Summary
The development of biohydrogen as an alternative energy source has had great economic and environmental benefits. Hydrogen production from microalgae is considered a clean and sustainable energy production method that can both alleviate fuel shortages and recycle waste. Although algal hydrogen production has low energy consumption and requires only simple pretreatment, it has not been commercialized because of low product yields. To increase microalgal biohydrogen production several technologies have been developed, although they struggle with the oxygen sensitivity of the hydrogenases responsible for hydrogen production and the complexity of the metabolic network. In this review, several genetic and metabolic engineering studies on enhancing microalgal biohydrogen production are discussed, and the economic feasibility and future direction of microalgal biohydrogen commercialization are also proposed.
Subject areas: Biochemistry, Ecological biochemistry, Techniques in genetics
Graphical abstract
Biochemistry; Ecological biochemistry; Techniques in genetics
Introduction
Fossil fuels are the most important energy source globally and have brought great economic benefits.1 Their excessive use, however, has negative environmental and human health impacts. At present, more than 80% of global energy use originates from fossil fuels.1 With continuous population growth, fossil fuel demand will only increase, and the output of limited fossil fuel resources will inevitably decline in this century.2,3 Excessive exploitation of fossil fuels has also caused deforestation, loss of farmland, and damage to ecosystems because of acid rain.4 Global pollution caused by fossil fuels has multiple negative consequences5; because the burning of fossil fuels contributes to about 75% of global carbon-dioxide emissions, it is estimated that carbon emissions from fossil fuels will increase to 39 billion tons by 2030.6 Global warming caused by carbon emissions from human activities has resulted in a global temperature 1°C higher than that before industrialization. Global warming has also led to increased frequency and severity of ice melt, leading to rising sea levels and extreme weather.7 The combustion of fossil fuels also negatively impacts human health, causing and exacerbating respiratory and nervous system diseases.8 Reducing reliance on fossil fuels and finding alternative energy sources are therefore crucial for the long-term sustainability of human development.9
Hydrogen has the highest energy content per unit of mass of all known fuels. It is the most abundant element in the universe and is a clean energy source because water is the only by-product of its combustion.10 The high energy efficiency, high energy density, and safety of solid hydrogen demonstrate promise for scaling hydrogen energy.11 As hydrogen is considered the most promising alternative energy source, hydrogen produced from renewable resources could be a sustainable and clean fuel.12,13 Biomass-derived hydrogen can be generated through thermochemical and microbial processes.14,15 The low concentration of hydrogen produced by thermochemical methods and the need for high-temperature complex operating conditions increase production costs and energy consumption, making them unsustainable.16,17 Compared with chemical hydrogen production, biological hydrogen production by microbes can be carried out under lower temperatures and pressures.18,19 Microorganisms can easily become hydrogen production sources.20
Microalgae are a kind of biomass with great potential for hydrogen production.21 Owing to their highly adaptable nature, they are widely distributed globally.22 Microalgae are the basis of the global carbon cycle and absorb 50% of global carbon dioxide.23 They also have rapid growth rates as well as high yields and carbohydrate contents.24 They can be cultured in a variety of substrates including fresh water, seawater, domestic sewage, and wastewater.25,26 As microalgae convert light into chemical energy and produce hydrogen by cracking water, they are considered a green production source. In addition, algal biomass can be cultured at a large scale, and could be used as a substrate fuel source for hydrogen production.27,28,29 Novel strategies for microalgal hydrogen production have been explored to improve competitiveness.30 Various pretreatments could also improve hydrogen recovery.31 Despite these successes, there are some obstacles to developing large-scale hydrogen production using algae.24 The production of hydrogen in these systems depends on interactions in the microalgal metabolic network, including photosynthesis, respiration, and fermentation; these natural processes are not enough to support large-scale industrial production.32,33 With the application of genetic engineering, microalgae are becoming the most promising microorganisms for hydrogen production.19 Directed transformation has been shown to improve hydrogen production capacity.34
This review provides an overview of pretreatment techniques for microalgal biomass and discusses methods to improve biohydrogen production using genetic and metabolic engineering. Many strategies, including adjusting environmental tolerance, metabolic remodeling, and the expression of synthetic pathways from high-yield strains have been adopted to improve hydrogen production on an industrial scale.35
Microalgal biomass as a raw material for biohydrogen production
Introduction of algal substrate
The selection of the right substrate for industrial-scale fuel production is important because it accounts for 60–70% of the cost of production.36 Traditional hydrogen production methods involving lignocellulose consumption lead to competition with farmland and utilization of land that could be used for other purposes.18 The crystalline structure of lignocellulose limits its utilization by microorganisms and enzymes. The scarcity of land also hampers energy crops.37 In contrast, microalgae have almost no lignin and low hemicellulose content, and the required pretreatment is not energy-intensive and does not require the use of ecologically toxic solvents that can cause secondary pollution.38 Biofuel production has shifted from using sugar and starch to lignocellulosic materials, and finally to the third generation of algal biomass in recent years.36 The transformation of microalgal biomass is an efficient and environmentally friendly strategy to produce hydrogen.39 It can be converted into hydrogen by microbial fermentation.40 ‘Microalgae’ is a general term for algae whose morphology can be distinguished under a microscope.41 They are extremely diverse single-cell populations that can be as small as 0.2 μm.42 Microalgae can be composed of one or a few cells, and there can also be structures formed by the aggregation of many cells.35,24 The composition of algal cells also varies with species and the surrounding environment, with carbohydrate content reaching as high as 70%, making them ideal substrates for biohydrogen fermentation.43,44
Pretreatment of algal biomass
Pretreatment increases biohydrogen production and is a key step used to extract the main components from cells for conversion into biofuels.25 Pretreatment of algal biomass to degrade cell walls enhances fermentative H2 productivity by 50–70%.45 Because microalgae lack lignin and hemicellulose, only simple pretreatment is required greatly reducing costs.46 Biohydrogen is mainly obtained from the conversion of carbohydrates, so the best pretreatment strategies aim to release more carbohydrates.43 The destruction of a cell wall composed of cellulose, pectin, and algaenan is the main obstacle.47
The properties of microalgae are varied and pretreatment conditions vary from strain to strain.36 Usually, pretreatments are classified into four categories: physical, chemical, biological, and combinatorial (a combination of several treatment methods).48 Common pretreatment methods and their advantages are shown in Table 1. Physical pretreatment uses external machinery or electricity to destroy cells, changing the surface area and biomass crystallinity; hydrothermal pretreatment, milling, ultrasonic, and microwave irradiation are the main methods used.49 Chemical pretreatment uses chemicals to extract biomolecules.49 The most commonly used acid pretreatments are HCl, H2SO4, and HNO3 with concentrations ranging between 0.1 and 6.0 M. The most commonly used alkali pretreatments are NaOH, KOH, and Ca(OH)2 at concentrations ranging between 1.0 and 8.0 M.50 Most studies have shown that acidic pretreatment is more effective.51 Physical and chemical pretreatments and their combinations generally require high-energy or corrosion-resistant high-pressure reactors. These traditional pretreatment methods produce various inhibitors, affecting enzymatic hydrolysis and yield. Biological pretreatment is low energy and low cost, involves no chemicals, generates no pollutants, and involves easy genetic manipulation.44 Biological pretreatment involves the lysis of microalgal cells with enzymes or microorganisms.36 Because this process can be carried out under mild conditions, the power requirements are low. The application of a variety of enzymes is necessary to achieve efficient cell disruption and starch hydrolysis.52,53 The enzymes used mainly include cellulase, hemicellulase, pectinase, protease, and amylase.36 Chlorella biomass co-pretreated with cellulase and macerozyme yielded 2350 mg L−1 of hydrogen; when sonicated yeast was used as a source of hydrolases, the hydrogen production efficiency was 4200 mL L−1.54
Table 1.
Comparison of different hydrogen production methods
Specific method | Time | Effect | Advantage | Disadvantage | Reference | |
---|---|---|---|---|---|---|
Hydrothermal pretreatment | Hydrothermal heating in the reactor, 80°C–160°C | 20-60 min | The chemical bonds between the cell walls are broken at high temperatures, thermal deformation occurs, and cellulose is decomposed | High decomposition efficiency Short decomposition time Simple operation Easy to control |
Monomer sugars degrade into inhibitory by-products High energy consumption | Chozhavendhan et al., Nagarajan et al.43,52 |
Grinding | Biomass is physically collided with ceramic, glass, or quartz beads under high agitation, comminution to fine size | A few minutes | Shear forces break down biomass, making it biodegradable | Does not contaminate products that require pretreatment Produces fewer inhibitory by-products | Unable to decompose macromolecular substances | Chozhavendhan et al., Shanmugam et al.52,55 |
Ultrasonic | Immersion of microalgal suspension in ultrasonic bath, destruction of the cell wall with ultrasound from 10 kHz to 20 MHz | Minutes to hours | Destroys algal cells and speeds up the hydrolysis process | High efficiency No other chemical reagents required | Heat generated by ultrasound needs to be cooled | Shanmugam et al. Snehya et al.55,56 |
Microwave irradiation | Electromagnetic waves in the frequency ranging 0.3–300 GHz. | Minutes to hours | Improves the solubility of cellular organic matter through thermal and non-thermal effects | Fast heating Short radiation time Non-contact heating Minimum space requirements | High energy consumption Unsuitable for scale expansion | Dinesh Kumar et al., Nagarajan et al., Yin et al.43,57,58 |
Acidolysis pretreatment | Pretreatment with inorganic acids, such as HCl, H2SO4 and HNO3 with concentrations ranging from 0.1 to 6.0 M | 20–90 min | Changes in pH changes cell membranes, causing cell rupture and dissolution of organic matter | Easy to enlarge Low energy consumption Simple to operate | Discharge of waste pollutes the environment Requires a lot of reagents Corrosion of the reactor | Chozhavendhan et al., Yin et al.31,52 |
Enzyme pretreatment | Utilization of purified or crude enzymes hydrolyze microalgal cells | Hours to a few days | Destroys the structure of cell walls and dissolves carbohydrates | Mild conditions Low energy consumption Less inhibitor production | High cost Enzyme is difficult to recover Enzyme is easily inactivated | Sriyod et al., Zabed et al.44,59 |
Fungal pretreatment | Exposure of biomass to fungi using liquid or solid culture techniques | weeks to months | Produces hydrolases to hydrolyze cellulose and lignin | Mild conditions Simple equipment Low cost Low energy consumption | Long pretreatment time Loss of carbohydrates | Barati et al., Zabed et al.44,47 |
Bacterial pretreatment | Inoculate bacteria into biomass | Several days | Bacteria continuously produce hydrolytic enzymes to hydrolyze cellulose in biomass | Short incubation time Easier to carry out genetic engineering transformation than fungi Higher adaptability More suitable for algae | Acid produced in the hydrolysis process changes the pH value | Barati et al.47 |
Conventional strategies and general requirements for biohydrogen production
Hydrogen production using microalgae is mainly done under ambient temperature and pressure with lower energy consumption than other methods.60 For decades, microalgal hydrogen production strategies have made great progress toward increasing hydrogen production.61,62 Typical methods of biohydrogen production include the direct or indirect decomposition of water into hydrogen using light energy, photo-fermentation, and anaerobic dark fermentation.44,63 Methods of producing biohydrogen using microalgae are shown in Table 2. Nowadays, biohydrogen production methods are more diversified and efficient.64 Hydrogen production from algal biomass remains at the laboratory scale, however, and is currently rarely used in large-scale commercial production.25 Commercial production of microalgal biohydrogen requires multi-stage optimization to achieve low-cost sustainable operations and needs to be carried out in bioreactors.65 Temperature primarily determines the accumulation of hydrogen, controlling the efficiency of the enzymatic reactions, and pH also plays a role.60,66 The optimal pH for different strains of microalgae varies but, generally, at pHs under 5, the activity of hydrogenase is inhibited.34 The most important component in photolysis and photofermentation is light.34 Bioreactors were used to investigate the effect of light intensity on the continuous production of hydrogen, and different light intensities led to different hydrogen yields.67,68 The response surface method optimizes several key factors of algal photobioreactor hydrogen production, and designs and builds pilot-scale reactors to produce biofuels.69
Table 2.
Comparison of different hydrogen production methods
Organism | Advantage | Disadvantage | Reference | |
---|---|---|---|---|
Photolysis | Green algae | Using light as energy The amount of solar energy fixed is ten times that of trees and crops Pure H2 produced |
Hydrogen production is inhibited by oxygen Short hydrogen production time High-intensity light is required | Bechara et al.70 |
Photofermentation | Chlorella, Scenedesmus obliquus | Spectrum range that can be used is wide Substrate is completely converted to H2 and CO2 Low energy required compared with photolysis Industrial waste can be used as the substrate | Need for light energy Pretreatment of substrate is required | Anwar et al., Liu et al.39,71 |
Dark fermentation | Lyngbya limnetica, Scenedesmus | Low price Does not depend on light Utilizes multiple carbon sources Utilizes waste Produces by-product acids |
Produces mixed gas Relatively low H2 production | Srivastava et al.72 |
Combination of light fermentation and dark fermentation | Microalgae and bacteria | Highest hydrogen production | Difficult to control the reactor conditions | Pandey et al.73 |
Direct photolysis of water
Biological photolysis can be categorized as either direct or indirect. Biohydrogen production, as a complex biochemical process, is affected by light, heat, reaction liquid composition, and mass transfer characteristics.39
Green algae (e.g., Chlamydomonas reinhardtii) exist in the ocean and humid soil and are photosynthetic autotrophs that have the same PS I and PS II systems as plants.74 Using the hydrogenase PS Ⅱ photosystem on the thylakoid membrane, microalgal cells decompose water into oxygen and hydrogen using solar energy75; the hydrogen ion is then converted into hydrogen by hydrogenases.66 Overall, the total reaction is as follows:
2H2O + light energy→ O2 + 2H2 | (Equation 1) |
Hydrogenase activity can be inhibited by 2% oxygen34; therefore, in the natural state, the reaction time for hydrogen production is reduced by oxygen inhibition, and efficiency becomes very low.39 To combat this, genetic engineering or the utilization of antioxidants has been employed to improve hydrogen yield. Particularly popular strategies are the use of stable gene expression and genetic evolution.76 Hydrogen production pathways from direct photolysis, indirect photolysis, and photo-fermentation are described in Figure 1.
