Next Article in Journal
Selection of Cutting Inserts in Dry Machining for Reducing Energy Consumption and CO2 Emissions
Next Article in Special Issue
Bioenergy and Food Supply: A Spatial-Agent Dynamic Model of Agricultural Land Use for Jiangsu Province in China
Previous Article in Journal
Adaptive Neuro-Fuzzy Inference Systems as a Strategy for Predicting and Controling the Energy Produced from Renewable Sources
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biohydrogen Production from Lignocellulosic Biomass: Technology and Sustainability

1
Government of India, Ministry of Science and Technology, Department of Scientific and Industrial Research (DSIR), Technology Bhawan, New Mehrauli Road, New Delhi 110016, India
2
Department of Chemical Engineering, Qatar University, Doha 2713, Qatar
3
Separation and Conversion Technologies, Flemish Institute for Technological Research (VITO), Mol 2400, Belgium
4
School of Environment and Sustainable Development, Central University of Gujarat, Gandhinagar 382030, Gujarat, India
*
Authors to whom correspondence should be addressed.
Energies 2015, 8(11), 13062-13080; https://doi.org/10.3390/en81112357
Submission received: 16 June 2015 / Revised: 4 November 2015 / Accepted: 10 November 2015 / Published: 17 November 2015
(This article belongs to the Special Issue Advances in Biomass for Energy Technology)

Abstract

:
Among the various renewable energy sources, biohydrogen is gaining a lot of traction as it has very high efficiency of conversion to usable power with less pollutant generation. The various technologies available for the production of biohydrogen from lignocellulosic biomass such as direct biophotolysis, indirect biophotolysis, photo, and dark fermentations have some drawbacks (e.g., low yield and slower production rate, etc.), which limits their practical application. Among these, metabolic engineering is presently the most promising for the production of biohydrogen as it overcomes most of the limitations in other technologies. Microbial electrolysis is another recent technology that is progressing very rapidly. However, it is the dark fermentation approach, followed by photo fermentation, which seem closer to commercialization. Biohydrogen production from lignocellulosic biomass is particularly suitable for relatively small and decentralized systems and it can be considered as an important sustainable and renewable energy source. The comprehensive life cycle assessment (LCA) of biohydrogen production from lignocellulosic biomass and its comparison with other biofuels can be a tool for policy decisions. In this paper, we discuss the various possible approaches for producing biohydrogen from lignocellulosic biomass which is an globally available abundant resource. The main technological challenges are discussed in detail, followed by potential solutions.

1. Introduction

Recent years have seen a rapid surge in research activities focusing intensely on alternative fuels in order to reduce the dependency on fossil fuels, mainly by providing local energetic resources. This is mainly due to two reasons, the first being that new fuels are needed to supplement and ultimately replace depleting oil reserves and secondly, fuels capable of low or nil CO2 emissions are urgently required to reduce the impact of global warming [1,2,3,4,5,6,7]. Hydrogen (H2), which can be used in fuel cells mainly to operate machines, is a fascinating alternative, particularly because its combustion provides high amounts of energy and water is the only reaction product. Among all biofuels, H2 has the highest gravimetric energy density at 141 MJ/kg. Despite this, its volumetric energy density, at only 12 MJ/m3 (at normal temperature and pressure) is low. This is an important aspect particularly in reference to transportation fuel. It is considered to be one of the cleanest energy carriers if produced using energy generated from renewable sources. In summary, H2 is interesting due to its potentially high efficiency of conversion to usable power, low generation of pollutants and high energy density [8]. Global H2 production today amounts to around 700 billion Nm3 and is based almost exclusively on fossil fuels [9]. However, for H2 to be accepted as a sustainable substitute for fossil fuels, it has to be produced from renewable feedstock other than fossil fuels [10].
Hydrogen has been suggested as the ideal fuel of the future. It is considered as one of the cleanest energy carriers to be generated from renewable sources [11]. It has a high energy yield (122 kJ/g) which is 2.75 times greater than hydrocarbon fuels. It can be easily used in fuel cells for generation of electricity. Though not a primary energy source, it serves as a medium through which primary energy sources (such as H2 produced from nuclear power and/or solar energy) can be stored, transported and utilized to fulfill our energy needs. The major problem facing H2 as a fuel is its unavailability in Nature. H2 can be produced safely, is environmentally friendly when combusted, and versatile i.e., has many potential energy uses, including powering non-polluting vehicles, heating homes and offices, and fueling aircraft. Current H2 production technologies such as steam reforming of natural gas, thermal cracking or coal gasification are not environmentally friendly. Biological H2 production is a promising alternative. There are two methods to produce H2 from microorganisms. The first method uses photosynthetic microorganisms such as bacteria or algae (photofermentative processes) and the second method uses fermentative organisms (dark fermentation processes). Fermentative H2 production has the advantage of producing H2 under mild conditions with the additional benefit of allowing residual biomass valorization. The dark fermentation process is more attractive as it has the potential to use wastewater and organic wastes and has higher production rates compared to photofermentative processes. So far, few studies have used real wastewater for the production of H2 due to inhibition by both substrate and/or product in the fermentation process [12]. Studies on bioH2 production have been focused on photodecomposition of organic compounds by photosynthetic bacteria, dark fermentation from organic compounds with anaerobes and biophotolysis of water using algae and cyanobacteria [13,14,15,16,17].
Lignocellulosic biomass is the most abundant in Nature and it is present in hardwood, softwood, grasses, and agricultural residues. The global annual yields of lignocellulosic biomass residues were estimated to exceed 220 billion tons, equivalent to about 60–80 billion tons of crude oil [12]. Lignocellulosic feedstocks consist mainly of glucose and xylose and thus microbial strains that can effectively degrade glucose and xylose are important for development of renewable H2 production processes [18]. Direct conversion of lignocellulosic biomass to H2 needs pretreatment to hydrolyze the incorporated heterogeneous and crystalline structure [19,20]. The lignocellulosic biomass hence presents an attractive, low-cost feed stock for H2 production.

Aim of the Paper

In recent past, several reviews have appeared which have discussed the prospects and challenges of biomass-based H2 [21,22]. Earlier, Kraemer and Bagley gave a thorough description of the yield improvement approaches in fermentative H2 production [23]. Wang and Wan summarized the main factors influencing fermentative H2 production [24]. A special issue of the journal International Journal of Hydrogen Energy, recently dealt with “Biohydrogen: From Basic Concepts to Technology” [25]. Biohydrogen can be generated by adopting different technologies and different technologies can perform differently. Thus, the aim of this paper is to discuss specifically the technological aspect of biohydrogen production from lignocellulosic biomass and its sustainability on the basis of a life cycle assessment (LCA).

2. Feed Stock for Biohydrogen Production

Glucose is the ideal substrate, but it is too costly at present. Many agricultural residues and food wastes are rich in carbohydrates that could serve as feedstock. Lignocellulosic biomass is another sustainable feedstock for H2 production [26]. The criteria for an ideal feedstock for sustainable H2 production, which include high carbohydrate content, minimum pre-treatment requirement, sustainable resources, low cost and sufficient concentration of carbohydrate for fermentative conversion, have been suggested by Bartacek [27]. The substrates usable for fermentative H2 production were further divided into four main groups, namely, pure substrates such as glucose; energy crops such as Miscanthus, solid wastes like food waste and industrial wastewaters such as wastewater from the pulp and paper industry.
A variety of substrates has been used as feedstock for H2 production. For example, the fermentation of household wastes under different temperature conditions has been well studied [28,29,30]. An increase in H2 yields as temperature increased to thermophilic regimes was reported.
Wastewaters and residual biomass with high carbohydrate content have also been demonstrated to be a suitable candidate for dark fermentation. This includes molasses [31,32] and cheese whey [33,34], which have been evaluated under continuous stirred tank reactor (CSTR) and immobilized system configurations. Besides, H2 production from soluble and particulate starch and cellulose [35,36], xylose [37], sugar beet [38], wastewater from a sugar beet refinery [39] and the bottom layer from a beer manufacturing plant [40] has also been demonstrated.

3. Technology

3.1. Biohydrogen Production Systems

The conventional methods for producing H2 gas include steam reforming of methane and hydrocarbons, non-catalytic partial oxidation of fossil fuels and autothermal reforming. However, most of these methods are energy intensive processes requiring high temperatures (>850 °C). A general scheme of H2 production from renewable sources is shown in Figure 1. Biological methods of H2 production are preferable to chemical methods because of the possibility to use sunlight, CO2 and organic wastes as substrates for environmentally benign conversions, under moderate conditions.
Figure 1. The main alternative methods of H2 production from energy sources.
Figure 1. The main alternative methods of H2 production from energy sources.
Energies 08 12357 g001
The biological production of H2 involves light-dependent methods: direct and indirect biophotolysis, and photo fermentation. The other routes are light-independent methods, including the dark fermentation process and water-gas shift reaction of photoheterotrophic bacteria. This biologically produced H2, generally referred to as “biohydrogen”, is characterized by low H2 yields which present a challenge for commercial applications.

