Next Article in Journal
Numerical Investigations of the Cooling Performance of an R410A Closed-Loop Spray Cooling System
Previous Article in Journal
Modeling and Simulation of Distribution Networks under Lightning Transients: A Comparative Study of Accuracy and Complexity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biotechnological Valorization of Waste Glycerol into Gaseous Biofuels—A Review

by
Joanna Kazimierowicz
1,
Marcin Dębowski
2,*,
Marcin Zieliński
2,
Sławomir Kasiński
3 and
Jordi Cruz Sanchez
4
1
Department of Water Supply and Sewage Systems, Faculty of Civil Engineering and Environmental Sciences, Bialystok University of Technology, 15-351 Bialystok, Poland
2
Department of Environment Engineering, Faculty of Geoengineering, University of Warmia and Mazury in Olsztyn, Str. Oczapowskiego 5, 10-719 Olsztyn, Poland
3
Department of Environmental Biotechnology, Faculty of Geoengineering, University of Warmia and Mazury in Olsztyn, Str. Oczapowskiego 5, 10-719 Olsztyn, Poland
4
Department of Basic Formation, Escola Universitària Salesiana de Sarrià, Passeig Sant Joan Bosco, 74, 08017 Barcelona, Spain
*
Author to whom correspondence should be addressed.
Energies 2024, 17(2), 338; https://doi.org/10.3390/en17020338
Submission received: 13 December 2023 / Revised: 5 January 2024 / Accepted: 8 January 2024 / Published: 9 January 2024
(This article belongs to the Section A4: Bio-Energy)

Abstract

:
The supply of waste glycerol is rising steadily, partially due to the increased global production of biodiesel. Global biodiesel production totals about 47.1 billion liters and is a process that involves the co-production of waste glycerol, which accounts for over 12% of total esters produced. Waste glycerol is also generated during bioethanol production and is estimated to account for 10% of the total sugar consumed on average. Therefore, there is a real need to seek new technologies for reusing and neutralizing glycerol waste, as well as refining the existing ones. Biotechnological means of valorizing waste glycerol include converting it into gas biofuels via anaerobic fermentation processes. Glycerol-to-bioenergy conversion can be improved through the implementation of new technologies, the use of carefully selected or genetically modified microbial strains, the improvement of their metabolic efficiency, and the synthesis of new enzymes. The present study aimed to describe the mechanisms of microbial and anaerobic glycerol-to-biogas valorization processes (including methane, hydrogen, and biohythane) and assess their efficiency, as well as examine the progress of research and implementation work on the subject and present future avenues of research.

1. Introduction

The supply of waste glycerol is rising steadily—spurred, in part, by the increased global production of biodiesel, which in turn is driven by biogenic fraction minimums required in conventional fuel [1]. Global biodiesel production totals about 47.1 billion liters and is a process that involves the co-production of waste glycerol, accounting for over 12% of total esters produced, regardless of the catalyst or technological process used [2]. Waste glycerol is also generated during bioethanol production and is estimated to account for 10% of the total sugar consumed on average [3]. The influx of surplus crude glycerol from the biofuels industry has led to a drastic drop in glycerol prices from 480 USD/ton in 2002 to 110 USD/ton in 2021, making it a waste product that requires significant expenditure to dispose of or reuse. This directly impacts the cost-effectiveness of biodiesel and bioethanol production [4]. Figure 1 shows the global growth of biodiesel and bioethanol production [5].
Valorization and processing of raw glycerol into value-added products is an urgent and growing priority, as well as a contributor to the viability of a biofuel-based economy [6,7]. Research on technologies for converting glycerol into value-added products has included both chemical and biological processes [8,9]. Biological pathways have several advantages over chemical conversion processes, including higher tolerance to contamination, greater environmental friendliness thanks to the lower temperatures and pressures involved, no consumption of chemicals, reduced secondary contamination, and, in many cases, a simple process design that provides versatility and ease of operation. Another added benefit is the large variety of glycerol-metabolizing microorganisms that have been identified [10].
Glycerol generates more reducing equivalents than other carbon sources (e.g., glucose or xylose), thus making it possible to produce biofuels and reduced chemicals at higher efficiency [11]. The abundance, relatively low price, and high reduction capacity of glycerol have made it an attractive feedstock to exploit for the production of various biofuels [12]. These features have significantly increased interest in developing microbial methods for the valorization of waste glycerol into gaseous biofuels, resulting in an abundance of available technologies. As shown in Figure 2, there has been a marked and fast-growing interest in research on the bioconversion of fuel to hydrogen (H2), methane (CH4), and, to a lesser extent, biohythane.
Although certain microorganisms that can process and metabolize glycerol into value-added products or biofuels have been selected and tested under natural conditions, their industrial applications have often been limited by relatively low yields and production rates [13]. Thus, interventions are needed to modify and increase the efficiency of natural enzymes and accelerate metabolic pathways to maximize industrial benefits [14]. Process improvement draws on research regarding metabolic engineering, genetic modification, and upgrading/optimization current processes [15,16]. The most common refinements include: eliminating transcriptional repression and inhibiting enzyme feedback; reducing the generation and accumulation of undesired by-products in the medium; enhancing and improving the medium via micro-nutrient amendment; reducing energy consumption and cost of the system; expanding the substrate and end product portfolio; and enhancing strain tolerance to inhibitors and environmental stress [17,18]. The integration of protein engineering, systems biology, and synthetic biology into metabolic engineering seems to have extended strain engineering from local modification to total system optimization. Combining insights from various omics technologies, such as genomics, transcriptomics, proteomics, and fluxomics, has provided an in-depth understanding of how glycerol metabolism proceeds in bioenergy systems and how whole microbial communities are regulated at the system level [19].
Despite the highly advanced research on waste glycerol bioconversion into gaseous biofuels and the fast proliferation of such processes, there is still a real need to seek new technologies as well as refine the existing ones. Earlier reports on the use of waste glycerol as a substrate for the production of gaseous biofuels focused mainly on methane fermentation. There are few papers analyzing the possibility of using this organic substrate in hydrogen production processes, while the biological conversion of glycerol into biohythane has not yet been presented. This study aimed to describe the mechanisms of microbial and anaerobic glycerol-to-biogas valorization processes (including CH4, H2, and biohythane), assess their efficiency, examine the progress of research and implementation work on the subject, and present future avenues of research. Glycerol, or propane-1,2,3-triol, also commonly known as glycerin, can be obtained through organic synthesis and biotechnological processing. Its considerable supply on the market comes from industrial biodiesel production [20].

2. Glycerol Production and Applications

Biodiesel—fatty acid methyl ester (FAME)—is an alternative fuel produced from vegetable oils, animal fats, waste fats, and other fatty feedstocks with suitable characteristics and properties. Typical feedstocks for biodiesel production are rapeseed oil, soya oil, sunflower oil, palm oil, and oil obtained from microalgae [21]. Cattle/sheep/poultry tallow, animal oil, used cooking oil, fish oil, jatropha oil, and coke have also been used [22,23].
Biodiesel is synthesized by transesterifying lipids (triacylglycerols) with a simple alcohol, such as methanol or ethanol (alcoholysis) (Figure 3). This reaction is usually catalyzed by NaOH, KOH, or sulfuric acid (VI) [24,25]. Chemical transesterification by alkaline catalysts is a fast-progressing reaction, which means that high volumes of triglycerides can be converted into their respective methyl or ethyl esters. The reaction will convert approx. 10% of the feedstock into glycerol [26,27]. However, it is difficult to separate the pure glycerol from the crude fraction due to the type and amount of impurities, such as methanol, fatty acid mono- and diacylglycerols, free fatty acids, and soaps produced during the biodiesel production process. Purified glycerol is a value-added product of great commercial and economic importance.
Glycerol is also formed during bioethanol production via alcoholic fermentation [28]. These biochemical pathways allow anaerobic microorganisms to regenerate the NAD (nicotinamide adenine dinucleotide) consumed during glycolysis. The pyruvate formed during the last step of glycolysis is then reduced to ethanol (ethyl alcohol) via a two-step process that also oxidizes the NADH (reduced form of NAD) formed during glycolysis to NAD while releasing carbon dioxide [29]. In the first step, pyruvate is converted to ethanal (acetaldehyde) and carbon dioxide by pyruvate decarboxylase. The second step reduces ethanal to ethanol by alcohol dehydrogenase (ADH), with NADH oxidized to NAD [30]. Alcoholic fermentation does not convert all of the glucose to bioethanol [31]. A portion remains unconverted to bioethanol, while at the same time some of the bioethanol is subject to further reactions, which generate glycerol as a by-product [32]. Yeast can also produce glycerol via osmotic regulation under stress [33]. It is important to note that the glycerol output of bioethanol production can vary depending on a multitude of factors, including the type of feedstock, fermentation parameters, type of yeast, etc. [34]. Bioethanol producers often strive to optimize their processes in order to minimize by-products, including glycerol [35].
The efficient use of crude glycerol is crucial for the further commercialization of biodiesel production and can significantly reduce the price of biodiesel [36]. One of the ways of neutralizing the waste glycerol fraction is to harness it for energy production, primarily by means of thermal conversion or bioconversion [37,38]. Major biofuel producers purify glycerin waste through filtration or fractional vacuum distillation [39,40], as well as by using chemical conversion or selective catalytic oxidation [41,42]. The literature also includes numerous reports on reusing the glycerin fraction in biological methods as a carbon source to fuel microbial growth. From a commercial standpoint, converting waste glycerol into gaseous biofuel via microbial bioconversion represents a major opportunity for the fuel industry. Well-selected microbial groups are capable of converting glycerol into value-added gas metabolites, and the process itself does not require that the glycerol fraction be purified [43]. The cost-effectiveness of the process can be further boosted by replacing other carbon sources with cheaper waste feedstock. In this light, crude glycerol—seen as industrial waste and available in large quantities—would seem to be a good candidate for the production of biogases with various compositions and properties [44,45]. A basic division of the methods used to valorize and reuse waste glycerol from biodiesel production is presented in Figure 4 [37,38,43].

