Next Article in Journal
Analysis and Research on the Influence of a Magnetic Field Concentrator on the Gear Heating Process Using a High-Frequency Resonant Inverter
Next Article in Special Issue
Performance Improvement of Proton Exchange Membrane Fuel Cells with a TiO2 Sputtered Gas Diffusion Layer Under Low-Humidity Conditions
Previous Article in Journal
The Application of Vibroacoustic Mean and Peak-to-Peak Estimates to Assess the Rapidly Changing Thermodynamic Process of Converting Energy Obtained from Various Fuel Compositions Using a CI Engine
Previous Article in Special Issue
Application of the Metalog Probability Distribution Family to Predict Energy Production by Photovoltaic Systems for the Purposes of Generating Green Hydrogen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Biohydrogen Production Through Dark Fermentation, Process Parameters and Simulation

by
Babak Mokhtarani
,
Jafar Zanganeh
* and
Behdad Moghtaderi
Centre for Innovative Energy Technologies, The University of Newcastle, Callaghan, NSW 2308, Australia
*
Author to whom correspondence should be addressed.
Energies 2025, 18(5), 1092; https://doi.org/10.3390/en18051092
Submission received: 14 January 2025 / Revised: 12 February 2025 / Accepted: 21 February 2025 / Published: 24 February 2025
(This article belongs to the Special Issue Sustainable Development of Fuel Cells and Hydrogen Technologies)

Abstract

:
This study explores biohydrogen production through dark fermentation, an alternative supporting sustainable hydrogen generation. Dark fermentation uses organic waste under anaerobic conditions to produce hydrogen in the absence of light. Key process parameters affecting hydrogen yield, including substrate type, microorganism selection, and fermentation conditions, were examined. Various substrates, such as organic wastes and carbohydrates, were tested, and the role of anaerobic and facultative anaerobic microorganisms in optimizing the process was analyzed. The research also focused on factors such as pH, temperature, and hydraulic retention time to enhance yields and scalability. Additionally, the study modelled the process using ASPEN Plus software 14. This simulation identifies the bottle necks of this process. Due to the lack of available data, modelling and simulation of the described processes in ASPEN Plus required certain approximations. The simulation provides insight into the key challenges that need to be addressed for hydrogen production. Future research should indeed explore current limitations, such as substrate efficiency, process scalability, and cost-effectiveness, as well as potential advancements like the genetic engineering of microbial strains and improved bioreactor designs.

1. Introduction

Burning fossil fuels has greatly increased greenhouse gases and polluted the environment. Since fossil fuels are non-renewable and will eventually be depleted, it is crucial to find alternative energy sources that are both renewable and environmentally friendly. Hydrogen is a promising alternative because it only produces water as a byproduct when combusted. Additionally, hydrogen can be produced through various methods, making it a renewable energy source. In recent years, hydrogen production has garnered increasing attention worldwide due to its potential role in advancing sustainable energy solutions. Hydrogen is poised to play a crucial role in the future of energy production, offering significant benefits over traditional fossil fuels. The energy released from hydrogen (122 kJ/g) is approximately 2.75 times higher than that produced by commonly used fossil fuels, highlighting its efficiency and potential as a cleaner energy source [1,2]. The literature review indicates that the majority of global hydrogen production is derived from fossil fuel combustion and steam reforming, with less than 5 percent coming from water electrolysis and less than 1 percent from biomass [3].
The production of biological hydrogen (bio-H2) is becoming more important due to the utilization of energy resources and its utility at ambient pressures and temperatures [3]. This process is less energy-intensive and more eco-friendly compared with conventional chemical methods. It can also use various agricultural waste products, which facilitates waste recycling [4].
Biohydrogen production can be achieved through four main technologies: photo fermentation, dark fermentation (DF), bio photolysis, and microbial electrolysis cells (MECs). Among these, DF stands out due to its higher hydrogen production rate and its adaptability to a wide range of organic substrates. This versatility not only reduces operational costs but also offers a sustainable method for revalorizing waste materials [5].
Photo fermentation converts organic acids into hydrogen and carbon dioxide through photosynthesis, facilitated by anaerobic bacterial strains. Hydrogen is produced when nitrogenase reduces molecular nitrogen, while also reducing protons to generate molecular hydrogen [6].
Bio photolysis is a process that harnesses light energy—whether from solar or artificial sources—to produce biohydrogen from water molecules through photosynthesis by photoautotrophic organisms. This process primarily occurs in green algae and cyanobacteria, with nitrogenase or hydrogenase being the key enzymes involved in biohydrogen production [7,8]. For optimal biohydrogen production via bio photolysis, the optimum light intensity, pH, and temperature should be applied, respectively [9].
MECs (microbial electrolysis cells) typically use two-chambered cells and operate on the principle of organic oxidation at the anode and reduction reactions at the cathode. This process, also known as electro-fermentation, involves bio-catalyzed electrolysis where microorganisms at the anode catalyzed the oxidation of an organic substrate, resulting in the generation of electrons, protons, and carbon dioxide [10].
Dark fermentation primarily utilizes anaerobic bacteria to produce biohydrogen. This process involves the use of anaerobic microorganisms in the absence of light, operating at temperatures ranging from mesophilic (25–40 °C) to thermophilic (50–65 °C) conditions, with some processes occurring at extreme thermophilic temperatures exceeding 80 °C [11]. It is one of the most extensively researched and promising technologies for biohydrogen production due to its higher production rates and its effectiveness in treating organic wastes. Various carbohydrate-rich substrates can be utilized in this process, including first-generation fuel crops like sugarcane, wheat, corn, and sugar beets, as well as second-generation biomass sources such as agricultural residues, industrial waste, and wastewater [12]. gasses such as CO2, carbon monoxide (CO), and hydrogen sulphide (H2S) are also released in this process.
Dark fermentation is a preferred biohydrogen production method due to its high hydrogen production rate, operational simplicity, and adaptability to various feedstocks, including organic waste. While other technologies such as photo fermentation, bio photolysis, and microbial electrolysis cells (MECs) offer advantages in hydrogen yield and efficiency under specific conditions, DF stands out for its practicality. Unlike photo fermentation and bio photolysis, DF produces a higher concentration of H2 and supports microorganisms with a shorter doubling time, enhancing overall process efficiency. DF is particularly scalable and cost-effective, requiring low energy inputs and allowing for waste valorization, making it an attractive option for sustainable hydrogen production. Although other technologies may achieve higher yields under controlled laboratory conditions, DF remains the most viable for real-world applications when considering energy efficiency, operational feasibility, and total lifecycle environmental impact. DF outperforms other biohydrogen production methods in waste-rich, low-energy-input, and high-rate production environments, making it ideal for industrial biorefineries, wastewater treatment plants, and decentralized energy systems. While photo fermentation and MECs have advantages in specific setups, DF is the most practical and scalable solution for large-scale, sustainable biohydrogen production [1,13,14].
Lignocellulosic materials and agricultural wastes are promising feedstocks for biohydrogen production. However, these feedstocks require the addition of enzymes to undergo fermentation [15]. For large-scale profitability, DF requires low-cost substrates, as sugar-based feedstocks might be more economically viable for other uses. Consequently, researchers have explored various organic residues as alternative feedstocks. Among these, agricultural, industrial, and food wastes are the most extensively studied, likely due to their carbohydrate-rich compositions [16,17]. Biohydrogen production through dark fermentation (DF) is currently one of the most established biological methods, demonstrating practical readiness and potential for scale-up [18].
Although fermentation technologies have improved in recent years, biohydrogen production through fermentation remains limited to the lab scale [19]. In recent years, research activities in dark fermentation (DF) have increased, as reflected by the growing number of reviewed articles. Ghimire et al. [14] reviewed improvements in hydrogen yield and production rates achieved through various seed inoculum enrichment methods, bioreactor design modifications, and optimization of operational conditions within DF bioreactors. They describe how hydrogen yield can be improved by optimizing the design and operation of DF bioreactors [20]. The cost of hydrogen production can be reduced by using low-cost renewable feeds, such as waste biomass, as feedstocks [21,22]. Additionally, inoculum enrichment methods [23,24,25,26,27] enhance hydrogen yield, while the pretreatment of feed can further boost productivity [28].
Lukajtis et al. [6] reviewed research on bacterial strains used for hydrogen production from agro-industrial waste and lignocellulosic biomass. They investigated the effects of various parameters, including the type of raw material, processing methods, temperature, pH, substrate concentration, hydrogen partial pressure, hydraulic retention time, inoculum preparation, and the type and operating conditions of the reactor, on the yield of DF.
Ahmad et al. conducted a literature review on biohydrogen production through dark fermentation. They examined the metabolic pathways involved in biohydrogen production and identified inhibitory substances that negatively impact dark fermentation. Additionally, they provided a brief review of pilot-scale biohydrogen production using various substrates through dark fermentation [13].
The mixed cultures composed of diverse bacteria are efficient for biohydrogen production and were compared with pure bacteria by some researchers [29,30,31,32,33]. They found that mixed cultures are more appropriate to overcome operational difficulties such as a metabolic shift in response to environmental stress, and the need for a sterile environment. Mohanakrishna and Pengadeth [34] reviewed the use of mixed-culture bacteria for biohydrogen production and found that mixed cultures are more suitable than pure cultures, particularly for scaling up DF.
According to the literature, most DF studies have been conducted at the lab scale, with no reports found on industrial or full-scale implementations. Only a limited number of studies have explored the pilot-scale application of DF processes [35,36,37]. Process modelling and simulation are appropriate tools for the identification of technical and fundamental challenges that exist in biohydrogen production. No study was found in the literature on process simulation for biohydrogen production. Furthermore, there is a lack of literature on the process modelling of biohydrogen production using the DF process.
The aim of this research is to develop a comprehensive process simulation for biohydrogen production through fermentation. Initially, the DF process is thoroughly explained, with a focus on identifying key process parameters crucial to optimizing hydrogen production. This includes an in-depth analysis of the type of microorganisms involved and the specific fermentation parameters that affect their activity. Based on these insights, a detailed process design for hydrogen production is created, followed by the modelling and simulation of the system to evaluate its performance. The goal is to use these simulations to better understand the process dynamics and optimize conditions for a maximum hydrogen yield. The process simulation is performed based on simple assumptions to highlight the bottlenecks of this process.