3.1.1. Dark/Anaerobic Fermentation

Dark fermentation is one of the most common processes for bio-H2 production. Although only 15%–20% of the theoretical H2 potential of carbohydrates can be harvested, dark fermentation is considered as a promising process in a two phase anaerobic treatment system [23]. Also, since the CO2 produced in dark fermentation has already been fixed by the waste treated originally from the atmosphere, the emissions associated with global climate change are virtually zero [41]. Microbial species analysis of hydrogen-producing cultures (using anaerobic sludge as inoculum) shows the presence of Clostridium cellulosi, Clostridium acetobutylicum, Clostridium tyrobutyricum, Enterobacteriaceae and Streptococcus bovis [42] Fermentative H2 production usually proceeds from the anaerobic glycolytic breakdown of sugars. The theoretical complete oxidation of 1 mole of hexose to CO2 can produce 12 moles of H2. Nevertheless, the theoretical yield of H2 via acetic acid fermentation cannot be higher than 4 moles. Actual H2 yields are quite lower, typically ranging from 1.0 to 2.5 moles per mole of hexose consumed. Recently, Varanasi et al. reported production of 2.95 mol H2/mol hexose equivalent by thermophillic dark fermentation using cellubiose as substrate [43]. If butyric acid is produced as the major fermentation product instead of acetic acid, only 2 moles of H2 can be produced [44]. H2 yield is even lower when more reduced organic compounds such as lactic acid, propionic acid and ethanol are produced, because these metabolites represent end products of metabolic pathways that bypass the major H2-producing reaction [45]. Recently, it was concluded that to maximize net energy gain via dark fermentation, appropriate cultures capable of high-H2 yield have to be employed and the process has to be operated at near-ambient temperatures with the lowest feedstock concentration as possible [10]. In an experiment with Thermotoga neapolitana sparged with N2 and supplemented with 40 mM sodium bicarbonate a 2.8 and 2.7 mol/mol glucose yield of hydrogen with a lactic acid/acetic acid ratio of 0.26 was obtained, challenging the currently accepted dark fermentation model that predicts reduction of this gas when glucose is converted into organic products different from acetate [46]. Pradhan et al. reviewed the hydrogen production efficiency of a similar bacterium (Thermotoga neapolitana) with different feedstocks and found 1.9–3.5 mol H2/mol hexose yields achievable with a range of feedstocks and variable substrate loads [47]. Byproducts of the reactions are acetic acid, lactic acid and ethanol.

3.1.2. Photo Fermentation

Photo fermentation is carried out by purple non-sulfur (PNS) photosynthetic bacteria which can grow as photoheterotrophs, photoautotrophs or chemoheterotrophs [48]. These bacteria produce H2 under photoheterotrophic conditions (light, anaerobiosis, organic electron donor) [49]. The advantages of this process over photolysis of water using green algae and cyanobacteria, are that oxygen does not inhibit the process and that these bacteria can be used in a wide variety of conditions (i.e., batch processes, continuous cultures, and immobilized systems) [50]. The hydrogenase and nitrogenase enzymes produced in photosynthesis by green algae and photosynthetic bacteria, respectively, play a crucial role in biohydrogen production. The main PNS bacteria that participate in H2 production are Rhodospirillum rubrum, Rhodopseudomonas palustris, Rhodobacter sphaeroides O.U 001, Rhodobacter sphaeroides RV, Rhodobacter sulfidophilus and Rhodobacter capsulatus. Kapdan et al. used three different pure strains of Rhodobacter sphaeroides (RV, NRLL and DSZM) in batch experiments to select the most suitable strain [51]. R. sphaeroides RV resulted in the highest cumulative hydrogen gas formation (178 mL), hydrogen yield (1.23 mol·H2·mol−1 glucose) and specific hydrogen production rate (46 mL·H2·g−1 biomass·h−1) at 5 g·L−1 initial total sugar concentration among the other pure cultures. Using Rhodobacter capsulatus JP91, Keskin and Hallenbeck compare the photofermentative biohydrogen yield of different feedstocks in a batch culture experiment [52]. Overall yield of biohydrogen was 10.5, 8 and 14.9 mol H2/mol sucrose using beet molasses, black strap molasses and sucrose respectively. Optimization of process parameters such as availability of solar light, bioreactor configuration and proper C/N ratio in substrate (synthetic and derived from waste products) still needs to be studied at higher scale.

3.1.3. Combined Biotechnologies

Combination of two or more of the abovementioned techniques have also been studied for improved H2 yields. Theoretically 12 moles of H2 per mole of glucose can be generated by combining dark fermentation with photo fermentation (using PNS bacteria) [48]. For instance, Nath et al. studied combined dark and photo fermentation using glucose as substrate [53]. The effluent from the dark process (containing unconverted metabolites, mainly acetic acid) underwent photo fermentation by Rhodobacter sphaeroides in a column photo-bioreactor demonstrating the feasibility of this combination to achieve higher yields of H2 by complete utilization of the chemical energy stored in the substrate. A sequential process using glucose as substrate and an immobilized system for the photo fermentation step evaluating key factors such as diluted ratio of dark fermentation effluent, ratio of dark and photo fermentation bacteria, light intensity, and light/dark cycle has also been studied [54]. During the combined process, maximum total H2 yield was 5.374 moles of H2/moles of glucose. However, the sterilization step applied to the dark fermentation effluent may pose a constraint to a scale-up the process. The combined system can also be run in continuous mode and achieve more combined H2 yield [49]. These further combinations will reduce the overall cost of H2 production but more field studies are required to obtain an economical H2 production process.

3.1.4. Bioelectrochemical Production

Bioelectrochemical production of H2 is the latest technology using systems called microbial electrolysis cells (MEC). This is an emerging field where the oxidation of organic material is carried out by the bacteria present at the anode and results in formation of protons, CO2 and electrons (Figure 2). Protons migrate through a proton exchange membrane (PEM) to the cathode and the electrons are transported through the external circuit to the cathode [55]. By applying an external voltage of approximately 0.5–0.9 V, these electrons combine at the cathode with protons producing H2 gas (Table 1). The advantage here is the low energy consumption (0.3–0.9 V) necessary for microbial electrohydrogenesis to produce H2 in comparison to the theoretical minimum voltage of 1.23 V required for water electrolysis [56]. An overall scheme of H2 production from lignocellulosic biomass is shown in Figure 3. Figure 3 summarized the pathways for biohydrogen production by using lignocellulosic biomass, depending on the nature of feed (solid or liquid), pre-treatment methods were used and then followed by the dark fermentation.
Figure 2. Schematic of hydrogen production in MEC (Adapted from [56]).
Figure 2. Schematic of hydrogen production in MEC (Adapted from [56]).
Energies 08 12357 g002
Figure 3. General scheme for biohydrogen production from lignocellulosic biomass (adapted from [12]).
Figure 3. General scheme for biohydrogen production from lignocellulosic biomass (adapted from [12]).
Energies 08 12357 g003
Table 1. A comparison between the three major routes for biological hydrogen production (adapted from [57]).
Table 1. A comparison between the three major routes for biological hydrogen production (adapted from [57]).
Production RoutesMain ReactionH2 Production Rates (mmol/h·L)Remark
Direct photolysis2H2O + “light energy” → 2H2 + O20.07Similar to the processes found in plants and algal photosynthesis.
Photo fermentationC6H12O6 + 6H2O + “light energy” → 12H2 + 6CO2, ΔG0 = +3.2 kJ145–160Bacteria evolve molecular H2 catalyzed by nitrogenase under N-deficient conditions using light energy and reduced compounds (organic acids).
Dark fermentationPyruvate + CoA → acetyl-CoA + formate OR Pyruvate + CoA + 2Fd(ox) → Acetyl-CoA + CO2 + 2Fd (red)77H2 is produced by anaerobic bacteria, grown in the dark on carbohydrate rich substrate.
Table 2 shows the various substrates used for H2 production in different MEC volumes. The rate of H2 production differs with substrate due to their degradation pathways. To date MECs have shown H2 production from initial volumes ranging from 5 mL to 1000 L (pilot plant) reactors, which shows that MECs can be a better solution for producing H2 in the cathodic chamber while treating wastewater in the anodic chamber [58]. Still, there are many issues to be addressed for the long term real time application such as electrode and membrane stability for longer duration and reactor configuration design for higher volumes. Before scale up, mathematical models are required, which need to be validated first for the present lab scale studies and then on the basis of data obtained, higher volume MECs may be designed and validated.
Table 2. Various substrates used in MEC for hydrogen production (adapted from [59]).
Table 2. Various substrates used in MEC for hydrogen production (adapted from [59]).
SubstrateConcentration (g/L)Applied Voltage (V)MEC Volume (mL)Hydrogen Production Rate (m3 H2/m3/day)Reference
A de-oiled refinery wastewater04–10.7579% (Hydrogen production based on COD removal)[60]
Sodium Acetate10.6182.0[61]
Glucose20.6260.25 ± 0.03[62]
Glucose20.8260.37 ± 0.04[62]
Fermentation effluent10.6261.41[63]
Sodium Acetate10.6281.99 ± 0.02[64]
Sodium Acetate10.8283.12 ± 0.002[64]
Sodium Acetate10.5281.7[65]
Glucose10.5280.83 ± 0.18[66]
Glucose10.9281.87 ± 0.30[66]
Potato wastewater1.9–2.5 (COD)0.9280.74[67]
Swine wastewater2 (COD)0.5280.9–1.0[68]
Sodium Acetate10.6480.76[65]
Sodium Acetate10.776-[69]
Sodium Acetate10.82400.0231 ± 0.003[70]
Sodium Acetate114001.58[71]
Sodium Acetate20.65000.53[72]
Winery wastewater80.91000 Lt0.19 ± 0.04[58]
Sodium Acetate10.566000.02[56]

3.2. Microbiology of Biohydrogen Production

Perera et al. evaluated three main routes for biological H2 production [10]. These are (1) direct photolysis, in which cyanobacteria decomposes water to generate hydrogen and oxygen in presence of light; (2) photo fermentation, where anoxygenic photoheterotrophic bacteria utilizes organic feedstock to produce H2 in presence of light and (3) dark fermentation, in which anaerobic heterotrophic bacteria utilizes organic feedstock without any light to produce H2. A comparison of these three main routes is shown in Table 1.
The microbiology and biochemistry of dark fermentative H2 production was discussed in detail by Hawkes et al. [42]. H2 production in Clostridia is due to the presence of hydrogenase enzymes. These transfer electrons from reduced ferredoxin or NADH to protons to regenerate the oxidized forms (Fdox and NAD+) required so that glycolysis and oxidative decarboxylation of pyruvate can proceed to generate ATP.
Pure microbial cultures have mainly been used in lab-scale reactors for studying the effect of environmental and operational parameters on fermentation profiles and carbon metabolism. One of the successful tests using pure culture in a pilot-scale bioreactor using a non-sterilized feedstock employed Caldicellulosiruptor saccharolyticus [73]. However, most studies on H2 production on biowaste have been performed using mixed cultures under mesophilic conditions [74,75]. Only a few studies have focused on mixed thermophilic consortia [76,77]. It has been demonstrated that the extreme thermophile C. saccharolyticus can produce H2 from mono- and disaccharides [78]. Hexose is the predominant component in the cellulose hydrolysates. A highest H2 yield of approximately 83% of the theoretical value (4 mol·mol−1 hexose) has been reported using thermophilic anaerobic bacteria [78].