3. Biomethane Production

A common method for neutralizing waste glycerol is by processing it via an anaerobic digestion (AD) process. AD utilizes biological conversion processes occurring under aerobic conditions, producing CH4-rich biogas and digestate (usually utilized as organic fertilizer) [46,47,48]. Each step of the digestion is mediated by specialized microorganisms that hydrolyze polymeric substances through enzyme action [49,50,51]. They are further broken down into fatty acids, alcohols, H2, and CO2 [52]. AD is a process that consists of the following steps: hydrolysis, acidogenesis, acetogenesis, and methanogenesis. Hydrolysis involves the enzymatic biological degradation of complex organic compounds. Enzymes produced by anaerobic, hydrolytic microorganisms are used in this process [53]. During acidogenesis, the hydrolysates are further degraded into simpler compounds, mainly organic acids, alcohols and aldehydes, CO2, and H2 [54]. Acetogenesis is the process by which organic acids are converted to acetic acid, CO2, and H2. Acetate is synthesized via one of two pathways. The first oxidizes the fatty acids produced in the acidogenesis phase, releasing H2; in the second phase, it uses H2 to reduce CO2. Acetogenesis is the most important phase of the AD process [55]. Methanogenesis is the last phase of AD. CH4 is produced via two pathways. In the first pathway, it is produced from acetic acid, and in the second from H2. In addition to CH4, methanogenesis also produces CO2, H2S, ammonia, and water [56]. The CH4-producing organisms involved in this process are archaea [57]. A step-by-step conversion diagram for the anaerobic digestion (AD) of glycerol (including the end products) is shown in Figure 5.
Waste glycerol is usually included as a part of a wider mix of organic feedstocks that can be readily biodegraded in anaerobic digesters. The process by which bioreactors are fed with a mixture of different feedstocks is called co-digestion [60]. Many previous studies and data from full-scale plants have shown amendment with waste glycerol to have a very positive effect on the AD process and its performance [61]. This is owing to the high energy value of the substance, which increases the organic load rate (OLR) in the digesters while also directly and significantly boosting biogas yield and quality profile [62]. Glycerol is readily biodegradable under anaerobic conditions, though it does typically require higher hydraulic retention times (HRT) in digesters [63]. In many cases, its incorporation into the organic feedstock mix results in optimal OLR values and improves the carbon-to-nitrogen (C/N) ratio, which in turn reduces the levels of ammonia—a major inhibitor of AD in bioreactors [64].
It has also been shown that waste glycerol-based processes produce digestate of better fertilizing quality [65]. If the digestate from glycerol (co-)digestion is to be reused as an organic fertilizer, humus balancing in the soil becomes an important consideration [66]. This is an important factor when growing cereals, rapeseed, root crops, and tuber crops—plantations that cause a negative organic matter balance in the soil. Such organic fertilizer can be successfully used in humus-poor, low-quality soils, including light, degraded, or marginal soils. The organic-rich digestate from glycerol digestion can be used as a first-line treatment for restoring the humus index, which has an enormous impact on soil productivity, improving the water and sorption (ion binding) capacity, structure, and biological activity [67]. The literature often references the concept of “residual effect”, which refers to the crop yield increase upon the use of organic fertilizers that cannot be achieved with mineral fertilizers. The residual effect—such as the one from digestate derived from AD of glycerol—improves the physical, chemical, and biological properties of the soil [67].
Because of the growing supply of waste glycerol and its promising properties as an AD feedstock, an increasing number of researchers and industrial-scale plant operators have been striving to optimize its use [68]. Mesophilic digestion is the most popular AD process for this purpose, as evidenced, for example, by the studies on mesophilic co-digestion of sewage sludge and glycerol [69,70]. Research has shown that the best results are achieved when glycerol accounts for 25 to 60% of the total OLR fed into the digestate. Keeping the proportions within this range can boost biogas production by 82–280%. Increasing the share of glycerol in total OLR beyond 70% inhibited the process and hamstrung the result, i.e., biogas yields, CH4 fractions, and digestate profile [71,72]. Baba et al. [73] compared the energy balance of anaerobic digestion in a digester with an active volume of 30 m3, fed with a mix of crude glycerol and excess sludge. Once a day, 3.5 m3 of sludge and 5 to 75 L of crude glycerol were fed into the reactor. The best-performing variant provided a net energy output 106% higher than when digesting sewage sludge alone. The added benefit of the process was that the soil fertilized with the digested sludge showed improved structure and yielded 20% more grass for haylage.
Athanasoulia et al. [74] used a continuous stirred-tank reactor (CSTR) to co-digest waste glycerol and sewage sludge under mesophilic conditions. The addition of up to 4% glycerol (v/v) improved biogas production by 3.8–4.7 times. Another study investigated the impact of glycerol addition on biogas yields at a pilot scale [75] and found a low dose of glycerol (0.63% v/v) to be the optimal one. This variant had a CH4 yield of 1.3 m3/L crude glycerol [75]. Silvestre et al. [76] studied the effect of adding crude glycerol to AD-treated sewage sludge under thermophilic and mesophilic conditions. The addition of glycerol at the thermophilic temperature range had a negative impact on the stability and performance of the process, even at low doses. In contrast, the process performed steadily within the mesophilic temperature range, with a 148% increase in CH4 output with 1% glycerol in the influent v/v (27% of influent COD). The addition of glycerol to sewage sludge can be used to effectively balance the C/N ratio. However, no improvements in biogas yield were observed when the glycerol content exceeded 1% v/v in the influent, likely due to the high C/N [76].
These results suggest that crude glycerol can be used as a co-feedstock for the AD of sewage sludge, although various parameters should be taken into account—depending on the glycerol characteristics and operating conditions—to ensure digester stability and optimal performance [76]. Another study [77] concluded that feeding glycerol at up to 2% v/v did not enhance methane production. A detailed analysis of the process kinetics and biochemical dynamics of the microbial community found synergisms between glycerol and sewage sludge, with the result being improved biogas and CH4 yields. Doses of up to 2% v/v glycerol provided a good balance between increasing the organic loading rate and minimizing the impact on hydraulic retention time. During continuous operation over 200 days, feeding glycerol at up to 2% v/v sharply increased the organic load by 70% and resulted in a 50% increase in CH4 production, with nominal yields of 1248 ± 58 mL CH4/L/d [77].
Alves et al. [78] investigated how the addition of 1 and 3% crude glycerol v/v from biodiesel production affected the AD primary sewage sludge. Process stability and performance were monitored using such parameters as pH, COD, VS, C/N, and volatile acid/alkalinity ratio. At 1% glycerol, the CH4 yield was 223.8 mL CH4/g VS, whereas at 3% glycerol, the CH4 output was 368.8 mL CH4/g VS. This translates to an increase in CH4 yield of 61% and 167%, respectively, against the sludge-only control (138.2 mL CH4/g VS). Peak daily biogas production (56.8 mL CH4/g VS) was achieved with the highest glycerol fraction. However, an initial instability period (with methanogenesis inhibition) of 5.8 days was noted, likely attributable to the accumulation of intermediate volatile acids [78]. Another study by Alves et al. [79] aimed to evaluate the production of biogas (including CH4) from the anaerobic co-digestion of primary sludge (PS) from sewage treatment, food waste (FW), and crude glycerol. To study the effect of glycerol on AD stability and performance, biochemical methane potential (BMP) tests were performed at different glycerol concentrations—1 and 3% v/v. A modified Gompertz model was used to describe biogas and CH4 production. The study demonstrated that the small increases in the organic load from glycerol addition led to significant improvements in biogas and CH4 production. Although methanogenesis was temporarily inhibited in the BMP tests with the highest glycerol concentration—suggesting a longer adaptation period for methanogenic Archaea—CH4 production was later restored and even maximized under these conditions. The biogas and CH4 production reached 432.4 mL/g VS removed and 343.3 mL/g VS at 1% glycerol, respectively, and 692.6 mL/g VS removed and 525.7 mL CH4/g VS at 3% glycerol, respectively. This translates to an increase in CH4 production of 45.4% and 122.7% (for 1 and 3% glycerol, respectively) compared to the results for PS + FW [79].
Waste glycerol has also been co-digested with wastewater. Siles et al. [80] co-digested wastewater from biodiesel production with 1% v/v glycerol, obtaining 310 mL of CH4/g COD removed. Anaerobic co-digestion was shown to require less clean water and nutrients, resulting in commercial and environmental benefits [80]. Fountoulakis and Manios [81] tested the co-digestion of olive mill wastewater and slaughterhouse wastewater with 1% v/v glycerol. CH4 production was 1210 ± 205 mL CH4/d, which represents a 731 mL CH4/d increase compared to the no-glycerol variant (479 ± 38 mL CH4/d) [81]. The authors of [81] also co-digested the wastewater organic fraction of municipal solid waste (40% fruit, 25% potatoes, 25% vegetables, 8% bread, and 2% paper) with 1% v/v glycerol. CH4 yield amounted to 2094 mL CH4/d—meaning an increase of 694 mL CH4/d compared to the no-glycerol variant (1400 mL CH4/d).
Another common feedstock mix with waste glycerol provides for co-digestion of the glycerol with manure/slurry of various origins and properties. Astals et al. [82] demonstrated that thermophilic digestion of a feedstock mix containing pig manure and crude glycerol (3% v/v) yielded 180% more biogas (0.47 L biogas/g VS) compared with a single-feedstock process (pig manure only). The increase was mainly attributed to a balanced C/N ratio, the energy value of the organic matter fed in, and the optimal and stable OLR [82].
However, the authors emphasize that the high levels of organic biodegradable matter in the digestate preclude it from being a universal fertilizer—rather, its use is limited to light soils, mineral soils, and other soil types with a low humus content [82,83]. A high glycerol content in the organic feedstock can directly lead to high levels of organics in the digestate, thus making it difficult to stabilize [84]. The presence of biodegradable organic compounds in natural fertilizers used on agricultural soils or on degraded land under remediation can lead to sanitary difficulties, the spread of nuisance odors, or negative environmental impacts (surface runoff, water erosion, wind erosion) [85]. The quality and profile of the digestate, the options for its use as a fertilizer, and the conditions of such fertilization should always be taken into account when designing systems for the AD of glycerol [86].
A study by Usack and Angenent [87] used glycerol as a co-digestion feedstock with dairy cattle manure in a CSTR for 900 days under mesophilic conditions. The highest biogas production was noted when the ratio of dairy manure to crude glycerol (VS basis) was 62:38. The CH4 production reached 549 ± 25 mL CH4/g VS at 3.2 g VS/L/d OLR. The authors noted encountering some processing hurdles, primarily foaming in the anaerobic reactor, which caused limited diffusion of gas metabolites produced by anaerobic bacteria, flotation of the reactor medium, and problems with removing the digested mass from the reactor [87]. Foaming may be triggered by the presence of proteins or surfactants, which can occur in waste from dairy cow farming. These substances can originate from the milking station wash water and enter digesters with the input manure [88]. Glycerol input also significantly increases the fraction of biodegradable organic matter in the feedstock [89]. This leads to faster production of gas but also faster formation of foam, which, in the long term, can coalesce into a thick layer of scum on the surface of digesters. The scum acts as a barrier that prevents gases from diffusing at the liquid–gas interface and can lead to flow restriction or, in extreme cases, completely inhibit the flow of gaseous metabolites produced by anaerobic bacteria [90]. Ultimately, this leads to the accumulation of digestion metabolites in the medium and the inhibition of organic compound breakdown under anaerobic conditions [91].
Inhibition of anaerobic digestion of swine manure with added crude glycerol has been the subject of a study by Fierro et al. [92]. Added glycerol at over 8% v/v caused medium acidification, methanogenesis failure, and a rise in the CO2 content in the gas metabolites. Significant increases in H2S levels were also observed. As a result, the system had to be modified and expanded. A post-stabilization stage had to be introduced to complete the biodegradation of the feedstock organics, especially proteins and lipids [92]. The observations of Fierro et al. [92] have been reproduced by other researchers [93,94]. Glycerol is an organic substrate that, despite its positive effect on AD and its performance, should be used with caution in feedstock mixes used in bioreactors [95]. Exceeding the threshold of organic levels, and thus the OLR value, overloads the system [96,97]. In the long term, this causes volatile fatty acids (VFAs)—produced during the acidogenesis step of AD—to accumulate [98]. This, in turn, decreases pH, inhibits the metabolism of methanogenic Archaea, kills anaerobic microorganisms in the microbial community, and halts methanogenesis altogether [99].
Nevertheless, crude glycerol can be an extremely attractive candidate for an AD co-substrate due to its widely reported boosting effect on biogas production and composition [78]. It is recommended that optimization procedures and multivariate biogas tests be conducted to select the best-performing variants and identify potential problems early [86]. This is important both in terms of the process itself and the later utilization of its products. Oliveira et al. [100] co-digested Sargassum sp. macroalgae with glycerol and waste frying oil (WFO), looking for optimal CH4 production parameters. Three variables were investigated: % of the total algal mass (%TS Sargassum sp.), type of co-feedstock (glycerol or WFO), and co-feedstock concentration (g/L). The biochemical methane potential (BMP) of Sargassum sp. was 181 ± 1 L CH4/kg COD. Co-digestion with glycerol or WFO increased this parameter by 56% and 46%, respectively, whereas the methane production rate (k) was 38% and 19% higher for glycerol and WFO, respectively. Peak BMP (283 ± 18 L CH4/kg COD) was obtained in the 0.5% TS, 3.0 g/L glycerol trial [100]. Optimization of anaerobic digestion has also been pursued by Takeda et al. [101]. A Central Composite Rotational Design (CCRD) was built to experimental variables such as time (16.6, 20, 25, 30, and 33.4 days), glycerol content (0.43, 0.70, 1.10, 1.50, and 1.77%), and substrate/inoculum ratio (0.23, 0.30, 0.40, 0.50, and 0.57 g COD/g VSS). The optimization was successful in maximizing the efficiency of organic matter removal (90.15%) and specific biogas production (403.15 mL/g VSS) for the parameters: 33.2 days, glycerol content of 1.71%, and substrate/inoculum ratio of 0.37 g COD/g VSS. Using a modified Gompertz model of the optimal condition performed, the specific biogas production obtained was, on average, 20.3 times higher compared to the mono-digestion of leachate [101].
All studies to date have shown that overloading a system with glycerol-derived organic compounds leads to the inhibition of their breakdown under anaerobic conditions [102], decreases CH4 in the biogas [93], increases hydrogen sulfide levels [103], produces unstable digestate [104], and, in extreme cases, causes failures of reactors, stirrers, pumping systems, or cogeneration units [105]. Each introduction of glycerol into a tried and tested substrate mix requires systematic and skillful adaptation of the system and acclimation of the bacterial community to be used for a multi-step AD process [106].
A well-adapted bacterial community allows operators to use higher glycerol contents in the feedstock for efficient CH4 production and value-added organic fertilizer [107]. Many have reported good results with two-step digestion processes, using physically separated sequences of acidogenic and methanogenic digestion [94]. Such separation provides better process control and reduces the impact of the intensive VFA production on the methanogenic step [108]. Researchers and operators should focus their efforts on optimizing digestion parameters and individual process steps, including C/N ratio, OLR, HRT control during hydrolysis (acidogenic and methanogenic steps), and monitoring of the bacterial community composition [109]. Examples of waste glycerol use in AD, key process parameters, and performance metrics are given in Table 1.

4. Biohydrogen Production

4.1. Photofermentation

H2 meets the criteria of a low-impact energy carrier [110,111]. Its calorific value ranges from 10.8 to 12.75 MJ/Nm3, making it suitable for use in heating, power generation, and air/car transportation [112,113]. Currently, H2 is used on a limited basis, mostly in the refining industry, fuel cells, and space technology [114]. The main barrier to use is the lack of viable methods of production and storage—ones that would be technologically feasible and cost-effective [115]. Conventional H2 production mainly consists of thermochemical methods, including combustion, gasification, pyrolysis, thermochemical liquefaction, or water pyrolysis [116]. However, such solutions are energy-intensive, pollution-generating, and mired in high investment costs [117]. It is estimated that almost 95% of H2 is produced by converting fossil fuels [118]. Biomass-based technologies, as well as methods that harness the biological processes of microorganisms, are becoming more and more viable as a means of H2 production [119]. They consist mostly of organic-feedstock fermentation, carried out by specialized groups of bacteria, or biochemical processes in the cells of selected microalgae species [120]. Crude glycerol can be converted to H2 via two different fermentation processes: photofermentation and dark fermentation. The former requires a light source to proceed.
Photofermentation is performed by anaerobic bacteria, mainly green/purple sulfur and non-sulfur bacteria such as Rhodobacter sphaeroides, Rhodobacter capsulatus, Rhodobacter sulidophilu, Rhodopseudomonas palustris, Rhodopseudomonas sphaeroides, and Halobacterium halobium [121]. These species are capable of breaking down organic acids into H2 and CO2 [122]. Nitrogenase—the primary photofermentative enzyme—is a metalloprotein complex responsible for the biological fixation of molecular nitrogen and commonly found in both archaea and bacteria. The nitrogen fixation process can produce H2, thus maintaining redox homeostasis [123,124]. Nitrogenase activity can be inhibited by excessive C/N, oxygen, or ammonia, among other factors [125]. Chemical energy in the form of a proton gradient is produced from sunlight. This energy can be used to power various cellular processes, including a reduction in nitrogenase via reverse electron flow. On the other hand, reduced nitrogenase uses the hydrolysis of ATP, produced by photophosphorylation, to reduce metabolically produced protons to H2. The normal physiological function of nitrogenase is to reduce dinitrogen to ammonia, with the reduced release of one H2 per N2. However, nitrogenase turnover can still occur even in the absence of other reducible substrates, reducing protons to H2. The process follows the flowchart presented in Figure 6 [122].
The literature data indicates that Rhodopseudomonas palustris is the most popular microorganism for the photofermentation of waste glycerol, owing to its particular enzymatic activity [126]. Pott et al. [127] found that an R. palustris community fed with waste glycerol grew at a rate of 0.074 h−1. In this particular process, glycerol was converted at a conversion efficiency of approximately 90% at a reaction rate of 34 mL H2/gdw/h. Inhibition of microbial growth and H2 production was noted during the experiment, which was attributed to the contaminants in the crude glycerol. Saponified fatty acids were determined to be the main inhibitors of the process [127].
Sabourin-Provost and Hallenbeck [128] conducted a comparative study with both pure and crude glycerol as feedstock for photofermentative H2 production by Rhodopseudomonas palustris. The yields were approx. 6 moles of H2/mole glycerol across all variants. The impurities present in crude glycerol were not found to negatively affect the process of microbial growth in this case [128]. In another study, Ghosh et al. [129] endeavored to optimize a photofermentative H2 production process by Rhodopseudomonas palustris. Response Surface Methodology (RSM) was applied to select the process conditions and parameters important for glycerol-to-H2 conversion. This optimization protocol was used to determine the interactive effects among the process parameters, such as the concentration of crude glycerol and glutamate, light intensity, nitrogen source, and N levels, on H2 production efficiency.
Under optimal conditions (light intensity of 175 W/m2, 30 mM glycerol, and 4.5 mM glutamate), 6.69 moles H2/mole of crude glycerol were obtained, which translates to 96% of the theoretical yield. Determination of nitrogenase activity and expression levels showed that process variables produced relatively little variation in nitrogenase protein levels, whereas nitrogenase activity varied considerably, with peak nitrogenase activity (228 nmol C2H4/mL/min) achieved at the optimal central point [129]. Zhang et al. [130] used modeling to simulate the entire growth phase of R. palustris in order to maximize H2 production. Two piecewise models were designed, both fitted via batch experiments, by solving the underlying optimal control problems through stable and accurate discretization techniques. The optimal initial glutamate-to-glycerol ratio was 0.25, regardless of the initial biomass concentration. The glycerol conversion efficiency was found to correlate with the initial biomass concentration, with a computational peak of 64.4%. By optimizing a 30 day industrial batch process, an H2 production rate of 37.7 mL g biomass/h was achieved, while the glycerol conversion efficiency was maintained at 58% [130].
Jiang et al. [131] investigated the correlation between H2 yield improvement during photofermentation by Arundo donax L. and glycerol amendment. Glycerol concentrations of 0, 10, 15, 20, and 30 g/L were tested. The highest H2 yield improvement (79.2 mL/g substrate) was obtained at 15 g/L added glycerol, translating to a 294% increase in production compared with the Arundo donax L. mono-substrate system. Under the optimal glycerol addition (15 g/L), process factors were as follows: glycerol/Arundo donax L. ratio: 1/1, C/N ratio: 25.1, initial medium redox potential (Eh): 57 mV, solid/liquid ratio: 1/68. A canonical correlation analysis (CCA) found that initial and final Eh had the strongest relationship with photofermentation yield improvement. Furthermore, Pearson’s correlation analysis showed that the Arundo donax L./glycerol ratio played a key role in the photofermentative H2 production process [131].
Process parameters play a key role in photofermentation. The process requires a temperature between 30 °C and 36 °C, a near-neutral pH (6.8–7.5), and a source of bright 400–1000 nm light (6–10 klux) [112,132]. Using the right illumination intensity boosts H2 production rates and yields. However, due to high running costs, alternating light–dark cycles are usually used, with a 12/12 h interval being the most common choice [133]. H2 production efficiency is also directly related to the type and design of the photobioreactor (PBR) [134]. The most common are tubular, column, and flat-plate PBRs [135]. PBRs are typically closed and hermetically sealed, preventing oxygen penetration, contamination, and the growth of competing microbial species. PBRs are similar in design to reactors used for cultivating and growing microalgae [136,137]. Ross and Pott [138] have demonstrated that one way to increase H2 production is to develop improved photobioreactor (PBR) systems. Du Toit and Pott [139] compared H2 production from glycerol by Rhodopseudomonas palustris using planktonic cultures and immobilized cells in packed-bed or fluidized-bed reactors. The fluidized-bed PBR achieved a maximum specific H2 production rate and substrate conversion efficiency of 15.74 ± 2.2 mL/g/h and 43%, respectively, outperforming the conventional planktonic culture and the packed-bed PBR. There have also been attempts to use transparent polyvinyl-alcohol (PVA) cryogel beads to immobilize Rhodopseudomonas palustris for long-term H2 production [139].

