trends in biotechnology 27 (2009) 287-297

Upload: diogo-lopes

Post on 04-Jun-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    1/11

    Advances in fermentative biohydrogenproduction: the way forward?Patrick C. Hallenbeck and Dipankar Ghosh

    Departement de Microbiologie et Immunologie, Universite de Montreal, CP 6128 Succursale Centre-ville, Montreal, Quebec H3C

    3J7, Canada

    A significant effort is underway to develop biofuels as

    replacements for non-renewable fossil fuels. Among the

    various options, hydrogen is an attractive future energy

    carrier due to its potentially higher efficiency of conver-

    sion to usable power, low generation of pollutants and

    high energy density. There are a variety of technologies

    for biological hydrogen production; here, we concentrate

    on fermentativehydrogen production and highlight some

    recently applied approaches, such as response surfacemethodology, different reactor configurations and organ-

    isms that have been used to maximize hydrogen pro-

    duction rates and yields. However, there are significant

    remaining barriers to practical application, such as low

    yields and production rates, and we discuss several

    methods, including two stage processes and metabolic

    engineering, that are aimed at overcoming these barriers.

    Introduction

    The joint challenges of anthropogenic climate change and

    dwindling fossil fuel reserves are driving intense research

    into alternative energy sources. One attractive avenue is to

    use a biological process to produce a biofuel. Among the

    various biofuel options, biohydrogen gas is an attractivefuture energy carrier due to its potentially higher efficiency

    of conversion to usable power, low to non-existent gener-

    ation of pollutants and high energy density. Biological

    hydrogen production processes can be classified into three

    major categories: biophotolysis of water using algae and

    cyanobacteria; photodecomposition of organic compounds

    by photosynthetic bacteria (photofermentation); and fer-

    mentative hydrogen production from organic wastes or

    energy crops (Box 1) [1]. However, low yields and pro-

    duction rates have been major barriers to the practical

    application of biohydrogen technologies.

    Intensive research on biohydrogen is underway, and in

    the last few years several novel approaches have beenproposed and studied to surpass these drawbacks. Here

    we concentrate mainly on fermentative hydrogen pro-

    duction, for which there has been a virtual explosion of

    research articles lately (but not hydrogen, not yet!), and

    many recent reviews on specific aspects are available [27].

    We discuss novel techniques, and improvements to exist-

    ing techniques, that have been used in attempts to increase

    hydrogen production rates and yields. Some of the dis-

    cussed techniques have permitted large increases in hydro-

    gen production rates, and this advance has brought

    fermentative hydrogen production to the point where pro-

    duction rates with real substrates (wastes) under realistic

    conditions are approaching practical levels. Of course, it is

    too premature to be able to exactly define what might be a

    practical level for biohydrogen production, but a rough

    yardstick can be obtained by examining a detailed process

    analysis of lignocellulosic ethanol production [8], where

    design metrics centered around a plant capable of produ-

    cing 50 to 100 kJ/l/h of ethanol; the hydrogen equivalent

    would be 5 to 10 l/l/h.The major remaining stumbling block is incomplete sub-

    strate conversion and the consequent low yields. Traditional

    techniques, such as the manipulation and optimization of

    bioprocess parameters or of the organisms used, have failed

    to achieve more than 25% conversion of substrate to H2.

    Indeed, this is a direct consequence of the thermodynamics

    of known metabolic processes, a fact pointed out theoreti-

    cally over thirty years ago [9] and now amply demonstrated

    in practice (for an example, see [10]). In addition to H2,

    hydrogen-fermenting microbes make other products to

    satisfy their metabolic needs and to further growth; these

    include acetate, which permits ATP synthesis, and a variety

    of reduced products (for example, ethanol, butanol and

    Review

    Glossary

    ANN (artificial neural network): a computational statistical modeling process

    based on the connectivity of biological neurons. It is nonlinear and adaptive

    (i.e. it learns).

    COD (chemical oxygen demand): commonly used to determine the amount of

    organic compounds in water.

    CSTR (continuous [flow] stirred tank reactor): the simplest reactor for

    continuous operation. Essentially, it is a chemostat.

    DOE (design of experiments): a structured method used in determining the

    relationship between a series of variables in a process and the output of that

    process. It should predict the important parameters and the relationship

    between them.

    FD (factorial design): a statistical method that shows the relationship between

    a response and two or more experimental factors that are varied simulta-

    neously.

    HRT (hydraulic retention time): the residence time of the liquid in a

    continuously fed reactor.

    MEC (microbial electrohydrogenesis cell): a bioelectrochemical cell, basically a

    modified microbial fuel cell (MFC), in which an applied voltage drives H2evolution.

    OLR (organic loading rate): the concentration of the substrate, usually

    expressed in g/l, that is fed to the reactor.

    Reverse micelle (also called inverted micelle): a micelle formed by amphiphilic

    surfactants in which the polar heads are oriented inwards encapsulating a

    polar (water) phase and the hydrophobic tails are in contact with a nonpolar

    solvent at the exterior.

    RSM (response surface methodology): a computational method that explores

    the relationship between variables that are subject to experimental manipula-

    tion and the output from the system, in this case hydrogen.

    UASB (upflow anaerobic sludge blanket): a type of reactor in which the

    biomass (functioning as a catalyst) is immobilized, thus rendering microbial

    growth independent of the liquid phase flow rate through the reactor.Corresponding author: Hallenbeck, P.C. ([email protected]).

    0167-7799/$ see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.tibtech.2009.02.004 Available online 28 March 2009 287

    mailto:[email protected]://dx.doi.org/10.1016/j.tibtech.2009.02.004http://dx.doi.org/10.1016/j.tibtech.2009.02.004mailto:[email protected]
  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    2/11

    butyric acid) that permit the reoxidation of NADH, which is

    necessary for continuing glycolysis (see Figures 1 and 2).

