Next Article in Journal
Improved Extraction of High Value-Added Polyphenols from Pomegranate Peel by Solid-State Fermentation
Previous Article in Journal
Fungicide Effect of a Novelty Antimicrobial Biosurfactant Extract Alone or Combined with Copper Oxychloride on Botrytis cinerea Cells
Previous Article in Special Issue
Glucose Conversion for Biobutanol Production from Fresh Chlorella sorokiniana via Direct Enzymatic Hydrolysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Microalgae to Bioenergy: Recent Advances in Biochemical Conversion Processes

by
Sheetal Kishor Parakh
1,2,*,†,
Zinong Tian
3,†,
Jonathan Zhi En Wong
3 and
Yen Wah Tong
1,2,3,*
1
NUS Environmental Research Institute, National University of Singapore, #02-01, T-Lab Building, 5A Engineering Drive 1, Singapore 117411, Singapore
2
Energy and Environmental Sustainability for Megacities (E2S2) Phase II, Campus for Research Excellence and Technological Enterprise (CREATE), 1 CREATE Way, Singapore 138602, Singapore
3
NUS Chemical and Biomolecular Engineering, College of Design and Engineering, National University of Singapore, E5 #02-09, 4 Engineering Drive 4, Singapore 117585, Singapore
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Fermentation 2023, 9(6), 529; https://doi.org/10.3390/fermentation9060529
Submission received: 31 March 2023 / Revised: 4 May 2023 / Accepted: 8 May 2023 / Published: 29 May 2023
(This article belongs to the Special Issue Algae—the Medium of Bioenergy Conversion)

Abstract

:
Concerns about rising energy demand, fossil fuel depletion, and global warming have increased interest in developing and utilizing alternate renewable energy sources. Among the available renewable resources, microalgae biomass, a third-generation feedstock, is promising for energy production due to its rich biochemical composition, metabolic elasticity, and ability to produce numerous bioenergy products, including biomethane, biohydrogen, and bioethanol. However, the true potential of microalgae biomass in the future bioenergy economy is yet to be realized. This review provides a comprehensive overview of various biochemical conversion processes (anaerobic digestion, direct biophotolysis, indirect biophotolysis, photo fermentation, dark fermentation, microalgae-catalyzed photo fermentation, microalgae-catalyzed dark fermentation, and traditional alcoholic fermentation by ethanologenic microorganisms) that could be adapted to transform microalgae biomass into different bioenergy products. Recent advances in biochemical conversion processes are compiled and critically analyzed, and their limitations in terms of process viability, efficacy, scalability, and economic and environmental sustainability are highlighted. Based on the current research stage and technological development, biomethane production from anaerobic digestion and bioethanol production from traditional fermentation are identified as promising methods for the future commercialization of microalgae-based bioenergy. However, significant challenges to these technologies’ commercialization remain, including the high microalgae production costs and low energy recovery efficiency. Future research should focus on reducing microalgae production costs, developing an integrated biorefinery approach, and effectively utilizing artificial intelligence tools for process optimization and scale-up to solve the current challenges and accelerate the development of microalgae-based bioenergy.

1. Introduction

Energy has played a crucial role in economic and social development [1]. However, over 80% of our energy demand is fulfilled by non-renewable and less environment-friendly fossil fuels such as coal, natural gas, and oil [2]. These fuels are the primary source of carbon dioxide emissions, responsible for over 90% of global carbon emissions [3]. This dependency on fossil fuels has created two crises: fossil fuel depletion and global climate change. Despite this, energy demand continues to rise and is projected to increase by 47% in the next 30 years, particularly in developing countries [4]. It is, thus, imperative to find alternative renewable and clean energy sources to ensure a sustainable future. Bioenergy, the oldest known form of energy derived from biomass, is a promising option for meeting growing energy demand sustainably [5]. Biomass can be replenished in a relatively shorter time when compared with fossil fuels, making bioenergy a renewable energy source [6]. Bioenergy is carbon-neutral, releasing only the carbon dioxide that biomass consumes during its growth. Compared to bioenergy, fossil fuels release new carbon dioxide into the atmosphere that was sequestered and stored under the earth’s crust for millions of years [7]. Another advantage of bioenergy is its local production, reducing the need for long-distance energy transportation [8]. Local bioenergy production reduces the environmental impact of energy production and stimulates the local economy by providing energy security and economic opportunities [9].
Bioenergy is a broad term that refers to energy derived from various biomass sources, including food/non-food crops, agricultural/forest residue, and even algae [10]. Biomass can be divided into three generations based on its type. First-generation biomass includes traditional food and energy crops, such as corn, sugarcane, maize, soybean, and palm. Second-generation biomass comprises non-food crops and waste, such as woody/grassy plants and forest product residues [11]. Currently, a majority of biofuel is derived from first-generation biomasses, such as sugar (36.3 billion liters/year), maize (61.8 billion liters/year), palm oil (18.3 billion liters/year), and soybean oil (13.6 billion liters/year) [12]. However, using first-generation feedstocks for bioenergy production presents a food vs. fuel dilemma and challenges the food security [13]. Although second-generation feedstocks do not include food crops, these still compete with food production systems for resources such as arable land, freshwater, and fertilizers [14], making them unsustainable for fuel production. Third-generation feedstock, algae, has the potential to meet future bioenergy demands without compromising resources used for food production [15].
Microalgae are highly efficient photosynthetic organisms that rapidly grow throughout the year in various habitats, including aquatic (fresh or marine) and terrestrial ecosystems [16]. Microalgae exhibit relatively higher growth rates than terrestrial plants, completing an entire growth cycle in just a few days [17]. Moreover, microalgae can be cultivated on non-arable land using wastewater or seawater as a source of nutrients [18]. High biomass yield, minimal resource requirements, and a consistent biomass supply make microalgae a leading candidate for bioenergy production. Initially, the interest in using microalgae for bioenergy production was primarily due to its high lipid content (up to 60–70%), which could be converted into biodiesel using transesterification processes [19]. However, microalgae biomass can also be converted into other forms of bioenergy using biochemical and thermochemical processes. Biochemical conversion mainly includes anaerobic digestion [20], alcoholic fermentation [21], and fermentation/biophotolysis [22] processes for biomethane, bioethanol, and biohydrogen production, respectively. Thermochemical conversion includes pyrolysis [23], gasification [24], and hydrothermal processes [25] for bio-oil/biochar/syngas, syngas, and bio-oil/hydrochar/biogas production, respectively. Both biochemical and thermochemical processes are efficient pathways to recover energy from microalgae biomass since these methods utilize the whole biomass and do not depend on extracting specific macromolecules [26]. However, biochemical processes have a competitive advantage over thermochemical processes due to their ability to process wet biomass, operate at ambient processing conditions (temperature and pressure), and exhibit high selectivity toward the desired product, making them more sustainable and environment-friendly in the long run [27].
Most review articles regarding microalgal bioenergy concentrate on a single biochemical conversion technology [28,29,30]. However, conducting a detailed review of the available biochemical conversion technologies is crucial to selecting appropriate and sustainable methods for converting microalgae into bioenergy efficiently. This article provides a comprehensive overview of various biochemical conversion technologies, highlighting the most recent developments and critically discussing their limitations. This review article aims to provide valuable information to scientists, entrepreneurs, and governments, assisting them in identifying further research and development opportunities in microalgae-based bioenergy production.

2. Microalgae Biomass and Its Components

Algae are efficient, sunlight-driven green cell factories that convert carbon dioxide into various biomolecules, including lipids, carbohydrates, and proteins [31]. The cellular content of these biomolecules varies significantly between different algae species [32]. Algae can be classified into two broad categories: microalgae and macroalgae. Microalgae are unicellular and smaller (up to a few millimeters), whereas macroalgae are large (up to a few meters) and multicellular. There are more than 50,000 microalgae species in nature, of which 4000 species have been identified, and only a few species (<50) have been commercialized [33]. Therefore, microalgae biomass has immense untapped and unexplored potential. Table 1 displays the biochemical composition of commonly used microalgae species.
All biomolecules in microalgae cells play crucial roles in growth, reproduction, metabolism, and other cellular functions [36]. Lipids, essential components of cell membranes, energy storage, and cell signaling molecules, can be categorized into two groups in microalgae: polar lipids and non-polar lipids [37]. Polar lipids, such as phospholipids and glycolipids, are structural lipids that help maintain cell shape and structure. Polar lipids constitute 41–92% of the total lipids in microalgae biomass. Non-polar lipids, or neutral lipids, such as sterols and free fatty acids (FFA), usually function as energy storage molecules and make up 5–51% of the total microalgal lipids [38]. Neutral lipids, stored as triacylglycerols (TAGs), are preferred for biodiesel production due to their lower degree of unsaturation and the fact that industrial-scale transesterification is designed to process acylglycerols and has limited efficacy on other lipid types [39]. This makes microalgae species with high concentrations of neutral lipids (TAGs) promising candidates for biodiesel production. Besides lipids, carbohydrates (polysaccharides or oligosaccharides) are energy-storage molecules and structural elements in all living cells [40]. Structural carbohydrates are primarily present in the cell wall, whereas storage components can accumulate inside or outside the chloroplast [19]. Some carbohydrates may also be excreted as exopolysaccharides [41]. Carbohydrates in microalgae cells consist mainly of cellulose and starch and are free of lignin [42], making them a suitable candidate for bioethanol production through fermentation or biomethane production through anaerobic digestion. In addition to carbohydrates and lipids, proteins play essential roles in living organisms. Proteins serve as the major structural components of cells and transport nutrients and other molecules in and out of cells. They also act as enzymes or catalysts for various cellular biochemical reactions [43]. However, the high protein content in microalgae biomass may lead to a low carbon/nitrogen (C/N) ratio, limiting the biomass conversion to bioenergy [44]. To address the issue of low C/N ratio, various physiochemical approaches have been developed (Table 2) to lower protein content and enhance the lipid and carbohydrate content of microalgae biomass. Nevertheless, as shown in Table 2, high lipid or carbohydrate accumulation in microalgae does not always coincide with high biomass production, posing challenges for using microalgae as a bioenergy feedstock.
With the advent of metabolic engineering and its suite of genome-editing tools, some researchers are now genetically modifying microalgae cells by overexpressing or knocking out specific genes to alter metabolic pathways and improving the productivity of the compound of interest [52]. For instance, knocking out the phospholipase A2 gene from Chlamydomonas reinhardtii using the CRISPR-Cas9 system resulted in a 64% increase in lipid productivity without compensating for the biomass growth rate [53]. In another study, the individual expression of three genes (glycerol-3-phosphate acyltransferase (GPAT), lysophosphatidic acid acyltransferase (LPAT), and diacylglycerol acyltransferase (DGAT)) in Neochloris oleoabundans resulted in a 52% increase in lipid content without significantly affecting the microalgae growth [54]. Additionally, the cloning and transforming of specific genes from other species of microorganisms like S. cerevisiae can help improve biomolecule biosynthesis pathways in microalgae cells [55]. Due to the potential for manipulating microalgae to produce high yields of energy-rich compounds like lipids and carbohydrates, it is considered one of the most important feedstocks for bioenergy production [56]. Figure 1 presents an overview of the conversion technologies that can transform microalgae biomass into bioenergy.

3. Biochemical Conversion

The essential part of the biochemical conversion process is to use microorganisms or enzymes to convert biomass into bioenergy [27]. Biochemical conversion technologies can handle biomass with high water content (>50%) [57], making it suitable for processing wet algae biomass. Since drying microalgae biomass is an energy-intensive step [58], biochemical conversion technologies are promising for transforming microalgae into bioenergy. The classical biochemical conversion processes include anaerobic digestion for biomethane production, alcoholic fermentation for bioethanol production, and biological hydrogen production.

3.1. Anaerobic Digestion

Anaerobic digestion (AD) is a complex process in which microorganisms degrade organic biomass under anaerobic conditions and convert it into biogas, majorly comprising methane and other products, such as carbon dioxide, trace amounts of hydrogen, and ammonia [59]. Biogas produced from the AD process can be directly combusted in gas boilers to generate heat or electricity or upgraded into natural gas-quality biomethane and injected into gas grids [60]. AD process consists of four steps, hydrolysis, acidogenesis, acetogenesis, and methanogenesis, where each step is led by a unique functional group of microorganisms [61]. In the first stage of AD (hydrolysis), existing macromolecules in the biomass, such as proteins, lipids, and carbohydrates, are broken down into simpler molecules, such as amino acids, long-chain fatty acids, and simple sugars, respectively. Hydrolysis is achieved by hydrolytic bacteria, mainly belonging to the phyla Firmicutes and Bacteroidetes, secreting a mixture of hydrolytic enzymes comprising cellulase, xylanase, pectinase, amylase, lipase, and protease [62]. In the second stage of AD (acidogenesis), hydrolyzed products obtained at the end of the hydrolysis stage are converted into volatile fatty acids, such as acetates, propionate, butyrate, valerate, and isobutyrate, using facultative and obligate anaerobic bacteria species, majorly belonging to the phyla Firmicutes, Bacteroidetes, Chloroflexi, Proteobacteria, and Atribacteria. Alcohol and other inorganic compounds, such as hydrogen, carbon dioxide, ammonia, and hydrogen sulfide, are also produced during the acidogenesis stage [63]. During the third stage of AD (acetogenesis), the acidogenesis products, such as propionate, butyrate, isobutyrate, valerate, and isovalerate, are further broken down into acetate as well as hydrogen and carbon dioxide using obligate anaerobic bacteria species, majorly belonging to phyla Firmicutes, Synergistota, and Myxococcota [64]. In the final step of AD (methanogenesis), methanogens (a specialized group of archaea belonging to the phyla Euryarchaeota, Bathyarchaeota, and Verstaraeteachaeota) convert acetic acid and hydrogen into methane and carbon dioxide [65]. Tuning these four metabolic stages in the AD process can influence final product yields [61]. AD is a well-established commercial technology that is currently being applied to a wide range of organic substrates (sewage waste [66], food waste [67], high-strength organic wastewater [68], and agricultural or forest residue [69]).
Research on anaerobic digestion (AD) of microalgal biomass began in 1957 when microalgae (Chlorella and Scenedesmus) were cultivated for wastewater treatment and subjected to AD for biomethane production [70]. However, the resulting biomethane yields were significantly lower due to two main factors: the rigidity of microalgae cell walls and the low carbon-to-nitrogen (C/N) ratio in microalgae biomass. Most microalgae cell walls are rigid, which limits anaerobic microorganisms’ access to biodegradable microalgal organic matter [28]. The rigidity of microalgae cell walls can be attributed to components such as hemicellulose, cellulose, glycoprotein structures, and certain carbohydrates (e.g., glucose, xylose, rhamnose, and galactose) [71]. Some types of microalgae species also contain algaenan in their outer cell walls. Algaenan is a heteropolymer compound, highly resistant to acidic and basic environments [72]. Several studies have reported intact microalgae cells in the AD effluent after a hydraulic retention time of 30–180 days [73,74], highlighting the recalcitrant nature of microalgal cell walls and their resistance to bacterial degradation during the AD process. Since insufficient biodegradation leads to lower biomethane production, a pretreatment step is necessary to disrupt the microalgae cell walls. Secondly, a C/N ratio of 20–30 (especially 25) is recommended for the optimal functioning of the AD process [75]. If the ratio falls below 20, high amounts of ammonia-nitrogen are released in the anaerobic digester due to the imbalance between microbial carbon and nitrogen requirements. Ammonia buildup can inhibit methanogen growth, leading to volatile fatty acid (VFA) accumulation and process failure [76]. However, the C/N ratio in microalgae biomass is usually between 4–8 [33]. To address the low C/N ratio in microalgae, several researchers have suggested co-digesting microalgae biomass with other biomass streams with a high C/N content [77]. Various biomass pretreatment and co-digestion technologies have been developed in the past decade to improve biomethane production from microalgae biomass. These strategies are discussed in detail in the following sections.

