated in another study [40] the fermentation of 

    the same concentration of sucrose under the same tempera-

    ture, and a wide range of pH (4.0e9.0) by the same bacterium

    and observed maximum hydrogen yield and hydrogen pro-

    duction rate of 2.53 mol H2 /mol hexose equivalent, and 6.5 L/L/

    d at a pH of 6.25. Therefore, it is presumed that if the contin-

    uous system was operated at the optimum pH, a higher

    hydrogen yield and production rate would have been realized.

    5. Feedstocks

    Hydrogen can be produced from a wide spectrum of carbo-hydrates. Nevertheless, 80% of the studies reported in the

    surveyed literature have investigated hydrogen production by

    dark fermentation from pure sugars, such as glucose, or su-

    crose as substrate. Only a few studies have focussed on sus-

    tainable substrate conversion (Fig. 1). However, for real value

    to the society and environment, biohydrogen should be pro-

    duced from renewable feedstocks (real waste)   [41]. The po-

    tential feedstocks include: biomass, agricultural waste bi-

    products, lignocellulosic products (wood and wood waste),

    waste from food processing, aquatic plants, algae, agricul-

    tural, and livestock effluents. If used under appropriate con-

    trol, these resources would become the major source of 

    energy in the future. In this study, different types of 

    feedstocks are discussed in terms of their applicability and

    operational challenges, as well as the motivation for their use

    in fermentative hydrogen production.

    5.1. Pure carbohydrates (synthetic waste)

    Pure carbohydrate sources are expensive raw materials for

    real scale hydrogen production (which can only be viablewhen based on renewable and low cost sources). Neverthe-

    less, the majority of the reviewed studies utilized pure car-

    bohydrates as substrate, including: monosaccharides

    (glucose, xylose, fructose, arabinose, mannose and ribose);

    disaccharides (sucrose, cellubiose, maltose, and lactose); or

    polysaccharides (starch, cellulose, and xylan).

    Simple sugars such as glucose, sucrose, and lactose are the

    most commonly used pure substrates due to their ease of 

    biodegradability, relatively simple structures, and presence in

    real industrial effluents  [42,43]. Unlike starch and cellulose,

    they require short fermentation times (i.e. process HRT),

    which makes these substrates preferred model substrates for

    hydrogen production studies. Model substrates are employedin fermentative hydrogen production processes to study bac-

    terial kinetics, assess adequate nutrients, and identify the

    optimized operational parameters for the process [19,44e48].

    However, real life applications involve complex sugars, and

    thus it is indispensable to employ these substrates in the dark

    fermentation process in order to provide relevant insight into

    system performance [1,49].

    Hydrolysis of real waste comprising different sugars was

    modelled by co-digestion of various pure substrates. The

    control experiments assessed the fermentation preferences of 

    the bacteria among the different types of sugars. Pan-

    agiotopoulos et al.   [50] conducted four experiments to eval-

    uate the hydrogen production and the main organic acids by

    Pure monosaccharides

    59%

    Pure polysaccharides

    11%

    Sustainable

    feedstocks

    (current+future)

    20%

    Fig. 1  e  Percentage of the usage of pure and real waste

    substrates in the reviewed literature.

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 6   4959

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    16/22

    C. saccharolyticus   from the hydrolysed sweet sugar bagasse

    (SSB) and from a mixture of pure sugars (glucose, sucrose and

    xylose). They employed SSB in 2 experiments at sugar con-

    centrations of 10 g/L and 20 g/L, and conducted two control

    experiments employing mixtures of pure sugars of glucose,

    xylose, and sucrose, at concentrations of 10 g/L and 20 g/L. At a

    sugar concentration of 10 g/L, consumption of pure sugars and

    sugars of SSB hydrolysate was complete within similarfermentation time. At the 20 g/L of pure sugars, consumption