2. Dark Fermentation Process

DF uses anaerobic bacteria to convert organic compounds into hydrogen without the need for light. This allows DF to produce hydrogen continuously, independent of light availability, which contributes to its higher biohydrogen production rate and hydrogen content compared to other biohydrogen methods. Additionally, DF is a simple process with stable operation, requiring less investment, thus making it an attractive option for biohydrogen production [38]. It has the potential to replace of fossil fuel for biohydrogen production as a renewable energy source.
Figure 1 shows the pathways for biological hydrogen production through DF. In this process, the biomass is broken down into sugar molecules, such as glucose. Hydrogen-producing microorganism then convert glucose into pyruvate via glycolytic pathways, generating adenosine triphosphate (ATP) from adenosine diphosphate (ADP) and reduced nicotinamide adenine dinucleotide (NADH). The subsequent conversion of pyruvate depends on the type of microorganism involved. Facultative anaerobes typically use the pyruvate formate lyase (PFL) pathway, while strict anaerobes generally employ the pyruvate ferredoxin oxidoreductase (PFOR) pathway [39]. Acetyl coenzyme A (acetyl-CoA) is produced in both pathways and can be further converted into various compounds, including acetate, butyrate, butanol, and ethanol, during the DF of complex carbohydrates by mixed anaerobic microbiota. The resulting intermediates and byproducts can vary widely depending on operational parameters such as substrate type, substrate loading rate, pH, temperature, and other environmental and operating conditions. These factors also significantly influence the microbial community structure within bioreactors [14].
The differences in metabolic pathways between facultative and strict anaerobes significantly impact hydrogen production efficiency and yield. Facultative anaerobes, capable of switching between aerobic and anaerobic metabolism, often consume oxygen rapidly, creating favourable anaerobic conditions for strict anaerobes that specialize in fermentative hydrogen production. Strict anaerobes, such as Clostridium spp., primarily rely on fermentative pathways, leading to higher hydrogen yields under optimal conditions. However, their efficiency is influenced by factors such as substrate utilization patterns and byproduct formation [40,41].
Variations in operational parameters directly affect the metabolic distribution of products in dark fermentation (DF) processes. Substrate type dictates the dominant metabolic pathways; for instance, carbohydrate-rich substrates typically favour acetate and butyrate pathways, promoting higher hydrogen yields, whereas protein- or lipid-rich substrates may lead to increased solvent production (e.g., ethanol and butanol), potentially reducing hydrogen output. Organic loading rate (OLR) influences microbial community dynamics; excessively high OLR can lead to substrate inhibition and the increased accumulation of reduced end-products like lactate and ethanol, which are less favourable for hydrogen production. pH plays a crucial role in regulating metabolic shifts; acidic conditions (<5.5) tend to favour solventogenic pathways (ethanol, butanol), whereas neutral to slightly acidic conditions (5.5–6.5) enhance acetate and butyrate production, optimizing hydrogen yields [39,41,42,43].
A stable DF process requires a balanced microbial consortium, where changes in substrate composition, OLR, and pH can cause shifts in community structure, favouring either hydrogen-producing bacteria or competing microbial groups (e.g., methanogens, lactic acid bacteria). Adaptation mechanisms such as selective pressure, microbial succession, and syntrophic interactions play key roles in maintaining system performance. Operational instability, such as pH fluctuations or substrate shocks, can lead to process failure due to microbial washout or the dominance of less efficient pathways. Thus, optimizing these parameters is essential for enhancing process efficiency and scalability in hydrogen production bioreactors [41,44].
The biochemical reactions involved in DF by facultative anaerobes can be illustrated by reactions (1) and (2), as shown.
C 6 H 12 O 6 + 2 H 2 O 2 C H 3 C O O H + 2 C O 2 + 4 H 2
C 6 H 12 O 6 C H 3 ( C H 2 ) 2 C O O H + 2 C O 2 + 2 H 2
Acetate and butyrate are the primary products of this process. As these reactions demonstrate, glucose can be converted into 4 moles of hydrogen per mole of substrate when acetic acid is produced as a co-product (Equations (1) and (2)) moles of hydrogen per mole of substrate when butyric acid is the co-product (Equation (2)). Therefore, the acetate pathway is preferred because it yields more hydrogen than the butyrate pathway [45]. However, the actual hydrogen yield is lower than the theoretical yield because a portion of the substrate is used for biomass production, and the degradation of substrates can follow alternative biochemical pathways that do not produce hydrogen [46,47]. Under certain conditions, metabolic pathways shift toward ethanol and acetate production, reducing the stoichiometric hydrogen yield to 2 moles of hydrogen per mole of glucose [23].
C 6 H 12 O 6 + 2 H 2 O C H 3 C O O H + C 2 H 5 O H + 2 C O 2 + 2 H 2
DF primarily utilizes anaerobic bacteria to ferment organic compounds into hydrogen without the need for light. As a result, DF can continuously generate hydrogen without relying on light availability. This advantage contributes to its high biohydrogen production rate and hydrogen content [38].

3. Process Parameters

To design an optimal process for hydrogen production via DF, it is essential to identify the key parameters that influence the process. Hydrogen production under DF is complex, with several factors affecting its efficiency. The type of substrate and microorganism are the primary factors influencing hydrogen production. Additionally, operating parameters for the bioreactor, such as temperature, pH, and hydrogen partial pressure, play crucial roles [14]. These parameters are reviewed below.

3.1. Types of Substrate

Selecting a suitable substrate has a significant effect on any fermentation process. In the DF process, raw materials containing organic compounds serve as substrates for hydrogen production. Monosaccharides and disaccharides such as glucose or lactose are the desired carbon sources for the conversion of microorganisms in DF. Renewable sugar sources, including starch, cellulose, and hemicellulose, are primarily found in plants as polymers. Using raw materials rich in starch, which can be easily hydrolyzed into simple carbohydrates, offers particular advantages [48].
Sugars like glucose, xylose, and sucrose are biodegradable substrates that have been applied for DF since the 1980s [49]. These substrates do not require complicated processing and produce the highest yields [50]. However, they are not suitable for large-scale applications because their costs are too high [6].
Renewable resources like lignocellulose or starch waste can be used for continuous large-scale hydrogen production. These materials, known as second-generation substrates, require pretreatment before use in the DF process due to their initially low hydrogen yield. To enhance hydrogen production, these substrates should be processed to change long polymer chains into simple or short oligomers [6].
The organic fraction of readily biodegradable municipal waste is suitable for the DF process. These materials are primarily from households, restaurants, and food processing plant residues, which contain materials rich in polysaccharides, proteins, lipids, and simple carbohydrates. Simple sugars like glucose and sucrose offer high bioavailability and efficient hydrogen production, but their cost can be prohibitive for large-scale applications. Conversely, complex substrates, such as lignocellulosic biomass, municipal waste, and industrial effluents, are more affordable but require extensive pretreatment to break down polymers into fermentable monomers [51]. For managing economic and efficiency Trade-offs, the pretreatment optimization of substrates is performed. Cost-effective pretreatment methods (e.g., enzymatic hydrolysis, mild acid hydrolysis, or anaerobic digestion) can enhance hydrolysis efficiency while minimizing operational costs and inhibitory byproducts. Furthermore, the co-digestion of mixed substrates—blending simple and complex substrates (e.g., glucose with agricultural residues)—balances bioavailability and cost, improving microbial adaptability and hydrogen yield. The utilization of industrial byproducts and readily available industrial effluents (e.g., dairy wastewater, food processing waste) as feedstock reduces raw material costs while providing valuable organic content [52,53]. The optimal substrate concentration can be determined by conducting laboratory-scale batch fermentations. Applying kinetic models (Monod or substrate-inhibition models) enables predictive control over substrate loading rates to prevent inhibitory effects. Furthermore, using online sensors for volatile fatty acids (VFAs), pH, and hydrogen production rates ensures real-time adjustments in feed rates to maintain optimal fermentation conditions [54].
Sewage sludge contains a high concentration of simple carbohydrates, disaccharides, and peptides. To generate hydrogen from sewage sludge, it is essential to eliminate bacteria that consume hydrogen during their metabolic processes, particularly methanogenic bacteria. This can be achieved through pretreatment methods such as heat shock, ultrasonication, acidic or alkaline treatment, or by supplementing with organic materials [55,56].
The composition of these wastes fluctuates and pretreatment methods such as filtration, dilution, or nutrient supplementation can help to standardize feedstock composition. Enriching microbial communities with diverse metabolic capabilities enhances resilience to substrate variability. Furthermore, implementing feedback loops for pH, OLR, and VFA concentrations enables real-time process optimization to counteract variations in feedstock composition [57,58].
Table 1 provides a brief classification of suitable substrates for the DF process, highlighting their advantages and disadvantages. According to the table, renewable resource substrates rich in lignocellulose or starch, such as organic municipal waste, manure, agricultural waste, industrial waste, and effluent, are identified as suitable sources. However, these renewable sources need pretreatment and hydrolyzation into simple molecules.
The substrate concentration is a critical factor in the DF process. High concentrations can lead to reduced hydrogen production, as is common in many biological processes. Optimal hydrogen yields are generally achieved with dilute substrate concentrations, highlighting the importance of diluting the substrate to an ideal range [48]. Furthermore, substrate recycling and wastewater treatment are vital components of the DF process and should consider for the scale-up process.

3.2. Microorganism Type

In designing an efficient biohydrogen production process, selecting an appropriate microorganism is the most important step. Various types of microorganisms capable of producing hydrogen are found in environments, such as soil, wastewater, sludge, and compost. These materials can therefore serve as potential sources of inoculum for fermentative hydrogen production. Hydrogen-producing microorganisms exist in environments either as single species or as mixtures of various species. The choice of strain largely depends on the type of substrate used. Bacteria involved in DF process are anaerobic and can be classified into two groups. The first group is based on oxygen sensitivity, divided them into strict anaerobes and facultative anaerobes. The second group is determined by the temperature ranges at which bacteria exhibit the highest growth rate and activity [6].
Strict anaerobic microorganisms require completely oxygen-free conditions. These include anaerobes such as Clostridia, Methylotrophs, Methanogenic bacteria, Archaea, and Rumen bacteria [69]. The most used obligate anaerobes belong to the genus Clostridium, producing hydrogen in the logarithmic growth phase. As these microorganisms enter the stationary phase, their metabolism shifts toward producing liquid organic compounds, particularly volatile fatty acids [22]. Facultative anaerobes are organisms that primarily generate ATP through aerobic respiration when oxygen is available, but can switch to fermentation when oxygen is absent. Their ability to tolerate oxygen makes them easier to use in dark fermentation processes. Furthermore, a high oxygen partial pressure in the reactor does not negatively affect the fermentation yield. This group includes Enterobacteriaceae, a large family of Gram-negative, non-spore-forming bacteria [6]. The metabolic pathway of these microorganisms involves glucose metabolism, leading to hydrogen production. The end-products of this process include butyric, lactic, and formic acids, along with small amounts of ethanol.
Mixed or pure cultures of these strains can be applied for hydrogen production. Mixed cultures are more practical than pure cultures because the production of hydrogen is simpler to operate and control [3]. However, some undesirable microorganisms that might be present in mixed cultures reduce the total hydrogen yield, either by consuming the hydrogen produced or by altering the biochemical pathways of the hydrogen [23]. Mixed bacterial cultures that can produce hydrogen abundantly are commonly found in municipal sewage, composts, and organic waste, from which they can be isolated.
The metabolic pathways for pure cultures are understood well, but they require a sterile environment and are sensitive to changes in operational parameters. In contrast, mixed cultures can operate under non-sterile conditions and are easier to manage on a large scale. However, due to the presence of hydrogen-consuming strains, pretreatment is necessary [34]. Pretreatment is necessary because mixed cultures often contain a variety of microbial species, some of which may compete with or inhibit the desired hydrogen production process. The pretreatment method is summarized in Table 2. The choice of pretreatment method depends on the specific characteristics of the mixed culture, the types of microorganisms present, and the operational conditions of the fermentation process.