3.3. Limiting Factors in Biohydrogen Production Systems

The most challenging barrier of fermentative H2 production is its low H2 molar yield [26]. Thauer et al. predicted that 4 moles of H2 per mole of glucose is the biological maximum in Clostridial microbes if acetate is the only waste by-product [79]. In practice, even that figure is rarely achieved. A number of factors adversely affect and inhibit H2 fermentation [44]. H2 itself, when it reaches high concentrations not only makes its production thermodynamically unfavourable but also acts as an inhibitory agent as do other metabolic products, such as acetic acid and propionic acid [17,80]. Partial pressure of H2 is one of the most critical parameters in fermentative production of H2 as high H2 partial pressures make H2 production thermodynamically unfavourable. Removal of produced H2 from the liquid phase lowers the H2 partial pressure which in turn increases H2 yield [81]. Moreover, the H2 remaining in the system might be consumed by some bacteria [82]. Removal of dissolved H2 and reduction of H2 partial pressure can be achieved by nitrogen flushing, adsorption of H2 by metals and H2 stripping by boiling or by introduction of steam [83,84,85]. Low H2 partial pressure also needs to be maintained because hydrogenases (such as NiFe-hydrogenase) may re-oxidize the produced hydrogen into protons and electrons [86]. Gas sparging has proved to be an efficient method to maintain maximum hydrogen production even though it leads to biogas dilution and higher cost for hydrogen recovery [87]. Depending on the nature of the flushing gas, the flow rate and the reactor configuration, volumetric production of biogas up to 120% has been achieved [85,88]. Non-sparging techniques such as headspace modification under vacuum, high pressure or gas adsorption (reviewed in [87]), hydrogen-separating membranes [83] and using mechanical stirring [89,90], have also showed significant improvements in hydrogen yield. Argon has been often used to flush both oxygen and nitrogen and to keep a low H2 partial pressure in the reactors, but it increases production costs and hinders H2 purification [91]. Some researchers have reported reduced pressure and CO2 for flushing the headspace and maintaining low H2 partial pressure in dark fermentation [92,93], but the information on photofermentation is deficite. Montiel-Corona et al. suggested that flushing with Ar could be replaced with reduced pressure, which can be less expensive and practical for hydrogen recuperation [91]. Coupling the dark and photo fermentation showed an increased total hydrogen yield. One of the major drawbacks in coupling the dark and photo fermentation processes is the need of keeping apart the H2-producing microflora and the presence of NH4+, which may be naturally present in wastewater and may also be generated in the dark fermentation process when hydraulic retention time (HRT) is high enough to achieve protein degradation, especially when particulate substrates (as in the case of food wastes) are being considered, since HRT may be as high as 5 days [94].
In case of bioelectrochemical production of H2 in MEC, the main challenge is avoiding methane formation via methanogenesis [24], though researchers are now shifting more towards methane formation rather than H2 in these systems [95,96]. Another issue limiting the large-scale application of this technology is the use of precious metal catalyst such as platinum which is usually used on the cathode [97]. Though there have been efforts to use low cost materials such as stainless steel [98] and Ni-based electrodes [61], the results are much lesser from the targets.

3.4. Role of Metabolic Engineering

The application of genomic and molecular tools has made it possible to steer the metabolic pathways towards maximal H2 production and avoid waste and by-product accumulation. This is especially true when genetic engineering is conducted on cellulolytic microbes [26]. The main principles of genetic engineering include: (1) overexpression of cellulases, hemicellulases and lignases to maximize substrate availability, (2) elimination of H2-consuming hydrogenases and (3) overexpression of H2-producing hydrogenases [53]. Metabolic engineering modifications have been used to increase H2 production in fermentative systems [99]. These include over-expression of H2-evolving enzymes [100], the knockout of metabolic pathways that compete for reducing equivalents [81] and the introduction/over-expression of genes (cellulases, hemicellulases and lignases) to enhance carbohydrate availability to the cell [101]. Inactivation of the gene lactate dehydrogenase (ldhA) in E. coli by introducing mutations could lead to a modest increase (20%–45%) in net hydrogen production (reviewed in [102]).
Ryu et al. combined several known approaches to construct a superior hydrogen-producing strain of the purple nonsulfur photosynthetic bacterium Rhodobacter sphaeroides HPCA* (mutant expressing NifA L62Q) [103]. In this strain maximum hydrogen levels are reached almost twice as fast as in wild type cells and final hydrogen levels are ~39% higher than in the wild type as well. As increased number of genomes for H2 producing microorganisms are sequenced and compared and as more specific enzymes are functionally characterized, the distinctive metabolic strategies used and enzymological contexts through which H2 evolution is controlled in different organisms will become clearer. This will allow researchers to construct more effective strategies to modulate competing pathways, and help in the designing molecular engineering strategies leading to enhanced H2 evolution.

4. Kinetic Models for Hydrogen Production by Fermentation

Different factors such as substrate and inhibitor concentrations, temperature, pH and reactor type affect H2 production by fermentation. Modeling of the H2 production is very important to improve, analyze and predict H2 production during fermentation. Mathematical models include the kinetic of cell growth and product(s) formation, substrate utilization and inhibition. In addition some models are developed to describe the effect of pH, temperature and dilution rate on H2 production. The obtained model kinetic constants can be used in the design, operation and optimization of the fermentative H2 production process. Different kinetic models have been proposed to describe growth of H2 producing bacteria, substrate degradation and H2 production. H2 production is reported as growth associated product.
Monod (or Michaelis–Menten equation) (Equation (1)) which is an unstructured, non-segregated model of microbial growth, fits a wide range of data. The kinetic constants of this equation, KS and μmax, can be obtained by linear regression. Wang and Wan reported on previous studies using a Monod model to describe H2 production with time in bio-H2 fermentation [104]:
μ = 1 X d X d t = μ m S K s + S
where µ is the specific growth rate, X is the biomass concentration, S is the substrate concentration, Ks is the saturation constant, µm is the maximum specific growth rate.
Recently, the logistic model (Equation (2)) became the most popular in describing cell growth. This equation has a sigmoidal shape that includes the lag phase, exponential and stationary phase of the batch growth:
μ = 1 X d X d t = μ m ( 1 X X m )
where Xm is the maximum biomass concentration.
At high substrate concentration, the cell growth is inhibited and production of H2 is reduced. Different substrate inhibition models have been proposed. The Haldane-Andrew model (Equation (3)) is widely used to describe the substrate dependence of the specific growth rate of H2 fermentations. Wang and Wan have reported that previous studies used an Andrews model to describe H2 production with time [104]. Other substrate inhibition models are used in the literature such as modified Han-Levenspiel model (Equation (4)):
μ = 1 X d X d t = μ m S K s + S + S 2 K i
where Ki is the inhibition constant.
The presence of other inhibitors such as salts and the product cause reduction of H2 production. Some models have been proposed to describe the effect of inhibitors such as the modified Han-Levenspiel model (Equation (4)):
μ = 1 X d X d t = μ m ( 1 C C m )
where C is the inhibitor concentration, Cm is the maximum inhibitor concentration or the concentration of inhibitor above which there is no biomass growth
The modified Gompertz model (Equation (5)) is widely used to describe the progress of cumulative H2 production in batch fermentations [104]:
H t = H max exp { exp [ R max × e H max ( λ t ) + 1 ] }
where Ht is the cumulative volume of H2 produced at any time (mL), Hmax is the gas production potential (mL), Rmax is the maximum gas production rate (mL/h), λ is the lag time (h). t is the incubation time (h)
The Luedeking-Piret model (Equation (6)) has been widely used to describe the relation between cell growth rate and H2 production:
d P d t = Y P / X d X d t + β X
where P is the product, YP/X is the growth associated yield coefficient; β is the non-growth associated product yield coefficient.
Wang and Wan reported that previous studies used the Luedeking–Piret model to relate cell growth rate and H2 production rate [104]. The effect of temperature on the fermentative H2 production has been widely described using the Arrhenius model, while the effect of pH on the substrate consumption rate is described by an Andrew model using the concentration of H+ as the limiting substrate concentration. According to this model, the rate of substrate consumption passes through maximum with increasing H+ concentration.