    was still incomplete at 72 h, while sugars of SSB hydrolysate

    were completely consumed at 70 h. The consumption pattern

    of 20 g/L sugars of SSB hydrolysate sugars differed markedly

    from that of pure sugars. Lactate production only occurred in

    fermentations on SSB hydrolysate. Therefore, hydrogen pro-

    duction and hydrogen yields were higher in fermentations on

    pure sugars than of SSB hydrolysate. The high rate of glucose

    consumption in the fermentation of 20 g/L of SSB hydrolysate

    sugars coincided with the high rate of lactate production in

    this fermentation. Basically, at sugar concentration of 10 g/L,

    the batch fermentations under controlled conditions

    confirmed the results of the fermentability tests, but at highersubstrate concentrations, lactate production increased

    dramatically at the expense of hydrogen production.

    Trus de Vrije et al. [51] investigated thermophilic hydrogen

    production using   C. saccharolyticus  and  T. neapolitana  on hy-

    drolysate of the lignocellulosic feedstock Miscanthus (obtained

    from enzymatic hydrolysis) in batch tests. Control experi-

    ments were also conducted at different mixing ratios of xylose

    and glucose to assess the utilization preference of the bacteria

    for a sugar type over the other and the optimum substrate

    concentration. The authors observed that   T. neapolitana

    showed a preference for glucose over xylose, which were the

    main sugars in the hydrolysate, while   C. saccarolyticus

    consumed both at a similar rate. Lactate production by   C.saccarolyticus was very low in fermentations on pure sugars, as

    well as on hydrolysate.   T. neapolitana  produced more lactate

    on the hydrolysate than on pure sugars. The optimum total

    sugars concentration was 17 g/L, and  C. saccharolyticus offered

    the advantage of nearly 10% higher hydrogen yield during 

    growth on Miscanthus hydrolysates as compared to T. neapo-

    litana, but the rates of substrate consumption and hydrogen

    production by T. neapolitana were 5e18% higher.

    Trus de Vrije et al.   [26]  investigated hydrogen production

    from carrot pulp hydrolysate (obtained from enzymatic hy-

    drolysis) by the same thermophilic bacteria  C. saccharolyticus

    and   T. neapolitana. The main sugars in the hydrolysate were

    glucose, fructose, and sucrose. Therefore, they initiallyinvestigated hydrogen production from different concentra-

    tions of glucose, fructose, and mixtures of glucose and fruc-

    tose as control experiments. in order to assess the adequate

    degree of hydrolysation and optimized substrate concentra-

    tion by determining the preferred sugar type for bacteria. They

    observed that in fermentations of 10 g/L glucose and 10 g/L

    fructose, C. saccharolyticus could virtually completely consume

    all substrates with almost identical rates of consumptions. In

    contrast,  T. neapolitana  consumption trend was different for

    glucose than fructose, suggesting a preference for glucose. In

    fermentations of 20 g/L of substrate, the consumption of 

    substrate was incomplete for both cultures even after 2 days.

    Also, they found that the cultures productivities were

    equivalent or higher than the productivities achieved with the

    corresponding pure sugars (mixtures of glucose and fructose)

    at 10 g/L sugars. Doubling the hydrolysate concentration had

    adverseeffect on hydrogen production, with a severe decrease

    in yield in  C.  saccharolyticus  cultures and a decrease in pro-

    ductivity with T. neapolitana.

    Nan-Qi Ren et al. [52] investigated the utilization of an agro-

    waste, corn stover, as a renewable lignocellulosic feedstockforthe fermentative H2 production by the moderate thermophile

    T. thermosaccharolyticum W16. The corn stover was hydrolysed

    by cellulase with supplementation of xylanase after delignifi-

    cation with 2% NaOH, producing glucose, xylose, and arabi-

    nose. To determine the fermentative behaviour of the

    bacterium, a set of control experiments supplemented with

    glucose, xylose, and a mixture of glucose, xylose, and arabi-

    nose at a fixed total sugar quantity of 10 g/Lwere undertaken.