3.1.1. Pretreatment Technologies

Biomass pretreatment technologies focus on enhancing the biodegradability of microalgae biomass by disrupting cell walls. These technologies can be categorized into three groups: physical, chemical, and biological (Table 3). Physical pretreatment technologies consist of mechanical (high-pressure homogenization, bead-milling, microwave, and ultrasonication) and thermal (steam explosion, thermal hydrolysis, and hydrothermal treatment) methods [78]. Mechanical pretreatment uses mechanical forces to disrupt cell walls through size reduction (high-pressure homogenization and milling) or physical damage induction (microwave and ultrasonication) [79]. Among the available mechanical pretreatments, ultrasonication is more commonly applied to microalgae biomass [44]. Ultrasonication pretreatment uses high-frequency sonic waves (>20 kHz) to initiate a cavitation process that propagates shock waves in the medium surrounding cells and causes cell wall disruption by high shear forces [80]. A study dealing with AD of Scenedesmus sp. and Pinnularia sp. reported 65–71% higher biomethane production when both microalgae species were ultrasonically pretreated before AD [81]. In addition to ultrasonication, microwave pretreatment is an effective cell wall disruption method. Microwaves are electromagnetic waves with shorter wavelengths ranging from one meter to one millimeter corresponding to frequencies of 0.3 to 300 GHz, respectively [82]. Microwaves damage cell walls through athermal and thermal effects [83]. In the athermal part of reactions, microwaves polarize and realign macromolecules along the electromagnetic field, altering the macromolecular structure through hydrogen bond breakage. On the other hand, the thermal effect generates heat inside the cells by absorbing microwave energy through cellular organic complexes or surrounding aqueous media, resulting in cell damage. Passos et al. [84] analyzed the effect of ultrasound (with an energy input of 26.7 MJ/kg TS (TS: Total solids)) and microwave (with an energy input of 34.3 MJ/kg TS) pretreatment on the mixed microalgae biomass grown in high-rate algal ponds and reported an 8% and 21% increase in biomethane production, respectively, compared to untreated biomass. Apart from ultrasonication and microwave pretreatment, other mechanical pretreatment methods, such as high-pressure homogenization and bead milling, have also been widely studied for the pretreatment of microalgae biomass. Córdova et al. [34] reported a 39% increase in biomethane production when Chlorella sorokiniana was homogenized before AD. A 51% higher biomethane was produced when Acutodesmus obliquus was pretreated with glass beads [85]. Recently, Straessner et al. [86] used pulsed electric field (PEF) for the first time as a pretreatment technology to improve biomethane yield from Auxenochlorella protothecoides. PEF is a well-established technology commonly used for extracting molecules from microalgae biomass. It can permeabilize microalgae cell walls by applying electrical pulses. The authors reported a 10% increase in biomethane production in PEF-treated biomass compared to raw biomass. However, more PEF studies using different microalgae species are needed to prove the practicality of this approach in biomethane improvement. Nevertheless, based on the examples listed in Table 3, it can be seen that higher power input or longer exposure time during the treatment has generally favored mechanical pretreatments, leading to higher biomethane production. The high energy required by mechanical pretreatment makes it unfavorable from an energy perspective [84]. Therefore, achieving a net positive energy balance (energy output > energy input) through improved biomethane yield is crucial to the success of mechanical pretreatment applications.
Thermal pretreatment is used to solubilize cell walls by exposing the biomass to temperatures ranging from 50–240 °C [79]. When thermal pretreatment is conducted under atmospheric pressure at and below 100 °C, it is called thermal hydrolysis, whereas pretreatment at temperatures between 100 °C and 180 °C under gradual pressure (<2 MPa) is called hydrothermal pretreatment. Córdova et al. [34] studied the efficiency of thermal hydrolysis pretreatment on Chlorella sorokiniana biomass at different temperatures (60, 70, and 80 °C) and revealed a 6, 9, and 18% higher biomethane production, respectively. Wang et al. [87] compared the efficiency of thermal hydrolysis and hydrothermal pretreatment in improving biomethane yield from Chlorella sp. and obtained 114% higher biomethane from hydrothermally treated biomass, while biomass pretreated with thermal hydrolysis yielded 39–47% higher biomethane. Although hydrothermal pretreatment seems promising, a significant amount of energy is required to heat the biomass from ambient temperature to the desired pretreatment temperature (>100 °C), hindering the commercial viability of this approach. In recent studies, solar-driven hydrothermal pretreatment methods have been proposed to reduce energy expenditure [88]. Xiao et al. [89] utilized parabolic trough collectors to concentrate solar radiations for hydrothermal pretreatment of microalgae biomass and obtained 57% higher biomethane production. Furthermore, Xiao et al. [90] evaluated the thermodynamic performance of the solar-driven hydrothermal system using exergy analysis, which considers both the quantity and quality of energy. The solar-driven hydrothermal system achieved the highest exergy efficiency of 41%, whereas the exergy efficiencies of hydrothermal and control (without pretreatment) systems were 36 and 26%, respectively. Despite these promising results, solar-driven hydrothermal systems are still in the nascent development phase and require further engineering innovations to reduce capital costs and allow better integration with the existing bioenergy infrastructure. Apart from thermal hydrolysis and hydrothermal pretreatments, the steam explosion is a widely used thermal pretreatment method, especially in commercial refineries dealing with lignocellulosic biomass. In the steam explosion method, the biomass is first exposed to saturated steam (180–240 °C and 1.03–3.45 MPa) for several minutes, followed by sudden depressurization to ambient conditions [91]. However, studies have shown that steam explosion may not be a good pretreatment option for microalgae biomass. Martín Juárez et al. [92] observed no significant improvement in biomethane production when microalgae biomass was pretreated with a steam explosion (130 °C for 0.08 h and 170 °C for 0.34 h) before AD. Another study dealing with AD of Chlorella sorokiniana reported a reduction in biomethane production by 28–60% when biomass was pretreated with a steam output of 75 kg/h at 4 bar for 0.08–0.25 h [34]. The steam explosion pretreatment of microalgae biomass may have produced or released inhibitory compounds that affected the activity of AD microorganisms, resulting in lower biomethane production. However, the inhibitory mechanism observed during the steam explosion pretreatment of microalgae biomass remains unknown, and further studies need to be conducted to understand the process better.
Chemical pretreatment involves using acidic, alkaline, oxidizing agents, or organic solvents at varying temperatures to solubilize microalgae biomass. Acidic pretreatment uses sulfuric, hydrochloric, and free nitrous acids to hydrolyze cellulosic and hemicellulosic matrices in the cell wall into simple sugars [79]. Marques et al. [93] achieved a 93% higher biomethane production by treating Scenedesmus obliquus biomass with 0.1% (v/v) sulfuric acid at 150 °C for an hour. Bai et al. [94] reported a 55% increase in biomethane production from Tetraselmis striata M8 using free nitrous acid pretreatment. In contrast to acidic pretreatments, alkaline pretreatment uses alkalis such as sodium, potassium, or calcium hydroxide to permeabilize cell walls through saponification of uronic acids and esters present in cell walls, inducing swelling and increasing specific surface areas available for microbial degradation and protein solubilization [44]. Fu et al. [95] observed a 77% increase in biomethane production when Chlorella pyrenoidosa was pretreated with 1.5% w/v sodium hydroxide solution at 90 °C for two hours. Biomethane yield was enhanced by 133% when a 2 M sodium hydroxide solution was used to pretreat mixed microalgae consortium at 121 °C for an hour [92]. The higher biomethane yield observed in the latter study could be attributed to the high concentration of sodium hydroxide used. However, alkalis, such as sodium hydroxide, are expensive. Recently, a cheaper alkali alternative (lime) was investigated, and 25% higher biomethane was obtained when the mixed culture of Chlorella sp. and Scenedesmus sp. was pretreated with 10% w/v lime at 72 °C for four hours [96]. Despite the promise of acidic and alkaline pretreatments in enhancing biomethane production, these methods may lead to process equipment corrosion and contaminate the biomass by introducing toxic ions and molecules during the pretreatment [44]. Besides acidic and alkaline pretreatments, some studies have used organic solvent pretreatment. For example, Caporgno et al. [97] studied the effect of N-methylmorpholine-N-oxide (NMMO) solvent to pretreat Nannochloropsis oculata and achieved a 42% increase in biomethane production. However, the authors used an additional step of evaporation to remove residual solvent from the biomass before anaerobic digestion, which may increase the process cost. Another type of chemical pretreatment is oxidative pretreatment, which uses oxidizing agents to react with aromatic and unsaturated compounds and break down cell walls [98]. A recent study used peroxymonosulfate oxidant to pretreat Microcystis sp. before AD, but it only improved the biomethane yield by 4% [99]. In contrast, Cardeña et al. [100] achieved up to a 66% improvement in biomethane yield by applying different doses of ozone pretreatment to a mixed microalgae consortium. In another study, a mixed microalgae culture was subjected to 0.5 w/w% hydrogen peroxide at 50 °C for an hour before AD, resulting in a 72% increase in the biomethane production [92]. Li et al. [101] studied the effect of zero-valent iron dosage on ultrasonically pretreated Microcystis sp. and achieved a 64% higher biomethane. Oxidative pretreatments can be conducted under mild temperature and pressure conditions without using concentrated acids or bases. This makes the oxidative pretreatment method a less costly and more environmentally friendly alternative to traditional chemical pretreatment methods [78]. However, further research is needed to determine the cost-effectiveness of the oxidative pretreatment method based on the chemical and energy input. Recently, a research study conducted by Wang et al. [102] demonstrated that free ammonia (FA), which can be directly obtained from anaerobic digester effluent, improved biomethane production by 17%. Although the reported results of FA pretreatment are low compared to other chemical pretreatment methods, it is a closed-loop technology that can significantly reduce the environmental impact of the pretreatment process. More investigations should be carried out to improve the performance of FA pretreatments.
Biological pretreatment uses hydrolytic enzymes, such as cellulase, hemicellulase, protease, lipase, α-amylase, and amyloglucosidase, to break down microalgae cells. The type of enzyme used in biological pretreatment depends mainly on the microalgae cell wall composition. It is advantageous over other pretreatments due to its low energy requirements, higher selectivity, mild operational conditions, and no production of inhibitory metabolites. Biological pretreatment can be divided into two types: external addition of commercial-specific enzymes and in situ production of crude enzymes by microbial activity. External addition of commercial-specific enzymes involves the addition of isolated and purified enzymes to the biomass. The commercial enzymes used for this purpose are typically selected or genetically engineered to possess high activity and specificity for the targeted compounds, making the process more efficient and cost-effective. In situ production of crude enzymes, on the other hand, involves utilizing microbial activity to produce enzymes on-site. This is typically conducted by inoculating the biomass with a single or mixed community of microorganisms, such as fungi or bacteria, that excrete hydrolytic enzymes naturally or as a result of genetic engineering. Kendir Çakmak et al. [103] studied the effect of commercial enzyme pretreatment on biomethane production from Phorphyridium cruentum and reported a 109 and 102% increase with single enzyme (protease) and enzyme mixture (protease and viscozyme) addition, respectively. Another study reported a 48–62% and 143–162% increase in biomethane production when Chlorella vulgaris biomass was pretreated with a single and mixture of enzymes, respectively [104]. Nevertheless, the high cost of commercial enzymes discourages their wide-scale application in biomethane enhancement. To reduce the high costs of enzymes, they can be either produced using low-cost substrates, such as waste [105] or immobilized to allow their recovery and reuse in several pretreatment cycles [106]. However, producing enzymes using low-cost substrates and immobilizing them for reuse remains a subject of high interest where intense research is ongoing. Another strategy to make biological pretreatment cost-effective is producing crude enzymes in situ using fungi or bacteria capable of secreting extracellular enzymes. Hom-Diaz et al. [107] pretreated mixed microalgae consortium with fungal broth and observed a 74% increase in biomethane production. Kavitha et al. [108] reported a 217–334% increase in biomethane production when a mixed microalgae consortium was pretreated using different bacterial population sets secreting protease, amylase, and cellulase enzymes. Aydin et al. [109] used 20% cow rumen fluid containing fungi, such as Anaeromyces sp., Orpinomyces sp., Piromyces sp., and Neocallimastix sp., to anaerobically digest Haematococcus pluvialis and observed a two-time increase in biomethane production. However, selecting appropriate microorganisms, preparing their starter culture, and fixing their inoculum ratio with microalgae biomass are the two main hurdles of this process. Another factor that needs to be considered is the possible loss of microalgae biomass due to its use as a substrate by other microorganisms. Both biological pretreatment methods have their advantages and disadvantages. The external addition of commercial-specific enzymes is more efficient and predictable, but it can also be more expensive due to the cost of the enzymes. The in-situ production of crude enzymes, on the other hand, is cheaper and more sustainable, but it is also less predictable and may require more time to achieve the desired results.
Based on the above discussion and the examples listed in Table 3, it can be concluded that each pretreatment type can effectively enhance biomethane production from microalgae to a certain extent. However, solar-driven thermal, microbial, and oxidative pretreatments seem more promising than traditional pretreatments, thus requiring further exploration. It should also be noted that the most efficient pretreatment method may not be the same for each microalgae species since cell wall structures and compositions vary with microalgae species [78]. A careful selection of the pretreatment method and its process parameters will be required to enhance biomethane production from different microalgae species.
Table 3. Effect of pretreatment technologies on biomethane production from microalgae biomass.
Table 3. Effect of pretreatment technologies on biomethane production from microalgae biomass.
Microalgae
Species
Pretreatment Strategy and
Operating Conditions
Biomethane Yield
(mL CH4/g VS)
References
Without
Pretreatment
With
Pretreatment
% Improvement in Yield
Physical pretreatment (Mechanical)
Scenedesmus sp.Ultrasound pretreatment
at 400 W power for 200 s
183 ± 25313 ± 1571 [81]
Pinnularia sp.152 ± 21250 ± 2165
Mixed microalgae and bacteria consortium dominated by green microalgae (Stigeoclonium sp. and Monoraphidium sp.) and diatoms (Nitzschia sp. and Navicula sp.Ultrasound pretreatment
at 70 W power for 0.5 h
106 ± 2114 ± 28 [84]
Microwave pretreatment
at 900 W power for 180 s
128 ± 521
Chlorella sorokinianaHomogenization pretreatment
at 200 W power for 0.5 h
318 ± 1442 ± 2939 [34]
Acutodesmus obliquusBead milling pretreatment using 0.35 mm glass beads at 40 g of glass beads/100 g of wet algae for 0.34 h at 8500 rpm19128951 [85]
Auxenochlorella protothecoidesPulsed electric field pretreatment at 40 kV/cm electric field and 1 μs pulse duration (3 Hz)42546710 [86]
Physical pretreatment (Thermal)
Chlorella
sorokiniana
Thermal hydrolysis pretreatment at 60 °C for 0.5 h318 ± 1337 ± 126 [34]
Thermal hydrolysis pretreatment at 70 °C for 0.5 h347 ± 399
Thermal hydrolysis pretreatment at 80 °C for 0.5 h375 ± 5318
Chlorella sp.Thermal hydrolysis pretreatment at 70 °C for 0.5 h15521539 [87]
Thermal hydrolysis pretreatment at 90 °C for 0.5 h22847
Hydrothermal hydrolysis pretreatment at 121 °C for 0.5 h332114
Chlorella
pyrenoidosa
Solar-driven hydrothermal pretreatment at 723 W/m2 irradiation (~160 °C) for 0.5 h22234857 [89]
Chlorella
sorokiniana
Steam explosion pretreatment at 4 bars for 0.08 h318 ± 1137 ± 5−57 [34]
Steam explosion pretreatment at 4 bars for 0.17 h128 ± 7−60
Steam explosion pretreatment at 4 bars for 0.25 h230 ± 4−28
Chemical pretreatment
Scenedesmus obliquusAcidic pretreatment with 0.1% v/v sulfuric acid at 150 °C for 1 h131 ± 26253 ± 5193 [93]
Tetraselmis striata M8Acidic pretreatment with 2.31 mg/L free nitrous acid at 5.5 pH for 48 h161 ± 7250 ± 255 [94]
Chlorella
pyrenoidosa
Alkaline pretreatment with 1.5% (w/v) NaOH at 90 °C for 2 h21838677 [95]
Mixed microalgae
consortium
Alkaline pretreatment with 0.5 M NaOH at 121 °C for 1 h1621737 [92]
Alkaline pretreatment with 2 M NaOH at 121 °C for 1 h377133
Oxidative pretreatment with 0.5% w/w hydrogen peroxide (11.5 pH) at 50 °C for 1 h 27972
Mixed microalgae
consortium of Chlorella sp. and Scenedesmus sp.
Alkaline pretreatment with 4% (w/v) CaO at 55 °C for 24 h260 ± 8255 ± 6−2 [96]
Alkaline pretreatment with 10% (w/v) CaO at 55 °C for 24 h292 ± 1112
Alkaline pretreatment with 4% (w/v) CaO at 72 °C for 24 h287 ± 411
Alkaline pretreatment with 10% (w/v) CaO at 72 °C for 24 h325 ± 1225
Nannochloropsis oculataOrganosolv treatment with N-methylmorpholine-N-oxide 238 ± 6339 ± 442 [97]
Microcystis sp.Oxidative pretreatment with 0.1 g peroxymonosulfate/g algae (TSS) 291 (mL CH4/g COD)303 (mL CH4/g COD)4 [99]
Mixed microalgae
consortium
Oxidative pretreatment with 96 mg of ozone/g algae VS at 23 °C26030618 [100]
Oxidative pretreatment with 191 mg of ozone/g algae VS
at 23 °C
33428
Oxidative pretreatment with 383 mg of ozone/g algae VS
at 23 °C
43366
Ultrasonically pretreated Microcystis sp.Oxidative pretreatment with 20 g of zero-valent iron/g algae (TS) 37 (mL CH4/g COD)61 (mL CH4/g COD)64 [101]
Mixed microalgae
consortium
Free ammonia pretreatment with 530 mg NH3-N/L at 22 °C for 24 h (pH 9.5)18821917 [102]
Biological pretreatment
Porphyridium cruentumSingle enzymatic pretreatment with 0.5 mL/g dry biomass commercial cellulase at 55 °C for 24 h (pH 5–5.5)13015217 [103]
Single enzymatic pretreatment with 0.5 mL/g dry biomass commercial protease at 55 °C for 24 h (pH 8–8.5)271109
Cocktail enzymatic pretreatment with 0.5 mL/g dry biomass commercial viscozyme (carbohydrase mix) at 55 °C for 24 h (pH 4–4.5)24286
Cocktail enzymatic pretreatment with 0.5 mL/g dry biomass enzyme mix (commercial protease and viscozyme) at 55 °C for 9 h (pH 8–8.5 for first 4.5 h and 4–4.5 for next 4.5 h)263102
Chlorella vulgarisSingle enzymatic pretreatment with 1% w/v commercial
cellulase at 55 °C for 24 h
120 ± 15 183 ± 1253 [104]
Single enzymatic pretreatment with 1% w/v commercial
protease at 55 °C for 24 h
194 ± 162
Single enzymatic pretreatment with 1% w/v commercial
amylase at 55 °C for 24 h
177 ± 1448
Cocktail enzymatic pretreatment with 1% w/v enzyme mix (commercial cellulase and protease) at 55 °C for 24 h 314 ± 11162
Cocktail enzymatic pretreatment with 1% w/v enzyme mix (commercial cellulase and amylase) at 55 °C for 24 h291 ± 5143
Mixed microalgae–bacteria
consortium
Single enzyme pretreatment with 100 U/L commercial laccase at 25 °C83 ± 1100 ± 7 21 [107]
Fungal pretreatment with 100 U/L laccase-rich broth from Trametes versicolor144 ± 274
Mixed microalgae
consortium
Bacterial pretreatment with bacterial consortium secreting protease, amylase (Bacillus jerish 03 and Bacillus jerish 04)0.06 (g CODconverted/g CODadded)0.19 (g CODconverted/g CODadded)217 [108]
Bacterial pretreatment with cellulase secreting bacteria
(Bacillus sp.)
0.21 (g CODconverted/g CODadded)250
Bacterial pretreatment with mixed bacteria population (Bacillus jerish 03, Bacillus jerish 04 and Bacillus sp.)0.26 (g CODconverted/g CODadded)334
Mixed microalgae
consortium
Cow rumen fluid mixed with anaerobic granular sludge inoculum at 1:4 v/v ratio300600100 [109]
Note: VS—volatile solids, TSS—total suspended solids, COD—chemical oxygen demand, TS—total solids, g CODconverted/g CODadded—biodegradable fraction of COD converted to biomethane.

3.1.2. Co-Digestion

Most microalgae species have a low C/N ratio [33], and it is often associated with destabilization and reduced biomethane production in the AD process due to ammonia release and inhibition [110]. As a result, anaerobic digesters that solely process microalgae biomass operate at low organic loading rates (<2 g VS/L.d) to avoid ammonia build-up and subsequent process failure [77]. However, operations at low organic loading rates compromise the economic feasibility of microalgal AD. Co-digestion of microalgae biomass with carbon-rich streams is an opportunity to overcome the risk of ammonia inhibition by improving the C/N ratio through the simultaneous digestion of microalgae biomass with one or more highly biodegradable carbon-rich materials. Among different carbon-rich materials, carbon-rich waste streams are considered ideal co-substrates for co-digestion because they can economically enhance biomethane production while readjusting the nutrient balance (C/N ratio) in the digester. Table 4 summarizes a few examples of microalgae co-digestion with carbon-rich waste streams. Primary sewage and waste-activated sludge are the most researched co-substrates for microalgae digestion. Wágner et al. [111] co-digested a mixed-microalgae consortium of Chlorella sorokiniana and Scenedesmus sp. with waste-activated sludge and observed a 21–69% increase in biomethane production. In another study [112], biomethane production was increased by 12% when Scenedesmus quadricauda was co-digested with thickened waste-activated sludge. Solé-Bundó et al. [113] varied the microalgae to the co-substrate ratio (VS basis) in the AD from 100:0 to 75:25, 50:50, and 25:75, and reported 48, 140, and 223% higher biomethane production, respectively. Microalgae biomass has been successfully digested with other types of organic substrates, including food waste [114], animal manure [115], agro-industrial waste [116,117], glycerol (a by-product of the biodiesel industry) [110], and fat, oil, and grease (FOG) waste [118] to produce 19–500% higher biomethane. Recent studies have suggested pretreating microalgae biomass before co-digestion to achieve better biomethane yields. For instance, Fu et al. [95] treated Chlorella pyrenoidosa with a thermo-alkaline method before co-digestion with sewage sludge and observed an 83% increase in biomethane production. Similarly, a 20% higher biomethane was produced when thermally pretreated microalgae biomass was co-digested with sewage sludge [119]. In these studies, biomass pretreatment may have improved microalgae biomass solubilization by rupturing rigid microalgae cell walls.
The co-digestion of microalgae biomass has shown promising results, but a reduction in biomethane production has also been reported when microalgae biomass was co-digested [120]. This highlights the need for a better understanding of co-substrate dosing strategies. For example, Avila et al. [121] observed a reduction of about four times in biomethane production when enzymatically pretreated microalgae biomass was co-digested with waste-activated sludge, mixed at a ratio of 7:93 (VS basis). Similarly, co-digestion of microalgae biomass with piggery wastewater (40:60 organic matter basis) produced 13% lesser biomethane [122]. Currently, the microalgae co-digestion approach primarily focuses on balancing the C/N ratio of the feedstock, but it may not be sufficient to enhance biomethane production. Other variables, such as operating conditions, digester configurations, and underlying microbial dynamics, may affect the biomethane output. Therefore, linking these factors with process monitoring, such as pH, daily biomethane production, and accumulation of intermediate metabolites, can be extremely helpful in understanding and optimizing the co-digestion process [123]. More lab and pilot-scale batch and continuous studies equipped with online process monitoring are needed to gain insights into the co-digestion process.

3.2. Biohydrogen Production

Biohydrogen is considered the fuel of the future due to its higher heating value (142 kJ/g) and ability to produce energy without emitting carbon dioxide [124]. It can be used in fuel cells to generate electricity and as a fuel for automobiles, providing a carbon-neutral solution to current energy crises [125]. To tap into the potential of biohydrogen as a future fuel, microalgae have emerged as a promising source for its production. Microalgae can produce biohydrogen through two different pathways: bio-photolysis and fermentation. During biophotolysis, microalgae cells act as a biocatalyst to directly produce biohydrogen, whereas in fermentation, microalgae biomass serves as a feedstock for biohydrogen-producing microorganisms [126]. The following sections will discuss these two biohydrogen production strategies in detail.