3.3. Fermentation Process Parameters

Anaerobic microorganisms see major application during hydrogen production in DF. The growth of microorganisms is highly dependent on the condition of the fermentation process. Temperature has a significant effect on the growth rate of microorganisms. The optimal temperature varies depending on the types of microorganism: mesophiles (25–40 °C), thermophiles (45–65 °C), extreme thermophiles (65–80 °C), and hyperthermophiles (above 80 °C) [80]. Literature studies have shown that the temperature can affect the metabolic pathways and alter the composition of the byproducts of DF [14]. Valdez-Vazquez et al. [81] reported that thermophilic fermentation yields higher H2 production compared to mesophilic conditions. Additionally, acetic acid was the dominant byproduct in thermophilic digestion, while butyrate was more prevalent in mesophilic digestion. In the DF of food waste, acetate was the primary end-product in mesophilic cultures, whereas thermophilic cultures produced higher levels of butyrate and hydrogen [82]. DF is limited at high temperatures due to economic considerations [83]. While thermophilic conditions can be cost-effective, offering higher hydrogen yields and the ability to process more complex substrates, extreme thermophilic and hyper thermophilic conditions require substantial energy inputs, which diminishes the profitability of the process [84].
Operational pH determines the optimal metabolic pathways for hydrogen production and helps to inhibit hydrogen-consuming processes that may occur simultaneously [85]. The pH value has a large effect on the activity of various microorganism in a mixed culture. It changes the metabolic pathways of microorganisms, as well as their morphology and cell structure [86]. All enzymes involved in bacterial metabolic processes are active within a specific pH range, with peak activity occurring at an optimal pH. Maintaining and controlling the pH at this optimal level is crucial during fermentation. This is because hydrogen production is complemented by the formation of acids such as acetic, lactic, butyric, and propionic acid. These acids lower the medium’s pH, inhibiting the enzymes responsible for hydrogen production [76]. A low pH (below 5) also impairs the bacteria’s ability to maintain appropriate intracellular pH levels [87]. According to the literature, the optimal initial and operational pH values depend on the specific bacterial strain or source of mixed cultures. Additionally, some studies suggest that the type of substrate in the medium influences the optimal pH [4]. Generally, the optimal pH for fermentative hydrogen production ranges from 5.0 to 7.0, which aligns with the pH range favourable for bacterial growth [23]. pH has a significant impact on hydrogen production yield and the concentration of other metabolic byproducts [88,89]. Studies show that at low pH values, acetic and butyric acids are the primary byproducts, while increasing the pH leads to higher concentrations of ethanol, propionic acid, and lactic acid [90]. The election of the operational pH is dependent on the substrate type and organic loading rate. The optimum pH for organic food waste varies from 4.5 to 7 and for lignocellulosic waste it varies from 6.5 to 7, whereas a neutral pH is optimal for animal manure [91].
The hydraulic retention time (HRT) influences hydrogen yield in DF processes. It measures the average time a substrate stays in the fermentation chamber for. HRT significantly impacts hydrogen productivity in continuous or semi-continuous DF operations. The rate of hydrogen production increases within a certain HRT range, but exceeding the optimal HRT value results in a decline in production [92]. The value of HRT is highly dependent on the substrate type. HRT can influence substrate hydrolysis, impacting the production of intermediates and products, thereby affecting fermentative H2 production. In addition to hydrolysis, HRT can also serve as a control parameter for methanogenic activity [14]. Several studies were conducted to determine the optimal HRT for various substrate types. They reported that HRT values vary within the range of 2 to 24 h [92,93,94]. During continuous culture growth, HRT is typically reduced gradually from longer to shorter intervals. This allows microorganisms to adapt to new conditions and prevents the target bacteria from being washed out. As HRT shifts, the microbial population undergoes dynamic changes, causing some species to disappear while others emerge [95].
The partial pressure of hydrogen is another factor influencing hydrogen production. During hydrogen production, the enzyme hydrogenase catalyzes the oxidation and reduction of ferredoxin in a reversible process. From a thermodynamic perspective, a rise in hydrogen partial pressure can promote ferredoxin reduction, which inhibits further substrate conversion into hydrogen. As hydrogen partial pressure increases, hydrogen production decreases, while the concentrations of other metabolic byproducts, such as lactic acid, ethanol, acetone, and butanol, increase [14,48]. Lowering the partial pressure in the reactor’s headspace enhances the mass transfer of hydrogen from the liquid to the gas phase [96]. Therefore, it is crucial to remove hydrogen as it forms to maintain a high and consistent production rate. Currently, several methods exist to reduce the pressure of hydrogen in a reactor, and thereby the dissolved hydrogen concentration in the liquid medium. Stirring the media is the easiest method for the removal of the gas phase from the fermenter [6].
Hydrogen yields in DF depend on the mode of fermentation, which can be batch, semi-continuous, or continuous. Batch reactors are mostly used in laboratory settings for preliminary research to determine the optimal parameters of dark fermentation [48]. These reactors have a straightforward and cost-effective design, allowing for the convenient control of fermentation parameters, such as temperature and pH. However, industrial-scale hydrogen production requires the use of continuous reactors, as they offer higher process efficiency. Continuous reactors are usually initiated in batch mode to allow for proper inoculum preparation and pretreatment. The success of the transition to a continuous mode depends significantly on the start-up strategy [95]. The most popular bioreactor configurations include continuous stirred tank reactors (CSTRs), upflow anaerobic sludge blanket reactors (UASBs), anaerobic fluidized bed reactors (AFBR), and membrane bioreactors (MBR) [6]. The advantages and disadvantages of various bioreactor configurations are presented in Table 3.
Recently, the Dark Fermentation Moving-Bed Biofilm Reactor (DF-MBBR) process was proposed for the first time as a coupled system integrating the DF-MBBR treatment of organic residues with the bioleaching of concrete debris. This innovative approach enables the simultaneous production of biohydrogen, aggregates, and calcium-based compounds for reuse in the construction industry while facilitating the disposal of both biodegradable and construction and demolition waste. A key advantage of this configuration is the elimination of physical or mechanical treatment for biomass separation and recirculation, making it highly compatible with bioleaching processes [33].
CSTRs are normally used for continuous fermentation. These reactors have a simple design and allow for the easy modification of operating conditions. Stirring ensures there is a homogeneous environment within the media and promotes effective contact between microorganisms and substrates [97,98]. The MBR is a commonly used reactor for fermentation processes; however, it includes a membrane or membrane system. Membranes can be positioned externally (side-stream MBRs) or immersed within the reactor. The immersed design offers advantages, including lower operating costs and more compact membranes. By retaining microorganisms inside the reactor, the membrane helps maintain a consistently high biomass concentration [23].
The membrane enables the selection of an optimal HRT, independent of the activated sludge retention time, allowing for better control of process parameters [59].
Hydrogen production rates were compared between UASBs and CSTRs. The results indicate that the UASB achieved higher production rates than the CSTR at low retention times, with 19.05 mmol H2/h/L versus 8.42 mmol H2/h/L at a 2 h HRT. However, the CSTR consistently produced higher hydrogen yields (mmol H2/mol glucose) across all tested HRTs. These findings suggest that a balance must be struck between technical efficiency (based on hydrogen yield) and economic efficiency (based on production rate) when selecting between these two systems [99].
Table 3. The advantages and disadvantages of different bioreactor configurations.
Table 3. The advantages and disadvantages of different bioreactor configurations.
BioreactorAdvantageDisadvantageReference
CSTR
  • Simple design.
  • Easy adjustment of operating conditions.
  • Homogeneous conditions with stirring.
  • Precise control of pH and temperature.
  • Limited biomass concentration.
  • Biomass removal and inactivation of microorganisms.
[23,48,97,98]
MBR
  • Lower operating costs.
  • Small membranes.
  • The membrane retains microorganisms inside the reactor.
  • Optimal HRT.
  • Control of process parameters.
  • Membrane fouling.
  • High operating cost.
[23,59]
UASB
  • High effectiveness of hydrogen.
  • Production.
  • Short HRT.
  • Stable operating conditions.
  • Good substrate conversion.
  • Sensitivity to channelling effects.
  • Long.
  • Initiation time.
[100,101]
AFBR
  • Good.
  • Mass and heat transfer.
  • Uniform particle mixing.
  • Can be operated continuously.
  • High energy consumption.
  • Difficulties in scale-up.
[102,103]

4. Modelling and Simulation

Most studies on hydrogen production using the DF process have been conducted at the laboratory scale. To the best of our knowledge, there are no studies in the literature regarding the industrial-scale application of this process. Only a few studies have explored pilot-scale applications of DF processes [35,36,37,104,105]. Process modelling and simulation methods facilitated the scale-up of this process, as laboratory-scale experiments alone were insufficient to fully understand the underlying science of the fermentation process. Therefore, there is a need to develop suitable models to comprehend the complex operation of biohydrogen production. In this regard, the process modelling of biohydrogen can help evaluate the design, scale-up, optimization, and control of the system. To achieve these objectives, a reliable process model is essential.
Process modelling allows us to create detailed representations of the fermentation process, which can be used to simulate different operating conditions and configurations. This approach helps to identify optimal parameters for maximize hydrogen yield and production rates. Furthermore, modelling and simulation enable a deeper understanding of the interactions between different process variables, providing insights into the underlying mechanisms of biohydrogen production.
When it comes to scaling up from laboratory or pilot-scale systems to industrial-scale operations, modelling and simulation become even more valuable. They provide a cost-effective way to predict the performance of large-scale reactors, minimizing the risks associated with scaling up. By using reliable models, it is possible to evaluate the effects of scale-up on process stability, efficiency, and product yield without the need for extensive experimental trials. Additionally, these models can assist in designing control strategies and optimizing reactor configurations, further enhancing the economic viability and technical efficiency of biohydrogen production at an industrial scale.