    The concentrations of glucose, xylose, and arabinose in the

    mixture were at the same levels as found in the corn stover

    hydrolysate. It was observed that the bacterium showed pref-

    erence for glucose over the other types of sugars. The bacte-

    rium grew well on the hydrolysate and reached a similaroptical density and maximum hydrogen production rate as on

    simulated medium, although hydrogen yield was slightly

    higher on hydrolysate. Although the molar carbon balances in

    the control experiments closed at 100%, carbon balances did

    not close in the hydrolysate, most probably due to the inter-

    ference of unidentified components in the hydrolysate.

    5.2. Sustainable feedstocks (real waste)

    For sustainable biohydrogen production, the feedstock has to

    be cheap and would have to meet the following criteria: car-

    bohydrate produced from sustainable resources; sufficient

    concentration that fermentative conversion and energy re-covery is energetically favourable; and minimum pretreat-

    ment [53].

    Biomass is a viable renewable resource. It includes agri-

    cultural residues, energy crops, and industrial wastes, which

    can be used for the production of power, heat and biofuels

    [50]. Producing hydrogen from biomass greatly enhances the

    security of supply   [26]. Therefore, most recent studies

    employed biomass for biological conversion via fermentation

    processes. As shown in   Table 4, sugar-containing crops

    (e.g.sweet sorghum and sugar beet), starch-based crops (e.g.

    corn and wheat), ligno-cellulosics (e.g. fodder grass and mis-

    canthus), and food industry by-products are all biomass types

    used as substrates in the literature [1,13,16,26,49e

    52,54e

    64].In view of the increasingly negative public reaction to the

    use of food for biofuel production, employing energy crops as

    feedstocks for biofuels generation (e.g. wheat straw, barely

    straw, corn stalk, miscanthus, and cassava) is widely accepted

    in the scientific community   [1,51,54,55,61]. They are

    commonly referred to as second generation cellulosic

    biomass. Additionally, utilizing industrial and agricultural

    waste residues (e.g. delignified wood fibres, and corn stalk

    waste) [16,35,49,52,60], or food industry waste (e.g. carrot pulp,

    potato steam peels, sugarcane waste, sweet sorghum syrup,

    corn starch, and sweet potato starch)   [13,26,50,56e59,63,64]

    addresses the concerns of skyrocketing food and energy

    prices.

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 64960

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    17/22

    Table 4 e Operational and performance parameters of the studies utilizing sustainable feedstocks.

    Substrate types Culture(s) Reactor type   TC Substrateconcentration

    (g/L)

    pH Hydrogen yi(mol H2 /mol glu

    hexose equiva

    1-Current sustainable feedstocks:

    Pretreated wheat straw   Caldicellulosiruptor saccharolyticus   Batch 70 10 7.2 3.8

    Pretreated barely straw   Caldicellulosiruptor saccharolyticus   Batch 70 20h 7 ND

    Carrot pulp hydrolysate   Caldicellulosiruptor saccharolyticus   Batch 70 10 7 2.8

    PSPa Caldicellulosiruptor saccharolyticus   Batch 72 10 7 3.5 PSP-H2b Caldicellulosiruptor saccharolyticus   Batch 72 10 7 3.4

    SSBc Caldicellulosiruptor saccharolyticus   Batch 72 20h 6.8i 2.8

    Miscanthus hydrolysate   Caldicellulosiruptor saccharolyticus   Batch 72 10 7 3.4

    Miscanthus hydrolysate   Caldicellulosiruptor saccharolyticus   Batch 72 14 7 3.3

    Miscanthus hydrolysate   Caldicellulosiruptor saccharolyticus   Batch 72 28 7 2.4

    PSP-H2b Thermotoga neapolitana   Batch 75 10 7 3.3

    PSPa Thermotoga neapolitana   Batch 75 10 7 3.8

    Carrot pulp hydrolysate   Thermotoga neapolitana   Batch 75 10 7 2.7

    Carrot pulp hydrolysate   Thermotoga neapolitana   Batch 75 20 7 2.4

    Miscanthus hydrolysate   Thermotoga neapolitana   Batch 80 10 7 2.9

    Miscanthus hydrolysate   Thermotoga neapolitana   Batch 80 14 7 3.2

    Miscanthus hydrolysate   Thermotoga neapolitana   Batch 80 28 7 2

    SCB hemicellulose

    hydrolysatef C. butyricum   Batch 37 20g  5.5 1.73e

    Pretreated straw

    hydrolysate

    C. butyricum CGS5   Batch 37 9.2 7.5 0.91

    Sugarcane juice   C. butyricum TISTR 1032   Batch 37 22.3

    (sucrose)