3.2.1. Biophotolysis

Biophotolysis can be classified into direct and indirect types and some of their examples have been listed in Table 5 [127]. During direct bio-photolysis, microalgae generate biohydrogen through light energy-driven water splitting (2H2O + light energy → 2H2 + O2) [128]. In brief, microalgae use photosystems II (PSII) to harvest light energy and split water into protons and electrons. These electrons flow linearly from water through two photosystems (PSII to PSI) to the hydrogen-producing enzyme, hydrogenase, under special conditions, via an electronic carrier (Ferredoxin). The hydrogenase enzyme is activated when the microalgae culture is exposed to anaerobic conditions. Finally, the hydrogenase enzyme catalyzes the reaction between protons and electrons to produce biohydrogen [129]. Gaffron and Rubin [130] first reported direct biophotolysis in 1942 while studying Scenedesmus obliquus. This phenomenon was later observed in other green microalgae species, such as Chlamydomonas reinhardtii, Chlorella fusca, Chlorella pyrenoidosa, Chlorella vulgaris, Chlorococcum littorale, Monoraphidium sp., Platymonas subcordiformis, and Tetraspora sp. [131]. The conversion of readily available substrates, solar energy, and water into biohydrogen makes direct biophotolysis promising. However, its practical application has been limited for the last 70 years due to the lack of efficient techniques to overcome the oxygen sensitivity of hydrogenase.
Even in small amounts (<2% v/v), oxygen generated as a by-product of PSII activity suppresses all hydrogenase-catalyzed reactions, allowing only transient biohydrogen production (lasting for a few minutes) from the direct biophotolysis [132]. To avoid oxygen inhibition of hydrogenase in Chlamydomonas reinhardtii, Reeves and Greenbaum [133] continuously purged inert gas in the system, extending the biohydrogen production for up to 160 h. Hamed et al. [134] reported a 2–10-fold increase in biohydrogen production from four microalgae species by purging the system with argon gas. Maswanna et al. [135,136] reported 6.5 times higher biohydrogen production under argon gas purging. However, continuous inert gas purging is an expensive and operationally impractical strategy for large-scale biohydrogen production systems. Pow and Krasna [137] trialed an alternative approach to remove the photosynthetically generated oxygen from the photobioreactor system. The authors used oxygen absorbers, such as Fieser’s reagent, 20% potassium hydroxide solution, sodium dithionite, and diuron. Although Fieser’s reagent, 20% potassium hydroxide solution, and diuron resulted in little to no significant biohydrogen production, sustained biohydrogen production for up to 6 h was reported when sodium dithionite was directly added to the microalgae culture. Paramesh and Chandrasekhar [138] tested the efficiency of three oxygen-scavenging agents (sodium sulfite, sodium metabisulfite, and sodium dithionate), and found that sodium sulfite was the most efficient agent in extending biohydrogen production. However, the direct addition of exogenous reductants to the algae culture may compromise the cell viability and biohydrogen production yield in the long term [139].
To combat the oxygen inhibition of hydrogenase and sustain biohydrogen production in microalgae during biophotolysis, a two-stage process (indirect biophotolysis) has been developed, separating hydrogen production activities from oxygen evolution [140]. In the first stage, cells undergo photosynthesis under aerobic conditions to fix carbon dioxide into biomass and release oxygen (6H2O + 6CO2 + light energy → C6H12O6 + 6O2). In the second stage, cells catabolize stored organic compounds under anaerobic conditions to produce hydrogen (C6H12O6 + 2H2O → 4H2 + 2CH3COOH + 2CO2 and 2CH3COOH + 4H2O + light energy → 8H2 + 4CO2). To induce anaerobiosis in the second stage, nutrient (sulfur, nitrogen, and phosphorus) deprivation has been extensively studied, where cells are first cultivated under nutrient-replete conditions (first stage) and then subjected to nutrient-deprived conditions in the second stage [141]. Daniel et al. [142], Maswanna et al. [135], and Pongpadung et al. [143] deprived the Coccomyxa chodatii, Tetraspora sp., and Chlorella sorokiniana cultures of sulfur and reported a 4-, 1.2-, and 69-fold increase in biohydrogen production, respectively. Sulfur is an essential component of amino acids, cysteine, and methionine, which play a vital role in protein synthesis during the PSII repair cycle [125]. When microalgae cells are deprived of sulfur, protein biosynthesis is inhibited, partially deactivating PSII and resulting in lower water-splitting activity. This slows down the oxygen evolution rate from PSII compared to the cell’s respirational oxygen consumption rate, and the cultures become anaerobic, enhancing the hydrogenase activity. Hence, the technique of inducing anaerobiosis through sulfur deprivation is an attractive option to sustain biohydrogen production from microalgae [126]. Besides sulfur, phosphorus deprivation also inhibits oxygen-evolving activity from PSII since phosphorus is an essential component of nucleic acids, including DNA (Deoxyribonucleic acid) and RNA (Ribonucleic acid) [144]. Maswanna et al. [135] reported a 1.1-fold increase in biohydrogen production when Tetraspora culture was deprived of sulfur and phosphorus compared to only sulfur deprivation. Moreover, phosphorus deprivation is a promising approach for enhancing biohydrogen production, especially from marine algae, as the sulfur deprivation strategy cannot be applied to seawater-based media due to its high sulfate concentrations (2649 mg/L) [145]. In addition to phosphorus and sulfur deprivation, nitrogen deprivation has been investigated to enhance biohydrogen production from microalgae. Pongpadung et al. [143], Hamed et al. [146], and Li et al. [147] deprived the Chlorella sorokiniana, Parachlorella kessleri, and Chlorella pyrenoidosa cultures of nitrogen and reported a 17, 1.2, and 8943-fold increase in biohydrogen production, respectively. Although sulfur, nitrogen, and phosphorus deprivation were the focus of many studies, magnesium and potassium deprivation have also exhibited promising results. Volgusheva et al. [148] compared the effect of sulfur and magnesium deprivation on Chlamydomonas reinhardtii cells and observed that magnesium-deprived cells produced 60% more biohydrogen. Tetraspora sp. generated 2.6 times higher biohydrogen under potassium deprivation conditions [149]. In addition to single-nutrient deprivation, some research studies used double- or multiple-nutrient stress to enhance biohydrogen production. Pongpadung et al. [143] exposed nitrogen-limited Chlorella sorokiniana culture to phosphorus and sulfur deprivation and reported 6.4-, 8.2-, and 10.4-times higher biohydrogen production when cultures were deprived of only sulfur, only phosphorus, and both sulfur and phosphorus, respectively.
Even though nutrient-deprived conditions seem promising in enhancing biohydrogen production from microalgae during two-stage growth, their overall impact can be limited by the amount of intracellular organic carbon present inside the autotrophically grown microalgae biomass [150]. During indirect biophotolysis, intracellular organic carbon compounds degrade to supply the electrons and protons required for biohydrogen production, linking the biohydrogen yield directly with the amount of intracellular organic compounds present in the biomass [125]. Photosynthetically fixed carbon may not be sufficient to produce desired levels of biohydrogen [150]. One approach to increasing the intracellular organic carbon content and the resulting electron supply is to utilize exogenic organic substrates [151]. Microalgae can be cultivated using hetero- and mixotrophic modes in addition to phototrophic cultivation. During phototrophic cultivation, cells harvest light energy and assimilate carbon dioxide, whereas in the absence of light (heterotrophic mode), cells utilize organic carbon as the sole energy source. In the mixotrophic method, both photosynthetic and heterotrophic metabolisms occur concurrently, and cells simultaneously assimilate both carbon dioxide and organic carbon sources in the presence of light. It has been reported that the mixotrophic mode of microalgae cultivation can yield higher biomass than the other two modes [152]. Liu et al. [150] demonstrated that the biohydrogen yield from sulfur-deprived Chlorella pyrenoidosa culture almost doubled from 65.6 to 121.1 mL/L with the addition of 0.7 g/L glucose, showing a positive correlation between biohydrogen production and glucose consumption by microalgae. Another study reported 2.7 times higher biohydrogen production from Chlorella vulgaris when the initial glucose concentration was increased from 5 to 10 g/L [153]. Biohydrogen production in Chlorella sp. was increased from 0 to 128 and 150 μmol/mg Chl (Chl: Chlorophyll) when glucose and acetate were supplied, respectively [154]. However, supplementing commercial carbon sources for biohydrogen production may be an economic burden for the overall process. Hence, it is essential to use alternate organic carbon sources that are cheaper and abundant. Sengmee et al. [155] supplemented Chlorella sp. culture with 16 g/L of crude glycerol (a byproduct of the biodiesel industry) and achieved 2.6 times higher biohydrogen production. Dudek et al. [156] used 50% diluted aerobically pretreated dairy wastewater to cultivate Tetraselmis subcordiformis and produced 1.3 times higher biohydrogen. These studies highlight the possibility of substituting commercial organic compounds with waste resources to enhance microalgae-based biohydrogen production.
Another challenge with the nutrient deprivation approach is that nutrient-deprived cells may suffer from acute oxidative stress, resulting in diminished biohydrogen production over time [157]. However, it is possible to establish an anaerobic environment without nutrient deprivation by adding acetate, using specific illumination protocols, and co-culturing microalgae with bacteria. Recently, Hwang et al. [158] used an acetate-rich fermenter effluent (at an acetate/Cl ratio of 150) as a natural PSII oxygen regulator in nutrient-replete Chlamydomonas reinhardtii and Chlorella sorokiniana cultures to demonstrate a continuous biohydrogen production process for 15 days. In another study, biohydrogen production was sustained for up to three days in Chlamydomonas reinhardtii culture without any nutrient deprivation when the culture was illuminated by a train of short pulses (1 s), followed by long dark periods (9 s) instead of continuous illumination [159]. Co-culturing microalgae with bacteria is an alternate approach to induce anaerobiosis through the bacterial partner’s respiration [160]. Ben et al. [161] cultured Chlamydomonas reinhardtii with different isolates of Pseudomonas sp. and found 12 times higher biohydrogen production compared to the pure algal culture. Shetty et al. [162] achieved a three-fold increase in biohydrogen production when 5% enriched microbial inoculum was added to the Chlorella vulgaris culture utilizing pretreated brewery effluent as a cultivation media. Although co-culturing microalgae with bacteria can benefit biohydrogen production, some bacteria may compete with microalgae for nutrients, negatively impacting microalgal growth [31]. This phenomenon was observed in the study conducted by Fakhimi and Tavakoli [163] in which Chlamydomonas reinhardtii was co-cultivated with Escherichia coli, Pseudomonas stutzeri, Pseudomonas putida, and an unknown bacterial consortium, resulting in up to 24, 46, 32, and 56%, higher biohydrogen production, respectively. However, in their study, P. stutzeri and P. putida negatively influenced the microalgae growth at all tested initial cell concentrations (optical density: 0.01–0.5), whereas E. coli reduced microalgae growth at higher concentrations (optical density >0.3). This highlights the importance of selecting appropriate bacterial cultures and cultivation conditions to benefit from algal-bacteria symbiosis in biohydrogen production. Biohydrogen production without nutrient deprivation is an emerging area of research in biophotolysis, and it should be further explored for direct and efficient biohydrogen production from microalgae.
Table 5. Effect of different strategies on biohydrogen production through microalgal biophotolysis.
Table 5. Effect of different strategies on biohydrogen production through microalgal biophotolysis.
Microalgae
Species
Biohydrogen Production
Strategy and Experimental Conditions
Biohydrogen ProductionBiohydrogen Production
Duration
References
Without
Pretreatment
With
Pretreatment
Oxygen removal by inert gas purging
Synechocystis sp. PCC 6803Direct photolysis with and without
anaerobic conditions
0.12–0.7 mmol/mg Chla.h1–4
mmol/mg Chla.h
168 h [134]
Parachlorella kessleri
EMCCN 3073
0.04–0.2 mmol/mg Chla.h0.4–1 mmol/mg Chla.h
Nostoc spongiaeforme0.1–0.8 mmol/mg Chla.h0.2–2.1 mmol/mg Chla.h
Nostoc sp.0.07–0.2 mmol/mg Chla.h0.2–0.9 mmol/mg Chla.h
Immobilized
Tetraspora sp.
CU2551 in sodium alginate beads
Two-stage growth (sulfur deprivation in the second stage) with and without
anaerobic conditions
182 nmol/mg DW.h1183 nmol/mg DW.h108 h
(Anaerobic);
1034 h
(Aerobic)
[135,136]
Oxygen removal by scavenging agents
Scenedesmus
obliquus 393
Direct photolysis with and without
sodium dithionite addition
≈0570 μL 6 h [137]
Chlorococcum
minutum
Direct photolysis with and without
sodium sulfite addition
NA
300 μmol 24 h [138]
Direct photolysis with and without
sodium metabisulfite addition
300 μmol
Direct photolysis with and without
sodium dithionite addition
135 μmol
Nutrient deprivation
Chlamydomonas
reinhardtii (CC425)
Two-stage growth with and without
sulfur deprivation in the second stage under anaerobic conditions
NA61 ± 7 mL/L
(17 ± 4 μmol/L.h)
204 h [141]
Chlamydomonas moewusii (SAG24.91)NA21 ± 3 mL/L
(5 ± 0.4 μmol/L.h)
Coccomyxa chodatii SAG 216–2Direct photolysis with and without
sulfur deprivation and malate
supplementation
<50 mL/L200 mL/L120 h [142]
Immobilized
Tetraspora sp.
CU2551 in sodium alginate beads
Two-stage growth with and without
sulfur deprivation in the second stage under aerobic conditions
0.46 mL per 25 mL of
medium
0.55 mL per 25 mL of
medium
48 h [135]
Two-stage growth with and without sulfur and phosphorus deprivation in the second stage under aerobic
conditions
0.61 mL per 25 mL of
medium
Chlorella sp. IOAC707STwo-stage growth with and without
phosphorus deprivation in the second stage under aerobic conditions
NAUp to 40 mL/L650 h [144]
Chlorella sorokiniana KU204Two-stage growth with and without
sulfur deprivation in the second stage under anaerobic conditions
0.7 mL/L48 mL/L84 h [143]
Two-stage growth with and without
sulfur and nitrogen deprivation in the second stage under anaerobic conditions
98 mL/L
Two-stage growth with and without
nitrogen deprivation in the second stage under anaerobic conditions
12 mL/L
Two-stage growth with and without
phosphorus and nitrogen deprivation in the second stage under anaerobic conditions
77 mL/L
Two-stage growth with and without
sulfur, nitrogen, and phosphorus deprivation in the second stage under anaerobic conditions
125 mL/L
Parachlorella kessleri
EMCCN 3073
Two-stage growth with and without
nitrogen deprivation in the second stage under anaerobic conditions
250 μL/L300 μL/L9 d [146]
Chlorella pyrenoidosa
IOAC707S
Two-stage growth with and without
nitrogen deprivation in the second stage under anaerobic conditions
0.003 mL/L 26.83 mL/L 92 h [147]
Chlamydomonas
Reinhardtii wild type strain 137+
Two-stage growth with and without
sulfur deprivation in the second stage under aerobic conditions
NA1.3 mmol/106 cells100 h [148]
Three-stage growth with and without
magnesium deprivation in the third stage under aerobic conditions
NA2.1 mmol/106 cells230 h
Tetraspora sp. CU2551Two-stage growth with and without
potassium deprivation in the second stage under anaerobic conditions
3.6  ±  0.1 μmol/mg DW9.2  ±  0.1 μmol/mg DW32 h [149]
Addition of exogenic substrates
Immobilized
Chlorella vulgaris in
sodium alginate beads
Two-stage growth (under purple light) with and without sulfur deprivation and exogenous organic carbon addition (10 g/L of glucose) in the second stage under anaerobic conditions NA60 mL/L
(39 mL/L.d)
50 h [151]
Immobilized
Scenedesmus obliquus in sodium alginate beads
NA128 mL/L
(205 mL/L.d)
70 h
Chlorella pyrenoidosaTwo-stage growth with and without sulfur deprivation and exogenous organic carbon addition (0.7 g/L of glucose) in the second stage under
anaerobic conditions
65.5 mL/L121.1 mL/L120 h [150]
Chlorella vulgarisDirect biophotolysis with and without exogenous organic carbon addition (5–10 g/L of glucose) under anaerobic
conditions
0 0.75–2 mL/h174 h [153]
Chlorella sp. KLSc59Two-stage growth with and without
exogenous organic carbon addition (glucose) in the second stage under anaerobic conditions
0128 μmol/mg Chl42 h [154]
Two-stage growth with and without
exogenous organic carbon addition (acetate) in the second stage under anaerobic conditions
150 μmol/mg Chl
Chlorella sp.Two-stage growth with and without
exogenous organic carbon addition (16 g/L of crude glycerol) in the second stage under anaerobic conditions
4 mL/L10.3 mL/L24 h [155]
Tetraselmis subcordiformisTwo-stage growth with and without
exogenous organic carbon addition (50% diluted aerobically pretreated dairy wastewater) in the second stage under anaerobic conditions
54 ± 2 mL/g DW69 ± 4 mL/g DW120 h [156]
Chlamydomonas reinhardtii
UTEX 2243
Supplementing acetate-rich fermenter effluent to achieve acetate/Cl ratio of 150 without nutrient deprivation under aerobic conditionsNA95 μmol/L15 d [158]
Chlorella sorokiniana
UTEX 2714
80 μmol/L
Light intensity manipulation
Chlamydomonas reinhardtiiPulse illumination (strong light pulse for 1 s, followed by dark period for 9 s) vs. continuous illumination under
anaerobic conditions
03 mmoL/L48 h [159]
Chlorella sp. KLSc59Illumination with and without light intensity of 7 μmol photons/m2.s5 μmol/mg Chl63 μmol/mg Chl60 h [154]
Illumination with and without light intensity of 14 μmol photons/m2.s130 μmol/mg Chl
Illumination with and without light intensity of 28 μmol photons/m2.s206 μmol/mg Chl
Algae–bacteria co-culture
Chlamydomonas reinhardtii
FACHB-265
Co-cultured with and without Pseudomonas sp. strain-D in sulfur-deprived media under aerobic conditions10 mL/L120 mL/L15 d [161]
Chlorella vulgaris MACC360Co-cultured with and without 5% enriched microbial consortium in pretreated brewery effluent media under aerobic conditions52 mL/L.d154 mL/L.d3 d [162]
Chlamydomonas reinhardtii
strain 704
Co-cultured with and without Escherichia coli K-12 MG1655 with acetic acid under aerobic and low light intensity conditionsNA24% higher5 d [163]
Co-cultured with and without Pseudomonas stutzeri A1501 with acetic acid under aerobic and low light intensity conditions46% higher
Co-cultured with and without Pseudomonas putida 12,264 with acetic acid under aerobic and low light 56 intensity conditions32% higher
Co-cultured with and without unknown bacterial consortium with acetic acid under aerobic and low light intensity conditions56% higher
Immobilization
Tetraspora sp.
CU2551
Two-stage growth with and without immobilized cells in the second stage under aerobic conditions0.0025 μmol/mg DW.h0.1
μmol/mg DW.h
NA [135]
Note: NA—not applicable, DW—dry weight, Chla—chlorophyll a, Chl—chlorophyll. Strategies used in each treatment group are highlighted.
Currently, biohydrogen yields from biophotolysis are low, ranging from 0.015 to 1.084 mmol/L.h (Table 5), which is not competitive with that of the fermentation approach (0.35 to 10.26 mmol/L.h) [22]. As a result, biophotolysis remains far from commercialization. However, understanding the underlying molecular mechanisms of biohydrogen production and genetically reengineering those metabolic pathways to achieve optimal biohydrogen productivity can accelerate its commercial implementation. The primary focus of genetic engineering is the development of oxygen-resistant hydrogenase, but this remains a significant challenge for the scientific community. Recent genetic engineering studies (Table 6) have shown that current approaches focus on redirecting electron pathways and promoting hydrogenase oxygen sensitivity. In addition to lower biohydrogen yields, capital and operational costs are additional issues for scaling up the biophotolysis process because separate photobioreactors are needed for biomass generation and biohydrogen production. This includes the energy-intensive process of centrifugation, which is conducted to exchange cultivation media and switch between oxygenic photosynthesis and biohydrogen production processes. Operational costs can be reduced by simplifying the media exchange process through microalgae immobilization. Some published studies have also reported that the immobilization of microalgae cells enhances their biohydrogen yield due to the self-shading-induced reduction in photosynthetic activity and subsequent oxygen evolution [135,151]. However, immobilization technology is still under development and warrants further research.