    Types and relative proportions of products depends upon

    the organism, environmental conditions (pH and the partial

    pressure of hydrogen) and the oxidation state of the

    substrate being degraded. Thus, fermentations are thought

    to be limited to producing at most either 2 H2/glucose

    (enteric bacterial type mixed acid fermentation) or 4 H2/

    glucose (clostridia-type fermentation) [2]. Therefore, hydro-

    gen yields are low, and two-thirds of the carbon and protons

    Box 1. Basic biohydrogen production technologies

    All biohydrogen technologies depend on either a hydrogenase or

    nitrogenase for hydrogen evolution and derive energy either directly

    from light energy or indirectly by consuming photosynthetically

    derived carbon compounds. Each approach has positive and negative

    aspects, and each has serious technical barriers that need to be

    overcome before it could become practical.

    Direct biophotolysis (Figure I)

    In this process, an organism, for example a green alga orcyanobacterium, that carries out plant-type photosynthesis uses

    captured solar energy to split water (producing O2) and reduce

    ferredoxin, which can in turn reduce a hydrogenase or nitrogenase,

    both of which are oxygen sensitive, producing H2.

    Advantages

    Abundant substrate (water); simple products (H2 and CO2).

    Disadvantages

    Low light conversion efficiencies; oxygen-sensitive hydrogenase;

    expensive hydrogen impermeable photobioreactors required.

    Photofermentation (Figure II)

    In this process, a non-sulfur purple photosynthetic bacterium,

    carrying out an anaerobic photosynthesis, uses captured solar energy

    to produce ATP and high energy electrons (through reverse electron

    flow) that reduce ferredoxin. ATP and reduced ferredoxin drive proton

    reduction to hydrogen by nitrogenase. These organisms cannotderive electrons from water and therefore use organic compounds,

    usually organic acids, as substrates.

    Advantages

    Complete conversion of organic acid wastes to H2and CO2; potential

    waste treatment credits.

    Disadvantages

    Low light conversion efficiencies; high energy demand by nitrogen-

    ase; expensive hydrogen impermeable photobioreactors required.

    Dark fermentation (Figure III)

    In this process, a variety of different microbes can be used

    anaerobically to breakdown carbohydrate-rich substrates to hydrogen

    and other products, principally acids (lactic, acetic, butyric, etc.) and

    alcohols (ethanol, butanol, etc.). Product distribution can be different

    dependent upon the microbe, oxidation state of the substrate and

    environmental conditions (pH, hydrogen partial pressure).

    Advantages

    No direct solar input needed; variety of waste streams/energy crops

    can be used; simple reactor technology.

    Disadvantages

    Low H2 yields; large quantities of side products formed.

    Figure II. Photofermentation.

    Figure III. Dark fermentation.

    Figure I. Biophotolysis.

    Review Trends in Biotechnology Vol.27 No.5

    288

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    3/11

    in the substrate are excreted as other products and wasted.

    Several approaches, including metabolic engineering and

    various two-stage processes, are being actively investigated

    in an attempt to solve this problem, and these are discussed

    below. First, we examine some traditional techniques that

    havebeen applied for biohydrogen production and discuss to

    what extent they have, or might, contribute to increasing

    biohydrogen production.

    Techniques for improved biohydrogen

    Reactor configurations

    Possible improvements to biohydrogen production have

    been sought through specialized bioreactor configurations

    (seeTable 1). This has led to systems with more robust,

    reliable performance that are stable over long periods of

    time (months) and resistant to short-term fluctuations in

    operational parameters. In addition, optimized volu-

    metric production rates could be obtained. Biohydrogen

    fermentations, as most other fermentations, can be car-

    ried out in either batch or continuous mode. Batch mode

    fermentations have been shown to be more suitable for

    initial optimization studies [4,11], but any industrially

    feasible process would most likely have to be performed on

    a continuous or at least semi-continuous (fed or sequen-

    cing batch) basis. Many studies have employed continuous

    stirred tank reactors (CSTRs, see Glossary) with either

    purified strains or microbial mixtures [4,5,11]. CSTRs,the

    most commonly used continuous reactor systems, offer

    simple construction, ease of operation and effective hom-

    ogenous mixing. However, in these reactors, hydraulicretention time (HRT, seeGlossary) controls the microbial

    growth rate and therefore HRTs must be greater than the

    maximum growth rate of the organism(s), because faster

    dilution rates cause washout. A second conceptual

    category of continuous flow reactors, characterized by

    the means of the physical retention of the microbial bio-

    mass, overcomes this problem and offers several advan-

    tages for a practical bioprocess. Because microbial growth

    and the concentration of microbial biomass are rendered

    independent of HRT, high cell concentrations can be

    achieved, fostering high volumetric production rates,

    and high throughput is possible, allowing the use (and

    Figure 1. Hydrogen production during mixed-acid fermentation by Escherchia coli. The mixed-acid fermentation carried out by E. coliis schematically shown. Different

    fermentation products can arise and their relative amounts can change depending upon the redox state of the substrate and the pH of the culture. Lactate formation and

    formate degradation to CO2 and H2 are induced at acidic pH, which serves to relieve the acid stress. Succinate production from phosphoenolpyruvate is also possible but

    usually represents only a minor fraction of total fermentation products. The relative proportions of products are balanced to maximize ATP production (via acetate

    production from acetyl-CoA by theptaackApathway) while at the same time reoxidizing NADH through the formation of reduced products, such as ethanol, to provide the

    NAD needed by the glycolytic pathway for further substrate utilization. Acetyl coenzyme A (acetyl-CoA) and formate are formed from pyruvate by pyruvateformate lyase.

    Pathways that have been eliminated with the aim of increasing hydrogen production are shown by dashed red crosses. Enzymes in the FHL (formate hydrogen lyase)

    pathway that have been upregulated to increase hydrogen production are shown by green arrows. The standard designations for the genes encoding the enzymes for the

    various steps shown are given in italics: adhE, alcohol dehydrogenase E; fdh, formate dehydrogenase; frdABCD, fumarate reductase A, B, C and D; fumBC, fumaratehydratase B and C; hyd1, (for convenience this stands for the operons necessary to express hydrogenase 1 activity); hyd2, (for convenience this stands for the operons

    necessary to express hydrogenase 2 activity); hyd3, (for convenience this stands for the operons necessary to express hydrogenase 3 activity); ldhA, lactate dehydrogenase

    A; mdh, malate dehydrogenase; pfl, pyruvateformate lyase; ptaackA, phosphotransacetylaseacetate kinase A.