Figure 1.
Overview of hydrogen production by microalgal chloroplasts
(A) Direct photolysis process of microalgae.
(B) Indirect photolysis process of microalgae.
(C) Photo fermentation process of microalgae. Figure redrew from reference (A–C).74,77,78
Indirect photolysis
In indirect photolysis, electrons produced by water decomposition are not directly used for hydrogen generation, as hydrogen production is temporarily separated from photosynthesis. Electrons are first used to synthesize carbohydrates and oxygen, after which the carbohydrates are catabolized to produce hydrogen.79 Indirect photolysis consists of two steps: carbon dioxide fixation by photosynthesis and carbohydrate decomposition.39 The reaction steps are as follows:
12H2O + 6CO2 + light →C6H12O6 + 6O2 + 6H2O | (Equation 2) |
C6H12O6 + 12H2O → 12H2 + 6CO2 + 6H2O | (Equation 3) |
Compared with green microalgae, indirect hydrogen production by cyanobacteria is more attractive.80 Indirect photolysis is observed in some nitrogen-fixing and non-nitrogen-fixing cyanobacteria, utilizing nitrogenase and hydrogenase enzymes.76 The problems surrounding the oxygen sensitivities of hydrogenase and nitrogenase are solved by temporal and spatial separation of the production of hydrogen.81 Non-nitrogen-fixing cyanobacteria such as Synechocystis, Synechococcus, and Myxococcus produce a bidirectional Ni-Fe hydrogenase.81 Generally, this enzyme exists in the cytoplasm loosely bound to thylakoids and acts effectively under low hydrogen partial pressure.82,79 These cyanobacteria carry out photosynthesis and carbon dioxide fixation under light, following which carbohydrates are fermented to produce hydrogen without light exposure.79 There are special structures known as ‘heterocysts’ in nitrogen-fixing cyanobacteria such as Chlorella and Anabaena.80 Compared with hydrogenase, nitrogenase catalyzes an irreversible process and is mainly responsible for reducing nitrogen in the atmosphere to ammonia accompanied by the production of hydrogen.83 Heterocysts spatially separate photosynthesis from the hydrogen production catalyzed by nitrogenase, keeping the enzyme from contacting oxygen to maintain its activity.34,84 Nitrogen-fixing cyanobacteria also have Ni-Fe hydrogenases, which catalyze the oxidation of hydrogen to effectively recover the electrons released during nitrogen fixation.79 Bidirectional hydrogenases are more efficient than nitrogenases, however, because it does not need ATP to generate hydrogen.79
Hydrogen production by photofermentation
Photofermentation degrades organic substrates into small molecules in the presence of light39:
C6H12O6 + 6H2O + light energy = 6CO2 + 12H2 | (Equation 4) |
Photofermentation is different from photolysis, which requires the reduction of substrates.39 Nitrogenase is the main enzyme in microalgae photofermentation, which obtains electrons through photo-catabolism of organic substrates and heterogeneous fermentation.85,86 Photofermentation provides electrons through the citric acid cycle and produces hydrogen under the catalysis of nitrogenase and hydrogenase9,87; in addition, it requires a large amount of ATP to promote electron transport and hydrogen production.39 The hydrogen produced by this method is relatively pure, however, so the need for purification of the generated gas is reduced.88 The fermentation efficiency is affected by light availability, light intensity, carbon source, and microbial fermentation ability.89 The main disadvantage of photo-fermentation is the inhibition of nitrogenase by oxygen, which requires cells to separate oxygen production and nitrogen fixation; another solution for this is to run the fermentation in the absence of nitrogen.75
Hydrogen production by dark fermentation
Dark fermentation is a light-independent heterotrophic hydrogen production pathway (Figure 2).90 Biomass with high carbon and water content such as sewage, food waste, and agricultural residues can be used as raw materials.91
Figure 2.
Hydrogen production by dark fermentation of microalgae
The three main dark fermentation processes of microalgae to produce hydrogen are shown. Formic acid catabolism, NADH re-oxidation, and pyruvate ferredoxin oxidoreductase (PFR) oxidation of pyruvate are the three main ways in which microalgae dark fermentation produces hydrogen. Figure redrew from ref.74,92,93,94,95
In the dark, microalgae obtain energy through the oxidation and reduction of organic compounds.96 Carbohydrates are first hydrolyzed to glucose, which is then converted with NAD+ into pyruvic acid and NADH through glycolysis; the pyruvate is then converted into acetyl coenzyme to produce formic acid and NADH.97 There are three processes for producing biohydrogen through dark fermentation: formic acid catabolism, NADH re-oxidation, and pyruvate ferredoxin oxidoreductase (PFR) oxidation of pyruvate.90 The redox reaction in microalgal cells also needs to be stabilized through the production of reducing compounds such as ethanol, butanol, and lactic acid.86 Slow hydrolysis is the rate-limiting step in the production of hydrogen by this method, and thus, pretreatment of biomass is required.96
The primary problem facing industrial hydrogen production is insufficient hydrogen production; the most fundamental reasons for this are the limitations of microalgal biology.86 Hydrogen production from dark fermentation can be optimized through the enhancement of biological activity and improvement of electron transfer during microbial fermentation.98 The application of genetic engineering, metabolic engineering, molecular biology, and other disciplines is very important to achieve these ends.
Genetic engineering for microalgae fermentation as substrates to produce hydrogen
The proportion of carbohydrates in microalgal organic matter determines its applicability for hydrogen production.97 Microalgae are photosynthetic autotrophs, and carbohydrate accumulation occurs through photosynthesis; carbohydrates manufactured by photosynthesis can be stored as subunits of cell walls or plastids as starch and cellulose.55 Ideal substrate strains exhibit high biomass productivity and carbohydrate content.99 Overexpression of carbohydrate synthetase, knockout of carbohydrate suppressor genes, and knock-in of carbohydrate synthesis genes can enhance intracellular carbohydrate productivity.61
Enhancement of microalgal carbohydrate content using genetic engineering
Increasing microalgal carbon fixation capacity is very important for increasing biomass.100 Overexpression of enzymes involved in carbohydrate synthesis can increase the carbohydrate content inside microalgae cells. A sucrose-phosphate synthase (SPS) encoding-gene was overexpressed in Anabaena sp. PCC 7120, and the resulting strains over-accumulated sucrose up to 10% (w/w) on a dry biomass basis under NaCl stress.101 Ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) is the most important enzyme for carbon fixation in photosynthetic organisms, and the rate of carbon fixation is determined by its catalytic ability102; in the air, however, it has only 25% of its catalytic capacity.100 Enhancing Rubisco activity has been shown to improve carbon fixation. An enzyme with ATPase activity, Rubisco Activase (RCA) uses ATP hydrolysis to remove various inhibitors at the Rubisco catalytic site (tightly bound sugar phosphates). Vectors were constructed that overexpressed putative RCA genes revealed through genome-wide analysis in Nannochloropsis oceanica, and the resulting transgenic strains had increased Rubisco activity, photosynthesis rates, and biomass productivity.103 Fructose 1,6-bisphosphate aldolase helps determine carbon allocation in the Calvin cycle, where it catalyzes the reversible conversion of dihydroxyacetone phosphate (DHAP) and glyceraldehyde 3-phosphate to fructose 1,6-bisphosphate (FBP). Aldolase plays an important role in the branching point of DHAP metabolism and is a key intermediate in starch and sucrose biosynthesis. When the cyanobacterial fructose 1,6-bisphosphate aldolase gene tied to a plastid transit peptide was introduced into Chlorella vulgaris, aldolase overexpression induced higher Rubisco activation, resulting in a 1.2-fold increase in photosynthetic capacity.102
The atmospheric expansion efficiency of carbon dioxide is very low, and the entry of carbon dioxide into the carbon cycle is very important for carbon fixation. The carbon dioxide fixation capacity of microalgae is 10–50 times higher than that of terrestrial plants.104 Microalgae prefer to fix carbon dioxide as carbonate.105 Carbonic anhydrase (CA) is present in most living organisms and catalyzes the conversion of carbon dioxide and carbonat.78 A CA with high activity (MlCA) from Mesorhizobium loti was identified in the BRaunschweig ENzyme DAtabase (BRENDA); when overexpressed in Chlorella sorokiniana and C. vulgaris, accelerated carbon capture and fixation was observed, resulting in a lipid content 2.2 times higher than that of the wild type strains.106
Reduce the breakdown of starch or competition for lipid synthesis
Carbon fixation competes with carbohydrate and lipid synthesis, as glycerol-3-phosphate from glycolysis is the precursor for both.107 The most important step in starch biosynthesis is the activation of glucose, which is converted into nucleoside-diphosphate-glucose (ADPGlc) by ADP-glucose pyrophosphorylase (AGPase).108 Overexpression of AGPase in microalgae enhances starch biosynthesis.109
Plastidial lysophosphatidic acid acyltransferase (LPAAT) catalyzes the acylation of lysophosphatidic acid (LPA) to produce phosphatidic acid, the second step in the Kennedy pathway responsible for glycerolipid biosynthesis. Nitrogen (N) deficiency is the most widely used strategy to trigger the accumulation of stored metabolites in microalgae.110 Overexpression of LPAAT in Chlamydomonas reinhardtii strains increased lipid content under adequate N, whereas lipid synthesis decreased and starch content increased by 50% when N was depleted. The putative regulatory gene encoding N-acetyltransferase (GNAT19) was identified by transcriptomic analysis and cloned, and was up-regulated 11- to 12-fold under N depletion. Overexpression of GNAT19 significantly increased the starch content of microalgae, showing great potential for enhancing biohydrogen production.111 Many microalgae alter their metabolism under phosphorus (P) starvation, inducing biosynthesis of stored metabolites. Pi Starvation Response1 (PSR1) is a transcription factor that regulates carbon storage metabolism under P starvation. In C. reinhardtii mutants lacking PSR1, P starvation-induced lipid and starch accumulation was inhibited; overexpression led to increased starch content and decreased neutral lipid content, which correlated with higher expression of specific starch metabolism genes.112
In microalgae, starch degradation provides nutrients for cell metabolism and growth in the dark; starch also provides a carbon backbone for lipid biosynthesis under adverse conditions.111 A decline in starch levels is largely attributed to starch catabolism. Manipulation of metabolism through genetic or metabolic engineering to reduce starch degradation is essential to improve biohydrogen production.113 Knockout of starch-inhibiting genes (glucan-water dikinases) and amylases can inhibit its degradation, increasing the accumulation of intracellular starch.61 So far, little is known about the regulation of starch metabolic changes in microalgae, and future efforts need to identify metabolism-specific transcriptional regulators.112
Increasing photosynthetic efficiency
Microalgae convert sunlight into organic molecules and biomass through photosynthesis. In regards to carbon dioxide fixation efficiency, improving light utilization efficiency helps to reduce the culturing and biomass pretreatment costs.114 Generally, it is difficult to improve the photosynthetic efficiency in traditional crops, but it is much simpler to alter microalgae through genetic engineering.66,115 Genetic engineering including knockouts, replacements, or insertions to the photosynthetic system has been shown to improve biomass accumulation.114
Microalgal cells have a powerful light-harvesting complex (LHC).115 The LHC complex in C. reinhardtii captures light energy before transferring it to the photosystem116; the complex also adapts to fluctuations in light quality. The LHC can both capture photons and disperse additional light energy to provide light protection.117 The light utilization efficiency of microalgae is largely limited by light transmittance. In one study, the efficiency of light energy absorption by large-scale cultured microalgal cells in the upper layer of the culture was higher than the rate of photosynthesis, meaning that 95% of light energy was wasted; the shielded microalgae cells in the lower layer could not carry out photosynthesis.62 Genetic manipulations overcoming the limitations of LHC antenna size and light saturation have been shown to improve the carbon fixation and light conversion efficiency of photosynthesis.86 Microalgae adapt to light conditions by regulating LHC expression; the cytoplasmic RNA binding protein NAB1 binds to the mRNA of the LHC gene to prevent its translation; the absence of NAB1 should therefore increase LHC antenna size. To test this, a plasmid expressing a mutated NAB1 coding sequence (C-terminal cysteine residues at positions 181 and 226 were replaced by serine) was constructed; the mutations in these two amino acids resulted in a permanently active variant of NAB1 that reduced antenna size. A suitable NAB1 promoter could then be used to control the antenna size, adjusting according to growth stage and light fluctuation status, improving light transmission.118 Alternately, an LHC-deficient mutant was generated in C. reinhardtii by DNA insertional mutagenesis. The smaller LHC antenna reduced excessive light absorption, allowing light energy to reach the lower microalgae cells in the culture; this mutant showed higher photosynthetic productivity in large-scale culture.119,120 Theoretically, truncated LHC antenna sizes could increase photosynthetic solar conversion efficiency and productivity by up to three times in large-scale cultivation.
Genetic engineering for microalgal hydrogen production using their metabolism
Compared with fossil fuels, algae biofuels are not competitive in cost.61 To reduce the cost of hydrogen production, microalgal genetic engineering strategies have attracted more attention.121 Genetic engineering techniques such as foreign gene expression, RNA interference, gene silencing, and gene disruption have been carried out in many species to improve biohydrogen production.61,62; These strategies aim to either change the key enzymes directly or indirectly involved in algal hydrogen production or increase the electron flow to enzymes.61 Different genetic engineering approaches for improving H2 yield in microalgae are described in Figure 3.
