Fact-checked by Grok 2 weeks ago

Biohydrogen


Biohydrogen is hydrogen gas (H₂) produced through biological processes mediated by microorganisms, including and , which convert organic substrates, , or gases into H₂ via pathways such as dark fermentation, photo-fermentation, and biophotolysis. These methods leverage microbial enzymes like hydrogenases and nitrogenases to generate H₂ as a byproduct of metabolic activities, often utilizing renewable feedstocks like , , or sunlight-driven .
Biohydrogen production stands out for its potential to yield a clean, from abundant , mitigating compared to steam methane reforming, the dominant chemical method that relies on fossil fuels. Key advantages include compatibility with , operation at ambient temperatures and pressures, and scalability through of microbes to enhance yields, as demonstrated in advances with like Chlorella and fermentative . Despite these benefits, significant hurdles persist, including thermodynamically limited H₂ yields (typically 1-4 mol H₂ per mol glucose in dark ), sensitivity of hydrogen-producing enzymes to oxygen, and high operational costs that render it non-competitive with gray at current scales. Research has progressed toward hybrid systems combining biological and electrocatalytic steps to boost efficiency, with pilot demonstrations achieving integrated bio-photoelectrochemical production, though commercial viability demands breakthroughs in reactor design and inhibitor mitigation. Defining characteristics include its role in circular economies by treating organic waste while generating energy, positioning biohydrogen as a promising yet underdeveloped vector for sustainable energy transitions amid global decarbonization efforts.

Fundamentals

Definition and Biochemical Mechanisms

Biohydrogen is gas (H₂) generated through biological processes mediated by microorganisms, such as , , and , distinguishing it from chemical or electrolytic methods by relying on enzymatic driven by or . These processes exploit natural metabolic pathways to produce H₂, often as a byproduct or engineered output, with potential yields influenced by factors like availability and environmental conditions. Unlike fuel-derived , biohydrogen pathways aim to utilize renewable inputs like or , though practical efficiencies remain constrained by biological limitations. At the core of biohydrogen production are metalloenzymes—primarily hydrogenases and nitrogenases—that catalyze the reversible reduction of protons to H₂ via the reaction 2H⁺ + 2e⁻ ⇌ H₂. Hydrogenases, found in diverse anaerobes and phototrophs, facilitate electron transfer from metabolic reductants (e.g., ferredoxin or NADH) to protons, enabling H₂ evolution under anaerobic conditions to dispose of excess reducing equivalents. Nitrogenases, typically involved in nitrogen fixation, generate H₂ stoichiometrically during the reduction of N₂ (or in its absence), consuming significant ATP: N₂ + 8H⁺ + 8e⁻ + 16ATP → 2NH₃ + H₂ + 16ADP + 16Pᵢ. These enzymes operate at ambient temperatures and pressures, offering kinetic advantages over synthetic catalysts, but are often inhibited by oxygen or require anoxic environments for sustained activity. Hydrogenases are classified into three main types based on metal content: [NiFe]-hydrogenases, prevalent in for H₂ uptake and sensing; [FeFe]-hydrogenases, dominant in and certain for high-rate H₂ production due to their low and turnover numbers exceeding 10,000 s⁻¹; and [Fe]-hydrogenases, specialized for methanogenic pathways. The [FeFe]-hydrogenase active site features a unique H-cluster (2Fe subcluster linked to [4Fe-4S]), enabling efficient via mechanisms like formation and intermediates. In biohydrogen contexts, [FeFe]-hydrogenases in organisms like couple to photosynthetic electron transport, while [NiFe]-variants in fermentative handle bidirectional . Engineering efforts focus on enhancing oxygen tolerance and eliminating uptake hydrogenases to boost net yields, as native systems often recycle produced H₂. Nitrogenases, with their MoFe or VFe/FeFe cofactors, exhibit lower specificity for H₂ production but contribute in diazotrophic microbes under nitrogen-limited conditions.

Thermodynamic and Efficiency Constraints

The production of biohydrogen is fundamentally limited by the thermodynamics of hydrogen evolution, particularly the endergonic water-splitting reaction (2H₂O → 2H₂ + O₂), which requires a standard input of +237 kJ/mol H₂ (or +474 kJ for two moles) at 25°C and , corresponding to a minimum cell potential of 1.23 V. In biological systems, this energy barrier is overcome either through light-driven or of organic substrates, but inherent losses in , overpotentials at active sites, and unfavorable reaction equilibria impose strict yield limits. Biological [FeFe]-hydrogenases exhibit low overpotentials (near the reversible H⁺/H₂ potential of -0.414 V at 7), enabling theoretically efficient , yet thermodynamic favorability decreases with rising H₂ , inhibiting further production above ~10⁻⁴ . In light-dependent pathways like biophotolysis and photofermentation, solar-to-hydrogen (STH) efficiency is constrained by the photosynthetic apparatus: require at least 8-10 quanta per H₂ molecule (4 for water oxidation via PSII and 4 for H₂ evolution via or ), with pigment absorption limited to ~45% of the solar spectrum (400-700 PAR) and thermodynamic losses in charge separation (~0.2-0.4 eV per step). The theoretical maximum STH efficiency for photobiological systems is 10-13%, factoring in limits, ceilings, and entropy penalties, but practical values rarely exceed 1-2% due to antenna shading, cyclic electron flow dissipation, and O₂-mediated inactivation necessitating spatial or temporal of O₂ and H₂ evolution. U.S. Department of Energy targets aim for 5.5% STH from organic feedstocks by 2030, reflecting compounded biological inefficiencies beyond pure photochemical conversion. Dark fermentation faces substrate-specific thermodynamic ceilings, where (e.g., glucose → 4H₂ + 2CO₂ + 2) is exergonic (ΔG° ≈ -184 kJ/mol glucose for the acetate pathway) but limited to ~33% of higher heating value (HHV) as H₂ , as more reduced end-products like butyrate or (yielding 2-3 H₂/mol ) are favored under high H₂ pressures to minimize . Yields seldom surpass 2 mol H₂/mol in practice (versus a stoichiometric maximum of 4), constrained by NADH/NAD⁺ imbalances and solventogenesis shifting pathways away from H₂-producing routes. efficiency hovers at 20-40%, further eroded by heat losses and incomplete COD removal (~50-70%). Two-stage hybrids (dark + photofermentation) theoretically access up to 12 mol H₂/mol by photo-oxidizing dark fermentation effluents, boosting combined toward 7-10% light energy conversion, yet inter-stage losses and microbial consortia imbalances cap real-world performance below 5% overall STH. These constraints underscore that while offers ambient-condition operation, thermodynamic and kinetic bottlenecks necessitate integrated to approach viability against electrolytic benchmarks (>70% ).

Biological Production Pathways

Light-Dependent Processes

Light-dependent processes for biohydrogen production utilize photosynthetic microorganisms to convert into stored as gas, drawing from water or organic substrates as electron donors. These pathways encompass biophotolysis in oxygenic phototrophs such as and , and photofermentation in anoxygenic phototrophs like purple non-sulfur bacteria. Unlike dark fermentation, these methods require illumination to drive through or bacteriochlorophyll-based reaction centers, achieving theoretical solar-to- efficiencies of 10-12% under optimal conditions, though practical yields remain below 5% due to kinetic limitations and sensitivities. In biophotolysis, light absorption by generates electrons from water oxidation, producing oxygen as a byproduct, with subsequent transfer to enzymes catalyzing H2 evolution. Direct biophotolysis couples this process simultaneously in the same compartment, but oxygen irreversibly inhibits the typically O2-sensitive [FeFe]-hydrogenases, restricting sustained production to or sulfur-deprived conditions, as demonstrated in where sulfur deprivation induces activity and yields up to 1-2% of solar energy conversion. Indirect biophotolysis separates carbohydrate accumulation via from subsequent hydrogen production, mitigating oxygen inhibition and enabling higher theoretical efficiencies, with achieving light conversion rates up to 16.3% in optimized strains. Photofermentation by purple non-sulfur bacteria, such as Rhodopseudomonas species, employs enzymes under limitation to reduce protons to using electrons from oxidized acids like , with light enhancing cyclic electron flow to regenerate ATP. This process tolerates broader substrates from dark fermentation effluents, yielding at rates of 1-7 mmol per liter per hour in lab-scale bioreactors, and integrates well in two-stage systems for complete waste valorization, though 's high ATP demand (16-24 ATP per ) imposes efficiency constraints around 1-3%.

Direct and Indirect Biophotolysis in Algae

Direct biophotolysis in algae involves the light-driven splitting of water molecules by photosystem II (PSII) in green microalgae, where absorbed photons generate electrons that are transferred through photosystem I (PSI) and ferredoxin to [Fe-Fe]-hydrogenase enzymes, catalyzing the reduction of protons to hydrogen gas (2H⁺ + 2e⁻ → H₂). This process theoretically achieves over 80% sunlight conversion efficiency but practically yields low rates due to oxygen inhibition of the oxygen-sensitive hydrogenase. Species such as Chlamydomonas reinhardtii are commonly studied, with hydrogen production induced under anaerobic conditions or sulfur deprivation to temporarily repress PSII activity and minimize O₂ evolution. Challenges in direct biophotolysis include the rapid inactivation of [Fe-Fe]-hydrogenase by O₂, a byproduct of water oxidation, limiting sustained yields to less than 2% solar-to-hydrogen efficiency in most setups. Reported production rates reach approximately 20 kg H₂ per 1,000 m² per day under optimized lab conditions, though scalability remains hindered by high light requirements and enzyme instability. Advances, such as genetic modifications introducing PSI-hydrogenase chimeras or flavodiiron protein knockouts, have enhanced tolerance to O₂ in C. reinhardtii, boosting yields in experimental strains. Indirect biophotolysis separates hydrogen production into two stages: first, fixes CO₂ into carbohydrates during an aerobic phase, followed by of these storage compounds (e.g., or ) to H₂ via under O₂-free conditions. This temporal or spatial decoupling mitigates O₂ inhibition, allowing higher sustainability than direct methods, with examples in species yielding up to 11.65 mL H₂ per liter. like and Scenedesmus obliquus demonstrate efficacy, often enhanced by nutrient deprivation strategies similar to direct processes. Indirect approaches benefit from continuous accumulation but face discontinuities between and dark phases, potentially reducing overall . Hybrid systems integrating with or techniques, such as alginate encapsulation of Tetraspora sp., have reported up to 10-fold yield improvements by facilitating microenvironments. Economic analyses estimate production costs at around $1.42 per kg H₂, lower than direct methods due to better O₂ management, though large-scale bioreactors are needed for viability.