    Review Trends in Biotechnology Vol.27 No.5

    289

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    4/11

    Figure 2. Different possible two-stage systems for the complete conversion of substrate. In the first stage (shown on the left) substrate is fermented to hydrogen and

    various fermentation products, which vary with organism and conditions. The two types of fermentation pathways are shown within the box on the left. Glycolysis of

    sugars produces pyruvate, from which hydrogen and other fermentation products are derived. Hydrogen is either produced from formate via an Ech (NiFe) hydrogenase

    (H2ase) in enterobacterial-type fermentation (right-hand side of the box) or from reduced ferredoxin via an FeFe hydrogenase in Clostridia-type fermentations (left-hand side

    of blue box). Additionally, hydrogen can be derived from NADH in Clostridia-type fermentations, although the molecular details are unclear and might involve an NADH-

    dependent FeFe hydrogenase, at least in a few cases (as shown by the dashed pink box). In both cases, the formation of acetate provides ATP synthesis. As discussed in the

    main text, acetate and the other fermentation products can then be fed into a second stage reactor (right panels), in which additional energy is extracted, either in the formof methane in the case of anaerobic digestion or hydrogen for photofermentations or MECs. In the latter case, additional energy is required for the second stage of the

    process: either light for photofermentation or electricity for MECs. Anaerobic digestion as a second stage is shown on the upper right side. Anaerobic digestors comprise a

    community of different organisms whose concerted action is to convert the various fermentation products of the first stage into methane. Archaea carry out

    methanogenesis (blue hexagon) and can directly use the acetate produced in the first stage. Associated syntrophic bacteria (purple trapezoid) oxidize other fermentation

    products from the first stage to substrates that can be used in methanogenesis, such as acetate, H 2and CO2. Photofermentation as a second-stage process is shown in the

    middle right scheme. Usually, a pure culture of a purple non-sulfur photosynthetic bacterium is used to convert the organic acids produced in the first stage dark

    fermentation to hydrogen. These organisms use the captured light energy to drive reverse electron flow (light blue cylinder), producing the necessary low potential

    electrons as well as the ATP required for the nitrogenase (N 2ase), which in this case is the hydrogen-producing enzyme system (shown in red). A variation on this theme

    would be inclusion of both metabolic types of organisms in a single photobioreactor. Indeed, it has been shown that glucose could be converted to hydrogen at a

    respectable yield of 7 H2/glucose (60%) by an immobilized co-culture consisting of Lactobacillus, which converted glucose to lactate, and the photosynthetic bacterium

    Rhodobacter sphaeroides, which converted the lactate to hydrogen [67]. A microbial electrohydrogenesis cell as a second stage is shown in the lower right scheme. Here,

    bacteria in the anodic chamber (shown in green) degrade the fermentation products from the first stage and donate electrons to the anode. Additional voltage is added via a

    power supply and hydrogen is evolved at the cathode in the cathodic chamber (shown in blue).

    Review Trends in Biotechnology Vol.27 No.5

    290

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    5/11

    treatment) of dilute waste streams with relatively small

    volume reactors.

    Indeed, many recent studies [3,5,1122] have shown

    that high volumetric production rates of hydrogen can

    be achieved in these reactors, and some examples are given

    in Table 1. Physical retention of microbial biomass has

    been accomplished by several different means, including

    the use of naturally forming flocs or granules of self-

    immobilized microbes, microbial immobilization on inertmaterials, microbial-based biofilms or retentive mem-

    branes. In fact, there are so many variations that there

    are almost as many different reactor types, with associated

    acronyms (UASB, FBBAC, AFBR, ASBR, CIGSB, FBR), as

    there are laboratories carrying out research in this area

    (Table 1). Unfortunately, this also creates a situation in

    which it is very difficult to ascertain whether the differ-

    ences in the various studies are due to different reactor

    configurations or to differences in operational parameters.

    Future studies could help to resolve this ambiguity by

    directly comparing different reactor configurations or, at

    the very least, by operating a novel configuration under

    conditions used in previous studies. One study that specifi-

    cally compared two different reactor types showed that

    granules gave superior performance over a biofilm, with an

    approximately threefold higher hydrogen production rate

    [21]. In general, the major advance of these types of

    reactors over a CSTR is their greatly increased volumetric

    production rate, which in some cases can be up to an

    astounding>50-fold larger (see Table 1). In addition, these

    reactor types present a favorable environment for the de-

    velopment and maintenance of mixed microbial consortia,

    which have advantages that are discussed below. A poten-

    tial problem with these types of reactors is the loss of

    hydrogen through the formation of methane. Because cell

    growth is no longer directly controlled by HRT in systems

    using retained microbial biomass, slow growing methano-gens can flourish even at high rates of liquid throughput. It

    should be noted that the hydrogen yields obtained with

    these systems are no greater than those achieved with

    CSTRs, and indeed there is no reason to think that the

    reactor type would influence the yield.

    The use of mixed microbial consortia

    Most of the reactors described above depend upon the

    formation of flocs or granules, which are macroscopic

    aggregates of microbial cells. The ability to form these

    particles is a rare trait in a pure culture, but it can be

    easily selected for when a mixed culture is used as an

    inoculum. There are also a number of other potential

    advantages of using microbial consortia instead of pure

    cultures. Industrial hydrogen fermentations will have to be

    carried out under non-sterile conditions using readily

    available complex feedstocks with only minimal pretreat-

    ment. Microbial consortia address these issues as they

    have been selected for growth and dominance under

    non-sterile conditions. As a complex community they are

    also likely to contain a suite of the necessary hydrolyticactivities, and they are potentially more robust to changes

    in environmental conditions[23].

    A large number of recent studies have examined hydro-

    gen production using microbial consortia, and some

    relevant examples are given in Table 1 (others can be

    found elsewhere[24]). The use of microbial consortia has

    indeed been proven useful in reactor systems that yield

    high volumetric rates of hydrogen production, as discussed

    above. However, there are also several issues associated

    with their use. As complex communities, their composition

    can vary over time, with changes in process parameters

    and from reactor to reactor, as was shown by molecular

    (16sRNA) studies[2528]. A possible way to overcome this

    issue might be to construct designer consortia [29] with

    the goal of creating a community of diverse members, each

    contributing a unique and essential metabolic capacity.