3.2.2. Fermentation

Fermentation can be categorized into photo and dark fermentation (Table 7). The photo fermentation (PF) process employs photosynthetic bacteria, mainly purple non-sulfur (PNS) bacteria, to convert organic compounds (acetate, butyrate, lactate) into biohydrogen and carbon dioxide using a light energy [140]. The ratio of biohydrogen to carbon dioxide produced during the PF process can vary depending on the type of substrate used [169]. PNS bacteria use a nitrogenase enzyme for simultaneous biohydrogen production and nitrogen fixation, making nitrogen-limited conditions a requirement for the evolution of biohydrogen. During this process, organic acids are first degraded using light energy to generate electrons and adenosine triphosphate (ATP) that drive nitrogenase to evolve biohydrogen and fix molecular nitrogen into ammonium ions [170]. Examples of the commonly used purple non-sulfur bacteria species for biohydrogen production are Rhodopseudomonas palustris, Rhodobacter capsulats, Rhodobacter sphaeroides, Rhodospirillum rubrum, and Rubrivivax gelatinosus [169]. However, PNS bacteria involved in the PF process can only consume small organic molecules, usually organic acids, requiring microalgae biomass to undergo hydrolysis before PF. Therefore, the PF process is most likely preceded by dark fermentation [171,172].
The dark fermentation (DF) process is similar to the AD process described in Section 3.1, in which organic substrates undergo hydrolysis, acidogenesis, and acetogenesis stages. However, the methanogenic stage is suppressed during DF. In the AD process, biohydrogen is typically produced as an intermediate metabolite during the acidogenesis and acetogenesis stages, and it is then converted into biomethane by hydrogen-utilizing methanotrophs. Unlike AD, during DF, special care is taken to stop the fermentation process at the acidogenesis–acetogenesis stage by inactivating methanotrophs from the initial inoculum [173]. Similar to AD, the major hurdle in DF is the poor biodegradability of microalgal biomass due to its recalcitrant cell wall [124]. Therefore, selecting a suitable pretreatment process before fermentation is vital to efficiently break down the cell wall and convert the complex carbohydrates into simpler monosaccharides. Pretreatment methods that could be applied for microalgal cell wall disruption and carbohydrate hydrolysis are discussed in Section 3.1.1, and their effect on biohydrogen production has been summarized in Table 7. DF exhibits a low biohydrogen production efficiency, but it can be drastically increased by pairing it with PF. In addition to biohydrogen, DF produces hydrogenic effluent containing a wide range of organic acids that PF can use for additional biohydrogen production, as reported in published studies [171,172]. However, the requirement to build customized anaerobic photobioreactors with large surface areas exposed to sunlight is a bottleneck in the combined DF–PF process, as it increases the cost and complexity of the system [174].
Table 7. Photo fermentation and dark fermentation of microalgae biomass for biohydrogen production.
Table 7. Photo fermentation and dark fermentation of microalgae biomass for biohydrogen production.
Microalgae
Species
PretreatmentFermentation
Microorganisms
Operating
Conditions
Biohydrogen Yield
(mL H2/g VS)
References
Photofermentation
Chlorella sp. Acid–hydrothermal treatment followed by dark fermentation Rhodobacter sphaeroides
TISTR 1952
Initial pH: 7
Light: 5000 lux
Inoculum size: 20% v/v
Temperature: 37 °C
125 [171]
Arthrospira platensisAcid–hydrothermal treatment followed by dark fermentation and NaCl-modified zeolite treatmentRhodopseudomonas palustrisInitial pH: 7
Light: 6000 lux
Inoculum size: NA
Temperature: 30 °C
333 [172]
Dark fermentation
Chlorella sp.1.5% v/v HCl, 180 °C, 0.25 h Heat-treated (105 °C for 3 h) anaerobic granules (dominated by Clostridium sp.) collected from brewery wastewater treatment plant Initial pH: 6
F/I: 1 (VS/VS)
Temperature: 37 °C
47 [171]
Arthrospira platensis1% v/v H2SO4, 135 °C, 0.25 h
Heat treated (100 °C, 0.5 h) anaerobic digestion sludge dominated by Clostridium sp.Initial pH: 6
F/I: NA
Temperature: 35 °C
96 [172]
Chlorella sp.4% v/v H2SO4, 2.5 h Heat-treated (105 °C for 3 h) anaerobic granules (dominated by Clostridium sp.) collected from brewery wastewater treatment plantInitial pH: 6
F/I: 3 (VS/VS)
Temperature: 35 °C
26 [173]
0.75% v/v H2SO4, 160 °C, 0.5 h54
Chlorella sp.No pretreatmentHeat treated (100 °C, 0.25 h) anaerobic sludge obtained from sewage treatment plant Initial pH: 6
F/I: NA
Temperature: 37 °C
8 [175]
Chlorella sp.No pretreatmentMixed anaerobic bacterial consortiaInitial pH: 7
F/I: NA
Temperature: 35 °C
22 [176]
Deoiled Scenedesmus obliquus UTEX 393No pretreatmentAcidogenic mixed consortia (dominated by Clostridium sp.) developed from heat-treated (100 °C for 0.34 h) cow dungInitial pH: 6.7
Inoculum size: 10% v/v
Temperature: 37 °C
10 [177]
Grinding15
Homogenization20
Autoclave30
Sonication36
1 N NaOH, 121 °C, 0.5 h40
1 N KOH, 121 °C, 0.5 h38
0.5 N H2SO4, 121 °C, 0.5 h89
10% w/v magnetic solid acid, 121 °C, 0.5 h53
Scenedesmus obtusiusculus
AT-UAM
No pretreatmentGranular sludge obtained from a full-scale up-flow anaerobic sludge blanket reactor fed with tequila vinassesInitial pH: 7.5
F/I: 12 (VS/VS)
Temperature: 37 °C
29 [178]
3% HCl, 100 °C, 1.7 h48
Algal bloom dominated by Microcystis sp. No pretreatmentHeat-treated (100 °C, 0.5 h) anaerobic digestion sludge dominated by Clostridium sp.Initial pH: 6
F/I: 0.5 (VS/VS)
Temperature: 35 °C
0.3 [179]
2% v/v H2SO4, 135 °C, 0.25 h (steam treatment)19
2% v/v H2SO4, 135 °C, 0.25 h
(hydrothermal treatment)
25
Wastewater-born microalgal biomassNo pretreatmentHeat-treated (90 °C, 0.5 h) anaerobic digestion sludge from sewage treatment plantInitial pH: 9.5
F/I: 1 (TS)
Temperature: 35 °C
18 [180]
240–530 mg NH3-N/L, 1 day, pH 9.5 (free ammonia pretreatment)20–22
Note: NA—not available, F/I—food/inoculum.

3.3. Alcoholic Fermentation

Bioethanol is the most extensively used renewable fuel for transportation and can be produced from microalgae using three pathways: dark fermentation (DF), photo fermentation (PF), and traditional fermentation (TF) [181]. Although DF and PF are similar process terminologies used to depict alcoholic fermentation and biohydrogen production pathways, the mechanisms involved in these pathways are different. Some examples of DF, PF, and TF for bioethanol production are listed in Table 8 and Table 9. In the DF process, microalgae species, such as Chlamydomonas reinhardtii, Chlamydomonas moewusii, Chlorococcum littorale, Chlorogonium elongatum, and Chlorella fusca, ferment intracellular polysaccharides into bioethanol under dark and anaerobic conditions [182,183,184]. Pyruvate, an intermediate compound, is generated through hydrolysis and glycolysis of intracellular polysaccharides (starch). Subsequently, pyruvate is converted into various end products, including acetate, ethanol, formate, glycerol, lactate, biohydrogen, and carbon dioxide, depending on the type of microalgae species and surrounding environmental conditions [185]. Studies on microalgae DF for bioethanol production reported low bioethanol yields of less than 2% w/w, as shown in Table 8. This could be attributed to the complex network of metabolic pathways involved in microalgal DF and difficulties associated with understanding and selectively manipulating those metabolic pathways to enhance bioethanol production. Therefore, the practical application of the DF pathway for bioethanol production has not received much attention.
Compared to the DF pathway, the PF pathway results in a more specific and efficient bioethanol production [186]. The PF pathway comprises two steps: photosynthesis and fermentation. During the first step of photosynthesis, inorganic carbon (carbon dioxide) is fixed into organic carbon (phosphoglycerate) through the Calvin cycle and later converted to pyruvate. In the second step, pyruvate is fermented into ethanol with the help of two key enzymes, pyruvate decarboxylase (pdc) and alcohol dehydrogenase II (adhII). pdc catalyzes the conversion of pyruvate into acetaldehyde and carbon dioxide through a nonoxidative decarboxylation reaction, whereas adhII oxidizes the resulting acetaldehyde into ethanol [187]. However, these enzymes are naturally missing or expressed in insufficient quantities in microalgae. Therefore, genetic engineering is used to heterologously express pdc and adhII genes in microalgae, preferably cyanobacteria, to enable direct bioethanol production [188]. Cyanobacteria have relatively well-characterized genetic backgrounds, demonstrate a high tolerance to foreign gene introduction, and exhibit amenability to genetic modifications [189]. Foreign DNA (Deoxyribonucleic acid) can be introduced into cyanobacteria under controlled conditions through shuttle vectors or by directly integrating it into the chromosome via targeted homologous recombination. A research study by Deng and Coleman [190] was the first to report the cyanobacteria Synechococcus elongatus PCC7942 strain as the platform of bioethanol production. The authors constructed a new S. elongatus strain using a shuttle vector pCB4, cloned from the coding sequences of pdc and adhII genes obtained from the bacterium Zymomonas mobilis. Following four weeks of culture, the transformed S. elongatus strain produced a bioethanol titer of 0.23 g/L. Later, Dexter and Fu [191] used the same two genes from Z. mobilis and integrated them into the chromosome of Synechocystis sp. PCC 6803 using a double homologous recombination system and produced a bioethanol titer of 0.46 g/L in six days of cultivation. Since then, several genetic engineering efforts have been made to improve the bioethanol yield from cyanobacteria while maintaining cell growth (Table 8). However, PF pathways produce much lower bioethanol yields (<6 g/L), making the ethanol separation process (distillation) too costly for large-scale applications. Poor bioethanol yields could be attributed to co-factor imbalance [192], low ethanol tolerance levels [193], competition for carbon usage between biomass synthesis and target product formation [194], and inefficient carbon fixation mechanisms [195,196]. However, there is still room for optimizing the bioethanol yield from cyanobacteria through alternate gene expression approaches.
Table 8. Dark fermentation and photo fermentation for microalgae-based bioethanol production.
Table 8. Dark fermentation and photo fermentation for microalgae-based bioethanol production.
Dark Fermentation
Microalgae
Species
Operating
Conditions
Starch
Content
(% of Dry Cell Weight)
% Starch
Decomposed
Bioethanol Yield
(% of Dry Cell Weight)
References
Chlamydomonas reinhardtii UTEX 2247Incubation under dark and anaerobic conditions at 25 °C for 46 h
Slurry concentration: 15% w/w
45NA1 [182]
Chlamydomonas sp. YA-SH-1Incubation under dark and anaerobic conditions at 30–35 °C for 44 h
Slurry concentration: 15–25% w/w
30NA1.3 [183]
Chlorococcum littoraleIncubation under dark and anaerobic conditions at 30 °C for 24 h
Slurry concentration: 1.4% w/w
15461.6 [184]
Photo Fermentation
CyanobacteriaGenes expressed (source of genes) and their expression mechanismPromotor usedGene deletion
(Effect)
Bioethanol titer (g/L)
and days of
cultivation (d)
Reference
Synechococcus elongatus PCC7942pdc (Zymomonas mobilis),
Shuttle vector;
adhII (Zymomonas mobilis),
Shuttle vector
rbcLSNA0.23 in 28 d [190]
Synechocystis
sp. PCC 6803
pdc (Zymomonas mobilis),
Homologous recombination;
adhII (Zymomonas mobilis),
Homologous recombination
psbA2NA0.46 in 6 d [191]
Synechocystis
sp. PCC6803
pdc (Zymomonas mobilis),
Homologous recombination;
adh, slr1192 (Endogenous overexpression),
Homologous recombination
PrbcphaA and phaB
(Disrupting PHB biosynthesis pathway)
5.5 in 26 d [197]
Synechocystis
sp. PCC6803
pdc (Zymomonas mobilis),
Homologous recombination;
adhII (Zymomonas mobilis),
Homologous recombination
nblAglgC
(Disrupting glycogen biosynthesis pathway) and
phaC + phaE (Disrupting PHB biosynthesis pathway)
3 in 3 d [194]
Synechocystis
sp. PCC6803
zwf (Endogenous
overexpression to enhance NADPH production)
Homologous recombination;
pdc (Zymomonas mobilis),
Homologous recombination;
yqhD, NADPH-dependent adh (Escherichia coli),
Homologous recombination
Pcpc560NA0.59 in 14 d [192]
Synechocystis
sp. PCC6803
pdc (Zymomonas mobilis),
Shuttle vector;
adh, slr1192 (Endogenous),
Shuttle vector
PnrsBNA0.45 in 7 d [196]
pdc (Zymomonas mobilis),
Shuttle vector;
adh, slr1192 (Endogenous),
Shuttle vector;
rbcSC, slr0009-slr0011-slr0012-FLAG with RuBisCO-encoding genes (Endogenous),
Shuttle vector
PnrsB (for pdc and adh) and psbA2 (for rbcSC, 70glpX, tktA, and fbaA)0.7 in 7 d
pdc (Zymomonas mobilis),
Shuttle vector;
adh, slr1192 (Endogenous),
Shuttle vector;
70glpX with FBP/SBPase-encoding genes (Synechococcus PCC 7002),
Shuttle vector
0.75 in 7 d
pdc (Zymomonas mobilis),
Shuttle vector;
adh, slr1192 (Endogenous),
Shuttle vector;
tktA, sll1070 with TK-encoding genes (Endogenous),
Shuttle vector
0.6 in 7 d
pdc (Zymomonas mobilis),
Shuttle vector;
adh, slr1192 (Endogenous),
Shuttle vector;
fbaA, sll0018 with FBA-encoding genes (Endogenous),
Shuttle vector
0.75 in 7 d
Synechocystis
sp. PCC6803
pdc (Zymomonas mobilis),
Shuttle vector;
adh, slr1192 (Endogenous),
Shuttle vector;
fbaA, sll0018 with FBA-encoding genes (Endogenous),
Shuttle vector;
tktA, sll1070 with TK-encoding genes (Endogenous),
Shuttle vector
PnrsB (for pdc and adh) and psbA2 (for tktA and fbaA)NA1.2 in 20 d [195]
Synechocystis
sp. PCC6803 (Fe2O3-treated culture)
pdc (Saccharomyces cerevisiae),
Shuttle vector;
adh (Endogenous),
Shuttle vector
psbA1NA4.9 in 25 d [198]
Synechocystis sp. PCC6803 (MgO-treated culture)5.1 in 25 d
Note: NA—not applicable, NADPH—nicotinamide adenine dinucleotide phosphate, pdc—pyruvate decarboxylase, adh—alcohol dehydrogenase, phaA—polyhydroxyalkanoate-specific β-ketothiolase, phaB—polyhydroxyalkanoate-specific acetoacetyl-CoA reductase, PHB—polyhydroxybutyrate, glgC—glucose-1-phosphate adenylyltransferase, phaC—polyhydroxyalkanoate synthase, phaE—polyhydroxyalkanoate polymerase subunit, zwf—glucose 6-phosphate dehydrogenase, RuBisCO (rbcSC)—ribulose-1,5-bisphosphate carboxylase/oxygenase, FBA (fbaA)—Fructose-1,6-bisphosphate aldolase, FBP/SBPase (70glpX)—fructose-1,6-/sedoheptulose-1,7-bisphosphatase, TK (tktA)—transketolase.
Traditional fermentation has been the most widely studied method of bioethanol production, as it typically yields higher bioethanol quantities (21–88% w/w (% of dry cell weight) and 5–43 g/L (based on the working volume)) compared to DF (<2% w/w) and PF (<6 g/L) (Table 8 and Table 9). In the traditional fermentation process, the carbohydrate content of microalgae biomass is used as a feedstock by ethanologenic microorganisms, such as yeast (Saccharomyces cerevisiae) and bacteria (Zymomonas mobilis). However, S. cerevisiae is more commonly used for bioethanol fermentation due to its tolerance towards low pH and high ethanol concentrations [199]. As described earlier in Section 3.1, microalgae biomass pretreatment is crucial before fermentation for easy access to intracellular microalgal compounds. Various biomass pretreatment methods are discussed in detail in Section 3.1.1, and the effects on bioethanol production have been summarized in Table 9. Shokrkar et al. [200] compared the effect of acidic and enzymatic pretreatments on the bioethanol production performance of a mixed microalgae culture and reported a 1.3 times higher bioethanol production when microalgae culture was pretreated with enzymes. De Farias Silva et al. [201] observed no significant variations between the acidic and enzymatic pretreatments for Chlorella vulgaris and Scenedesmus obliquus biomass. Another study demonstrated an alternate approach for bioethanol production without any acidic or enzymatic pretreatments [202]. The authors combined the extraction and fermentation process in which a lysozyme and calcium chloride mixture was used to extract glycogen from Arthrospira platensis. Extracted glycogen was simultaneously degraded to glucose with the help of a recombinant Saccharomyces cerevisiae culture, which produced alpha-amylase and glucoamylase. However, such an approach can only work for cyanobacteria, which lack robust cell wall structure. The examples listed in Table 9 confirm that the effectiveness of microalgae pretreatment varies depending on the species.
The traditional fermentation process can be divided into two groups: separate hydrolysis and fermentation (SHF) and simultaneous saccharification and fermentation (SSF) [203]. During SHF, hydrolysis and fermentation processes are conducted separately in different reactors, whereas, during SSF, hydrolysis and fermentation processes proceed simultaneously in the same reactor. The main advantage of the SHF process is the possibility to separately optimize the operating conditions of the hydrolysis and fermentation processes. El-Mekkawi et al. [21] used the response surface method (RSM) during SHF to optimize the process variables, such as algal biomass, yeast loading, and fermentation time, achieving a higher bioethanol concentration of 19 g/L. Other advantages of the SHF process are the potential use of cheaper chemicals, shorter residence time, and easy operation. However, its high capital cost is moving the research direction toward SSF. Kim et al. [204] compared the effect of SHF and SSF on bioethanol production from Phorphydium cruentum and observed that SSF produced slightly better bioethanol yields (74–80%) over SHF (70–78%). Similarly, Megawati et al. [205] observed a slightly better bioethanol production result with SSF (48.5%) compared to SHF (46%). Although most of the bioethanol production studies have used a single strain of yeast to study the fermentation process, some studies have applied a co-fermentation approach in which two or more different strains of yeast are used with a capacity to simultaneously degrade pentose and hexose sugar [201]. However, using a combination of different yeast strains would still require careful and thorough investigation.
Table 9. Pretreatment of microalgae biomass for bioethanol production through traditional fermentation.
Table 9. Pretreatment of microalgae biomass for bioethanol production through traditional fermentation.
Microalgae
Species
PretreatmentFermentation
Microorganisms
Fermentation Operating
Conditions
Bioethanol Concentration (g/L) and Yield
(% of Dry Cell Weight)
References
Separate hydrolysis and fermentation
Carotenoid-free Chromochloris zofingiensis SAG 211-14Autoclave (120 °C for 0.34 h) followed by two-stage enzymatic pretreatment with α-Amylase (90 °C, 2 h, 4.5 pH) and glucoamylase (60 °C, 22 h, 6.5 pH)Saccharomyces cerevisiae CCUG 53310Initial pH: 4.8
Inoculum size: NA
Temperature: 37 °C
Time:
25 ± 2% [206]
Autoclave (120 °C for 0.34 h) followed by three-stage enzymatic pretreatment with Cellic Ctec2 and Cellic Htec2 (45 °C, 48 h, 5 pH), α-Amylase (90 °C, 2 h, 5 pH) and glucoamylase (60 °C, 22 h, 5 pH)62 ± 2%
Carotenoid-free
Haematococcus pluvialis
SAG 192.80
Autoclave (120 °C for 0.34 h) followed by two-stage enzymatic pretreatment with α-Amylase (90 °C, 2 h, 4.5 pH) and glucoamylase (60 °C, 22 h, 6.5 pH)Saccharomyces cerevisiae CCUG 53310Initial pH: 4.8
Inoculum size: NA
Temperature: 37 °C
Time:
35 ± 0.3% [207]
Autoclave (120 °C for 0.34 h) followed by three-stage enzymatic pretreatment with Cellic Ctec2 and Cellic Htec2 (45 °C, 49 h, 5 pH), α-Amylase (90 °C, 2 h, 5 pH) and glucoamylase (60 °C, 22 h, 5 pH)88.1 ± 0.5%
Chlorella vulgaris1 N HCl, 90 °C, 1 hSaccharomyces cerevisiaeInitial pH: 5
Inoculum size: 3% v/v
Temperature: 30 °C
Time:
46% [205]
Chlorella vulgaris FSP-E2% H2SO4, 121 °C, 0.34 hSaccharomyces cerevisiae
FAY-1
Initial pH: NA
Inoculum size: NA
Temperature: 30 °C
21% (43 g/L) [208]
Mixed microalgae consortium0.5 M H2SO4 and 2.5% (w/v) MgSO4 at 121 °C, 0.67 hSaccharomyces cerevisiae
ATCC 7921
Initial pH: 6.5
Inoculum size: 3% v/v
Temperature: 30 °C
5 g/L [200]
Three-stage enzymatic pretreatment with β-glucosidase/cellulase (65 °C, 3 h), α-amylase (95 °C, 3 h) and amyloglucosidase (55 °C, 3 h) 6.4 g/L
Arthrospira platensis NIES-391 g/L lysozyme and 100 mM CaCl2 Saccharomyces cerevisiae
strain BY4741 AASS/GASS
Initial pH:
5.2–5.4
Inoculum size: 5% v/v
Temperature: 38–40 °C
32% [202]
Wastewater-grown microalgae biomass dominated by Microcystis 0.5 N H2SO4, 120 °C, 4 h Immobilized
Saccharomyces cereviciae ATCC 4126
Initial pH:
4.5
Inoculum size: 15% v/v
Temperature: 30 °C
19 g/L [21]
Porphyridium cruentum KMMCC-1061 One-stage enzymatic hydrolysis with pectinase and cellulase (37 °C, 7 h, 4.8 pH)Saccharomyces cerevisiae KCTC 7906Initial pH:
4.5
Inoculum size: 0.1% w/v
Temperature: 37 °C
70–78% (based on initial glucose content) [204]
Simultaneous Saccharification and Fermentation
Chlorella vulgarisTwo-stage enzymatic pretreatment with α-Amylase (90 °C, 6 pH) and glucoamylase (80 °C, 5 h, 6 pH)Saccharomyces cerevisiaeInitial pH: 5
Inoculum size: 3% v/v
Temperature: 30 °C
49% [205]
Porphyridium cruentum KMMCC-1061 One-stage enzymatic hydrolysis with pectinase and cellulase (37 °C, 10 h, 4.8 pH)Saccharomyces cerevisiae KCTC 7906Initial pH:
4.5
Inoculum size: 0.1% w/v
Temperature: 37 °C
74–80% (based on initial glucose content) [204]
Co-fermentation
Dried and milled Chlorella vulgaris biomass powder Neoalgae® (Micro seaweed products B-52501749). 3% H2SO4, 120 °C, 0.5 h75% Saccharomyces cerevisiae Cameo S.p.A. and 25% Pichia stipitis ATCC 58, 785 Initial pH:
5–6
Inoculum size: 7.5 g/L
Temperature: 30 °C
49 ± 5% [201]
One-stage enzymatic hydrolysis with Viscozyme® L, AMG 300 L, and Pectinex Ultra SP-L (50 °C, 4 h, 5 pH)49 ± 0.5%
Scenedesmus obliquus
SAG 276.7
3% H2SO4, 120 °C, 0.5 h87 ± 6%
Ultrasonication followed by One-stage enzymatic hydrolysis with Viscozyme® L, AMG 300 L, and Pectinex Ultra SP-L (50 °C, 8 h, 5 pH)41 ± 1.5%
Note: NA—Not available.