Biophotolysis in Cyanobacteria

Biophotolysis in utilizes photosynthetic light energy to split water molecules, generating electrons that reduce protons to molecular hydrogen (H₂) via enzymes such as or nitrogenases.00105-6) This process occurs through direct or indirect pathways, with distinguished by their prokaryotic nature and ability to perform , unlike eukaryotic . Direct biophotolysis involves the immediate transfer of electrons from (PSII) and (PSI) to a enzyme, potentially achieving theoretical solar-to-hydrogen efficiencies up to 10%, though practical yields remain below 2% due to oxygen inhibition of the oxygen-sensitive [NiFe]-hydrogenase. In indirect biophotolysis, first store photosynthetic products like or carbohydrates in specialized cells, then ferment these reserves to produce H₂, mitigating oxygen interference. Filamentous species such as Anabaena variabilis employ spatial separation: vegetative cells perform oxygenic , while heterocysts— cells comprising 5-10% of the filament—host , which evolves H₂ as a byproduct of , with rates up to 10-20 μmol H₂ mg⁻¹ h⁻¹ under nitrogen-limited conditions. Unicellular like Cyanothece sp. ATCC 51142 achieve temporal separation, accumulating carbohydrates during the day and producing H₂ at night via , yielding up to 400 μmol H₂ mg⁻¹ protein h⁻¹ in optimized shake-flask cultures. Efficiency constraints in both pathways stem from competition for electrons by CO₂ fixation and the bidirectional nature of hydrogenases, which also consume H₂, limiting net production to 1-5% of theoretical maxima. Genetic engineering strategies, including maturation enhancements and PSII downregulation, have improved yields in strains like Synechocystis sp. PCC 6803, but scalability remains challenged by light saturation and nutrient demands. Ongoing research emphasizes integrating indirect biophotolysis with waste CO₂ feeds to enhance sustainability.

Photofermentation by Purple Non-Sulfur Bacteria

Photofermentation by purple non-sulfur bacteria (PNSB) utilizes light energy to convert organic substrates into molecular under conditions, without . These anoxygenic phototrophs, including species such as , Rhodobacter sphaeroides, , and Rhodobacter capsulatus, absorb light via bacteriochlorophylls and to generate ATP through cyclic . This ATP drives the enzyme complex, which reduces protons to H₂ using electrons from the catabolism of organic compounds like , , or glucose, yielding CO₂ as a byproduct. The process is catalyzed primarily by , with uptake potentially reconsuming H₂ unless inhibited by conditions such as limitation or specific inhibitors. Substrates are typically volatile fatty acids from dark fermentation effluents or wastes like and agricultural residues, enabling two-stage systems that achieve near-complete substrate oxidation. Hydrogen yields in photofermentation exceed those of dark fermentation alone, with theoretical maxima of up to 12 H₂ per glucose due to the avoidance of oxygen inhibition and efficient electron diversion to . Practical rates vary by strain and conditions; for instance, R. sphaeroides achieved 8.7 mmol H₂/L/h from 40 mM , while R. capsulatus produced 2.6 mmol H₂/L/h from 35 mM glucose. In R. rubrum, dynamic CO feeding in fed-batch mode enhanced rates to 27.2 mmol H₂/L/h at 30°C and pH 7.0–7.2, compared to 11 mmol H₂/L/h in batch with . Scale-up studies with Rhodopseudomonas sp. S16-VOGS3 in 4 L photobioreactors yielded 1642 mL H₂ total, with productivities of 0.717 mL/L/h and light conversion efficiencies around 0.72% using optimized spiral mixing. This pathway offers advantages including high H₂ purity (>95%), operation at ambient temperatures and pressures, and compatibility with waste streams for , but faces challenges like light penetration limits in dense cultures, low volumetric rates (typically <10 mmol/L/h), and sensitivity to substrate inhibition or ammonium excess. Ongoing optimizations, such as co-cultures (e.g., R. rubrum with R. capsulatus) and genetic engineering to disrupt hydrogenase, aim to boost net yields toward thermodynamic limits.

Light-Independent Processes

Light-independent processes for biohydrogen production encompass anaerobic microbial pathways that generate hydrogen without requiring photosynthetic light energy, enabling potential continuous operation but constrained by lower thermodynamic yields compared to light-dependent methods. These primarily include dark fermentation and microbial electrolysis cells (MECs), both leveraging organic substrates such as biomass or waste for H2 evolution through bacterial metabolism. Dark fermentation yields typically range from 1 to 2.5 moles of H2 per mole of glucose, limited by electron diversion to biomass growth and byproduct formation, while MECs can enhance recovery by applying a low external voltage to drive cathodic H2 production from anodic oxidation products.

Dark Fermentation by Anaerobic Bacteria

Dark fermentation involves the anaerobic catabolism of carbohydrates by fermentative bacteria, primarily strict anaerobes like Clostridium species and facultative anaerobes such as Enterobacter, producing alongside volatile fatty acids (e.g., acetate, butyrate) and CO2. The process proceeds via glycolytic breakdown of hexoses to pyruvate, followed by ferredoxin-mediated hydrogenase activity that releases from reduced ferredoxin, with pathways favoring acetate (theoretical maximum of 4 mol /mol glucose) over butyrate (2 mol /mol glucose) for higher yields. Actual yields seldom exceed 25-30% of the theoretical due to factors including pH sensitivity (optimal 5.5-6.5), temperature (mesophilic 30-40°C or thermophilic >50°C), and inhibition by partial pressure or methanogenic competitors, which consume to form CH4. Substrates like starch-rich wastes or lignocellulosic hydrolysates are commonly used, with pretreatment (e.g., acid or enzymatic) enhancing accessibility, though end-product accumulation often necessitates downstream for yield optimization. Recent efforts focus on mixed consortia from natural sources or heat-pretreated to suppress non- producers, achieving up to 2.8 mol /mol glucose in batch systems.

Microbial Electrolysis Cells

Microbial electrolysis cells integrate bioelectrochemical systems where anode-respiring bacteria (e.g., or ) oxidize organic matter, transferring electrons extracellularly to the ; these electrons migrate to the upon a minimal applied voltage (typically 0.3-0.8 V), reducing protons to via hydrogen-evolving reactions. Unlike pure , MECs circumvent thermodynamic barriers to acetate oxidation, potentially recovering up to 8 /mol glucose when fed effluents, with efficiencies approaching 90-100% under optimized conditions like gas-sparging or biocathodes. Key challenges include electrode overpotentials, stability, and energy input for voltage supply, though recent advances incorporate or wind-derived power and stacked cell designs to achieve energy-neutral operation, with reported production rates of 0.5-2 m³/m³ reactor/day in continuous-flow setups. -MEC systems sequentially process substrates, converting acids to additional , yielding 5-7 total /mol glucose and treating simultaneously, as demonstrated in pilot studies since 2010. Scalability remains limited by and catalysts, prompting research into non-precious metal alternatives like foams.

Dark Fermentation by Anaerobic Bacteria

Dark fermentation involves the anaerobic conversion of organic substrates into biohydrogen by strictly anaerobic or facultative anaerobic , occurring without light input and relying on fermentative to generate H₂ as a byproduct. This process typically utilizes carbohydrate-rich feedstocks such as glucose, starch, or , where break down substrates via to pyruvate, followed by pyruvate oxidation to and formate or acetate, with hydrogen evolution mediated by [Fe-Fe]-hydrogenases or ferredoxin-dependent pathways. Unlike light-dependent methods, dark fermentation enables continuous operation independent of diurnal cycles, making it suitable for integrating with systems. The primary biochemical pathway begins with substrate and acidogenesis, where hexoses like glucose are metabolized to yield up to 4 moles of H₂ per mole of glucose theoretically under optimal acetate-forming conditions, as electrons from pyruvate reduce protons via enzymes. However, competing pathways leading to butyrate, , or formation reduce yields to 1–2.5 moles H₂ per mole glucose in practice, with thermodynamic constraints limiting further H₂ recovery from reduced end products like without additional processes. Key enzymes include pyruvate: (PFOR) for ferredoxin reduction and bidirectional for H₂ production, though sensitivity to oxygen and inhibitors like can impair activity. Optimal conditions favor mesophilic (30–40°C) or thermophilic (55–60°C) temperatures, acidic (5.0–6.0) to suppress methanogens, and short hydraulic retention times to favor acidogens over acetogens. Prominent hydrogen-producing anaerobes include species from the genus , such as C. butyricum, C. beijerinckii, and C. acetobutylicum, which dominate mixed consortia due to their robust glycolytic and capabilities; thermophilic strains like Thermoanaerobacterium spp. offer advantages in suppression and higher rates at elevated temperatures. Facultative anaerobes like and can initiate H₂ production in mixed cultures but yield less efficiently than strict anaerobes. Enrichment strategies often involve heat-shock pretreatment (e.g., 100°C for 30 minutes) of inocula to select spore-forming while eliminating H₂-consuming methanogens and homoacetogens. Substrates from agricultural residues, food waste, or enhance feasibility, with reported yields up to 2.8 mol H₂/mol from glucose using optimized Clostridium strains, though real-world mixed wastes yield 0.5–1.5 mol H₂/mol due to inhibitory compounds like lignins or high . Challenges include low (10–20% of to H₂) and byproduct accumulation requiring downstream valorization, yet dark fermentation's simplicity positions it as a foundational step in two-stage systems combining with photofermentation or for higher overall yields. Recent advances, such as additives or of hydrogenases, aim to mitigate bottlenecks but remain lab-scale as of 2024.