Figure 3.
Different Genetic engineering approaches for improvement the H2 yield
Genetic engineering approaches for microalgae to improve biohydrogen are shown. The genetic engineering approaches shown here include enhancing carbohydrate content, increasing photosynthetic efficiency, altering hydrogenase or nitrogenase activity and optimizing electronic sources. Figure redrew from ref.122
Improving light capture efficiency
As has been previously discussed, reducing the size of LHC antennas to improve light energy utilization efficiency to increase hydrogen production has shown promise.114 In addition, a mutant of C. reinhardtii with a smaller light capture antenna system was constructed by transforming LHC translation inhibitors; here, the hydrogen yield of the mutant increased by 50% with heterotrophic growth with exogenous glucose.118
Reducing the size of chloroplast can also improve photosynthetic efficiency. The chlorophyll content per unit cell volume of a chloroplast in a C. reinhardtii mutant was decreased, leading to better light transmittance and higher solar energy conversion efficiency. The maximum hydrogen accumulation of the mutant in pure culture as well as in co-culture with Pseudomonas was 1.8–5.2 times and 2.7–3.1 times that of the wild-type, respectively.123
Altering hydrogenase and nitrogenase activity
Hydrogenase and nitrogenase are two key enzymes responsible for hydrogen production in microalgae; as these enzymes are easily inhibited by oxygen, genetic engineering has been attempted to improve their activities, including changing the structure of the enzyme, foreign gene expression, altering the absorption of oxygen, and reducing the consumption of hydrogen.40,98
Hydrogenase enzymes can be divided into Fe-Fe, Ni-Fe, and Fe types according to the metals present in the enzyme’s active site.73 The activity of Fe-Fe hydrogenase has been demonstrated to be 100 times higher than that of other hydrogenases.117,124 Of interest, transcription, translation, and enzyme activity of microalgal hydrogenases are controlled by oxygen.121 Screening for oxygen-tolerant hydrogenases created by random mutation is the most common method currently used to find improved enzymes in green algae40; however, this method is time-consuming and high-throughput.125,126 Overexpression of hydrogenase can also increase hydrogen production, and hydrogenases of different microbes exhibit different levels of activity.84 As an example, Clostridium acetobutylicum is capable of overexpressing various Fe-Fe hydrogenases, however, protein production is quite low; expression vectors from C. reinhardtii were successfully heterologously expressed in C. acetobutylicum leading to high hydrogenase content (1 mg protein per liter of cells).124
The inhibition of hydrogenases by oxygen is caused by the degradation of 2Fe and 4Fe-4S clusters when oxygen binds the enzyme active sites.126 To help circumvent this, the oxygen gas channel structure of microalgae can be reduced to prevent the entry of oxygen. Overexpression of hydrogenases with high O2 tolerance and H2 yield as well as site-directed mutagenesis to replace amino acid residues around the hydrogenase gas tunnel to restrict oxygen access to the active site have been tested; the resulting mutant had improved tolerance to oxygen and 30 times the hydrogen production of the wild-type strain.127 Hydrogenase activity can also be increased by increasing oxygen consumption; although a sulfur-deficient culture can solve this problem, the cost of this method is high.128 Under normal conditions intracellular oxidases cannot consume all the oxygen produced by photosynthesis. Decreasing cellular oxygen levels is a major area of research. Full-field fragments of pyruvate oxidase genes from E. coli (POX) and catalase from Synechococcus elongatus (CAT) to prevent toxicity of H2O2, a byproduct of POX, were cloned into vectors. Expression in C. reinhardtii, resulted in reduced oxygen content and increased hydrogen production in sealed culture; after 48 h, hydrogen production in low light was 100 times higher than that of the wild type.129 The advantage of this method is that it allows a reduction in oxygen concentration without impacting the photosynthetic rate, enabling the production of more hydrogen.45 Similarly, genetic manipulation of microalgae using genes from methanotrophs to increase oxygen tolerance and enhance hydrogen production may also be effective.129
When sulfur is deficient, hydrogen can be continuously produced. The lack of sulfur limits the synthesis of methionine and cysteine, inactivating photosystem II (PSII) and resulting in photosynthetic oxygen release at a lower rate than respiration.130 miRNAs are a group of small noncoding RNAs that are ubiquitous in various gene regulatory pathways in eukaryotes that are involved in post-transcriptional gene expression regulation in plants and animals.131 Changing miRNA expression may affect many metabolic processes.126 Under sulfur-deficient culture conditions, C. reinhardtii microRNA (miRNA) expression was found to be up-regulated, with expression changes being closely related to the level of sulfur deficiency.131 Artificial miRNA gene sequences were also constructed into a highly expressed blue light-inducible system to target a PSII reaction-center protein in C. reinhardtii; after blue light treatment, the PSII response center protein was down-regulated.132
Artificial miRNA could also potentially be used to control the expression of the oxygen-evolving enhancer (OEE2) gene in PSII to increase hydrogen production in C. reinhardtii. OEE2 is a potent PS II-related protein involved in photosynthetic oxygen precipitation, and previous studies have shown that OEE2 expression decreased significantly after sulfur deprivation39,27; A heat-shock inducible expression vector containing miRNA targeting OEE2 was constructed and successfully transformed into C. reinhardtii. After repeated heat induction of the same microalgae culture, hydrogen accumulation was doubled when compared to the control.133 Compared with sulfur-deficient culture conditions, genetic engineering is advantageous as culture media does not need to be changed, making this method more suitable for large-scale production.133
In addition to hydrogenases, cyanobacteria contain nitrogenase to fix nitrogen.134 Different from the reversible formation of hydrogen using protons, nitrogenase catalyzes the unidirectional production of hydrogen.135 Owing to the unidirectional nature of the catalytic reaction, it was previously thought that hydrogen could be produced more easily with nitrogenases compared to hydrogenases83; however, it has since been shown that the catalytic efficiency of nitrogenase is much lower. Nitrogenase requires a large number of electrons, a reducing agent, and ATP to produce large volumes of hydrogen117; therefore, genetic engineering is key to hydrogen production utilizing this enzyme.136
In nitrogen-fixing cyanobacteria, a Hup uptake hydrogenase reduces the net hydrogen production of nitrogen-fixing cyanobacteria.137 A promising method for improving biohydrogen production in nitrogen-fixing cyanobacteria in the presence of oxygen is to select strains with high nitrogenase activity and eliminate the activity of the Hup uptake hydrogenase.135 A mutant lacking Hup was created by chemical mutagenesis (Nostoc sp. PCC 7120)138; elimination of Hup activity was very effective in increasing hydrogen accumulation by the strain.139 Nostoc sp. PCC 7422 had the highest nitrogenase activity out of the strains examined in the study, and a mutant of this strain was constructed that contained a hupL gene insertional disruption (ΔhupL). The knock-out improved the hydrogen production rate in the optimal hydrogen production stage to 4–7 times higher than that of the wild-type strain.140
Hydrogen production can also be increased by reducing the fixed electron distribution to nitrogen by replacing nitrogen gas with argon, but the cost of this method is very high, making it unsuitable for large-scale production.135 Although the inactivation of uptake hydrogenase can increase nitrogenase activity, the high nitrogenase activity of the mutant in the previous study was only stable in the air for around 10 h. Citrate combined with the Fe-Mo cofactor of nitrogenase is important for efficient nitrogen fixation, but not necessary for hydrogen production. There are two homocitrate synthase genes (nifV1 and nifV2) in cyanobacteria139; using these, homocitrate synthase gene knock-out mutations were performed in the Hup mutant of Nostoc sp. PCC 7120 (ΔhupL). Compared with the ΔhupL mutant, the double mutant had higher hydrogen production capacity and nitrogenase activity. Changing homocitrate synthase activity may therefore be an effective strategy to improve the sustained hydrogen production capacity of nitrogen-fixing cyanobacteria and other microalgal strains.141
To achieve high expression of this gene, nitrogenase must be expressed in cyanobacteria lacking the nif gene cluster.142 The minimal nif gene cluster from Bacteroides was integrated into the genome of the non-nitrogen-fixing S. elongatus PCC 7942 for nitrogenase expression. This minimal nif gene cluster is divided into six segments, each of which has an independent promoter and terminator; this solves the problem of decreased gene expression levels caused by the long nif gene cluster. A bioelectrochemical nitrogen fixation (e-BNF) system was formed by coupling the biological and electrochemical systems of the recombinant strain; methyl viologen (MV) provided the reducing power for nitrogenase, and PSII inhibitor 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) was used to inhibit photosynthetic oxygen production. The recombined S. elongatus PCC 7942 was able to continuously carry out nitrogen fixation and hydrogen production; after 48 h, the nitrogen fixation efficiency was 21 times that of purely photosynthetic nitrogen fixation systems.143
Hemoglobin is an oxygen-binding protein that exists widely in nature. Leguminous hemoglobin (encoded by the lba gene) combines with oxygen in root nodules during nitrogen fixation in leguminous plants and plays an important role in nitrogen metabolism.144 The heterologous expression of hemoglobin could provide a solution for the inactivation of microalgal nitrogenase when oxygen is present145; however, few studies have examined recombinant lba expressed in transgenic algae. After optimizing codon bias, lba and hemH (ferrochelatase gene from Bradyrhizobium japonicum) were transformed into the chloroplasts of C. reinhardtii; the hydrogen yield of the codon-optimized transgenic C. reinhardtii was 22% higher than that of the non-optimized strain.146
Electronic supply optimization
There are two main ways microalgae produce hydrogen from light; PSII water decomposition and electron transfer to hydrogenase, or oxidation of organic substrates.147 The first pathway uses photosynthesis, which consists of a light-utilizing reaction and a reaction that does not use light. The light reaction obtains electrons from water decomposition that are transferred from PSII to ferredoxin (FDX) through the electron transport chain, resulting in the production of ATP and NAD(P)H during the dark reaction; carbon dioxide is therefore reduced and fixed to an organic form as starch or glycogen.148 Overall, during direct photolysis, FDX provides electrons for hydrogenase and H2 come from protons and electrons produced by PSII water decomposition. Indirect photolysis hydrogen production is the result of intracellular carbohydrate degradation.148
Insufficient electron supply is a barrier to H2 synthesis.149 The hydrogenase iron-sulfur cluster (FeS) with redox activity in the center of FDX plays a role in electron transfer in microalgal cells.150 Through photosynthesis, H+ is released into the thylakoid lumen to establish a proton gradient.151 Electrons are transferred from water to FDX through the electron transfer chain.152 Proton carrier enhancement combined with thylakoid proton power dissipation promotes the availability of protons and electrons to Fe-Fe hydrogenase, forming hydrogen.147 FDX acts as a switch between carbon fixation and hydrogen production, as well as other environment-related functions.149 Hydrogenase needs to compete with other metabolic processes for electrons153; to circumvent this, genetic manipulation can increase electron flow to hydrogenase.134
Naturally expressed hydrogenases exist as a defense mechanism against excess electrons in microalgae.154 FDX directly supplies electrons to hydrogenase, but FDX availability can be limited because of other competitive processes155,150; Reducing electron competition at the FDX level can improve the affinity of hydrogenase for FDX.156 In one study, FDX was fused with Fe-Fe hydrogenase and expressed in C. reinhardtii; the resulting mutation improved the electron competitiveness of the hydrogenase, and the photosynthetic hydrogen production rate was 4.5 times higher than that of the strain containing the natural enzyme.154 Other genes can also be modified to improve electron flow to hydrogenase. Ferredoxin NADP+ reductase (FNR) is an enzyme in microalgae that catalyzes the conversion of FDX to NADP+; however, under anaerobic conditions, the reaction catalyzed by this enzyme competes with that of hydrogenase. Using RNA interference (RNAi), double-stranded RNA targeting the FNR gene was introduced into C. reinhardtii to inhibit the expression of FNR; hydrogen production in the mutant was 2.5 times higher than that in the wild-type.157
The fixation of carbon dioxide is the main competitor to hydrogenase electron flow.158 Rubisco, as discussed earlier, is a key enzyme responsible for carbon dioxide fixation. Although overexpression of Rubisco greatly improved carbon fixation and biomass production, it also led to an increase in oxygen release, which negatively impacts hydrogenase activity.159 When Rubisco was inhibited, the corresponding hydrogenase activity increased.160 In Rubisco mutant Y67A of C. reinhardtii, the tyrosine residues at positions 67, 68, and 72 were replaced by alanine, weakening the enzyme’s stability; the resulting mutant had a high potential for hydrogen production under normal light without nutrition, with 10–15 times higher hydrogen production than that of the wild-type. At the beginning of the culture, Rubisco protein content increased before decreasing rapidly as the protein degraded. The degradation of Rubisco inhibited carbon fixation, resulting in an excessive reduction in photosynthetic electron transport; hydrogenase was then widely expressed to relieve the metabolic pressure in cells. With the same photosynthetic and respiratory rates, Y67A could reach anaerobic conditions under normal light without the deactivation of hydrogenase.158