4.2. Dark Fermentation

Bacteria-mediated dark fermentation is one of the most effective means of bioconverting biomass (including glycerol) to H2 [140]. The primary sub-types of such fermentation are butyrate/butanol fermentation, conducted by the genus Clostridium sp., and mixed-acid fermentation, typical of the family Enterobacteriaceae (Enterobacter aerogenes, Escherichia coli, Vibrio cholerae, Klebsiella pneumoniae, Shigella dysenteriae), Klebsiella sp., Bacillus sp., Clostridium pasteurianum, Clostridium butyricum, Thermotoga neapolitana, or mixed microbial communities [141]. In commonly used technological processes, microbial cultures are fed with glycerol medium, containing waste glycerol, nutrients (nitrogen, potassium, and other compounds), and sometimes yeast. The process is conducted in an anaerobic environment at a temperature of approx. 30–40 °C [142].
H2 release by bacteria results from pyruvate molecules being broken down into acetyl-CoA. This reaction is catalyzed by pyruvate-ferredoxin oxidoreductase. Ferredoxin plays a key role in electron distribution within the cell as an electron carrier. Reduced ferredoxin can transfer electrons to iron-containing hydrogenase, allowing protons to be used as a final electron acceptor, thus facilitating H2 production. An example flowchart of H2 production by Clostridium sp. bacteria is presented in Figure 7 [143].
Research to date has shown that the efficiency of glycerol-to-H2 conversion via dark fermentation is not affected by impurities in the starting substrate, regardless of the degree of contamination [141]. Zahedi et al. [144] found that adding 1% v/v crude glycerol to the organic fraction of municipal waste almost doubled H2 production in a process mediated by Eubacteria and Archaea. A trial using municipal waste as the sole feedstock resulted in a yield of 1.25 ± 0.29 LH2/L. By comparison, H2 production after amendment with glycerol (1% v/v) amounted to 2.32 ± 0.21 LH2/L [144]. Trchounian K. and Trchounian A. [145] conducted a dynamic and kinetic analysis of glycerol-to-H2 conversion by E. coli. Metabolic engineering of E. coli proved to directly improve dark fermentation efficiency. It was demonstrated that it was possible to redirect metabolic pathways, induce the DHAP production pathway, and block lactic, acetic acid, and ethanol production [145]. Hu and Wood [146] obtained an improved H2-producing strain that yielded 0.68 mmol H2L/hL in a glycerol medium. This represented a 20-fold increase in H2 production compared to the precursor [146]. What is more, Sanchez-Torres et al. [147] and Tran et al. [148] identified uncharacterized genes (54% of the E. coli genome has been experimentally demonstrated; the remainder is uncharacterized or has been modeled through computational methods) that can be inactivated to boost H2 production from glycerol. Some of the identified strains were able to produce 1.6 times more H2 [147,148].
There have also been studies comparing E. coli with other bacteria in terms of the optimal fermentation conditions, process parameters, and H2 yields [149]. Lo et al. [150] tested crude glycerol from the biodiesel industry for H2 production using seven isolated H2-producing bacterial strains (Clostridium pasteurianum, Clostridium butyricum, and Klebsiella sp.). Among the strains tested, C. pasteurianum CH4 had the highest H2 production performance under the optimal conditions at a temperature of 35 °C, an initial pH = 7.0, an agitation rate of 200 rpm, and a glycerol concentration of 10 g/L. When using pure glycerol as the carbon source for continuous fermentative hydrogen production, the average H2 production rate was 103.1 ± 8.1 mL/h/L, with yields of 0.50 ± 0.02 moles H2/mole glycerol. By comparison, when crude glycerol was used as the carbon source, the H2 production rate and H2 yields rose to 166.0 ± 8.7 mL/h/L and 0.77 ± 0.05 moles H2/mole glycerol, respectively [150].
Kumar et al. [151] used Bacillus thuringiensis EGU45H-2 for fermentative hydrogen production from crude glycerol, which yielded H2 at 0.646 mol H2/mole glycerol consumed in a batch culture. Under continuous culture, after 2 days’ hydraulic retention time, B. thuringiensis immobilized on lignocellulosic materials (banana leaves—BL, 10% v/v) produced 0.386 mol H2/mole glycerol consumed. The maximum yield was 0.393 mol H2/mole glycerol consumed [151]. Ito et al. [152] evaluated the production of H2 and ethanol from glycerol-containing biodiesel waste using Enterobacter aerogenes HU-101. H2 and ethanol yields were inversely correlated with glycerol levels. The highest production of H2 and ethanol—1.12 and 0.96 moles H2/mole glycerol, respectively—was obtained at 1.7 g/L glycerol [152].
Experimental reports confirm that co-digestion of mixed carbon sources—for example, formate and glycerol—can significantly increase H2 production. Trchounian et al. [153] found that H2 production by E. coli BW 25113 wild type, supplemented with glycerol, reached 0.75 ± 0.03 mmol H2/L at the end of the exponential growth phase (7 h growth). The action of formate (10 mM) and glycerol (10 g/L) at pH 6.5 led to a yield of 0.83 ± 0.05 mmol H2/L at the early exponential phase (2 h growth). The same stimulatory effect of formate and glycerol on H2 production was observed at pH 7.5 [153].
There have also been studies on how the type and concentration of heavy metals affect H2 production and E. coli growth rate [154]. It was found that E. coli possessed four [Ni-Fe]-Hyd enzymes, the activity of which requires heavy metals. Furthermore, Mo is required to activate formate hydrogen lyase (FHL). It was found that low concentrations of Ni, Fe, and Mo ions, as well as combinations thereof (Ni2+ + Fe3+ (50 mM), Ni2+ + Fe3+Mo6+ (20 mM), and Fe3+ + Mo6+ (20 mM)), could increase bacterial biomass and H2 production, mostly by way of glycerol fermentation under acidic conditions (pH 5.5 and pH 6.5). In contrast, Cu+ and Cu2+ (100 mM) were found to have no effect [154].
Pachapur et al. [155] verified the effect of added surface-active substances on H2 production from glycerol in a mixed Enterobacter aerogenes and Clostridium butyricum culture. The process yielded up to 32.1 ± 0.03 mmol H2/L medium under the optimized conditions of 17.5 g/L crude glycerol and 15 mg/L non-ionic surfactant (Tween 80). The increase in H2 production was around 1.25-fold higher with Tween 80 than without, with a production of 25.56 ± 0.91 mmol/L [155]. Faber and Ferreira-Leitão [156] endeavored to optimize the conditions of the dark fermentation of waste glycerol. A statistical analysis showed that the optimal conditions were: pH = 5.5, glycerol concentration of 0.5 g/L, and volatile suspended solids of 8.7 g/L. H2 production in the optimized process was 2.44 mol H2/mole glycerol [156].
H2 is often combined with microbial fuel cells. Sharma et al. [157] investigated the possibilities of converting glycerol into H2 and electricity. Hydrogen bioreactors (HPB) and microbial fuel cells (MFC) were used for this purpose. A comparative analysis of pure and waste glycerol was carried out. A comparable H2 production efficiency was achieved with 0.17–0.18 mol H2/mol glycerol, regardless of the source of the substrate. The highest power density of 4579 mW/m3 was achieved with 2 g/L of pure glycerol. The observed power densities of waste glycerol ranged from 1614 to 2324 mW/m3 [157]. Chookaew et al. [158] conducted analog studies with a two-stage process combining dark fermentation with an MFC or a microbial electrolysis cell (MEC). The efficiency of H2 production was 332 mL/L. 20% of the organic matter was biodegraded in the process. The two-chamber MFC produced a power density of 92 mW/m2 and removed 50% of the COD. The columbic efficiency (CE) was 14%. When the cell was fed with 50% diluted fermentation product, similar power output (90 mW/m2) and COD removal (49%) values were achieved, but the CE was almost doubled (27%). When similar substrates were used to produce H2 in two-chamber MECs, the diluted influent provided better performance, with the yield reaching a maximum of 106 mL H2/g COD and 24% CE. Thus, dark fermentation coupled with MFC/MEC proved to be a viable option for converting waste glycerol into bioenergy [158].
Fermentative H2 production from glycerol is influenced by many factors and system parameters, the most important being: glycerol type, origin and composition, presence of impurities, hydraulic retention time, type of digester, pH, temperature, and microbial strain [159]. Even trace amounts of oxygen in the system can inhibit hydrogenase activity when obligate anaerobes are used. Therefore, it is usually a safer choice to use facultative anaerobes, such as Clostridium sp. and Enterobacter sp., which better tolerate oxygen in the bioreactor [160]. The optimal pH range for efficient hydrogen-producing fermentation of glycerol is 5.0 to 6.0. Any increase in pH above that range can induce methanogenic bacteria to grow, consuming H2 to produce CH4 [161]. On the other hand, lower pH values cause microbes to switch their metabolism towards other biochemical processes. This leads to a change in the composition of the gaseous products and impaired H2 production. Additionally, a pH below 4.0 can inhibit microbial growth [162].
Heat treatment at 80–104 °C is a common processing step used to eliminate methanogenic bacteria from the anaerobic sludge communities [163]. This conditioning of anaerobic microflora ensures the survival of hydrogenous spore-forming microbes, including Bacillus sp. and Clostridium sp. [164]. Short hydraulic retention times (HRT of 12 h or less) can be used to limit the growth of microbes that compete with H2-producing bacteria [165].
The efficiency of H2 production can be hampered by an excess of undissociated volatile fatty acids (VFAs) generated in the reactor [166]. Excessive molecular pressure in the gas phase of the reactor is another inhibitor of fermentative hydrogen production. Too much H2 in the system causes the accumulation of propionic acid and butyric acid, reducing H2 yields. Reducing the pressure—and thus the H2 concentration—can significantly improve performance [167]. Another crucial parameter for fermentative hydrogen production is the concentration of iron (involved in hydrogenase activity), which should be in the range of 10 to 100 mg/L. Nitrogen levels are also important since N is an essential nutrient for microbes [168].

4.3. Mixed Fermentation

A combination of dark and photofermentation has been used to enhance H2 production [169]. Dark fermentation produces alcohols and organic acids as by-products, which serve as a carbon source for the bacteria involved in photofermentative H2 production [170]. Integrated biological processes significantly improve the ratio of energy stored in H2 to the energy needed to maintain the culture. The literature data show that ratios of up to 3.0 are achievable [171,172]. Chookaew et al. [158,173] studied H2 production from crude glycerol and experimented with a two-step process involving dark fermentation followed by photofermentation [174]. The strains used for the two-steps were Klebsiella sp. TR17 and Rhodopseudomonas palustris TN1, respectively. The optimal results of photofermentation were obtained with diluted wastewater, added glutamate, and no added yeast. Under these conditions, total H2 production across the two-step process was 6.42 mmol/g COD—10.4% of the theoretical yield [174].
Rodrigues et al. [175] evaluated H2 production in an integrated biosystem through dark fermentation followed by photofermentation, using crude glycerol from biodiesel production from waste cooking oils (WCO). A mixed Clostridiales culture was able to take up crude glycerol during dark fermentation at 37 °C and pH 5.5, generating a high H2 output of 22.38 mmol H2/L with yields of 1.75 mol H2/mole glycerol (which corresponds to 24.06 mmol H2/g COD). The dark fermentation effluent (DFE) in concentrations of 1.0, 2.0, and 3.0 g COD/L was used to grow a phototrophic bacteria culture of the order Rhizobiales and Clostridiales at 37 °C, pH 7.0, and 18.50 W/m2. The optimum conditions were achieved with 1.0 g COD/L DFE, generating 3.94 mmol H2/g COD consumed, and removing 76.10% COD. The culture was able to absorb ethanol (76.86%), acetic acid (95.73%), butyric acid (94.76%), and particularly methanol (99.18%) [175].
Sarma et al. [176] investigated the technical and economic aspects of two-step H2 production using dark fermentation followed by photofermentation. The authors concluded that 1 kg of crude glycerol can be converted into the equivalent of 2.56 L, but at a cost of USD 330. The media components accounted for 82% of production costs, meaning that the reduction in these costs is crucial to making the process viable. The process was also found to be very environmentally beneficial. It was demonstrated that the conversion of 1 kg of crude glycerol reduces greenhouse gases by 7.66 kg [176]. Examples of glycerol use in fermentative hydrogen production, key process parameters, and performance metrics are given in Table 2.

5. Biohythane Production

Coupling H2 production with CH4 production—a solution known as “biohythane”—can optimize the reuse of leftover organic matter from fermentative hydrogen production and thus help recover gas energy [187]. The leftover spent media from dark fermentative H2 production contains high levels of short-chain fatty acids such as acetate, butyrate, propionate, etc., which could be a suitable feedstock for methanogens [188]. Before the spent media are subjected to biomethanation, their pH should be adjusted to a range of 7 to 7.8. Additionally, the H2 dissolved in the media also promotes the growth of hydrogenotrophic methanogens. The recovery of gaseous energy in the form of H2 alone might not be enough to make the process commercially viable, since only 20–30% of total energy can be recovered through H2 production. In this light, integrated biohythane production seems to be a good choice since it provides an attractive method of converting organic residues rich in carbohydrates, lipids, and proteins for clean energy generation. On the other hand, these systems present challenges of their own, including the shading effect of pigments produced by photofermentative organisms and problems with scale-up [188].
Two-step AD has been acknowledged to be a promising method of biohythane production, as it enables reduced organic loading and increased total energy conversion efficiency by generating two gases with high fuel-burning capacity [189]. By separating the hydrolysis/acidogenesis phase from the methanogenesis phase, the stability of the whole process can be improved—controlling the acidification phase at the H2 generation stage (first stage) can prevent the inhibition of the methanogen population during the methane generation phase (second stage) [190]. Two-step biohythane production has been carried out using feedstocks such as maize silage [191], algal biomass [192], food waste [193], dairy waste [194], organic fraction of municipal sewage wastewater [195], cassava stillage [196], tequila vinasses [197], and skim latex serum [198]. There are few reports on the use of glycerol for biohythane production, even though glycerol is a very promising substrate due to its high content of carbon and macronutrients. Jiang et al. [199] conducted a two-step AD process to degrade glycerol. The system produced H2 and CH4 yields of 0.026 L H2/g COD and 0.29 L CH4/g COD removed, respectively. In addition, 93% COD removal was noted [199].
Co-digestion with glycerol can boost biohythane production at the optimal C/N range of 19 to 41 and the right H2/CH4 ratio of 0.06 to 0.3 [200]. Jehlee et al. [200] used glycerol as a co-substrate with Chlorella sp. TISTR 8411 biomass for biohythane production. Mono-digestion of Chlorella sp. TISTR 8411 biomass produced H2 and CH4 yields of 36.4 mL H2/g VS and 166.18 mL CH4/g VS, respectively. Co-digestion of Chlorella biomass brought a marked improvement in biohythane production. The optimal dose of 2% TS glycerol waste resulted in 39.8 mL H2/g VS and 577.33 mL CH4/g VS being produced [200]. Dounavis et al. [201] developed a continuous process for the production of biohythane from crude glycerol in a two-step reactor system. In the first step, H2 production was tested using mixed acidogenic communities in an up-flow column bioreactor. Cylindrical ceramic beads with a porosity of 600 m2/L were used as an attachment matrix for bacterial cells. The experimental parameters were: HRT 24 h; pH 6, 6.5, and 7; and a glycerol concentration of 20 g/L. The effluent from the hydrogenic reactor was fed to a methanogenic continuous stirred reactor (CSTR), which was used to examine the effect of organic loading on CH4 yield. The gaseous phase of the reactors was mixed to produce biohythane. At 20 g/L glycerol and 6, 6.5, and 7 feedstock pH, the H2 output was 0.051, 0.070, and 0.094 L/g COD, respectively. CH4 was produced at yields of 0.257 L/g COD (commercial glycerol), 0.283 L/g COD (crude glycerol), 0.198, 0.242, and 0.273 L/g COD (effluents from the hydrogenogenic stage (1st stage), diluted with water to 5, 7.5, and 10 g COD/L, respectively) [201].
Sittijunda et al. [202] produced biohythane by co-fermentation of crude glycerol and microalgae biomass. The focus was on optimizing the technological parameters that determine the efficiency of H2 production, including the initial glycerol concentration, the amount of anaerobic bacterial inoculum used, and the microalgae biomass. The highest fermentation efficiency of 655.1 mL H2/L was achieved with 13.83 g/L glycerol, 23.1 g VS/L algal biomass, and 10.3% (v/v) anaerobic sludge. The substrate for CH4 production was the effluent from the bioreactor for hydrogen fermentation. The highest yield and kinetics of CH4 production were 868.7 mL CH4/L and 2.95 mL CH4/L−h, respectively. The use of a sequence of hydrogen and methane reactors proved to be a more effective solution than the single-stage process and provided a total additional energy of 1.27 kJ/g VS [202].
Silva et al. [94] studied H2 and CH4 production in a two-step AD process with co-digestion of food waste, sewage sludge, and glycerol under mesophilic conditions (35 °C). The study evaluated the effect of adding 1 and 3% v/v glycerol as a co-substrate in three-component mixtures, keeping the concentration of all substrates at 10 g VS/L. Glycerol was found to increase H2 production under all tested conditions. The highest H2 yield (179.3 mL H2/g VS) was obtained for the 3% glycerol mixture. In terms of CH4 output, however, the 1% glycerol batch performed the best at 342 mL CH4/g VS generated. At 3% glycerol, abrupt reductions in biogas pH and CH4 content were observed, stemming from unstable methanogenesis. When taking into account both the H2 and CH4 production stages, the three-component mixture with 1% glycerol showed the highest energy production, yielding 15.5 kJ/g VS [94].
Microbes capable of producing H2 and CH4 have been identified across a wide temperature range. The majority of the literature reports on H2 and CH4 production focus on mesophilic dark fermentation [203]. The digester operating temperature should be adapted to the microbes used. Methanogens are the most sensitive to temperature changes. Even minute temperature shifts by 2–3 °C can lead to methanogen suppression and promote acidogen formation, causing VFA accumulation in the system. This, in turn, can cause pH to drop below 7 and inhibit methanogenesis [204]. Kanchanasuta and Sillaparassamee [205] have posited that crude glycerol could be used not only for co-digestion but also as a pH-adjuster for fermentative hydrogen production (to maintain pH > 5 throughout the fermentation process). Examples of glycerol use in biohythane production, key process parameters, and performance metrics are given in Table 3.

6. Summary and Conclusions

Applying waste organic substrates, such as glycerol, for industrial purposes can help reduce industrial greenhouse gas emissions and significantly decrease the cost of commercial production of alternative fuels. It is also fully in line with the principles of the circular economy and the ongoing trend of limiting and fixing CO2 emissions. Biodiesel production is one of the major sources of waste glycerol. Given that the glycerol generated in the process is very costly to purify, alternative solutions are being sought. Waste glycerol can be harnessed for bioenergy production by using it as a feedstock for the fermentative production of gas biofuels, including CH4, H2, and biohythane.
Though our understanding of glycerol fermentation is quite extensive and well systematized, there is still a need for further research with a view to optimizing processes, upgrading existing technologies, and improving final performance. With regard to the microbial fermentation of crude glycerol into energy products, it is crucial to develop more efficient technologies by utilizing new microbial strains and adapting existing ones, improving their selectivity and durability, and, in particular, bolstering their tolerance to high levels of volatile fatty acids (VFAs) in the medium and to impurities in the glycerol. Important considerations include selecting the best ratio of crude glycerol to other organic waste in anaerobic processing and developing innovative projects, as well as modifying operational guidelines for glycerol-to-biogas bioconversion systems. It is important to optimize operation conditions and running parameters, such as pH, temperature, substrate concentration, type of dosing/stirring, hydraulic retention time, and the organic load rate.
Another prerequisite to ensuring stability and good performance of digestion processes lies in minimizing the adverse effect of by-products/intermediates accumulating due to imbalances in anaerobic biochemical processes. Currently, the most popular solution is to adopt two-step digestion processes, hybrid systems, systems that integrate dark fermentation with photofermentation processes, and mixtures of multiple bacterial strains. Another important avenue of research is to explore the use of pre-treatment techniques to boost the performance of biological anaerobic processes, both in terms of biogas yield/quality and the degree of digestate mineralization/stabilization (an important factor for subsequent digestate reuse). An important aspect in this regard is to improve the overall cost-effectiveness of the process.
It should be emphasized that the use of waste glycerol as a substrate in methane fermentation processes is well recognized and is characterized by a high technological readiness level. This is proven by plants operating efficiently on a technical scale. In this case, the basic technological parameters of the process have already been determined, the final results obtained have already been estimated fairly accurately, and the functional features of the plant have been verified. The use of this organic substrate for the production of biohydrogen and biohythane is still at the laboratory-scale research level. Due to the environmentally friendly properties of these gaseous biofuels and, thus, their high competitiveness compared to methane, this is a promising research direction. The next step towards their realization should be to increase their level of technological maturity by starting with pilot-scale research and then developing technical and technological guidelines for large-scale plants. Raising the technological readiness level will allow an objective and reliable assessment of the technical and environmental effectiveness as well as the economic viability of these solutions.