    The total community metabolic range would be greater

    than any individual member, while at the same time

    mutual interdependence would assure stable maintenance

    of individual members. However, little is known about the

    complex interactions that occur in natural consortia or how

    stable synthetic microbial communities could be built[30].

    Thus, much additional fundamental work might be

    required before practically useful synthetic hydrogen-pro-

    ducing consortia could become a reality. It seems that

    spatial organization within a consortium might be import-

    ant[31].

    Improving biohydrogen production by metabolic

    engineering of existing pathways

    Metabolic engineering has recently received increased

    attention in attempts to increase biohydrogen production,

    in particular hydrogen yields [3240]. Improvements in

    hydrogen production by existing pathways can be sought

    by increasing the flux through gene knockouts of competing

    pathways or increased homologous expression of enzymes

    involved in the hydrogen-generating pathways. The

    majority of studies employing these strategies have used

    the laboratory workhorse for metabolic engineering,

    Table 1. Available dark fermentation reactors

    Microorganisms Substrate Type of reactor H2 rate (l H2/l/h) Refs

    Sludge (wastewater treatment plant) M olasses Cont inuous stirred- tan k r eactor (CSTR) 0.20 [12]

    Sludge (wastewater treatment plant) Glucose Ana erobic sequencing batch rea ctor (ASBR) 0. 23 [13]

    Sludge (wastewater treatment plant) Sucrose Fixed bed bioreactor with activated carbon (FBBAC) 1.2 [14]

    Activated sludge and digested sludge Glucose Anaerobic fluidized bed reactor (AFBR) 2.4 [15]

    Sludge (wastewater treatment plant) Sucrose Upflow anaerobic sludge blanket reactor (UASB) 0.27 [16]

    Anaerobic sludge Sucrose Polymethymethacrylate (PMMA) immobilized cells 1.8 [17]

    Sludge (wastewater treatment plant) Sucrose Carrier-induced granular sludge bed (CIGSB) 9.3 [18]

    Sludge (wastewater treatment plant) Sucrose Fluidized bed reactor (FBR) 1.4 [19]

    Sludge (wastewater treatment plant) Glucose Anaerobic fluidi ze d bed reactor (AF BR) 7 .6 biofilm; 6.6 gra nul es [20]

    Sludge (wastewater treatment plant) Sucrose Continuously stirred anaerobic bioreactor (CSABR) 15.0 [19]

    Heat-treated soil Glucose Membrane bioreactor (MBR) 0.38 [21]

    Review Trends in Biotechnology Vol.27 No.5

    291

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    6/11

    Escherchia coli (Table 2, Figure 1). The rationale behind

    usingE. coli has been: (i) its genome can be easily manipu-

    lated; (ii) its metabolism is the best understood of all

    bacteria; and (iii) it readily degrades a variety of sugars.

    On the downside is the fact that its metabolism restricts

    hydrogen yields to 2 H2/glucose[2].

    Nevertheless, the use ofE. coli has served as a proof of

    principal that metabolic engineering can be used to achieve

    the maximal yields of hydrogen predicted by theory (see

    Table 2 and Figure 1 for the genes that have been tar-geted). The inactivation of pathways that drain the pyr-

    uvate pool, the source of electrons (via formate) for proton

    reduction, is expected to increase hydrogen production,

    and modest increases in H2yields were found when ldhA

    (lactate dehydrogenase) [3336] or frd (fumarate

    reductase) [35,36] were inactivated. In E. coli, formate,

    which is formed from pyruvate, is degraded to H2and CO2by the FHL (formatehydrogen lyase) pathway, and

    manipulations that led to an increase in the expression

    of enzymes, such as FdhH (formate dehydrogenase H) and

    Hyd3 (hydrogenase 3), involved in this pathway also

    increased H2 yields [3336]. Moreover, the total amount

    of hydrogen produced can be increased by inactivating the

    hydrogenases Hyd1 and Hyd2, which are potentially able

    to oxidize it [2]. These effects are additive, and mutated

    strains of E. coli harboring multiple relevant mutations

    demonstrated H2yields that were close to the theoretical

    maximum of 2 H2/glucose for the metabolic pathways that

    exist in this organism [2]. Although less studied so far,

    similar metabolic engineering strategies could be applied

    to other organisms. However, it would not seem possible to

    surpass present metabolic limits (2 to 4 H2/glucose) by this

    approach.

    Other attempts

    Modeling and optimization have been carried out in

    attempts to improve biohydrogen production. Rates andyields of hydrogen production, as for many other biopro-

    cesses, are a function of several variables, including pH,

    temperature, substrate concentration and nutrient avail-

    ability, among others. Although many studies have

    reported the effects of varying individual parameters

    one-at-a-time (for a partial compilation, see [11]), modeling

    and analysis could be used to determine the optimal values

    of the important relevant parameters. A variety of model-

    ing methods have been developed that are broadly

    applicable to a spectrum of diverse fields, including engin-

    eering, biology, environmental science, food processing and

    industrial processing, and their application in biohydrogen

    production is the subject of a very recent review [41].

    Although the metabolic pathways and fluxes are relatively

    well understood for a pure culture fermenting a defined

    substrate, this type of analysis can nevertheless have some

    value for modeling fermentations of complex substrates by

    microbial consortia that involve multiple metabolic types

    with unknown interactions.

    One type of treatment, principal component analysis

    [42], was recently used to assess the effects of pH, HRT and

    mixing on hydrogen production [43]. Two main types ofwidely used modeling approaches that are pertinent to

    BioH2 production are artificial neural networks (ANNs,

    seeGlossary) and design of experiments (DOE, see Glos-

    sary). ANNs were successfully used to model the results of

    hydrogen production and chemical oxygen demand (COD,

    see Glossary) removal with an upflow anaerobic sludge

    blanket reactor (UASB, see Glossary) [44] and an expanded

    granular sludge bed (EGSB) reactor[45]. DOE uses stat-

    istical modeling to analyze the relationship between a set

    of controllable independent experimental factors and to

    propose an appropriate design for their experimental ver-

    ification[46]. A common and powerful DOE approach is

    afforded by factorial design (FD) combined with response

    surface methodology (RSM, seeGlossary) [47]. FD inves-

    tigates how responses of multiple factors depend on each

    other and permits the identification of the most important

    parameters that control the process and the degree of

    interaction among them. FD approaches can be coupled

    to experimental systems by RSM, which uses a response

    function to fit the obtained experimental data to the theor-

    etical design.