ATP synthase is a multi-subunit complex that combines proton transfer and ATP synthesis. Although the synthesis of ATP is very important for carbon dioxide fixation, the demand for ATP decreases in the hydrogen generation stage.150 When the electron transport chain is disconnected from ATP generation, carbon fixation is reduced because of insufficient energy being available.149 Robertson et al. characterized a chloroplast mutant of C. reinhardtii that produced an epsilon truncated polypeptide because of a frameshift mutation; the epsilon subunit protein is necessary for ATP synthase, preventing proton leakage and aiding assembly. Owing to this mutation, the assembly and function of ATP synthase were impaired, leading to the inactivation of the main metabolic pathway competing with hydrogenase for electrons, causing more electron flow to be used for hydrogen production.161,162
The photosynthetic electron transport chain driven by PSII, cytochrome b6f (Cytb6f) complex, and photosystem I (PSI) is on the thylakoid membrane163; PSII and PSI are located in the stacked thylakoid membrane and stromal layer, respectively. The reason for the spatial separation of PSII and PSI is to prevent energy from being transmitted from PSII to PSI because of the close contact of their antenna systems. In anaerobic microalgal culture, the protein on PSII attaches to PSI, sacrificing PSII to increase light absorption capacity. When an increase in PSI cyclic electron flow (CEF) compared with PSII linear electron flow (LEF), the proton gradient produced by CEF was found to limit the electron supply of hydrogenase leading to an increase in ATP synthesis and a decrease in hydrogen production.151,155 Therefore, inhibition of CEF can improve photosynthetic hydrogen production efficiency.151 Random insertion of resistance genes to produce PRGL1 knockout mutants (pgrl1-ko) of C. reinhardtii resulted in a lack of photosynthetic CEF in the mutant; electron flow then went to hydrogenase, acting as a safety mechanism for cellular health. Under higher light intensity, the hydrogen production rate of the mutant remained high, whereas that of their wild-type counterparts decreased rapidly.163 The same findings were also found using a mutant strain of C. reinhardtii where proton gradient regulatory protein (PGR5) was affected; the mutant lacked CEF and had high respiratory activity, breaking through the two bottlenecks of hydrogen production. The mutant was able to prolong hydrogen production for 12 days under mixed nutrition conditions and was an ideal strain for hydrogen production.164
Metabolic engineering regulation
Genome editing enables metabolic engineering to reconstruct metabolic pathways to produce biological products and has been used to build highly efficient microbial factories.165 In recent years, great progress has been made in the production of bio-based chemicals with the help of advanced tools and strategies for systems metabolic engineering.166
Gene manipulation of hydrogenase can improve enzyme activity, and is suitable for use with both light and dark fermentation. Biohydrogen production by microalgae is transient under natural conditions and is mainly controlled by the cell’s metabolic network.167 Metabolic engineering has great potential for increasing the production of microbial biological hydrogen.94 Microalgal metabolic networks are highly complex, nonlinear, and closely coordinated; microalgal cells carry out various metabolic processes organized to meet the needs caused by their environment. Cellular metabolism is first applied to growth rather than hydrogen production, but with the help of metabolic engineering, the hydrogen production pathway can be systematically and quantitatively improved.94,168 Microalgae have been transformed to eliminate bottlenecks, increase substrate utilization, and alternate substrate metabolic pathways can be introduced into the metabolic network. The modified microalgae have all been shown to continuously produce hydrogen in a reactor environment.10 Some metabolic engineering approaches have also achieved good results by emphasizing hydrogen production by blocking other metabolic processes. A wide range of combined strategies have been shown to significantly increase hydrogen, though combinatorial approaches have been primarily confirmed with hydrogen production in Escherichia coli.94 The application of advanced metabolic engineering in microalgae is expected to remove the limitation of low metabolic efficiency. These engineering approaches include providing energy to break thermodynamic barriers, expressing heterologous proteins, achieving more complete substrate conversions, and increasing proton reduction electron flux.169
The natural metabolic pathways of microalgae are diverse. Dark fermentation in particular has a low energy demand and is simple to run.20,170 In industrial production, dark fermentation is easier to control and has shown higher industrial stability and feasibility.171 The primary problem with this approach is that only one-third of the substrate can be used for hydrogen production, as the remaining two-thirds are used to produce additional fermentation products.172 The nitrate assimilation pathway in unicellular cyanobacteria (Synechocystis sp.) has the potential to compete with hydrogenase for electrons in dark fermentation. To improve hydrogen production, Baebprasert et al. deleted the nitrate assimilation pathway; the hydrogen production of the mutant strain was 140 times higher than that of the wild-type.173
The hexose transporter is a membrane protein that transports glucose into the cytoplasm. The plasma membrane of C. reinhardtii lacks hexose transporter proteins, but some microalgae can obtain nutrition through the overexpression of some hexose transporters.174 Expression of Chlorella HUP1 hexose transporter in C. reinhardtii allowed for carbohydrates from external sources to provide protons and electrons to hydrogenase, improving hydrogen production.175 Glucose transporters 1 (GLUT1) are a group of human uniporter protein, which facilitates bidirectional glucose transport. The GLUT1 on the human erythrocyte membrane is superior to that found in plants or algae; when introduced into C. reinhardtii, it led to very high glucose transport efficiency.174
With the development of a microalgal metabolism model (AlgaGEM) based on C. reinhardtii, it has become simpler to predict the effect of known mutations on hydrogen production. Using this metabolic model, the physiological pathway for hydrogen production under dark conditions was predicted, showing that when the circulating electron flow is prevented, hydrogen production will increase. The impact of this result on increasing hydrogen production is consistent with the results of studies in the literature.176
Economic and technical feasibility
The commercialization of biohydrogen faces severe economic and technical challenges. The production cost is mainly determined by the bioreactor and the selected process.177 In microalgal large-scale cultivation, open pond systems are used more frequently as they are lower cost, however, the high water quality requirements increase operating costs. Cultivating microalgae for hydrogen production in 140 ha of open ponds or 14 ha of photobioreactors has been reported to cost US$6 m−2 and US$100 m−2, respectively, with annual operating costs of US$43 million and US$10 million; the cost of producing biohydrogen was calculated to be US$10 GJ−1, much higher than that of gasoline (US$2.5 GJ−1).86,178 Unfortunately, this analysis was too simplistic, as it did not include the cost of harvesting, purifying, and transporting hydrogen. With the current hydrogen production process, biohydrogen from microalgae is rather costly. An economic analysis report shows that the cost of hydrogen production by photolysis is between US$1.42 kg−1 and US$7.24 kg−1, and the cost of hydrogen production by fermentation is between US$7.54 kg−1 and US$7.61 kg−186 In addition, cultivation and reactor designs play a decisive role in the overall cost; overall, the average cost of biohydrogen production from algae is no less than US$10 GJ−1 (1.3kg−1) under current conditions.86,178
For biohydrogen to be used as fuel, purification is required. Purification using a two-stage carbon film system can produce >99.5 vol % of high-purity biohydrogen at US$0.67/kg, which is more competitive with the state-of-the-art hydrogen purification technologies of pressure swing adsorption and cryogenic distillation.179 With the expected future increase in hydrogen demand, hydrogen purification will develop scalable production at industrial capacities.
Cost-effective biohydrogen requires low operational cost and high yield; the main limitation to commercialization is production cost. Although progress has been made, improving production efficiency using genetic or metabolic engineering remains challenging.117 Genetic engineering can also be applied to other aspects of microalgal commercialization to reduce harvest costs, improve bioflocculant production, and allow biofuel production from wastewater180; expression of limonene synthase in microalgae has been shown to increase cell hydrophobicity and drive cell aggregation for sedimentation and harvest.181 Future research should focus on the use of high-throughput genome-wide analysis in combination with the design of efficient bioreactors.122 This requires the joint efforts of scientists to make fully establish biohydrogen as a future energy source.19
Conclusion
To combat resource depletion and environmental concerns, microalgal biohydrogen must be commercialized. The application of genetic and metabolic engineering along with synthetic biology has been shown to make biohydrogen production more cost-effective. With the understanding of algal genome sequences and metabolic functions, as well as the development of advanced tools and software, the combination of genetic engineering and bioinformatics will no doubt help to develop microalgae strains that are more suitable for biohydrogen production than those available currently. Although measures have been taken to improve the competitiveness of biohydrogen production, the infant stage of this technology has not been applied to large-scale hydrogen production. A tremendous amount of effort is still required before this technology could occupy a large market share.
Future directions and challenges
As a sustainable energy source, microalgal biohydrogen has attracted increasing international attention. It can be produced in two ways, with the most common being hydrogen fermentation using microalgal biomass as raw material as a sustainable and robust feedstock for large-scale energy production. Genetic and metabolic engineering can be used to improve the efficiency of photosynthesis, improve carbon sequestration by driving carbon fluxes into energy-rich compounds that can be used as hydrogen energy sources, and develop robust microalgae strains to allow for low-cost large-scale cultivation to reduce operating costs.115,182
Hydrogen can also be produced from microalgae through metabolism. There are two main processes involved in the production of biohydrogen; increasing biomass, and photolysis, photofermentation, or dark fermentation under anaerobic conditions.183 Improving the efficiency of photosynthesis, rapidly accumulating carbohydrates, and maintaining cellular stability are important for the efficient use of water and organic matter. In addition, it is necessary to increase the activity of enzymes associated with hydrogen production through genetic manipulation, and reconstruct metabolic pathways through metabolic engineering to establish more efficient production strains.121,165 Currently, there are bottlenecks in the large-scale production of biohydrogen, as production costs depend not only on the microorganisms used and metabolic pathways present84; it is also necessary to improve the technology used for microalgae biomass cultivation and harvesting. Adjustments to culture conditions may have a large impact on the biomass and carbohydrate contents, and optimal production can be achieved by a proper combination of genetic manipulation and culture condition engineering.115
Gene editing technology can overcome many obstacles, but the stability and safety of genetically-modified microalgae are also primary factors to consider, requiring further evaluation to ensure that they do not pose future risks to human and environmental health.115 Another primary challenge in facing metabolic engineering is the limited number of available genes and the complex genotype-phenotypic relationships resulting from their combination with host organism metabolic pathways.165 For downstream biohydrogen production, it is necessary to weaken the demand for light in downstream processes and consider more energy-efficient dark fermentation strategies. Biological hydrogen production also does not produce only one gas product, impacting hydrogen purity. Low-cost gas separation technologies need to be developed to remove by-products such as carbon dioxide.184 For continuous production, periodicity is required, and the integration of steps in the production process is crucial. It is necessary to specifically understand the impact of each variable in each step on hydrogen production to achieve the highest possible hydrogen production.108
Recommendations
Application of genetic engineering in the field of biofuels is expanding rapidly. Genetic engineering tools solve many problems, and provide industrially capable microalgae strains. A systematic review of biofuel production in genetically modified (GM) algae between 2008 and 2019 found that one of the areas that received the least attention was environmental risk.185 There are differences in biotechnology laws and regulations between countries; because of the lack of internationally harmonized regulations, the commercialization of genetically modified biofuels cannot be effectively applied. The use of genetic engineering can provide high yields, high growth rates, and insect resistance, but the impact of released modified plasmid or chromosomal DNA into surrounding water bodies has been overlooked.186 Farming without a comprehensive risk assessment of GM strains can pose a serious threat to human and environmental health. More work should be done in regards to safety testing of GM organisms to assess the potential risks posed to the environment from the growth and processing of GM microalgae.
Acknowledgments
This work was funded by the National Natural Science Foundation of China (32070107). We would like to thank MogoEdit (https://www.mogoedit.com) for its English editing during the preparation of this manuscript.
Author contributions
J.Z.: Conceptualization, Methodology, Software, Data curation and processing, Validation, Writing–original draft. D.X.: Data curation, Investigation. C.W.: Validation, Writing–review and editing. D.F.: Software, Validation. L.C.: Methodology, Formal analysis. C.G.: Conceptualization, Methodology, Validation, Writing–review and editing, Supervision, Funding.
Declaration of interests
The authors declare no competing interests.
Inclusion and diversity
We support inclusive, diverse, and equitable conduct of research.