Author Contributions

Conceptualization, J.K. and M.D.; methodology, J.K. and M.D.; validation, J.K.; formal analysis, J.K. and M.D.; investigation, J.K., M.D., M.Z., S.K. and J.C.S.; resources, J.K., M.D., M.Z., S.K. and J.C.S.; data curation, J.K., M.D., M.Z., S.K. and J.C.S.; supervision, J.K. and M.D.; writing—original draft preparation, J.K. and M.D.; writing—review and editing, J.K., M.D., M.Z., S.K. and J.C.S.; visualization, J.K. and M.D.; funding acquisition, J.K., M.D. and M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by works no. 29.610.023-110 of the University of Warmia and Mazury in Olsztyn and WZ/WB-IIŚ/3/2022 of the Bialystok University of Technology, funded by the Minister of Education and Science.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ballesteros, R.; Ramos, Á.; Sánchez-Valdepeñas, J. Particle-Bound PAH Emissions from a Waste Glycerine-Derived Fuel Blend in a Typical Automotive Diesel Engine. J. Energy Inst. 2020, 93, 1970–1977. [Google Scholar] [CrossRef]
  2. Mizik, T.; Gyarmati, G. Economic and Sustainability of Biodiesel Production—A Systematic Literature Review. Clean Technol. 2021, 3, 19–36. [Google Scholar] [CrossRef]
  3. Polprasert, S.; Choopakar, O.; Elefsiniotis, P. Bioethanol Production from Pretreated Palm Empty Fruit Bunch (PEFB) Using Sequential Enzymatic Hydrolysis and Yeast Fermentation. Biomass Bioenergy 2021, 149, 106088. [Google Scholar] [CrossRef]
  4. Chilakamarry, C.R.; Sakinah, A.M.M.; Zularisam, A.W.; Pandey, A. Glycerol Waste to Value Added Products and Its Potential Applications. Syst. Microbiol. Biomanufact. 2021, 1, 378–396. [Google Scholar] [CrossRef]
  5. OECD-FAO. Agricultural Outlook 2022–2031 by Variable. Available online: https://stats.oecd.org/index.aspx?queryid=30104 (accessed on 1 December 2023).
  6. Chilakamarry, C.R.; Sakinah, A.M.M.; Zularisam, A.W. Opportunities of Biodiesel Industry Waste Conversion into Value-Added Products. Mater. Today Proc. 2022, 57, 1014–1020. [Google Scholar] [CrossRef]
  7. Moklis, M.H.; Cheng, S.; Cross, J.S. Current and Future Trends for Crude Glycerol Upgrading to High Value-Added Products. Sustainability 2023, 15, 2979. [Google Scholar] [CrossRef]
  8. Lima, P.J.M.; da Silva, R.M.; Neto, C.A.C.G.; Gomes e Silva, N.C.; da Souza, J.E.S.; Nunes, Y.L.; Sousa dos Santos, J.C. An Overview on the Conversion of Glycerol to Value-Added Industrial Products via Chemical and Biochemical Routes. Biotechnol. Appl. Biochem. 2022, 69, 2794–2818. [Google Scholar] [CrossRef]
  9. Malaika, A.; Ptaszyńska, K.; Kozłowski, M. Conversion of Renewable Feedstock to Bio-Carbons Dedicated for the Production of Green Fuel Additives from Glycerol. Fuel 2021, 288, 119609. [Google Scholar] [CrossRef]
  10. Tabassum, N.; Pothu, R.; Pattnaik, A.; Boddula, R.; Balla, P.; Gundeboyina, R.; Challa, P.; Rajesh, R.; Perugopu, V.; Mameda, N.; et al. Heterogeneous Catalysts for Conversion of Biodiesel-Waste Glycerol into High-Added-Value Chemicals. Catalysts 2022, 12, 767. [Google Scholar] [CrossRef]
  11. Le, H.T.Q.; Nguyen, A.D.; Park, Y.R.; Lee, E.Y. Sustainable Biosynthesis of Chemicals from Methane and Glycerol via Reconstruction of Multi-Carbon Utilizing Pathway in Obligate Methanotrophic Bacteria. Microb. Biotechnol. 2021, 14, 2552–2565. [Google Scholar] [CrossRef]
  12. Minh Loy, A.C.; Ng, W.L.; Samudrala, S.P.; Bhattacharya, S. Technical, Economic, and Environmental Potential of Glycerol Hydrogenolysis: A Roadmap towards Sustainable Green Chemistry Future. Sustain. Energy Fuels 2023, 7, 2653–2669. [Google Scholar] [CrossRef]
  13. Tey, K.Y.; Tan, J.P.; Yeap, S.K.; He, N.; Bukhari, N.A.; Hui, Y.W.; Luthfi, A.A.I.; Manaf, S.F.A. Current Analysis on 1,3-Propanediol Production from Glycerol via Pure Wild Strain Fermentation. J. Environ. Chem. Eng. 2023, 11, 110998. [Google Scholar] [CrossRef]
  14. Mahendrasinh Kosamia, N.; Samavi, M.; Piok, K.; Kumar Rakshit, S. Perspectives for Scale up of Biorefineries Using Biochemical Conversion Pathways: Technology Status, Techno-Economic, and Sustainable Approaches. Fuel 2022, 324, 124532. [Google Scholar] [CrossRef]
  15. Ko, Y.S.; Kim, J.W.; Lee, J.A.; Han, T.; Kim, G.B.; Park, J.E.; Lee, S.Y. Tools and Strategies of Systems Metabolic Engineering for the Development of Microbial Cell Factories for Chemical Production. Chem. Soc. Rev. 2020, 49, 4615–4636. [Google Scholar] [CrossRef] [PubMed]
  16. Ye, D.Y.; Moon, J.H.; Jung, G.Y. Recent Progress in Metabolic Engineering of Escherichia Coli for the Production of Various C4 and C5-Dicarboxylic Acids. J. Agric. Food Chem. 2023, 71, 10916–10931. [Google Scholar] [CrossRef] [PubMed]
  17. Agrawal, D.; Budakoti, M.; Kumar, V. Strategies and Tools for the Biotechnological Valorization of Glycerol to 1, 3-Propanediol: Challenges, Recent Advancements and Future Outlook. Biotechnol. Adv. 2023, 66, 108177. [Google Scholar] [CrossRef] [PubMed]
  18. Igbokwe, V.C.; Ezugworie, F.N.; Onwosi, C.O.; Aliyu, G.O.; Obi, C.J. Biochemical Biorefinery: A Low-Cost and Non-Waste Concept for Promoting Sustainable Circular Bioeconomy. J. Environ. Manag. 2022, 305, 114333. [Google Scholar] [CrossRef]
  19. Pathania, R.; Srivastava, A.; Srivastava, S.; Shukla, P. Metabolic Systems Biology and Multi-Omics of Cyanobacteria: Perspectives and Future Directions. Bioresour. Technol. 2022, 343, 126007. [Google Scholar] [CrossRef]
  20. Miyuranga, K.A.V.; Arachchige, U.S.P.R.; Jayasinghe, R.A.; Samarakoon, G. Purification of Residual Glycerol from Biodiesel Production as a Value-Added Raw Material for Glycerolysis of Free Fatty Acids in Waste Cooking Oil. Energies 2022, 15, 8856. [Google Scholar] [CrossRef]
  21. García Martín, J.F.; Torres García, M.; Álvarez Mateos, P. Special Issue on “Biodiesel Production Processes and Technology”. Processes 2022, 11, 25. [Google Scholar] [CrossRef]
  22. Fasanya, O.O.; Osigbesan, A.A.; Avbenake, O.P. Biodiesel Production from Non-Edible and Waste Lipid Sources. In Biodiesel Technology and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2021; pp. 389–427. [Google Scholar] [CrossRef]
  23. Nayab, R.; Imran, M.; Ramzan, M.; Tariq, M.; Taj, M.B.; Akhtar, M.N.; Iqbal, H.M.N. Sustainable Biodiesel Production via Catalytic and Non-Catalytic Transesterification of Feedstock Materials—A Review. Fuel 2022, 328, 125254. [Google Scholar] [CrossRef]
  24. Alagha, S.M.; Salih, R. Review the Studies of Mass Transfer and Kinetic Modeling for Production the Biodiesel by the Transesterification Method and the Impact of Some Selected Factors. IOP Conf. Ser. Earth Environ. Sci. 2023, 1232, 012014. [Google Scholar] [CrossRef]
  25. Shunaia, A.A.; Jazie, A.A. Applications of Super Critical Technology in Biodiesel Production. IOP Conf. Ser. Earth Environ. Sci. 2023, 1232, 012011. [Google Scholar] [CrossRef]
  26. Ma, Q.; Fu, K.; Zhang, J.; Li, M.; Han, X.; Chen, Z.; Ma, L.; Chang, C. New Bio-Based Polyurethane (PU) Foams Synthesized Using Crude Glycerol-Based Biopolyol and Humin-Based Byproducts from Biomass Hydrolysis. Ind. Crops Prod. 2023, 205, 117548. [Google Scholar] [CrossRef]
  27. Tardiolo, G.; Nicolò, M.S.; Drago, C.; Genovese, C.; Fava, G.; Gugliandolo, C.; D’antona, N. Orange Peel Waste as Feedstock for the Production of Glycerol-Free Biodiesel by the Microalgae Nannochloropsis Oculata. Molecules 2023, 28, 6846. [Google Scholar] [CrossRef] [PubMed]
  28. Sawasdee, V.; Vikromvarasiri, N.; Pisutpaisal, N. Optimization of Ethanol Production from Co-Substrate of Waste Glycerol and Acetic Acid by Enterobacter Aerogenes. Biomass Convers. Biorefinery 2023, 13, 10505–10512. [Google Scholar] [CrossRef]
  29. Vivek, N.; Sindhu, R.; Madhavan, A.; Anju, A.J.; Castro, E.; Faraco, V.; Pandey, A.; Binod, P. Recent Advances in the Production of Value Added Chemicals and Lipids Utilizing Biodiesel Industry Generated Crude Glycerol as a Substrate—Metabolic Aspects, Challenges and Possibilities: An Overview. Bioresour. Technol. 2017, 239, 507–517. [Google Scholar] [CrossRef]
  30. Murashchenko, L.; Abbas, C.; Dmytruk, K.; Sibirny, A. Overexpression of the Truncated Version of ILV2 Enhances Glycerol Production in Saccharomyces Cerevisiae. Yeast 2016, 33, 463–469. [Google Scholar] [CrossRef]
  31. Stoumpou, V.; Novakovic, J.; Kontogianni, N.; Barampouti, E.M.; Mai, S.; Moustakas, K.; Malamis, D.; Loizidou, M. Assessing Straw Digestate as Feedstock for Bioethanol Production. Renew. Energy 2020, 153, 261–269. [Google Scholar] [CrossRef]
  32. Osman, A.I.; Qasim, U.; Jamil, F.; Al-Muhtaseb, A.H.; Jrai, A.A.; Al-Riyami, M.; Al-Maawali, S.; Al-Haj, L.; Al-Hinai, A.; Al-Abri, M.; et al. Bioethanol and Biodiesel: Bibliometric Mapping, Policies and Future Needs. Renew. Sustain. Energy Rev. 2021, 152, 111677. [Google Scholar] [CrossRef]
  33. Blomberg, A.; Morrissey, J. Yeast Osmoregulation—Glycerol Still in Pole Position. FEMS Yeast Res. 2022, 22, foac035. [Google Scholar] [CrossRef] [PubMed]
  34. Tse, T.J.; Wiens, D.J.; Reaney, M.J.T. Production of Bioethanol—A Review of Factors Affecting Ethanol Yield. Fermentation 2021, 7, 268. [Google Scholar] [CrossRef]
  35. Moshtagh, B.; Hawboldt, K.; Zhang, B. Biosurfactant Production by Native Marine Bacteria (Acinetobacter Calcoaceticus P1-1A) Using Waste Carbon Sources: Impact of Process Conditions. Can. J. Chem. Eng. 2021, 99, 2386–2397. [Google Scholar] [CrossRef]
  36. Zahid, I.; Ayoub, M.; Abdullah, B.B.; Nazir, M.H.; Ameen, M.; Zulqarnain; Mohd Yusoff, M.H.; Inayat, A.; Danish, M. Production of Fuel Additive Solketal via Catalytic Conversion of Biodiesel-Derived Glycerol. Ind. Eng. Chem. Res. 2020, 59, 20961–20978. [Google Scholar] [CrossRef]
  37. Asopa, R.P.; Bhoi, R.; Saharan, V.K. Valorization of Glycerol into Value-Added Products: A Comprehensive Review on Biochemical Route. Bioresour. Technol. Rep. 2022, 20, 101290. [Google Scholar] [CrossRef]
  38. Pirzadi, Z.; Meshkani, F. From Glycerol Production to Its Value-Added Uses: A Critical Review. Fuel 2022, 329, 125044. [Google Scholar] [CrossRef]
  39. Dimian, A.C.; Kiss, A.A. Eco-Efficient Processes for Biodiesel Production from Waste Lipids. J. Clean. Prod. 2019, 239, 118073. [Google Scholar] [CrossRef]
  40. Kongjan, P.; Reungsang, A.; Sittijunda, S. Conversion of Glycerol Derived from Biodiesel Production to Butanol and 1,3-Propanediol. In Chemical Substitutes from Agricultural and Industrial By-Products; John Wiley & Sons: Hoboken, NJ, USA, 2023; pp. 337–353. [Google Scholar] [CrossRef]
  41. Ayude, M.A.; Doumic, L.I.; Cassanello, M.C.; Nigam, K.D.P. Clean Catalytic Oxidation for Derivatization of Key Biobased Platform Chemicals: Ethanol, Glycerol, and Hydroxymethyl Furfural. Ind. Eng. Chem. Res. 2019, 58, 16077–16095. [Google Scholar] [CrossRef]
  42. Sedghi, R.; Shahbeik, H.; Rastegari, H.; Rafiee, S.; Peng, W.; Nizami, A.S.; Gupta, V.K.; Chen, W.H.; Lam, S.S.; Pan, J.; et al. Turning Biodiesel Glycerol into Oxygenated Fuel Additives and Their Effects on the Behavior of Internal Combustion Engines: A Comprehensive Systematic Review. Renew. Sustain. Energy Rev. 2022, 167, 112805. [Google Scholar] [CrossRef]
  43. Salgaonkar, N.; Kadamkode, V.; Kumaran, S.; Mallemala, P.; Christy, E.; Appavoo, S.; Majumdar, A.; Mitra, R.; Dasgupta, A. Glycerol Fermentation by Skin Bacteria Generates Lactic Acid and Upregulates the Expression Levels of Genes Associated with the Skin Barrier Function. Exp. Dermatol. 2022, 31, 1364–1372. [Google Scholar] [CrossRef]
  44. Guerfali, M.; Ayadi, I.; Sassi, H.E.; Belhassen, A.; Gargouri, A.; Belghith, H. Biodiesel-Derived Crude Glycerol as Alternative Feedstock for Single Cell Oil Production by the Oleaginous Yeast Candida Viswanathii Y-E4. Ind. Crops Prod. 2020, 145, 112103. [Google Scholar] [CrossRef]
  45. Marciniak, P.; Możejko-Ciesielska, J. What Is New in the Field of Industrial Wastes Conversion into Polyhydroxyalkanoates by Bacteria? Polymers 2021, 13, 1731. [Google Scholar] [CrossRef]
  46. Banaszuk, P.; Kamocki, A.K.; Wysocka-Czubaszek, A.; Czubaszek, R.; Roj-Rojewski, S. Closing the Loop—Recovery of Nutrients and Energy from Wetland Biomass. Ecol. Eng. 2020, 143, 105643. [Google Scholar] [CrossRef]
  47. Kisielewska, M.; Dębowski, M.; Zieliński, M.; Kazimierowicz, J.; Quattrocelli, P.; Bordiean, A. Effects of Liquid Digestate Treatment on Sustainable Microalgae Biomass Production. Bioenergy Res. 2022, 15, 357–370. [Google Scholar] [CrossRef]
  48. Teleszewski, T.J.; Z’ukowski, M. Analysis of Heat Loss of a Biogas Anaerobic Digester in Weather Conditions in Poland. J. Ecol. Eng. 2018, 19, 242–250. [Google Scholar] [CrossRef]
  49. Czubaszek, R.; Wysocka-Czubaszek, A.; Wichtmann, W.; Banaszuk, P. Specific Methane Yield of Wetland Biomass in Dry and Wet Fermentation Technologies. Energies 2021, 14, 8373. [Google Scholar] [CrossRef]
  50. Czubaszek, R.; Wysocka-Czubaszek, A.; Banaszuk, P. GHG Emissions and Efficiency of Energy Generation through Anaerobic Fermentation of Wetland Biomass. Energies 2020, 13, 6497. [Google Scholar] [CrossRef]
  51. Teleszewski, T.; Żukowski, M. The Influence of Sludge on Thermal Performance of Heat Exchanger Tubes Inside in an Anaerobic Digester. Annu. Set Environ. Prot. 2018, 20, 763–779. [Google Scholar]
  52. Kazimierowicz, J.; Dębowski, M. Aerobic Granular Sludge as a Substrate in Anaerobic Digestion—Current Status and Perspectives. Sustainability 2022, 14, 10904. [Google Scholar] [CrossRef]