    Several recent studies investigating dark fermentative

    hydrogen production have applied FD and RSM with the

    aim of optimizing hydrogen production (Table 3). Their

    results demonstrate that this type of analysis can indeed

    be used to optimize various aspects of a biohydrogen

    process, including pretreatment strategies [48,49], con-ditions for spore germination [50], micronutrient formu-

    lations [5153], substrate composition [34], carbon/

    nitrogen (C/N) and carbon/phosphate (C/P) [53]. The most

    useful application has been to determine the optimal oper-

    ating conditions with regard to HRT [13,55,56], pH [5761]

    and substrate concentration, that is, the organic loading

    rate (OLR, seeGlossary)[5661]. Although all mesophilic

    fermentations will require similar conditions, there are

    sufficient differences in some of the parameters to justify

    the need for optimization of a particular system.

    In addition, these methods could prove helpful in establish-

    ing process parameters for novel consortia or complex

    Table 2. Targets for modification of existing pathways in Escherchia coli

    Gene Function Mode of action Effect on H2productiona Refs

    ldhA Lactate dehydrogenase Inactivation eliminates a drain on the pyruvate pool Modest increase (2030%) [3336]

    frdBC Fumarate reductase Inactivation eliminates side reaction in EMP thereby increasing

    pyruvate pool

    Modest increase (2030%) [35,36]

    hycA Inhibitor of fhlA expression Inactivation increases synthesis of FhlA, increase in FHL complex No increase [35]

    fhlA Activator of Fhl expression Introduction of constitutive allele (FhlA*), increase in FHL complex Small increase (510%) [33,34]

    hyd1, hyd2 Uptake hydrogenases Inactivation prevents H2oxidation Modest increase (35%) [33,34]

    hyd3 Ech hydrogenase H2 evolving, FHL H2ase Large increase over low

    basal level

    [37]

    aLaboratory E. colistrains can vary widely in their unmodified hydrogen production capacity, possibly because some carry unrecognized mutations. All the strains reported

    here, with the exception of that used in [37], had relatively high native hydrogen evolution. Thus, in most cases, the reported increases are quantitatively significant.

    Review Trends in Biotechnology Vol.27 No.5

    292

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    7/11

    substrates (waste streams). Finally, it should be noted that

    although these methods can quickly optimize a particular

    fermentation, they cannot further improve the yield over

    what could be obtained by more classical methods of

    optimization, as can been seen by the results reported in

    Table 3.

    Reverse micelles have also been proposed as a possible

    means to increase hydrogen production. Reverse micelles,

    essentially self-assembling nano-structures in which an

    amphiphilic surfactant is used to entrap a hydrophilic

    biological catalyst, have been used primarily to investigate

    their ability to increase enzyme stability, enhance enzyme

    kinetics and promote partitioning of substrates or products

    between the organic and aqueous phase [62,63]. The encap-

    sulation of whole cells in reverse micelles has been much

    less studied, and the possible benefits to whole-cell-

    mediated catalysis are much less evident. Several studies

    have examined hydrogen production by different co-cul-

    tures in reverse micelles in small-scale batch studies[64

    66]. However, the reasons for the higher yields observed

    are unclear. Much work, including very substantial scale

    up, remains before this system could be exploited as apractical biohydrogen-producing system that could func-

    tion continuously on a long-term basis. At any rate, it is

    difficult to see how the use of reverse micelles could further

    improve yields or surpass the improvements in volumetric

    production rates already achieved by reactor configur-

    ations with retained microbial biomass (see above).

    Moving towards complete substrate conversion

    However, even with the improvements noted above, hydro-

    gen yields are restricted by the existing metabolic path-

    ways to either 2 H2/glucose (Enterobacteracae) or, at most,

    4 H2/glucose (Clostridia) at very low hydrogen partial

    pressures. The techniques already discussed have not,

    and cannot, increase yields beyond these limits. These

    yields are, however, practically unacceptable for several

    reasons. For one, they are not competitive with the pro-

    duction of other biofuels, such as biomethane or bioethanol,

    from the same starting materials, because the efficiency of

    substrate conversion to these biofuels is 80% or greater.

    Furthermore, incomplete substrate conversions during

    hydrogen fermentations (two-thirds of the substrate is

    used to make other products; see Box 1 and Figures 1

    and 2) lead to the production of a large amount of side

    products (i.e. COD), which subsequently need to be dis-

    posed of. Therefore, development of a practical biohydro-

    gen fermentation process requires either the introduction

    of additional pathways that would allow near stoichio-

    metric conversion (12 H2/glucose) within the microbial

    cell or the development of a two-stage system that would

    allow nearly complete energy recovery by introducing a

    second process to generate energy, preferably hydrogen,

    from the fermentation side products that are produced in

    the first hydrogen-producing dark fermentation stage.

    Metabolic engineering of new hydrogen-producing

    pathways

    One approach to surmounting the metabolic barrier is to

    construct in E. coli an alternative hydrogen-producing

    pathway from pathways found in other organisms; for

    example, the E. coli construct could express highly active

    FeFe hydrogenase (hydA), an enzyme that is normally

    responsible for hydrogen evolution in the fermenting Clos-

    tridia (Figure 2). However, the insertion of additional

    genes would be required because E. coli does not possess

    the genes hydE, hydF and hydG, which are required for

    FeFe hydrogenase maturation, nor any obvious pathways

    Table 3. Application of FD and RSM for biohydrogen production

    Bioprocess aspect Parameters studied Results RefsOptimized value H2 yield (mol H2/mol hexose) H2 rate (l/l/day)

    a

    Pretreatment Temp, cavitation, alkalinization Alkalinizat ion most ef fective 1.55 0.63 [48]

    Temp, pH, enzyme/substrate 35 8C, 7.0, 2.5% 0.98 2.1 [49]