4.1. Process Description

A general block diagram for the hydrogen production is designed and depicted in Figure 2.
Biomass is mixed with water in a mixing tank. Straw, a common agricultural residue, is considered a valuable biomass. Wheat, maize, and rice straw are among the largest agricultural residues worldwide. Consequently, utilizing this abundant biomass as a feedstock for dark fermentative hydrogen production could have significant potential for future feedstock supply [14].
Straw is a lignocellulosic residue composed of cellulose, hemicellulose, and lignin. Due to its complex structure, the bioconversion of straw is challenging, making pretreatment essential for effective utilization [106]. Acid and alkali pretreatment methods are commonly used to degrade straw. These methods alter the cellulose structure and enhance sugar yield in a relatively short time [107]. Acid or base treatment takes place in the mixing tank, after which the slurry is pumped into the hydrolyser reactor. In this reactor, hydrolytic enzymes are added to convert carbohydrates into monomeric sugar molecules through a batch process.
The effluent from the hydrolysis reactor passes through a filter to separate solids from the liquid stream. The resulting sugar solution is then fed into a fermenter, where microorganisms convert the sugars into hydrogen according to their metabolic pathways. The produced gases are separated from the liquid phase and sent to a gas separation unit for hydrogen purification. A gas separation unit is used to separate hydrogen from carbon dioxide and water. This unit can be designed as a membrane or an adsorption column, utilizing a suitable adsorbent. The liquid waste from the fermenter is directed to the wastewater treatment section, where the water is treated and recycled back into the process.

4.2. Process Simulation

ASPEN Plus V14.0 is used for simulation of hydrogen production. The software has a large data bank that enables us to calculate the materials’ properties. The model was employed to simultaneously solve the steady-state governing equations of mass and energy for different compositions of solid streams utilizing the integrated “SOLIDS” property package. A steady state model is developed for this process. gasses are modelled as conventional components, and the biomass and ash are modelled as non-conventional components. Also, the “Mixed”, and “NCPSD” packages were chosen as the sub-stream class in the simulations. The biomass, as a non-conventional component, is modelled in two stages. In the first stage, biomass is decomposed into its ultimate and proximate component in a RYield reactor and, in the second stage, the component is converted into sugar molecules in a Rstoich reactor. The biomass section was simulated and validated with the literature results. Due to the lack of experimental data for the DF process, the model’s reliability has only been partially confirmed. The models were examined to ensure that the calculations were conducted without any ‘error’ or ‘warning’ messages, indicating that the energy and mass balances were within acceptable limits. The benefit of this method is that it allows for a preliminary validation of the model, even in the absence of experimental data for the DF. By ensuring that the calculations are free of ‘error’ or ‘warning’ messages, the model demonstrates that the energy and mass balances are within acceptable limits. This suggests the model is functioning correctly in terms of basic conservation laws, which provides a foundation for its reliability, even though its full accuracy has not yet been confirmed through direct experimental comparison.
Figure 3 represent the process simulation model, which comprises four units: hydrolyzation and filtration, the fermenter, water treatment and recycling, and gas treatment and recycling. In this figure, straw is crushed and mixed with water in a mixer (CH-101) and then fed into the biomass reactors (BIOMA-R1 and R2), where it is converted into glucose. The resulting mixture of glucose, water, and unconverted straw is then separated using a solid–liquid separation unit (SP-101). DF focuses more on substrate breakdown and gas production in anaerobic conditions. The model simulates different operational parameters (temperature, pH, substrate concentration) and their influence on hydrogen production rates, allowing for the optimization of process conditions.
Straw is considered as a biomass and the proximate and ultimate analyses are reported in Table 4 [108]. After the hydrolysis of straw (BIOMA-R), the resulting glucose solution is fed into the fermenter, where it is converted into hydrogen via reaction (1) (BC-101). The acetate pathway is chosen for this step because it yields a higher amount of hydrogen, producing 4 moles of hydrogen per mole of glucose. Phosphate-buffered solution (PBS) is added to the fermenter to prevent pH fluctuation.
Due to the high complexity of metabolic pathways, formal kinetic studies are rarely reported in the literature, although some kinetic data are available for specific microorganisms. For simplicity, in this research, hydrogen production is modelled as first-order kinetics [109].
R H 2 = k · S
where RH2 is the rate of hydrogen production (mol/L·h), k is the first-order rate constant (h−1), and S is substrate concentration (glucose, mole/L). The rate constant is expressed via the Arrhenius equation, as follows:
k = k 0 e E R T
where k0 is the Arrhenius constant (pre-exponential factor), E is the activation energy, R is the universal gas constant (8.314 J/mol·K), and T is the absolute temperature. The activation energy and Arrhenius constant of the DF process for glucose were extracted from the literature [110,111] and set to 67.3 kJ/mol, the value which was derived for E. aerogenes (NCIMB 10102) [110].
It is worth noting that hydrogen production increases up to 40 °C, after which it decreases linearly. The activation energy calculated for this range, associated with thermal inactivation, is approximately 118 kJ/mol—which is significantly lower than the values reported for microbial death, which range between 290 and 380 kJ/mol [112].
The output from the fermenter is fed into a separator where the gas and liquid are separated (S-101). The gas stream passes through a condenser unit to reduce its water content (C-102), and then it is directed to a gas separation unit to separate carbon dioxide from hydrogen (GS-01). The resulting hydrogen is then sent to a compressor to be pressurized (CO-102). Meanwhile, the liquid stream from the first separator is directed to another separator to separate acetic acid from water and other products (GS-02). The remaining liquid is sent to a water treatment and recycling unit, where water is separated from the residual contents of the fermenter and reused in the process.

4.3. Simulation Results

The simulation was conducted based on a hydrogen production rate of 5 tons per day. Due to the lack of available data, the modelling and simulation of the described processes in ASPEN Plus required certain approximations. It was assumed that the straw was directly converted into glucose molecules, with the pretreatment process being ignored. Straw was used as the biomass source and mixed with water to form a slurry, which was stored in a holding tank and treated with either acid or base. The slurry is then pumped into an acid hydrolysis reactor, where enzymes break down carbohydrates into monomeric sugar molecules, primarily glucose. The concentration of glucose was set to fall in the range reported by the literature for glucose (1–10 g glucose/L) [4]. The solids are removed using a filter press, and the liquid phase is sent to a makeup tank where nutrients and a phosphate-buffered solution are added to regulate pH. The nitrogen source plays a crucial role in the fermentation process and is provided by the addition of ammonium chloride.
The prepared feed is then pumped into a fermenter, where the glucose solution is converted into hydrogen and acetic acid under a first-order reaction until all glucose is consumed. The temperature of the fermentation process is set at 25 °C. During fermentation, the gas is separated from the liquid. The gas phase, containing hydrogen, carbon dioxide, and water vapor, is sent to a condenser to reduce humidity. The gas mixture is then processed through an adsorption or membrane separator to remove carbon dioxide from the hydrogen. It was assumed that 98% of carbon dioxide was separated from hydrogen. Furthermore, the hydrogen product passes through a compressor to increase its pressure to 4 bar.
The liquid phase from the fermenter, containing acetic acid, is directed to a separation column, where the acetic acid is separated. The remaining solution is sent to a wastewater treatment system, where water is separated from other materials and recycled back into the process. The simulation results are presented in Table 5. In this Table, the stream number are indicated based on the numbers in Figure 2.
The simulation results indicate that 10 tons per hour of straw can produce approximately 5694 kg per hour of glucose. This amount of glucose is reasonable and aligns with the typical composition of straw. As previously mentioned, straw is composed of cellulose, hemicellulose, and lignin, which typically constitute 40–60%, 20–40%, and 10–25% of straw, respectively. Cellulose is a linear polymer made up of repeating glucose units, and hemicellulose also contains glucose [113]. Therefore, the amount of glucose obtained after treatment is consistent with these compositional ranges. The hydrogen production amount is 212 kg/h and 3412 kg/h acetic acid as a subsidiary product. The results indicated that for one mole of glucose, 3.5 moles of hydrogen were produced, which aligns with the experimental data [19].
Since acetate and butyrate are the primary products in DF, the second scenario included both the butyrate and acetate pathways (Equations (1) and (2)), and the simulation was conducted accordingly. The simulation results are presented in Table 6.
The results indicate a significant shift in product distribution during the process. Hydrogen production decreases to 169 kg/h, showing a reduction compared to previous levels. Similarly, acetic acid output drops to 1792 kg/h. However, an increase in butyric acid production is observed, reaching 1190 kg/h. This suggest that the metabolic pathway favours the production of butyric acid. The stoichiometric yield of hydrogen is 2 moles per mole of glucose, as outlined in Equation (2), when butyric acid is the final product. However, the actual hydrogen yield is lower than this theoretical value. This discrepancy arises because a portion of the substrate is diverted for biomass production, and some of the substrate may follow alternative biochemical pathways that do not generate hydrogen. Consequently, the overall hydrogen output is reduced compared to the expected theoretical yield [14]. Based on these results, 1.8 moles of hydrogen are produced per mole of glucose, which is consistent with the literature data for this metabolic pathway [4].
To mitigate efficiency losses and optimize process conditions, pH control and process optimization significantly influence metabolic pathway selection. Maintaining an optimal pH range (typically 5.5–6.0) can favour acetate production over butyrate, thereby enhancing hydrogen yield. Selecting or genetically modifying microbial strains with higher hydrogen-producing efficiency can help direct metabolic flux towards the acetate pathway while minimizing the butyrate shift. Adjusting nutrient composition, C/N ratio, and supplementing trace metals (e.g., iron, nickel) can enhance enzymatic activity in the hydrogenase pathway, promoting acetate fermentation.

5. Possible Future Directions

Future research in dark fermentation for hydrogen production is a promising area that seeks to overcome several key challenges. Biological processes such as fermentation still face application problems because of the existing bottlenecks. The main problem is cost, which is higher because of lower hydrogen yields and the instability of the process. Addressing these limitations requires innovative approaches to enhance both efficiency and stability.
One major focus of future research is the optimization of microbial strains involved in dark fermentation. By employing genetic engineering and synthetic biology techniques, researchers aim to develop more robust strains with higher hydrogen yields and better resistance to fluctuations in environmental conditions. Furthermore, identifying novel microorganisms or consortia capable of operating under more diverse substrates or conditions can significantly improve the overall sustainability of the process. It is necessary to consider the ethical and regulatory rules. All genetically engineered strains are developed following biosafety guidelines established by organizations such as the EPA, FDA, and EFSA, ensuring their controlled use in research and industrial applications. Comprehensive risk assessments evaluate potential ecological impacts before deployment, and continuous monitoring programmes are implemented to track microbial behaviour in operational settings.
Another critical area is improving reactor design and operation strategies. Current bioreactors often struggle with maintaining stable fermentation conditions, which directly impacts hydrogen production efficiency. Research efforts are directed toward advanced reactor technologies, such as immobilized cell systems and continuous-flow reactors, to stabilize the process and enhance yields. Integrating fermentation with other processes, such as bio-electrochemical systems, could also help address inefficiencies and improve energy output.
Substrate utilization and pretreatment are also important research areas. By investigating the specific relationships between substrate concentration, microbial activity, and hydrogen production rates, the optimal conditions for maximizing hydrogen yield are identified. This includes studying the kinetics of different substrates, such as complex lignocellulosic biomass, and how their degradation rates influence hydrogen evolution. Developing cost-effective methods to utilize a wide variety of waste materials and biomass can reduce feedstock costs and contribute to a more sustainable hydrogen economy. Innovations in pretreatment technologies, including enzymatic and thermal approaches, can make substrates more accessible for microbial degradation, improving overall hydrogen yield.
Finally, integrating dark fermentation with other renewable hydrogen production methods, like photo-fermentation or water-splitting techniques, offers a hybrid approach that may overcome current bottlenecks. Combining different methods can increase overall efficiency and lower costs, pushing dark fermentation closer to large-scale, commercial viability.
In summary, while dark fermentation faces challenges related to cost, yield, and stability, future research focused on microbial strain improvement, reactor design, and process integration holds great potential to overcome these barriers and unlock its potential as a sustainable hydrogen production method.