  53. Khalid, S.A.; Elsherif, W.M. Types of Microorganisms for Biodegradation. In Handbook of Biodegradable Materials; Springer: Cham, Switzerland, 2022; pp. 1–27. [Google Scholar] [CrossRef]
  54. Gnanasekaran, L.; Priya, A.K.; Thanigaivel, S.; Hoang, T.K.A.; Soto-Moscoso, M. The Conversion of Biomass to Fuels via Cutting-Edge Technologies: Explorations from Natural Utilization Systems. Fuel 2023, 331, 125668. [Google Scholar] [CrossRef]
  55. 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]
  56. Cardoso, W.; di Felice, R.; Baptista, R.C. A Critical Overview of Development and Innovations in Biogas Upgrading. Smart Innov. Syst. Technol. 2022, 295 SIST, 42–50. [Google Scholar] [CrossRef]
  57. Vítězová, M.; Kohoutová, A.; Vítěz, T.; Hanišáková, N.; Kushkevych, I. Methanogenic Microorganisms in Industrial Wastewater Anaerobic Treatment. Processes 2020, 8, 1546. [Google Scholar] [CrossRef]
  58. Ahmad, I. Implementation and Optimizing Methane Content in Biogas for the Production of Electricity. Int. J. Eng. Res. Technol. 2015, 4, 1326–1329. [Google Scholar]
  59. Santibáñez, C.; Varnero, M.T.; Bustamante, M. Glicerol Residual de La Producción de Biodiesel, Residuo o Potencial Fuente de Energía: Una Revisión. Chil. J. Agric. Res. 2011, 71, 469–475. [Google Scholar] [CrossRef]
  60. Battista, F.; Strazzera, G.; Valentino, F.; Gottardo, M.; Villano, M.; Matos, M.; Silva, F.M.; Reis, M.A.; Mata-Alvarez, J.; Astals, S.; et al. New Insights in Food Waste, Sewage Sludge and Green Waste Anaerobic Fermentation for Short-Chain Volatile Fatty Acids Production: A Review. J. Environ. Chem. Eng. 2022, 10, 108319. [Google Scholar] [CrossRef]
  61. Adames, L.V.; Jacobus, A.P.; Sakamoto, I.K.; Lazaro, C.Z.; Pires, L.O.; Maintinguer, S.I. Bioenergy Recovery from Anaerobic Co-Digestion of Crude Glycerol and Domestic Sewage In-Series Reactor: Microbial Characterization and System Performance. Bioenergy Res. 2022, 15, 2145–2158. [Google Scholar] [CrossRef]
  62. Ferreira, C.M.; Akisue, R.A.; de Sousa Júnior, R. Mathematical Modeling and Computational Simulation Applied to the Study of Glycerol and/or Molasses Anaerobic Co-Digestion Processes. Processes 2023, 11, 2121. [Google Scholar] [CrossRef]
  63. Adames, L.V.; Pires, L.O.; Maintinguer, S.I. Continuous Long-Term Anaerobic Co-Digestion of Crude Glycerol and Domestic Sewage: Plug-Flow In-Series Reactor Performance and Microbiota Acclimatization. Bioenergy Res. 2023, 16, 1876–1888. [Google Scholar] [CrossRef]
  64. Bułkowska, K.; Mikucka, W.; Pokój, T. Enhancement of Biogas Production from Cattle Manure Using Glycerine Phase as a Co-Substrate in Anaerobic Digestion. Fuel 2022, 317, 123456. [Google Scholar] [CrossRef]
  65. Dębowski, M.; Zieliński, M.; Kazimierowicz, J.; Kujawska, N.; Talbierz, S. Microalgae Cultivation Technologies as an Opportunity for Bioenergetic System Development—Advantages and Limitations. Sustainability 2020, 12, 9980. [Google Scholar] [CrossRef]
  66. Catenacci, A.; Boniardi, G.; Mainardis, M.; Gievers, F.; Farru, G.; Asunis, F.; Malpei, F.; Goi, D.; Cappai, G.; Canziani, R. Processes, Applications and Legislative Framework for Carbonized Anaerobic Digestate: Opportunities and Bottlenecks. A Critical Review. Energy Convers. Manag. 2022, 263, 115691. [Google Scholar] [CrossRef]
  67. Fan, Q. Methane Potential of Municipal Sludge in Anaerobic Co- Digestion Process Boosted with Glycerol. Ph.D. Thesis, Savonia-ammattikorkeakoulu, Kuopio, Finland, 2018. [Google Scholar]
  68. Sittijunda, S.; Reungsang, A. Methane Production from the Co-Digestion of Algal Biomass with Crude Glycerol by Anaerobic Mixed Cultures. Waste Biomass Valorization 2020, 11, 1873–1881. [Google Scholar] [CrossRef]
  69. Dos Santos Ferreira, J.; Volschan, I.; Cammarota, M.C. Co-Digestion of Sewage Sludge with Crude or Pretreated Glycerol to Increase Biogas Production. Environ. Sci. Pollut. Res. 2018, 25, 21811–21821. [Google Scholar] [CrossRef] [PubMed]
  70. Zahedi, S.; Rivero, M.; Solera, R.; Perez, M. Mesophilic Anaerobic Co-Digestion of Sewage Sludge with Glycerine: Effect of Solids Retention Time. Fuel 2018, 215, 285–289. [Google Scholar] [CrossRef]
  71. He, Q.; McNutt, J.; Yang, J. Utilization of the Residual Glycerol from Biodiesel Production for Renewable Energy Generation. Renew. Sustain. Energy Rev. 2017, 71, 63–76. [Google Scholar] [CrossRef]
  72. Nartker, S.; Ammerman, M.; Aurandt, J.; Stogsdil, M.; Hayden, O.; Antle, C. Increasing Biogas Production from Sewage Sludge Anaerobic Co-Digestion Process by Adding Crude Glycerol from Biodiesel Industry. Waste Manag. 2014, 34, 2567–2571. [Google Scholar] [CrossRef]
  73. Baba, Y.; Tada, C.; Watanabe, R.; Fukuda, Y.; Chida, N.; Nakai, Y. Anaerobic Digestion of Crude Glycerol from Biodiesel Manufacturing Using a Large-Scale Pilot Plant: Methane Production and Application of Digested Sludge as Fertilizer. Bioresour. Technol. 2013, 140, 342–348. [Google Scholar] [CrossRef]
  74. Athanasoulia, E.; Melidis, P.; Aivasidis, A. Co-Digestion of Sewage Sludge and Crude Glycerol from Biodiesel Production. Renew. Energy 2014, 62, 73–78. [Google Scholar] [CrossRef]
  75. Nghiem, L.D.; Nguyen, T.T.; Manassa, P.; Fitzgerald, S.K.; Dawson, M.; Vierboom, S. Co-Digestion of Sewage Sludge and Crude Glycerol for on-Demand Biogas Production. Int. Biodeterior. Biodegrad. 2014, 95, 160–166. [Google Scholar] [CrossRef]
  76. Silvestre, G.; Fernández, B.; Bonmatí, A. Addition of Crude Glycerine as Strategy to Balance the C/N Ratio on Sewage Sludge Thermophilic and Mesophilic Anaerobic Co-Digestion. Bioresour. Technol. 2015, 193, 377–385. [Google Scholar] [CrossRef] [PubMed]
  77. Jensen, P.D.; Astals, S.; Lu, Y.; Devadas, M.; Batstone, D.J. Anaerobic Codigestion of Sewage Sludge and Glycerol, Focusing on Process Kinetics, Microbial Dynamics and Sludge Dewaterability. Water Res. 2014, 67, 355–366. [Google Scholar] [CrossRef] [PubMed]
  78. Alves, I.R.F.S.; Mahler, C.F.; Oliveira, L.B.; Reis, M.M.; Bassin, J.P. Assessing the Use of Crude Glycerol from Biodiesel Production as an Alternative to Boost Methane Generation by Anaerobic Co-Digestion of Sewage Sludge. Biomass Bioenergy 2020, 143, 105831. [Google Scholar] [CrossRef]
  79. Alves, I.R.F.S.; Mahler, C.F.; Oliveira, L.B.; Reis, M.M.; Bassin, J.P. Investigating the Effect of Crude Glycerol from Biodiesel Industry on the Anaerobic Co-Digestion of Sewage Sludge and Food Waste in Ternary Mixtures. Energy 2022, 241, 122818. [Google Scholar] [CrossRef]
  80. Siles, J.A.; Martín, M.A.; Chica, A.F.; Martín, A. Anaerobic Co-Digestion of Glycerol and Wastewater Derived from Biodiesel Manufacturing. Bioresour. Technol. 2010, 101, 6315–6321. [Google Scholar] [CrossRef] [PubMed]
  81. Fountoulakis, M.S.; Manios, T. Enhanced Methane and Hydrogen Production from Municipal Solid Waste and Agro-Industrial by-Products Co-Digested with Crude Glycerol. Bioresour. Technol. 2009, 100, 3043–3047. [Google Scholar] [CrossRef] [PubMed]
  82. Astals, S.; Nolla-Ardèvol, V.; Mata-Alvarez, J. Thermophilic Co-Digestion of Pig Manure and Crude Glycerol: Process Performance and Digestate Stability. J. Biotechnol. 2013, 166, 97–104. [Google Scholar] [CrossRef] [PubMed]
  83. Astals, S.; Nolla-Ardèvol, V.; Mata-Alvarez, J. Anaerobic Co-Digestion of Pig Manure and Crude Glycerol at Mesophilic Conditions: Biogas and Digestate. Bioresour. Technol. 2012, 110, 63–70. [Google Scholar] [CrossRef]
  84. Jasińska, A.; Grosser, A.; Meers, E. Possibilities and Limitations of Anaerobic Co-Digestion of Animal Manure—A Critical Review. Energies 2023, 16, 3885. [Google Scholar] [CrossRef]
  85. Logan, M.; Visvanathan, C. Management Strategies for Anaerobic Digestate of Organic Fraction of Municipal Solid Waste: Current Status and Future Prospects. Waste Manag. Res. 2019, 37 (Suppl. S1), 27–39. [Google Scholar] [CrossRef]
  86. Romaniuk, W.; Rogovskii, I.; Polishchuk, V.; Titova, L.; Borek, K.; Wardal, W.J.; Shvorov, S.; Dvornyk, Y.; Sivak, I.; Drahniev, S.; et al. Study of Methane Fermentation of Cattle Manure in the Mesophilic Regime with the Addition of Crude Glycerine. Energies 2022, 15, 3439. [Google Scholar] [CrossRef]
  87. Usack, J.G.; Angenent, L.T. Comparing the Inhibitory Thresholds of Dairy Manure Co-Digesters after Prolonged Acclimation Periods: Part 1—Performance and Operating Limits. Water Res. 2015, 87, 446–457. [Google Scholar] [CrossRef]
  88. Patra, F.; Duary, R.K. Waste from Dairy Processing Industries and Its Sustainable Utilization. In Sustainable Food Waste Management Concepts and Innovations; Springer: Cham, Switzerland, 2021; pp. 127–154. [Google Scholar] [CrossRef]
  89. Do Borges, A.V.; Fuess, L.T.; Takeda, P.Y.; Alves, I.; Dias, M.E.S.; Damianovic, M.H.R.Z. Co-Digestion of Biofuel by-Products: Enhanced Biofilm Formation Maintains High Organic Matter Removal When Methanogenesis Fails. J. Environ. Manag. 2022, 310, 114768. [Google Scholar] [CrossRef] [PubMed]
  90. Haider, U.; Munir, A.; Ghafoor, A. Design of Biogas Fermentation Chamber And Techniques To Enrich Bio-Methanation. Artic. Pak. J. Agric. Res. 2021, 56, 1617–1627. [Google Scholar]
  91. Li, Y.; Chen, Y.; Wu, J. Enhancement of Methane Production in Anaerobic Digestion Process: A Review. Appl. Energy 2019, 240, 120–137. [Google Scholar] [CrossRef]
  92. Fierro, J.; Martinez, E.J.; Rosas, J.G.; Fernández, R.A.; López, R.; Gomez, X. Co-Digestion of Swine Manure and Crude Glycerine: Increasing Glycerine Ratio Results in Preferential Degradation of Labile Compounds. Water Air Soil Pollut. 2016, 227, 78. [Google Scholar] [CrossRef]
  93. Prasertsan, P.; Leamdum, C.; Chantong, S.; Mamimin, C.; Kongjan, P.; O-Thong, S. Enhanced Biogas Production by Co-Digestion of Crude Glycerol and Ethanol with Palm Oil Mill Effluent and Microbial Community Analysis. Biomass Bioenergy 2021, 148, 106037. [Google Scholar] [CrossRef]
  94. Silva, F.M.S.; Mahler, C.F.; Oliveira, L.B.; Bassin, J.P. Hydrogen and Methane Production in a Two-Stage Anaerobic Digestion System by Co-Digestion of Food Waste, Sewage Sludge and Glycerol. Waste Manag. 2018, 76, 339–349. [Google Scholar] [CrossRef] [PubMed]
  95. Sevillano, C.A.; Pesantes, A.A.; Peña Carpio, E.; Martínez, E.J.; Gómez, X. Anaerobic Digestion for Producing Renewable Energy—The Evolution of This Technology in a New Uncertain Scenario. Entropy 2021, 23, 145. [Google Scholar] [CrossRef]
  96. Kazimierowicz, J.; Dzienis, L.; Dębowski, M.; Zieliński, M. Optimisation of Methane Fermentation as a Valorisation Method for Food Waste Products. Biomass Bioenergy 2021, 144, 105913. [Google Scholar] [CrossRef]
  97. Zieliński, M.; Kazimierowicz, J.; Dębowski, M. Advantages and Limitations of Anaerobic Wastewater Treatment—Technological Basics, Development Directions, and Technological Innovations. Energies 2022, 16, 83. [Google Scholar] [CrossRef]
  98. Wang, S.; Li, D.; Zhang, K.; Ma, Y.; Liu, F.; Li, Z.; Gao, X.; Gao, W.; Du, L. Effects of Initial Volatile Fatty Acid Concentrations on Process Characteristics, Microbial Communities, and Metabolic Pathways on Solid-State Anaerobic Digestion. Bioresour. Technol. 2023, 369, 128461. [Google Scholar] [CrossRef] [PubMed]
  99. Viancelli, A.; Schneider, T.M.; Demczuk, T.; Delmoral, A.P.G.; Petry, B.; Collato, M.M.; Michelon, W. Unlocking the Value of Biomass: Exploring Microbial Strategies for Biogas and Volatile Fatty Acids Generation. Bioresour. Technol. Rep. 2023, 23, 101552. [Google Scholar] [CrossRef]
  100. Oliveira, J.V.; Alves, M.M.; Costa, J.C. Optimization of Biogas Production from Sargassum Sp. Using a Design of Experiments to Assess the Co-Digestion with Glycerol and Waste Frying Oil. Bioresour. Technol. 2015, 175, 480–485. [Google Scholar] [CrossRef] [PubMed]
  101. Takeda, P.Y.; Gotardo, J.T.; Gomes, S.D. Anaerobic Co-Digestion of Leachate and Glycerol for Renewable Energy Generation. Environ. Technol. 2022, 43, 1118–1128. [Google Scholar] [CrossRef] [PubMed]
  102. Farghali, M.; Mohamed, I.M.A.; Hassan, D.; Iwasaki, M.; Yoshida, G.; Umetsu, K.; Ihara, I. Kinetic Modeling of Anaerobic Co-Digestion with Glycerol: Implications for Process Stability and Organic Overloads. Biochem. Eng. J. 2023, 199, 109061. [Google Scholar] [CrossRef]
  103. Fernández-Palacios, E.; Lafuente, J.; Mora, M.; Gabriel, D. Exploring the Performance Limits of a Sulfidogenic UASB during the Long-Term Use of Crude Glycerol as Electron Donor. Sci. Total Environ. 2019, 688, 1184–1192. [Google Scholar] [CrossRef]
  104. Strzelec, E.; Chmiel, M.J. Laboratory-Scale Fermentation of Raw Glycerol with Lactobacillus Brevis WLP 672. In Contemporary Problems of Power Engineering and Environmental Protection 2022; Silesian University of Technology (PolSL): Gliwice, Poland, 2023; p. 87. [Google Scholar]
  105. González, R.; Smith, R.; Blanco, D.; Fierro, J.; Gómez, X. Application of Thermal Analysis for Evaluating the Effect of Glycerine Addition on the Digestion of Swine Manure. J. Therm. Anal. Calorim. 2019, 135, 2277–2286. [Google Scholar] [CrossRef]
  106. Kumar, B.; Verma, P. Biomass-Based Biorefineries: An Important Architype towards a Circular Economy. Fuel 2021, 288, 119622. [Google Scholar] [CrossRef]