4. Challenges and Future Prospects

With rising energy demand and the gradual fossil fuel depletion, it is necessary to develop renewable and sustainable alternative energy sources, including bioenergy. Among the different available biomass resources, microalgae biomass possesses comparative advantages due to its high growth rate and ability to grow using flue gas and waste as a nutrient source. Despite its advantages, microalgae-based bioenergy is not yet commercialized. One of the main challenges is the high production cost of microalgae biomass, making it less promising than other renewable and non-renewable energy feedstocks. However, research efforts are underway to reduce microalgae biomass production costs by using waste cultivation media, redesigning photobioreactors, and developing cost-effective harvesting techniques [209].
The second challenge is the low energy recovery efficiency. Single biochemical conversion process often exhibits low energy recovery efficiencies, even under optimal processing conditions. This could be attributed to the complex biochemical composition of microalgae biomass (as shown in Table 1), limiting the energy recovered by a single biomass conversion process. To fully exploit the microalgae biomass and extract a maximum amount of bioenergy, a biorefinery framework can be applied to produce multiple bioenergy products simultaneously [210]. Table 10 lists examples of studies that have combined two or more biomass conversion processes to generate multiple bioenergy products. Such approaches help recover the maximum possible bioenergy from microalgae biomass while simplifying the downstream handling of those bioenergy products. For instance, studies have combined biohydrogen and biomethane production to increase energy recovery [171,177]. Such sequential production of biohydrogen and biomethane makes it feasible to produce biohythane (a mixture of 90–75% biomethane and 10–25% biohydrogen) on site, which can be stored and transported using the existing natural gas infrastructure, avoiding the limitation of building dedicated storage and transport infrastructures for pure hydrogen energy [211]. However, such integrated bioenergy production processes are still in their early stage of development. Although integrated processes can maximize bioenergy production from microalgal biomass, their complexity and excessive energy requirements to execute certain steps (such as biomass pretreatment and photo fermentation) will limit their commercial applicability. Further investigations on enhancing energy efficiency and simplifying the procedures, such as the development of novel energy-saving biomass pretreatment methods, high-performance photobioreactors for photo fermentation utilizing solar energy, biogas upgrading through carbon dioxide fixation during microalgae cultivation using anaerobic digestate, etc., remain to be conducted in future studies to make this integrated process more energetically and economically feasible for industrial applications.
In addition to bioenergy, microalgae biomass can also be used to produce high-value-added bioproducts, such as pigments (astaxanthin, phycocyanin, lutein, and β-carotene), polyunsaturated fatty acids (Docosahexaenoic acid and Eicosapentaenoic acid), and protein supplements [212]. Co-production of high-value-added products with bioenergy can significantly reduce the overall cost of microalgae-based biorefineries. As shown in Table 10, Mirzaei et al. [206] and Hosseini et al. [207] successfully demonstrated the co-production of astaxanthin with bioenergy (biomethane and bioethanol) using Chromochloris zofingiensis and Haematococcus pluvialis, respectively. However, techno-economical and life cycle analyses must be conducted to determine the best possible scenarios for microalgae biomass valorization in a biorefinery, concurrently producing bioenergy and high-value-added bioproducts. It is necessary to continue building new relevant solutions based on the experiences from recent advancements and challenges encountered to fully exploit the microalgae biomass and balance the sustainability aspect of microalgal biotechnology with economic gains.
Even though bioenergy production from microalgae biomass has increasingly exhibited promising results at the laboratory and pilot scale, studies focusing on process optimization and the industrial scale-up are scarce. Optimizing process parameters for biochemical processes is complex because it involves many permutations and combinations of operating conditions [213]. Moreover, the composition of feedstock can influence the bioenergy yield. Conventionally, trial and error or one variable at a time (OVAT) analyses are conducted to decipher the correlation between output (bioenergy yield) and input (governing factors) variables. However, these analyses involve time, cost, and labor-intensive laboratory studies. To overcome the shortcomings of conventional strategies, theory-driven (hypothesis-driven) models are developed by deriving empirical judgments from multiple experiments. However, theory-driven models often fail to accurately predict the outcomes for the bioenergy systems due to their complex and non-linear nature [214]. With the emergence of artificial intelligence (AI) tools, it is now possible to identify and use the patterns in available datasets to predict the outcome for a new input without conducting detailed laboratory studies [215]. Machine learning models (data-driven models), such as artificial neural networks, random forests, support vector machines, multilinear regression, and decision trees, have been successfully applied in microalgae biomass conversions technologies such as pyrolysis [23,216], gasification [217], hydrothermal liquefaction [218] and biological hydrogen production [219] for the prediction and optimization of the bioenergy yield. Despite their success, machine learning-assisted predictions in the bioenergy field are still in the initial stages of development. More developments in machine learning studies are required to expand the overall understanding of microalgae biomass conversion processes and obtain new insights to improve the bioenergy yield. More research should be conducted to improve the interpretability and predictability of machine learning models by developing high-quality datasets to test and apply novel machine learning algorithms, promoting the application of state-of-the-art algorithms such as multi-view and deep learning, and integrating theory-driven models.
In addition to technological breakthroughs, bioenergy cannot replace fossil fuels without significant policy changes. Government policies such as taxation for greenhouse gas emissions, subsidies for bioenergy production, and incentives for bioenergy utilization may help to relieve the cost pressure to some extent. It is still too early to predict which of the many developments summarized in this article will succeed on a large scale. However, when we contemplate the myriad of possibilities from microalgae biomass exploitation and the likelihood of continued “crises” arising from non-renewable energy usage, there is little doubt that this field will be vital in shaping the future of new and clean energy technologies. The current literature review on microalgae-based bioenergy production indicates that long-term research and development plans are required to translate laboratory studies into sustainable real-scale facilities.

5. Conclusions

This article reviewed various biochemical conversion technologies, such as anaerobic digestion, biohydrogen production (direct biophotolysis, indirect biophotolysis, photo fermentation, and dark fermentation), and alcoholic fermentation (microalgae-catalyzed photo fermentation, microalgae-catalyzed dark fermentation, and traditional fermentation by ethanologenic microorganisms) for biomethane, biohydrogen, and bioethanol production, respectively. Compared to other biochemical conversion processes, anaerobic digestion and traditional alcoholic fermentation are simple, easy to operate, and more technically advanced technologies that can pave the way for commercializing microalgae-based bioenergy. Nevertheless, the high cost of microalgae biomass production and low energy recovery efficiencies are the major bottlenecks in anaerobic digestion and traditional alcoholic fermentation technologies. To reduce biomass production costs and improve energy recovery, future research should focus on cultivating microalgae using waste resources, designing efficient photobioreactor systems, and developing cost-effective biomass harvesting and pretreatment technologies. Artificial intelligence tools can be used to accelerate process optimization and scale-up. In addition, a biorefinery approach must be explored to fully exploit the microalgae biomass and produce high-value-added products with bioenergy. Techno-economical and life cycle analyses must be conducted to determine the best scenarios for microalgae valorization in an integrated biorefinery. Along with continuous research and development efforts, changes in government policies are also needed to incentivize bioenergy production and consumption.