References
- 1.Kalair A., Abas N., Saleem M.S., Kalair A.R., Khan N. Role of energy storage systems in energy transition from fossil fuels to renewables. Energy Storage. 2021;3:e135. doi: 10.1002/est2.135. [DOI] [Google Scholar]
- 2.Gong C., Cao L., Fang D., Zhang J., Kumar Awasthi M., Xue D. Genetic manipulation strategies for ethanol production from bioconversion of lignocellulose waste. Bioresour. Technol. 2022;352 doi: 10.1016/j.biortech.2022.127105. [DOI] [PubMed] [Google Scholar]
- 3.Warner K.J., Jones G.A. The 21st century coal question: China, India, development, and climate change. Atmosphere. 2019;10:476. doi: 10.3390/atmos10080476. [DOI] [Google Scholar]
- 4.Ghahari S., Ghahari S., Ghahari S., Nematzadeh G.A., Sarma H. Springer; 2021. Environmental Biotechnology: Toward a Sustainable Future. Biotechnology for Sustainable Environment; pp. 1–31. [DOI] [Google Scholar]
- 5.Johnsson F., Kjärstad J., Rootzén J. The threat to climate change mitigation posed by the abundance of fossil fuels. Clim. Pol. 2019;19:258–274. doi: 10.1080/14693062.2018.1483885. [DOI] [Google Scholar]
- 6.Gani A. Fossil fuel energy and environmental performance in an extended STIRPAT model. J. Clean. Prod. 2021;297 doi: 10.1016/j.jclepro.2021.126526. [DOI] [Google Scholar]
- 7.Hanna E., Hall R.J. Science, Faith and the Climate Crisis. Emerald Publishing Limited); 2020. Earth, air, fire and ice: exploring links between human-induced global warming, polar ice melt and local scale extreme weather; pp. 47–64. [DOI] [Google Scholar]
- 8.Lelieveld J., Klingmüller K., Pozzer A., Burnett R.T., Haines A., Ramanathan V. Effects of fossil fuel and total anthropogenic emission removal on public health and climate. Proc. Natl. Acad. Sci. USA. 2019;116:7192–7197. doi: 10.1073/pnas.1819989116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Gong C., Singh A., Singh P., Singh A. Anaerobic digestion of agri-food wastes for generating biofuels. Indian J. Microbiol. 2021;61:427–440. doi: 10.1007/s12088-021-00977-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Singh V., Das D. Science and engineering of hydrogen-based energy technologies. Elsevier; 2019. Potential of hydrogen production from biomass; pp. 123–164. [DOI] [Google Scholar]
- 11.Xu Y., Deng Y., Liu W., Zhao X., Xu J., Yuan Z. Research progress of hydrogen energy and metal hydrogen storage materials. Sustain. Energy Technol. Assessments. 2023;55 doi: 10.1016/j.seta.2022.102974. [DOI] [Google Scholar]
- 12.Li Z., Yue Y., Peng J., Luo Z. Phase engineering two-dimensional nanostructures for electrocatalytic hydrogen evolution reaction. Chin. Chem. Lett. 2023;34 doi: 10.1016/j.cclet.2022.01.012. [DOI] [Google Scholar]
- 13.Shah S.A.A. Feasibility study of renewable energy sources for developing the hydrogen economy in Pakistan. Int. J. Hydrogen Energy. 2020;45:15841–15854. doi: 10.1016/j.ijhydene.2019.09.153. [DOI] [Google Scholar]
- 14.Milani D., Kiani A., McNaughton R. Renewable-powered hydrogen economy from Australia's perspective. Int. J. Hydrogen Energy. 2020;45:24125–24145. doi: 10.1016/j.ijhydene.2020.06.041. [DOI] [Google Scholar]
- 15.Moravvej Z., Soroush E., Makarem M.A., Rahimpour M.R. Advances in bioenergy and microfluidic applications. Elsevier; 2021. Thermochemical routes for hydrogen production from biomass; pp. 193–208. [DOI] [Google Scholar]
- 16.Lee J.E., Shafiq I., Hussain M., Lam S.S., Rhee G.H., Park Y.K. A review on integrated thermochemical hydrogen production from water. Int. J. Hydrogen Energy. 2022;47:4346–4356. doi: 10.1016/j.ijhydene.2021.11.065. [DOI] [Google Scholar]
- 17.Dou B., Zhang H., Song Y., Zhao L., Jiang B., He M., Ruan C., Chen H., Xu Y. Hydrogen production from the thermochemical conversion of biomass: issues and challenges. Sustain. Energy Fuels. 2019;3:314–342. doi: 10.1039/c8se00535d. [DOI] [Google Scholar]
- 18.Yang G., Wang J. Changes in microbial community structure during dark fermentative hydrogen production. Int. J. Hydrogen Energy. 2019;44:25542–25550. doi: 10.1016/j.ijhydene.2019.08.039. [DOI] [Google Scholar]
- 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. doi: 10.1016/j.ijhydene.2022.01.156. [DOI] [Google Scholar]
- 20.Mahidhara G., Burrow H., Sasikala C., Ramana C.V. Biological hydrogen production: Molecular and electrolytic perspectives. World J. Microbiol. Biotechnol. 2019;35:116–213. doi: 10.1007/s11274-019-2692-z. [DOI] [PubMed] [Google Scholar]
- 21.Zhang K., Kim W.J., Park A.H.A. Alkaline thermal treatment of seaweed for high-purity hydrogen production with carbon capture and storage potential. Nat. Commun. 2020;11:3783–3789. doi: 10.1038/s41467-020-17627-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Saad M.G., Dosoky N.S., Zoromba M.S., Shafik H.M. Algal biofuels: current status and key challenges. Energies. 2019;12:1920. doi: 10.3390/en12101920. [DOI] [Google Scholar]
- 23.Queiroz M.I., Vieira J.G., Maroneze M.M. Handbook of Microalgae-Based Processes and Products. Elsevier; 2020. Morphophysiological, structural, and metabolic aspects of microalgae; pp. 25–48. [DOI] [Google Scholar]
- 24.Yap J.K., Sankaran R., Chew K.W., Halimatul Munawaroh H.S., Ho S.H., Rajesh Banu J., Show P.L. Advancement of green technologies: A comprehensive review on the potential application of microalgae biomass. Chemosphere. 2021;281 doi: 10.1016/j.chemosphere.2021.130886. [DOI] [PubMed] [Google Scholar]
- 25.Vij R.K., Subramanian D., Pandian S., Krishna S., Hari S. A review of different technologies to produce fuel from microalgal feedstock. Environ. Technol. Innov. 2021;22 doi: 10.1016/j.eti.2021.101389. [DOI] [Google Scholar]
- 26.Xiao C., Yuan J., Li L., Zhong N., Zhong D., Xie Q., Chang H., Xu Y., He X., Li M. Photocatalytic synergistic biofilms enhance tetracycline degradation and conversion. Environ. Sci. Ecotechnol. 2023;14 doi: 10.1016/j.ese.2022.100234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Banerjee S., Banerjee S., Ghosh A.K., Das D. Maneuvering the genetic and metabolic pathway for improving biofuel production in algae: Present status and future prospective. Renew. Sustain. Energy Rev. 2020;133 doi: 10.1016/j.rser.2020.110155. [DOI] [Google Scholar]
- 28.Chi N.T.L., Anto S., Ahamed T.S., Kumar S.S., Shanmugam S., Samuel M.S., Mathimani T., Brindhadevi K., Pugazhendhi A. A review on biochar production techniques and biochar based catalyst for biofuel production from algae. Fuel. 2021;287 doi: 10.1016/j.fuel.2020.119411. [DOI] [Google Scholar]
- 29.Chauton M.S., Forbord S., Mäkinen S., Sarno A., Slizyte R., Mozuraityte R., Standal I.B., Skjermo J. Sustainable resource production for manufacturing bioactives from micro-and macroalgae: Examples from harvesting and cultivation in the Nordic region. Physiol. Plantarum. 2021;173:495–506. doi: 10.1111/ppl.13391. [DOI] [PubMed] [Google Scholar]
- 30.Kumar Sharma A., Kumar Ghodke P., Manna S., Chen W.H. Emerging technologies for sustainable production of biohydrogen production from microalgae: A state-of-the-art review of upstream and downstream processes. Bioresour. Technol. 2021;342 doi: 10.1016/j.biortech.2021.126057. [DOI] [PubMed] [Google Scholar]
- 31.Yin Y., Wang J. Hydrogen production and energy recovery from macroalgae Saccharina japonica by different pretreatment methods. Renew. Energy. 2019;141:1–8. doi: 10.1016/j.renene.2019.03.139. [DOI] [Google Scholar]
- 32.Kovač A., Paranos M., Marciuš D. Hydrogen in energy transition: A review. Int. J. Hydrogen Energy. 2021;46:10016–10035. doi: 10.1016/j.ijhydene.2020.11.256. [DOI] [Google Scholar]
- 33.Wang Y., Yang H., Zhang X., Han F., Tu W., Yang W. Microalgal hydrogen production. Small Methods. 2020;4 doi: 10.1002/smtd.201900514. [DOI] [Google Scholar]
- 34.Akhlaghi N., Najafpour-Darzi G. A comprehensive review on biological hydrogen production. Int. J. Hydrogen Energy. 2020;45:22492–22512. doi: 10.1016/j.ijhydene.2020.06.182. [DOI] [Google Scholar]
- 35.Ravichandran S.R., Venkatachalam C.D., Sengottian M., Sekar S., Kandasamy S., Ramasamy Subramanian K.P., Purushothaman K., Lavanya Chandrasekaran A., Narayanan M. A review on hydrothermal liquefaction of algal biomass on process parameters, purification and applications. Fuel. 2022;313 doi: 10.1016/j.fuel.2021.122679. [DOI] [Google Scholar]
- 36.Bhatia S.K., Rajesh Banu J., Singh V., Kumar G., Yang Y.H. Algal biomass to biohydrogen: pretreatment, influencing factors, and conversion strategies. Bioresour. Technol. 2023;368 doi: 10.1016/j.biortech.2022.128332. [DOI] [PubMed] [Google Scholar]
- 37.Goswami R.K., Agrawal K., Verma P. Micro-algae: Next-generation Feedstock for Biorefineries: Cultivation and Refining Processes. Springer; 2022. Microalgae Biomass Biorefinery: A Sustainable Renewable Energy Feedstock of the Future; pp. 1–29. [DOI] [Google Scholar]
- 38.Saravanan A., Senthil Kumar P., Khoo K.S., Show P.L., Femina Carolin C., Fetcia Jackulin C., Jeevanantham S., Karishma S., Show K.Y., Lee D.J., Chang J.S. Biohydrogen from organic wastes as a clean and environment-friendly energy source: Production pathways, feedstock types, and future prospects. Bioresour. Technol. 2021;342 doi: 10.1016/j.biortech.2021.126021. [DOI] [PubMed] [Google Scholar]
- 39.Anwar M., Lou S., Chen L., Li H., Hu Z. Recent advancement and strategy on bio-hydrogen production from photosynthetic microalgae. Bioresour. Technol. 2019;292 doi: 10.1016/j.biortech.2019.121972. [DOI] [PubMed] [Google Scholar]
- 40.Li S., Li F., Zhu X., Liao Q., Chang J.S., Ho S.H. Biohydrogen production from microalgae for environmental sustainability. Chemosphere. 2022;291 doi: 10.1016/j.chemosphere.2021.132717. [DOI] [PubMed] [Google Scholar]
- 41.Tang D.Y.Y., Khoo K.S., Chew K.W., Tao Y., Ho S.-H., Show P.L. Potential utilization of bioproducts from microalgae for the quality enhancement of natural products. Bioresour. Technol. 2020;304 doi: 10.1016/j.biortech.2020.122997. [DOI] [PubMed] [Google Scholar]
- 42.Koyande A.K., Chew K.W., Rambabu K., Tao Y., Chu D.T., Show P.L. Microalgae: A potential alternative to health supplementation for humans. Food Sci. Hum. Wellness. 2019;8:16–24. doi: 10.1016/j.fshw.2019.03.001. [DOI] [Google Scholar]
- 43.Nagarajan D., Chang J.S., Lee D.J. Pretreatment of microalgal biomass for efficient biohydrogen production–Recent insights and future perspectives. Bioresour. Technol. 2020;302 doi: 10.1016/j.biortech.2020.122871. [DOI] [PubMed] [Google Scholar]
- 44.Zabed H.M., Akter S., Yun J., Zhang G., Awad F.N., Qi X., Sahu J. Recent advances in biological pretreatment of microalgae and lignocellulosic biomass for biofuel production. Renew. Sustain. Energy Rev. 2019;105:105–128. doi: 10.1016/j.rser.2019.01.048. [DOI] [Google Scholar]
- 45.Griffiths G., Hossain A.K., Sharma V., Duraisamy G. Key Targets for Improving Algal Biofuel Production. Cleanroom Technol. 2021;3:711–742. doi: 10.3390/cleantechnol3040043. [DOI] [Google Scholar]
- 46.Wu Y., Xiang W., Li L., Liu H., Zhong N., Chang H., Rittmann B.E. A novel photoelectrochemical system to disrupt microalgae for maximizing lipid-extraction efficiency. Chem. Eng. J. 2021;420 doi: 10.1016/j.cej.2021.130517. [DOI] [Google Scholar]
- 47.Barati B., Zafar F.F., Rupani P.F., Wang S. Bacterial pretreatment of microalgae and the potential of novel nature hydrolytic sources. Environ. Technol. Innov. 2021;21 doi: 10.1016/j.eti.2021.101362. [DOI] [Google Scholar]
- 48.Wang J., Yin Y. Fermentative hydrogen production using various biomass-based materials as feedstock. Renew. Sustain. Energy Rev. 2018;92:284–306. doi: 10.1016/j.rser.2018.04.033. [DOI] [Google Scholar]
- 49.Zhang Y., Ding Z., Shahadat Hossain M., Maurya R., Yang Y., Singh V., Kumar D., Salama E.S., Sun X., Sindhu R., et al. Recent advances in lignocellulosic and algal biomass pretreatment and its biorefinery approaches for biochemicals and bioenergy conversion. Bioresour. Technol. 2023;367 doi: 10.1016/j.biortech.2022.128281. [DOI] [PubMed] [Google Scholar]