  107. Vasconcelos, E.A.F.; Santaella, S.T.; Viana, M.B.; dos Santos, A.B.; Pinheiro, G.C.; Leitão, R.C. Composition and Ecology of Bacterial and Archaeal Communities in Anaerobic Reactor Fed with Residual Glycerol. Anaerobe 2019, 59, 145–153. [Google Scholar] [CrossRef]
  108. Wang, Z.; Watson, J.; Wang, T.; Yi, S.; Si, B.; Zhang, Y. Enhancing Energy Recovery via Two Stage Co-Fermentation of Hydrothermal Liquefaction Aqueous Phase and Crude Glycerol. Energy Convers. Manag. 2021, 231, 113855. [Google Scholar] [CrossRef]
  109. Seekao, N.; Sangsri, S.; Rakmak, N.; Dechapanya, W.; Siripatana, C. Co-Digestion of Palm Oil Mill Effluent with Chicken Manure and Crude Glycerol: Biochemical Methane Potential by Monod Kinetics. Heliyon 2021, 7, e06204. [Google Scholar] [CrossRef] [PubMed]
  110. Farias, C.B.B.; Barreiros, R.C.S.; da Silva, M.F.; Casazza, A.A.; Converti, A.; Sarubbo, L.A. Use of Hydrogen as Fuel: A Trend of the 21st Century. Energies 2022, 15, 311. [Google Scholar] [CrossRef]
  111. Kazimierowicz, J.; Dębowski, M.; Zieliński, M. Effectiveness of Hydrogen Production by Bacteroides Vulgatus in Psychrophilic Fermentation of Cattle Slurry. Clean Technol. 2022, 4, 806–814. [Google Scholar] [CrossRef]
  112. Dębowski, M.; Dudek, M.; Zieliński, M.; Nowicka, A.; Kazimierowicz, J. Microalgal Hydrogen Production in Relation to Other Biomass-Based Technologies—A Review. Energies 2021, 14, 6025. [Google Scholar] [CrossRef]
  113. Meier, K.; Kurtz, C.; Weckerle, C.; Hubner, M.; Bürger, I. Air-Conditioning System for Vehicles with on-Board Hydrogen. Appl. Therm. Eng. 2018, 129, 1150–1159. [Google Scholar] [CrossRef]
  114. Capurso, T.; Stefanizzi, M.; Torresi, M.; Camporeale, S.M. Perspective of the Role of Hydrogen in the 21st Century Energy Transition. Energy Convers. Manag. 2022, 251, 114898. [Google Scholar] [CrossRef]
  115. Tashie-Lewis, B.C.; Nnabuife, S.G. Hydrogen Production, Distribution, Storage and Power Conversion in a Hydrogen Economy—A Technology Review. Chem. Eng. J. Adv. 2021, 8, 100172. [Google Scholar] [CrossRef]
  116. Younas, M.; Shafique, S.; Hafeez, A.; Javed, F.; Rehman, F. An Overview of Hydrogen Production: Current Status, Potential, and Challenges. Fuel 2022, 316, 123317. [Google Scholar] [CrossRef]
  117. Ahmed, S.F.; Mofijur, M.; Nuzhat, S.; Rafa, N.; Musharrat, A.; Lam, S.S.; Boretti, A. Sustainable Hydrogen Production: Technological Advancements and Economic Analysis. Int. J. Hydrogen Energy 2022, 47, 37227–37255. [Google Scholar] [CrossRef]
  118. Pareek, A.; Dom, R.; Gupta, J.; Chandran, J.; Adepu, V.; Borse, P.H. Insights into Renewable Hydrogen Energy: Recent Advances and Prospects. Mater. Sci. Energy Technol. 2020, 3, 319–327. [Google Scholar] [CrossRef]
  119. Bhatia, S.K.; Jagtap, S.S.; Bedekar, A.A.; Bhatia, R.K.; Rajendran, K.; Pugazhendhi, A.; Rao, C.V.; Atabani, A.E.; Kumar, G.; Yang, Y.H. Renewable Biohydrogen Production from Lignocellulosic Biomass Using Fermentation and Integration of Systems with Other Energy Generation Technologies. Sci. Total Environ. 2021, 765, 144429. [Google Scholar] [CrossRef] [PubMed]
  120. Martinez-Burgos, W.J.; de Souza Candeo, E.; Pedroni Medeiros, A.B.; Cesar de Carvalho, J.; Oliveira de Andrade Tanobe, V.; Soccol, C.R.; Sydney, E.B. Hydrogen: Current Advances and Patented Technologies of Its Renewable Production. J. Clean. Prod. 2021, 286, 124970. [Google Scholar] [CrossRef]
  121. Akroum-Amrouche, D.; Akroum, H.; Lounici, H. Green Hydrogen Production by Rhodobacter Sphaeroides. Energy Sources Part A Recover. Util. Environ. Eff. 2023, 45, 2862–2880. [Google Scholar] [CrossRef]
  122. Sağır, E.; Hallenbeck, P.C. Photofermentative Hydrogen Production. In Biomass, Biofuels and Biochemical: Biohydrogen, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 141–157. [Google Scholar] [CrossRef]
  123. Ghiasian, M. Biophotolysis-Based Hydrogen Production by Cyanobacteria. Prospect. Renew. Bioprocess. Future Energy Syst. 2019, 10, 161–184. [Google Scholar] [CrossRef]
  124. Budiman, P.M.; Wu, T.Y. Role of Chemicals Addition in Affecting Biohydrogen Production through Photofermentation. Energy Convers. Manag. 2018, 165, 509–527. [Google Scholar] [CrossRef]
  125. Akhlaghi, N.; Najafpour-Darzi, G. A Comprehensive Review on Biological Hydrogen Production. Int. J. Hydrogen Energy 2020, 45, 22492–22512. [Google Scholar] [CrossRef]
  126. Du Toit, J.P.; Pott, R.W.M. Heat-Acclimatised Strains of Rhodopseudomonas palustris Reveal Higher Temperature Optima with Concomitantly Enhanced Biohydrogen Production Rates. Int. J. Hydrogen Energy 2021, 46, 11564–11572. [Google Scholar] [CrossRef]
  127. Pott, R.W.M.; Howe, C.J.; Dennis, J.S. Photofermentation of Crude Glycerol from Biodiesel Using Rhodopseudomonas palustris: Comparison with Organic Acids and the Identification of Inhibitory Compounds. Bioresour. Technol. 2013, 130, 725–730. [Google Scholar] [CrossRef]
  128. Sabourin-Provost, G.; Hallenbeck, P.C. High Yield Conversion of a Crude Glycerol Fraction from Biodiesel Production to Hydrogen by Photofermentation. Bioresour. Technol. 2009, 100, 3513–3517. [Google Scholar] [CrossRef]
  129. Ghosh, D.; Sobro, I.F.; Hallenbeck, P.C. Stoichiometric Conversion of Biodiesel Derived Crude Glycerol to Hydrogen: Response Surface Methodology Study of the Effects of Light Intensity and Crude Glycerol and Glutamate Concentration. Bioresour. Technol. 2012, 106, 154–160. [Google Scholar] [CrossRef] [PubMed]
  130. Zhang, D.; Xiao, N.; Mahbubani, K.T.; del Rio-Chanona, E.A.; Slater, N.K.H.; Vassiliadis, V.S. Bioprocess Modelling of Biohydrogen Production by Rhodopseudomonas palustris: Model Development and Effects of Operating Conditions on Hydrogen Yield and Glycerol Conversion Efficiency. Chem. Eng. Sci. 2015, 130, 68–78. [Google Scholar] [CrossRef]
  131. Jiang, D.; Zhang, X.; Ge, X.; Yue, T.; Zhang, T.; Zhang, Y.; Zhang, Z.; He, C.; Lu, C.; Zhang, Q. Insights into Correlation between Hydrogen Yield Improvement and Glycerol Addition in Photo-Fermentation of Arundo donax, L. Bioresour. Technol. 2021, 321, 124467. [Google Scholar] [CrossRef] [PubMed]
  132. Fu, Q.; Wang, D.; Li, X.; Yang, Q.; Xu, Q.; Ni, B.J.; Wang, Q.; Liu, X. Towards Hydrogen Production from Waste Activated Sludge: Principles, Challenges and Perspectives. Renew. Sustain. Energy Rev. 2021, 135, 110283. [Google Scholar] [CrossRef]
  133. Montiel Corona, V.; Le Borgne, S.; Revah, S.; Morales, M. Effect of Light-Dark Cycles on Hydrogen and Poly-β-Hydroxybutyrate Production by a Photoheterotrophic Culture and Rhodobacter capsulatus Using a Dark Fermentation Effluent as Substrate. Bioresour. Technol. 2017, 226, 238–246. [Google Scholar] [CrossRef] [PubMed]
  134. Carolin Christopher, F.; Kumar, P.S.; Vo, D.V.N.; Joshiba, G.J. A Review on Critical Assessment of Advanced Bioreactor Options for Sustainable Hydrogen Production. Int. J. Hydrogen Energy 2021, 46, 7113–7136. [Google Scholar] [CrossRef]
  135. Bolatkhan, K.; Kossalbayev, B.D.; Zayadan, B.K.; Tomo, T.; Veziroglu, T.N.; Allakhverdiev, S.I. Hydrogen Production from Phototrophic Microorganisms: Reality and Perspectives. Int. J. Hydrogen Energy 2019, 44, 5799–5811. [Google Scholar] [CrossRef]
  136. Bosman, C.E.; Pott, R.W.M.C.; Bradshaw, S.M. Design, Modelling and Simulation of a Thermosiphon Photobioreactor for Photofermentative Hydrogen Production. Biochem. Eng. J. 2022, 186, 108582. [Google Scholar] [CrossRef]
  137. Sirohi, R.; Kumar Pandey, A.; Ranganathan, P.; Singh, S.; Udayan, A.; Kumar Awasthi, M.; Hoang, A.T.; Chilakamarry, C.R.; Kim, S.H.; Sim, S.J. Design and Applications of Photobioreactors—A Review. Bioresour. Technol. 2022, 349, 126858. [Google Scholar] [CrossRef]
  138. Ross, B.S.; Pott, R.W.M. Hydrogen Production by Immobilized Rhodopseudomonas palustris in Packed or Fluidized Bed Photobioreactor Systems. Int. J. Hydrogen Energy 2021, 46, 1715–1727. [Google Scholar] [CrossRef]
  139. Du Toit, J.P.; Pott, R.W.M. Transparent Polyvinyl-Alcohol Cryogel as Immobilisation Matrix for Continuous Biohydrogen Production by Phototrophic Bacteria. Biotechnol. Biofuels 2020, 13, 105. [Google Scholar] [CrossRef]
  140. Aziz, M.; Darmawan, A.; Juangsa, F.B. Hydrogen Production from Biomasses and Wastes: A Technological Review. Int. J. Hydrogen Energy 2021, 46, 33756–33781. [Google Scholar] [CrossRef]
  141. Rodrigues, C.V.; Nespeca, M.G.; Sakamoto, I.K.; de Oliveira, J.E.; Amâncio Varesche, M.B.; Maintinguer, S.I. Bioconversion of Crude Glycerol from Waste Cooking Oils into Hydrogen by Sub-Tropical Mixed and Pure Cultures. Int. J. Hydrogen Energy 2019, 44, 144–154. [Google Scholar] [CrossRef]
  142. Baeyens, J.; Zhang, H.; Nie, J.; Appels, L.; Dewil, R.; Ansart, R.; Deng, Y. Reviewing the Potential of Bio-Hydrogen Production by Fermentation. Renew. Sustain. Energy Rev. 2020, 131, 110023. [Google Scholar] [CrossRef]
  143. Bernal, M.; Tinoco, L.K.; Torres, L.; Malagón-Romero, D.; Montoya, D. Evaluating Colombian Clostridium Spp. Strains’ Hydrogen Production Using Glycerol as Substrate. Electron. J. Biotechnol. 2013, 16, 6. [Google Scholar] [CrossRef]
  144. Zahedi, S.; Solera, R.; García-Morales, J.L.; Sales, D. Effect of the Addition of Glycerol on Hydrogen Production from Industrial Municipal Solid Waste. Fuel 2016, 180, 343–347. [Google Scholar] [CrossRef]
  145. Trchounian, K.; Trchounian, A. Hydrogen Production from Glycerol by Escherichia Coli and Other Bacteria: An Overview and Perspectives. Appl. Energy 2015, 156, 174–184. [Google Scholar] [CrossRef]
  146. Hu, H.; Wood, T.K. An Evolved Escherichia Coli Strain for Producing Hydrogen and Ethanol from Glycerol. Biochem. Biophys. Res. Commun. 2010, 391, 1033–1038. [Google Scholar] [CrossRef] [PubMed]
  147. Sanchez-Torres, V.; Mohd Yusoff, M.Z.; Nakano, C.; Maeda, T.; Ogawa, H.I.; Wood, T.K. Influence of Escherichia Coli Hydrogenases on Hydrogen Fermentation from Glycerol. Int. J. Hydrogen Energy 2013, 38, 3905–3912. [Google Scholar] [CrossRef]
  148. Tran, K.T.; Maeda, T.; Sanchez-Torres, V.; Wood, T.K. Beneficial Knockouts in Escherichia Coli for Producing Hydrogen from Glycerol. Appl. Microbiol. Biotechnol. 2015, 99, 2573–2581. [Google Scholar] [CrossRef]
  149. Trchounian, A. Mechanisms for Hydrogen Production by Different Bacteria during Mixed-Acid and Photo-Fermentation and Perspectives of Hydrogen Production Biotechnology. Crit. Rev. Biotechnol. 2015, 35, 103–113. [Google Scholar] [CrossRef] [PubMed]
  150. Lo, Y.C.; Chen, X.J.; Huang, C.Y.; Yuan, Y.J.; Chang, J.S. Dark Fermentative Hydrogen Production with Crude Glycerol from Biodiesel Industry Using Indigenous Hydrogen-Producing Bacteria. Int. J. Hydrogen Energy 2013, 38, 15815–15822. [Google Scholar] [CrossRef]
  151. Kumar, P.; Sharma, R.; Ray, S.; Mehariya, S.; Patel, S.K.S.; Lee, J.K.; Kalia, V.C. Dark Fermentative Bioconversion of Glycerol to Hydrogen by Bacillus Thuringiensis. Bioresour. Technol. 2015, 182, 383–388. [Google Scholar] [CrossRef] [PubMed]
  152. Ito, T.; Nakashimada, Y.; Senba, K.; Matsui, T.; Nishio, N. Hydrogen and Ethanol Production from Glycerol-Containing Wastes Discharged after Biodiesel Manufacturing Process. J. Biosci. Bioeng. 2005, 100, 260–265. [Google Scholar] [CrossRef] [PubMed]
  153. Trchounian, K.; Abrahamyan, V.; Poladyan, A.; Trchounian, A. Escherichia Coli Growth and Hydrogen Production in Batch Culture upon Formate Alone and with Glycerol Co-Fermentation at Different PHs. Int. J. Hydrogen Energy 2015, 40, 9935–9941. [Google Scholar] [CrossRef]
  154. Trchounian, K.; Poladyan, A.; Trchounian, A. Enhancement of Escherichia Coli Bacterial Biomass and Hydrogen Production by Some Heavy Metal Ions and Their Mixtures during Glycerol vs. Glucose Fermentation at a Relatively Wide Range of PH. Int. J. Hydrogen Energy 2017, 42, 6590–6597. [Google Scholar] [CrossRef]
  155. Pachapur, V.L.; Sarma, S.J.; Brar, S.K.; Le Bihan, Y.; Buelna, G.; Verma, M. Surfactant Mediated Enhanced Glycerol Uptake and Hydrogen Production from Biodiesel Waste Using Co-Culture of Enterobacter Aerogenes and Clostridium Butyricum. Renew. Energy 2016, 95, 542–551. [Google Scholar] [CrossRef]
  156. De Faber, M.O.; Ferreira-Leitão, V.S. Optimization of Biohydrogen Yield Produced by Bacterial Consortia Using Residual Glycerin from Biodiesel Production. Bioresour. Technol. 2016, 219, 365–370. [Google Scholar] [CrossRef]
  157. Sharma, Y.; Parnas, R.; Li, B. Bioenergy Production from Glycerol in Hydrogen Producing Bioreactors (HPBs) and Microbial Fuel Cells (MFCs). Int. J. Hydrogen Energy 2011, 36, 3853–3861. [Google Scholar] [CrossRef]
  158. Chookaew, T.; Prasertsan, P.; Ren, Z.J. Two-Stage Conversion of Crude Glycerol to Energy Using Dark Fermentation Linked with Microbial Fuel Cell or Microbial Electrolysis Cell. New Biotechnol. 2014, 31, 179–184. [Google Scholar] [CrossRef]
  159. Khalil, M.E.; Jain, A.; Das, E.; Yang, K.L.; Rajagopalan, G. A Robust and Efficient Bioprocess of Hydrogen Production from Crude Glycerol by Clostridium Beijerinckii G117. Int. J. Hydrogen Energy 2023, 48, 7604–7620. [Google Scholar] [CrossRef]
  160. Sinharoy, A.; Kumar, M.; Pakshirajan, K. An Overview of Bioreactor Configurations and Operational Strategies for Dark Fermentative Biohydrogen Production. In Bioreactors: Sustainable Design and Industrial Applications in Mitigation of GHG Emissions; Academic Press: Cambridge, MA, USA, 2020; pp. 249–288. [Google Scholar] [CrossRef]