6. Conclusions

Hydrogen production through dark fermentation (DF) depends on substrate type, microorganism selection, and key bioreactor conditions such as temperature, pH, and hydrogen partial pressure. Easily degradable carbohydrates generally yield higher hydrogen levels, while different bacterial strains influence production efficiency through varied metabolic pathways. Operational conditions like temperature, pH, and hydrogen partial pressure significantly impact process efficiency, with mesophilic (30–40 °C) and thermophilic (50–60 °C) conditions being optimal for microbial activity. Maintaining pH within the range of 5.0 to 6.5 supports bacterial growth and enzyme function, while controlling hydrogen partial pressure prevents inhibitory feedback effects.
This study focuses on hydrogen production from straw via dark fermentation, analyzing the metabolic pathways involved. A simple process model and simulation were conducted for the first time to better understand and optimize hydrogen production in this context. The simulation framework allows researchers to examine interactions between variables, such as substrate concentration, hydrogen production rate, and operating conditions, providing insights into optimizing bioreactor performance. By predicting how changes in one parameter affect overall system efficiency, the model enables precise control over the fermentation process to enhance hydrogen yield.
The results confirm that straw is a viable biomass source for hydrogen production due to its abundance and high carbohydrate content. However, the efficiency of hydrogen yield depends on factors like microbial consortia, pH control, and temperature optimization.
One of the key findings of the simulation is the identification of bottlenecks that can hinder large-scale hydrogen production, such as substrate availability, microbial activity, and operational inefficiencies related to temperature and pH fluctuations. Addressing these challenges is critical for successfully scaling up dark fermentation from laboratory to industrial applications.
Additionally, the findings indicate that byproducts from the fermentation process can significantly impact the overall cost of hydrogen production. Byproducts, such as acetic or butyric acids, may require additional processing, which increases the complexity and expense of the system. Reducing or managing these byproducts is essential for making hydrogen production more economically viable.

Author Contributions

B.M. (Babak Mokhtarani): writing—original draft, editing, software, data curation, methodology, validation, investigation. J.Z.: conceptualization, resources, supervision, writing—review and editing. B.M. (Behdad Moghtaderi): funding acquisition, supervision, project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript.
ADPadenosine diphosphate
AFBRanaerobic fluidized bed reactors
ATPadenosine triphosphate
CSTRcontinuous stirred tank reactors
DFdark fermentation
HRThydraulic retention time
LHVlower heating value
MBRmembrane bioreactors
MECmicrobial electrolysis cells
NADHnicotinamide adenine dinucleotide
PFLpyruvate formate lyase
PFORpyruvate ferredoxin oxidoreductase
TStotal solid
TVStotal volatile solid
UASBupflow anaerobic sludge blanket reactor