5. Sustainability and Life Cycle Assessment

The concept of sustainable development is an attempt to combine growing concerns about a range of environmental issues with socio-economic issues and implies smooth transition to more effective technologies from a point view of an environmental impact and energy efficiency [105,106]. H2 can be considered one of the pillars of a future sustainable energy system [107]. H2 production could be a possible avenue for the large-scale sustainable generation of H2 needed to fuel a future H2 economy [106]. Despite its many obvious advantages, there remains a problem with storage and transportation. Pressurized H2 gas occupies a great deal of volume compared with other fuels. For example, gasoline that with equal energy content, needs about 30 times less volume at 100 bar gas pressure. Due to its high explosivity there are also obvious safety concerns with the use of pressurized or liquefied H2 in vehicles as well as additional energy use for pressurizing or liquefaction. Furthermore, the overall energy balance of using H2 as vehicle fuel does indeed seem to be less beneficial than gasoline, but being the only non-carbon fuel it may still make sense to produce H2 from waste streams if some of the obstacles can be solved and it can be used effectively for energy production to feed into grid or to use in stationary requirements, e.g., industries, etc.
Though this paper is focused on bio-H2 production from lignocellulosic biomass, it is important to compare it to other production methods using various substrates. Such a comparison has been made in Table 3 by presenting the various H2 production systems, which show different H2 yields from different feedstocks by adopting different production systems. Therefore, life cycle assessment (LCA) could be a tool to scrutinize the best H2 production system for a particular feedstock in terms of environmental impact and indirect natural resource costs towards different services and commodities [108]. LCA allows the possibility of comparing different H2 production approaches and identifying the environmental “hot spot” of the whole process, which helps in development of a sustainable H2 production process [106,109]. Investigations of the environmental benefits and impacts from a life cycle perspective are scarce. Only a few LCA-studies have been performed specifically on H2 production. The feedstocks investigated so far are steamed potato peel, wheat straw and sweet sorghum stalks [110,111,112].
Table 3. Comparison of different biohydrogen production systems.
Table 3. Comparison of different biohydrogen production systems.
ReactorFeed StockMaximum H2 YieldReference
Fermentation
Dark fermentation
CSTRStarch0.52 L/h/L and 13.2 mmol H2/g total sugar[113]
BatchGlycerol0.41 mol H2/mol glycerol[114]
FBRSucrose4.26 mol H2/mol sucrose[115]
BatchFood waste593 mL H2/g carbohydrate[116]
Fed-batchSwine manure18.7 × 10−3 g H2 per g TVS[117]
BatchSucrose4.3 mol H2/mol sucrose[118]
BatchFructose, sorbitol, glucose1.27, 1.46 and 1.51 mol H2/substrate[119]
Fed-batchStarch, glucose465 mL H2/g starch, 3.1 mol H2/mol glucose[120]
BatchFood waste39.14 mL H2/g food waste (219.91 mL H2/VSadded)[121]
BatchCrude Glycerol64.24 mmol H2/L and 5.74 mmol H2/g COD consumed[122]
BatchDistillery wastewaters1 L H2/L medium[123]
BatchCheese whey94.2 L H2/kgvs[124]
BatchWater hyacinth (leaves and stems)76.7 mL H2/g TVS was obtained at 20 g/L of water hyacinth[125]
Batchwaste ground wheat solutionSHPR = 25.7 mL H2/g cells/h[126]
Photo fermentation
CSTRSucrose5.81 mol H2/mol hexose[127]
Fed-batch operationWheat starch201 mL H2 g/L starch[128]
BatchMolasses0.50 mmol H2/Lc h[129]
BatchBeet molasses10.5 mol H2/mol sucrose[52]
BatchBlack strap8 mol H2/mol sucrose[52]
BatchSucrose14 mol H2/mol sucrose[52]
BatchGround wheat starch46 mL H2/g biomass/h, 1.23 mol H2/mol glucose[51]
Batchlignocellulose-derived organic acids7 mL H2/mL of the fermentation effluent[130]
Photosynthesis
Direct Photolysis
BatchLactate0.07 mmol H2 (l × h) or 54 mL/h·g dry weight[131]
Indirect Photolysis
Batcharabinose and xylose14.55 mmol/g (arabinose); 13.73 mmol/g (xylose)[132]
Thermochemical
Gasification
Continuous supercritical water gasificationglucose10.5–11.2 mol/mol glucose[133]
Partial Oxidation
Batchmunicipal sludgeNot reported the amount[134]
Steam reforming
molten carbonate fuel cell (MCFC) systemethanol5 mol H2/mol fed ethanol[135]
Cracking
fixed-bed quartz micro reactorMethane500 µmoles/min[136]
Pyrolysis
stainless steel tank reactorBiomass (redwood sawdust; cole stalk and rice husk) feed65.39 g/Kg biomass for redwood sawdust; 40.0 g/Kg biomass for cole stalk and rice husk[137]
Thermoelectrochemical
membrane electrode assemblysulfur dioxide0.4 A/cm2 at 0.835 V (H2 production rate did not reported)[138]
membrane electrode assemblyanhydrous hydrogen bromide2.0 A/cm2 at 1.91 V (H2 production rate did not reported)[138]
Electrochemical
Electrolysis
The BiOx−TiO2 electrode and stainless steel (SS, Hastelloy C-22) were used as an anode and a cathode in the electrochemical system, respectivelyarsenite (As(III))9.4 µmoles/min[139]
Photoelectrolysis
The TiO2(ns) was prepared in the form of a sol-gelphotoelectrode system TiO2(ns)–VO26 L·h−1·m−2 for the TiO2(ns); 13.0 L·h−1·m−2 for the TiO2(ns)–VO2 photoelectrode[140]
In connection with a European research study, HYVOLUTION, the life cycle environmental impacts of pilot production of H2 through thermophilic fermentation, and photo fermentation of potato peel was compared to production of H2 from natural gas through steam methane reforming (SMR) [112]. It was demonstrated that the bio-H2 production had approximately 5.7 times higher environmental impacts (negative impacts on the environment) than a centralized SMR. The processes involved in steam (pretreatment), phosphate buffer (used in photo fermentation) and potassium hydroxide (used in thermophilic fermentation), were the main causes of the environmental impact (98.3%). Recirculation of the sewage reduces the environmental impacts considerably to having only approximately two times more environmental impact than SMR. If instead biomethane were produced for use in the SMR the environmental impact would be reduced to less than 1/3 of the traditional SMR [112]. On the other hand alternative use of the peel would be as animal feed and Djomo et al. showed that the production of bio-H2 is more beneficial than the use as animal feed by a factor of 2–3 [110]. In a more recent study Djomo and Blumberga investigated potential differences in environmental performance between the three different feedstocks [111]. They performed a “well-to-tank” study i.e., the system boundary is at supplying H2 to road vehicles meaning that the combustion and transportation of H2 in the vehicles, was not included. Further, the production of feedstock was excluded as they are considered wastes. Their conclusion is in contrast to the earlier study they find that H2 produced from any of the feedstock reduced GHG-emissions by approximately 55% compared to SMR and a few percent less for gasoline. When the subsequent use of the remains from the H2 production were considered as animal feed, an environmental benefit could be observed. The energy ratio calculated was 1.08–1.17, i.e., the energy gain is between 8 and 17%. Though steamed potato peel was slightly better, no significant environmental differences were observed between the feedstocks [111]. The results compare well with those of Manish and Banerjee who investigated the energy balance of H2 and found an energy ratio of 3.1 (excluding the gas treatment and the compressing) [57].
The conclusion from these studies from an environmental view point is that the production of H2 for renewable energy production from potato peel could be preferred to using SMR or as direct animal feed due to the lesser environmental impacts. The LCA studies can further be used for identification of the main environmental improvements in the technology development (e.g., recirculation of the sewage and reuse of the remains for animal feed). The LCA of H2 is very important before taking them into consideration for commercial scale production and policy decisions on H2 promotion.

6. Future Directions and Perspectives

One option proposed to lower feedstock costs is to identify microbes that can directly utilize hemicellulose and cellulose [26]. This would eliminate the need for cellulase enzymes and simplify biomass pretreatment. As cellulose is the most abundant biopolymer in the world [141], its bioconversion provides a viable approach to produce renewable H2 from organic matter. The combined dark fermentation coupling with photo fermentation, or dark fermentation coupling with bioelectrohydrogenesis is a promising H2 production process from lignocellulosic biomass if the technological barriers can be overcome [12]. Overall, to develop a mature H2 production technology, bioconversion performance from lignocellulosic biomass need to be further improved in terms of production rates, cost-effectiveness, and system scale-up. Based on the limited number of LCA studies done on H2 production, it can be assumed that the bioconversion of lignocelluloses-to-H2 on industrial scale is a feasible option to produce H2 via biotechnology. However, more in-depth studies need to be carried out to confirm this.

7. Conclusions

Although considerable progress has been made on H2 production from lignocellulosic biomass, several challenges remain for its commercial application. Among the various techniques available for H2 production from lignocellulosic biomass, dark fermentation seems to have an edge over the others and is the closest to commercialization. Photo fermentation is the next best option, though it has to overcome the problems associated with reactor design and operation. Bioelectrochemical H2 production is still in its infancy and needs much more research and development. The kinetic models for H2 production provide insights on substrate utilization and factors limiting higher yields. The models will help in scale up studies for validating the proposed data and later on with the experimental data. The few environmental assessment studies performed from a LCA perspective show that H2 production from lignocellulosic biomass also may be preferable to other renewable energy production pathways. Such studies can furthermore help identifying technological improvement options. The results of LCA studies could also help policy makers in taking decision on policies related to promotion of renewable energy.