  161. Toledo-Alarcón, J.; Cabrol, L.; Jeison, D.; Trably, E.; Steyer, J.P.; Tapia-Venegas, E. Impact of the Microbial Inoculum Source on Pre-Treatment Efficiency for Fermentative H2 Production from Glycerol. Int. J. Hydrogen Energy 2020, 45, 1597–1607. [Google Scholar] [CrossRef]
  162. Sivaramakrishnan, R.; Shanmugam, S.; Sekar, M.; Mathimani, T.; Incharoensakdi, A.; Kim, S.H.; Parthiban, A.; Edwin Geo, V.; Brindhadevi, K.; Pugazhendhi, A. Insights on Biological Hydrogen Production Routes and Potential Microorganisms for High Hydrogen Yield. Fuel 2021, 291, 120136. [Google Scholar] [CrossRef]
  163. Magrini, F.E.; de Almeida, G.M.; da Maia Soares, D.; Fuentes, L.; Ecthebehere, C.; Beal, L.L.; da Silveira, M.M.; Paesi, S. Effect of Different Heat Treatments of Inoculum on the Production of Hydrogen and Volatile Fatty Acids by Dark Fermentation of Sugarcane Vinasse. Biomass Convers. Biorefinery 2021, 11, 2443–2456. [Google Scholar] [CrossRef]
  164. Hernández, C.; Alamilla-Ortiz, Z.L.; Escalante, A.E.; Navarro-Díaz, M.; Carrillo-Reyes, J.; Moreno-Andrade, I.; Valdez-Vazquez, I. Heat-Shock Treatment Applied to Inocula for H2 Production Decreases Microbial Diversities, Interspecific Interactions and Performance Using Cellulose as Substrate. Int. J. Hydrogen Energy 2019, 44, 13126–13134. [Google Scholar] [CrossRef]
  165. Wang, J.; Yin, Y. Progress in Microbiology for Fermentative Hydrogen Production from Organic Wastes. Crit. Rev. Environ. Sci. Technol. 2019, 49, 825–865. [Google Scholar] [CrossRef]
  166. Mıynat, M.E.; Argun, H. Prevention of Substrate and Product Inhibitions by Using a Dilution Strategy during Dark Fermentative Hydrogen Production from Molasses. Int. J. Hydrogen Energy 2020, 45, 34695–34706. [Google Scholar] [CrossRef]
  167. Chen, Y.; Yin, Y.; Wang, J. Influence of Butyrate on Fermentative Hydrogen Production and Microbial Community Analysis. Int. J. Hydrogen Energy 2021, 46, 26825–26833. [Google Scholar] [CrossRef]
  168. Sun, Y.; Ma, Y.; Zhang, B.; Sun, H.; Wang, N.; Wang, L.; Zhang, J.; Xue, R. Comparison of Magnetite/Reduced Graphene Oxide Nanocomposites and Magnetite Nanoparticles on Enhancing Hydrogen Production in Dark Fermentation. Int. J. Hydrogen Energy 2022, 47, 22359–22370. [Google Scholar] [CrossRef]
  169. Niño-Navarro, C.; Chairez, I.; Christen, P.; Canul-Chan, M.; García-Peña, E.I. Enhanced Hydrogen Production by a Sequential Dark and Photo Fermentation Process: Effects of Initial Feedstock Composition, Dilution and Microbial Population. Renew. Energy 2020, 147, 924–936. [Google Scholar] [CrossRef]
  170. 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 Strateg. Rev. 2019, 24, 27–37. [Google Scholar] [CrossRef]
  171. Manish, S.; Banerjee, R. Comparison of Biohydrogen Production Processes. Int. J. Hydrogen Energy 2008, 33, 279–286. [Google Scholar] [CrossRef]
  172. Szewczyk, K.W. Biological Production of Hydrogen. Prog. Microbiol. 2008, 47, 241–247. [Google Scholar]
  173. Chookaew, T.; O-Thong, S.; Prasertsan, P. Biohydrogen Production from Crude Glycerol by Immobilized Klebsiella Sp. TR17 in a UASB Reactor and Bacterial Quantification under Non-Sterile Conditions. Int. J. Hydrogen Energy 2014, 39, 9580–9587. [Google Scholar] [CrossRef]
  174. Chookaew, T.; O-Thong, S.; Prasertsan, P. Biohydrogen Production from Crude Glycerol by Two Stage of Dark and Photo Fermentation. Int. J. Hydrogen Energy 2015, 40, 7433–7438. [Google Scholar] [CrossRef]
  175. Rodrigues, C.V.; Rios Alcaraz, F.A.; Nespeca, M.G.; Rodrigues, A.V.; Motteran, F.; Tallarico Adorno, M.A.; Varesche, M.B.A.; Maintinguer, S.I. Biohydrogen Production in an Integrated Biosystem Using Crude Glycerol from Waste Cooking Oils. Renew. Energy 2020, 162, 701–711. [Google Scholar] [CrossRef]
  176. Sarma, S.J.; Brar, S.K.; Le Bihan, Y.; Buelna, G. Bio-Hydrogen Production by Biodiesel-Derived Crude Glycerol Bioconversion: A Techno-Economic Evaluation. Bioprocess Biosyst. Eng. 2013, 36, 1–10. [Google Scholar] [CrossRef]
  177. Ghosh, D.; Tourigny, A.; Hallenbeck, P.C. Near Stoichiometric Reforming of Biodiesel Derived Crude Glycerol to Hydrogen by Photofermentation. Int. J. Hydrogen Energy 2012, 37, 2273–2277. [Google Scholar] [CrossRef]
  178. Ngo, T.A.; Sim, S.J. Dark Fermentation of Hydrogen from Waste Glycerol Using Hyperthermophilic Eubacterium Thermotoga Neapolitana. Environ. Prog. Sustain. Energy 2012, 31, 466–473. [Google Scholar] [CrossRef]
  179. Varrone, C.; Giussani, B.; Izzo, G.; Massini, G.; Marone, A.; Signorini, A.; Wang, A. Statistical Optimization of Biohydrogen and Ethanol Production from Crude Glycerol by Microbial Mixed Culture. Int. J. Hydrogen Energy 2012, 37, 16479–16488. [Google Scholar] [CrossRef]
  180. Varrone, C.; Rosa, S.; Fiocchetti, F.; Giussani, B.; Izzo, G.; Massini, G.; Marone, A.; Signorini, A.; Wang, A. Enrichment of Activated Sludge for Enhanced Hydrogen Production from Crude Glycerol. Int. J. Hydrogen Energy 2013, 38, 1319–1331. [Google Scholar] [CrossRef]
  181. Mangayil, R.; Aho, T.; Karp, M.; Santala, V. Improved Bioconversion of Crude Glycerol to Hydrogen by Statistical Optimization of Media Components. Renew. Energy 2015, 75, 583–589. [Google Scholar] [CrossRef]
  182. Dounavis, A.S.; Ntaikou, I.; Lyberatos, G. Production of Biohydrogen from Crude Glycerol in an Upflow Column Bioreactor. Bioresour. Technol. 2015, 198, 701–708. [Google Scholar] [CrossRef] [PubMed]
  183. Liu, B.; Christiansen, K.; Parnas, R.; Xu, Z.; Li, B. Optimizing the Production of Hydrogen and 1,3-Propanediol in Anaerobic Fermentation of Biodiesel Glycerol. Int. J. Hydrogen Energy 2013, 38, 3196–3205. [Google Scholar] [CrossRef]
  184. Sarma, S.; Ortega, D.; Minton, N.P.; Dubey, V.K.; Moholkar, V.S. Homologous Overexpression of Hydrogenase and Glycerol Dehydrogenase in Clostridium Pasteurianum to Enhance Hydrogen Production from Crude Glycerol. Bioresour. Technol. 2019, 284, 168–177. [Google Scholar] [CrossRef] [PubMed]
  185. De Paranhos, A.G.O.; Silva, E.L. Statistical Optimization of H2, 1,3-Propanediol and Propionic Acid Production from Crude Glycerol Using an Anaerobic Fluidized Bed Reactor: Interaction Effects of Substrate Concentration and Hydraulic Retention Time. Biomass Bioenergy 2020, 138, 105575. [Google Scholar] [CrossRef]
  186. De Almeida Silva, M.C.; Monteggia, L.O.; Alves Barroso Júnior, J.C.; Granada, C.E.; Giongo, A. Evaluation of Semi-Continuous Operation to Hydrogen and Volatile Fatty Acids Production Using Raw Glycerol as Substrate. Renew. Energy 2020, 153, 701–710. [Google Scholar] [CrossRef]
  187. Kazimierowicz, J.; Dębowski, M.; Zieliński, M. Biohythane Production in Hydrogen-Oriented Dark Fermentation of Aerobic Granular Sludge (AGS) Pretreated with Solidified Carbon Dioxide (SCO2). Int. J. Mol. Sci. 2023, 24, 4442. [Google Scholar] [CrossRef]
  188. Roy, S.; Das, D. Biohythane Production from Organic Wastes: Present State of Art. Environ. Sci. Pollut. Res. 2016, 23, 9391–9410. [Google Scholar] [CrossRef]
  189. Liu, X.; Li, R.; Ji, M.; Han, L. Hydrogen and Methane Production by Co-Digestion of Waste Activated Sludge and Food Waste in the Two-Stage Fermentation Process: Substrate Conversion and Energy Yield. Bioresour. Technol. 2013, 146, 317–323. [Google Scholar] [CrossRef]
  190. Fu, S.F.; Xu, X.H.; Dai, M.; Yuan, X.Z.; Guo, R.B. Hydrogen and Methane Production from Vinasse Using Two-Stage Anaerobic Digestion. Process Saf. Environ. Prot. 2017, 107, 81–86. [Google Scholar] [CrossRef]
  191. Nkemka, V.N.; Gilroyed, B.; Yanke, J.; Gruninger, R.; Vedres, D.; McAllister, T.; Hao, X. Bioaugmentation with an Anaerobic Fungus in a Two-Stage Process for Biohydrogen and Biogas Production Using Corn Silage and Cattail. Bioresour. Technol. 2015, 185, 79–88. [Google Scholar] [CrossRef] [PubMed]
  192. Cheng, J.; Liu, Y.; Lin, R.; Xia, A.; Zhou, J.; Cen, K. Cogeneration of Hydrogen and Methane from the Pretreated Biomass of Algae Bloom in Taihu Lake. Int. J. Hydrogen Energy 2014, 39, 18793–18802. [Google Scholar] [CrossRef]
  193. Wang, X.; Zhao, Y.-c. A Bench Scale Study of Fermentative Hydrogen and Methane Production from Food Waste in Integrated Two-Stage Process. Int. J. Hydrogen Energy 2009, 34, 245–254. [Google Scholar] [CrossRef]
  194. Lateef, S.A.; Beneragama, N.; Yamashiro, T.; Iwasaki, M.; Umetsu, K. Batch Anaerobic Co-Digestion of Cow Manure and Waste Milk in Two-Stage Process for Hydrogen and Methane Productions. Bioprocess Biosyst. Eng. 2014, 37, 355–363. [Google Scholar] [CrossRef] [PubMed]
  195. Chu, C.F.; Li, Y.Y.; Xu, K.Q.; Ebie, Y.; Inamori, Y.; Kong, H.N. A PH- and Temperature-Phased Two-Stage Process for Hydrogen and Methane Production from Food Waste. Int. J. Hydrogen Energy 2008, 33, 4739–4746. [Google Scholar] [CrossRef]
  196. Luo, G.; Xie, L.; Zou, Z.; Wang, W.; Zhou, Q.; Shim, H. Anaerobic Treatment of Cassava Stillage for Hydrogen and Methane Production in Continuously Stirred Tank Reactor (CSTR) under High Organic Loading Rate (OLR). Int. J. Hydrogen Energy 2010, 35, 11733–11737. [Google Scholar] [CrossRef]
  197. Buitrón, G.; Kumar, G.; Martinez-Arce, A.; Moreno, G. Hydrogen and Methane Production via a Two-Stage Processes (H2-SBR + CH4-UASB) Using Tequila Vinasses. Int. J. Hydrogen Energy 2014, 39, 19249–19255. [Google Scholar] [CrossRef]
  198. Kongjan, P.; Jariyaboon, R.; O-Thong, S. Anaerobic Digestion of Skim Latex Serum (SLS) for Hydrogen and Methane Production Using a Two-Stage Process in a Series of up-Flow Anaerobic Sludge Blanket (UASB) Reactor. Int. J. Hydrogen Energy 2014, 39, 19343–19348. [Google Scholar] [CrossRef]
  199. Rivero, M.; Solera, R.; Perez, M. Anaerobic Mesophilic Co-Digestion of Sewage Sludge with Glycerol: Enhanced Biohydrogen Production. Int. J. Hydrogen Energy 2014, 39, 2481–2488. [Google Scholar] [CrossRef]
  200. Jehlee, A.; Rodjaroen, S.; Waewsak, J.; Reungsang, A.; O-Thong, S. Improvement of Biohythane Production from Chlorella Sp. TISTR 8411 Biomass by Co-Digestion with Organic Wastes in a Two-Stage Fermentation. Int. J. Hydrogen Energy 2019, 44, 17238–17247. [Google Scholar] [CrossRef]
  201. Dounavis, A.; Νtaikou, Ι.; Lyberatos, G. Production of Advanced Biobased Hydrogen Enriched Methane from Waste Glycerol in a Two-Stage Continuous System. Waste Biomass Valorization 2016, 7, 677–689. [Google Scholar] [CrossRef]
  202. Sittijunda, S.; Sitthikitpanya, N.; Plangklang, P.; Reungsang, A. Two-Stage Anaerobic Codigestion of Crude Glycerol and Micro-Algal Biomass for Biohydrogen and Methane Production by Anaerobic Sludge Consortium. Fermentation 2021, 7, 175. [Google Scholar] [CrossRef]
  203. Sarkar, O.; Santhosh, J.; Dhar, A.; Venkata Mohan, S. Green Hythane Production from Food Waste: Integration of Dark-Fermentation and Methanogenic Process towards Biogas up-Gradation. Int. J. Hydrogen Energy 2021, 46, 18832–18843. [Google Scholar] [CrossRef]
  204. Eryildiz, B.; Lukitawesa; Taherzadeh, M.J. Effect of PH, Substrate Loading, Oxygen, and Methanogens Inhibitors on Volatile Fatty Acid (VFA) Production from Citrus Waste by Anaerobic Digestion. Bioresour. Technol. 2020, 302, 122800. [Google Scholar] [CrossRef]
  205. Kanchanasuta, S.; Sillaparassamee, O. Enhancement of Hydrogen and Methane Production from Co-Digestion of Palm Oil Decanter Cake and Crude Glycerol Using Two Stage Thermophilic and Mesophilic Fermentation. Int. J. Hydrogen Energy 2017, 42, 3440–3446. [Google Scholar] [CrossRef]
Figure 1. Trends in global biodiesel and bioethanol production in 2010–2021 with a forecast for 2022–2031, own elaboration based on data from [5].
Figure 1. Trends in global biodiesel and bioethanol production in 2010–2021 with a forecast for 2022–2031, own elaboration based on data from [5].
Energies 17 00338 g001
Figure 2. Report from searches on (a) Google Scholar, (b) Scopus, (c) Scilit, and (d) Science Direct for the keywords “biotechnological glycerol valorization”, “methane fermentation of glycerol”, “hydrogen fermentation of glycerol”, and “biohythane production from glycerol” between 2010 and 2022. Retrieved on 19 November 2023.
Figure 2. Report from searches on (a) Google Scholar, (b) Scopus, (c) Scilit, and (d) Science Direct for the keywords “biotechnological glycerol valorization”, “methane fermentation of glycerol”, “hydrogen fermentation of glycerol”, and “biohythane production from glycerol” between 2010 and 2022. Retrieved on 19 November 2023.
Energies 17 00338 g002
Figure 3. The triglyceride transesterification reaction.
Figure 3. The triglyceride transesterification reaction.
Energies 17 00338 g003
Figure 4. Primary uses of crude glycerol, own elaboration based on data from [37,38,43].
Figure 4. Primary uses of crude glycerol, own elaboration based on data from [37,38,43].
Energies 17 00338 g004
Figure 5. Process flowchart for the AD of waste glycerol, own elaboration based on data from [58,59]. (1) organics are hydrolyzed, (2) monomers are digested, (3) propionic and butyric acids are oxidized by obligate H2 producers (acetogenic bacteria), (4) hydrogen carbonates are oxidized by acetogenic bacteria, (5) propionic and butyric acids are oxidized by sulfate-reducing and nitrogen-reducing bacteria, (6) acetic acid is oxidized by sulfate-reducing and nitrogen-reducing bacteria, (7) H2S is oxidized by nitrogen-reducing bacteria, (8) acetic acid is decarboxylated, and (9) CO2 is reduced to CH4.