Author Contributions

Conceptualization, S.K.P.; resources, Y.W.T.; writing—original draft preparation, S.K.P., Z.T. and J.Z.E.W.; writing—review and editing, S.K.P. and Y.W.T.; supervision, Y.W.T.; project administration, Y.W.T.; funding acquisition, Y.W.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by The National Research Foundation, Prime Minister’s Office, Singapore, under its Campus for Research Excellence and Technological Enterprise (CREATE) program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yilanci, V.; Haouas, I.; Ozgur, O.; Sarkodie, S.A. Energy diversification and economic development in emergent countries: Evidence from fourier function-driven bootstrap panel causality test. Front. Energy Res. 2021, 9, 632712. [Google Scholar] [CrossRef]
  2. International Energy Agency. World Energy Outlook 2022; OECD Publishing: Paris, France, 2022. [Google Scholar]
  3. Friedlingstein, P.; O’Sullivan, M.; Jones, M.W.; Andrew, R.M.; Gregor, L.; Hauck, J.; Le Quéré, C.; Luijkx, I.T.; Olsen, A.; Peters, G.P.; et al. Global Carbon Budget 2022. Earth Syst. Sci. Data 2022, 14, 4811–4900. [Google Scholar] [CrossRef]
  4. U.S. Energy Information Administration (EIA). International Energy Outlook 2021 Narrative; U.S. Department of Energy: Washington, DC, USA, 2021. [Google Scholar]
  5. Al-Bawwat, A.a.K.; Jurado, F.; Gomaa, M.R.; Cano, A. Availability and the possibility of employing wastes and biomass materials energy in Jordan. Sustainability 2023, 15, 5879. [Google Scholar] [CrossRef]
  6. Yadav, K.K.; Krishnan, S.; Gupta, N.; Prasad, S.; Amin, M.A.; Cabral-Pinto, M.M.S.; Sharma, G.K.; Marzouki, R.; Jeon, B.-H.; Kumar, S.; et al. Review on evaluation of renewable bioenergy potential for sustainable development: Bright future in energy practice in India. ACS Sustain. Chem. Eng. 2021, 9, 16007–16030. [Google Scholar] [CrossRef]
  7. Wang, Y.; Guan, W.; Liu, L.; Ma, X. Biomass energy consumption and carbon neutrality in OECD countries: Testing pollution haven hypothesis and environmental Kuznets curve. Front. Environ. Sci. 2022, 10, 1691. [Google Scholar] [CrossRef]
  8. Duarah, P.; Haldar, D.; Patel, A.K.; Dong, C.-D.; Singhania, R.R.; Purkait, M.K. A review on global perspectives of sustainable development in bioenergy generation. Bioresour. Technol. 2022, 348, 126791. [Google Scholar] [CrossRef] [PubMed]
  9. Al-Bawwat, A.a.K.; Cano, A.; Gomaa, M.R.; Jurado, F. Availability of biomass and potential of nanotechnologies for bioenergy production in Jordan. Processes 2023, 11, 992. [Google Scholar] [CrossRef]
  10. Maliha, A.; Abu-Hijleh, B. A review on the current status and post-pandemic prospects of third-generation biofuels. Energy Syst. 2022, 1–32. [Google Scholar] [CrossRef]
  11. Malode, S.J.; Prabhu, K.K.; Mascarenhas, R.J.; Shetti, N.P.; Aminabhavi, T.M. Recent advances and viability in biofuel production. Energy Convers. Manag. X 2021, 10, 100070. [Google Scholar] [CrossRef]
  12. International Energy Agency. Renewables 2022: Analysis and Forecast to 2027; OECD Publishing: Paris, France, 2022. [Google Scholar]
  13. Ahmed, S.; Warne, T.; Smith, E.; Goemann, H.; Linse, G.; Greenwood, M.; Kedziora, J.; Sapp, M.; Kraner, D.; Roemer, K.; et al. Systematic review on effects of bioenergy from edible versus inedible feedstocks on food security. NPJ. Sci. Food 2021, 5, 9. [Google Scholar] [CrossRef]
  14. Jeswani, H.K.; Chilvers, A.; Azapagic, A. Environmental sustainability of biofuels: A review. Proc. Math. Phys. Eng. Sci. 2020, 476, 20200351. [Google Scholar] [CrossRef]
  15. Olabi, A.G.; Shehata, N.; Sayed, E.T.; Rodriguez, C.; Anyanwu, R.C.; Russell, C.; Abdelkareem, M.A. Role of microalgae in achieving sustainable development goals and circular economy. Sci. Total Environ. 2023, 854, 158689. [Google Scholar] [CrossRef]
  16. Elalami, D.; Oukarroum, A.; Barakat, A. Anaerobic digestion and agronomic applications of microalgae for its sustainable valorization. RSC Adv. 2021, 11, 26444–26462. [Google Scholar] [CrossRef]
  17. Sharma, P.K.; Saharia, M.; Srivstava, R.; Kumar, S.; Sahoo, L. Tailoring microalgae for efficient biofuel production. Front. Mar. Sci. 2018, 5, 382. [Google Scholar] [CrossRef]
  18. Parakh, S.K.; Praveen, P.; Loh, K.C.; Tong, Y.W. Integrating gravity settler with an algal membrane photobioreactor for in situ biomass concentration and harvesting. Bioresour. Technol. 2020, 315, 123822. [Google Scholar] [CrossRef] [PubMed]
  19. Sajjadi, B.; Chen, W.-Y.; Raman, A.A.A.; Ibrahim, S. Microalgae lipid and biomass for biofuel production: A comprehensive review on lipid enhancement strategies and their effects on fatty acid composition. Renew. Sustain. Energy Rev. 2018, 97, 200–232. [Google Scholar] [CrossRef]
  20. Zabed, H.M.; Akter, S.; Yun, J.; Zhang, G.; Zhang, Y.; Qi, X. Biogas from microalgae: Technologies, challenges and opportunities. Renew. Sustain. Energy Rev. 2020, 117, 109503. [Google Scholar] [CrossRef]
  21. El-Mekkawi, S.A.; Abdo, S.M.; Samhan, F.A.; Ali, G.H. Optimization of some fermentation conditions for bioethanol production from microalgae using response surface method. Bull. Natl. Res. Cent. 2019, 43, 164. [Google Scholar] [CrossRef]
  22. 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, 753878. [Google Scholar] [CrossRef]
  23. Mustapha, S.I.; Mohammed, U.A.; Rawat, I.; Bux, F.; Isa, Y.M. Production of high-quality pyrolytic bio-oils from nutrient-stressed Scenedesmus obliquus microalgae. Fuel 2023, 332, 126299. [Google Scholar] [CrossRef]
  24. Raheem, A.; Abbasi, S.A.; Mangi, F.H.; Ahmed, S.; He, Q.; Ding, L.; Memon, A.A.; Zhao, M.; Yu, G. Gasification of algal residue for synthesis gas production. Algal. Res. 2021, 58, 102411. [Google Scholar] [CrossRef]
  25. Shahi, T.; Beheshti, B.; Zenouzi, A.; Almasi, M. Bio-oil production from residual biomass of microalgae after lipid extraction: The case of Dunaliella sp. Biocatal. Agric. Biotechnol. 2020, 23, 101494. [Google Scholar] [CrossRef]
  26. Gomaa, M.R.; Mustafa, R.J.; Al-Dmour, N. Solar thermochemical conversion of carbonaceous materials into syngas by Co-Gasification. J. Cleaner Prod. 2020, 248, 119185. [Google Scholar] [CrossRef]
  27. Hossain, S.M.Z. Biochemical conversion of microalgae biomass into biofuel. Chem. Eng. Technol. 2019, 42, 2594–2607. [Google Scholar] [CrossRef]
  28. Magdalena, J.A.; Ballesteros, M.; González-Fernandez, C. Efficient anaerobic digestion of microalgae biomass: Proteins as a key macromolecule. Molecules 2018, 23, 1098. [Google Scholar] [CrossRef]
  29. Maity, S.; Mallick, N. Trends and advances in sustainable bioethanol production by marine microalgae: A critical review. J. Cleaner Prod. 2022, 345, 131153. [Google Scholar] [CrossRef]
  30. 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, 660399. [Google Scholar] [CrossRef]
  31. Parakh, S.K.; Praveen, P.; Loh, K.-C.; Tong, Y.W. Wastewater treatment and microbial community dynamics in a sequencing batch reactor operating under photosynthetic aeration. Chemosphere 2019, 215, 893–903. [Google Scholar] [CrossRef]
  32. Niccolai, A.; Chini Zittelli, G.; Rodolfi, L.; Biondi, N.; Tredici, M.R. Microalgae of interest as food source: Biochemical composition and digestibility. Algal. Res. 2019, 42, 101617. [Google Scholar] [CrossRef]
  33. Parakh, S.K. Enhancing the Sustainability of Microalgae Production through Novel Photobioreactor Engineering and Harvesting Strategies; National University of Singapore: Singapore, 2019. [Google Scholar]
  34. Córdova, O.; Passos, F.; Chamy, R. Physical pretreatment methods for improving microalgae anaerobic biodegradability. Appl. Biochem. Biotechnol. 2018, 185, 114–126. [Google Scholar] [CrossRef]
  35. Barkia, I.; Saari, N.; Manning, S.R. Microalgae for high-value products towards human health and nutrition. Mar. Drugs 2019, 17, 304. [Google Scholar] [CrossRef] [PubMed]
  36. Alishah Aratboni, H.; Rafiei, N.; Garcia-Granados, R.; Alemzadeh, A.; Morones-Ramírez, J.R. Biomass and lipid induction strategies in microalgae for biofuel production and other applications. Microb. Cell Fact. 2019, 18, 178. [Google Scholar] [CrossRef] [PubMed]
  37. Arif, M.; Bai, Y.; Usman, M.; Jalalah, M.; Harraz, F.A.; Al-Assiri, M.S.; Li, X.; Salama, E.-S.; Zhang, C. Highest accumulated microalgal lipids (polar and non-polar) for biodiesel production with advanced wastewater treatment: Role of lipidomics. Bioresour. Technol. 2020, 298, 122299. [Google Scholar] [CrossRef]
  38. Mimouni, V.; Couzinet-Mossion, A.; Ulmann, L.; Wielgosz-Collin, G. Chapter 5-Lipids from microalgae. In Microalgae in Health and Disease Prevention; Levine, I.A., Fleurence, J., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 109–131. [Google Scholar]
  39. Chhandama, M.V.L.; Satyan, K.B.; Changmai, B.; Vanlalveni, C.; Rokhum, S.L. Microalgae as a feedstock for the production of biodiesel: A review. Bioresour. Technol. Rep. 2021, 15, 100771. [Google Scholar] [CrossRef]
  40. Bligh, M.; Nguyen, N.; Buck-Wiese, H.; Vidal-Melgosa, S.; Hehemann, J.-H. Structures and functions of algal glycans shape their capacity to sequester carbon in the ocean. Curr. Opin. Chem. Biol. 2022, 71, 102204. [Google Scholar] [CrossRef] [PubMed]
  41. Udayan, A.; Pandey, A.K.; Sirohi, R.; Sreekumar, N.; Sang, B.-I.; Sim, S.J.; Kim, S.H.; Pandey, A. Production of microalgae with high lipid content and their potential as sources of nutraceuticals. Phytochem. Rev. 2022, 1–28. [Google Scholar] [CrossRef]
  42. 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, 126304. [Google Scholar] [CrossRef] [PubMed]
  43. Bruce, A.; Heald, R.; Johnson, A.; Morgan, D.; Raff, M.; Roberts, K.; Walter, P. Molecular Biology of the Cell, 7th ed.; W. W. Norton & Company: New York, NY, USA, 2022. [Google Scholar]
  44. de Oliveira, M.C.; Bassin, I.D.; Cammarota, M.C. Microalgae and cyanobacteria biomass pretreatment methods: A comparative analysis of chemical and thermochemical pretreatment methods aimed at methane production. Fermentation 2022, 8, 497. [Google Scholar] [CrossRef]
  45. Shokravi, Z.; Shokravi, H.; Chyuan, O.H.; Lau, W.J.; Koloor, S.S.R.; Petrů, M.; Ismail, A.F. Improving ‘lipid productivity’ in microalgae by bilateral enhancement of biomass and lipid contents: A review. Sustainability 2020, 12, 9083. [Google Scholar] [CrossRef]
  46. Zaparoli, M.; Ziemniczak, F.G.; Mantovani, L.; Costa, J.A.V.; Colla, L.M. Cellular stress conditions as a strategy to increase carbohydrate productivity in Spirulina platensis. BioEnergy Res. 2020, 13, 1221–1234. [Google Scholar] [CrossRef]
  47. Ghosh, A.; Sarkar, S.; Gayen, K.; Bhowmick, T.K. Effects of carbon, nitrogen, and phosphorus supplements on growth and biochemical composition of Podohedriella sp. (MCC44) isolated from northeast India. Environ. Prog. Sustain. Energy 2020, 39, e13378. [Google Scholar] [CrossRef]
  48. Bibi, F.; Yasmin, H.; Jamal, A.; Al-Harbi, M.S.; Ahmad, M.; Zafar, M.; Ahmad, B.; Samra, B.N.; Ahmed, A.F.; Ali, M.I. Deciphering role of technical bioprocess parameters for bioethanol production using microalgae. Saudi J. Biol. Sci. 2021, 28, 7595–7606. [Google Scholar] [CrossRef] [PubMed]
  49. Jaiswal, K.K.; Banerjee, I.; Singh, D.; Sajwan, P.; Chhetri, V. Ecological stress stimulus to improve microalgae biofuel generation: A review. Octa. J. Biosci. 2020, 8, 48–54. [Google Scholar]
  50. Nzayisenga, J.C.; Farge, X.; Groll, S.L.; Sellstedt, A. Effects of light intensity on growth and lipid production in microalgae grown in wastewater. Biotechnol. Biofuels 2020, 13, 4. [Google Scholar] [CrossRef]
  51. Yun, C.-J.; Hwang, K.-O.; Han, S.-S.; Ri, H.-G. The effect of salinity stress on the biofuel production potential of freshwater microalgae Chlorella vulgaris YH703. Biomass Bioenergy 2019, 127, 105277. [Google Scholar] [CrossRef]
  52. Choudhary, S.; Tripathi, S.; Poluri, K.M. Microalgal-based bioenergy: Strategies, prospects, and sustainability. Energy Fuels 2022, 36, 14584–14612. [Google Scholar] [CrossRef]
  53. Shin, Y.S.; Jeong, J.; Nguyen, T.H.T.; Kim, J.Y.H.; Jin, E.; Sim, S.J. Targeted knockout of phospholipase A2 to increase lipid productivity in Chlamydomonas reinhardtii for biodiesel production. Bioresour. Technol. 2019, 271, 368–374. [Google Scholar] [CrossRef]
  54. Muñoz, C.F.; Weusthuis, R.A.; D’Adamo, S.; Wijffels, R.H. Effect of single and combined expression of lysophosphatidic acid acyltransferase, glycerol-3-phosphate acyltransferase, and diacylglycerol acyltransferase on lipid accumulation and composition in Neochloris oleoabundans. Front. Plant Sci. 2019, 10, 1573. [Google Scholar] [CrossRef]
  55. Brar, A.; Kumar, M.; Soni, T.; Vivekanand, V.; Pareek, N. Insights into the genetic and metabolic engineering approaches to enhance the competence of microalgae as biofuel resource: A review. Bioresour. Technol. 2021, 339, 125597. [Google Scholar] [CrossRef]
  56. Olguín, E.J.; Sánchez-Galván, G.; Arias-Olguín, I.I.; Melo, F.J.; González-Portela, R.E.; Cruz, L.; De Philippis, R.; Adessi, A. Microalgae-based biorefineries: Challenges and future trends to produce carbohydrate enriched biomass, high-added value products and bioactive compounds. Biology 2022, 11, 1146. [Google Scholar] [CrossRef]
  57. Paul, R.; Silkina, A.; Melville, L.; Suhartini, S.; Sulu, M. Optimization of ultrasound pretreatment of microalgal biomass for effective biogas production through anaerobic digestion process. Energies 2023, 16, 553. [Google Scholar] [CrossRef]
  58. Schmid, B.; Navalho, S.; Schulze, P.S.C.; Van De Walle, S.; Van Royen, G.; Schüler, L.M.; Maia, I.B.; Bastos, C.R.V.; Baune, M.-C.; Januschewski, E.; et al. Drying microalgae using an industrial solar dryer: A biomass quality assessment. Foods 2022, 11, 1873. [Google Scholar] [CrossRef] [PubMed]
  59. Workie, E.; Kumar, V.; Bhatnagar, A.; He, Y.; Dai, Y.; Wah Tong, Y.; Peng, Y.; Zhang, J.; Fu, C. Advancing the bioconversion process of food waste into methane: A systematic review. Waste Manag. 2023, 156, 187–197. [Google Scholar] [CrossRef] [PubMed]
  60. Abanades, S.; Abbaspour, H.; Ahmadi, A.; Das, B.; Ehyaei, M.A.; Esmaeilion, F.; El Haj Assad, M.; Hajilounezhad, T.; Jamali, D.H.; Hmida, A.; et al. A critical review of biogas production and usage with legislations framework across the globe. Int. J. Environ. Sci. Technol. 2022, 19, 3377–3400. [Google Scholar] [CrossRef] [PubMed]
  61. Richard, E.N.; Hilonga, A.; Machunda, R.L.; Njau, K.N. A review on strategies to optimize metabolic stages of anaerobic digestion of municipal solid wastes towards enhanced resources recovery. Sustain. Environ. Res. 2019, 29, 36. [Google Scholar] [CrossRef]
  62. Lim, J.W.; Park, T.; Tong, Y.W.; Yu, Z. Chapter One: The microbiome driving anaerobic digestion and microbial analysis. In Advances in Bioenergy; Li, Y., Khanal, S.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; Volume 5, pp. 1–61. [Google Scholar]
  63. Zhang, L.; Yan, M.; Tsui, T.-H.; Lee, J.T.E.; Loh, K.-C.; Dai, Y.; Tong, Y.W. Chapter 16-Functional microbial characteristics in acidogenic fermenters of organic wastes for production of volatile fatty acids. In Biomass, Biofuels, Biochemicals; Pandey, A., Tong, Y.W., Zhang, L., Zhang, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 367–394. [Google Scholar]
  64. Detman, A.; Bucha, M.; Treu, L.; Chojnacka, A.; Pleśniak, Ł.; Salamon, A.; Łupikasza, E.; Gromadka, R.; Gawor, J.; Gromadka, A.; et al. Evaluation of acidogenesis products’ effect on biogas production performed with metagenomics and isotopic approaches. Biotechnol. Biofuels 2021, 14, 125. [Google Scholar] [CrossRef]
  65. Meegoda, J.N.; Li, B.; Patel, K.; Wang, L.B. A review of the processes, parameters, and optimization of anaerobic digestion. Int. J. Environ. Res. Public Health 2018, 15, 2224. [Google Scholar] [CrossRef]
  66. Nguyen, P.-D.; Tran, N.-S.T.; Nguyen, T.-T.; Dang, B.-T.; Le, M.-T.T.; Bui, X.-T.; Mukai, F.; Kobayashi, H.; Ngo, H.H. Long-term operation of the pilot scale two-stage anaerobic digestion of municipal biowaste in Ho Chi Minh City. Sci. Total Environ. 2021, 766, 142562. [Google Scholar] [CrossRef]
  67. Tiong, Y.W.; Sharma, P.; Tian, H.; Tsui, T.-H.; Lam, H.T.; Tong, Y.W. Startup performance and microbial communities of a decentralized anaerobic digestion of food waste. Chemosphere 2023, 318, 137937. [Google Scholar] [CrossRef]
  68. Paulo, L.M.; Castilla-Archilla, J.; Ramiro-Garcia, J.; Escamez-Picón, J.A.; Hughes, D.; Mahony, T.; Murray, M.; Wilmes, P.; O’Flaherty, V. Microbial community redundancy and resilience underpins high-rate anaerobic treatment of dairy-processing wastewater at ambient temperatures. Front. Bioeng. Biotechnol. 2020, 8, 192. [Google Scholar] [CrossRef]
  69. Heitkamp, K.; Latorre-Pérez, A.; Nefigmann, S.; Gimeno-Valero, H.; Vilanova, C.; Jahmad, E.; Abendroth, C. Monitoring of seven industrial anaerobic digesters supplied with biochar. Biotechnol. Biofuels 2021, 14, 185. [Google Scholar] [CrossRef] [PubMed]
  70. Golueke, C.G.; Oswald, W.J.; Gotaas, H.B. Anaerobic digestion of algae. Appl. Microbiol. 1957, 5, 47–55. [Google Scholar] [CrossRef] [PubMed]
  71. Machado, L.; Carvalho, G.; Pereira, R.N. Effects of innovative processing methods on microalgae cell wall: Prospects towards digestibility of protein-rich biomass. Biomass 2022, 2, 80–102. [Google Scholar] [CrossRef]
  72. Dunker, S.; Wilhelm, C. Cell wall structure of coccoid green algae as an important trade-off between biotic interference mechanisms and multidimensional cell growth. Front. Microbiol. 2018, 9, 719. [Google Scholar] [CrossRef] [PubMed]
  73. Mussgnug, J.H.; Klassen, V.; Schlüter, A.; Kruse, O. Microalgae as substrates for fermentative biogas production in a combined biorefinery concept. J. Biotechnol. 2010, 150, 51–56. [Google Scholar] [CrossRef] [PubMed]
  74. Ras, M.; Lardon, L.; Bruno, S.; Bernet, N.; Steyer, J.-P. Experimental study on a coupled process of production and anaerobic digestion of Chlorella vulgaris. Bioresour. Technol. 2011, 102, 200–206. [Google Scholar] [CrossRef]
  75. Tg, I.; Haq, I.; Kalamdhad, A.S. 14-Factors affecting anaerobic digestion for biogas production: A review. In Advanced Organic Waste Management; Hussain, C., Hait, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 223–233. [Google Scholar]
  76. Shahbaz, M.; Ammar, M.; Korai, R.M.; Ahmad, N.; Ali, A.; Khalid, M.S.; Zou, D.; Li, X. Impact of C/N ratios and organic loading rates of paper, cardboard and tissue wastes in batch and CSTR anaerobic digestion with food waste on their biogas production and digester stability. SN Appl. Sci. 2020, 2, 1436. [Google Scholar] [CrossRef]
  77. Solé-Bundó, M.; Passos, F.; Romero-Güiza, M.S.; Ferrer, I.; Astals, S. Co-digestion strategies to enhance microalgae anaerobic digestion: A review. Renew. Sustain. Energy Rev. 2019, 112, 471–482. [Google Scholar] [CrossRef]
  78. Kendir, E.; Ugurlu, A. A comprehensive review on pretreatment of microalgae for biogas production. Int. J. Energy Res. 2018, 42, 3711–3731. [Google Scholar] [CrossRef]
  79. Agarwalla, A.; Komandur, J.; Mohanty, K. Current trends in the pretreatment of microalgal biomass for efficient and enhanced bioenergy production. Bioresour. Technol. 2023, 369, 128330. [Google Scholar] [CrossRef]
  80. Rokicka, M.; Zieliński, M.; Dudek, M.; Dębowski, M. Effects of ultrasonic and microwave pretreatment on lipid extraction of microalgae and methane production from the residual extracted biomass. BioEnergy Res. 2021, 14, 752–760. [Google Scholar] [CrossRef]
  81. Dębowski, M.; Kazimierowicz, J.; Świca, I.; Zieliński, M. Ultrasonic disintegration to improve anaerobic digestion of microalgae with hard cell walls-Scenedesmus sp. and Pinnularia sp. Plants 2023, 12, 53. [Google Scholar] [CrossRef] [PubMed]
  82. Hassan, S.S.; Williams, G.A.; Jaiswal, A.K. Emerging technologies for the pretreatment of lignocellulosic biomass. Bioresour. Technol. 2018, 262, 310–318. [Google Scholar] [CrossRef] [PubMed]
  83. Krishnamoorthy, A.; Rodriguez, C.; Durrant, A. Sustainable approaches to microalgal pre-treatment techniques for biodiesel production: A review. Sustainability 2022, 14, 9953. [Google Scholar] [CrossRef]
  84. Passos, F.; Carretero, J.; Ferrer, I. Comparing pretreatment methods for improving microalgae anaerobic digestion: Thermal, hydrothermal, microwave and ultrasound. Chem. Eng. J. 2015, 279, 667–672. [Google Scholar] [CrossRef]
  85. Gruber-Brunhumer, M.R.; Jerney, J.; Zohar, E.; Nussbaumer, M.; Hieger, C.; Bochmann, G.; Schagerl, M.; Obbard, J.P.; Fuchs, W.; Drosg, B. Acutodesmus obliquus as a benchmark strain for evaluating methane production from microalgae: Influence of different storage and pretreatment methods on biogas yield. Algal Res. 2015, 12, 230–238. [Google Scholar] [CrossRef]