Author Contributions

Anoop Singh and Deepak Pant originated the manuscript idea, created the outline, wrote parts of the manuscript and did the final editing. Surajbhan Sevda and Ibrahim M. Abu Reesh wrote about the modelling aspect and prepared the final revised draft. Karolien Vanbroekhoven and Dheeraj Rathore contributed to different sections of the manuscript mainly microbial electrolysis and sustainability.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Prasad, S.; Singh, A.; Joshi, H.C. Ethanol as an alternative fuel from agricultural, industrial and urban residues. Resour. Conserv. Recycl. 2007, 50, 1–39. [Google Scholar] [CrossRef]
  2. Prasad, S.; Singh, A.; Joshi, H.C. Ethanol production from sweet sorghum syrup for utilization as automotive fuel in India. Energy Fuels 2007, 21, 2415–2420. [Google Scholar] [CrossRef]
  3. Pant, D.; Van Bogaert, G.; Diels, L.; Vanbroekhoven, K. A review of the substrates used in microbial fuel cells (MFCs) for sustainable energy production. Bioresour. Technol. 2010, 101, 1533–1543. [Google Scholar] [CrossRef] [PubMed]
  4. Pant, D.; Singh, A.; van Bogaert, G.; Gallego, Y.A.; Diels, L.; Vanbroekhoven, K. An introduction to the life cycle assessment (LCA) of bioelectrochemical systems (BES) for sustainable energy and product generation: Relevance and key aspects. Renew. Sustain. Energy Rev. 2011, 15, 1305–1313. [Google Scholar] [CrossRef]
  5. Singh, A.; Smyth, B.M.; Murphy, J.D. A biofuel strategy for Ireland with an emphasis on production of biomethane and minimization of land-take. Renew. Sustain. Energy Rev. 2010, 14, 277–288. [Google Scholar] [CrossRef]
  6. Singh, A.; Pant, D.; Korres, N.E.; Nizami, A.; Prasad, S.; Murphy, J.D. Key issues in life cycle assessment of ethanol production from lignocellulosic biomass: Challenges and perspectives. Bioresour. Technol. 2010, 101, 5003–5012. [Google Scholar] [CrossRef] [PubMed]
  7. Singh, A.; Nigam, P.S.; Murphy, J.D. Renewable fuels from algae: An answer to debatable land based fuels. Bioresour. Technol. 2011, 102, 10–16. [Google Scholar] [CrossRef] [PubMed]
  8. Hallenbeck, P.C.; Ghosh, D. Advances in fermentative biohydrogen production: The way forward? Trends Biotechnol. 2009, 27, 287–297. [Google Scholar] [CrossRef] [PubMed]
  9. Ball, M.; Wietschel, M. The future of hydrogen—Opportunities and challenges. Int. J. Hydrog. Energy 2009, 34, 615–627. [Google Scholar] [CrossRef]
  10. Perera, K.R.J.; Ketheesan, B.; Gadhamshetty, V.; Nirmalakhandan, N. Fermentative biohydrogen production: Evaluation of net energy gain. Int. J. Hydrog. Energy 2010, 35, 12224–12233. [Google Scholar] [CrossRef]
  11. Kovacs, K.; Maroti, G.; Rakhely, G. A novel approach for biohydrogen production. Int. J. Hydrog. Energy 2006, 31, 1460–1468. [Google Scholar] [CrossRef]
  12. Ren, N.; Wang, A.; Cao, G.; Xu, J.; Gao, L. Bioconversion of lignocellulosic biomass to hydrogen: Potential and challenges. Biotechnol. Adv. 2009, 27, 1051–1060. [Google Scholar] [CrossRef] [PubMed]
  13. Kumar, N.; Das, D. Enhancement of hydrogen production by Enterobacter cloacae IIT-BT 08. Process Biochem. 2000, 35, 589–593. [Google Scholar] [CrossRef]
  14. Chen, C.C.; Lin, C.Y.; Chang, J.S. Kinetics of hydrogen production with continuous anaerobic cultures utilizing sucrose as the limiting substrate. Appl. Microbiol. Biotechnol. 2001, 57, 56–64. [Google Scholar] [PubMed]
  15. Ginkel, S.V.; Sung, S.; Lay, J.J. Biohydrogen production as a function of pH and substrate concentration. Environ. Sci. Technol. 2001, 35, 4726–4730. [Google Scholar] [CrossRef] [PubMed]
  16. Xing, D.; Ren, N.; Li, Q.; Lin, M.; Wang, A.; Zhao, L. Ethanoligenens harbinense gen. nov., sp. nov., isolated from molasses wastewater. Int. J. Syst. Evolut. Microbiol. 2006, 56, 755–760. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, L.; Zhou, Q.; Li, F. Avoiding propionic acid accumulation in the anaerobic process for biohydrogen production. Biomass Bioenergy 2006, 30, 177–182. [Google Scholar] [CrossRef]
  18. Ren, N.; Cao, G.; Guo, W.; Wang, A.; Zhu, Y.; Liu, B.; Xu, J. Biological hydrogen production from corn stover by moderately thermophile Thermoanaerobacterium thermosaccharolyticum W16. Int. J. Hydrog. Energy 2010, 35, 2708–2712. [Google Scholar] [CrossRef]
  19. De Vrije, T.; de Haas, G.; Tan, G.B.; Keijsers, E.R.; Claassen, P.A.M. Pretreatment of Miscanthus for hydrogen production by Thermotoga elfii. Int. J. Hydrog. Energy 2002, 27, 1381–1390. [Google Scholar] [CrossRef]
  20. Ntaikou, I.; Gavala, H.N.; Kornaros, M.; Lyberatos, G. Hydrogen production from sugars and sweet sorghum biomass using Ruminococcus albus. Int. J. Hydrog. Energy 2008, 33, 1153–1163. [Google Scholar] [CrossRef]
  21. Kapdan, I.K.; Kargi, F. Bio-hydrogen production from waste materials. Enzym. Microb. Technol. 2006, 38, 569–582. [Google Scholar] [CrossRef]
  22. Kalinci, Y.; Hepbasli, A.; Dincer, I. Biomass-based hydrogen production: A review and analysis. Int. J. Hydrog. Energy 2009, 34, 8799–8817. [Google Scholar] [CrossRef]
  23. Kraemer, J.T.; Bagley, D.M. Improving the yield from fermentative hydrogen production. Biotechnol. Lett. 2007, 29, 685–695. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, J.; Wan, W. Factors influencing fermentative hydrogen production: A review. Int. J. Hydrog. Energy 2009, 34, 799–811. [Google Scholar] [CrossRef]
  25. Cournac, L.; Sarma, P.M.; Fontecave, M. Biohydrogen: From Basic Concepts to Technology. Int. J. Hydrog. Energy 2010, 35, 10638. [Google Scholar] [CrossRef]
  26. Turner, J.; Sverdrup, G.; Mann, M.K.; Maness, P.C.; Kroposki, B.; Ghirardi, M.; Evans, R.J.; Blake, D. Renewable hydrogen production. Int. J. Energy Res. 2008, 32, 379–407. [Google Scholar] [CrossRef]
  27. Bartacek, J.; Zabranska, J.; Lens, P.N. Developments and constraints in fermentative hydrogen production. Biofuels Bioprod. Biorefining 2007, 1, 201–214. [Google Scholar] [CrossRef]
  28. Nielsen, A.T.; Amandusson, H.; Bjorklund, R.; Dannetun, H.; Ejlertsson, J.; Ekedahl, L.G.; Lundström, I.; Svensson, B.H. Hydrogen production from organic waste. Int. J. Hydrog. Energy 2001, 26, 547–550. [Google Scholar] [CrossRef]
  29. Kim, S.H.; Shin, H.S. Effects of base-pretreatment on continuous enriched culture for hydrogen production from food waste. Int. J. Hydrog. Energy 2008, 33, 5266–5274. [Google Scholar] [CrossRef]
  30. Kim, D.H.; Kim, S.H.; Shin, H.S. Hydrogen fermentation of food waste without inoculum addition. Enzyme Microb. Technol. 2009, 45, 181–187. [Google Scholar] [CrossRef]
  31. Ren, N.; Li, J.; Li, B.; Wang, Y.; Liu, S. Biohydrogen production from molasses by anaerobic fermentation with a pilot-scale bioreactor system. Int. J. Hydrog. Energy 2006, 31, 2147–2157. [Google Scholar] [CrossRef]
  32. Li, J.; Li, B.; Zhu, G.; Ren, N.; Bo, L.; He, J. Hydrogen production from diluted molasses by anaerobic hydrogen producing bacteria in an anaerobic baffled reactor (ABR). Int. J. Hydrog. Energy 2007, 32, 3274–3283. [Google Scholar] [CrossRef]
  33. Yang, P.; Zhang, R.; Mcgarvey, J.; Benemann, J. Biohydrogen production from cheese processing wastewater by anaerobic fermentation using mixed microbial communities. Int. J. Hydrog. Energy 2007, 32, 4761–4771. [Google Scholar] [CrossRef]
  34. Azbar, N.; Çetinkaya Dokgöz, F.T.; Keskin, T.; Korkmaz, K.S.; Syed, H.M. Continuous fermentative hydrogen production from cheese whey wastewater under thermophilic anaerobic conditions. Int. J. Hydrog. Energy 2009, 34, 7441–7447. [Google Scholar] [CrossRef]
  35. Lay, J.J. Modeling and optimization of anaerobic digested sludge converting starch to hydrogen. Biotechnol. Bioeng. 2000, 63, 269–278. [Google Scholar] [CrossRef]
  36. Ueno, Y.; Haruta, S.; Ishii, M.; Igarashi, Y. Microbial community in anaerobic hydrogen-producing microflora enriched from sludge compost. Appl. Microbiol. Biotechnol. 2001, 57, 555–562. [Google Scholar] [PubMed]
  37. Lin, C.Y.; Cheng, C.H. Fermentative hydrogen production from xylose using anaerobic mixed microflora. Int. J. Hydrog. Energy 2006, 31, 832–840. [Google Scholar] [CrossRef]