- 50.Singh H., Tomar S., Qureshi K.A., Jaremko M., Rai P.K. Recent advances in biomass pretreatment technologies for biohydrogen production. Energies. 2022;15:999. doi: 10.3390/en15030999. [DOI] [Google Scholar]
- 51.Ahmed S.F., Mofijur M., Nahrin M., Chowdhury S.N., Nuzhat S., Alherek M., Rafa N., Ong H.C., Nghiem L., Mahlia T. Biohydrogen production from wastewater-based microalgae: Progresses and challenges. Int. J. Hydrogen Energy. 2022;47:37321–37342. doi: 10.1016/j.ijhydene.2021.09.178. [DOI] [Google Scholar]
- 52.Chozhavendhan S., Rajamehala M., Karthigadevi G., Praveenkumar R., Bharathiraja B. A review on feedstock, pretreatment methods, influencing factors, production and purification processes of bio-hydrogen production. Case Stud. Chem. Environ. Eng. 2020;2 doi: 10.1016/j.cscee.2020.100038. [DOI] [Google Scholar]
- 53.Sriyod K., Reungsang A., Plangklang P. One-step multi enzyme pretreatment and biohydrogen production from Chlorella sp. biomass. Int. J. Hydrogen Energy. 2021;46:39675–39687. doi: 10.1016/j.ijhydene.2021.09.232. [DOI] [Google Scholar]
- 54.Abdel-Kader H.A., Abdel-Basset R., Danial A.W. Yeast and enzymatic hydrolysis in converting Chlorella biomass into hydrogen gas by Rhodobacter sp. and Rhodopseudomonas palustris. Int. J. Hydrogen Energy. 2022;47:1516–1528. doi: 10.1016/j.ijhydene.2021.10.126. [DOI] [Google Scholar]
- 55.Shanmugam S., Sekar M., Sivaramakrishnan R., Raj T., Ong E.S., Rabbani A.H., Rene E.R., Mathimani T., Brindhadevi K., Pugazhendhi A. Pretreatment of second and third generation feedstock for enhanced biohythane production: Challenges, recent trends and perspectives. Int. J. Hydrogen Energy. 2021;46:11252–11268. doi: 10.1016/j.ijhydene.2020.12.083. [DOI] [Google Scholar]
- 56.Snehya A., Sundaramahalingam M., Rajeshbanu J., Sivashanmugam P. Studies on evaluation of pretreatment efficiency of surfactant mediated ultrasonic pretreatment of Sargassum tennerimum (marine macroalgae) for achieving profitable biohydrogen production. Int. J. Hydrogen Energy. 2022;47:17184–17193. doi: 10.1016/j.ijhydene.2022.03.216. [DOI] [Google Scholar]
- 57.Dinesh Kumar M., Yukesh Kannah R., Kumar G., Sivashanmugam P., Rajesh Banu J. A novel energetically efficient combinative microwave pretreatment for achieving profitable hydrogen production from marine macro algae (Ulva reticulate) Bioresour. Technol. 2020;301 doi: 10.1016/j.biortech.2020.122759. [DOI] [PubMed] [Google Scholar]
- 58.Yin Y., Hu J., Wang J. Fermentative hydrogen production from macroalgae Laminaria japonica pretreated by microwave irradiation. Int. J. Hydrogen Energy. 2019;44:10398–10406. doi: 10.1016/j.ijhydene.2019.03.034. [DOI] [Google Scholar]
- 59.Sriyod K., Reungsang A., Plangklang P. Single and Combined Enzymatic Saccharification and Biohydrogen Production from Chlorella sp. Bioenergy Res. 2021;14:940–953. doi: 10.1007/s12155-020-10191-y. [DOI] [Google Scholar]
- 60.Thanigaivel S., Rajendran S., Hoang T.K.A., Ahmad A., Luque R. Photobiological effects of converting biomass into hydrogen-Challenges and prospects. Bioresour. Technol. 2023;367 doi: 10.1016/j.biortech.2022.128278. [DOI] [PubMed] [Google Scholar]
- 61.Chaturvedi V., Goswami R.K., Verma P. Biorefineries: a step towards renewable and clean energy. Springer; 2020. Genetic engineering for enhancement of biofuel production in microalgae; pp. 539–559. [DOI] [Google Scholar]
- 62.Fayyaz M., Chew K.W., Show P.L., Ling T.C., Ng I.S., Chang J.S. Genetic engineering of microalgae for enhanced biorefinery capabilities. Biotechnol. Adv. 2020;43 doi: 10.1016/j.biotechadv.2020.107554. [DOI] [PubMed] [Google Scholar]
- 63.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. doi: 10.1016/j.rser.2018.04.043. [DOI] [Google Scholar]
- 64.Cao L., Yu I.K.M., Xiong X., Tsang D.C.W., Zhang S., Clark J.H., Hu C., Ng Y.H., Shang J., Ok Y.S. Biorenewable hydrogen production through biomass gasification: A review and future prospects. Environ. Res. 2020;186 doi: 10.1016/j.envres.2020.109547. [DOI] [PubMed] [Google Scholar]
- 65.Jiménez-Llanos J., Ramírez-Carmona M., Rendón-Castrillón L., Ocampo-López C. Sustainable biohydrogen production by Chlorella sp. microalgae: A review. Int. J. Hydrogen Energy. 2020;45:8310–8328. doi: 10.1016/j.ijhydene.2020.01.059. [DOI] [Google Scholar]
- 66.Bolatkhan K., Kossalbayev B.D., Zayadan B.K., Tomo T., Veziroglu T.N., Allakhverdiev S.I. Hydrogen production from phototrophic microorganisms: reality and perspectives. Int. J. Hydrogen Energy. 2019;44:5799–5811. doi: 10.1016/j.ijhydene.2019.01.092. [DOI] [Google Scholar]
- 67.Nazarpour M., Taghizadeh-Alisaraei A., Asghari A., Abbaszadeh-Mayvan A., Tatari A. Optimization of biohydrogen production from microalgae by response surface methodology (RSM) Energy. 2022;253 doi: 10.1016/j.energy.2022.124059. [DOI] [Google Scholar]
- 68.Ruiz-Marin A., Canedo-López Y., Chávez-Fuentes P. Biohydrogen production by Chlorella vulgaris and Scenedesmus obliquus immobilized cultivated in artificial wastewater under different light quality. Amb. Express. 2020;10:1–7. doi: 10.1186/s13568-020-01129-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Jain I. Hydrogen the fuel for 21st century. Int. J. Hydrogen Energy. 2009;34:7368–7378. doi: 10.1016/j.ijhydene.2009.05.093. [DOI] [Google Scholar]
- 70.Bechara R., Azizi F., Boyadjian C. Process simulation and optimization for enhanced biophotolytic hydrogen production from green algae using the sulfur deprivation method. Int. J. Hydrogen Energy. 2021;46:14096–14108. doi: 10.1016/j.ijhydene.2021.01.115. [DOI] [Google Scholar]
- 71.Liu H., Zhang Z., Zhang H., Lee D.J., Zhang Q., Lu C., He C. Evaluation of hydrogen yield potential from Chlorella by photo-fermentation under diverse substrate concentration and enzyme loading. Bioresour. Technol. 2020;303 doi: 10.1016/j.biortech.2020.122956. [DOI] [PubMed] [Google Scholar]
- 72.Srivastava N., Hussain A., Kushwaha D., Haque S., Mishra P., Gupta V.K., Srivastava M. Nickel ferrite nanoparticles induced improved fungal cellulase production using residual algal biomass and subsequent hydrogen production following dark fermentation. Fuel. 2021;304 doi: 10.1016/j.fuel.2021.121391. [DOI] [Google Scholar]
- 73.Pandey A., Sinha P., Pandey A. Hydrogen production by sequential dark and photofermentation using wet biomass hydrolysate of Spirulina platensis: Response surface methodological approach. Int. J. Hydrogen Energy. 2021;46:7137–7146. doi: 10.1016/j.ijhydene.2020.11.205. [DOI] [Google Scholar]
- 74.Petrova E.V., Kukarskikh G.P., Krendeleva T.E., Antal T.K. The mechanisms and role of photosynthetic hydrogen production by green microalgae. Microbiol. 2020;89:251–265. doi: 10.1134/S0026261720030169. [DOI] [Google Scholar]
- 75.Singla S., Shetti N.P., Basu S., Mondal K., Aminabhavi T.M. Hydrogen production technologies-membrane based separation, storage and challenges. J. Environ. Manag. 2022;302 doi: 10.1016/j.jenvman.2021.113963. [DOI] [PubMed] [Google Scholar]
- 76.Javed M.A., Zafar A.M., Aly Hassan 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 doi: 10.1016/j.jece.2021.107003. [DOI] [Google Scholar]
- 77.Choi K.R., Jiao S., Lee S.Y. Metabolic engineering strategies toward production of biofuels. Curr. Opin. Chem. Biol. 2020;59:1–14. doi: 10.1016/j.cbpa.2020.02.009. [DOI] [PubMed] [Google Scholar]
- 78.Zhang S., Liu Z. Advances in the biological fixation of carbon dioxide by microalgae. J. Chem. Technol. Biotechnol. 2021;96:1475–1495. doi: 10.1002/jctb.6714. [DOI] [Google Scholar]
- 79.Singh H., Das D. Handbook of Microalgae-Based Processes and Products. Elsevier; 2020. Biohydrogen from microalgae; pp. 391–418. [DOI] [Google Scholar]
- 80.Singh H., Das D. Energy from microalgae. Springer; 2018. Biofuels from microalgae: biohydrogen; pp. 201–228. [DOI] [Google Scholar]
- 81.Nabgan W., Tuan Abdullah T.A., Nabgan B., Jalil A.A., Nordin A.H., Ul-Hamid A., Hassan N.S., Hussain I., Coelho A., Amin A., Ikram M. Catalytic biohydrogen production from organic waste materials: A literature review and bibliometric analysis. Int. J. Hydrogen Energy. 2021;46:30903–30925. doi: 10.1016/j.ijhydene.2021.04.100. [DOI] [Google Scholar]
- 82.Kufryk G. Advances in utilizing cyanobacteria for hydrogen production. Adv. Microbiol. 2013;03:60–68. doi: 10.4236/aim.2013.36A008. [DOI] [Google Scholar]
- 83.Sadvakasova A.K., Kossalbayev B.D., Zayadan B.K., Bolatkhan K., Alwasel S., Najafpour M.M., Tomo T., Allakhverdiev S.I. Bioprocesses of hydrogen production by cyanobacteria cells and possible ways to increase their productivity. Renew. Sustain. Energy Rev. 2020;133 doi: 10.1016/j.rser.2020.110054. [DOI] [Google Scholar]
- 84.Sivaramakrishnan R., Shanmugam S., Sekar M., Mathimani T., Incharoensakdi A., Kim S.H., Parthiban A., Edwin Geo V., Brindhadevi K., Pugazhendhi A. Insights on biological hydrogen production routes and potential microorganisms for high hydrogen yield. Fuel. 2021;291 doi: 10.1016/j.fuel.2021.120136. [DOI] [Google Scholar]
- 85.Osman A.I., Deka T.J., Baruah D.C., Rooney D.W. Critical challenges in biohydrogen production processes from the organic feedstocks. Biomass Convers. Biorefin. 2020;13:8383–8401. doi: 10.1007/s13399-020-00965-x. [DOI] [Google Scholar]
- 86.Show K.Y., Yan Y., Zong C., Guo N., Chang J.S., Lee D.J. State of the art and challenges of biohydrogen from microalgae. Bioresour. Technol. 2019;289 doi: 10.1016/j.biortech.2019.121747. [DOI] [PubMed] [Google Scholar]
- 87.Baeyens J., Zhang H., Nie J., Appels L., Dewil R., Ansart R., Deng Y. Reviewing the potential of bio-hydrogen production by fermentation. Renew. Sustain. Energy Rev. 2020;131 doi: 10.1016/j.rser.2020.110023. [DOI] [Google Scholar]
- 88.Kucharska K., Makoś-Chełstowska P., Słupek E., Gębicki J. Management of dark fermentation broth via bio refining and photo fermentation. Energies. 2021;14:6268. doi: 10.3390/en14196268. [DOI] [Google Scholar]
- 89.Abdalla A.M., Hossain S., Nisfindy O.B., Azad A.T., Dawood M., Azad A.K. Hydrogen production, storage, transportation and key challenges with applications: A review. Energy Convers. Manag. 2018;165:602–627. doi: 10.1016/j.enconman.2018.03.088. [DOI] [Google Scholar]
- 90.El-Dalatony M.M., Zheng Y., Ji M.K., Li X., Salama E.S. Metabolic pathways for microalgal biohydrogen production: Current progress and future prospectives. Bioresour. Technol. 2020;318 doi: 10.1016/j.biortech.2020.124253. [DOI] [PubMed] [Google Scholar]
- 91.Lopez-Hidalgo A.M., Smoliński A., Sanchez A. A meta-analysis of research trends on hydrogen production via dark fermentation. Int. J. Hydrogen Energy. 2022;47:13300–13339. doi: 10.1016/j.ijhydene.2022.02.106. [DOI] [Google Scholar]
- 92.Hallenbeck P.C., Ghosh D. Improvements in fermentative biological hydrogen production through metabolic engineering. J. Environ. Manag. 2012;95:S360–S364. doi: 10.1016/j.jenvman.2010.07.021. [DOI] [PubMed] [Google Scholar]
- 93.Hu Q., Sommerfeld M., Jarvis E., Ghirardi M., Posewitz M., Seibert M., Darzins A. Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances. Plant J. 2008;54:621–639. doi: 10.1111/j.1365-313X.2008.03492.x. [DOI] [PubMed] [Google Scholar]
- 94.Mathews J., Wang G. Metabolic pathway engineering for enhanced biohydrogen production. Int. J. Hydrogen Energy. 2009;34:7404–7416. doi: 10.1016/j.ijhydene.2009.05.078. [DOI] [Google Scholar]
- 95.Xia A., Cheng J., Song W., Su H., Ding L., Lin R., Lu H., Liu J., Zhou J., Cen K. Fermentative hydrogen production using algal biomass as feedstock. Renew. Sustain. Energy Rev. 2015;51:209–230. doi: 10.1016/j.rser.2015.05.076. [DOI] [Google Scholar]
- 96.Sarangi P.K., Nanda S. Biohydrogen production through dark fermentation. Chem. Eng. Technol. 2020;43:601–612. doi: 10.1002/ceat.201900452. [DOI] [Google Scholar]
- 97.Salakkam A., Sittijunda S., Mamimin C., Phanduang O., Reungsang A. Valorization of microalgal biomass for biohydrogen generation: A review. Bioresour. Technol. 2021;322 doi: 10.1016/j.biortech.2020.124533. [DOI] [PubMed] [Google Scholar]