    6.5 1.33

    Sugarcane juice   C. butyricum TISTR 1032

    ( immobilized)

    Repeated

    batch

    37 22.3

    (sucrose)

    6.5 1.52

    Molass   C. butyricum W5   Batch 37 100 7 1.63

    Sweet potato starch

    residiue

    C. butyricum and Enterobacter

    aerogenes HO-39

    Repeated

    batch

    37 ND 5.25i 2.7

    Hydrolysed corn stover   Thermoanaerobacterium

    thermosaccharolyticum W16

    Batch 60 ND 7 2.24e

    Hydrolyzed corn stover   Thermoanaerobacterium

    thermosaccharolyticum W16

    Batch 60 ND 7 ND

    Cassava wastewater   Clostridium acetobutylicum

    ATCC 824

    Batch 36 5g  7 2.41

    Cron stalk   Clostridium thermocellum 7072   CSTR, 100 L 55 30 7.4 0.45

    Cron stalk   Clostridium thermocellum 7072   CSTR, 10 L 55 30 7.4 0.43

    Cron stalk   Clostridium thermocellum 7072   Batch 55 5 7.4 ND

    Delignefied wood fibres   Clostridium thermocellum

    ATCC27405

    Batch 55 0.1 6.5 1.6

    Corn stalk waste   Clostridium thermocellum and

    C. thermosaccharolyticum

    Batch 55 10 7.2 ND

    Corn stalk waste   Clostridium thermocellum and

    C. thermosaccharolyticum

    CSTR 55 10 7.2 ND

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    18/22

    Table 4 e ( continued )

    Substrate types Culture(s) Reactor type   TC Substrateconcentration

    (g/L)

    pH Hydrogen yi(mol H2 /mol glu

    hexose equiva

    Corn starch hydrolysate   Enterobacter aerogenes NCIMB 10102   Continuous

    packed col.

    40 ND 5.5 2.55

    Starch hydrolysate   Enterobacter aerogenes

    NCIMB 10102

    Batch 40 20 6.5 1.09e

    2-Future sustainablefeedstocks:

    POMEd C. butyricum EB6   Batch 37 ND 5.5 0.22

    Glycerol   Klebsiella pneumoniae DSM2026   Batch 37 20 6.5 0.53

    Glycerol   Enterobacter aerogenes   Batch 37 20 7 0.2

    Glycorel   Enterobacter aerogenes

    ATCC35029

    Batch 37 21 ND 1.22

    Chlorella vulgaris ESP6

    (microalgal hydrolysate)

    C. butyricum CGS5   Batch 37 9 5.5i ND

    ND: Not defined.

    a Untreated potato steam peels, Molar yields were based on the amount of starch in untreated PSP assuming 100% starch consumption.

    b The starch in the PSP was liquefied with alpha-amylase, and then the liquefied starch was further hydrolyzed to glucose by amyloglucosidase.

    c Sweet sorgham bagasse.

    d Palm oil mill effluent.

    e mol H2 /mol total sugar.

    f Sugarcane bagasse hemicellulose hydrolysate.

    g g COD/L.

    h g sugars/L.

    i Controlled pH.

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/17/2019 Review Paper on Biohydrogen

    19/22

    The reported hydrogen yields from biomass used as sub-

    strates varied greatly from approximately 20% to more than

    90% of the theoretical 4 moles of H2  per mol of hexose. The

    diversity of the applied feedstocks and pretreatment methods

    hardly allow a comparison of hydrogen production efficiency.