  38. Hussy, I.; Hawkes, F.R.; Dinsdale, R.; Hawkes, D.L. Continuous fermentative hydrogen production from sucrose and sugarbeet. Int. J. Hydrog. Energy 2005, 30, 471–483. [Google Scholar] [CrossRef]
  39. Wang, C.H.; Lin, P.J.; Chang, J.S. Fermentative conversion of sucrose and pineapple waste into hydrogen gas in phosphate-buffered culture seeded with municipal sewage sludge. Process Biochem. 2006, 41, 1353–1358. [Google Scholar] [CrossRef]
  40. Fan, K.S.; Kan, N.R.; Lay, J.J. Effect of hydraulic retention time on anaerobic hydrogenesis in CSTR. Bioresour. Technol. 2006, 97, 84–89. [Google Scholar] [CrossRef] [PubMed]
  41. Hawkes, F.R.; Dinsdale, R.; Hawkes, D.L.; Hussy, I. Sustainable fermentative hydrogen production: Challenges for process optimisation. Int. J. Hydrog. Energy 2002, 27, 1339–1347. [Google Scholar] [CrossRef]
  42. Hawkes, F.R.; Hussy, I.; Kyazze, G.; Dinsdale, R.; Hawkes, D.L. Continuous dark fermentative hydrogen production by mesophilic microflora: Principles and progress. Int. J. Hydrog. Energy 2007, 32, 172–184. [Google Scholar] [CrossRef]
  43. Varanasi, J.L.; Roy, S.; Pandit, S.; Das, D. Improvement of energy recovery from cellobiose by thermophillic dark fermentative hydrogen production followed by microbial fuel cell. Int. J. Hydrog. Energy 2015, 40, 8311–8321. [Google Scholar] [CrossRef]
  44. Angenent, L.T.; Karim, K.; Al-Dahhan, M.H.; Wrenn, B.A.; Domíguez-Espinosa, R. Production of bioenergy and biochemicals from industrial and agricultural wastewater. Trends Biotechnol. 2004, 22, 477–485. [Google Scholar] [CrossRef] [PubMed]
  45. Levin, D. Biohydrogen production: Prospects and limitations to practical application. Int. J. Hydrog. Energy 2004, 29, 173–185. [Google Scholar] [CrossRef]
  46. Dipasquale, L.; D’Ippolito, G.; Fontana, A. Capnophilic lactic fermentation and hydrogen synthesis by Thermotoga neapolitana: An unexpected deviation from the dark fermentation model. Int. J. Hydrog. Energy 2014, 39, 4857–4862. [Google Scholar] [CrossRef]
  47. Pradhan, N.; Dipasquale, L.; D’Ippolito, G.; Panico, A.; Lens, P.N.L.; Esposito, G.; Fontana, A. Hydrogen production by the thermophilic bacterium thermotoga neapolitana. Int. J. Mol. Sci. 2015, 16, 12578–12600. [Google Scholar] [CrossRef] [PubMed]
  48. Basak, N.; Das, D. The prospect of purple non-sulfur (PNS) photosynthetic bacteria for hydrogen production: The present state of the art. World J. Microbiol. Biotechnol. 2007, 23, 31–42. [Google Scholar] [CrossRef]
  49. Redwood, M.D.; Paterson-Beedle, M.; MacAskie, L.E. Integrating dark and light bio-hydrogen production strategies: Towards the hydrogen economy. Rev. Environ. Sci. Biotechnol. 2009, 8, 149–185. [Google Scholar] [CrossRef]
  50. Holladay, J.D.; Hu, J.; King, D.L.; Wang, Y. An overview of hydrogen production technologies. Catal. Today 2009, 139, 244–260. [Google Scholar] [CrossRef]
  51. Kapdan, I.K.; Kargi, F.; Oztekin, R.; Argun, H. Bio-hydrogen production from acid hydrolyzed wheat starch by photo-fermentation using different Rhodobacter sp. Int. J. Hydrog. Energy 2009, 34, 2201–2207. [Google Scholar] [CrossRef]
  52. Keskin, T.; Hallenbeck, P.C. Hydrogen production from sugar industry wastes using single-stage photofermentation. Bioresour. Technol. 2012, 112, 131–136. [Google Scholar] [CrossRef] [PubMed]
  53. Nath, K.; Das, D. Improvement of fermentative hydrogen production: Various approaches. Appl. Microbiol. Biotechnol. 2004, 65, 520–529. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, B.F.; Ren, N.Q.; Xie, G.J.; Ding, J.; Guo, W.Q.; Xing, D.F. Enhanced bio-hydrogen production by the combination of dark- and photo-fermentation in batch culture. Bioresour. Technol. 2010, 101, 5325–5329. [Google Scholar] [CrossRef] [PubMed]
  55. Logan, B.E. Microbial Fuel Cell; Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008. [Google Scholar]
  56. Rozendal, R.; Hamelers, H.; Euverink, G.; Metz, S.; Buisman, C. Principle and perspectives of hydrogen production through biocatalyzed electrolysis. Int. J. Hydrog. Energy 2006, 31, 1632–1640. [Google Scholar] [CrossRef]
  57. Manish, S.; Banerjee, R. Comparison of biohydrogen production processes. Int. J. Hydrog. Energy 2008, 33, 279–286. [Google Scholar] [CrossRef]
  58. Cusick, R.D.; Bryan, B.; Parker, D.S.; Merrill, M.D.; Mehanna, M.; Kiely, P.D.; Liu, G.; Logan, B.E. Performance of a pilot-scale continuous flow microbial electrolysis cell fed winery wastewater. Appl. Microbiol. Biotechnol. 2011, 89, 2053–2063. [Google Scholar] [CrossRef] [PubMed]
  59. Kadier, A.; Simayi, Y.; Kalil, M.S.; Abdeshahian, P.; Hamid, A.A. A review of the substrates used in microbial electrolysis cells (MECs) for producing sustainable and clean hydrogen gas. Renew. Energy 2014, 71, 466–472. [Google Scholar] [CrossRef]
  60. Ren, L.; Siegert, M.; Ivanov, I.; Pisciotta, J.M.; Logan, B.E. Treatability studies on different refinery wastewater samples using high-throughput microbial electrolysis cells (MECs). Bioresour. Technol. 2013, 136, 322–328. [Google Scholar] [CrossRef] [PubMed]
  61. Hu, H.; Fan, Y.; Liu, H. Hydrogen production in single-chamber tubular microbial electrolysis cells using non-precious-metal catalysts. Int. J. Hydrog. Energy 2009, 34, 8535–8542. [Google Scholar] [CrossRef]
  62. Lu, L.; Xing, D.; Ren, N.; Logan, B.E. Syntrophic interactions drive the hydrogen production from glucose at low temperature in microbial electrolysis cells. Bioresour. Technol. 2012, 124, 68–76. [Google Scholar] [CrossRef] [PubMed]
  63. Lu, L.; Ren, N.; Xing, D.; Logan, B.E. Hydrogen production with effluent from an ethanol-H2-coproducing fermentation reactor using a single-chamber microbial electrolysis cell. Biosens. Bioelectron. 2009, 24, 3055–3060. [Google Scholar] [CrossRef] [PubMed]
  64. Call, D.; Logan, B.E. Hydrogen production in a single chamber microbial electrolysis cell lacking a membrane. Environ. Sci. Technol. 2008, 42, 3401–3406. [Google Scholar] [CrossRef] [PubMed]
  65. Call, D.F.; Merrill, M.D.; Logan, B.E. High surface area stainless steel brushes as cathodes in microbial electrolysis cells. Environ. Sci. Technol. 2009, 43, 2179–2183. [Google Scholar] [CrossRef] [PubMed]
  66. Selembo, P.A.; Perez, J.M.; Lloyd, W.A.; Logan, B.E. High hydrogen production from glycerol or glucose by electrohydrogenesis using microbial electrolysis cells. Int. J. Hydrog. Energy 2009, 34, 5373–5381. [Google Scholar] [CrossRef]
  67. Wang, X.; Cheng, S.; Feng, Y.; Merrill, M.D.; Saito, T.; Logan, B.E. Use of carbon mesh anodes and the effect of different pretreatment methods on power production in microbial fuel cells. Environ. Sci. Technol. 2009, 43, 6870–6874. [Google Scholar] [CrossRef] [PubMed]
  68. Wagner, R.C.; Regan, J.M.; Oh, S.-E.; Zuo, Y.; Logan, B.E. Hydrogen and methane production from swine wastewater using microbial electrolysis cells. Water Res. 2009, 43, 1480–1488. [Google Scholar] [CrossRef] [PubMed]
  69. Pisciotta, J.M.; Zaybak, Z.; Call, D.F.; Nam, J.Y.; Logan, B.E. Enrichment of microbial electrolysis cell biocathodes from sediment microbial fuel cell bioanodes. Appl. Environ. Microbiol. 2012, 78, 5212–5219. [Google Scholar] [CrossRef] [PubMed]
  70. Xiao, L.; Wen, Z.; Ci, S.; Chen, J.; He, Z. Carbon/iron-based nanorod catalysts for hydrogen production in microbial electrolysis cells. Nano Energy 2012, 1, 751–756. [Google Scholar] [CrossRef]
  71. Guo, K.; Tang, X.; Du, Z.; Li, H. Hydrogen production from acetate in a cathode-on-top single-chamber microbial electrolysis cell with a mipor cathode. Biochem. Eng. J. 2010, 51, 48–52. [Google Scholar] [CrossRef]
  72. Hu, H.; Fan, Y.; Liu, H. Hydrogen production using single-chamber membrane-free microbial electrolysis cells. Water Res. 2008, 42, 4172–4178. [Google Scholar] [CrossRef] [PubMed]
  73. Van Groenestijn, J.W.; Geelhoed, J.S.; Goorissen, H.P.; Meesters, K.P.; Stams, A.J.; Claassen, P.A. Performance and population analysis of a non-sterile trickle bed reactor inoculated with caldicellulosiruptor saccharolyticus, a thermophilic hydrogen producer. Biotechnol. Bioeng. 2009, 102, 1361–1367. [Google Scholar] [CrossRef] [PubMed]