Microbial Electrolysis Cells

Microbial electrolysis cells (MECs) are bioelectrochemical systems that utilize electroactive microorganisms to convert organic substrates into gas, requiring an external voltage input of typically 0.3–1.0 V to thermodynamically favor the process. In the anode compartment, oxidize organics such as or wastewater-derived carboxylates, transferring electrons extracellularly via mechanisms like direct contact or mediated shuttles to the . These electrons migrate through an external circuit to the , where protons from the anolyte combine to evolve H₂ through water reduction, often catalyzed by abiotic materials like or biological hydrogenases. Unlike dark , MECs achieve near-theoretical yields of 8–12 moles H₂ per mole of glucose equivalent by avoiding energy losses to biomass growth or , provided methanogens are suppressed. Key microbial players at the anode include Geobacter sulfurreducens and Shewanella species, which form biofilms capable of high electron transfer rates, while cathode performance depends on pH, electrode spacing, and materials to minimize overpotentials. Operational parameters such as substrate concentration (e.g., 1–10 g/L COD), hydraulic retention time (4–24 hours), and applied voltage critically influence Coulombic efficiency (up to 90%) and hydrogen recovery (70–95%), with optimal pH around 7 for balanced microbial activity. For instance, in a 2023 study using acetate-fed MECs, a voltage of 0.8 V yielded 2.5–3.0 m³ H₂ per m³ anolyte per day, demonstrating scalability potential when integrated with membrane separators to prevent gas crossover. MECs excel in valorizing dark fermentation effluents, where volatile fatty acids accumulate, boosting overall biohydrogen yields from complex biomass by up to 50% compared to standalone . Recent advancements include self-sustaining designs powered by microbial fuel cells or inputs, reducing net energy demands to near-zero, as reported in 2025 reviews achieving 1.5–2.0 kg H₂ per kg COD removed. However, challenges persist, including biofouling of electrodes, high capital costs for materials like carbon cloth or cathodes, and sensitivity to inhibitors like , limiting commercial yields to 10–20% of theoretical maxima without optimization. Ongoing research emphasizes catalysts and stacked configurations to enhance volumetric rates toward 10–20 L H₂/L reactor/day.

Technological Developments

Strain Engineering and Optimization Strategies

Strain engineering for biohydrogen production primarily targets metabolic pathways to increase hydrogenase activity, redirect electron flux toward hydrogen evolution, and mitigate inhibitory factors such as oxygen sensitivity or competing fermentative products. In photobiological systems, modifications focus on enhancing photosynthetic to [FeFe]-hydrogenases while reducing or uptake hydrogenase activity. Metabolic engineering techniques, including gene overexpression, knockouts via CRISPR-Cas9, and fusion proteins, have demonstrated yield improvements of up to 15-fold in model strains. In microalgae like , fusing (FDX) to has increased rates by 4.5-fold by improving supply efficiency. (RNAi) silencing of ferredoxin-NADP⁺ reductase (FNR) elevates yields by 2.5-fold through diversion of reducing equivalents from linear flow. of (Y67A variant) achieves 10- to 15-fold higher accumulation by minimizing CO₂ fixation competition. (LHC) mutants yield 50% greater output under sulfur-deprived conditions, as reduced antenna size limits photodamage and oxygen production. Codon-optimized expression of bacterial genes like lba and hemH in chloroplasts boosts yields by 22% via enhanced maturation. Cyanobacterial engineering emphasizes nitrogenase-based production, with inactivation of uptake hydrogenase (hupL knockout) in strains such as Anabaena sp. PCC 7120 and Nostoc sp. PCC 7422 resulting in 4- to 7-fold higher rates under or atmospheres. Combined ΔhupL and homocitrate (nifV) disruptions sustain elevated production by preventing reconsumption and optimizing reductant allocation, achieving 20-30% (v/v) accumulation. Mutations in the nifD subunit (e.g., R284H) further enhance yields to 87% under N₂-fixing conditions by favoring over synthesis. For dark fermentative bacteria, metabolic rerouting in species targets glycolytic and solventogenesis pathways. Overexpression of in strains increases hydrogen yields by 1.15- to 1.39-fold compared to wild-type by amplifying NADH availability for [FeFe]-hydrogenase. CRISPR-Cas9-mediated knockouts of and genes in Enterobacter aerogenes redirect flux, improving yields through reduced byproduct formation. In Clostridium pasteurianum, engineering phosphotransferase systems enhances substrate uptake and hydrogen output, addressing thermodynamic limits in acetate-butyrate pathways. These modifications, often combined with adaptive evolution, prioritize theoretical maximum yields approaching 4 mol /mol glucose, though practical gains remain constrained by and balance.

Bioreactor Configurations and Scale-Up Challenges

Photobioreactors (PBRs) are essential for light-dependent biohydrogen production pathways, such as direct and indirect biophotolysis in and , as well as photofermentation by non-sulfur . Common configurations include tubular PBRs, which circulate algal suspensions through transparent tubes to optimize light exposure but suffer from high energy demands for pumping; flat-plate PBRs, offering better light distribution and mixing via air sparging; and vertical-column PBRs, which leverage natural for reduced on cells. These closed systems enable precise control of environmental parameters like , temperature, and nutrient supply, achieving hydrogen yields up to 1.6 mol H₂ per mol glucose equivalent in optimized lab-scale setups with . For dark fermentation by , configurations prioritize retention and suppression; continuous stirred-tank reactors (CSTRs) provide uniform mixing and stable operation at hydraulic retention times of 6-24 hours, yielding 1.5-2.5 mol H₂ per mol glucose, while (UASB) reactors and packed-bed designs enhance for higher cell densities but risk clogging. bioreactors (MBRs) integrate to prevent washout, boosting yields by 20-50% in continuous modes compared to batch systems. Hybrid configurations, such as immobilized cell systems in fluidized-bed or gas-lift reactors, address limitations in both pathways by improving gas-liquid and reducing inhibition from byproducts like oxygen or organic acids; for instance, alginate-immobilized Clostridium species in fluidized beds have demonstrated sustained rates of 10-15 L H₂/L reactor/day under mesophilic conditions. Operational strategies often involve two-stage systems, where dark fermentation effluents feed photofermenters, potentially increasing overall yields to 8-12 H₂ per glucose through complementary microbial consortia. Scale-up to industrial volumes exceeding 100 m³ encounters engineering hurdles rooted in biophysical constraints. In PBRs, light attenuation follows the Beer-Lambert law, limiting effective penetration to outer cell layers in dense cultures (>10 g/L ), resulting in productivity drops of 50-70% beyond 0.1 m depth and necessitating costly artificial illumination or thin-layer designs that inflate land and material costs. Hydrodynamic introduces forces that damage fragile algal cells or disrupt bacterial biofilms, with power inputs cubically while volumes scale linearly, elevating to 5-10 kWh/m³ in large systems. For fermentative bioreactors, challenges include contamination by methanogens, which compete for s and reduce net hydrogen yields by up to 40% without stringent sterilization; dead zones and channeling in UASB or packed beds exacerbate uneven distribution at scales >10 m³, lowering efficiencies from 30-40% in lab tests to <20% in pilots. Gas handling poses further issues, as poor H₂ solubility demands efficient sparging and separation to avoid explosive mixtures, while byproduct accumulation (e.g., VFAs) inhibits enzymes like hydrogenases, necessitating pH control and effluent recycling that complicate continuous operation. Economic analyses indicate capital costs for PBR scale-up at $200-500/m² surface area, with operational expenses dominated by mixing and cooling, rendering current yields (typically <5% of theoretical maxima) insufficient for competitiveness against steam methane reforming. Mitigation efforts focus on computational fluid dynamics modeling for design optimization and genetic strain enhancements to tolerate scale-induced stresses, though pilot demonstrations remain limited to <1 m³ volumes as of 2023.

Integration with Biomass Waste Streams

Biohydrogen production integrates with biomass waste streams primarily through dark fermentation processes, where anaerobic bacteria convert organic components of waste into hydrogen gas, acetic acid, and other byproducts, thereby valorizing otherwise discarded materials and mitigating environmental burdens from landfilling or incineration. This approach leverages the high carbohydrate and lignocellulosic content in wastes such as agricultural residues (e.g., crop stalks, straw), food scraps, and sewage sludge, which serve as low-cost feedstocks without competing with food production. Integration facilitates a circular economy by coupling hydrogen recovery with waste pretreatment steps like hydrolysis and acidogenesis, though practical hydrogen yields remain constrained by microbial thermodynamics and substrate complexity. Agricultural biomass wastes, including lignocellulosic materials like maize straw and rice husks, undergo dark fermentation after pretreatment (e.g., alkaline or enzymatic hydrolysis) to break down recalcitrant structures, yielding 1.2–2.3 mol H₂ per mol hexose equivalent, equivalent to 30–50% of the theoretical maximum under the Thauer limit of 4 mol H₂/mol glucose. For instance, optimized dark fermentation of raw at 36°C, 20 g/L biomass loading, and controlled pH achieves modeled hydrogen outputs tied to hydrolytic efficiency, though inhibition from lignocellulose-derived phenolics often limits scalability. Vegetable and crop wastes further demonstrate potential, with recent studies reporting enhanced yields via microbial consortia, addressing waste management while producing clean energy carriers. Co-digestion of complementary streams, such as food waste and sewage sludge, boosts integration by balancing nutrient profiles (e.g., high carbon in food waste offsetting nitrogen in sludge), resulting in hydrogen yields up to 92.5 mL H₂/g volatile solids—a 90% increase over mono-digestion of food waste alone—through synergistic microbial activity and reduced inhibition. Pretreatments like heat shock or ultrasonication further optimize these systems; for sewage sludge, combined acidic-alkaline methods have been shown to elevate yields by disrupting microbial flocs and enhancing substrate accessibility. Such hybrid feedstocks from municipal sources enable decentralized production, but persistent challenges include variable waste composition, methane contamination risks, and the need for downstream effluent treatment to recover residual biogas potential. Overall, while integration reduces reliance on virgin biomass and aligns with sustainability goals, empirical data underscore yield gaps—often below 2 mol H₂/mol substrate—due to incomplete hydrolysis and competing metabolic pathways favoring solvents over gas, necessitating strain engineering and process refinements for viability. Post-fermentation management, including broth recycling or anaerobic digestion of residues, closes material loops but requires validation at pilot scales to confirm net energy gains over conventional waste handling.