  86. Straessner, R.; Nikolausz, M.; Silve, A.; Nazarova, N.; Wuestner, R.; Papachristou, I.; Akaberi, S.; Leber, K.; Mueller, G.; Frey, W. Holistic exploitation of pulsed electric field (PEF)-treated and lipid extracted microalgae Auxenochlorella protothecoides, utilizing anaerobic digestion (AD). Algal Res. 2023, 69, 102950. [Google Scholar] [CrossRef]
  87. Wang, M.; Lee, E.; Dilbeck, M.P.; Liebelt, M.; Zhang, Q.; Ergas, S.J. Thermal pretreatment of microalgae for biomethane production: Experimental studies, kinetics and energy analysis. J. Chem. Technol. Biotechnol. 2017, 92, 399–407. [Google Scholar] [CrossRef]
  88. Ayala-Cortés, A.; Arcelus-Arrillaga, P.; Millan, M.; Arancibia-Bulnes, C.A.; Valadés-Pelayo, P.J.; Villafán-Vidales, H.I. Solar integrated hydrothermal processes: A review. Renew. Sustain. Energy Rev. 2021, 139, 110575. [Google Scholar] [CrossRef]
  89. Xiao, C.; Liao, Q.; Fu, Q.; Huang, Y.; Chen, H.; Zhang, H.; Xia, A.; Zhu, X.; Reungsang, A.; Liu, Z. A solar-driven continuous hydrothermal pretreatment system for biomethane production from microalgae biomass. Appl. Energy 2019, 236, 1011–1018. [Google Scholar] [CrossRef]
  90. Xiao, C.; Liao, Q.; Fu, Q.; Huang, Y.; Xia, A.; Shen, W.; Chen, H.; Zhu, X. Exergy analyses of biogas production from microalgae biomass via anaerobic digestion. Bioresour. Technol. 2019, 289, 121709. [Google Scholar] [CrossRef] [PubMed]
  91. Ziegler-Devin, I.; Chrusciel, L.; Brosse, N. Steam explosion pretreatment of lignocellulosic biomass: A mini-review of theoretical and experimental approaches. Front. Chem. 2021, 9, 860. [Google Scholar] [CrossRef] [PubMed]
  92. Martín Juárez, J.; Riol Pastor, E.; Fernández Sevilla, J.M.; Muñoz Torre, R.; García-Encina, P.A.; Bolado Rodríguez, S. Effect of pretreatments on biogas production from microalgae biomass grown in pig manure treatment plants. Bioresour. Technol. 2018, 257, 30–38. [Google Scholar] [CrossRef]
  93. Marques, A.d.L.; Pinto, F.P.; Araujo, O.Q.d.F.; Cammarota, M.C. Assessment of methods to pretreat microalgal biomass for enhanced biogas production. J. Sustain. Dev. Energy Water Environ. Syst. 2018, 6, 394–404. [Google Scholar] [CrossRef]
  94. Bai, X.; Lant, P.A.; Jensen, P.D.; Astals, S.; Pratt, S. Enhanced methane production from algal digestion using free nitrous acid pre-treatment. Renew. Energy 2016, 88, 383–390. [Google Scholar] [CrossRef]
  95. Fu, J.; Yan, B.; Gui, S.; Fu, Y.; Xia, S. Anaerobic co-digestion of thermo-alkaline pretreated microalgae and sewage sludge: Methane potential and microbial community. J. Environ. Sci. 2023, 127, 133–142. [Google Scholar] [CrossRef] [PubMed]
  96. Solé-Bundó, M.; Carrère, H.; Garfí, M.; Ferrer, I. Enhancement of microalgae anaerobic digestion by thermo-alkaline pretreatment with lime (CaO). Algal Res. 2017, 24, 199–206. [Google Scholar] [CrossRef]
  97. Caporgno, M.P.; Olkiewicz, M.; Pruvost, J.; Lepine, O.; Legrand, J.; Font, J.; Bengoa, C. A novel pre-treatment for the methane production from microalgae by using N-methylmorpholine-N-oxide (NMMO). Bioresour. Technol. 2016, 201, 370–373. [Google Scholar] [CrossRef]
  98. M’Arimi, M.M.; Mecha, C.A.; Kiprop, A.K.; Ramkat, R. Recent trends in applications of advanced oxidation processes (AOPs) in bioenergy production: Review. Renew. Sustain. Energy Rev. 2020, 121, 109669. [Google Scholar] [CrossRef]
  99. Song, K.; Li, Z.; Li, L.; Zhao, X.; Deng, M.; Zhou, X.; Xu, Y.; Peng, L.; Li, R.; Wang, Q. Methane production from peroxymonosulfate pretreated algae biomass: Insights into microbial mechanisms, microcystin detoxification and heavy metal partitioning behavior. Sci. Total Environ. 2022, 834, 155500. [Google Scholar] [CrossRef]
  100. Cardeña, R.; Moreno, G.; Bakonyi, P.; Buitrón, G. Enhancement of methane production from various microalgae cultures via novel ozonation pretreatment. Chem. Eng. J. 2017, 307, 948–954. [Google Scholar] [CrossRef]
  101. Li, L.; Li, Z.; Song, K.; Gu, Y.; Gao, X. Improving methane production from algal sludge based anaerobic digestion by co-pretreatment with ultrasound and zero-valent iron. J. Cleaner Prod. 2020, 255, 120214. [Google Scholar] [CrossRef]
  102. Wang, Q.; Sun, J.; Liu, S.; Gao, L.; Zhou, X.; Wang, D.; Song, K.; Nghiem, L.D. Free ammonia pretreatment improves anaerobic methane generation from algae. Water Res. 2019, 162, 269–275. [Google Scholar] [CrossRef] [PubMed]
  103. Kendir Çakmak, E.; Ugurlu, A. Enhanced biogas production of red microalgae via enzymatic pretreatment and preliminary economic assessment. Algal Res. 2020, 50, 101979. [Google Scholar] [CrossRef]
  104. Zhang, Y.; Caldwell, G.S.; Sallis, P.J. Renewable energy: Evaluation of low energy demand pre-treatments to optimise methane production from microalgae. IET Renew. Power Gener. 2019, 13, 1701–1710. [Google Scholar] [CrossRef]
  105. Sakhuja, D.; Ghai, H.; Rathour, R.K.; Kumar, P.; Bhatt, A.K.; Bhatia, R.K. Cost-effective production of biocatalysts using inexpensive plant biomass: A review. 3 Biotech 2021, 11, 280. [Google Scholar] [CrossRef]
  106. Bolivar, J.M.; Woodley, J.M.; Fernandez-Lafuente, R. Is enzyme immobilization a mature discipline? Some critical considerations to capitalize on the benefits of immobilization. Chem. Soc. Rev. 2022, 51, 6251–6290. [Google Scholar] [CrossRef]
  107. Hom-Diaz, A.; Passos, F.; Ferrer, I.; Vicent, T.; Blánquez, P. Enzymatic pretreatment of microalgae using fungal broth from Trametes versicolor and commercial laccase for improved biogas production. Algal Res. 2016, 19, 184–188. [Google Scholar] [CrossRef]
  108. Kavitha, S.; Yukesh Kannah, R.; Rajesh Banu, J.; Kaliappan, S.; Johnson, M. Biological disintegration of microalgae for biomethane recovery-prediction of biodegradability and computation of energy balance. Bioresour. Technol. 2017, 244, 1367–1375. [Google Scholar] [CrossRef]
  109. Aydin, S.; Yıldırım, E.; Ince, O.; Ince, B. Rumen anaerobic fungi create new opportunities for enhanced methane production from microalgae biomass. Algal Res. 2017, 23, 150–160. [Google Scholar] [CrossRef]
  110. Zhang, Y.; Caldwell, G.S.; Blythe, P.T.; Zealand, A.M.; Li, S.; Edwards, S.; Xing, J.; Goodman, P.; Whitworth, P.; Sallis, P.J. Co-digestion of microalgae with potato processing waste and glycerol: Effect of glycerol addition on methane production and the microbial community. RSC Adv. 2020, 10, 37391. [Google Scholar] [CrossRef] [PubMed]
  111. Wágner, D.S.; Radovici, M.; Smets, B.F.; Angelidaki, I.; Valverde-Pérez, B.; Plósz, B.G. Harvesting microalgae using activated sludge can decrease polymer dosing and enhance methane production via co-digestion in a bacterial-microalgal process. Algal Res 2016, 20, 197–204. [Google Scholar] [CrossRef]
  112. Garoma, T.; Nguyen, D. Anaerobic co-digestion of microalgae Scenedesmus sp. and TWAS for biomethane Production. Water Environ. Res 2016, 88, 13–20. [Google Scholar] [CrossRef] [PubMed]
  113. Solé-Bundó, M.; Salvadó, H.; Passos, F.; Garfí, M.; Ferrer, I. Strategies to optimize microalgae conversion to biogas: Co-digestion, pretreatment and hydraulic retention time. Molecules 2018, 23, 2096. [Google Scholar] [CrossRef]
  114. Ferreira, L.O.; Astals, S.; Passos, F. Anaerobic co-digestion of food waste and microalgae in an integrated treatment plant. J. Chem. Technol. Biotechnol. 2022, 97, 1545–1554. [Google Scholar] [CrossRef]
  115. Panyaping, K.; Khiewwijit, R.; Wongpankamol, P. Enhanced biogas production potential of microalgae and swine wastewater using co-digestion and alkaline pretreatment. Water Sci. Technol. 2018, 78, 92–102. [Google Scholar] [CrossRef]
  116. Zhang, Y.; Caldwell, G.S.; Zealand, A.M.; Sallis, P.J. Anaerobic co-digestion of microalgae Chlorella vulgaris and potato processing waste: Effect of mixing ratio, waste type and substrate to inoculum ratio. Biochem. Eng. J. 2019, 143, 91–100. [Google Scholar] [CrossRef]
  117. Carminati, P.; Gusmini, D.; Pizzera, A.; Catenacci, A.; Parati, K.; Ficara, E. Biogas from mono- and co-digestion of microalgal biomass grown on piggery wastewater. Water Sci. Technol. 2018, 78, 103–113. [Google Scholar] [CrossRef]
  118. Solé-Bundó, M.; Garfí, M.; Ferrer, I. Pretreatment and co-digestion of microalgae, sludge and fat oil and grease (FOG) from microalgae-based wastewater treatment plants. Bioresour. Technol. 2020, 298, 122563. [Google Scholar] [CrossRef]
  119. Vassalle, L.; Passos, F.; Rosa-Machado, A.T.; Moreira, C.; Reis, M.; Pascoal de Freitas, M.; Ferrer, I.; Mota, C.R. The use of solar pre-treatment as a strategy to improve the anaerobic biodegradability of microalgal biomass in co-digestion with sewage. Chemosphere 2022, 286, 131929. [Google Scholar] [CrossRef]
  120. Mahdy, A.; Fotidis, I.A.; Mancini, E.; Ballesteros, M.; González-Fernández, C.; Angelidaki, I. Ammonia tolerant inocula provide a good base for anaerobic digestion of microalgae in third generation biogas process. Bioresour. Technol. 2017, 225, 272–278. [Google Scholar] [CrossRef]
  121. Avila, R.; Carrero, E.; Vicent, T.; Blánquez, P. Integration of enzymatic pretreatment and sludge co-digestion in biogas production from microalgae. Waste Manag. 2021, 124, 254–263. [Google Scholar] [CrossRef]
  122. Tsapekos, P.; Kougias, P.G.; Alvarado-Morales, M.; Kovalovszki, A.; Corbière, M.; Angelidaki, I. Energy recovery from wastewater microalgae through anaerobic digestion process: Methane potential, continuous reactor operation and modelling aspects. Biochem. Eng. J. 2018, 139, 1–7. [Google Scholar] [CrossRef]
  123. Wu, D.; Li, L.; Zhao, X.; Peng, Y.; Yang, P.; Peng, X. Anaerobic digestion: A review on process monitoring. Renew. Sustain. Energy Rev. 2019, 103, 1–12. [Google Scholar] [CrossRef]
  124. 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, 122871. [Google Scholar] [CrossRef]
  125. 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, 121972. [Google Scholar] [CrossRef]
  126. Li, S.; Li, F.; Zhu, X.; Liao, Q.; Chang, J.-S.; Ho, S.-H. Biohydrogen production from microalgae for environmental sustainability. Chemosphere 2022, 291, 132717. [Google Scholar] [CrossRef] [PubMed]
  127. Touloupakis, E.; Faraloni, C.; Benavides, A.M.S.; Torzillo, G. Recent achievements in microalgal photobiological hydrogen production. Energies 2021, 14, 7170. [Google Scholar] [CrossRef]
  128. Jiménez-Llanos, J.; Ramírez-Carmona, M.; Rendón-Castrillón, L.; Ocampo-López, C. Sustainable biohydrogen production by Chlorella sp. microalgae: A review. Int. J. Hydrog. Energy 2020, 45, 8310–8328. [Google Scholar] [CrossRef]
  129. 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, 121747. [Google Scholar] [CrossRef]
  130. Gaffron, H.; Rubin, J. Fermentative and photochemical production of hydrogen In algae. J. Gen. Physiol. 1942, 26, 219–240. [Google Scholar] [CrossRef] [PubMed]
  131. 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, 2000124. [Google Scholar] [CrossRef] [PubMed]
  132. Ghirardi, M.L.; Togasaki, R.K.; Seibert, M. Oxygen sensitivity of algal H2- production. Appl. Biochem. Biotechnol. 1997, 63–65, 141–151. [Google Scholar] [CrossRef] [PubMed]
  133. Reeves, M.; Greenbaum, E. Long-term endurance and selection studies in hydrogen and oxygen photoproduction by Chlamydomonas reinhardtii. Enzym. Microb. Technol. 1985, 7, 169–174. [Google Scholar] [CrossRef]
  134. Hamed, S.M.; Raut, M.P.; Jaffé, S.R.P.; Wright, P.C. Evaluation of the effect of aerobic–anaerobic conditions on photohydrogen and chlorophyll a production by environmental Egyptian cyanobacterial and green algal species. Int. J. Hydrog. Energy 2017, 42, 6567–6577. [Google Scholar] [CrossRef]
  135. Maswanna, T.; Lindblad, P.; Maneeruttanarungroj, C. Improved biohydrogen production by immobilized cells of the green alga Tetraspora sp. CU2551 incubated under aerobic condition. J. Appl. Phycol. 2020, 32, 2937–2945. [Google Scholar] [CrossRef]
  136. Maswanna, T.; Phunpruch, S.; Lindblad, P.; Maneeruttanarungroj, C. Enhanced hydrogen production by optimization of immobilized cells of the green alga Tetraspora sp. CU2551 grown under anaerobic condition. Biomass Bioenergy 2018, 111, 88–95. [Google Scholar] [CrossRef]
  137. Pow, T.; Krasna, A.I. Photoproduction of hydrogen from water in hydrogenase-containing algae. Arch. Biochem. Biophys. 1979, 194, 413–421. [Google Scholar] [CrossRef]
  138. Paramesh, K.; Chandrasekhar, T. Improvement of photobiological hydrogen production in Chlorococcum minutum using various oxygen scavengers. Int. J. Hydrog. Energy 2020, 45, 7641–7646. [Google Scholar] [CrossRef]
  139. Márquez-Reyes, L.A.; Sánchez-Saavedra, M.d.P.; Valdez-Vazquez, I. Improvement of hydrogen production by reduction of the photosynthetic oxygen in microalgae cultures of Chlamydomonas gloeopara and Scenedesmus obliquus. Int. J. Hydrog. Energy 2015, 40, 7291–7300. [Google Scholar] [CrossRef]
  140. Sung, Y.J.; Yu, B.S.; Yang, H.E.; Kim, D.H.; Lee, J.Y.; Sim, S.J. Microalgae-derived hydrogen production towards low carbon emissions via large-scale outdoor systems. Bioresour. Technol. 2022, 364, 128134. [Google Scholar] [CrossRef] [PubMed]
  141. Vargas, S.R.; Santos, P.V.d.; Giraldi, L.A.; Zaiat, M.; Calijuri, M.d.C. Anaerobic phototrophic processes of hydrogen production by different strains of microalgae Chlamydomonas sp. FEMS Microbiol. Lett. 2018, 365, fny073. [Google Scholar] [CrossRef] [PubMed]
  142. Danial, A.W.; Abdel-Basset, R.; Abdel-Kader, H.A.A. Tuning photosynthetic oxygen for hydrogen evolution in synergistically integrated, sulfur deprived consortia of Coccomyxa chodatii and Rhodobium gokarnense at dim and high light. Photosynth. Res. 2023, 155, 203–218. [Google Scholar] [CrossRef] [PubMed]
  143. Pongpadung, P.; Zhang, L.; Sathasivam, R.; Yokthongwattana, K.; Niran, J.; Jianguo, L. Stimulation of hydrogen photoproduction in Chlorella sorokiniana subjected to simultaneous nitrogen limitation and sulfur- and/or phosphorus-deprivation. J. Pure. Appl. Microbiol. 2018, 12, 1719–1727. [Google Scholar] [CrossRef]
  144. Batyrova, K.; Gavrisheva, A.; Ivanova, E.; Liu, J.; Tsygankov, A. Sustainable hydrogen photoproduction by phosphorus-deprived marine green microalgae Chlorella sp. Int. J. Mol. Sci. 2015, 16, 2705–2716. [Google Scholar] [CrossRef]
  145. Jeong, S.; Naidu, G.; Leiknes, T.; Vigneswaran, S. 4.3 Membrane biofouling: Biofouling assessment and reduction strategies in seawater reverse osmosis desalination. In Comprehensive Membrane Science and Engineering, 2nd ed.; Drioli, E., Giorno, L., Fontananova, E., Eds.; Elsevier: Oxford, UK, 2017; pp. 48–71. [Google Scholar]
  146. Hamed, S.M.; Kapoore, R.V.; Raut, M.P.; Vaidyanathan, S.; Wright, P.C. Influence of nutrient status on the biohydrogen and lipid productivity in Parachlorella kessleri: A biorefinery approach. Appl. Microbiol. Biotechnol. 2020, 104, 10293–10305. [Google Scholar] [CrossRef]
  147. Li, L.; Zhang, L.; Liu, J. Proteomic analysis of hydrogen production in Chlorella pyrenoidosa under nitrogen deprivation. Algal Res. 2021, 53, 102143. [Google Scholar] [CrossRef]
  148. Volgusheva, A.A.; Jokel, M.; Allahverdiyeva, Y.; Kukarskikh, G.P.; Lukashev, E.P.; Lambreva, M.D.; Krendeleva, T.E.; Antal, T.K. Comparative analyses of H2 photoproduction in magnesium- and sulfur-starved Chlamydomonas reinhardtii cultures. Physiol. Plant. 2017, 161, 124–137. [Google Scholar] [CrossRef]
  149. Pewnual, T.; Jampapetch, N.; Saladtook, S.; Raksajit, W.; Klinsalee, R.; Maneeruttanarungroj, C. Response of green alga Tetraspora sp. CU2551 under potassium deprivation: A new promising strategy for hydrogen production. J. Appl. Phycol. 2022, 34, 811–819. [Google Scholar] [CrossRef]
  150. Liu, J.-Z.; Ge, Y.-M.; Sun, J.-Y.; Chen, P.; Addy, M.; Huo, S.-H.; Li, K.; Cheng, P.-F.; Ruan, R. Exogenic glucose as an electron donor for algal hydrogenases to promote hydrogen photoproduction by Chlorella pyrenoidosa. Bioresour. Technol. 2019, 289, 121762. [Google Scholar] [CrossRef]
  151. 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 Expr. 2020, 10, 191. [Google Scholar] [CrossRef]
  152. Lacroux, J.; Seira, J.; Trably, E.; Bernet, N.; Steyer, J.-P.; van Lis, R. Mixotrophic growth of Chlorella sorokiniana on acetate and butyrate: Interplay between substrate, C:N ratio and pH. Front. Microbiol. 2021, 12, 703614. [Google Scholar] [CrossRef] [PubMed]
  153. Alalayah, W.M.; Al-Zahrani, A.; Edris, G.; Demirbas, A. Kinetics of biological hydrogen production from green microalgae Chlorella vulgaris using glucose as initial substrate. Energy Sources Part A 2017, 39, 1210–1215. [Google Scholar] [CrossRef]
  154. Sirawattanamongkol, T.; Maswanna, T.; Maneeruttanarungroj, C. A newly isolated green alga Chlorella sp. KLSc59: Potential for biohydrogen production. J. Appl. Phycol. 2020, 32, 2927–2936. [Google Scholar] [CrossRef]
  155. Sengmee, D.; Cheirsilp, B.; Suksaroge, T.T.; Prasertsan, P. Biophotolysis-based hydrogen and lipid production by oleaginous microalgae using crude glycerol as exogenous carbon source. Int. J. Hydrog. Energy 2017, 42, 1970–1976. [Google Scholar] [CrossRef]
  156. Dudek, M.; Dębowski, M.; Kazimierowicz, J.; Zieliński, M.; Quattrocelli, P.; Nowicka, A. The cultivation of biohydrogen-producing Tetraselmis subcordiformis microalgae as the third stage of dairy wastewater aerobic treatment system. Sustainability 2022, 14, 12085. [Google Scholar] [CrossRef]
  157. Khosravitabar, F. Microalgal biohydrogen photoproduction: Scaling up challenges and the ways forward. J. Appl. Phycol. 2020, 32, 277–289. [Google Scholar] [CrossRef]
  158. Hwang, J.-H.; Lee, M.; Kang, E.H.; Lee, W.H. Renewable algal photo H2 production without S control using acetate enriched fermenter effluents. Int. J. Hydrog. Energy 2021, 46, 1740–1751. [Google Scholar] [CrossRef]
  159. Kosourov, S.; Jokel, M.; Aro, E.-M.; Allahverdiyeva, Y. A new approach for sustained and efficient H2 photoproduction by Chlamydomonas reinhardtii. Energy Environ. Sci. 2018, 11, 1431–1436. [Google Scholar] [CrossRef]
  160. Fakhimi, N.; Gonzalez-Ballester, D.; Fernández, E.; Galván, A.; Dubini, A. Algae-bacteria consortia as a strategy to enhance H2 production. Cells 2020, 9, 1353. [Google Scholar] [CrossRef]
  161. Ban, S.; Lin, W.; Wu, F.; Luo, J. Algal-bacterial cooperation improves algal photolysis-mediated hydrogen production. Bioresour. Technol. 2018, 251, 350–357. [Google Scholar] [CrossRef] [PubMed]
  162. Shetty, P.; Boboescu, I.Z.; Pap, B.; Wirth, R.; Kovács, K.L.; Bíró, T.; Futó, Z.; White, R.A.; Maróti, G. Exploitation of algal-bacterial consortia in combined biohydrogen generation and wastewater treatment. Front. Energy Res. 2019, 7, 52. [Google Scholar] [CrossRef]
  163. Fakhimi, N.; Tavakoli, O. Improving hydrogen production using co-cultivation of bacteria with Chlamydomonas reinhardtii microalga. Mater. Sci. Energy Technol. 2019, 2, 1–7. [Google Scholar] [CrossRef]
  164. Liu, P.; Ye, D.M.; Chen, M.; Zhang, J.; Huang, X.H.; Shen, L.L.; Xia, K.K.; Xu, X.J.; Xu, Y.C.; Guo, Y.L.; et al. Scaling-up and proteomic analysis reveals photosynthetic and metabolic insights toward prolonged H(2) photoproduction in Chlamydomonas hpm91 mutant lacking proton gradient regulation 5 (PGR5). Photosynth. Res. 2022, 154, 397–411. [Google Scholar] [CrossRef] [PubMed]
  165. Krishna, P.S.; Styring, S.; Mamedov, F. Photosystem ratio imbalance promotes direct sustainable H2 production in Chlamydomonas reinhardtii. Green Chem. 2019, 21, 4683–4690. [Google Scholar] [CrossRef]
  166. 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. DT. Int. J. Hydrog. Energy 2019, 44, 2533–2545. [Google Scholar] [CrossRef]
  167. 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. [Google Scholar] [CrossRef]
  168. Li, H.; Liu, Y.; Wang, Y.; Chen, M.; Zhuang, X.; Wang, C.; Wang, J.; Hu, Z. Improved photobio-H2 production regulated by artificial miRNA targeting psbA in green microalga Chlamydomonas reinhardtii. Biotechnol. Biofuels 2018, 11, 36. [Google Scholar] [CrossRef]
  169. Zhang, Y.; Cheng, J.; He, Y.; Yuan, J. Photo-fermentative hydrogen production performance of a newly isolated Rubrivivax gelatinosus YP03 strain with acid tolerance. Int. J. Hydrog. Energy 2022, 47, 20784–20792. [Google Scholar] [CrossRef]
  170. Arimbrathodi, S.P.; Javed, M.A.; Hamouda, M.A.; Aly Hassan, A.; Ahmed, M.E. BioH2 production using microalgae: Highlights on recent advancements from a bibliometric analysis. Water 2023, 15, 185. [Google Scholar] [CrossRef]
  171. Phanduang, O.; Lunprom, S.; Salakkam, A.; Liao, Q.; Reungsang, A. Improvement in energy recovery from Chlorella sp. biomass by integrated dark-photo biohydrogen production and dark fermentation-anaerobic digestion processes. Int. J. Hydrog. Energy 2019, 44, 23899–23911. [Google Scholar] [CrossRef]