References

  1. Singh, L.; Wahid, Z.A. Methods for enhancing bio-hydrogen production from biological process: A review. J. Ind. Eng. Chem. 2015, 21, 70–80. [Google Scholar] [CrossRef]
  2. Reungsang, A.; Sreela-or, C. Bio-hydrogen production from pineapple waste extract by anaerobic mixed cultures. Energies 2013, 6, 2175–2190. [Google Scholar] [CrossRef]
  3. Sinha, P.; Pandey, A. An evaluative report and challenges for fermentative biohydrogen production. Int. J. Hydrogen Energy 2011, 36, 7460–7478. [Google Scholar] [CrossRef]
  4. Wang, J.; Wan, W. Factors influencing fermentative hydrogen production: A review. Int. J. Hydrogen Energy 2009, 34, 799–811. [Google Scholar] [CrossRef]
  5. Srivastava, P.; García-Quismondo, E.; Palma, J.; González-Fernández, C. Coupling dark fermentation and microbial electrolysis cells for higher hydrogen yield: Technological competitiveness and challenges. Int. J. Hydrogen Energy 2024, 52, 223–239. [Google Scholar] [CrossRef]
  6. Łukajtis, R.; Hołowacz, I.; Kucharska, K.; Glinka, M.; Rybarczyk, P.; Przyjazny, A.; Kamiński, M. Hydrogen production from biomass using dark fermentation. Renew. Sustain. Energy Rev. 2018, 91, 665–694. [Google Scholar] [CrossRef]
  7. Javed, M.A.; Zafar, A.M.; Hassan, A.A.; Zaidi, A.A.; Farooq, M.; El Badawy, A.; Lundquist, T.; Mohamed, M.M.A.; Al-Zuhair, S. The role of oxygen regulation and algal growth parameters in hydrogen production via biophotolysis. J. Environ. Chem. Eng. 2022, 10, 107003. [Google Scholar] [CrossRef]
  8. Akhlaghi, N.; Najafpour-Darzi, G. A comprehensive review on biological hydrogen production. Int. J. Hydrogen Energy 2020, 45, 22492–22512. [Google Scholar] [CrossRef]
  9. Azwar, M.; Hussain, M.; Abdul-Wahab, A. Development of biohydrogen production by photobiological, fermentation and electrochemical processes: A review. Renew. Sustain. Energy Rev. 2014, 31, 158–173. [Google Scholar] [CrossRef]
  10. Ndayisenga, F.; Yu, Z.; Zheng, J.; Wang, B.; Liang, H.; Phulpoto, I.A.; Habiyakare, T.; Zhou, D. Microbial electrohydrogenesis cell and dark fermentation integrated system enhances biohydrogen production from lignocellulosic agricultural wastes: Substrate pretreatment towards optimization. Renew. Sustain. Energy Rev. 2021, 145, 111078. [Google Scholar] [CrossRef]
  11. Müller, L.J.; Kätelhön, A.; Bringezu, S.; McCoy, S.; Suh, S.; Edwards, R.; Sick, V.; Kaiser, S.; Cuéllar-Franca, R.; El Khamlichi, A. The carbon footprint of the carbon feedstock CO2. Energy Environ. Sci. 2020, 13, 2979–2992. [Google Scholar] [CrossRef]
  12. Das, D.; Veziroglu, T.N. Advances in biological hydrogen production processes. Int. J. Hydrogen Energy 2008, 33, 6046–6057. [Google Scholar] [CrossRef]
  13. Ahmad, A.; Rambabu, K.; Hasan, S.W.; Show, P.L.; Banat, F. Biohydrogen production through dark fermentation: Recent trends and advances in transition to a circular bioeconomy. Int. J. Hydrogen Energy 2024, 52, 335–357. [Google Scholar] [CrossRef]
  14. Ghimire, A.; Frunzo, L.; Pirozzi, F.; Trably, E.; Escudie, R.; Lens, P.N.; Esposito, G. A review on dark fermentative biohydrogen production from organic biomass: Process parameters and use of by-products. Appl. Energy 2015, 144, 73–95. [Google Scholar] [CrossRef]
  15. Yogalakshmi, K.; Sivashanmugam, P.; Kavitha, S.; Kannah, Y.; Varjani, S.; AdishKumar, S.; Kumar, G. Lignocellulosic biomass-based pyrolysis: A comprehensive review. Chemosphere 2022, 286, 131824. [Google Scholar]
  16. Greses, S.; Tomás-Pejó, E.; Gónzalez-Fernández, C. Agroindustrial waste as a resource for volatile fatty acids production via anaerobic fermentation. Bioresour. Technol. 2020, 297, 122486. [Google Scholar] [CrossRef] [PubMed]
  17. Trancone, G.; Spasiano, D.; Race, M.; Luongo, V.; Petrella, A.; Pirozzi, F.; Fratino, U.; Piccinni, A. A combined system for asbestos-cement waste degradation by dark fermentation and resulting supernatant valorization in anaerobic digestion. Chemosphere 2022, 300, 134500. [Google Scholar] [CrossRef]
  18. Nemestóthy, N.; Bélafi-Bakó, K.; Bakonyi, P. Enhancement of dark fermentative H2 production by gas separation membranes: A review. Bioresour. Technol. 2020, 302, 122828. [Google Scholar] [CrossRef] [PubMed]
  19. Brar, K.K.; Cortez, A.A.; Pellegrini, V.O.; Amulya, K.; Polikarpov, I.; Magdouli, S.; Kumar, M.; Yang, Y.-H.; Bhatia, S.K.; Brar, S.K. An overview on progress, advances, and future outlook for biohydrogen production technology. Int. J. Hydrogen Energy 2022, 47, 37264–37281. [Google Scholar] [CrossRef]
  20. Show, K.-Y.; Lee, D.-J.; Chang, J.-S. Bioreactor and process design for biohydrogen production. Bioresour. Technol. 2011, 102, 8524–8533. [Google Scholar] [CrossRef] [PubMed]
  21. Ren, N.; Guo, W.; Liu, B.; Cao, G.; Ding, J. Biological hydrogen production by dark fermentation: Challenges and prospects towards scaled-up production. Curr. Opin. Biotechnol. 2011, 22, 365–370. [Google Scholar] [CrossRef] [PubMed]
  22. Kapdan, I.K.; Kargi, F. Bio-hydrogen production from waste materials. Enzym. Microb. Technol. 2006, 38, 569–582. [Google Scholar] [CrossRef]
  23. Li, C.; Fang, H.H. Fermentative hydrogen production from wastewater and solid wastes by mixed cultures. Crit. Rev. Environ. Sci. Technol. 2007, 37, 1–39. [Google Scholar] [CrossRef]
  24. Show, K.; Lee, D.; Tay, J.; Lin, C.; Chang, J.S. Biohydrogen production: Current perspectives and the way forward. Int. J. Hydrogen Energy 2012, 37, 15616–15631. [Google Scholar] [CrossRef]
  25. Wong, Y.M.; Wu, T.Y.; Juan, J.C. A review of sustainable hydrogen production using seed sludge via dark fermentation. Renew. Sustain. Energy Rev. 2014, 34, 471–482. [Google Scholar] [CrossRef]
  26. Ntaikou, I.; Antonopoulou, G.; Lyberatos, G. Biohydrogen production from biomass and wastes via dark fermentation: A review. Waste Biomass Valorization 2010, 1, 21–39. [Google Scholar] [CrossRef]
  27. De Gioannis, G.; Muntoni, A.; Polettini, A.; Pomi, R. A review of dark fermentative hydrogen production from biodegradable municipal waste fractions. Waste Manag. 2013, 33, 1345–1361. [Google Scholar] [CrossRef]
  28. Ariunbaatar, J.; Panico, A.; Esposito, G.; Pirozzi, F.; Lens, P.N. Pretreatment methods to enhance anaerobic digestion of organic solid waste. Appl. Energy 2014, 123, 143–156. [Google Scholar] [CrossRef]
  29. Mishra, P.; Roy, S.; Das, D. Comparative evaluation of the hydrogen production by mixed consortium, synthetic co-culture and pure culture using distillery effluent. Bioresour. Technol. 2015, 198, 593–602. [Google Scholar] [CrossRef] [PubMed]
  30. Rodrigues, C.V.; Nespeca, M.G.; Sakamoto, I.K.; de Oliveira, J.E.; Varesche, M.B.A.; Maintinguer, S.I. Bioconversion of crude glycerol from waste cooking oils into hydrogen by sub-tropical mixed and pure cultures. Int. J. Hydrogen Energy 2019, 44, 144–154. [Google Scholar] [CrossRef]
  31. Mthethwa, N.P.; Nasr, M.; Kiambi, S.L.; Bux, F.; Kumari, S. Biohydrogen fermentation from Pistia stratiotes (aquatic weed) using mixed and pure bacterial cultures. Int. J. Hydrogen Energy 2019, 44, 17720–17731. [Google Scholar] [CrossRef]
  32. Rosa, D.; Medeiros, A.B.P.; Martinez-Burgos, W.J.; do Nascimento Junior, J.R.; de Carvalho, J.C.; Sydney, E.B.; Soccol, C.R. Biological hydrogen production from palm oil mill effluent (POME) by anaerobic consortia and Clostridium beijerinckii. J. Biotechnol. 2020, 323, 17–23. [Google Scholar] [CrossRef]
  33. Trancone, G.; Policastro, G.; Spasiano, D.; Race, M.; Parrino, F.; Fratino, U.; Fabbricino, M.; Pirozzi, F. Treatment of concrete waste from construction and demolition activities: Application of organic acids from continuous dark fermentation in moving bed biofilm reactors. Chem. Eng. J. 2025, 505, 159536. [Google Scholar] [CrossRef]
  34. Mohanakrishna, G.; Pengadeth, D. Mixed culture biotechnology and its versatility in dark fermentative hydrogen production. Bioresour. Technol. 2024, 394, 130286. [Google Scholar] [CrossRef]
  35. Cavinato, C.; Giuliano, A.; Bolzonella, D.; Pavan, P.; Cecchi, F. Bio-hythane production from food waste by dark fermentation coupled with anaerobic digestion process: A long-term pilot scale experience. Int. J. Hydrogen Energy 2012, 37, 11549–11555. [Google Scholar] [CrossRef]
  36. 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. Hydrogen Energy 2006, 31, 2147–2157. [Google Scholar] [CrossRef]
  37. Jayalakshmi, S.; Joseph, K.; Sukumaran, V. Bio hydrogen generation from kitchen waste in an inclined plug flow reactor. Int. J. Hydrogen Energy 2009, 34, 8854–8858. [Google Scholar] [CrossRef]
  38. Yahaya, E.; Lim, S.W.; Yeo, W.S.; Nandong, J. A review on process modeling and design of biohydrogen. Int. J. Hydrogen Energy 2022, 47, 30404–30427. [Google Scholar] [CrossRef]
  39. Cao, Y.; Liu, H.; Liu, W.; Guo, J.; Xian, M. Debottlenecking the biological hydrogen production pathway of dark fermentation: Insight into the impact of strain improvement. Microb. Cell Factories 2022, 21, 166. [Google Scholar] [CrossRef] [PubMed]
  40. Meegoda, J.N.; Li, B.; Patel, K.; Wang, L.B. A review of the processes, parameters, and optimization of anaerobic digestion. Int. J. Environ. Res. Public Health 2018, 15, 2224. [Google Scholar] [CrossRef] [PubMed]
  41. Tapia-Venegas, E.; Ramirez-Morales, J.E.; Silva-Illanes, F.; Toledo-Alarcón, J.; Paillet, F.; Escudie, R.; Lay, C.-H.; Chu, C.-Y.; Leu, H.-J.; Marone, A. Biohydrogen production by dark fermentation: Scaling-up and technologies integration for a sustainable system. Rev. Environ. Sci. Bio/Technol. 2015, 14, 761–785. [Google Scholar] [CrossRef]
  42. Xu, R.-z.; Fang, S.; Zhang, L.; Huang, W.; Shao, Q.; Fang, F.; Feng, Q.; Cao, J.; Luo, J. Distribution patterns of functional microbial community in anaerobic digesters under different operational circumstances: A review. Bioresour. Technol. 2021, 341, 125823. [Google Scholar] [CrossRef]
  43. Detman, A.; Laubitz, D.; Chojnacka, A.; Wiktorowska-Sowa, E.; Piotrowski, J.; Salamon, A.; Kaźmierczak, W.; Błaszczyk, M.K.; Barberan, A.; Chen, Y. Dynamics and complexity of dark fermentation microbial communities producing hydrogen from sugar beet molasses in continuously operating packed bed reactors. Front. Microbiol. 2021, 11, 612344. [Google Scholar] [CrossRef]
  44. Navarro-Díaz, M.; Aparicio-Trejo, V.; Valdez-Vazquez, I.; Carrillo-Reyes, J.; Avitia, M.; Escalante, A.E. Levels of microbial diversity affect the stability and function of dark fermentation bioreactors. Front. Ind. Microbiol. 2024, 2, 1386726. [Google Scholar] [CrossRef]
  45. Chandrasekhar, K.; Lee, Y.-J.; Lee, D.-W. Biohydrogen production: Strategies to improve process efficiency through microbial routes. Int. J. Mol. Sci. 2015, 16, 8266–8293. [Google Scholar] [CrossRef]
  46. Hallenbeck, P.C.; Benemann, J.R. Biological hydrogen production; fundamentals and limiting processes. Int. J. Hydrogen Energy 2002, 27, 1185–1193. [Google Scholar] [CrossRef]