Economic and Practical Viability

Cost Structures and Yield Limitations

The primary cost structures in biohydrogen production encompass high capital expenditures for specialized bioreactors and photobioreactors, which can account for up to 90% of total costs in light-dependent processes due to requirements for light distribution, gas separation, and sterility maintenance. Operational expenditures include nutrient media (e.g., nitrogen sources like yeast extract), substrate pretreatment (up to 32% of costs for lignocellulosic feedstocks), and energy for mixing and purification, with overall production costs ranging from $3.2–48.96/kg H₂ for dark fermentation and $3.7–7.61/kg H₂ for photo-fermentation. These figures render biohydrogen uneconomical compared to steam methane reforming at $1–2/kg H₂, as biological processes demand continuous microbial culturing and yield low gas volumes necessitating expensive downstream separation. Yield limitations stem from inherent biological constraints, including thermodynamic barriers in dark fermentation, where maximum theoretical yields of 4 mol H₂/mol glucose are rarely exceeded due to competing pathways producing volatile fatty acids and alcohols, resulting in actual yields of 1–3.9 mol/mol in optimized strains. In photobiological methods, oxygen sensitivity of hydrogenase enzymes inhibits production, capping direct biophotolysis yields at 0.015–1.084 mmol H₂/L/h and solar-to-hydrogen efficiencies below 10%, while high partial pressure of H₂ feedback-represses nitrogenase activity. Substrate complexity further reduces yields, as lignocellulosic or waste-derived inputs require hydrolysis but generate inhibitors like furans and heavy metals that disrupt microbial metabolism.
ProcessTheoretical Yield (mol H₂/mol glucose equiv.)Typical Actual YieldKey Limitation
Dark Fermentation41–3.9Byproduct diversion to acids
Photo-fermentation120.21–2.03Light saturation and N₂ase inhibition
Direct Biophotolysis12<1 (rates 0.015–1.084 mmol/L/h)O₂ inactivation of enzymes
These constraints, compounded by low production rates (e.g., 20–2292 mL H₂/L/h in dark systems), preclude industrial scalability without genetic or process engineering breakthroughs, as current yields achieve only 25–50% of theoretical maxima.

Comparisons to Conventional Hydrogen Production

Conventional hydrogen production is dominated by steam methane reforming (SMR), which accounts for approximately 95% of global hydrogen output, utilizing natural gas as feedstock to achieve efficiencies of 74-85% on a lower heating value basis. In contrast, biohydrogen production via microbial processes like dark or photofermentation yields efficiencies typically ranging from 20-30%, constrained by thermodynamic limitations that cap hydrogen output at 2-4 moles per mole of glucose equivalent, far below the theoretical maximum of 12 moles. These biological pathways divert carbon to biomass growth and byproducts like acetate, reducing net hydrogen recovery and necessitating downstream purification that further diminishes overall process viability. Production costs for SMR average $0.75-2 per kg of hydrogen, benefiting from mature infrastructure and economies of scale, while electrolysis costs range from $2.6-7 per kg, dependent on electricity pricing and renewable integration. Biohydrogen costs, however, span $2-10 per kg in optimistic projections but often exceed $12 per kg in practice, driven by low yields, feedstock pretreatment needs, and bioreactor operational challenges such as microbial inhibition and contamination. Recent U.S. Department of Energy efforts have reduced biohydrogen costs from over $58 per kg to around $12.4 per kg through strain engineering, yet this remains uncompetitive without subsidies.
MetricSMRElectrolysis (Renewable)Biohydrogen (Fermentation)
Efficiency (%)74-85 [web:30]62-82 [web:30]20-30 [web:31]
Cost ($/kg H₂)0.75-2 [web:30]2.6-7 [web:30]2-12+ [web:36][web:32]
GHG Emissions (kg CO₂ eq/kg H₂)11-12 [web:30]0.5-1 [web:30]1-4.7 [web:30]
Biohydrogen offers environmental advantages, with lifecycle greenhouse gas emissions of 1-4.7 kg CO₂ equivalent per kg hydrogen versus 11-12 kg for SMR without carbon capture, particularly when utilizing waste biomass to avoid net carbon addition. However, scalability remains a barrier for biohydrogen due to biological rate limitations and feedstock logistics, whereas SMR and electrolysis support gigawatt-scale plants with established supply chains. Electrolysis edges biohydrogen in green credentials when powered by renewables, as biological processes still incur indirect emissions from cultivation or pretreatment, underscoring biohydrogen's niche potential in waste valorization rather than bulk production.

Barriers to Industrial Scalability

Biohydrogen production faces significant hurdles in achieving industrial scalability, primarily due to inherently low hydrogen yields across biological pathways. In dark fermentation, practical yields typically range from 1 to 2 mol H₂ per mol of glucose equivalent, far below the theoretical maximum of 4 mol, owing to thermodynamic constraints and the accumulation of volatile fatty acids (VFAs) like acetate and butyrate that inhibit further hydrogenase activity. Photobiological processes, such as direct biophotolysis, suffer even lower productivities (e.g., 0.21 mmol H₂/L/h), exacerbated by oxygen sensitivity requiring levels below 0.1% to prevent enzyme deactivation. These yield limitations necessitate oversized reactors and high substrate inputs, rendering processes inefficient for large-scale output comparable to chemical methods like steam methane reforming, which achieve yields exceeding 90% efficiency. Technical challenges in bioreactor design and operation further impede scale-up from laboratory (0.1–3 L) to pilot or industrial volumes (e.g., >1000 L). Continuous stirred-tank reactors (CSTRs) experience washout at hydraulic retention times under 6–12 hours, while photobioreactors (PBRs) grapple with light attenuation, uneven distribution, and , reducing by up to 50% in deeper cultures. from mixing damages sensitive algal or bacterial cells, and gas-liquid mass transfer limitations hinder CO₂ stripping and H₂ recovery, with contamination by diverting substrates to non-hydrogen pathways. Outdoor deployments amplify variability from fluctuating light, temperature, and inhibitors like or furans from lignocellulosic feedstocks, as demonstrated in pilot trials where yields dropped 30–40% under real conditions. Economic viability remains elusive, with production costs estimated at $3.2–48.96/kg H₂ for dark fermentation and up to $1362/kg for photofermentation, driven by feedstock pretreatment (up to 32–50% of total costs), nutrient supplementation, and downstream purification to separate H₂ from CO₂ and impurities. Capital expenditures for large PBRs or digesters are prohibitive due to needs (e.g., corrosion-resistant linings) and requirements, while operational expenses include for stirring, heating, and gas handling, often resulting in negative energy balances without waste integration. These figures contrast sharply with conventional at $1–2/kg, underscoring the need for yields >50 L H₂/L reactor/day—rarely achieved beyond lab scales—to compete. Biological constraints, including microbial sensitivity and process , compound scalability issues. Hydrogen-producing strains like Clostridium spp. or Rhodobacter are prone to metabolic shifts toward solvent production under , and genetic limits long-term performance in continuous cultures. Strict anaerobiosis demands costly gas sparging and sealing, while substrate inhibition by high organic loads (e.g., >20 g/L ) halts , as seen in wastewater-based pilots. Hybrid systems combining dark and photo improve yields (e.g., 8–12 mol H₂/mol glucose) but introduce complexity in sequential reactor management and microbial compatibility, delaying commercialization. Overall, these intertwined barriers have confined biohydrogen to demonstration plants, with no widespread industrial adoption as of 2024 despite decades of research.

Historical Context

Early Scientific Foundations (Pre-1970s)

The foundational understanding of biohydrogen production prior to the 1970s stemmed from microbiological discoveries revealing microbial enzymes and pathways capable of evolving molecular hydrogen from organic substrates or water. In 1931, Marjory Stephenson and Leonard H. Stickland isolated and characterized hydrogenase from bacteria including Escherichia coli, describing it as an enzyme that reversibly activates H2 to enable its oxidation or reduction of compounds such as sulfate to sulfide or methylene blue. This work established the enzymatic basis for biological H2 metabolism, primarily through observations of H2-dependent reductions in cell-free extracts and intact cultures under anaerobic conditions. Early investigations into dark fermentation demonstrated H2 evolution as a byproduct of carbohydrate breakdown by strict anaerobes like Clostridium pasteurianum, where pyruvate is cleaved via ferredoxin-dependent to yield H2 and , with yields approaching 2-3 mol H2/mol glucose under optimal mesophilic conditions. These processes, noted in bacterial cultures as early as , underscored the thermodynamic favorability of H2 release to regenerate NAD+ and maintain balance during , though rates were low without genetic or environmental enhancements. Pioneering algal studies by Hans Gaffron in the 1940s further expanded the scope to photosynthetic organisms. In 1942, Gaffron reported that hydrogen-adapted Scenedesmus obliquus produced H2 fermentatively in darkness at rates of several microliters per hour per mg dry weight and accelerated photoproduction upon illumination, reaching up to 10 times higher yields after adaptation. This revealed [FeFe]-hydrogenase activity in chloroplasts, enabling H2 evolution from protons and electrons derived from or starch reserves, challenging the then-dominant view of as solely CO2-fixing. Such findings highlighted oxygen-sensitive biophotolysis but noted inhibitions by O2 accumulation, limiting practical yields to under 1% of theoretical solar conversion efficiency.