  74. Kleerebezem, R.; van Loosdrecht, M.C. Mixed culture biotechnology for bioenergy production. Curr. Opin. Biotechnol. 2007, 18, 207–212. [Google Scholar] [CrossRef] [PubMed]
  75. Kengen, S.W.M.; Goorissen, H.P.; Verhaart, M.; Stams, A.J.M.; van Niel, E.W.J.; Claassen, P.A.M. Biological Hydrogen Production by Anaerobic Microorganisms. In Biofuels; John Wiley & Sons, Ltd.: Chichester, UK, 2009; pp. 197–221. [Google Scholar]
  76. Kotsopoulos, T.A.; Zeng, R.J.; Angelidaki, I. Biohydrogen production in granular up-flow anaerobic sludge blanket (UASB) reactors with mixed cultures under hyper-thermophilic temperature (70 C). Biotechnol. Bioeng. 2006, 94, 296–302. [Google Scholar] [CrossRef] [PubMed]
  77. De Vrije, T.; Bakker, R.R.; Budde, M.A.; Lai, M.H.; Mars, A.E.; Claassen, P.A. Efficient hydrogen production from the lignocellulosic energy crop Miscanthus by the extreme thermophilic bacteria Caldicellulosiruptor saccharolyticus and Thermotoga neapolitana. Biotechnol. Biofuels 2009, 2, 12. [Google Scholar] [CrossRef] [PubMed]
  78. Van Niel, E.W.; Budde, M.A.; De Haas, G.; Van der Wal, F.J.; Claassen, P.A.; Stams, A.J. Distinctive properties of high hydrogen producing extreme thermophiles, Caldicellulosiruptor saccharolyticus and Thermotoga elfii. Int. J. Hydrog. Energy 2002, 27, 1391–1398. [Google Scholar] [CrossRef]
  79. Thauer, R.K.; Jungermann, K.; Decker, K. Energy conservation in chemotrophic anaerobic bacteria. Bacteriol. Rev. 1977, 41, 100–180. [Google Scholar] [PubMed]
  80. Vanginkel, S.; Oh, S.; Logan, B. Biohydrogen gas production from food processing and domestic wastewaters. Int. J. Hydrog. Energy 2005, 30, 1535–1542. [Google Scholar] [CrossRef]
  81. Liu, D.; Liu, D.; Zeng, R.J.; Angelidaki, I. Hydrogen and methane production from household solid waste in the two-stage fermentation process. Water Res. 2006, 40, 2230–2236. [Google Scholar] [CrossRef] [PubMed]
  82. Oh, S.-E.; van Ginkel, S.; Logan, B.E. The relative effectiveness of pH control and heat treatment for enhancing biohydrogen gas production. Environ. Sci. Technol. 2003, 37, 5186–5190. [Google Scholar] [CrossRef] [PubMed]
  83. Liang, T.M.; Cheng, S.S.; Wu, K.L. Behavioral study on hydrogen fermentation reactor installed with silicone rubber membrane. Int. J. Hydrog. Energy 2002, 27, 1157–1165. [Google Scholar] [CrossRef]
  84. Van Groenestijn, J.W.; Hazewinkel, J.H.; Nienoord, M.; Bussmann, P.J. Energy aspects of biological hydrogen production in high rate bioreactors operated in the thermophilic temperature range. Int. J. Hydrog. Energy 2002, 27, 1141–1147. [Google Scholar] [CrossRef]
  85. Kim, D.H.; Han, S.K.; Kim, S.H.; Shin, H.S. Effect of gas sparging on continuous fermentative hydrogen production. Int. J. Hydrog. Energy 2006, 31, 2158–2169. [Google Scholar] [CrossRef]
  86. Koku, H. Aspects of the metabolism of hydrogen production by Rhodobacter sphaeroides. Int. J. Hydrog. Energy 2002, 27, 1315–1329. [Google Scholar] [CrossRef]
  87. Beckers, L.; Masset, J.; Hamilton, C.; Delvigne, F.; Toye, D.; Crine, M.; Thonart, P.; Hiligsmann, S. Investigation of the links between mass transfer conditions, dissolved hydrogen concentration and biohydrogen production by the pure strain Clostridium butyricum CWBI1009. Biochem. Eng. J. 2015, 98, 18–28. [Google Scholar] [CrossRef]
  88. Mizuno, O.; Dinsdale, R.; Hawkes, F.R.; Hawkes, D.L.; Noike, T. Enhancement of hydrogen production from glucose by nitrogen gas sparging. Bioresour. Technol. 2000, 73, 59–65. [Google Scholar] [CrossRef]
  89. Chou, C.; Wang, C.; Huang, C.; Lay, J. Pilot study of the influence of stirring and pH on anaerobes converting high-solid organic wastes to hydrogen. Int. J. Hydrog. Energy 2008, 33, 1550–1558. [Google Scholar] [CrossRef]
  90. Fontes Lima, D.M.; Zaiat, M. The influence of the degree of back-mixing on hydrogen production in an anaerobic fixed-bed reactor. Int. J. Hydrog. Energy 2012, 37, 9630–9635. [Google Scholar] [CrossRef]
  91. Montiel-Corona, V.; Revah, S.; Morales, M. Hydrogen production by an enriched photoheterotrophic culture using dark fermentation effluent as substrate: Effect of flushing method, bicarbonate addition, and outdoor–indoor conditions. Int. J. Hydrog. Energy 2015, 40, 9096–9105. [Google Scholar] [CrossRef]
  92. Clark, I.C.; Zhang, R.H.; Upadhyaya, S.K. The effect of low pressure and mixing on biological hydrogen production via anaerobic fermentation. Int. J. Hydrog. Energy 2012, 37, 11504–11513. [Google Scholar] [CrossRef]
  93. Kim, D.-H.; Shin, H.-S.; Kim, S.-H. Enhanced H2 fermentation of organic waste by CO2 sparging. Int. J. Hydrog. Energy 2012, 37, 15563–15568. [Google Scholar] [CrossRef]
  94. Gómez, X.; Morán, A.; Cuetos, M.J.; Sánchez, M.E. The production of hydrogen by dark fermentation of municipal solid wastes and slaughterhouse waste: A two-phase process. J. Power Sources 2006, 157, 727–732. [Google Scholar] [CrossRef]
  95. Clauwaert, P.; Verstraete, W. Methanogenesis in membraneless microbial electrolysis cells. Appl. Microbiol. Biotechnol. 2009, 82, 829–836. [Google Scholar] [CrossRef] [PubMed]
  96. Cheng, S.; Xing, D.; Call, D.F.; Logan, B.E. Direct biological conversion of electrical current into methane by electromethanogenesis. Environ. Sci. Technol. 2009, 43, 3953–3958. [Google Scholar] [CrossRef] [PubMed]
  97. Pant, D.; van Bogaert, G.; de Smet, M.; Diels, L.; Vanbroekhoven, K. Use of novel permeable membrane and air cathodes in acetate microbial fuel cells. Electrochimica Acta 2010, 55, 7710–7716. [Google Scholar] [CrossRef]
  98. Selembo, P.A.; Merrill, M.D.; Logan, B.E. The use of stainless steel and nickel alloys as low-cost cathodes in microbial electrolysis cells. J. Power Sour. 2009, 190, 271–278. [Google Scholar] [CrossRef]
  99. Maeda, T.; Sanchez-Torres, V.; Wood, T.K. Metabolic engineering to enhance bacterial hydrogen production. Microb. Biotechnol. 2008, 1, 30–39. [Google Scholar] [CrossRef] [PubMed]
  100. Morimoto, K.; Kimura, T.; Sakka, K.; Ohmiya, K. Overexpression of a hydrogenase gene in Clostridium paraputrificum to enhance hydrogen gas production. FEMS Microbiol. Lett. 2005, 246, 229–234. [Google Scholar] [CrossRef] [PubMed]
  101. Vardar-Schara, G.; Maeda, T.; Wood, T.K. Metabolically engineered bacteria for producing hydrogen via fermentation. Microb. Biotechnol. 2008, 1, 107–125. [Google Scholar] [CrossRef] [PubMed]
  102. Hallenbeck, P.C.; Ghosh, D. Improvements in fermentative biological hydrogen production through metabolic engineering. J. Environ. Manag. 2012, 95, S360–S364. [Google Scholar] [CrossRef] [PubMed]
  103. Ryu, M.-H.; Hull, N.C.; Gomelsky, M. Metabolic engineering of Rhodobacter sphaeroides for improved hydrogen production. Int. J. Hydrog. Energy 2014, 39, 6384–6390. [Google Scholar] [CrossRef]
  104. Wang, J.; Wan, W. Kinetic models for fermentative hydrogen production: A review. Int. J. Hydrog. Energy 2009, 34, 3313–3323. [Google Scholar] [CrossRef]
  105. Dincer, I. Hydrogen and Fuel Cell Technologies for Sustainable Future. Jordan J. Mech. Ind. Eng. 2008, 2, 1–14. [Google Scholar]
  106. Rathore, D.; Singh, A. Biohydrogen production from microalgae. In Biofuels Technologies Recent Developments; Gupta, V., Tuohy, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 317–333. [Google Scholar]
  107. Kwak, H.Y.; Lee, H.S.; Jung, J.Y.; Jeon, J.S.; Park, D.R. Exergetic and thermoeconomic analysis of a 200-kW phosphoric acid fuel cell plant. Fuel 2004, 83, 2087–2094. [Google Scholar] [CrossRef]
  108. Rubio Rodríguez, M.A.; de Ruyck, J.; Díaz, P.R.; Verma, V.K.; Bram, S. An LCA based indicator for evaluation of alternative energy routes. Appl. Energy 2011, 88, 630–635. [Google Scholar] [CrossRef]
  109. Romagnoli, F.; Blumberga, D.; Pilicka, I. Life cycle assessment of biohydrogen production in photosynthetic processes. Int. J. Hydrog. Energy 2011, 36, 7866–7871. [Google Scholar] [CrossRef]
  110. Djomo, S.N.; Humbert, S. Dagnija Blumberga Life cycle assessment of hydrogen produced from potato steam peels. Int. J. Hydrog. Energy 2008, 33, 3067–3072. [Google Scholar] [CrossRef]
  111. Djomo, S.N.; Blumberga, D. Comparative life cycle assessment of three biohydrogen pathways. Bioresour. Technol. 2011, 102, 2684–2694. [Google Scholar] [CrossRef] [PubMed]
  112. Ochs, D.; Wukovits, W.; Ahrer, W. Life cycle inventory analysis of biological hydrogen production by thermophilic and photo fermentation of potato steam peels (PSP). J. Clean. Prod. 2010, 18, S88–S94. [Google Scholar] [CrossRef]
  113. Chen, S.D.; Lo, Y.C.; Lee, K.S.; Huang, T.I.; Chang, J.S. Sequencing batch reactor enhances bacterial hydrolysis of starch promoting continuous bio-hydrogen production from starch feedstock. Int. J. Hydrog. Energy 2009, 34, 8549–8557. [Google Scholar] [CrossRef]
  114. Seifert, K.; Waligorska, M.; Wojtowski, M.; Laniecki, M. Hydrogen generation from glycerol in batch fermentation process. Int. J. Hydrog. Energy 2009, 34, 3671–3678. [Google Scholar] [CrossRef]
  115. Lin, C.N.; Wu, S.Y.; Chang, J.S.; Chang, J.S. Biohydrogen production in a three-phase fluidized bed bioreactor using sewage sludge immobilized by ethylene-vinyl acetate copolymer. Bioresour. Technol. 2009, 100, 3298–3301. [Google Scholar] [CrossRef] [PubMed]
  116. Yasin Nazlina, H.M.; Aini, R.; Ismail, F.; Zulkhairi, M.; Hassan, M.A. Effect of different temperature, initial pH and substrate composition on biohydrogen production from food waste in batch fermentation. Asian J. Biotechnol. 2009, 1, 42–50. [Google Scholar] [CrossRef]
  117. Zhu, J.; Li, Y.; Wu, X.; Miller, C.; Chen, P.; Ruan, R. Swine manure fermentation for hydrogen production. Bioresour. Technol. 2009, 100, 5472–5477. [Google Scholar] [CrossRef] [PubMed]
  118. Gadhamshetty, V.; Johnson, D.C.; Nirmalakhandan, N.; Smith, G.B.; Deng, S. Feasibility of biohydrogen production at low temperatures in unbuffered reactors. Int. J. Hydrog. Energy 2009, 34, 1233–1243. [Google Scholar] [CrossRef]
  119. Ghosh, D.; Hallenbeck, P.C. Fermentative hydrogen yields from different sugars by batch cultures of metabolically engineered Escherichia coli DJT135. Int. J. Hydrog. Energy 2009, 34, 7979–7982. [Google Scholar] [CrossRef]
  120. Kargi, F.; Pamukoglu, M.Y. Dark fermentation of ground wheat starch for bio-hydrogen production by fed-batch operation. Int. J. Hydrog. Energy 2009, 34, 2940–2946. [Google Scholar] [CrossRef]
  121. Han, W.; Ye, M.; Zhu, A.J.; Zhao, H.T.; Li, Y.F. Batch dark fermentation from enzymatic hydrolyzed food waste for hydrogen production. Bioresour. Technol. 2015, 191, 24–29. [Google Scholar] [CrossRef] [PubMed]
  122. Chookaew, T.; O-Thong, S.; Prasertsan, P. Biohydrogen production from crude glycerol by two stage of dark and photo fermentation. Int. J. Hydrog. Energy 2015, 40, 7433–7438. [Google Scholar] [CrossRef]
  123. Wicher, E.; Seifert, K.; Zagrodnik, R.; Pietrzyk, B.; Laniecki, M. Hydrogen gas production from distillery wastewater by dark fermentation. Int. J. Hydrog. Energy 2013, 38, 7767–7773. [Google Scholar] [CrossRef]
  124. Moreno, R.; Escapa, A.; Cara, J.; Carracedo, B.; Gómez, X. A two-stage process for hydrogen production from cheese whey: Integration of dark fermentation and biocatalyzed electrolysis. Int. J. Hydrog. Energy 2015, 40, 168–175. [Google Scholar] [CrossRef]
  125. Su, H.; Cheng, J.; Zhou, J.; Song, W.; Cen, K. Hydrogen production from water hyacinth through dark- and photo- fermentation. Int. J. Hydrog. Energy 2010, 35, 8929–8937. [Google Scholar] [CrossRef]
  126. Argun, H.; Kargi, F. Effects of sludge pre-treatment method on bio-hydrogen production by dark fermentation of waste ground wheat. Int. J. Hydrog. Energy 2009, 34, 8543–8548. [Google Scholar] [CrossRef]
  127. Laurinavichene, T.V.; Belokopytov, B.F.; Laurinavichius, K.S.; Tekucheva, D.N.; Seibert, M.; Tsygankov, A.A. Towards the integration of dark- and photo-fermentative waste treatment. 3. Potato as substrate for sequential dark fermentation and light-driven H2 production. Int. J. Hydrog. Energy 2010, 35, 8536–8543. [Google Scholar] [CrossRef]
  128. Ozmihci, S.; Kargi, F. Effects of starch loading rate on performance of combined fed-batch fermentation of ground wheat for bio-hydrogen production. Int. J. Hydrog. Energy 2010, 35, 1106–1111. [Google Scholar] [CrossRef]
  129. Avcioglu, S.G.; Ozgur, E.; Eroglu, I.; Yucel, M.; Gunduz, U. Biohydrogen production in an outdoor panel photobioreactor on dark fermentation effluent of molasses. Int. J. Hydrog. Energy 2011, 36, 11360–11368. [Google Scholar] [CrossRef]
  130. Zhu, Z.; Shi, J.; Zhou, Z.; Hu, F.; Bao, J. Photo-fermentation of Rhodobacter sphaeroides for hydrogen production using lignocellulose-derived organic acids. Process Biochem. 2010, 45, 1894–1898. [Google Scholar] [CrossRef]
  131. Francou, N.; Vignais, P.M. Hydrogen production by Rhodopseudomonas capsulata cells entrapped in carrageenan beads. Biotechnol. Lett. 1984, 6, 639–644. [Google Scholar] [CrossRef]
  132. Taguchi, F.; Mizukami, N.; Hasegawa, K.; Saito-Taki, T. Microbial conversion of arabinose and xylose to hydrogen by a newly isolated clostridium sp. No. 2. Can. J. Microbiol. 1994, 40, 228–233. [Google Scholar] [CrossRef]
  133. Susanti, R.F.; Dianningrum, L.W.; Yum, T.; Kim, Y.; Lee, B.G.; Kim, J. High-yield hydrogen production from glucose by supercritical water gasification without added catalyst. Int. J. Hydrog. Energy 2012, 37, 11677–11690. [Google Scholar] [CrossRef]
  134. Qinming, Z.; Shuzhong, W.; Liang, W.; Donghai, X. Catalytic Hydrogen Production from Municipal Sludge in Supercritical Water with Partial Oxidation. Chall. Power Eng. Environ. 2007, 1, 1252–1255. [Google Scholar]
  135. Frusteri, F.; Freni, S.; Chiodo, V.; Spadaro, L.; Bonura, G.; Cavallaro, S. Potassium improved stability of Ni/MgO in the steam reforming of ethanol for the production of hydrogen for MCFC. J. Power Sour. 2004, 132, 139–144. [Google Scholar] [CrossRef]
  136. Aiello, R.; Fiscus, J.E.; Loye, H.; Amiridis, M.D. Hydrogen production via the direct cracking of methane over Ni/SiO2: Catalyst deactivation and regeneration. Appl. Catal. A Gen. 2000, 192, 227–234. [Google Scholar] [CrossRef]
  137. Deng, W.; Jiang, H.; Wu, Y.; Fan, H.; Ji, J. Hydrogen production from biomass pyrolysis in molten alkali. AASRI Procedia 2012, 3, 217–223. [Google Scholar] [CrossRef]
  138. Sivasubramanian, P.; Ramasamy, R.P.; Freire, F.J.; Holland, C.E.; Weidner, J.W. Electrochemical hydrogen production from thermochemical cycles using a proton exchange membrane electrolyzer. Int. J. Hydrog. Energy 2007, 32, 463–468. [Google Scholar] [CrossRef]
  139. Kim, J.; Kwon, D.; Kim, K.; Hoffmann, M.R. Electrochemical Production of Hydrogen Coupled with the Oxidation of Arsenite. Environ. Sci. Technol. 2014, 48, 2059–2066. [Google Scholar] [CrossRef] [PubMed]
  140. Karn, R.K.; Srivastava, O.N. On the synthesis and photochemical studies of nanostructured TiO 1 and TiO 1 admixed VO 1 photoelectrodes in regard to hydrogen production through photoelectrolysis. Int. J. Hydrog. Energy 1999, 24, 965–971. [Google Scholar] [CrossRef]
  141. Perlack, R.D.; Wright, L.L.; Turhollow, A.F.; Graham, R.L.; Stokes, B.J.; Erbach, D.C. Biomass as Feedstock For a Bioenergy and Bioproducts Industry: The Technical Feasibility of a Billion-Ton Annual Supply; U.S. Department of Energy: Oak Ridge, TN, USA, 2005. [Google Scholar]

Share and Cite

MDPI and ACS Style

Singh, A.; Sevda, S.; Abu Reesh, I.M.; Vanbroekhoven, K.; Rathore, D.; Pant, D. Biohydrogen Production from Lignocellulosic Biomass: Technology and Sustainability. Energies 2015, 8, 13062-13080. https://doi.org/10.3390/en81112357

AMA Style

Singh A, Sevda S, Abu Reesh IM, Vanbroekhoven K, Rathore D, Pant D. Biohydrogen Production from Lignocellulosic Biomass: Technology and Sustainability. Energies. 2015; 8(11):13062-13080. https://doi.org/10.3390/en81112357

Chicago/Turabian Style

Singh, Anoop, Surajbhan Sevda, Ibrahim M. Abu Reesh, Karolien Vanbroekhoven, Dheeraj Rathore, and Deepak Pant. 2015. "Biohydrogen Production from Lignocellulosic Biomass: Technology and Sustainability" Energies 8, no. 11: 13062-13080. https://doi.org/10.3390/en81112357

Article Metrics

Back to TopTop