  172. Ding, L.; Cheng, J.; Lu, H.; Yue, L.; Zhou, J.; Cen, K. Three-stage gaseous biofuel production combining dark hydrogen, photo hydrogen, and methane fermentation using wet Arthrospira platensis cultivated under high CO2 and sodium stress. Energy Convers. Manag. 2017, 148, 394–404. [Google Scholar] [CrossRef]
  173. Giang, T.T.; Lunprom, S.; Liao, Q.; Reungsang, A.; Salakkam, A. Improvement of hydrogen production from Chlorella sp. biomass by acid-thermal pretreatment. PeerJ 2019, 7, e6637. [Google Scholar] [CrossRef]
  174. Mishra, P.; Krishnan, S.; Rana, S.; Singh, L.; Sakinah, M.; Ab Wahid, Z. Outlook of fermentative hydrogen production techniques: An overview of dark, photo and integrated dark-photo fermentative approach to biomass. Energy Strategy Rev. 2019, 24, 27–37. [Google Scholar] [CrossRef]
  175. Wu, H.; Li, J.; Liao, Q.; Fu, Q.; Liu, Z. Enhanced biohydrogen and biomethane production from Chlorella sp. with hydrothermal treatment. Energy Convers. Manag. 2020, 205, 112373. [Google Scholar] [CrossRef]
  176. Usmanbaha, N.; Jariyaboon, R.; Reungsang, A.; Kongjan, P.; Chu, C.-Y. Optimization of batch dark fermentation of Chlorella sp. using mixed-cultures for simultaneous hydrogen and butyric acid production. Energies 2019, 12, 2529. [Google Scholar] [CrossRef]
  177. Singh, H.; Rout, S.; Das, D. Dark fermentative biohydrogen production using pretreated Scenedesmus obliquus biomass under an integrated paradigm of biorefinery. Int. J. Hydrog. Energy 2022, 47, 102–116. [Google Scholar] [CrossRef]
  178. Rincón-Pérez, J.; Razo-Flores, E.; Morales, M.; Alatriste-Mondragón, F.; Celis, L.B. Improving the biodegradability of Scenedesmus obtusiusculus by thermochemical pretreatment to produce hydrogen and methane. BioEnergy Res. 2020, 13, 477–486. [Google Scholar] [CrossRef]
  179. Cheng, J.; Yue, L.; Ding, L.; Li, Y.-Y.; Ye, Q.; Zhou, J.; Cen, K.; Lin, R. Improving fermentative hydrogen and methane production from an algal bloom through hydrothermal/steam acid pretreatment. Int. J. Hydrog. Energy 2019, 44, 5812–5820. [Google Scholar] [CrossRef]
  180. Wang, Q.; Gong, Y.; Liu, S.; Wang, D.; Liu, R.; Zhou, X.; Nghiem, L.D.; Zhao, Y. Free ammonia pretreatment to improve bio-hydrogen production from anaerobic dark fermentation of microalgae. ACS Sustain. Chem. Eng. 2019, 7, 1642–1647. [Google Scholar] [CrossRef]
  181. Lakatos, G.E.; Ranglová, K.; Manoel, J.C.; Grivalský, T.; Kopecký, J.; Masojídek, J. Bioethanol production from microalgae polysaccharides. Folia Microbiol. 2019, 64, 627–644. [Google Scholar] [CrossRef] [PubMed]
  182. Hirano, A.; Ueda, R.; Hirayama, S.; Ogushi, Y. CO2 fixation and ethanol production with microalgal photosynthesis and intracellular anaerobic fermentation. Energy 1997, 22, 137–142. [Google Scholar] [CrossRef]
  183. Hirayama, S.; Ueda, R.; Ogushi, Y.; Hirano, A.; Samejima, Y.; Hon-Nami, K.; Kunito, S. Ethanol production from carbon dioxide by fermentative microalgae. In Studies in Surface Science and Catalysis; Inui, T., Anpo, M., Izui, K., Yanagida, S., Yamaguchi, T., Eds.; Elsevier: Amsterdam, The Netherlands, 1998; Volume 114, pp. 657–660. [Google Scholar]
  184. Ueno, Y.; Kurano, N.; Miyachi, S. Ethanol production by dark fermentation in the marine green alga, Chlorococcum littorale. J. Ferment. Bioeng. 1998, 86, 38–43. [Google Scholar] [CrossRef]
  185. de Farias Silva, C.E.; Bertucco, A. Bioethanol from microalgae and cyanobacteria: A review and technological outlook. Process Biochem. 2016, 51, 1833–1842. [Google Scholar] [CrossRef]
  186. Gundolf, R.; Oberleitner, S.; Richter, J. Evaluation of new genetic toolkits and their role for ethanol production in cyanobacteria. Energies 2019, 12, 3515. [Google Scholar] [CrossRef]
  187. Mund, N.K.; Liu, Y.; Chen, S. Advances in metabolic engineering of cyanobacteria for production of biofuels. Fuel 2022, 322, 124117. [Google Scholar] [CrossRef]
  188. Kopka, J.; Schmidt, S.; Dethloff, F.; Pade, N.; Berendt, S.; Schottkowski, M.; Martin, N.; Dühring, U.; Kuchmina, E.; Enke, H.; et al. Systems analysis of ethanol production in the genetically engineered cyanobacterium Synechococcus sp. PCC 7002. Biotechnol. Biofuels 2017, 10, 56. [Google Scholar] [CrossRef]
  189. Chou, H.-H.; Su, H.-Y.; Chow, T.-J.; Lee, T.-M.; Cheng, W.-H.; Chang, J.-S.; Chen, H.-J. Engineering cyanobacteria with enhanced growth in simulated flue gases for high-yield bioethanol production. Biochem. Eng. J. 2021, 165, 107823. [Google Scholar] [CrossRef]
  190. Deng, M.-D.; Coleman John, R. Ethanol synthesis by genetic engineering in cyanobacteria. Appl. Environ. Microbiol. 1999, 65, 523–528. [Google Scholar] [CrossRef]
  191. Dexter, J.; Fu, P. Metabolic engineering of cyanobacteria for ethanol production. Energy Environ. Sci. 2009, 2, 857–864. [Google Scholar] [CrossRef]
  192. Choi, Y.-N.; Park, J.M. Enhancing biomass and ethanol production by increasing NADPH production in Synechocystis sp. PCC 6803. Bioresour. Technol. 2016, 213, 54–57. [Google Scholar] [CrossRef] [PubMed]
  193. Lopes da Silva, T.; Passarinho, P.C.; Galriça, R.; Zenóglio, A.; Armshaw, P.; Pembroke, J.T.; Sheahan, C.; Reis, A.; Gírio, F. Evaluation of the ethanol tolerance for wild and mutant Synechocystis strains by flow cytometry. Biotechnol. Rep. 2018, 17, 137–147. [Google Scholar] [CrossRef] [PubMed]
  194. Namakoshi, K.; Nakajima, T.; Yoshikawa, K.; Toya, Y.; Shimizu, H. Combinatorial deletions of glgC and phaCE enhance ethanol production in Synechocystis sp. PCC 6803. J. Biotechnol. 2016, 239, 13–19. [Google Scholar] [CrossRef] [PubMed]
  195. Roussou, S.; Albergati, A.; Liang, F.; Lindblad, P. Engineered cyanobacteria with additional overexpression of selected Calvin-Benson-Bassham enzymes show further increased ethanol production. Metab. Eng. Commun. 2021, 12, e00161. [Google Scholar] [CrossRef] [PubMed]
  196. Liang, F.; Englund, E.; Lindberg, P.; Lindblad, P. Engineered cyanobacteria with enhanced growth show increased ethanol production and higher biofuel to biomass ratio. Metab. Eng. 2018, 46, 51–59. [Google Scholar] [CrossRef] [PubMed]
  197. Gao, Z.; Zhao, H.; Li, Z.; Tan, X.; Lu, X. Photosynthetic production of ethanol from carbon dioxide in genetically engineered cyanobacteria. Energy Environ. Sci. 2012, 5, 9857–9865. [Google Scholar] [CrossRef]
  198. Velmurugan, R.; Incharoensakdi, A. Metal oxide mediated extracellular NADPH regeneration improves ethanol production by engineered Synechocystis sp. PCC 6803. Front. Bioeng. Biotechnol. 2019, 7, 148. [Google Scholar] [CrossRef]
  199. Mohd Azhar, S.H.; Abdulla, R.; Jambo, S.A.; Marbawi, H.; Gansau, J.A.; Mohd Faik, A.A.; Rodrigues, K.F. Yeasts in sustainable bioethanol production: A review. Biochem. Biophys. Rep. 2017, 10, 52–61. [Google Scholar] [CrossRef]
  200. Shokrkar, H.; Ebrahimi, S.; Zamani, M. Bioethanol production from acidic and enzymatic hydrolysates of mixed microalgae culture. Fuel 2017, 200, 380–386. [Google Scholar] [CrossRef]
  201. de Farias Silva, C.E.; Meneghello, D.; Bertucco, A. A systematic study regarding hydrolysis and ethanol fermentation from microalgal biomass. Biocatal. Agric. Biotechnol. 2018, 14, 172–182. [Google Scholar] [CrossRef]
  202. Aikawa, S.; Inokuma, K.; Wakai, S.; Sasaki, K.; Ogino, C.; Chang, J.S.; Hasunuma, T.; Kondo, A. Direct and highly productive conversion of cyanobacteria Arthrospira platensis to ethanol with CaCl2 addition. Biotechnol. Biofuels 2018, 11, 50. [Google Scholar] [CrossRef]
  203. Phwan, C.K.; Ong, H.C.; Chen, W.H.; Ling, T.C.; Ng, E.P.; Show, P.L. Overview: Comparison of pretreatment technologies and fermentation processes of bioethanol from microalgae. Energy Convers. Manag. 2018, 173, 81–94. [Google Scholar] [CrossRef]
  204. Kim, H.M.; Oh, C.H.; Bae, H.J. Comparison of red microalgae (Porphyridium cruentum) culture conditions for bioethanol production. Bioresour. Technol. 2017, 233, 44–50. [Google Scholar] [CrossRef] [PubMed]
  205. Megawati, M.; Bahlawan, Z.A.S.; Damayanti, A.; Putri, R.D.A.; Triwibowo, B.; Prasetiawan, H. Comparative study on the various hydrolysis and fermentation methods of Chlorella vulgaris biomass for the production of bioethanol. Int. J. Renew. Energy Dev. 2022, 11, 515–522. [Google Scholar] [CrossRef]
  206. Mirzaei, D.; Jazini, M.; Rahimi, M.; Mahdieh, M.; Karimi, K. Production of astaxanthin, ethanol and methane from Chromochloris zofingiensis microalga in an integrated biorefinery. Algal Res. 2022, 68, 102905. [Google Scholar] [CrossRef]
  207. Hosseini, A.; Jazini, M.; Mahdieh, M.; Karimi, K. Efficient superantioxidant and biofuel production from microalga Haematococcus pluvialis via a biorefinery approach. Bioresour. Technol. 2020, 306, 123100. [Google Scholar] [CrossRef] [PubMed]
  208. Condor, B.E.; de Luna, M.D.G.; Chang, Y.-H.; Chen, J.-H.; Leong, Y.K.; Chen, P.-T.; Chen, C.-Y.; Lee, D.-J.; Chang, J.-S. Bioethanol production from microalgae biomass at high-solids loadings. Bioresour. Technol. 2022, 363, 128002. [Google Scholar] [CrossRef]
  209. Rafa, N.; Ahmed, S.F.; Badruddin, I.A.; Mofijur, M.; Kamangar, S. Strategies to produce cost-effective third-generation biofuel from microalgae. Front. Energy Res. 2021, 9, 749968. [Google Scholar] [CrossRef]
  210. Sivaramakrishnan, R.; Suresh, S.; Kanwal, S.; Ramadoss, G.; Ramprakash, B.; Incharoensakdi, A. Microalgal biorefinery concepts’ developments for biofuel and bioproducts: Current perspective and bottlenecks. Int. J. Mol. Sci. 2022, 23, 2623. [Google Scholar] [CrossRef] [PubMed]
  211. Mahant, B.; Linga, P.; Kumar, R. Hydrogen economy and role of hythane as a bridging solution: A perspective review. Energy Fuels 2021, 35, 15424–15454. [Google Scholar] [CrossRef]
  212. Mehariya, S.; Goswami, R.K.; Karthikeysan, O.P.; Verma, P. Microalgae for high-value products: A way towards green nutraceutical and pharmaceutical compounds. Chemosphere 2021, 280, 130553. [Google Scholar] [CrossRef] [PubMed]
  213. 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, 124059. [Google Scholar] [CrossRef]
  214. Wang, Z.; Peng, X.; Xia, A.; Shah, A.A.; Huang, Y.; Zhu, X.; Zhu, X.; Liao, Q. The role of machine learning to boost the bioenergy and biofuels conversion. Bioresour. Technol. 2022, 343, 126099. [Google Scholar] [CrossRef]
  215. Ahmad Sobri, M.Z.; Redhwan, A.; Ameen, F.; Lim, J.W.; Liew, C.S.; Mong, G.R.; Daud, H.; Sokkalingam, R.; Ho, C.-D.; Usman, A.; et al. A review unveiling various machine learning algorithms adopted for biohydrogen productions from microalgae. Fermentation 2023, 9, 243. [Google Scholar] [CrossRef]
  216. Ullah, H.; Haq, Z.U.; Naqvi, S.R.; Khan, M.N.A.; Ahsan, M.; Wang, J. Optimization based comparative study of machine learning methods for the prediction of bio-oil produced from microalgae via pyrolysis. J. Anal. Appl. Pyrolysis 2023, 170, 105879. [Google Scholar] [CrossRef]
  217. Mutlu, A.Y.; Yucel, O. An artificial intelligence based approach to predicting syngas composition for downdraft biomass gasification. Energy 2018, 165, 895–901. [Google Scholar] [CrossRef]
  218. Shafizadeh, A.; Shahbeig, H.; Nadian, M.H.; Mobli, H.; Dowlati, M.; Gupta, V.K.; Peng, W.; Lam, S.S.; Tabatabaei, M.; Aghbashlo, M. Machine learning predicts and optimizes hydrothermal liquefaction of biomass. Chem. Eng. J. 2022, 445, 136579. [Google Scholar] [CrossRef]
  219. Salameh, T.; Sayed, E.T.; Olabi, A.G.; Hdaib, I.I.; Allan, Y.; Alkasrawi, M.; Abdelkareem, M.A. Adaptive network fuzzy inference system and particle swarm optimization of biohydrogen production process. Fermentation 2022, 8, 483. [Google Scholar] [CrossRef]
Figure 1. Different biomass conversion pathways to produce bioenergy from microalgae.
Figure 1. Different biomass conversion pathways to produce bioenergy from microalgae.
Fermentation 09 00529 g001
Table 1. Biochemical composition of some microalgae species (% w/w, on a dry mass basis).
Table 1. Biochemical composition of some microalgae species (% w/w, on a dry mass basis).
Microalgae SpeciesLipid ProteinCarbohydrateReferences
Botryococcus braunii25–751.54–55 [16]
Chlorella emersonii23–633641 [19]
Chlorella protothecoides40–6010–2811–15 [19]
Chlamydomonas reinhardtii15–189.259.7 [16]
Chlorella sorokiniana26.245.523.7 [34]
Chlorella vulgaris41–5851–5812–17 [19]
Dunaliella salina6–255732 [16]
Euglena gracilis4–2039–6114–18 [16]
Isochrysis galbana112734 [35]
Nannochloropsis gaditana23.348.39.3 [20]
Nannochloropsis granulata24–2827–3618–34 [35]
Neochloris oleoabundans35–6510–2717–27 [19]
Porphyridium cruentum9–1428–3940–57 [20]
Scenedesmus dimorphus16–408–1821–52 [19]
Scenedesmus obliquus30–5010–4520–40 [19]
Tetraselmis chuii1231–4612 [35]
Table 2. Different approaches to enhance lipid and carbohydrate content in the microalgae biomass as well as biomass yield and their outcomes (↑: increase, ↓: decrease, and -: not available).
Table 2. Different approaches to enhance lipid and carbohydrate content in the microalgae biomass as well as biomass yield and their outcomes (↑: increase, ↓: decrease, and -: not available).
Approach Outcomes References
Lipid CarbohydrateBiomass
Nutrient stress
Nitrogen deprivation/starvation [19,45]
Phosphorus deprivation/starvation↑/↓-/↓ [19,45]
Trace metal availability↑/↓↑/↓ [19,46]
Carbon availability- [19,47]
pH stress
Acidic↑/↓↑/↓↑/↓ [19,48]
Alkaline↑/↓↑/↓↑/↓ [19,48]
Temperature stress
High temperature↑/↓-↑/↓ [49]
Low temperature↑/↓-↑/↓ [49]
Light stress
Low Light↑/↓↑/↓ [50]
High Light↑/↓↑/↓↑/↓ [50]
Saline stress↑/↓ [19,51]
Table 4. Co-digestion of microalgae biomass with carbon-rich waste streams.
Table 4. Co-digestion of microalgae biomass with carbon-rich waste streams.
Microalgae
Species
Co-SubstrateMicroalgae/
Co-Substrate Ratio
Operating
Conditions
Biomethane Yield
(mL CH4/g VS)
Improvement in Biomethane Yield with Co-Digestion (%)References
Mixed microalgae
consortium
of
Chlorella
sorokiniana and Scenedesmus sp.
NoneNABatch;
Mesophilic
(37 °C)
331 ± 76NA [111]
Waste-activated sludge from the
aerobic phase of the wastewater
treatment plant
0.1 g of
algae/1 g of sludge (TS)
400 ± 2221
Waste-activated sludge from the
anaerobic phase of the wastewater
treatment plant
0.1 g of
algae/1 g of sludge (TS)
560 ± 2469
Scenedesmus quadricaudaNoneNABatch;
Mesophilic
(35 °C)
197 NA [112]
Thickened
waste-activated sludge from the wastewater
treatment plant
49 g of algae/51 g of sludge (VS)22212
Mixed microalgae–bacteria
consortium
NoneNASemi-continuous;
30 d HRT;
Mesophilic
(37 °C)
90 ± 2NA [113]
Thickened primary sludge from the wastewater
treatment plant
75 g of algae/25 g of sludge (VS) 133 ± 648
50 g of algae/50 g of sludge (VS)216 ± 1140
25 g of algae/75 g of sludge (VS)291 ± 9223
Mixed microalgae–bacteria
consortium,
dominated by
Dictyosphaerium sp.
NoneNABatch;
Mesophilic
(35 °C)
332NA [114]
Synthetic food waste (20% rice, 15% meat, 20% beans, 25% lettuce, 10% carrot, and 10% tomato) 50 g of algae/50 g of food waste (VS) 40923
25 g of algae/75 g of food waste (VS)51455
Mixed microalgae–bacteria
consortium
dominated by Chlorella sp.,
Scenedesmus sp., and pennate
diatoms
NoneNABatch;
Mesophilic
(25–32 °C)
274NA [115]
Swine wastewater50: 50 (v/v)32619
Mixed microalgae–bacteria
consortium
NoneNABatch;
Mesophilic
(34.5–35.5)
174 ± 2NA [117]
Deproteinated cheese whey17 g algae/83 g of whey (VS)302 ± 774
Cellulose16 g of algae/84 g of cellulose (VS)272 ± 1256
NoneNASemi-continuous;
30 d HRT;
Mesophilic
(33–35 °C)
36–81 mL CH4/g CODNA
Deproteinated cheese whey17 g algae/83 g of whey (VS)210–222 mL CH4/g COD167–500
Cellulose16 g of algae/84 g of cellulose (VS)97–122 mL CH4/g COD51–169
Chlorella vulgarisNoneNABatch;
Mesophilic
(35 °C)
167NA [116]
Potato processing waste (discarded parts and peels)25 g of algae/75 g of potato waste (VS)26659
Chlorella vulgarisPotato processing waste (discarded parts)25 g of algae/75 g of potato waste (VS)Semi-continuous;
20 d HRT;
Mesophilic
(37 °C)
300 mL CH4/g CODNA [110]
Potato processing waste (discarded parts) + glycerol25 g of algae/75 g of potato waste (VS) + 1% glycerol (v/v)730 mL CH4/g COD143
Potato processing waste (peels) 25 g of algae/75 g of potato waste (VS)330 mL CH4/g CODNA
Potato processing waste (peels) +
glycerol
25 g of algae/75 g of potato waste (VS) + 1% glycerol (v/v)550 mL CH4/g COD67
Mixed microalgae–bacteria
consortium
None NABatch;
Mesophilic
(35 °C)
140 ± 4NA [118]
Thickened primary sludge50 g of algae/50 g of sludge (VS)207 ± 548
Thickened primary sludge and fat, oil, and grease (FOG) 50 g of algae/50 g of sludge + 10% FOG (VS)259 ± 1385
50 g of algae/50 g of sludge + 20% FOG (VS)293 ± 8109
Note: NA—not applicable, TS—total solids, VS—volatile solids, HRT—hydraulic retention time, COD—chemical oxygen demand.
Table 6. Genetic engineering approaches to enhance biohydrogen production from microalgal biophotolysis.
Table 6. Genetic engineering approaches to enhance biohydrogen production from microalgal biophotolysis.
Microalgae
Species
Genetic Engineering ApproachOutcomeReferences
Chlamydomonas reinhardtii
hpm91 mutant
Mutant lacking PGR5 (Proton Gradient Regulation 5)Produced 7287 mL/10 Lreactor biohydrogen in 26 days under sulfur deprivation [164]
Chlamydomonas reinhardtii
C3 mutant
Altered the ratio between PSI and PSII from 0.85 to 0.33Produced biohydrogen with a rate of 3 mL/L.d for 42 days [165]
Chlorella sp.
DT mutant
Modified amino acid residues A105I, V265W, G113I, or V273I around the hydrogenase gas tunnel to prevent oxygen from accessing the enzyme active site via site-directed mutagenesisProduced 7 times more biohydrogen than the wild type in the presence of 5% oxygen [166]
Chlamydomonas reinhardtii
FACHB-265
Randomly mutated by atmospheric and room temperature plasma (ARTP)Produced 2.7–3.1 times higher biohydrogen than the wild type [167]
Chlamydomonas
reinhardtii
Used artificial miRNA (amiRNA) to regulate the function of D1-encoded gene, psbA Produced 60% more biohydrogen content than the wild type [168]
Table 10. Integrated biorefinery approach for combined production of bioenergy and bioproducts from microalgae biomass.
Table 10. Integrated biorefinery approach for combined production of bioenergy and bioproducts from microalgae biomass.
Microalgae
Species
ProcessProductEnergy/Product Yield or
Recovery
References
Chlorella sp.Dark fermentationBiohydrogen0.5 kJ/g VS [171]
Dark fermentation and photo fermentationBiohydrogen1.9 kJ/g VS
Dark fermentation and anaerobic digestionBiohydrogen and Biomethane6 kJ/g VS
Chlorella sp.Dark fermentationBiohydrogen0.4%/g VS [175]
Dark fermentation and anaerobic digestionBiohydrogen and Biomethane57%/g VS
Algal bloom dominated by Microcystis sp.Dark fermentationBiohydrogen0.4%/g VS [179]
Dark fermentation and anaerobic digestionBiohydrogen and Biomethane39–44%/g VS
Scenedesmus obliquus UTEX 393Lipid extraction/transesterificationBiodiesel13%/g VS [177]
Lipid extraction/transesterification and dark fermentationBiodiesel and Biohydrogen19%/g VS
Lipid extraction/transesterification, dark fermentation, and anaerobic digestionBiodiesel, Biohydrogen, and Biomethane30%/g VS
Arthrospira platensisDark fermentationBiohydrogen0.5–1 kJ/g VS [172]
Dark fermentation and photo fermentationBiohydrogen2.4–4.6 kJ/g VS
Dark fermentation, photo fermentation, and anaerobic digestionBiohydrogen and biomethane9.9–10.5 kJ/g VS
Chromochloris zofingiensis SAG 211-14Anaerobic digestionBiomethane287 L/kg TS
(10,343 kJ)
[206]
Carotenoid extraction and
anaerobic digestion
Carotenoids
(Mainly astaxanthin)
10 g/kg TS
Biomethane198 L/kg TS (7153 kJ)
Carotenoid extraction, yeast-based fermentation, and anaerobic digestionCarotenoids
(Mainly astaxanthin)
10 g/kg TS
Bioethanol143 g/kg TS (3832 kJ)
Biomethane123 L/kg TS (4428 kJ)
Haematococcus pluvialis
SAG 192.80
Carotenoid extraction and
anaerobic digestion
Astaxanthin39 g/kg TS [207]
Biomethane192 L/kg TS
(6939 kJ)
Carotenoid extraction, yeast-based fermentation, and anaerobic digestionAstaxanthin39 g/kg TS
Bioethanol170 g/kg TS
(4666 kJ)
Biomethane67 L/kg TS
(2430 kJ)
Note: VS—volatile solids, TS—total solids.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Parakh, S.K.; Tian, Z.; Wong, J.Z.E.; Tong, Y.W. From Microalgae to Bioenergy: Recent Advances in Biochemical Conversion Processes. Fermentation 2023, 9, 529. https://doi.org/10.3390/fermentation9060529

AMA Style

Parakh SK, Tian Z, Wong JZE, Tong YW. From Microalgae to Bioenergy: Recent Advances in Biochemical Conversion Processes. Fermentation. 2023; 9(6):529. https://doi.org/10.3390/fermentation9060529

Chicago/Turabian Style

Parakh, Sheetal Kishor, Zinong Tian, Jonathan Zhi En Wong, and Yen Wah Tong. 2023. "From Microalgae to Bioenergy: Recent Advances in Biochemical Conversion Processes" Fermentation 9, no. 6: 529. https://doi.org/10.3390/fermentation9060529

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

Article Metrics

Back to TopTop