Post-Oil Crisis Advancements (1970s-2000s)

The 1973 and 1979 oil crises catalyzed renewed investment in alternative energy pathways, including biological , as nations sought to reduce dependence on fossil fuels through solar-driven microbial processes. In the United States and , government agencies like the U.S. Department of Energy initiated funding for photobiological hydrogen research, emphasizing and photosynthetic bacteria for their ability to split water or ferment organics into H₂ without external energy inputs beyond . Early 1970s efforts focused on direct biophotolysis in eukaryotic such as , where enzymes catalyze H₂ evolution under anaerobic conditions following oxygenic . Building on pre-1970s discoveries, applied studies from 1970 onward quantified photosynthetic H₂ and O₂ production , revealing turnover times of 0.1 to 3 milliseconds in species like Scenedesmus obliquus and Chlorella fusca, though oxygen sensitivity severely limited net yields to below 1% of theoretical solar conversion efficiency. Researchers explored sulfur deprivation protocols to induce anaerobiosis, enabling transient H₂ production rates of approximately 20-50 μmol H₂ per mg per hour in lab-scale cultures. The 1980s saw diversification into indirect biophotolysis and photofermentation using and purple non-sulfur bacteria like , which avoided O₂ inhibition by separating from H₂ evolution in two-stage systems. Historical reviews note initial demonstrations of hybrid processes, where dark-fermenting anaerobes like species produced H₂ from , followed by bacterial photo-upgrading of effluents, achieving combined yields up to 20-30% of substrate energy content as H₂. innovations, including immobilized cell systems, emerged to maintain microbial activity, with pilot experiments reporting volumetric H₂ production of 1-5 L H₂ per L reactor per day under continuous illumination. By the 1990s, strain optimization and early targeted O₂ tolerance and substrate utilization, with genetic knockouts in C. reinhardtii enhancing sustained photoproduction. Japanese and U.S. programs reported incremental gains, such as H₂ yields of 100-200 mL per g dry in algal suspensions, alongside explorations of waste-derived feedstocks for dark . Despite these developments, persistent challenges like low quantum efficiencies (typically 1-5%) and instability confined progress to laboratory scales, underscoring the gap between mechanistic insights and practical viability.

Contemporary Research (2010s-Present)

Research since the 2010s has advanced through optimizations in and , systems, and microbial engineering, targeting higher yields from diverse feedstocks like organic wastes and . using Clostridium species on carbohydrates achieves yields of 1–2 mol H₂ per mol substrate, with process parameters such as 5–6 and additives like nano zero-valent iron enhancing and reducing inhibition. In 2023, reactors with and Clostridium cultures reached 208.3 L H₂/L/day at 4.4 and 8-hour hydraulic retention time, demonstrating improved volumetric productivity over continuous stirred tank reactors' 39.65 L H₂/L/day. Photofermentation with Rhodobacter species converts organic acids to up to 3.5–4 mol H₂ per mol at light intensities of 4000–6000 , while two-stage dark-photo systems attained 77% of theoretical maximum yield in 15 days during the . Integration of dark fermentation effluents into microbial cells has boosted overall hydrogen recovery by 50–70% through applied voltage and conductive materials, as reported in studies on food waste and effluents yielding 2.1 mol H₂/mol glucose with Clostridium beijerinckii in 2024. Food waste co-fermentation produced up to 127 mL H₂/g volatile solids in optimized conditions by 2023. Microalgal biophotolysis, particularly in , has seen to improve oxygen-tolerant enzymes and deprivation protocols, addressing sensitivity bottlenecks since 2010, though solar-to-hydrogen efficiencies remain below 5% without breakthroughs in scaling. and additives in enhance substrate conversion, with pretreated to yield up to 17.5 g H₂/100 g biomass. models, including artificial neural networks with R² > 0.987, enable prediction and adjustment of operational parameters like organic loading rates across reactor types. Efforts in strain include co-cultures and genetic modifications, such as 2017 enhancements in activity for higher accumulation in fermentative . These strategies, applied to waste streams like and cheese , prioritize carbon-neutral pathways but face persistent challenges in inhibitor management and economic scalability, with lifecycle analyses highlighting energy inputs for pretreatment as key hurdles.

Criticisms and Realistic Assessments

Technical and Biological Shortcomings

Biological processes, including photobiological systems and dark fermentation, suffer from inherently low yields due to thermodynamic and metabolic constraints. In dark fermentation, the maximum theoretical hydrogen yield is 4 s of H₂ per of glucose, but practical yields rarely exceed 1-2 s, limited by the cessation of the process at formation and the diversion of electrons to non-hydrogen pathways like solvent production. Photobiological via or achieves even lower efficiencies, with solar-to- conversion rates typically below 1-5%, far short of the 10-12% theoretical limit for , due to inefficient light harvesting and . A primary biological limitation is the oxygen sensitivity of key enzymes such as [FeFe]-hydrogenases, which catalyze H₂ evolution but are irreversibly inhibited by O₂ levels as low as 0.1-1%, conflicting with the O₂-generating nature of in direct photolysis systems. This necessitates indirect strategies like deprivation in (e.g., ), which temporarily represses but yields transient H₂ production rates of only 10-50 mL/L/h, declining after days due to nutrient depletion and cell stress. In anaerobic dark fermenters, microbial consortia often include H₂-consuming methanogens or homoacetogens, reducing net yields by up to 50% unless suppressed via heat pretreatment or selective inhibitors, though these add complexity and incomplete efficacy. Technically, biohydrogen processes face scalability barriers from poor and reactor instability. Continuous bioreactors experience washout at hydraulic retention times below 6-12 hours, limiting productivity to 1-10 L H₂/L reactor/day, while batch systems suffer from startup delays and inconsistent performance. -dependent systems require uniform illumination, but dense algal cultures (>1 g/L ) attenuate penetration beyond 1-5 cm, necessitating energy-intensive mixing or thin-layer designs that increase and capital costs. Downstream purification is hindered by H₂ comprising only 20-60% of the mixture (with CO₂ and traces of H₂S or NH₃), demanding costly or membrane separation, which can consume 20-30% of the produced H₂'s energy value. End-product inhibition further constrains biology: in dark fermentation, accumulated volatile fatty acids (e.g., at >5 g/L) lower and inhibit activity, capping conversion at 30-50% without control or effluent removal. efforts to enhance yields, such as knocking out competing pathways in species, have boosted lab-scale outputs by 20-50% but falter in mixed cultures due to risks and regulatory hurdles for open systems. Overall, these factors result in biohydrogen yields orders of magnitude below chemical methods like steam reforming (50-70% ), underscoring the need for hybrid integrations rather than standalone viability.

Overhyped Promises and Environmental Trade-offs

Proponents of biohydrogen have frequently promoted it as a pathway to scalable, carbon-neutral using , , or , potentially revolutionizing without fossil fuel dependence. However, practical yields remain constrained by biological s, such as the oxygen sensitivity of enzymes in photobiological systems, resulting in solar-to- efficiencies typically below 1-5%, far short of the 10% threshold needed for economic viability. Dark processes, while avoiding light dependency, achieve yields of only 1-2 s per mole of glucose under optimal conditions, well below the theoretical Thauer of 4 moles, due to competing metabolic pathways favoring growth over gas production. These persistent shortfalls, documented across decades of , underscore how initial optimism overlooked inherent thermodynamic inefficiencies in diverting photosynthetic or fermentative toward rather than cellular maintenance. Energy balance assessments further reveal overhyped expectations, with many integrated biohydrogen systems exhibiting net energy ratios (NER) below 1, indicating greater input for cultivation, processing, and purification than output. For instance, a modeled biological hydrogen-methane system from organic waste yielded an NER of 77.8% and a negative net balance of -738.4 kWh per batch, highlighting the energetic penalties of pretreatment, mixing, and gas separation. Such results contrast with promotional claims of self-sustaining cycles, as external inputs like heating or often dominate, rendering biohydrogen a net energy sink in real-world deployments rather than a surplus provider. Environmental trade-offs compound these issues, as photobiological production via demands extensive open ponds or closed reactors covering vast land areas—potentially competing with or ecosystems—while consuming significant through evaporation and nutrient inputs like and . Life-cycle analyses indicate that reduced amplify impacts across categories such as and , with a 20% yield drop raising overall environmental burdens by 25-26%. Although integrating streams can offset some demands and enable circular use, scaling to industrial levels risks localized stress in arid regions and runoff, negating purported gains without stringent management. These trade-offs, often downplayed in advocacy, prioritize biological novelty over the lower land and footprints of alternatives like - or solar-powered .

Policy and Market Distortions

Government subsidies for clean hydrogen technologies, including potential applicability to biohydrogen under frameworks like the U.S. Inflation Reduction Act's Section 45V —finalized on January 8, 2025, and offering up to $3 per kilogram for production with lifecycle below 0.45 kg CO2e per kg —create incentives that overlook biohydrogen's inherent inefficiencies and high costs. Biohydrogen production, reliant on microbial processes with yields typically below 4 mol per mol glucose and requiring costly bioreactors and feedstock preprocessing, remains uncompetitive against steam reforming at $1-2 per kg without such support, distorting capital allocation toward biologically constrained pathways over scalable alternatives like paired with renewables. These policies, driven by decarbonization mandates, fail to reward biohydrogen's potential carbon negativity—achievable via waste but unaccounted in credit multipliers—perpetuating underpricing of environmental and economic trade-offs. In the , biohydrogen garners limited but targeted funding through research programs like , yet regulatory emphasis on electrolytic undervalues biological routes, fostering uneven market signals that inflate research hype without commercialization pathways. Broad support mechanisms, such as premiums and cap-and-trade interactions, can exacerbate distortions by incentivizing intermittent renewable integration for over direct , indirectly subsidizing biohydrogen's oxygen-sensitive fermentations that demand controlled, energy-intensive conditions. Analogous to mandates, which since the have diverted and inflated without proportional emissions reductions, biohydrogen subsidies risk similar inefficiencies by propping up low-efficiency processes (e.g., 1-5% overall solar-to- in photobiological systems) amid academic and institutional biases toward "sustainable" narratives. Critics contend that such interventions, exemplified by U.S. Department of Energy's historical outlays for -derived R&D (e.g., $15 million in 2010s-era grants for integrated , including biological steps), mask fundamental barriers like microbial inhibition and contamination, channeling public funds into perennial pilot-scale demonstrations rather than cost-reflective innovation. Recent fiscal restraint, including DOE's October 2025 cancellation of remaining hub grants totaling billions—prioritizing viable electrolytic over niche biological production—highlights growing recognition of these distortions, potentially reallocating resources to technologies with proven industrial yields exceeding biohydrogen's lab-constrained outputs. Without phase-outs, however, policies continue to hinder true market pricing, delaying 's role in energy systems where bio routes contribute marginally to global supply projections below 1% by 2050.