  47. Nath, K.; Das, D. Improvement of fermentative hydrogen production: Various approaches. Appl. Microbiol. Biotechnol. 2004, 65, 520–529. [Google Scholar] [CrossRef]
  48. Hallenbeck, P.C. Microbial Technologies in Advanced Biofuels Production; Springer: Berlin/Heidelberg, Germany, 2012; Volume 1. [Google Scholar]
  49. Yokoi, H.; Ohkawara, T.; Hirose, J.; Hayashi, S.; Takasaki, Y. Characteristics of hydrogen production by aciduric Enterobacter aerogenes strain HO-39. J. Ferment. Bioeng. 1995, 80, 571–574. [Google Scholar] [CrossRef]
  50. Kumar, N.; Das, D. Enhancement of hydrogen production by Enterobacter cloacae IIT-BT 08. Process Biochem. 2000, 35, 589–593. [Google Scholar] [CrossRef]
  51. Segers, B.; Nimmegeers, P.; Spiller, M.; Tofani, G.; Grojzdek, E.J.; Dace, E.; Kikas, T.; Marchetti, J.M.; Rajić, M.; Yildiz, G. Lignocellulosic biomass valorisation: A review of feedstocks, processes and potential value chains, and their implications for the decision-making process. RSC Sustain. 2024, 2, 3730–3749. [Google Scholar] [CrossRef]
  52. Basera, P.; Chakraborty, S.; Sharma, N. Lignocellulosic biomass: Insights into enzymatic hydrolysis, influential factors, and economic viability. Discov. Sustain. 2024, 5, 311. [Google Scholar] [CrossRef]
  53. Mumtha, C.; Mahalingam, P.U. Biohydrogen production from co-substrates through dark fermentation by bacterial consortium. 3 Biotech 2024, 14, 281. [Google Scholar] [CrossRef]
  54. Romão, B.; Batista, F.; Ferreira, J.; Costa, H.; Resende, M.; Cardoso, V. Biohydrogen production through dark fermentation by a microbial consortium using whey permeate as substrate. Appl. Biochem. Biotechnol. 2014, 172, 3670–3685. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, C.; Chang, C.; Chu, C.; Lee, D.; Chang, B.-V.; Liao, C. Producing hydrogen from wastewater sludge by Clostridium bifermentans. J. Biotechnol. 2003, 102, 83–92. [Google Scholar] [CrossRef] [PubMed]
  56. Cai, M.; Liu, J.; Wei, Y. Enhanced biohydrogen production from sewage sludge with alkaline pretreatment. Environ. Sci. Technol. 2004, 38, 3195–3202. [Google Scholar] [CrossRef]
  57. Rawat, S.; Rautela, A.; Yadav, I.; Misra, S.; Kumar, S. A comprehensive review on enhanced biohydrogen production: Pretreatment, applied strategies, techno-economic assessment, and future perspective. BioEnergy Res. 2023, 16, 2131–2154. [Google Scholar] [CrossRef]
  58. Carrillo-Reyes, J.; Valdez-Vazquez, I.; Vital-Jácome, M.; Vargas, A.; Navarro-Díaz, M.; Cortez-Cervantes, J.; Chango-Cañola, A.P. Microbial Communities in Dark Fermentation, Analytical Tools to Elucidate Key Microorganisms and Metabolic Profiles. In Wastewater Exploitation: From Microbiological Activity to Energy; Springer: Berlin/Heidelberg, Germany, 2024; pp. 107–132. [Google Scholar]
  59. Oh, S.E.; Iyer, P.; Bruns, M.A.; Logan, B.E. Biological hydrogen production using a membrane bioreactor. Biotechnol. Bioeng. 2004, 87, 119–127. [Google Scholar] [CrossRef]
  60. Chu, C.-F.; Xu, K.-Q.; Li, Y.-Y.; Inamori, Y. Hydrogen and methane potential based on the nature of food waste materials in a two-stage thermophilic fermentation process. Int. J. Hydrogen Energy 2012, 37, 10611–10618. [Google Scholar] [CrossRef]
  61. Wu, X.; Zhu, J.; Dong, C.; Miller, C.; Li, Y.; Wang, L.; Yao, W. Continuous biohydrogen production from liquid swine manure supplemented with glucose using an anaerobic sequencing batch reactor. Int. J. Hydrogen Energy 2009, 34, 6636–6645. [Google Scholar] [CrossRef]
  62. Xing, Y.; Li, Z.; Fan, Y.; Hou, H. Biohydrogen production from dairy manures with acidification pretreatment by anaerobic fermentation. Environ. Sci. Pollut. Res. 2010, 17, 392–399. [Google Scholar] [CrossRef] [PubMed]
  63. Sheng, T.; Gao, L.; Zhao, L.; Liu, W.; Wang, A. Direct hydrogen production from lignocellulose by the newly isolated Thermoanaerobacterium thermosaccharolyticum strain DD32. RSC Adv. 2015, 5, 99781–99788. [Google Scholar] [CrossRef]
  64. Cui, M.; Shen, J. Effects of acid and alkaline pretreatments on the biohydrogen production from grass by anaerobic dark fermentation. Int. J. Hydrogen Energy 2012, 37, 1120–1124. [Google Scholar] [CrossRef]
  65. Ivanova, G.; Rákhely, G.; Kovács, K.L. Thermophilic biohydrogen production from energy plants by Caldicellulosiruptor saccharolyticus and comparison with related studies. Int. J. Hydrogen Energy 2009, 34, 3659–3670. [Google Scholar] [CrossRef]
  66. Venetsaneas, N.; Antonopoulou, G.; Stamatelatou, K.; Kornaros, M.; Lyberatos, G. Using cheese whey for hydrogen and methane generation in a two-stage continuous process with alternative pH controlling approaches. Bioresour. Technol. 2009, 100, 3713–3717. [Google Scholar] [CrossRef] [PubMed]
  67. Reungsang, A.; Sittijunda, S.; Sompong, O. Bio-hydrogen production from glycerol by immobilized Enterobacter aerogenes ATCC 13048 on heat-treated UASB granules as affected by organic loading rate. Int. J. Hydrogen Energy 2013, 38, 6970–6979. [Google Scholar] [CrossRef]
  68. Chu, C.-Y.; Tung, L.; Lin, C.-Y. Effect of substrate concentration and pH on biohydrogen production kinetics from food industry wastewater by mixed culture. Int. J. Hydrogen Energy 2013, 38, 15849–15855. [Google Scholar] [CrossRef]
  69. Nandi, R.; Sengupta, S. Microbial production of hydrogen: An overview. Crit. Rev. Microbiol. 1998, 24, 61–84. [Google Scholar] [CrossRef] [PubMed]
  70. Valdez-Vazquez, I.; Poggi-Varaldo, H.M. Hydrogen production by fermentative consortia. Renew. Sustain. Energy Rev. 2009, 13, 1000–1013. [Google Scholar] [CrossRef]
  71. Wang, J.; Wan, W. Comparison of different pretreatment methods for enriching hydrogen-producing bacteria from digested sludge. Int. J. Hydrogen Energy 2008, 33, 2934–2941. [Google Scholar] [CrossRef]
  72. Kotay, S.M.; Das, D. Novel dark fermentation involving bioaugmentation with constructed bacterial consortium for enhanced biohydrogen production from pretreated sewage sludge. Int. J. Hydrogen Energy 2009, 34, 7489–7496. [Google Scholar] [CrossRef]
  73. Zhang, J.-N.; Li, Y.-H.; Zheng, H.-Q.; Fan, Y.-T.; Hou, H.-W. Direct degradation of cellulosic biomass to bio-hydrogen from a newly isolated strain Clostridium sartagoforme FZ11. Bioresour. Technol. 2015, 192, 60–67. [Google Scholar] [CrossRef] [PubMed]
  74. Kogo, T.; Yoshida, Y.; Koganei, K.; Matsumoto, H.; Watanabe, T.; Ogihara, J.; Kasumi, T. Production of rice straw hydrolysis enzymes by the fungi Trichoderma reesei and Humicola insolens using rice straw as a carbon source. Bioresour. Technol. 2017, 233, 67–73. [Google Scholar] [CrossRef] [PubMed]
  75. Chang, S.; Li, J.-Z.; Liu, F. Evaluation of different pretreatment methods for preparing hydrogen-producing seed inocula from waste activated sludge. Renew. Energy 2011, 36, 1517–1522. [Google Scholar] [CrossRef]
  76. Deublein, D.; Steinhauser, A. Biogas from Waste and Renewable Resources: An Introduction; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  77. Mohan, S.V.; Babu, V.L.; Sarma, P. Effect of various pretreatment methods on anaerobic mixed microflora to enhance biohydrogen production utilizing dairy wastewater as substrate. Bioresour. Technol. 2008, 99, 59–67. [Google Scholar] [CrossRef] [PubMed]
  78. Liu, H.; Wang, J.; Wang, A.; Chen, J. Chemical inhibitors of methanogenesis and putative applications. Appl. Microbiol. Biotechnol. 2011, 89, 1333–1340. [Google Scholar] [CrossRef] [PubMed]
  79. Bundhoo, M.Z.; Mohee, R.; Hassan, M.A. Effects of pre-treatment technologies on dark fermentative biohydrogen production: A review. J. Environ. Manag. 2015, 157, 20–48. [Google Scholar] [CrossRef]
  80. Levin, D.B.; Pitt, L.; Love, M. Biohydrogen production: Prospects and limitations to practical application. Int. J. Hydrogen Energy 2004, 29, 173–185. [Google Scholar] [CrossRef]
  81. Valdez-Vazquez, I.; Ríos-Leal, E.; Esparza-García, F.; Cecchi, F.; Poggi-Varaldo, H.M. Semi-continuous solid substrate anaerobic reactors for H2 production from organic waste: Mesophilic versus thermophilic regime. Int. J. Hydrogen Energy 2005, 30, 1383–1391. [Google Scholar] [CrossRef]
  82. Shin, H.-S.; Youn, J.-H.; Kim, S.-H. Hydrogen production from food waste in anaerobic mesophilic and thermophilic acidogenesis. Int. J. Hydrogen Energy 2004, 29, 1355–1363. [Google Scholar] [CrossRef]
  83. Lee, K.-S.; Lin, P.-J.; Chang, J.-S. Temperature effects on biohydrogen production in a granular sludge bed induced by activated carbon carriers. Int. J. Hydrogen Energy 2006, 31, 465–472. [Google Scholar] [CrossRef]
  84. Foglia, D.; Wukovits, W.; Friedl, A.; de Vrije, T.; Claassen, P.A. Fermentative Hydrogen Production: Influence of Application of Mesophilic and Thermophilic Bacteria on Mass and Energy Balances. 2011. Available online: https://research.wur.nl/en/publications/fermentative-hydrogen-production-influence-of-application-of-meso (accessed on 21 February 2025).
  85. Khanal, S.K.; Chen, W.-H.; Li, L.; Sung, S. Biological hydrogen production: Effects of pH and intermediate products. Int. J. Hydrogen Energy 2004, 29, 1123–1131. [Google Scholar] [CrossRef]
  86. Lin, C.-Y.; Lay, C.-H.; Sen, B.; Chu, C.-Y.; Kumar, G.; Chen, C.-C.; Chang, J.-S. Fermentative hydrogen production from wastewaters: A review and prognosis. Int. J. Hydrogen Energy 2012, 37, 15632–15642. [Google Scholar] [CrossRef]
  87. Bowles, L.K.; Ellefson, W. Effects of butanol on Clostridium acetobutylicum. Appl. Environ. Microbiol. 1985, 50, 1165–1170. [Google Scholar] [CrossRef]
  88. 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]
  89. Temudo, M.F.; Kleerebezem, R.; van Loosdrecht, M. Influence of the pH on (open) mixed culture fermentation of glucose: A chemostat study. Biotechnol. Bioeng. 2007, 98, 69–79. [Google Scholar] [CrossRef] [PubMed]
  90. Fang, H.H.; Liu, H. Effect of pH on hydrogen production from glucose by a mixed culture. Bioresour. Technol. 2002, 82, 87–93. [Google Scholar] [CrossRef]
  91. Guo, X.M.; Trably, E.; Latrille, E.; Carrère, H.; Steyer, J.-P. Hydrogen production from agricultural waste by dark fermentation: A review. Int. J. Hydrogen Energy 2010, 35, 10660–10673. [Google Scholar] [CrossRef]
  92. Chen, C.-C.; Chen, H.-P.; Wu, J.-H.; Lin, C.-Y. Fermentative hydrogen production at high sulfate concentration. Int. J. Hydrogen Energy 2008, 33, 1573–1578. [Google Scholar] [CrossRef]
  93. Wu, S.Y.; Lin, C.N.; Chang, J.S. Hydrogen production with immobilized sewage sludge in three-phase fluidized-bed bioreactors. Biotechnol. Prog. 2003, 19, 828–832. [Google Scholar] [CrossRef] [PubMed]
  94. Tapia-Venegas, E.; Ramirez, J.E.; Donoso-Bravo, A.; Jorquera, L.; Steyer, J.-P.; Ruiz-Filippi, G. Bio-hydrogen production during acidogenic fermentation in a multistage stirred tank reactor. Int. J. Hydrogen Energy 2013, 38, 2185–2190. [Google Scholar] [CrossRef]