- 98.Li X., Patena W., Fauser F., Jinkerson R.E., Saroussi S., Meyer M.T., Ivanova N., Robertson J.M., Yue R., Zhang R., et al. A genome-wide algal mutant library and functional screen identifies genes required for eukaryotic photosynthesis. Nat. Genet. 2019;51:627–635. doi: 10.1038/s41588-019-0370-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Singh H., Varanasi J.L., Banerjee S., Das D. Production of carbohydrate enrich microalgal biomass as a bioenergy feedstock. Energy. 2019;188 doi: 10.1016/j.energy.2019.116039. [DOI] [Google Scholar]
- 100.Barati B., Zeng K., Baeyens J., Wang S., Addy M., Gan S.Y., El-Fatah Abomohra A. Recent progress in genetically modified microalgae for enhanced carbon dioxide sequestration. Biomass Bioenergy. 2021;145 doi: 10.1016/j.biombioe.2020.105927. [DOI] [Google Scholar]
- 101.Sanz Smachetti M.E., Perez Cenci M., Salerno G.L., Curatti L. Ethanol and protein production from minimally processed biomass of a genetically-modified cyanobacterium over-accumulating sucrose. Bioresour. Technol. Rep. 2019;5:230–237. doi: 10.1016/j.biteb.2019.01.019. [DOI] [Google Scholar]
- 102.Yang B., Liu J., Ma X., Guo B., Liu B., Wu T., Jiang Y., Chen F. Genetic engineering of the Calvin cycle toward enhanced photosynthetic CO2 fixation in microalgae. Biotechnol. Biofuels. 2017;10:229–313. doi: 10.1186/s13068-017-0916-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Wei L., Wang Q., Xin Y., Lu Y., Xu J. Enhancing photosynthetic biomass productivity of industrial oleaginous microalgae by overexpression of RuBisCO activase. Algal Res. 2017;27:366–375. doi: 10.1016/j.algal.2017.07.023. [DOI] [Google Scholar]
- 104.Priyadharsini P., Nirmala N., Dawn S.S., Baskaran A., SundarRajan P., Gopinath K.P., Arun J. Genetic improvement of microalgae for enhanced carbon dioxide sequestration and enriched biomass productivity: Review on CO2 bio-fixation pathways modifications. Algal Res. 2022;66 doi: 10.1016/j.algal.2022.102810. [DOI] [Google Scholar]
- 105.Onyeaka H., Miri T., Obileke K., Hart A., Anumudu C., Al-Sharify Z.T. Minimizing carbon footprint via microalgae as a biological capture. Carbon Capture Sci. Technol. 2021;1 doi: 10.1016/j.ccst.2021.100007. [DOI] [Google Scholar]
- 106.Lin W.R., Lai Y.C., Sung P.K., Tan S.I., Chang C.H., Chen C.Y., Chang J.S., Ng I.S. Enhancing carbon capture and lipid accumulation by genetic carbonic anhydrase in microalgae. J. Taiwan Inst. Chem. Eng. 2018;93:131–141. doi: 10.1016/j.jtice.2018.10.010. [DOI] [Google Scholar]
- 107.Shanmugam S., Hari A., Kumar D., Rajendran K., Mathimani T., Atabani A., Brindhadevi K., Pugazhendhi A. Recent developments and strategies in genome engineering and integrated fermentation approaches for biobutanol production from microalgae. Fuel. 2021;285 doi: 10.1016/j.fuel.2020.119052. [DOI] [Google Scholar]
- 108.de Carvalho Silvello M.A., Severo Gonçalves I., Patrícia Held Azambuja S., Silva Costa S., Garcia Pereira Silva P., Oliveira Santos L., Goldbeck R. Microalgae-based carbohydrates: A green innovative source of bioenergy. Bioresour. Technol. 2022;344 doi: 10.1016/j.biortech.2021.126304. [DOI] [PubMed] [Google Scholar]
- 109.Goswami R.K., Agrawal K., Verma P. Bioenergy research: commercial opportunities & challenges. Springer; 2021. Microalgae-based biofuel-integrated biorefinery approach as sustainable feedstock for resolving energy crisis; pp. 267–293. [DOI] [Google Scholar]
- 110.Ran W., Wang H., Liu Y., Qi M., Xiang Q., Yao C., Zhang Y., Lan X. Storage of starch and lipids in microalgae: Biosynthesis and manipulation by nutrients. Bioresour. Technol. 2019;291 doi: 10.1016/j.biortech.2019.121894. [DOI] [PubMed] [Google Scholar]
- 111.Li Z., Cao L., Zhao L., Yu L., Chen Y., Yoon K.S., Hu Q., Han D. Identification and Biotechnical Potential of a Gcn5-Related N-Acetyltransferase Gene in Enhancing Microalgal Biomass and Starch Production. Front. Plant Sci. 2020;11 doi: 10.3389/fpls.2020.544827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Bajhaiya A.K., Dean A.P., Zeef L.A.H., Webster R.E., Pittman J.K. PSR1 is a global transcriptional regulator of phosphorus deficiency responses and carbon storage metabolism in Chlamydomonas reinhardtii. Plant Physiol. 2016;170:1216–1234. doi: 10.1104/pp.15.01907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Bastos R.G. Energy from microalgae. Springer; 2018. Biofuels from microalgae: bioethanol; pp. 229–246. [DOI] [Google Scholar]
- 114.Kumar V., Sharma N., Jaiswal K.K., Vlaskin M.S., Nanda M., Tripathi M.K., Kumar S. Microalgae with a truncated light-harvesting antenna to maximize photosynthetic efficiency and biomass productivity: Recent advances and current challenges. Process Biochem. 2021;104:83–91. doi: 10.1016/j.procbio.2021.03.006. [DOI] [Google Scholar]
- 115.Grama S.B., Liu Z., Li J. Emerging trends in genetic engineering of microalgae for commercial applications. Mar. Drugs. 2022;20:285. doi: 10.3390/md20050285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Pan X., Cao P., Su X., Liu Z., Li M. Structural analysis and comparison of light-harvesting complexes I and II. Biochim. Biophys. Acta Bioenerg. 2020;1861 doi: 10.1016/j.bbabio.2019.06.010. [DOI] [PubMed] [Google Scholar]
- 117.Razu M.H., Hossain F., Khan M. Microalgae Biotechnology for Development of Biofuel and Wastewater Treatment. Springer; 2019. Advancement of bio-hydrogen production from microalgae; pp. 423–462. [DOI] [Google Scholar]
- 118.Beckmann J., Lehr F., Finazzi G., Hankamer B., Posten C., Wobbe L., Kruse O. Improvement of light to biomass conversion by de-regulation of light-harvesting protein translation in Chlamydomonas reinhardtii. J. Biotechnol. 2009;142:70–77. doi: 10.1016/j.jbiotec.2009.02.015. [DOI] [PubMed] [Google Scholar]
- 119.Polle J.E.W., Kanakagiri S.D., Melis A. tla1, a DNA insertional transformant of the green alga Chlamydomonas reinhardtii with a truncated light-harvesting chlorophyll antenna size. Planta. 2003;217:49–59. doi: 10.1007/s00425-002-0968-1. [DOI] [PubMed] [Google Scholar]
- 120.Tetali S.D., Mitra M., Melis A. Development of the light-harvesting chlorophyll antenna in the green alga Chlamydomonas reinhardtii is regulated by the novel Tla1 gene. Planta. 2007;225:813–829. doi: 10.1007/s00425-006-0392-z. [DOI] [PubMed] [Google Scholar]
- 121.Khan S., Fu P. Biotechnological perspectives on algae: a viable option for next generation biofuels. Curr. Opin. Biotechnol. 2020;62:146–152. doi: 10.1016/j.copbio.2019.09.020. [DOI] [PubMed] [Google Scholar]
- 122.Goswami R.K., Mehariya S., Obulisamy P.K., Verma P. Advanced microalgae-based renewable biohydrogen production systems: A review. Bioresour. Technol. 2021;320 doi: 10.1016/j.biortech.2020.124301. [DOI] [PubMed] [Google Scholar]
- 123.Ban S., Lin W., Luo Z., Luo J. Improving hydrogen production of Chlamydomonas reinhardtii by reducing chlorophyll content via atmospheric and room temperature plasma. Bioresour. Technol. 2019;275:425–429. doi: 10.1016/j.biortech.2018.12.062. [DOI] [PubMed] [Google Scholar]
- 124.Von Abendroth G., Stripp S., Silakov A., Croux C., Soucaille P., Girbal L., Happe T. Optimized over-expression of [FeFe] hydrogenases with high specific activity in Clostridium acetobutylicum. Int. J. Hydrogen Energy. 2008;33:6076–6081. doi: 10.1016/j.ijhydene.2008.07.122. [DOI] [Google Scholar]
- 125.El-Mohsnawy E., Abdelaal A.S., El-Sheekh M. Handbook of Algal Biofuels. Elsevier; 2022. Direct biohydrogen production from algae; pp. 303–330. [DOI] [Google Scholar]
- 126.Lu Y., Koo J. O2 sensitivity and H2 production activity of hydrogenases-A review. Biotechnol. Bioeng. 2019;116:3124–3135. doi: 10.1002/bit.27136. [DOI] [PubMed] [Google Scholar]
- 127.Yang D.W., Syn J.W., Hsieh C.H., Huang C.C., Chien L.F. Genetically engineered hydrogenases promote biophotocatalysis-mediated H2 production in the green alga Chlorella sp. Int. J. Hydrogen Energy. 2019;44:2533–2545. doi: 10.1016/j.ijhydene.2018.11.088. [DOI] [Google Scholar]
- 128.Khosravitabar F., Hippler M. A new approach for improving microalgal biohydrogen photoproduction based on safe & fast oxygen consumption. Int. J. Hydrogen Energy. 2019;44:17835–17844. doi: 10.1016/j.ijhydene.2019.05.038. [DOI] [Google Scholar]
- 129.Xu F.Q., Ma W.M., Zhu X.G. Introducing pyruvate oxidase into the chloroplast of Chlamydomonas reinhardtii increases oxygen consumption and promotes hydrogen production. Int. J. Hydrogen Energy. 2011;36:10648–10654. doi: 10.1016/j.ijhydene.2011.05.130. [DOI] [Google Scholar]
- 130.Sivaramakrishnan R., Incharoensakdi A. Ecophysiology and Biochemistry of Cyanobacteria. Springer; 2021. Cyanobacteria as Renewable Sources of Bioenergy (Biohydrogen, Bioethanol, and Bio-Oil Production) pp. 431–454. [DOI] [Google Scholar]
- 131.Xu L., Fan J., Wang Q. Omics application of bio-hydrogen production through green alga Chlamydomonas reinhardtii. Front. Bioeng. Biotechnol. 2019;7:201. doi: 10.3389/fbioe.2019.00201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Wang Y., Jiang X., Hu C., Sun T., Zeng Z., Cai X., Li H., Hu Z. Optogenetic regulation of artificial microRNA improves H2 production in green alga Chlamydomonas reinhardtii. Biotechnol. Biofuels. 2017;10:257–259. doi: 10.1186/s13068-017-0941-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Li H., Zhang L., Shu L., Zhuang X., Liu Y., Chen J., Hu Z. Sustainable photosynthetic H2-production mediated by artificial miRNA silencing of OEE2 gene in green alga Chlamydomonas reinhardtii. Int. J. Hydrogen Energy. 2015;40:5609–5616. doi: 10.1016/j.ijhydene.2015.02.073. [DOI] [Google Scholar]
- 134.Giri D.D., Dwivedi H., Khalaf D Alsukaibi A., Pal D.B., Otaibi A.A., Areeshi M.Y., Haque S., Gupta V.K. Sustainable production of algae-bacteria granular consortia based biological hydrogen: New insights. Bioresour. Technol. 2022;352 doi: 10.1016/j.biortech.2022.127036. [DOI] [PubMed] [Google Scholar]
- 135.Masukawa H., Kitashima M., Inoue K., Sakurai H., Hausinger R.P. Genetic engineering of cyanobacteria to enhance biohydrogen production from sunlight and water. Ambio. 2012;41:169–173. doi: 10.1007/s13280-012-0275-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Mishra A.K., Kaushik M.S., Tiwari D. Cyanobacteria. Elsevier; 2019. Nitrogenase and hydrogenase: enzymes for nitrogen fixation and hydrogen production in cyanobacteria; pp. 173–191. [DOI] [Google Scholar]
- 137.Tanvir R.U., Zhang J., Canter T., Chen D., Lu J., Hu Z. Harnessing solar energy using phototrophic microorganisms: A sustainable pathway to bioenergy, biomaterials, and environmental solutions. Renew. Sustain. Energy Rev. 2021;146:111181. doi: 10.1016/j.rser.2021.111181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Masukawa H., Mochimaru M., Sakurai H. Disruption of the uptake hydrogenase gene, but not of the bidirectional hydrogenase gene, leads to enhanced photobiological hydrogen production by the nitrogen-fixing cyanobacterium Anabaena sp. PCC 7120. Appl. Microbiol. Biotechnol. 2002;58:618–624. doi: 10.1007/s00253-002-0934-7. [DOI] [PubMed] [Google Scholar]
- 139.Sakurai H., Masukawa H., Zhang X., Ikeda H., Inoue K. Photosynthesis. Energy from the Sun. Springer; 2008. Improvement of nitrogenase-based photobiological hydrogen production by cyanobacteria by gene engineering–hydrogenases and homocitrate synthase; pp. 1277–1280. [DOI] [Google Scholar]
- 140.Yoshino F., Ikeda H., Masukawa H., Sakurai H. High photobiological hydrogen production activity of a Nostoc sp. PCC 7422 uptake hydrogenase-deficient mutant with high nitrogenase activity. Mar. Biotechnol. 2007;9:101–112. doi: 10.1007/s10126-006-6035-3. [DOI] [PubMed] [Google Scholar]