    Seed spore germination pH, [substrate] 5.5, 5 g/l 0.66 2.2 [50]

    Micronutrient formulation [Fe2+] 3 mg/l 1.72 3.2 [51]

    180 mg/l 2.2 n/a [52]

    260 mg/l 2.6 0.58 [53]

    Substrate composition Food waste/sewage 87:13 0.9 3.0 [54]

    C/N, C/P 74, 450 n/a n/a [53]

    Fermenter operation HRT 48 h n/a n/a [55]

    16 h 1.6 2.9 [56]

    8 h 0.77 5 [13]

    pH 7.5b, 7.5c, 6.0d 1.8b, 1.5c, 1.6d 3.3b, 5.2c, 4.9d [57]

    6.1 n/a n/a [58]

    6.5 n/a n/a [59]

    5.5 1.74 0.5 [60]

    7.2 2.05 6.8 [61]

    Substrate concentration (g/l) 14.5e 1.6 2.95 [56]

    5b, 5c, 15d 1.8b, 1.5c, 1.6d 3.3b, 5.2c, 4.9d [57]

    21c n/a n/a [58]

    18c n/a n/a [59]

    25c 1.74 0.5 [60]

    28c 2.05 6.8 [61]

    n/a could not be calculated either due to the use of a complex substrate, or to lack of sufficient information.aNote that inTable 1the volumetric rates are per hour, whereas here they are per day.bLactose.cGlucose.dCheese whey powder.eSucrose.

    Review Trends in Biotechnology Vol.27 No.5

    293

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    8/11

    for reducing this hydrogenase, which functions with

    reduced ferredoxin. At a minimum, two more genes would

    probably be required to drive significant hydrogen evol-

    ution: fdxA (clostridial-type ferredoxin) and porA (pyru-

    vate ferredoxin/flavodoxin oxidoreductase). Recently, an

    artificial operon approach [38] was used to construct a

    ferredoxin-dependent NAD(P)HH2 pathway in E. coli[39]. Two compatible plasmids were used to introduce

    six genes (hydF, -E, -G and -A, fdxA and yumC, whichencodes an NAD(P)H:ferredoxin oxidoreductase) under the

    control of IPTG-inducible T7 promotors. Hydrogen pro-

    duction, which was sensitive to variation in hydrogen

    partial pressures as expected, was observed [39]. However,

    the levels of hydrogen produced were quite low and insig-

    nificant compared to what would be observed with a wild-

    type E. coli. However, the host strain, BL21(DE3), is

    incapable of hydrogen metabolism, providing a proof of

    principle for the introduction of a de novo hydrogen-produ-

    cing pathway. Additional strategies, including increasing

    the activity of FeFe hydrogenase and modifying the metab-

    olism to generate higher levels of reduced nucleotides,

    might bring further improvements. This area clearly needsfurther research efforts if present metabolic limits are to be

    surpassed, which would constitute a pivotal breakthrough.

    Some suggestions, based on computational modeling, of

    how to achieve this goal have recently been presented and

    include greatly increasing glucose oxidation through the

    pentose phosphate pathway with conversion of the gener-

    ated NADPH to NADH by transhydrogenase [40].

    Additionally, it has been suggested that limited respiration

    could be used to completely dissimilate the substrate-

    derived pyruvate, producing energy to drive reverse elec-

    tron flow and produce more than 4 H2/glucose)[1,2], but

    this has never been demonstrated experimentally.

    Hybrid two-stage systems

    The basic principle of a two-stage process is as follows: (i) in

    the first stage, the fermentation of the substrate to hydro-

    gen and organic acids takes place in a process that has been

    optimized using one or several of the advanced technol-

    ogies described above; (ii) then, in the second stage,

    additional gaseous energy, either methane or (more pre-

    ferably) hydrogen, is extracted from the effluent of the first

    stage reactor. Three different two-stage systems that are

    theoretically capable of complete energy extraction have

    been proposed and are under active investigation in small

    laboratory-scale experiments (Figure 2). One approach is

    to use a different reactor for the second stage that is

    operated under different conditions, such as higher pHand longer HRT, than the first reactor, thus favoring

    methanogenesis. Despite the disadvantage of generating

    two different gas streams, hydrogen and methane, in

    practical terms this might be useful because hydrogen-

    methane mixtures are cleaner fuels for internal combus-

    tion engines than methane alone in that they produce less

    NOX[68]. Although in the long term the goal should be to

    produce only a hydrogen stream, this hybrid two-stage

    system, producing both hydrogen and methane, is nearly

    ready to be put into practice and has already been scaled up

    to the pilot plant stage [69]. Such a two-stage system might

    offer several advantages over a traditional simple methane

    fermentation, including an effective solubilization of sub-

    strates such as organic solid wastes and increased toler-

    ance to high OLR. Several recent studies have reported the

    successful operation of such two-stage systems using

    actual wastes [6971], including a pilot plant-scale fermen-

    tation of kitchen waste that was able to generate relatively

    high production rates of hydrogen (5.4 m3/m3/day) and

    methane (6.1 m3/m3/day) while at the same time achieving

    removal of 80% of the COD. The efficiency of this process isdemonstrated by the fact that methane yields were twofold

    higher than a comparable single-stage process [69].

    Another possible process for increasing the overall

    energy extraction is the use of photofermentation in the

    second stage, with the aim of recovering additional hydro-

    gen from the products of a dark hydrogen-generating

    fermentation. In this reaction, non-sulfur purple photosyn-

    thetic bacteria capture light energy and use it to quanti-

    tatively convert organic acids to hydrogen. A great amount

    of research has been performed in this area and has been

    recently reviewed[72,73]. The main conclusions were that

    a large numberof these organisms are capable of degrading

    a variety of organic acids to hydrogen, sometimes achievinghigh yields with pure substrates. Indeed, in principle,

    photofermentations are capable of completely converting

    organic compounds into hydrogen, even against relatively

    high hydrogen partial pressures, because hydrogen evol-

    ution is driven by ATP-dependent nitrogenase and the

    requisite ATP is formed via the photosynthetic capture

    of light energy. In fact, in practice, nearly stoichiometric

    conversion of some organic acids to hydrogen can be

    obtained. Some recent studies, in which the second stage

    was actually fed effluent from a hydrogen-producing reac-

    tor, have demonstrated the feasibility of such a two-stage

    system, although yields were well below stoichiometric

    [7476]. However, low light conversion efficiencies coupled

    with low volumetric rates of production translate to a

    requirement for impossibly large surface areas. Although

    further research could very well help to increase yields, the

    technical hurdles to the practical application of such pro-

    cesses are severe given the rather low photosynthetic

    efficiencies of these organisms at moderate to high light

    intensities and the requirement for high cost photobior-

    eactors that are transparent and hydrogen-impermeable.

    Another approach employs microbial electrohydrogen-

    esis cells (MECs), in which electricity applied to a microbial

    fuel cell provides the necessary energy to convert organic

    acids, which are typical side products of a hydrogen fer-

    mentation, to hydrogen [7782]. This area has been

    thoroughly covered in an excellent recent review [83], soonly a summary of the major points is given here. This

    technology represents an elegant application of thermo-

    dynamic principles to the manipulation of microbial

    metabolism for use in the biotechnological production of

    fuels. MECs are in fact quite versatile; not only is it

    possible to select the appropriate bacteria during operation

    but also other substrates, such as wastewater, can be used.

    Thus, because of the versatility of the microbial community

    in the anodic chamber, MECs are not restricted to using

    fermentation products as substrate but can also, at least in

    principle, completely decompose glucose or glucose-con-

    taining substrates, such as cellulose, to hydrogen [80].

    Review Trends in Biotechnology Vol.27 No.5

    294

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    9/11

    Here, the mixed community carries out dark fermentation

    in situ, and the resulting fermentation products are used

    by the electrogenic members of the community. The elec-

    trogenic bacteria in the bioelectrochemical cells catabolize

    their substrate using the electrode (anode) as terminal

    electron acceptor. Supplementary voltage (>200 mV) is

    added to that already generated from the substrate

    (acetate 300 mV) to drive hydrogen evolution at the

    cathode. Thus, in principle, a second stage MEC after aninitial fermentative hydrogen stage could completely con-

    vert a substrate to hydrogen, achieving 12 H2/glucose with

    only a small investment of electricity. In the six years since

    MECs were first described, several innovations have been

    developed and consequently their performance has been

    dramatically improved. Currently reported volumetric

    rates (6 m3H2/m3 reactor/day) are much higher than

    those originally obtained. However, these rates are still

    almost two orders of magnitude lower than those obtained

    with dark fermentations and, moreover, have only been

    observed at high applied voltages (800 mV). It remains to

    be seen whether or not sufficiently high volumetric rates

    can be achieved with nominal electrical inputs. Thus,substantial challenges to the practical implementation of

    this promising technology remain, including the replace-

    ment of expensive platinum for the cathode, the increase of

    current densities and the reduction of the electrical input

    required.

    Outlook and future directions

    Many say that hydrogen is the fuel of the future; some add

    and it always will be! Indeed, in spite of a large amount of

    research in the past and at present, major hurdles remain

    to be overcome before a feasible practical process can be

    demonstrated for any approach to biological hydrogen

    production. Here we have concentrated on recent progress

    in hydrogen production by dark fermentation. As shown

    here, there have been substantial recent improvements in

    both the yield and volumetric production rates of hydrogen

    fermentations. Yet, to be practical, yields must consider-

    ably extend past the present metabolic limitation of 4 H2/

    glucose. Thus, the major outstanding question is Can a

    practical biological process be created to extract nearly all

    the H2from the substrate (12 H2/glucose)? Attempting to

    address this challenge should be the focus of future biohy-

    drogen research.

    Thermodynamics[9]and metabolic constraints suggest

    that it would be impossible to find an organism capable of

    the complete conversion of sugar-based substrates to

    hydrogen by fermentation and that human interventionis needed to solve this problem. Several possibilities exist,

    including further metabolic engineering or various two-

    stage systems, as discussed here. However, there are out-

    standing questions that must be answered for each of these

    different approaches if they are to succeed. For example,

    with regard to metabolic engineering, can novel pathways

    be introduced that are capable of the nearly complete

    proton extraction from the substrate and its reduction to

    H2? Although already demonstrable on a pilot scale,

    coupled hydrogen fermentation and methane digestion

    (which is one of the two-stage systems discussed here) does

    not produce a pure hydrogen stream and is therefore less

    desirable. In order for the other two-stage systems dis-

    cussed here to move forward, major impediments must be

    overcome. For a hybrid system using photofermentation,

    the key questions are: (i) can more efficient photosynthetic

    bacteria be created?; and (ii) can materials scientists

    develop sufficiently low cost transparent and hydrogen-

    impermeable photobioreactors? For a hybrid system using

    MECs, a different outstanding question arises: can MECs

    be developed that have sufficient current densities, requirelower voltages and use inexpensive cathodes? Solving the

    outstanding problem of increasing hydrogen yields could

    lead to the development of systems capable of the complete

    conversion of waste streams and energy crops to hydrogen,

    a potential sustainable fuel of the future.

    AcknowledgementsBiohydrogen research in the laboratory of P.C.H. is supported by NSERC

    (Natural Sciences and Engineering Council of Canada) and NRCan

    (Natural Resources Canada). D.G.s stay as a visiting scholar from the

    Indian Institute of Technology, Khargpur was supported by a GSEP

    (Graduate Student Exchange Program) grant from the Department of

    Foreign Affairs and International Trade, Canada.

    References1 Hallenbeck, P.C. and Benemann, J.R. (2002) Biological hydrogen

    production; fundamentals and limiting processes. Int. J. Hydrogen

    Energy 27, 11851193

    2 Hallenbeck, P.C. (2005) Fundamentals of the fermentative production

    of hydrogen. Water Sci. Technol. 52, 2129

    3 Hawkes, F.R. et al. (2007) Continuous dark fermentative hydrogen

    production by mesophilic microflora: principles and progress. Int. J.

    Hydrogen Energy 32, 172184

    4 Kapdan, I.K. and Kargi, F. (2006) Bio-hydrogen production from waste

    materials. Enz. Micro. Tech. 38, 569582

    5 Bartacek, J. (2007) Developments and constraints in fermentative

    hydrogen production. Biofuels Bioprod. Bioref. 1, 201214

    6 Das, D. and Veziroglu, T.N. (2008) Advances in biological hydrogen

    production processes.Int. J. Hydrogen Energy 33, 60466057

    7 Kraemer, J.T. and Bagley, D.M. (2007) Improving the yield fromfermentative hydrogen production. Biotechnol. Lett. 29, 685695

    8 Aden,A.et al. (2002) Lignocellulosic Biomass to Ethanol Process Design

    and Economics Utilizing Co-Current Dilute Acid Prehydrolysis and

    Enzymatic Hydrolysis for Corn Stover. National Renewable Energy

    Laboratory(http://www.nrel.gov/docs/fy02osti/32438.pdf)

    9 Thauer, R.K. et al. (1977) Energy conservation in chemotrophic

    anaerobic bacteria. Bact. Rev. 41, 100180

    10 Lee, H-S.et al.(2008) Thermodynamic evaluation on H2production in

    glucose fermentation. Environ. Sci. Technol. 42, 24012407

    11 Davila-Vazquez, G.et al. (2008) Fermentativebiohydrogen production:

    trends and perspectives. Rev. Environ. Sci. Biotechnol.7, 2745

    12 Ren, N.Q. et al. (2007) Assessing optimal fermentation type for bio-

    hydrogen production in continuous-flow acidogenic reactors.Bioresour.

    Technol. 98, 17741780

    13 Cheong, D-Y. et al. (2007) Production of biohydrogen by mesophilic

    anaerobic fermentation in an acid-phase sequencing batch reactor.Biotechnol. Bioeng.96, 421432

    14 Chang, J-S. et al. (2002) Biohydrogen production with fixed-bed

    bioreactors.Int. J. Hydrogen Energy 27, 11671174

    15 Zhang, Z-P. et al. (2007) Biohydrogen production in a granular

    activated carbon anaerobic fluidized bed reactor. Int. J. Hydrogen

    Energy 32, 185191

    16 Chang, F-Y. et al. (2004) Biohydrogen production using up-flow

    anaerobic sludge blanket reactor. Int. J. Hydrogen Energy 29, 3339

    17 Wu, K-J. and Chang, J-S. (2007) Batch and continuous fermentative

    production of hydrogen with anaerobic sludgeentrapped in a composite

    polymeric matrix. Process Biochem. 42, 279284

    18 Lee, K-S. et al. (2006) Improving biohydrogen production in a carrier-

    induced granular sludge bed by altering physical configuration and

    agitation pattern of the bioreactor. Int. J. Hydrogen Energy 31, 1648

    1657

    Review Trends in Biotechnology Vol.27 No.5

    295

    http://www.nrel.gov/docs/fy02osti/32438.pdfhttp://www.nrel.gov/docs/fy02osti/32438.pdfhttp://www.nrel.gov/docs/fy02osti/32438.pdf
  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    10/11

  • 8/14/2019 Trends in Biotechnology 27 (2009) 287-297

    11/11

    69 Ueno, Y.et al. (2007) Operation of a two-stage fermentation process

    producing hydrogen and methane from organic waste. Environ. Sci.

    Technol. 41, 14131419

    70 Liu, D.et al. (2006) Hydrogen and methane production from household

    solid wastein the two-stage fermentationprocess. Water Res. 40, 2230

    2236

    71 Ueno, Y.et al. (2007) Production of hydrogen and methane from organic

    solid wastes by phase-separation of anaerobic process. Bioresour.

    Technol. 98, 18611865

    72 Basak, N. and Das, D. (2007) The prospect of purple non-sulfur (PNS)

    photosynthetic bacteria for hydrogen production: the present state ofthe art. World J. Microbiol. Biotechnol. 23, 3142

    73 Redwood, M.D.et al. (2009) Integrating dark and light bio-hydrogen

    production strategies: towards the hydrogen economy. Rev. Environ.

    Sci. Biotechnol., DOI: 10.1007/s11157-008-9144-9 (http://www.

    springerlink.com/content/108922/)

    74 Tao, Y. et al. (2007) High hydrogen yield from a two-step process of

    dark- and photo-fermentation of sucrose. Int. J. Hydrogen Energy 32,

    200206

    75 Nath, K. et al. (2008) Kinetics of two-stage fermentation process

    for the production of hydrogen. Int. J. Hydrogen Energy 33, 1195

    1203

    76 Chen, C.Y.et al.(2008) Biohydrogen production using sequential two-

    stage dark and photo fermentation processes. Int. J. Hydrogen Energy

    33, 47554762

    77 Liu, H.et al.(2005) Electrochemically assisted microbial production of

    hydrogen from acetate. Environ. Sci. Technol. 39, 43174320

    78 Rozendal, R.A. et al. (2006) Principle and perspectives of hydrogen

    production through biocatalyzed electrolysis.Int. J. Hydrogen Energy

    31, 16321640

    79 Ditzig, J. et al. (2007) Production of hydrogen from domestic

    wastewater using a bioelectrochemically assisted microbial reactor.

    Int. J. Hydrogen Energy 32, 2296

    230480 Cheng,S. and Logan, B.E.(2007) Sustainableand efficient biohydrogen

    production via electrohydrogenesis.Proc. Natl. Acad. Sci. U. S. A.104,

    1887118873

    81 Rozendal, R.A. et al. (2008) Hydrogen production with a microbial

    biocathode. Environ. Sci. Technol. 42, 629634

    82 Tartakovskya, B. et al. (2008) High rate membrane-less microbial

    electrolysis cell for continuous hydrogen production. Int. J.

    Hydrogen Energy 34, 672677

    83 Logan, B.E. et al. (2008) Microbial electrolysis cells for high yield

    hydrogen gas production from organic matter. Environ. Sci. Technol.

    42, 86308640

    Review Trends in Biotechnology Vol.27 No.5

    http://dx.doi.org/10.1007/s11157-008-9144-9http://www.springerlink.com/content/108922/http://www.springerlink.com/content/108922/http://www.springerlink.com/content/108922/http://www.springerlink.com/content/108922/http://www.springerlink.com/content/108922/http://dx.doi.org/10.1007/s11157-008-9144-9