References

  1. [1]
    Hydrogen Production Processes | Department of Energy
    Microbes such as bacteria and microalgae can produce hydrogen through biological reactions, using sunlight or organic matter. These technology pathways are in ...Missing: facts | Show results with:facts
  2. [2]
    Towards industrial biological hydrogen production: a review - PMC
    Dec 7, 2023 · This technology uses micro-organisms to convert organic matter into hydrogen gas, a clean and versatile fuel that can be used in a wide range of applications.
  3. [3]
    A review on biohydrogen production technology - AIP Publishing
    Oct 3, 2024 · This article reviews the recent investigations on biohydrogen production methods, as well as sources of this valuable fuel.
  4. [4]
    A Review on Biohydrogen Sources, Production Routes, and Its ...
    Biohydrogen is a natural or transient byproduct of several microbial-mediated biochemical reactions. It can be produced either by a biological process or the ...
  5. [5]
    Advancements in biohydrogen production - RSC Publishing
    Nov 18, 2024 · Biohydrogen production offers significant environmental and operational benefits over conventional chemical hydrogen generation methods.
  6. [6]
    Biological Routes for Biohydrogen Production: A Clean and Carbon ...
    Jul 28, 2025 · Bio‐H2 production utilizes water, organic substances, and gases as substrates, with various technologies available based on different microbial ...Missing: facts | Show results with:facts
  7. [7]
    Genetic engineering for biohydrogen production from microalgae - NIH
    Hydrogen production from microalgae is considered a clean and sustainable energy production method that can both alleviate fuel shortages and recycle waste.
  8. [8]
    Debottlenecking the biological hydrogen production pathway of dark ...
    Aug 19, 2022 · Biological hydrogen production from renewable biomass or waste materials by dark fermentation is a promising alternative to conventional routes ...
  9. [9]
    Trends in biohydrogen production: major challenges and ... - PubMed
    Current biohydrogen production technologies, however, face two major challenges such as low-yield and high production cost. Advances have been made in recent ...
  10. [10]
    Advancements in biohydrogen production – a comprehensive ... - NIH
    Nov 18, 2024 · Biohydrogen production offers significant environmental and operational benefits ... Batch reactors face several challenges, especially in ...
  11. [11]
    A review on biological biohydrogen production: Outlook on genetic ...
    Oct 20, 2023 · This review mainly focuses on bio‐hydrogen production via biological pathways, genetic improvements, knowledge gap, economics, and future directions.
  12. [12]
    Biological Hydrogen Production: A Comprehensive Review for ...
    Jun 13, 2024 · This study reviews the main types and technical principles of microbial hydrogen production from waste, available waste types, research progress ...
  13. [13]
    Biohydrogen production from waste materials: benefits and challenges
    Nov 6, 2019 · Cheap and abundant material availability is the key advantage of MSW over other wastes. In addition, it contains both macro- and micronutrients ...Missing: advantages | Show results with:advantages
  14. [14]
    Hydrogenase and Nitrogenase: Key Catalysts in Biohydrogen ...
    Feb 1, 2023 · The core catalysts of biohydrogen production are two types of enzymes, namely, hydrogenase and nitrogenase, and each type of enzyme can be ...
  15. [15]
    Hydrogen Production: Microbial Biomass Conversion
    In direct hydrogen fermentation, the microbes produce the hydrogen themselves. These microbes can break down complex molecules through many different pathways, ...
  16. [16]
    Enzymatic and Bioinspired Systems for Hydrogen Production - PMC
    Hydrogenases promote reversible proton reduction to hydrogen in a variety of anoxic bacteria and algae, displaying unparallel catalytic performances. Attempts ...
  17. [17]
    Solved 2H2O(I) 2H2( g)+O2( g) The standard free energy - Chegg
    Sep 21, 2023 · 2H2O(I)⟶2H2( g)+O2( g) The standard free energy change for this reaction is 474.2 kJ. The free energv change when 1.52 moles of H2O(I) react at ...
  18. [18]
    Thermophilic biohydrogen production: how far are we? - PMC
    Aug 16, 2013 · However, re-oxidation of the latter to H2 is inherent to a thermodynamic constraint and thus instead it might easily be oxidized to ...
  19. [19]
    Photobiological H2 Production: Theoretical Maximum Light ...
    Mar 1, 2020 · The above- described physical parameters set the theoretical maximum solar-conversion efficiency of biological systems to 13%. However, due ...<|separator|>
  20. [20]
    Photobiological hydrogen production : photochemical efficiency and ...
    The main drawback of the photoautotrophic hydrogen production process is oxygen inhibition. The few efficiencies reported on the conversion of light energy into ...Missing: limits | Show results with:limits
  21. [21]
    DOE Technical Targets for Photobiological Hydrogen Production
    Technology readiness targets (beyond 2020) are 5.5% efficiency of incident solar light energy to H2 (E0*E1*E2) from organic acids, 80% of maximum molar yield of ...Missing: limits | Show results with:limits
  22. [22]
    Biohydrogen Production: Strategies to Improve Process Efficiency ...
    Biological pathways for H2 production and the technical limitations. Type of Bioprocess, Technical Challenges. Dark fermentation. low substrate conversion ...
  23. [23]
    Light-dependent biohydrogen production: Progress and perspectives
    Light-dependent H2 production is a promising method derived from nature's most copious resources: solar energy, water and biomass. Reduced environmental impacts ...
  24. [24]
    Biological Routes for Biohydrogen Production: A Clean and Carbon ...
    Jul 28, 2025 · Biohydrogen (bio-H2) production pathways are well established and can be categorized into four main processes: (1) direct biological photolysis ...
  25. [25]
    A review of bioreactor configurations for hydrogen production by ...
    Jan 2, 2024 · 2.1. Microalgae. Green algae can produce hydrogen through direct and indirect biophotolysis. In direct biophotolysis, they split water into H2 ...
  26. [26]
    Biohydrogen production from microalgae for environmental ...
    This review explored the mechanism of biohydrogen production from microalgae containing direct biophotolysis, indirect biophotolysis, photo fermentation, and ...
  27. [27]
    Waste to Sustainable Biohydrogen Production Via Photo ...
    Nov 5, 2021 · In this review article we focus on two methods that are not widely used at industrial scale but have many future possibilities and growth margins.
  28. [28]
    effects of protonophores and inhibitors of responsible enzymes
    Sep 4, 2015 · Biohydrogen (H2) production by purple bacteria during photofermentation is a very promising way among biological H2 production methods.
  29. [29]
    Comparison of Photofermentative Hydrogen Production in ... - NIH
    Jun 14, 2025 · This study presents a scale-up from 0.2 L to 4.0 L of the photofermentation process using the purple non-sulfur bacterium Rhodopseudomonas sp. S16-VOGS3.
  30. [30]
    Genetic characterization of biohydrogen-producing purple non ...
    Oct 5, 2024 · Biohydrogen, a potential alternative energy source, is a widely studied product of PNSB photofermentation. It is mainly produced through the ...
  31. [31]
    Biohydrogen Production From Biomass Sources: Metabolic ...
    Sep 9, 2021 · The biophotolysis process is divided into two categories: indirect and direct biophotolysis. Endogenous substrates catabolize and produce ...
  32. [32]
    BioH2 Production Using Microalgae: Highlights on Recent ... - MDPI
    Biophotolysis is the initial step of microalgal bioH2 production. In direct biophotolysis, microalgae convert solar energy into chemical energy, and H2 is ...<|separator|>
  33. [33]
    Biophotolysis-based Hydrogen Production by Cyanobacteria and ...
    This mini review examines the technology status of direct and indirect biophotolysis for hydrogen production, and technical challenges for sustained ...
  34. [34]
    How Close We Are to Achieving Commercially Viable Large-Scale ...
    Photobiological H2 production by cyanobacteria is one of the options, and this review briefly discusses nitrogenase (N2ase)-based photobiological H2 production ...<|separator|>
  35. [35]
    High rates of photobiological H2 production by a cyanobacterium ...
    Dec 14, 2010 · In the present report, we describe Cyanothece sp. ATCC 51142, a cyanobacterial strain with the ability to produce remarkably high amounts of H2 ...
  36. [36]
    Bioprocesses of hydrogen production by cyanobacteria cells and ...
    The purpose of this review was to analyze the current state of research in the field of hydrogen release by cyanobacteria; systematize available evidence-based ...
  37. [37]
    Outlook on Synthetic Biology-Driven Hydrogen Production
    Photobiological hydrogen production offers a sustainable route to clean energy by harnessing solar energy through photosynthetic microorganisms.
  38. [38]
    Optimizing cyanobacterial hydrogen production: metabolic and ...
    Apr 8, 2025 · Improving hydrogen yields requires strategic metabolic and genetic modifications to optimize energy flow and overcome photosynthetic limitations.
  39. [39]
    Photo-Fermentative Bacteria Used for Hydrogen Production - MDPI
    Jan 31, 2024 · Photo-fermentation is an efficient hydrogen production pathway in which purple non-sulfur bacteria (PNSB) play an active role and produce ...
  40. [40]
    Metabolic pathways to sustainability: review of purple non-sulfur ...
    This review explores the potential of PNSB in upcycling AFW streams derived from various sources, such as fruit and vegetable residues, as well as effluents.Introduction · Carbohydrate metabolism in... · Volatile fatty acid metabolism...
  41. [41]
    Enhancement of biohydrogen production rate in Rhodospirillum ...
    Aug 6, 2021 · Rhodospirillum rubrum is a purple non-sulphur bacterium that produces H2 by photofermentation of several organic compounds or by water gas-shift ...
  42. [42]
    Biohydrogen Produced via Dark Fermentation: A Review - MDPI
    Dark fermentation is the most efficient and cost-effective method for producing biohydrogen, making it a key research focus. This article offers a comprehensive ...
  43. [43]
  44. [44]
    Microbes and Parameters Influencing Dark Fermentation for ... - MDPI
    Hydrogen production by dark fermentative bacteria from biomass mainly occurs via the acetate and butyrate pathways [25]. The maximum H2 yield through the ...<|separator|>
  45. [45]
    Bio-hydrogen production by dark anaerobic fermentation of organic ...
    Carbohydrate-rich substrates are degraded anaerobically by hydrogen-producing microorganisms such as facultative anaerobes and obligate anaerobes in dark ...
  46. [46]
    Microbiomes of biohydrogen production from dark fermentation of ...
    Mar 5, 2022 · In anaerobic digestion, hydrogen (H2) is produced during acidogenesis and acetogenesis, by hydrolytic and fermentative bacteria. It is later ...
  47. [47]
    Microbial electrolysis: a promising approach for treatment and ...
    Mar 17, 2022 · The microbial electrolysis cell (MEC) is one of the most efficient technologies for waste-to-product conversion that uses electrochemically active bacteria.
  48. [48]
    Recent Developments in Microbial Electrolysis Cell-Based ... - MDPI
    This paper examines the evolution of microbial electrolysis cell technology in terms of hydrogen yield, operational aspects that impact total hydrogen output ...
  49. [49]
    Recent advances in powering microbial electrolysis cells for ...
    Jul 11, 2025 · This review provides a comprehensive overview of the recent advancements in reducing and eliminating the electricity power demand of MECs for sustainable H 2 ...
  50. [50]
    Microbial electrolysis cells for the production of biohydrogen in dark ...
    Aug 10, 2025 · To assess biohydrogen for future green energy, this review revisited dark fermentation and microbial electrolysis cells (MECs).
  51. [51]
    Performance efficiency comparison of microbial electrolysis cells for ...
    Jan 5, 2022 · This review attempts to categories various MECs, used for sustainable biohydrogen production through organic matters or wastewater treatment.
  52. [52]
    Optimizing Hydrogen Production Through Efficient Organic Matter ...
    Microbial electrolysis cells (MECs) represent a pioneering technology for sustainable hydrogen production by leveraging bioelectrochemical processes.
  53. [53]
    A Review on Biohydrogen Production Through Dark Fermentation ...
    This study explores biohydrogen production through dark fermentation, an alternative supporting sustainable hydrogen generation.
  54. [54]
    Bio-hydrogen production by dark anaerobic fermentation of organic ...
    Using organic wastewater to produce hydrogen by fermentation can generate clean energy while treating wastewater. At present, there are many inhibitory ...
  55. [55]
    Biohydrogen production through dark fermentation of agricultural ...
    Dark fermentation (DF) is a biological hydrogen production method in which microorganisms break down complex organic matter into simpler compounds, producing ...
  56. [56]
    The key to unlocking enhanced dark fermentative processes
    Microorganisms play a pivotal role in hydrogen production efficiency. Strains from genus Clostridium have been the most widely detected and used microorganisms ...
  57. [57]
    Biological fermentation pilot-scale systems and evaluation for ...
    May 28, 2024 · Dark fermentation hydrogen production is the process in which bacteria convert organic matter into hydrogen and volatile fatty acids in the ...<|separator|>
  58. [58]
    Enhanced Biohydrogen Production through Dark Fermentation ... - NIH
    The dark fermentation process is conducted by anaerobic bacteria utilizing complex organic compounds to produce biohydrogen through the oxidative ...
  59. [59]
    Microbial electrolysis cells: Fuelling the future with biohydrogen
    The present review elucidates the mechanism, different configurations and substrates, and scaling-up potential for biohydrogen production via MEC.
  60. [60]
    Microbial electrolysis cells for hydrogen production - AIP Publishing
    Jun 1, 2020 · Microbial electrolysis cells (MECs) present an attractive route for energy-saving hydrogen (H 2 ) production along with treatment of various wastewaters.
  61. [61]
    Bio-hydrogen production through microbial electrolysis cell
    Jan 1, 2023 · This review discusses the principle, main components, and major operational parameters of MEC for significant performance.
  62. [62]
    Factors affecting hydrogen production in microbial electrolysis cell ...
    Apr 3, 2024 · This review examines the various operational parameters in optimization studies that influence overall hydrogen generation.
  63. [63]
    Microbial electrolysis cells designed for BioH 2 production
    Apr 10, 2024 · This study investigates a newly designed biohydrogen reactor by studying the effects of several variables on the production of bioH 2.
  64. [64]
  65. [65]
    Genetic Engineering of Cyanobacteria to Enhance Biohydrogen ...
    By combining several effective improvements through genetic engineering, high-H2-producing cyanobacterial strains suitable for large-scale production could be ...Missing: algae yield
  66. [66]
    The CRISPR technology: A promising strategy for improving dark ...
    This review highlighted the state-of-the-art in CRISPR-Cas systems for bacterial genome editing while paying attention to bioprocess optimization strategies.
  67. [67]
  68. [68]
    Improved bio-hydrogen production by overexpression of glucose-6 ...
    The results showed that the engineered strains produced 1.15 and 1.39-fold higher hydrogen yield, respectively, than the wild type. Furthermore, when pH and ...
  69. [69]
    Integrated strategy of CRISPR-Cas9 gene editing and small RNA ...
    We describe three strategies to improve hydrogen production by effectively regulating the anaerobic metabolism of E. aerogenes through genetic modification.1. Introduction · 2. Experimental Design · 3. Procedure
  70. [70]
    Metabolic engineering of Clostridium pasteurianum for enhanced ...
    Jul 24, 2025 · The hydrogen yield of dark fermentation is largely influenced by the microbial strain and its metabolic characteristics.
  71. [71]
    An overview of bioreactor configurations and operational strategies ...
    Bioreactors for biohydrogen production include continuous stirred tank, upflow anaerobic sludge blanket, packed bed, gas lift, membrane, fluidized bed, and ...
  72. [72]
    A comparative evaluation of dark fermentative bioreactor ...
    Jan 8, 2025 · The results revealed that continuously stirred reactors are widely utilized for bioH2 production as a cost-effective reactor configuration.
  73. [73]
    [PDF] Design of bioreactors for biohydrogen production
    This paper reviews reactor configurations (fixed-bed, fluidized-bed, upflow anaerobic sludge blanket and continuous stirred tank reactors) and operating ...
  74. [74]
    Biohydrogen production through dark fermentation: Recent trends ...
    Jan 2, 2024 · A dark fermentation process produces an impressive yield of bio-H2 using a simple, robust operation, unlike other non-fermentative hydrogen ...
  75. [75]
    A review on pretreatment methods, photobioreactor design and ...
    This review focuses on the novel and recent metabolic approaches for enhanced algal based biohydrogen production.
  76. [76]
    Scale-up and optimization of biohydrogen production reactor from ...
    As the reactor size increases, various problems such as breakages, wash out, and dead zones may become more severe owing to increasing inhomogeneity in the ...
  77. [77]
    Biological fermentation pilot-scale systems and evaluation for ... - NIH
    May 28, 2024 · Dark fermentation hydrogen production is the process in which bacteria convert organic matter into hydrogen and volatile fatty acids in the ...
  78. [78]
    Challenges and Mitigation Strategies Related to Biohydrogen ...
    Feb 1, 2023 · To achieve its implementation at an industrial scale, higher productivity is critical. Research on various bioreactor configurations and factors ...<|separator|>
  79. [79]
    Challenges and Mitigation Strategies Related to Biohydrogen ...
    Four challenges have been identified, namely, physical, biological, chemical, and economical, which enhancement strategies for improving biohydrogen ...
  80. [80]
    Innovative approaches to biohydrogen production from organic waste
    Hydrogen generation via biological processes encompasses microbial electrolysis, biophotolysis, and fermentation ... peer-reviewed literature to ensure ...
  81. [81]
    Renewable hydrogen production from biomass and wastes ...
    This article highlights the major research progress on biohydrogen production from renewable bioresources such as organic wastes, lignocellulosic biomass, algal ...
  82. [82]
    Quantitative Modelling of Biohydrogen Production from Indian ...
    Mar 29, 2025 · The operating conditions for dark fermentation of raw maize straw include a temperature of 36 °C, a biomass concentration of 20 g/L, and a pH of ...
  83. [83]
    Recent advances in dark fermentative hydrogen production from ...
    Dec 6, 2024 · This review aims to offer a comprehensive overview on the potential of vegetable waste as a feedstock for dark fermentative biohydrogen production.
  84. [84]
    [PDF] Organic waste to biohydrogen A critical review from technological ...
    Peer reviewed version. Link back to DTU Orbit. Citation (APA):. Tian, H., Li, J ... 2, the current widely developed and discussed technologies that produce ...<|separator|>
  85. [85]
    Optimized Biohydrogen Production from Sewage Sludge
    Oct 17, 2025 · This study systematically evaluates the efficiency of integrating various pretreatment methods, heat, ultrasonic, acidic, alkaline, and a novel ...
  86. [86]
    Bio-hydrogen production from food waste and sewage sludge in the ...
    At present, the dark fermentation of food wastes with sewage SL for the biogas production is attracting considerable attention as the present trend about waste ...Missing: integration | Show results with:integration
  87. [87]
    Closing the loop on biohydrogen production - ScienceDirect.com
    Sep 4, 2024 · This review aims to elucidate the various practices employed in post-fermentation broth treatment and management, shedding light on current methodologies.
  88. [88]
    A Comparative Analysis of Different Hydrogen Production Methods ...
    The article compares several hydrogen production processes in terms of scalability, cost-effectiveness, and technical improvements. It also investigates the ...
  89. [89]
    The Potential Role of Biohydrogen in Creating a Net-Zero World
    Currently, production costs for Bio-H2 range widely ($1.25–$10.51/kg-H2). The lower-range costs are comparable to the mean costs of gray hydrogen ($0.5–$2/kg-H2) ...Missing: per | Show results with:per
  90. [90]
    [PDF] Production - Hydrogen Program - Department of Energy
    May 7, 2024 · Achieved bio-H2 production cost reduction from > $58/kg-H2 to ~$12.4/kg H2 (TRL 2-4) o R&D priorities and cost reduction strategies:.Missing: per | Show results with:per
  91. [91]
    Hydrogenase: a bacterial enzyme activating molecular hydrogen
    Hydrogenase: a bacterial enzyme activating molecular hydrogen: The properties of the enzyme Available. Marjory Stephenson;. Marjory Stephenson.
  92. [92]
    Hydrogenase: a bacterial enzyme activating molecular hydrogen - NIH
    Hydrogenase: a bacterial enzyme activating molecular hydrogen. The properties of the enzyme. Marjory Stephenson.
  93. [93]
    [PDF] Biological Hydrogen Production: A Comprehensive Review for ...
    Sep 30, 2024 · However, as early as the 1930s, scientists observed that different bacteria released hydrogen under light and dark conditions [11-12]. The.
  94. [94]
    FERMENTATIVE AND PHOTOCHEMICAL PRODUCTION OF ...
    3. Algae which have been fermenting for several hours in the dark produce upon illumination free hydrogen at several times the rate observed in the dark, ...
  95. [95]
    Hydrogen metabolism of green algae: discovery and early research
    The detection of hydrogen metabolism in green algae more than 60 years ago by Hans Gaffron dispelled the widely accepted dogma at that time that this ...
  96. [96]
    Hydrogen metabolism of green algae: discovery and early research
    The detection of hydrogen metabolism in green algae more than 60 years ago by Hans Gaffron dispelled the widely accepted dogma at that time that this ...Missing: microbial | Show results with:microbial<|separator|>
  97. [97]
    Photosynthetic Hydrogen and Oxygen Production by Green Algae
    Aug 22, 1999 · ... 1970s that focused on photosynthetic hydrogen production from an applied perspective. From a scientific and technical point of view, current ...
  98. [98]
    Production of Hydrogen via Biological Processes - ResearchGate
    Aug 6, 2025 · When the oil crisis broke out in 1970s, the technology started receiving attention, especially in hydrogen production by photosynthetic process.
  99. [99]
    Photosynthetic Hydrogen and Oxygen Production by Green Algae
    Beginning with its discovery by Gaffron and Rubin in 1942, then motivated by curiosity-driven laboratory research, studies were initiated in the early 1970s ...
  100. [100]
    Photosynthetic Hydrogen and Oxygen Production: Kinetic Studies
    Steady-state turnover times for simultaneous photosynthetic production of hydrogen and oxygen have been measured for two systems.
  101. [101]
    Biohydrogen Production - Sources and Methods: A Review
    This article is a review of available methods of biohydrogen production. In the introduction, properties of different fuels and yields of hydrogen are ...Missing: definition | Show results with:definition
  102. [102]
    Photobiological production of hydrogen - ScienceDirect.com
    Among the topics included is a brief historical overview of hydrogen metabolism in photosynthetic bacteria, eucaryotic algae, and cyanobacteria (blue-green ...
  103. [103]
    Advancements of Biohydrogen Production Based on Anaerobic ...
    This review aims to comprehensively evaluate biohydrogen production via anaerobic digestion, addressing gaps in previous studies focusing on a single ...
  104. [104]
    Biohydrogen production from microalgae—Major bottlenecks and ...
    Nov 29, 2020 · This review provides comprehensive information on the mechanisms of hydrogen production by microalgae and the enzymes involved.<|separator|>
  105. [105]
    Advances and challenges in photosynthetic hydrogen production
    Sustainable hydrogen production remains challenging. The most elegant approach is to utilize photosynthesis for water splitting and to subsequently save solar ...
  106. [106]
    Challenges and opportunities for hydrogen production from ...
    Some of the greatest include market penetration given the dominance of fossil fuels, the high cost of replacing existing carbon‐based fuel infrastructure and ...
  107. [107]
    Photobiological hydrogen production: Bioenergetics and challenges ...
    We discuss the improvements in hydrogen production efficiency and the advances in related technologies that are needed before phototrophs can be used for ...
  108. [108]
    Dark Fermentative Biohydrogen Production: Recent Advances and ...
    Sep 11, 2023 · (43) The exact threshold above which the dark fermentation is inhibited is dependent on nature of feedstock and substrates, heavy metals, etc.
  109. [109]
    A critical review on limitations and enhancement strategies ...
    Biological hydrogen production is often constrained by less productivity. However, to obtain industrial level implementation, greater productivity is essential.
  110. [110]
    Photobiological Hydrogen Production at Scale: Integrating ... - bioRxiv
    Apr 30, 2025 · The improved H2 production rates reported in this study were achieved solely through engineering adjustments to the photobioreactor design.
  111. [111]
    Hydrogen Production: Photobiological | Department of Energy
    The photobiological hydrogen production process uses microorganisms and sunlight to turn water, and sometimes organic matter, into hydrogen.
  112. [112]
    Biohydrogen production beyond the Thauer limit by precision design ...
    Aug 14, 2020 · Until now, the H2 yield is restricted to 4 moles of H2 per mole of glucose, referred to as the “Thauer limit”. Here we show, that precision ...
  113. [113]
    Material and energy balances of an integrated biological hydrogen ...
    The net energy balance (NEB) and net energy ratio (NER) are −738.4 kWh and 77.8%, respectively, both of which imply an unfavorable energy production system.
  114. [114]
    Advances and bottlenecks in microbial hydrogen production - Stephen
    Aug 22, 2017 · This review highlights recent advances and bottlenecks in various approaches to biohydrogen processes, often in concert with management of ...
  115. [115]
    Environmental Sustainability of Biohydrogen Production from ...
    Sep 1, 2024 · With a 20% reduction of biohydrogen yield, the environmental impacts showed an increase from 25% to 26.1% in all impact categories. Conversely, ...
  116. [116]
    Sustainability considerations in bio-hydrogen from bio-algae with ...
    Jul 15, 2024 · This paper delves into several aspects concerning hydrogen synthesis in algae, encompassing microalgae anatomy and physiology, hydrogen synthesis via ...Bio-Algae Cultivation... · Bio-Hydrogen Production From... · Environmental Sustainability
  117. [117]
    Global land and water limits to electrolytic hydrogen production ...
    Sep 8, 2023 · Large-scale deployment of electrolytic hydrogen raises concerns about the availability of sufficient land and water resources for the ...
  118. [118]
    US finalizes 45V clean hydrogen subsidy scheme - C&EN
    Jan 8, 2025 · The 45V program, part of the Inflation Reduction Act of 2022, offers a tax credit for hydrogen if the total greenhouse gas (GHG) generated in its production is ...Missing: biohydrogen government funding<|separator|>
  119. [119]
    A Review of Techno-economics of Bio-hydrogen Production ...
    Biohydrogen is not competitive with natural gas based hydrogen in Western Canada. Carbon credits and other subsidies can improve the economics of hydrogen ...Missing: criticisms | Show results with:criticisms<|control11|><|separator|>
  120. [120]
    Biohydrogen Market Size, Share, Growth Report 2034
    Apr 26, 2024 · High production cost is a major factor limiting the adoption of biohydrogen. The majority of expenses involved in the production of biohydrogen ...
  121. [121]
    Carbon 'negativeness' of biohydrogen not rewarded by US policy ...
    Mar 15, 2023 · The biohydrogen production process produces a clean fuel with negative carbon emissions, but that carbon negative co-benefit is not incentivized ...Missing: criticisms | Show results with:criticisms
  122. [122]
    Bio and green hydrogen: A comparative review of regulatory trends ...
    Aug 1, 2025 · Currently, biohydrogen production by biomass pyrolysis is in the “Research, Commercialization, and Development” phase, operating at a medium ...
  123. [123]
    Interactions and distortions of different support policies for green ...
    Our analysis shows that mechanisms remunerating hydrogen production can distort spot prices of electricity and hydrogen more strongly than mechanisms that ...
  124. [124]
    How Good Politics Results in Bad Policy: The Case of Biofuel ...
    This paper argues that the growing list of concerns about the impact of biofuel targets and mandates are the predictable result of a failure to follow the ...
  125. [125]
  126. [126]
    Should Governments Subsidize Hydrogen?
    Jan 19, 2021 · First, hydrogen subsidies risk being environmentally counterproductive. Blue hydrogen—produced from fossil fuels but accompanied by carbon ...Missing: criticisms biohydrogen
  127. [127]
    DOE to Cancel Hydrogen Hub Grants, Leaked Docs Reveal
    Oct 8, 2025 · The US Department of Energy (DOE) is reportedly preparing to cancel all remaining grants under the federally funded hydrogen hubs programme, ...Missing: biohydrogen subsidies
  128. [128]
    Bio Hydrogen Market Size, Share and Analysis, 2025-2032
    May 26, 2025 · Bio Hydrogen Market size is growing with a CAGR of 7.2% in the prediction period & it crosses USD 127.8 Mn by 2032 from USD 78.6 Mn in 2025.