    5.3. Future feedstocks

    Based on the reviewed literature, It is evident that the current

    focus is primarily on food, agriculture, and industry-related

    substances to provide sustainable feedstocks. The bio-

    hydrogen technology would have a rather limited scope if the

    feedstock range and sources cannot be expanded intensively

    [41]. Furthermore, some wastewaters have promising poten-

    tial for biohydrogen production process (Table 4), including:

    oil industry wastewaters [65]; and biodiesel wastes containing 

    glycerol   [66e68]. In addition, microalgal biomass, which is

    produced by CO2  fixation through photosynthesis of micro-

    algae, was proven to be a good sustainable feedstock   [69].

    Although these industrial by-products are considered prom-

    ising approaches for sustainable biohydrogen production, theyields reported from utilizing these wastewaters were still low

    compared to the traditional sustainable feed stocks. Further

    research in utilizing these substrates via dark fermentation,

    and the adequate pretreatments methods is required.

    6. Concluding remarks

    Based on the findings of this literature review, the following 

    remarks can be drawn:

     Attaining technical and economic efficiencies is the main

    drive behind employing co-cultures of pure bacteria infermentative hydrogen production.

     There are three types of co-cultures of pure isolates

    a) Co-cultures of strict high hydrogen producers and

    facultative anaerobes, which is used to attain anaer-

    obic conditions without the need to add expensive

    reducing agents. These co-cultures yield better per-

    formance parameters, especially for complex sugars

    substrates.

    b) Co-cultures of cellulose-degrading anaerobes and high

    hydrogen producers capable of producing hydrogen

    from simpler forms of sugars. These co-cultures offer

    economical and technical advantages over cellulose

    degrading anaerobe solely or enzymatically hydro-lysed cellulose. They produce hydrogen in two steps;

    cellulose degrading anaerobe via the initial degrada-

    tion followed by high hydrogen-producing anaerobe

    from the degraded sugars.

    c) Co-cultures of aciduric microorganisms and hydrogen

    producers which reduces alkali consumption, hence

    reducing or eliminating the need for a buffer to

    maintain a neutral or weak acidic pH.

     The perceived advantages of thermophiles over mesophiles

    appear to be substrate-dependant:

    a) For simple sugars (monosaccharides and di-

    saccharides), either mesophiles or thermophiles could

    produce more hydrogen from fermenting simple

    sugars depending on bacterial kinetics, and the sub-

    strate type.

    b) For complex sugars, thermophiles outperform meso-

    philes in terms of hydrogen production due to their

    ability to degrade complex substrates, in addition to

    the increased hydrolysis and fermentation rates

    associated with the high operating temperature.

      It is essential to first determine optimal operational condi-tions batch studies. Continuous systems can then be oper-

    ated under these optimal operational conditions to achieve

    a sustainable system with same or better performance pa-

    rameters than its batch counterpart.

      Biodiesel wastes, oil industry wastewaters, and microalgal

    biomass have significant potential as sustainable feed

    stocks in the near future.

      Certain aspects of biohydrogen production merit further

    research, including:

    a) Employing co-cultures of high hydrogen producers of 

    facultative and strict anaerobes with different opti-

    mum pH ranges.

    b) Use of co-cultures of complex sugars degrading an-aerobes and high hydrogen producers.

    c) Enhancement of the hydrogen yields and hydrogen

    production rates from dark fermentation of emerging 

    sustainable feedstocks.

    Acknowledgement 

    The authors acknowledge NSERC, GreenField Ethanol, Union

    Gas, and Admira Energy for their financial support of the

    project, as well as the Ontario Trillium Ph.D. Scholarship

    Program awarded to Ms. Omneya Elsharnouby.

    r e f e r e n c e s

    [1] Cheng X, Liu C. Hydrogen production via thermophilicfermentation of cornstalk by  Clostridium thermocellum.Energy Fuels 2011;25:1714e20.

    [2] Das D, Veziroglu TN. Advances in biological hydrogenproduction processes. Int J Hydrog Energy 2008;33:6046e57.

    [3] Nishio N, Nakashimada Y. High rate production of hydrogen/methane from various substrates and wastes.Adv Biochem Eng Biotechnol 2004;90:63e87.

    [4] Wang J, Wan W. Factors influencing fermentative hydrogenproduction: a review. Int J Hydrog Energy 2009;34:799e811.

    [5] Hafez H, Baghchehsaraee B, Nakhla G, Karamanev D,Margaritis A, El Naggar H. Comparative assessment of decoupling of biomass and hydraulic retention times inhydrogen production bioreactors. Int J Hydrog Energy 2009;34:7603e11.

    [6] Hafez H, Nakhla G, El Naggar H. An integrated system forhydrogen and methane production during landfill leachatetreatment. Int J Hydrog Energy 2010;35:5010e4.

    [7] Ginkel S, Sung S, Lay J. Biohydrogen production as afunction of pH and substrate concentration. Environ SciTechnol 2001;35:4726e30.

    [8] Hafez H, Nakhla G, El Naggar H. Biological hydrogenproduction. In: Sherif SA, editor. Handbook of hydrogen

    energy. Boca Raton: CRC Press, Taylor and Francis; 2012.

    i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 4 9 4 5 e4 9 6 6   4963

    http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032http://dx.doi.org/10.1016/j.ijhydene.2013.02.032

  • 8/17/2019 Review Paper on Biohydrogen

    20/22

    [9] Lee D, Show K, Su A. Dark fermentation on biohydrogenproduction: pure culture. Bioresour Technol 2011;102:8393e402.

    [10] Carere CR, Sparling R, Cicek N, Levin DB. Third generationbiofuels via direct cellulose fermentation. Int J Mol Sci 2008;9:1342e60.

    [11] Seppa ¨ la ¨  JJ, Puhakka JA, Yli-Harja O, Karp MT, Santala V.Fermentative hydrogen production by Clostridium butyricum

    and  Escherichia coli  in pure and cocultures. Int J Hydrog Energy 2011;36:10701e8.

    [12] Yokoi H, Tokushige T, Hirose J, Hayashi S, Takasaki Y. H2production from starch by a mixed culture of  Clostridiumbutyricum and Enterobacter aerogenes. Biotechnol Lett 1998;20:143e7.

    [13] Yokoi H, Maki R, Hirose J, Hayashi S. Microbial production of hydrogen from starch-manufacturing wastes. BiomassBioenerg 2002;22:389e95.

    [14] Wang A, Ren N, Shi Y, Lee D. Bioaugmented hydrogenproduction from microcrystalline cellulose using co-cultureedClostridium acetobutylicum and  Ethanoigenensharbinense. Int J Hydrog Energy 2008;33:912e7.

    [15] Liu Y, Yu P, Song X, Qu Y. Hydrogen production fromcellulose by co-culture of  Clostridium thermocellum  JN4 and

    Thermoanaerobacterium thermosaccharolyticum GD17. Int JHydrog Energy 2008;33:2927e33.

    [16] Li Q, Liu C. Co-culture of  Clostridium thermocellum andClostridium thermosaccharolyticum  for enhancing hydrogenproduction via thermophilic fermentation of cornstalkwaste. Int J Hydrog Energy 2012;37:10648e54.

    [17] Geng A, He Y, Qian C, Yan X, Zhou Z. Effect of key factors onhydrogen production from cellulose in a co-culture of Clostridium thermocellum and  Clostridium thermopalmarium.Bioresour Technol 2010;101:4029e33.

    [18] LuW,WenJ,ChenY,SunB,JiaX,LiuM,etal.Synergisticeffectof  Candida maltosa HY-35 and Enterobacter aerogenes W-23 onhydrogen production. Int J Hydrog Energy 2007;32:1059e66.

    [19] Kamalaskar LB, Dhakephalkar PK, Meher KK, Ranade DR.High biohydrogen yielding Clostridium sp. DMHC-10

    isolated from sludge of distillery waste treatment plant. Int JHydrog Energy 2010;35:10639e44.