Figure 5. Process flowchart for the AD of waste glycerol, own elaboration based on data from [58,59]. (1) organics are hydrolyzed, (2) monomers are digested, (3) propionic and butyric acids are oxidized by obligate H2 producers (acetogenic bacteria), (4) hydrogen carbonates are oxidized by acetogenic bacteria, (5) propionic and butyric acids are oxidized by sulfate-reducing and nitrogen-reducing bacteria, (6) acetic acid is oxidized by sulfate-reducing and nitrogen-reducing bacteria, (7) H2S is oxidized by nitrogen-reducing bacteria, (8) acetic acid is decarboxylated, and (9) CO2 is reduced to CH4.
Energies 17 00338 g005
Figure 6. Flowchart of photofermentative H2 production by microbes, own elaboration based on data from [122].
Figure 6. Flowchart of photofermentative H2 production by microbes, own elaboration based on data from [122].
Energies 17 00338 g006
Figure 7. Flowchart of glycerol-to-H2 production by Clostridium spp., own elaboration based on data from [143].
Figure 7. Flowchart of glycerol-to-H2 production by Clostridium spp., own elaboration based on data from [143].
Energies 17 00338 g007
Table 1. Comparison of the literature data on glycerol used in AD.
Table 1. Comparison of the literature data on glycerol used in AD.
Base SubstrateExperimental ParametersGlycerol VolumeMethane/Biogas Production/YieldRef.
70% primary sludge and 30% activated sludge5.5 L CSTR reactor, 56 °C, 25 rpm, pH 6.8 ± 0.1, HRT 20 ± 1 days, OLR 2.5 g COD/L/d1.5% v/v0.25 ± 0.04 L CH4/L/d[76]
5.5 L CSTR reactor, 56 ± 1 °C, 25 rpm, pH 7.1 ± 0.2, HRT 22 ± 2 days, OLR 2.4 ± 0.1 g COD/L/d1.6% v/v0.23 ± 0.01 L CH4/L/d
5.5 L CSTR reactor, 56 ± 6 °C, 25 rpm, pH 6.9 ± 0.1, HRT 18 ± 3 days, OLR 3.6 ± 0.6 g COD/L/d2% v/v0.53 ± 0.14 L CH4/L/d
5.5 L CSTR reactor, 55 °C, 25 rpm, pH 6.4 ± 0.2, HRT 19 ± 1 days, OLR 3.4 ± 0.5 g COD/L/d3% v/v0.33 ± 0.15 L CH4/L/d
5.5 L CSTR reactor, 36 ± 1 °C, 25 rpm, pH 7.6 ± 0.2, HRT 20 ± 2 days, OLR 3.0 ± 0.3 g COD/L/d1.2% v/v0.62 ± 0.07 L CH4/L/d
5.5 L CSTR reactor, 36 ± 1 °C, 25 rpm, pH 7.2 ± 0.2, HRT 20 ± 1 days, OLR 3.2 ± 0.4 g COD/L/d2.4% v/v0.66 ± 0.09 L CH4/L/d
Waste activated sludge (WAS)60 L CSTR, 37 °C, pH 6.8–7.2, HRT 14 days2% v/v100 ± 8 L biogas/d[74]
60 L CSTR, 37 °C, pH 6.8–7.2, HRT 16.4 days3% v/v90 ± 2.8 L biogas/d
60 L CSTR, 37 °C, pH 6.8–7.2, HRT 19.7 days4% v/v80 ± 2.6 L biogas/d
Primary sewage sludge50 L CSTR, 35 ± 0.1 °C, HRT 20 days0.63% v/v1.3 m3 CH4/L crude glycerol[75]
Primary sewage sludge37 °C, pH 7.24, 20 days1% v/v223.8 mL CH4/g VS[78]
3% v/v368.8 mL CH4/g VS
Primary sewage sludge and food waste37 °C, pH 7.48–7.13, 20 days1% v/v432.4 mL biogas/g VS;
343.3 mL CH4/g VS
[79]
3% v/v692.6 mL biogas/g VS;
525.7 mL CH4/g VS
Wastewater from the biodiesel production process1 L Pyrex reactors, 37 °C, pH 7.3815% v/v310 mL CH4/g COD removed[80]
Olive mill wastewater and slaughterhouse wastewater4 L CSTR with 3 L working volume, 35 °C, pH 7–7.5, 46 days, HRT 23–25 d1% v/v1210 ± 205 mL CH4/d[81]
Organic fraction of municipal solid waste (40% fruit, 25% potatoes, 25% vegetables, 8% bread, and 2% paper)4 L CSTR with 3 L working volume, 35 °C, pH 7–7.5, 46 days, HRT 23–25 d1% v/v2094 ± 92 mL CH4/d
Landfill leachate 1 L batch reactor, 0.8 L reaction volume, 37 °C, pH 7, 33.2 days1.71%403.15 mL biogas/g VSS[101]
Sargassum sp.Work volume of 50 mL, 37 °C, pH 7.24, 42 days3 g/L283 ± 18 mL CH4/d[100]
Cattle manure and noodle factorysludge50 m3 pilot plant (30-m3 working volume), 35 °C1 mL glycerol/L/d, (47% purity) 358 mL CH4/g COD removed;
149.1 mL CH4/g COD
[73]
Dairy manure4.5 L reactor, 37 °C, 900 days, HRT 25 days, OLR 2.93–3.90 g VS/L/dDairy manure– glycerol proportion:
91:9; 75:25; 62:38; 51:49
1.4 L CH4/L/d
CH4 yield: 549 L/kg VS
[87]
Pig manure5.5 L semi-CSTR, 55 °C, 60 rpm3% v/v0.47 L biogas/g VS (180% higher)[82]
Swine manure25 L reactor, 34 °C, 80 days2–8% v/v1.4 L CH4/L/d
CH4 yield: 380 L/kg VS
[92]
Swine effluent 5.5 L semi-continuous stirred tank reactors with a working volume of 4 L, 35 °C, pH 6.5, 60 rpm, 20 days, OLR 1.9 ± 0.1 g VS/L/d4% v/v0.74 ± 0.03 L biogas/g VS[83]
5.5 L semi-continuous stirred tank reactors with a working volume of 4 L, 35 °C, pH 6.5, 60 rpm, 20 days, OLR 1.7 ± 0.1 g VS/L/d0.78 ± 0.02 L biogas/g VS
Table 2. Use of crude glycerol for fermentative hydrogen production.
Table 2. Use of crude glycerol for fermentative hydrogen production.
MethodBase SubstrateExperimental ParametersGlycerol VolumeProduction/YieldRef.
PhotofermentationR. palustris125 mL serum bottles, 30 °C, 200 W/m210 mM6.1 moles H2/mole crude glycerol (87%)[177]
R. palustris125 mL serum bottles, 30 °C, 50 W halogen light 10 mM6 moles H2/mole glycerol[128]
R. palustris125 mL serum bottles, 30 °C, 200 W/m230 mM6.69 moles H2/mole crude glycerol[129]
Arundo donax200 mL reactor, 30 ± 1 °C, pH 7.0, incubator with 2000 Lux illumination for 72 h10 g/L32.14 mL/g substrate[131]
15 g/L79.15 mL/g substrate
20 g/L51.06 mL/g substrate
30 g/L12.43 mL/g substrate
R. palustrisFluidized bed PBR (FB), immobilized,
28 ± 1 °C, pH 7, 100 W/m2
50 mM14.7 mL/g/h[138]
Packed bed PBR (PB), immobilized,
28 ± 1 °C, pH 7, 100 W/m2
4.53 mL/g/h
Column PBR, planktonic, 28 ± 1 °C,
pH 7, 100 W/m2
12.68 mL/g/h
R. palustris500 mL bioreactors, planktonic,
35 ± 0.2 °C, pH 7, 250 ± 20 W/m2
50 mM5.1 mL/g/h[139]
500 mL bioreactors, immobilized,
35 ± 0.2 °C, pH 7, 250 ± 20 W/m2
8 mL/g/h
Dark fermentationEnterobacter aerogenes and C. butyricum125 mL serum bottles, 36 °C, pH 6.5, 150 rpm17.5 g/L1.8 mmol H2/g glycerol[155]
Industrial and municipal solid waste
(Eubacteria and Archaea)
2 L reactor with a useful volume of 1.7 L, pH 6.61 ± 0.12, 23 rpm1% v/v51 ± 4 LH2/kg VS[144]
Enterobacter aerogenes HU-101Cylindrical glass column reactor with a working volume of 60 mL, 37 °C, pH 6.81.7 g/L1.12 moles H2/mole glycerol[152]
3.3 g/L0.9 moles H2/mole glycerol
10 g/L0.71 moles H2/mole glycerol
25 g/L0.71 moles H2/mole glycerol
Bacillus thuringiensis EGU45 Batch, 37 °C, pH 7, 1% w/v NH4Cl1% v/v0.261 moles H2/mole glycerol[151]
2% v/v0.452 moles H2/mole glycerol
3% v/v0.252 moles H2/mole glycerol
Batch, 37 °C, pH 7, 1% w/v NaNO31% v/v0.748 moles H2/mole glycerol
2% v/v0.570 moles H2/mole glycerol
3% v/v0.314 moles H2/mole glycerol
Batch, 37 °C, pH 7, 1% w/v NH4NO31% v/v0.547 moles H2/mole glycerol
2% v/v0.646 moles H2/mole glycerol
3% v/v0.299 moles H2/mole glycerol
Aspirator bottles, 1.2 L capacity with working a volume of 1.0 L, 37 °C, pH 7, 1% (w/v) NH4Cl,
support material: free-floating (FF), HRT 2 days
2% v/v0.188 moles H2/mole glycerol
5% v/v0.046 moles H2/mole glycerol
Aspirator bottles, 1.2 L capacity with working a volume of 1.0 L, 37 °C, pH 7, 1% (w/v) NH4Cl,
support material: banana leaves (BL), HRT 2 days
2% v/v0.273 moles H2/mole glycerol
5% v/v0.113 moles H2/mole glycerol
Aspirator bottles, 1.2 L capacity with working a volume of 1.0 L, 37 °C, pH 7, 1% (w/v) NH4Cl,
support material: coconut coir (CC), HRT 2 days
2% v/v0.237 moles H2/mole glycerol
5% v/v0.116 moles H2/mole glycerol
Aspirator bottles, 1.2 L capacity with working a volume of 1.0 L, 37 °C, pH 7, 1% (w/v) NH4Cl,
support material: free-floating (FF), HRT 4 days
2% v/v0.283 moles H2/mole glycerol
5% v/v0.146 moles H2/mole glycerol
Aspirator bottles, 1.2 L capacity with working a volume of 1.0 L, 37 °C, pH 7, 1% (w/v) NH4Cl,
support material: banana leaves (BL), HRT 4 days
2% v/v0.410 moles H2/mole glycerol
5% v/v0.366 moles H2/mole glycerol
Aspirator bottles, 1.2 L capacity with working a volume of 1.0 L, 37 °C, pH 7, 1% (w/v) NH4Cl,
support material: coconut coir (CC), HRT 4 days
2% v/v0.288 moles H2/mole glycerol
5% v/v0.286 moles H2/mole glycerol
C. pasteurianum CH4Batch, 35 °C, pH 7, 200 rpm 10 g/L of pure glycerol1.11 ± 0.16 moles H2/mole glycerol, 256.8 ± 8.1 mL/h/L[150]
CSTR, HRT 12 h, 35 °C10 g/L of pure glycerol0.50 ± 0.02 moles H2/mole glycerol, 103.1 ± 8.1 mL/h/L
10 g/L of crude glycerol0.77 ± 0.05 moles H2/mole glycerol,
166.0 ± 8.7 mL/h/L
Klebsiella sp. TR17Immobilized, 40 °C, up-flow
anaerobic sludge blanket reactor, 4 h
50% glycerol29.00–44.27 mmol H2/g glycerol[173]
T. neapolitana75 °C, pH 7.0–7.5,
itaconic acid
--[178]
bd125 mL serum bottles, 37 °C, pH 7.915 g/L0.96 moles H2/mole glycerol
(2.2 L/L/d)
7.92 g ethanol/L
[179]
Enterobacteriaceae (90%), with the genera Klebsiella and Escherichia/Shigella, followed by Burkholderiaceae (10%) with the genus Cupriavidus3 L bioreactor, 37 °C,
pH 6.8, 120 rpm
15 g/L0.9 moles H2/mole glycerol
(2960 mL H2/L/d)
[180]
bd120 mL serum bottles, 40 °C, 150 rpm1 g/L1.41 moles H2/mole glycerol consumed[181]
bdUp-flow column reactor, 35 °C, HRT 24 h, pH 7,
OLR 29.7 ± 0.5 g COD/d/L
4.2 ± 1.2 g/L107.3 L H2/kg waste glycerol[182]
bd160 mL serum bottles, 30 °C, pH 5.5, 350 rpm,
GCI/IGCT 2.5 h/20 h
7 g/L0.75 moles H2/mole glycerol [183]
bd120 mL serum bottles,
35 °C, 48 h
10 g/L0.55 moles H2/mole glycerol (332.04 mL H2/L)[158]
C. pasteurianumBatch, 37 °C, pH 760 g/L0.93 moles H2/mole glycerol[184]
Granular sludge from a mesophilic UASB reactor used for the treatment of poultry slaughterhouse wastewaterAnaerobic fluidized bed reactor, 30 °C, pH 4.518 g/L0.17 moles H2/mole glycerol[185]
Granular sludge from a UASB reactor from an effluent treatment plant of a soybean processing companySemi-continuous, 35 °C, pH 5.550 g/L0.01 moles H2/mole glycerol[186]
Dark fermentation and photofermentationDark fermentation: Klebsiella sp.,
Photofermentation: Rhodopseudomonas palustris TN1
Dark fermentation: 60 mL serum bottle with a 36 mL working volume, 40 °C, pH 8;
Photofermentation: 60 mL of serum bottle with a 36 mL working volume, 30 °C, pH 7, anaerobic-light (3000 lux) condition
11.14 g/L glycerol (50% purity)6.42 mmol H2/g COD consumed (80.21% glycerol conversion rate)[174]
R. palustris30 °C, 175 W/m285% glycerol,
<0.5% methanol
6.69 moles H2/mole glycerol[176]
Dark fermentation: Clostridiales, Photofermentation: Rhizobiales and ClostridialesDark fermentation feed: peptone 5 g/L, meat extract 5 g/L, and yeast extract 5 g/L, pH 5.5, 37 °C.
Photofermentation: 37 °C, pH 7.0, and 18.50 W/m2
Dark fermentation: 20 g/L COD from crude glycerol; photofermentation: 1 g/L COD from DFE28 mmol H2/g COD consumed[175]
Dark fermentation: 20 g/L COD from crude glycerol; Photofermentation: 2 g/L COD from DFE25.51 mmol H2/g COD consumed
Dark fermentation: 20 g/L COD from crude glycerol; photofermentation: 3 g/L COD from DFE24.91 mmol H2/g COD consumed
Table 3. Use of glycerol in biohythane production.
Table 3. Use of glycerol in biohythane production.
Base SubstrateExperimental Parameters (Step 1)Glycerol VolumeH2 Production/YieldExperimental Parameters (Step 2)CH4 Production/YieldRef.
Chlorella sp. TISTR 8411250 mL serum bottles with a working volume of 200 mL, 55 °C, pH 5.5 2%39.8 mL/g VS250 mL serum bottles with a working volume of 200 mL, 55 °C577.33 mL/g VS[200]
4%29.44 mL/g VS214.98 mL/g VS
6%20.93 mL/g VS130.67 mL/g VS
8%8.86 mL/g VS78.11 mL/g VS
11%4.28 mL/g VS54.32 mL/g VS
15%2.4 mL/g VS41.02 mL/g VS
GlycerolPVC up-flow, packed bed column bioreactor,
35 ± 0.5 °C, 600 m2/L;
pH 620 g/L, purity 92.2 ± 0.3%0.051 L/g CODContinuous Stirred Tank Reactor (CSTR),
35 °C, HRT 20 day
Effluents from the hydrogenogenic stage (1st stage) diluted with water to 5 g COD/L0.198 L/g COD[201]
pH 6.50.070 L/g CODEffluents from the hydrogenogenic stage (1st stage) diluted with water to 7.5 g COD/L0.242 L/g COD
pH 70.094 L/g CODEffluents from the hydrogenogenic stage (1st stage) diluted with water to 10 g COD/L0.273 L/g COD
Chlorella sp.Serum bottles of 120 mL with 70 mL working volumes,
35 ± 4 °C, pH 6, 150 rpm
Inoculum concentration of 6.17% v/v,
algal biomass concentration of 4.62 g VS/L
10 g/L252.56 ± 2.27 mL H2/LSerum bottles of 120 mL with 70 mL working volumes, pH 7.5, inoculum concentration 25% v/vEffluent from fermentative hydrogen production at 10 g/L of glycerol387.4 ± 20.04 mL CH4/L[202]
Inoculum concentration of 25% v/v,
algal biomass concentration of 23.1 g VS/L
40 g/L140.67 ± 3.41 mL H2/LEffluent from fermentative hydrogen production at 40 g/L glycerol428.0 ± 13.12 mL CH4/L
Inoculum concentration of 10.31% v/v, algal biomass concentration of 23.1 g VS/L13.83 g/L655.12 ± 1.64 m LH2/LEffluent from fermentative hydrogen production at 13.83 g/L glycerol868.7 ± 19.98 mL CH4/L
Food waste, sewage sludge250 mL glass bottles, 35 ± 1 °C, pH 5.5, 150 rpm1% v/v140.2 mL H2/g VS250 mL glass bottles, 35 °C, pH 7, 150 rpm342 mL CH4/g VS[94]
3% v/v177 mL H2/g VS224.4 mL CH4/g VS
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

Kazimierowicz, J.; Dębowski, M.; Zieliński, M.; Kasiński, S.; Cruz Sanchez, J. Biotechnological Valorization of Waste Glycerol into Gaseous Biofuels—A Review. Energies 2024, 17, 338. https://doi.org/10.3390/en17020338

AMA Style

Kazimierowicz J, Dębowski M, Zieliński M, Kasiński S, Cruz Sanchez J. Biotechnological Valorization of Waste Glycerol into Gaseous Biofuels—A Review. Energies. 2024; 17(2):338. https://doi.org/10.3390/en17020338

Chicago/Turabian Style

Kazimierowicz, Joanna, Marcin Dębowski, Marcin Zieliński, Sławomir Kasiński, and Jordi Cruz Sanchez. 2024. "Biotechnological Valorization of Waste Glycerol into Gaseous Biofuels—A Review" Energies 17, no. 2: 338. https://doi.org/10.3390/en17020338

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