- 141.Masukawa H., Inoue K., Sakurai H. Effects of disruption of homocitrate synthase genes on Nostoc sp. strain PCC 7120 photobiological hydrogen production and nitrogenase. Appl. Environ. Microbiol. 2007;73:7562–7570. doi: 10.1128/AEM.01160-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Thiel T. Organization and regulation of cyanobacterial nif gene clusters: implications for nitrogenase expression in plant cells. FEMS Microbiol. Lett. 2019;366:fnz077. doi: 10.1093/femsle/fnz077. [DOI] [PubMed] [Google Scholar]
- 143.Dong F., Lee Y.S., Gaffney E.M., Grattieri M., Haddadin H., Minteer S.D., Chen H. An engineered, non-diazotrophic cyanobacterium and its application in bioelectrochemical nitrogen fixation. Cell Rep. Phys. Sci. 2021;2 doi: 10.1016/j.xcrp.2021.100444. [DOI] [Google Scholar]
- 144.Larrainzar E., Villar I., Rubio M.C., Pérez-Rontomé C., Huertas R., Sato S., Mun J.H., Becana M. Hemoglobins in the legume–Rhizobium symbiosis. New Phytol. 2020;228:472–484. doi: 10.1111/nph.16673. [DOI] [PubMed] [Google Scholar]
- 145.Ding X.T., Fan Y., Jiang E.Y., Shi X.Y., Krautter E., Hu G.R., Li F.L. Expression of the Vitreoscilla hemoglobin gene in Nannochloropsis oceanica regulates intracellular oxygen balance under high-light. J. Photochem. Photobiol., B. 2021;221 doi: 10.1016/j.jphotobiol.2021.112237. [DOI] [PubMed] [Google Scholar]
- 146.Wu S., Xu L., Huang R., Wang Q. Improved biohydrogen production with an expression of codon-optimized hemH and lba genes in the chloroplast of Chlamydomonas reinhardtii. Bioresour. Technol. 2011;102:2610–2616. doi: 10.1016/j.biortech.2010.09.123. [DOI] [PubMed] [Google Scholar]
- 147.Manoyan J., Gabrielyan L., Kozel N., Trchounian A. Regulation of biohydrogen production by protonophores in novel green microalgae Parachlorella kessleri. J. Photochem. Photobiol., B. 2019;199 doi: 10.1016/j.jphotobiol.2019.111597. [DOI] [PubMed] [Google Scholar]
- 148.Khetkorn W., Rastogi R.P., Incharoensakdi A., Lindblad P., Madamwar D., Pandey A., Larroche C. Microalgal hydrogen production–A review. Bioresour. Technol. 2017;243:1194–1206. doi: 10.1016/j.biortech.2017.07.085. [DOI] [PubMed] [Google Scholar]
- 149.Esper B., Badura A., Rögner M. Photosynthesis as a power supply for (bio-) hydrogen production. Trends Plant Sci. 2006;11:543–549. doi: 10.1016/j.tplants.2006.09.001. [DOI] [PubMed] [Google Scholar]
- 150.Oey M., Sawyer A.L., Ross I.L., Hankamer B. Challenges and opportunities for hydrogen production from microalgae. Plant Biotechnol. J. 2016;14:1487–1499. doi: 10.1111/pbi.12516. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Kruse O., Rupprecht J., Mussgnug J.H., Dismukes G.C., Hankamer B. Photosynthesis: a blueprint for solar energy capture and biohydrogen production technologies. Photochem. Photobiol. Sci. 2005;4:957–970. doi: 10.1039/b506923h. [DOI] [PubMed] [Google Scholar]
- 152.Magnuson A. Heterocyst thylakoid bioenergetics. Life. 2019;9:13. doi: 10.3390/life9010013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Kumar K., Ghosh S., Angelidaki I., Holdt S.L., Karakashev D.B., Morales M.A., Das D. Recent developments on biofuels production from microalgae and macroalgae. Renew. Sustain. Energy Rev. 2016;65:235–249. doi: 10.1016/j.rser.2016.06.055. [DOI] [Google Scholar]
- 154.Eilenberg H., Weiner I., Ben-Zvi O., Pundak C., Marmari A., Liran O., Wecker M.S., Milrad Y., Yacoby I. The dual effect of a ferredoxin-hydrogenase fusion protein in vivo: successful divergence of the photosynthetic electron flux towards hydrogen production and elevated oxygen tolerance. Biotechnol. Biofuels. 2016;9:182–210. doi: 10.1186/s13068-016-0601-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Dubini A., Gonzalez-Ballester D. Algae Biotechnology. Springer; 2016. Biohydrogen from microalgae; pp. 165–193. [DOI] [Google Scholar]
- 156.Dubini A., Ghirardi M.L. Engineering photosynthetic organisms for the production of biohydrogen. Photosynth. Res. 2015;123:241–253. doi: 10.1007/s11120-014-9991-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Sun Y., Chen M., Yang H., Zhang J., Kuang T., Huang F. Enhanced H2 photoproduction by down-regulation of ferredoxin-NADP+ reductase (FNR) in the green alga Chlamydomonas reinhardtii. Int. J. Hydrogen Energy. 2013;38:16029–16037. doi: 10.1016/j.ijhydene.2013.10.011. [DOI] [Google Scholar]
- 158.Pinto T.S., Malcata F.X., Arrabaça J.D., Silva J.M., Spreitzer R.J., Esquível M.G. Rubisco mutants of Chlamydomonas reinhardtii enhance photosynthetic hydrogen production. Appl. Microbiol. Biotechnol. 2013;97:5635–5643. doi: 10.1007/s00253-013-4920-z. [DOI] [PubMed] [Google Scholar]
- 159.Liang F., Lindblad P. Effects of overexpressing photosynthetic carbon flux control enzymes in the cyanobacterium Synechocystis PCC 6803. Metab. Eng. 2016;38:56–64. doi: 10.1016/j.ymben.2016.06.005. [DOI] [PubMed] [Google Scholar]
- 160.Eroglu E., Melis A. Microalgal hydrogen production research. Int. J. Hydrogen Energy. 2016;41:12772–12798. doi: 10.1016/j.ijhydene.2016.05.115. [DOI] [Google Scholar]
- 161.Johnson E., Anastasios M. Functional characterization of Chlamydomonas reinhardtii with alterations in the atpE gene. Photosynth. Res. 2004;82:131–140. doi: 10.1007/s11120-004-6567-1. [DOI] [PubMed] [Google Scholar]
- 162.Robertson D., Boynton J.E., Gillham N.W. Cotranscription of the wild-type chloroplast atpE gene encoding the CF1/CF0 epsilon subunit with the 3′ half of the rps7 gene in Chlamydomonas reinhardtii and characterization of frameshift mutations in atpE. Mol. Gen. Genet. 1990;221:155–163. doi: 10.1007/BF00261715. [DOI] [PubMed] [Google Scholar]
- 163.Tolleter D., Ghysels B., Alric J., Petroutsos D., Tolstygina I., Krawietz D., Happe T., Auroy P., Adriano J.M., Beyly A., et al. Control of hydrogen photoproduction by the proton gradient generated by cyclic electron flow in Chlamydomonas reinhardtii. Plant Cell. 2011;23:2619–2630. doi: 10.1105/tpc.111.086876. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Elman T., Hoai Ho T.T., Milrad Y., Hippler M., Yacoby I. Enhanced chloroplast-mitochondria crosstalk promotes ambient algal-H2 production. Cell Rep. Phys. Sci. 2022;3 doi: 10.1016/j.xcrp.2022.100828. [DOI] [Google Scholar]
- 165.Aamer Mehmood M., Shahid A., Malik S., Wang N., Rizwan Javed M., Nabeel Haider M., Verma P., Umer Farooq Ashraf M., Habib N., Syafiuddin A., Boopathy R. Advances in developing metabolically engineered microbial platforms to produce fourth-generation biofuels and high-value biochemicals. Bioresour. Technol. 2021;337 doi: 10.1016/j.biortech.2021.125510. [DOI] [PubMed] [Google Scholar]
- 166.Kim J.Y., Ahn Y.J., Lee J.A., Lee S.Y. Recent advances in the production of platform chemicals using metabolically engineered microorganisms. Curr. Opin. Green Sustain. Chem. 2023;40 doi: 10.1016/j.cogsc.2023.100777. [DOI] [Google Scholar]
- 167.Nagarajan D., Dong C.D., Chen C.Y., Lee D.J., Chang J.S. Biohydrogen production from microalgae-Major bottlenecks and future research perspectives. Biotechnol. J. 2021;16 doi: 10.1002/biot.202000124. [DOI] [PubMed] [Google Scholar]
- 168.Oh Y.-K., Raj S.M., Jung G.Y., Park S. Current status of the metabolic engineering of microorganisms for biohydrogen production. Bioresour. Technol. 2011;102:8357–8367. doi: 10.1016/j.biortech.2011.04.054. [DOI] [PubMed] [Google Scholar]
- 169.Hallenbeck P.C., Abo-Hashesh M., Ghosh D. Strategies for improving biological hydrogen production. Bioresour. Technol. 2012;110:1–9. doi: 10.1016/j.biortech.2012.01.103. [DOI] [PubMed] [Google Scholar]
- 170.Margareta W., Nagarajan D., Chang J.S., Lee D.J. Dark fermentative hydrogen production using macroalgae (Ulva sp.) as the renewable feedstock. Appl. Energy. 2020;262 doi: 10.1016/j.apenergy.2020.114574. [DOI] [Google Scholar]
- 171.Boshagh F. Measurement methods of carbohydrates in dark fermentative hydrogen production-A review. Int. J. Hydrogen Energy. 2021;46:24028–24050. doi: 10.1016/j.ijhydene.2021.04.204. [DOI] [Google Scholar]
- 172.Jain R., Grogan D.W., Jain S.K., Gupta T., Agarwal C., Meena S.S. Bio-hydrogen production through dark fermentation: an overview. Extremophiles. 2022;27:1–26. doi: 10.1007/s13399-022-03282-7. [DOI] [Google Scholar]
- 173.Baebprasert W., Jantaro S., Khetkorn W., Lindblad P., Incharoensakdi A. Increased H2 production in the cyanobacterium Synechocystis sp. strain PCC 6803 by redirecting the electron supply via genetic engineering of the nitrate assimilation pathway. Metab. Eng. 2011;13:610–616. doi: 10.1016/j.ymben.2011.07.004. [DOI] [PubMed] [Google Scholar]
- 174.Huang G., Zhao D., Lan C., Wu B., Li X., Lou S., Zheng Y., Huang Y., Hu Z., Jia B. Glucose-assisted trophic conversion of Chlamydomonas reinhardtii by expression of glucose transporter GLUT1. Algal Res. 2022;62 doi: 10.1016/j.algal.2021.102626. [DOI] [Google Scholar]
- 175.Doebbe A., Rupprecht J., Beckmann J., Mussgnug J.H., Hallmann A., Hankamer B., Kruse O. Functional integration of the HUP1 hexose symporter gene into the genome of C. reinhardtii: impacts on biological H2 production. J. Biotechnol. 2007;131:27–33. doi: 10.1016/j.jbiotec.2007.05.017. [DOI] [PubMed] [Google Scholar]
- 176.Gomes de Oliveira Dal’Molin C., Quek L.E., Palfreyman R.W., Nielsen L.K. AlgaGEM–a genome-scale metabolic reconstruction of algae based on the Chlamydomonas reinhardtii genome. BMC Genom. 2011:1–10. doi: 10.1186/1471-2164-12-S4-S5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Mona S., Kumar S.S., Kumar V., Parveen K., Saini N., Deepak B., Pugazhendhi A. Green technology for sustainable biohydrogen production (waste to energy): a review. Sci. Total Environ. 2020;728 doi: 10.1016/j.scitotenv.2020.138481. [DOI] [PubMed] [Google Scholar]
- 178.Ahmed S.F., Rafa N., Mofijur M., Badruddin I.A., Inayat A., Ali M.S., Farrok O., Yunus Khan T.M. Biohydrogen production from biomass sources: metabolic pathways and economic analysis. Front. Energy Res. 2021;9 doi: 10.3389/fenrg.2021.753878. [DOI] [Google Scholar]
- 179.He X., Lei L., Dai Z. Green hydrogen enrichment with carbon membrane processes: Techno-economic feasibility and sensitivity analysis. Sep. Purif. Technol. 2021;276 doi: 10.1016/j.seppur.2021.119346. [DOI] [Google Scholar]
- 180.Bhatia S.K., Mehariya S., Bhatia R.K., Kumar M., Pugazhendhi A., Awasthi M.K., Atabani A.E., Kumar G., Kim W., Seo S.O., Yang Y.H. Wastewater based microalgal biorefinery for bioenergy production: Progress and challenges. Sci. Total Environ. 2021;751 doi: 10.1016/j.scitotenv.2020.141599. [DOI] [PubMed] [Google Scholar]
- 181.Long B., Fischer B., Zeng Y., Amerigian Z., Li Q., Bryant H., Li M., Dai S.Y., Yuan J.S. Machine learning-informed and synthetic biology-enabled semi-continuous algal cultivation to unleash renewable fuel productivity. Nat. Commun. 2022;13:541–611. doi: 10.1038/s41467-021-27665-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Khoo K.S., Ahmad I., Chew K.W., Iwamoto K., Bhatnagar A., Show P.L. Enhanced microalgal lipid production for biofuel using different strategies including genetic modification of microalgae: A review. Prog. Energy Combust. Sci. 2023;96 doi: 10.1016/j.pecs.2023.101071. [DOI] [Google Scholar]
- 183.Vargas J., Kava V., Balmant W., Mariano A., Ordonez J. Modeling microalgae derived hydrogen production enhancement via genetic modification. Int. J. Hydrogen Energy. 2016;41:8101–8110. doi: 10.1016/j.ijhydene.2015.12.217. [DOI] [Google Scholar]
- 184.Wang K., Khoo K.S., Chew K.W., Selvarajoo A., Chen W.H., Chang J.S., Show P.L. Microalgae: the future supply house of biohydrogen and biogas. Front. Energy Res. 2021;9 doi: 10.3389/fenrg.2021.660399. [DOI] [Google Scholar]
- 185.Shokravi H., Shokravi Z., Aziz M.A., Shokravi H. Fossil Free Fuels. CRC Press; 2019. Algal biofuel: a promising alternative for fossil fuel; pp. 187–211. [DOI] [Google Scholar]
- 186.Shokravi H., Shokravi Z., Heidarrezaei M., Ong H.C., Rahimian Koloor S.S., Petrů M., Lau W.J., Ismail A.F. Fourth generation biofuel from genetically modified algal biomass: Challenges and future directions. Chemosphere. 2021;285 doi: 10.1016/j.chemosphere.2021.131535. [DOI] [PubMed] [Google Scholar]