    [20] Nath K, Kumar A, Das D. Effect of some environmentalparameters on fermentative hydrogen production byEnterobacter cloacae DM11. Can J Microbiol 2006;56:525.

    [21] Sinha P, Pandey A. An evaluative report and challenges forfermentative biohydrogen production. Int J Hydrog Energy2011;36:7460e78.

    [22] Munro SA, Zinder SH, Walker LP. The fermentationstoichiometry of   Thermotoga neapolitana and influence of temperature, oxygen, and pH on hydrogen production.Biotechnol Prog 2009;25:1035e42.

    [23] d’Ippolito G, Dipasquale L, Vella FM, Romano I,Gambacorta A, Cutignano A, et al. Hydrogen metabolism inthe extreme thermophile   Thermotoga neapolitana. Int J

    Hydrog Energy 2010;35:2290e5.[24] Zhang T, Liu H, Fang HHP. Biohydrogen production from

    starch in wastewater under thermophilic condition. J Environ Manage 2003;69:149e56.

    [25] Zhu D, Wang G, Qiao H, Cai J. Fermentative hydrogenproduction by the new marine  Pantoea agglomerans  isolatedfrom the mangrove sludge. Int J Hydrog Energy 2008;33:6116e23.

    [26] de Vrije T, Budde MAW, Lips SJ, Bakker RR, Mars AE,Claassen PAM. Hydrogen production from carrot pulp bythe extreme thermophiles  Caldicellulosiruptor saccharolyticusand  Thermotoga neapolitana. Int J Hydrog Energy 2010;35:13206e13.

    [27] Ghosh D, Hallenbeck PC. Fermentative hydrogen yieldsfrom different sugars by batch cultures of metabolically

    engineered Escherichia coli DJT135. Int J Hydrog Energy 2009;34:7979e82.

    [28] Niela E, Buddeb M, Haasb G, Walb F, Claassenb P, Stamsa A.Distinctive properties of high hydrogen producing extremethermophiles,  Caldicellulosiruptor saccharolyticus andThermotaga elfii. Int J Hydrog Energy 2002;27:1391e8.

    [29] Kumar N, Das D. Enhancement of hydrogen production byEnterobacter cloacae  IIT-BT 08. Process Biochem 2000;35:

    589e93.[30] Ngo TA, Nguyen TH, Bui HTV. Thermophilic fermentative

    hydrogen production from xylose by Thermotoga neapolitanaDSM 4359. Renew Energy 2012;37:174e9.

    [31] Lin P, Whang L, Wu Y, Ren W, Hsiao C, Li S, et al. Biologicalhydrogen production of the genus Clostridium: metabolicstudy and mathematical model simulation. Int J Hydrog Energy 2007;32:1728e35.

    [32] Islam R, Cicek N, Sparling R, Levin D. Influence of initialcellulose concentration on the carbon flow distributionduring batch fermentation by Clostridium thermocellum ATCC27405. Appl Microbiol Biotechnol 2009;82:141e8.

    [33] Guo XM, Trably E, Latrille E, Carrère H, Steyer J. Hydrogenproduction from agricultural waste by dark fermentation: areview. Int J Hydrog Energy 2010;35:10660e73.

    [34] Magnusson L, Cicek N, Sparling R, Levin D. Continuoushydrogen production during fermentation of   a-cellulose bythe thermophillic bacterium Clostridium thermocellum.Biotechnol Bioeng 2009;102:759e66. 2008;102:759-66.

    [35] Whang L, Lin C, Liu I, Wu C, Cheng H. Metabolic andenergetic aspects of biohydrogen production of Clostridium tyrobutyricum: the effects of hydraulicretention time and peptone addition. Bioresour Technol2011;102:8378e83.

    [36] Jo JH, Lee DS, Park D, Park JM. Biological hydrogenproduction by immobilized cells of  Clostridium tyrobutyricum

     JM1 isolated from a food waste treatment process. BioresourTechnol 2008;99:6666e72.

    [37] Jo JH, Lee DS, Park JM. The effects of pH on carbon materialand energy balances in hydrogen-pr