  95. Bakonyi, P.; Nemestóthy, N.; Simon, V.; Bélafi-Bakó, K. Review on the start-up experiences of continuous fermentative hydrogen producing bioreactors. Renew. Sustain. Energy Rev. 2014, 40, 806–813. [Google Scholar] [CrossRef]
  96. Mandal, B.; Nath, K.; Das, D. Improvement of biohydrogen production under decreased partial pressure of H2 by Enterobacter cloacae. Biotechnol. Lett. 2006, 28, 831–835. [Google Scholar] [CrossRef]
  97. 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. Hydrogen Energy 2007, 32, 172–184. [Google Scholar] [CrossRef]
  98. Balachandar, G.; Khanna, N.; Das, D. Biohydrogen production from organic wastes by dark fermentation. In Biohydrogen; Elsevier: Amsterdam, The Netherlands, 2013; pp. 103–144. [Google Scholar]
  99. Gavala, H.N.; Skiadas, I.V.; Ahring, B.K. Biological hydrogen production in suspended and attached growth anaerobic reactor systems. Int. J. Hydrogen Energy 2006, 31, 1164–1175. [Google Scholar] [CrossRef]
  100. Jung, K.-W.; Kim, D.-H.; Kim, S.-H.; Shin, H.-S. Bioreactor design for continuous dark fermentative hydrogen production. Bioresour. Technol. 2011, 102, 8612–8620. [Google Scholar] [CrossRef]
  101. Lee, K.-S.; Lo, Y.-C.; Lin, P.-J.; Chang, J.-S. Improving biohydrogen production in a carrier-induced granular sludge bed by altering physical configuration and agitation pattern of the bioreactor. Int. J. Hydrogen Energy 2006, 31, 1648–1657. [Google Scholar] [CrossRef]
  102. Zhang, Z.-P.; Tay, J.-H.; Show, K.-Y.; Yan, R.; Liang, D.T.; Lee, D.-J.; Jiang, W.-J. Biohydrogen production in a granular activated carbon anaerobic fluidized bed reactor. Int. J. Hydrogen Energy 2007, 32, 185–191. [Google Scholar] [CrossRef]
  103. Zhang, Z.-P.; Show, K.-Y.; Tay, J.-H.; Liang, D.T.; Lee, D.-J. Biohydrogen production with anaerobic fluidized bed reactors—A comparison of biofilm-based and granule-based systems. Int. J. Hydrogen Energy 2008, 33, 1559–1564. [Google Scholar] [CrossRef]
  104. Lee, Y.-W.; Chung, J. Bioproduction of hydrogen from food waste by pilot-scale combined hydrogen/methane fermentation. Int. J. Hydrogen Energy 2010, 35, 11746–11755. [Google Scholar] [CrossRef]
  105. Lin, C.-Y.; Wu, S.-Y.; Lin, P.-J.; Chang, J.-S.; Hung, C.-H.; Lee, K.-S.; Lay, C.-H.; Chu, C.-Y.; Cheng, C.-H.; Chang, A.C. A pilot-scale high-rate biohydrogen production system with mixed microflora. Int. J. Hydrogen Energy 2011, 36, 8758–8764. [Google Scholar] [CrossRef]
  106. Mechery, J.; Thomas, D.M.; Kumar, C.P.; Joseph, L.; Sylas, V. Biohydrogen production from acidic and alkaline hydrolysates of paddy straw using locally isolated facultative bacteria through dark fermentation. Biomass Convers. Biorefinery 2021, 11, 1263–1272. [Google Scholar] [CrossRef]
  107. Kumar, K.; Roy, S.; Das, D. Continuous mode of carbon dioxide sequestration by C. sorokiniana and subsequent use of its biomass for hydrogen production by E. cloacae IIT-BT 08. Bioresour. Technol. 2013, 145, 116–122. [Google Scholar] [CrossRef]
  108. Niu, M.; Jin, B.; Huang, Y.; Wang, H.; Dong, Q.; Gu, H.; Yang, J. Co-gasification of high-ash sewage sludge and straw in a bubbling fluidized bed with oxygen-enriched air. Int. J. Chem. React. Eng. 2018, 16, 20170044. [Google Scholar] [CrossRef]
  109. Ruggeri, B.; Tommasi, T.; Sassi, G. Experimental kinetics and dynamics of hydrogen production on glucose by hydrogen forming bacteria (HFB) culture. Int. J. Hydrogen Energy 2009, 34, 753–763. [Google Scholar] [CrossRef]
  110. Fabiano, B.; Perego, P. Thermodynamic study and optimization of hydrogen production by Enterobacter aerogenes. Int. J. Hydrogen Energy 2002, 27, 149–156. [Google Scholar] [CrossRef]
  111. Wong, Y.M.; Juan, J.C.; Ting, A.; Wu, T.Y. High efficiency bio-hydrogen production from glucose revealed in an inoculum of heat-pretreated landfill leachate sludge. Energy 2014, 72, 628–635. [Google Scholar] [CrossRef]
  112. Posten, C.H.; Cooney, C.L. Growth of microorganisms. Biotechnology 1993, 1, 111–162. [Google Scholar]
  113. Cecílio, D.M.; Gonçalves, J.R.M.; Correia, M.J.N.; Mateus, M.M. Aspen Plus® Modeling and Simulation of an Industrial Biomass Direct Liquefaction Process. Fuels 2023, 4, 221–242. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the pathways of hydrogen production by dark fermentation.
Figure 1. Schematic diagram of the pathways of hydrogen production by dark fermentation.
Energies 18 01092 g001
Figure 2. A simplified block diagram for hydrogen production with a dark fermentation process.
Figure 2. A simplified block diagram for hydrogen production with a dark fermentation process.
Energies 18 01092 g002
Figure 3. Process overview for hydrogen production with ASPEN Plus.
Figure 3. Process overview for hydrogen production with ASPEN Plus.
Energies 18 01092 g003
Table 1. Pros and cons of different types of substrates for dark fermentation.
Table 1. Pros and cons of different types of substrates for dark fermentation.
SubstrateExampleAdvantageDisadvantageHydrogen Yield (Mole H2/Mole Substrate)Reference
MonosaccharidesGlucose,
D-Xylose
Biodegradable substrate, high yieldHigh cost for industrial level4,
0.95
[59]
DisaccharidesSucrose, lactoseBiodegradable substrate, high yieldHigh cost for industrial level6,
3
[50]
Organic municipal wasteKitchen waste,
kitchen garbage
Low cost and availabilityPretreatment is needed72 cm3 H2/g VS,
66 cm3 H2/g VS
[37]
[60]
ManureSwine manure,
dairy manure
Reduce the gas pollution, low costWastewater treatment is needed1.63,
31.5 cm3 H2/g TVS
[61]
[62]
Agriculture wasteCornstalk,
bagasse,
wheat straw
Availability, low costLignin present in waste resists on biodegradation6.38 mole/kg substrate
2.3,
3.8
[63]
[64]
[65]
Industrial wasteCheese whey,
glycerol waste,
brewery wastewater
Reduce the cost of waste treatment, low costPresence unwanted component inhibit the process0.78,
0.33,
2
[66]
[67]
[68]
TVS: total volatile solid; VS: volatile solid.
Table 2. Pretreatment method for mixed cultural dark fermentation.
Table 2. Pretreatment method for mixed cultural dark fermentation.
Pretreatment MethodOperating ConditionAdvantage Disadvantage CommentReference
Heat shock80–121 °CIncrease cell permeability
Simple and cost-effective
Cell damage
Energy consumption
Treating the
culture with a high temperature
kills non-spore-forming bacteria.
[70,71]
Freezing and thawing−25 to −10 °C freezing
Thawing
Incubating 20–30 °C
Cell disruption
Low cost
Time-consuming
Limited scalability
Inoculum is frozen
and this is maintained, followed by thawing and the incubating
[72]
AerationTime: 2 hours–14 daysEnhance microbial growth
Low cost
Scaleable
Energy-intensiveuse of the air to eliminate the bacteria sensitive to oxygen.[73,74]
Acid and alkaline treatmentAcid: pH 2–4
Alkaline:pH 10–12
Efficient biomass breakdown
Widely used
Corrosive
chemical disposal
Adjustment of the pH of the inoculum to a value that microorganisms
cannot survive.
[75,76]
Chemical treatment2-bromoethane sulfonate (2-BES)
2-bromoethane sulfonic acid (2-BESA)
Selective breakdown
High efficiency
Toxic byproduct
High cost
Chemical compounds can block metabolic pathways of methanogenic
bacteria selectively.
[77,78]
MicrowaveFrequencies ranging from 300 MHz to 300 GHzRapid and efficient
Scalable
Uneven heating
Energy consumption
This applies for the pretreatment of lignocellulosic biomass
for enhanced hydrolysis
[79]
UltravioletWavelengths varying from 10 nm to 400 nmEffective for sterilization
Environmental friendly
Limited penetration
Cost
Ultraviolet lies between visible light and x-rays in the electromagnetic spectrum
It applies for inactivation of bacterial population
[79]
Table 4. Proximate and ultimate analysis of straw [108].
Table 4. Proximate and ultimate analysis of straw [108].
Ultimate Analysis (wt%)Proximate Analysis (wt%)
C42.5Moisture8.3
H6.3Volatiles70.3
N0.8Fixed carbon18.7
O38.7Ash11
S0.2
Cl0.5
LHV (MJ/kg)16.32
Table 5. ASPEN Plus simulation results for temperature, pressure, and composition of various streams in kg/h for hydrogen production from straw in dark fermentation process with consideration of acetate pathway.
Table 5. ASPEN Plus simulation results for temperature, pressure, and composition of various streams in kg/h for hydrogen production from straw in dark fermentation process with consideration of acetate pathway.
Streams No.12345678910111213
Temp. (°C)25252550503420202025252525
P (Bar)1111124441113
F (kg/h)10,000105,144115,144115,14435,26779,8772733275245877,144398473,16065,144
PhaseLiquidLiquidLiquidLiquidLiquidLiquidVaporVaporVaporLiquidLiquidLiquid Liquid
mBiomass10,000010,0000000000000
mash0001016101600000000
mCO200000025045024540000
mH2O0105,144105,144105,91231,77374,13900073,11473172,38165,144
mH200000022922540000
msolid0002433243300000000
mCH3COOH000000000341732461710
mglucose00056940569400056955640
mNH4Cl000683434000341340
msulphur000211110000101100
Table 6. ASPEN Plus simulation results for temperature, pressure and composition of various streams in kg/h for hydrogen production from straw in dark fermentation process with consideration of acetate and butyrate pathway.
Table 6. ASPEN Plus simulation results for temperature, pressure and composition of various streams in kg/h for hydrogen production from straw in dark fermentation process with consideration of acetate and butyrate pathway.
Streams No.12345678910111213
Temp. (°C)25252550503420202025252525
P (Bar)1111124441113
F (kg/h)10,000106,358116,358116,35835,63280,7272676219245778,051356574,48666,358
Phase LiquidLiquidLiquidLiquidLiquidLiquidVaporVaporVaporLiquidLiquidLiquid Liquid
mBiomass 10,000010,0000000000000
mash0001016101600000000
mCO200000025045024540000
mH2O0106,358106,358107,12632,13874,98800074,47674573,73166,358
mH200000017216930000
msolid0002433243300000000
mCH3COOH00000000017081623850
mbutyrate00000000012531190630
mglucose00056940569400056955640
mNH4Cl000683434000341340
msulphur000211110000101100
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mokhtarani, B.; Zanganeh, J.; Moghtaderi, B. A Review on Biohydrogen Production Through Dark Fermentation, Process Parameters and Simulation. Energies 2025, 18, 1092. https://doi.org/10.3390/en18051092

AMA Style

Mokhtarani B, Zanganeh J, Moghtaderi B. A Review on Biohydrogen Production Through Dark Fermentation, Process Parameters and Simulation. Energies. 2025; 18(5):1092. https://doi.org/10.3390/en18051092

Chicago/Turabian Style

Mokhtarani, Babak, Jafar Zanganeh, and Behdad Moghtaderi. 2025. "A Review on Biohydrogen Production Through Dark Fermentation, Process Parameters and Simulation" Energies 18, no. 5: 1092. https://doi.org/10.3390/en18051092

APA Style

Mokhtarani, B., Zanganeh, J., & Moghtaderi, B. (2025). A Review on Biohydrogen Production Through Dark Fermentation, Process Parameters and Simulation. Energies, 18(5), 1092. https://doi.org/10.3390/en18051092

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop