Next Article in Journal
Revealing Commercial Epoxy Resins’ Antimicrobial Activity: A Combined Chemical–Physical, Mechanical, and Biological Study
Previous Article in Journal
Monolithic Polyepoxide Membranes for Nanofiltration Applications and Sustainable Membrane Manufacture
Previous Article in Special Issue
Manipulating Microbial Cell Morphology for the Sustainable Production of Biopolymers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polyhydroxyalkanoate Production by Methanotrophs: Recent Updates and Perspectives

1
Department of Chemical Engineering, Konkuk University, 120 Neungdong-ro, Gwangjin-gu, Seoul 05029, Republic of Korea
2
Department of Biotechnology, Hemvati Nandan Bahuguna Garhwal University (A Central University), Srinagar 246174, Uttarakhand, India
3
Department of Microbiology, Pondicherry University, Pondicherry 605014, Kalapet, India
*
Author to whom correspondence should be addressed.
Polymers 2024, 16(18), 2570; https://doi.org/10.3390/polym16182570
Submission received: 9 August 2024 / Revised: 8 September 2024 / Accepted: 9 September 2024 / Published: 11 September 2024

Abstract

:
Methanotrophs are bacteria that consume methane (CH4) as their sole carbon and energy source. These microorganisms play a crucial role in the carbon cycle by metabolizing CH4 (the greenhouse gas), into cellular biomass and carbon dioxide (CO2). Polyhydroxyalkanoates (PHAs) are biopolymers produced by various microorganisms, including methanotrophs. PHA production using methanotrophs is a promising strategy to address growing concerns regarding plastic pollution and the need for sustainable, biodegradable materials. Various factors, including nutrient availability, environmental conditions, and metabolic engineering strategies, influence methanotrophic production. Nutrient limitations, particularly those of nitrogen or phosphorus, enhance PHA production by methanotrophs. Metabolic engineering approaches, such as the overexpression of key enzymes involved in PHA biosynthesis or the disruption of competing pathways, can also enhance PHA yields by methanotrophs. Overall, PHA production by methanotrophs represents a sustainable and versatile approach for developing biomedical materials with numerous potential applications. Additionally, alternative feedstocks, such as industrial waste streams or byproducts can be explored to improve the economic feasibility of PHA production. This review briefly describes the potential of methanotrophs to produce PHAs, with recent updates and perspectives.

1. Introduction

Synthetic plastics are an indispensable component of daily life, and a primary environmental concern is associated with their non-biodegradable nature. As the global energy demand continues to grow, the search for alternative and sustainable sources of energy has become increasingly important, for example, the conversion of biowaste into biogas, such as methane (CH4) by anaerobic digestion (AD) [1,2]. One promising avenue is the bioconversion of CH4, which is the primary component of natural gas, along with carbon dioxide (CO2), into liquid fuels and other valuable biochemicals [3,4,5]. Advances in bioreactor design and metabolic pathways engineering could allow microbes, such as methanotrophs, to convert CH4 into a variety of products, including polyhydroxyalkanoates (PHAs), fatty acids, single-cell proteins, exopolysaccharides (EPS), ectoine, putrescine, and α-humulene [6,7,8]. The natural gas reserve of CH4 is approximately 1.13 × 1014 m3 with high content (>95%), and global CH4 emissions are approximately 770 Tg/y. Methane (CH4) exhibits a global warming potential approximately 25-fold higher than CO2 [9,10,11]. The bioconversion of CH4 into liquid fuels and other valuable products, such as methanotrophs, has several advantages over traditional chemical conversion methods [12,13]. Biological conversion processes can operate at milder temperatures and pressures, potentially lowering the energy consumption and capital costs. Additionally, using CH4 for biofuel production can help mitigate greenhouse gas (GHGs) emissions by capturing and converting potent GHGs into sustainable fuel sources [6,10,14]. However, the large-scale implementation of CH4 bioconversion technology faces several challenges. Achieving high product yields and selectivity and addressing the inherent difficulties in gas–liquid mass transfer are critical hurdles that must be overcome [15,16,17]. Ongoing research and development in areas such as bioreactor design, enzyme engineering, and process optimization is crucial for realizing the full potential of methanotroph-based bioconversion of CH4 [18].
Polyhydroxyalkanoates (PHAs) are polymers with hydroxyalkanoic acid monomer units that can accumulate in a range of microbes, including heterotrophic bacteria and photoautotrophic organisms, such as cyanobacteria [19,20,21]. Methanotrophs are a unique group of microbes that use CH4 as their leading energy and carbon source [1,22]. These microbes flourish under harsh environmental conditions, such as low temperature, high hydrostatic pressure, and inadequate nutrients [23,24]. Methanotrophs utilize diverse substrates, such as CH4, formaldehyde, methanol, and methylamine, to produce numerous value-added biomolecules, including PHAs [3,6]. This property has drawn considerable attention from the scientific community because of its potential applications in bioremediation, biofuel production, and GHG mitigation. Methanotrophs such as Methylosinus, Methylocystis, Methylocapsa, Methylomicrobium Methylococcus, Methylobacterium, and Methylocaldum can accumulate these biopolymers (PHAs) as intracellular carbon (C) and energy storage compounds, amounting to more than 80% of cell dry weight (cdw) when grown under a nutrient-limiting environment, i.e., nitrogen (N) or phosphorus (P) limitation, with an excess of C source [1,6]. The AD process is a well-established technology for the conversion of organic waste into biogas (CH4 and CO2) [2]. The production of PHAs from inexpensive C sources, such as biogas-derived CH4, is vital for reducing costs. Acetogens, autotrophs, and methanotrophs can synthesize PHA using CH4, carbon monoxide, and CO2 [6,25]. Methanotrophs produce vital biomolecules for industrial and biotechnological applications [23,24]. The incorporation of unsaturated fatty acids into a ruminant’s diet can reduce CH4 production, which can indirectly affect the availability of CH4 for PHA production by methanotrophs. The notable challenges in converting CH4 into PHAs by methanotrophs include less efficient PHA-producing methanotrophs, low gas-to-liquid mass transfer efficiency, and inadequate properties of the synthesized PHAs, especially polyhydroxybutyrate (P(3HB)) [6,23,26]. Various strategies can be beneficial for achieving high PHAs using methanotrophs, including engineering microbes, high-molecular-weight (Mw) PHAs, the use of a gas-permeable membrane, two-phase partitioning bioreactors, co-polymers such as poly(3-hydroxybutyrate-co-hydroxyvalerate) (P(3HB-co-3HV)), and fast famine regimes [1,27]. Therefore, CH4 serves as a unique platform for the production of various chemicals using engineered methanotrophs [28,29]. In addition, methanotrophic bacteria are helpful in wastewater treatment systems, with the potential to produce biomolecules such as PHAs, ectoine, and methanobactins for broad biomedical applications, including drug carriers, electronic devices, heart stents, prosthetics, and vaccines [30,31]. The techno-economics of large-scale biorefinery (100,000 t/y) PHA production from CH4 by methanotroph analysis suggested that the production cost of PHAs varied between USD 4.1 and 6.8/kg assuming 25% of uncertainty [7]. Ongoing research efforts have focused on engineering methanotrophs to enhance their CH4 conversion capabilities further and to improve bioreactor design and operational strategies to optimize the overall process efficiency [1,5,6]. Scientists have developed robust and sustainable bioconversion systems that can transform CH4 into valuable chemicals and fuels by harnessing the unique metabolic capabilities of methanotrophs. Recent reviews have focused on the diverse aspects of methanotrophs to produce diverse types of bioproducts, with minor details and updates on PHA production [1,3,5,8,11,13]. In this review, we provide a brief update on the production of PHAs by methanotrophs from a future perspective.

2. Methanotrophs and Their Metabolism

Methanotrophs play a vital role in the global C cycle as they oxidize CH4 and potent GHGs to CO2 [32,33]. The two main types of methanotrophs are type I and type II, which differ in their physiological and metabolic properties. Type I methanotrophs, including Methylomonas, Methylobacter, and Methylomicrobium, are typically found in environments with high CH4 concentrations, such as landfills, rice paddies, and the rumens of ruminant animals [18,34,35]. These microbes are characterized by their potential to rapidly grow and reproduce, making them attractive candidates for biofuel and biopolymer production. In contrast, type II methanotrophs, such as Methylocystis and Methylosinus, are generally found in environments with lower CH4 concentrations and are better adapted to nutrient-limited conditions [1,6]. The metabolism of methanotrophs is characterized by a complex network of enzymatic reactions and pathways that enable them to convert CH4 into energy and essential cellular components efficiently. This process typically begins with CH4 oxidation to methanol, which is further oxidized to formaldehyde [18,36]. Methanotrophs possess specialized enzymes, such as CH4 monooxygenases (MMOs), particulate MMOs (pMMOs), and soluble MMOs (sMMOs), which catalyze the initial oxidation step. The produced formaldehyde is either assimilated into biomass via the ribulose monophosphate (RuMP)/serine pathway or oxidized to CO2 for energy production [6,7]. A key aspect of the metabolism of methanotrophs is the interspecies electron transfer that occurs during the anaerobic decomposition of organic matter. Methanotrophs are majorly classified into a subdivision of α-proteobacteria/γ-proteobacteria and exhibit a range of adaptations that allow them to thrive in diverse environments, including acidic conditions [1,13]. A diverse thermoacidophilic group of methanotrophs that is involved in CO2 fixation belongs to Verrucomicrobia, such as Acidimethylosilex, and Methyloacidiphilium [6,7]. Type II methanotrophs (α-Proteobacteria) produce PHAs via the serine pathway involving 3-ketothiolase (encoded by phaA), acetoacetyl-CoA-reductase (encoded by phaB), and PHA synthetase (encoded by phaC). In contrast, type I methanotrophs do not exhibit PHA accumulation [37,38,39]. The CH4-based PHA production by methanotrophs depends on the culture type (pure methanotrophs or mixed cultures), process parameters, such as CH4 concentration, O2 availability, age of inoculum, pH value, temperature, presence of copolymer precursors (propionate, valerate, or others), and the extent of nutrient limitations [1,8,37]. Alterations in the metabolism of methanotrophs and co-feeding with desirable copolymer precursors can be beneficial for synthesizing various types of PHAs with suitable properties for broad biotechnological applications [6,28,40]. The details of PHA synthesis by methanotrophs are presented in Figure 1.

3. Production of Polyhydroxyalkanoates by Methanotrophs from Methane and Their Physical Characteristics

CH4-based PHA production is a promising approach to mitigate CH4 emissions and replace petroleum-derived synthetic plastics [27,41]. The process parameters strongly influence PHA production by methanotrophs, especially type II, including nutrient limitations, the ratio of CH4 and O2, co-substrates (propionate and valerate), and the incubation temperature and period [31]. CH4-rich biogas is an accessible feedstock for PHAs by methanotrophs. The production of PHAs by Methylocystis-dominated mixed inoculums was evaluated after incubation for 175 d under nonsterile conditions using ammonium (NH4+; N-source) and CH4 [42]. A maximum accumulation of PHAs up to 40% of the cdw by the enriched methanotrophs was noted. Furthermore, supplementation with low (100 mg/L) and high (400 mg/L) valerate resulted in up to 21 and 40% (3-hydroxyvalerate—3HV fraction, mol%) PHA contents of 45 and 30% of the cdw, respectively. The Mw analysis suggested that higher 40 mol% 3HV resulted in lower Mw of 0.93 × 106 than the control (1.20 × 106) and 20 mol% 3HV (1.15 × 106). In contrast, the Mw distribution (polydispersity index, PDI) was higher, up to 2.14, for 40 mol% 3HV compared to the control (1.76). The addition of reducing equivalents, such as formate, exhibited up to a 21% decreased CH4 requirement in the case of the valerate supplement system compared to the control (14%) [42]. Although CH4 is an inexpensive feedstock, the methanotrophic production of PHAs is limited by slow growth and low CH4 solubility. The enhancement of CH4 mass transfer can be improved by increasing the agitation, which requires energy inputs and incurs high operational expenses. The use of water-in-oil emulsions can support the superior growth of methanotrophs, and subsequently, PHA production [43]. Methylocystis parvus OBBPs showed maximum growth rates of 0.009 and 0.052 h−1 under non-shaking and shaking conditions (150 rpm), respectively. Based on Nile red staining, the cells were grown in droplets, and the produced PHAs amounted to 32% of the cdw compared to no PHA production under control conditions. Overall, M. parvus OBBP-led PHA production was 67-fold better P(3HB) in the emulsions than in the control [43]. A type II methanotroph, M. parvus OBBP, grown with CH4 as the sole feed, produced P(3HB) only under nutrient-limiting conditions [44]. Using various co-substrates (100 mg/L), such as 3-hydroxybutyrate, propionate, and valerate, M. parvus OBBP produced maximum PHAs contents of 60, 32, and 54% of the cdw compared to the control, with 50% of the cdw (as CH4 feed) after 48 h of incubation. Among these co-substrates, propionate and valerate exhibited 8 and 22 mol% 3HV in the produced PHAs. Furthermore, increasing the valerate concentration to 2000 mg/L resulted in a higher 40 mol% for 3HV in the accumulated PHAs. Under similar conditions, another type II methanotroph, Methylosinus trichosporium OB3b, using CH4 and valerate (100 mg/L), produced PHA contents (50% of the cdw) with 20 mol% of 3HV. The produced PHAs using CH4 only and with valerate co-substrate as 22 and 37 mol% of 3HV showed melting temperatures of 178, 150, and 134 °C with glass transition temperature (Tg) values of 8 (typical for P(3HB)), −2, and −6 °C (typical for P(3HB-co-3HV)), respectively. Furthermore, the produced P(3HB), P(3HB-co-3HV) with 22 mol% 3HV, and P(3HB-co-3HV) with 37 mol% 3HV exhibited corresponding Mw values of 3.24 × 106, 1.83 × 106, and 1.70 × 106, with PDI values of 1.67, 1.88, and 2.23, respectively [44]. Under N limitations for incubation up to 310 d, methanotrophic communities were enriched using different CH4 concentrations of 20.0, 2.0, and 0.2 g of CH4/m3 in stirred tank reactors (STRs). R1, R2, and R3 were evaluated for PHA production, respectively [45]. Proteobacteria dominated the microbial population, and the CH4-oxidizing culture belonged to type I methanotrophs; Methylobacter, Methylosarcina, Methylosoma, and Methylomicrobium were especially abundant in the R1 and R2 STRs. Their populations gradually declined over the incubation period under fed-batch cultivation. Methylocystis (type II methanotrophs) were present in these STRs, and their abundance gradually increased in the R2 and R3 STRs. Here, the higher abundance of type I than type II methanotrophs may be associated with high copper content. The methanotroph PHA content (w/w) in these STRs varied from 0.3 to 0.5% in R1, from 2.9 to 9.7% in R2, and from 0.1 to 0.8% in R3. The composition of PHAs in P(3HV):P(3HB) was 12:1 in R1 and 4:1 in R3. In contrast, a low P(3HV):P(3HB) ratio of 1:7 was observed for R2. The maximum PHA production of 12.6% was recorded in R2, which was cultivated at a higher CH4/biomass ratio under N limitation conditions. These differences in the PHA contents in the STRs may be associated with diverse methanotrophic communities [45]. In another study, the highest PHA contents of 1.0, 12.6, and 1.0% of the cdw by enriched inoculums of methanotrophs was supported by corresponding feedings of 20, 2.0, and 0.2 g of CH4/m3, respectively [46]. Under P-limiting conditions, Methylocystis sp. GB25 showed production of PHAs of 46.2% with a productivity of 1.13 g/L·h and Mw of 2.41 × 106 [47]. Owing to the limitations of K, S, and Fe, the Methylocystis sp. GB 25 DSM 7674-dominated mixed culture resulted in PHA accumulations of 33.6, 32.6, and 10.4% of the cdw, respectively [48]. Under P deficiency, the synthesized PHA achieved an Mw of 3.1 × 106.
Methylocystis sp. WRRC can produce co-polymers P(3HB-co-3HV) of PHAs by co-feeding CH4 with valerate or pentanol [49]. Using pure CH4, a maximum PHA production of 0.20 g/L was observed with 0.38 g of PHA/g of CH4. The co-feeding of pentanol and valerate exhibited higher PHA production rates of 0.31 and 0.57 g/L, with PHA contents up to 78% of the cdw, respectively. The synthesized P(3HB), P(3HB-co-15%-3HV), and P(3HB-co-43%-3HV) showed melting temperatures of 180, 161, and 170 °C with % crystallinity values of 52, 23, and 5%, respectively. These findings suggest that the PHA yield and polymeric characteristics can be altered by cofeeding valerate and CH4 [49]. García-Pérez et al. [50] evaluated the influence of nutrient-limiting environments, such as N, K, and Mn, with a surplus of Fe on P(3HB) production by Methylocystis hirsuta DSMZ 18500. The accumulation of PHAs was higher at 28% of the cdw under these nutrient-limiting conditions than in the control (7.5% of the cdw). Overall, the maximum CH4 elimination capacity and PHA content and productivity were recorded as 16.2 g/m3·h, 34.6% of the cdw, and 1.4.4 kg/m3·d [50]. Cantera et al. [51] assessed the feasibility of various value-added products using CH4, including PHAs, by mixing cultures in bubble-column bioreactors with different magnesium concentrations (Mg2+; 0.2, 0.02, and 0.002 g/L). High concentrations of Mg2+ were effective in producing ectoine (94.2 mg/g of the cdw) and EPS (2.6 g/L). In contrast, low Mg2+ favored PHA production of 14.3 mg/L. The dominant organisms in the reactors are Halomonas, Marinobacter, Methylophaga, and Methylomicrobium [51]. Similarly, Wendlandt et al. [52] reported PHA production rates of 28.3–51.3% of the cdw by Methylocystys sp. GB25 under Mg2+-limiting conditions. In another study, M. trichosporium OB3b and M. parvus OBBP under N- and Mg2+-limiting conditions exhibited higher PHA accumulation rates of 29–60% of the cdw [53].
The prices of selected biodegradable plastics are approximately USD 2.60–5.80/kg for starch-based bioplastics, USD 2.00–3.45/kg for poly(lactic acid), and USD 2.00–6.50/kg for P(3HB) [54]. The pilot production design and economic evaluation for 100,000 t/a of P(3HB) by methanotrophs and utilizing extraction by acetone/water solvent methods suggested that the estimated PHA expenses varied in ranges from USD 4.1/kg to USD 6.8/kg [54]. The production of PHAs by thermophilic methanotrophs operating at 60 °C can be beneficial over mesophilic methanotrophs (30–45 °C) by reducing costs to USD 3.2–5.4/kg. High-temperature fermentation decreases the mass transfer of CH4. A few thermophilic methanotrophs can accumulate PHAs. Additionally, thermophilic PHA production has several advantages, such as higher substrate solubility, increased diffusion rates, and low contamination risks [54]. Alternatively, biogas feed as a source of CH4 to produce PHAs by M. hirsuta could be helpful in developing anaerobic digestion biorefineries [55]. The artificially designed biogas feed (CH4, 70%; CO2, 29.5%; and H2S, 0.5%) showed a PHA production rate of 45% of the cdw by M. hirsuta DSM 18500. Here, H2S did not negatively influence M. hirsuta DSM 18,500 growth on the additional supplementation of volatile fatty acids, and a maximum PHA content and yield were recorded up to 54% of the cdw and 0.63 g of PHAs/g of substrate. Further, valerate supplementation resulted in 13.5 mol% of 3HV [55]. The enrichment temperature influences PHA production by mixed methanotroph inoculums. The mixed culture enrichment at temperatures of 30 and 37 °C resulted in PHA production rates of up to 35.1% and 34.1% of the cdw, respectively [56]. Here, high PHA production in these methanotroph-mixed inocula was associated with the dominance of Methylocystis at ~30% of the total microbial population. In another study, a Methylosinus-dominated mixed culture from CH4 showed a low PHA production rate of 8.6% of the cdw as P(3HB) [57]. Furthermore, supplementing propionate and valerate as co-substrates with CH4 enhanced it to 22.6 and 65.0 mol%, respectively, with a maximum PHA production rate of 14.1% of the cdw.
The enriched methanotrophic cultures from diverse inocula, Sphagnum peat moss, and Sphagnum with activated sludge, dominated by Proteobacteria (35.7–62.1%) and Bacteroidetes (31.7–39.8%), exhibited PHA production of up to 13.6% of the cdw [58]. In contrast, thermophilic conditions mixed with CH4-utilizing culture showed PHA production of up to 10.0 and 8.0% of the cdw, with 3HV contents of up to 70 and 85 mol% at 55 and 58 °C, respectively [59]. Eam et al. [60] examined bioaugmentation of P(3HB)-producing methanotrophs in activated sludge using M. trichosporium OB3b to improve PHA production. The 1:1 ratio of activated sludge to M. trichosporium OB3b exhibited the maximum PHA production of 37.1% of the cdw. In M. trichosporium OB3b-amended cultures, measurements of the microbial dynamics revealed that the dominant type II methanotrophs Methylocystis and Methylophilus spp. were crucial for the production of PHAs [60]. The co-feeding of CH4 and CO2 positively influenced PHA production by type II methanotrophs and enhanced PHA production by up to 162% [61]. The maximum reported PHAs produced was 38.0% of the cdw by Methylocystis sp. MJC1. Metabolic flux analysis revealed that 45% of the total PHA production occurred from CO2 through the vital roles of phosphoenolpyruvate carboxylase and crotonyl-CoA carboxylase/reductase [61]. Genomic analysis of the thermophilic methanotrophs Methylocaldum spp. metabolism (M. szegediense (O-12 and Norfolk) and M. marinum S8) revealed them as PHA producers [62]. In addition, M. parvus OBBP produced PHAs up to 35% of the cdw from ethane as the sole C source, compared to 48.0% of the cdw using CH4 as a feed [63]. Furthermore, co-feeding with ethane and valerate resulted in 25 mol% of the 3HV contents in the total PHA production of 12.9% of the cdw. Nitrate proved to be a better N source to produce PHAs by up to 27% of the cdw in mixed methanotrophic cultures dominated by Methylocystis sp. than the NH4+ [64]. The PHA-producing mixed cultures enriched from landfill biocover soil, peat bog soil, and waste-activated sludge produced PHA copolymers with diverse ratios of CH4 and O2. These cultures showed a maximum PHA production of up to 41 mol% 3HV after supplementing with valerate. The waste-activated sludge-based culture exhibited the highest PHA yield of 0.42 g/g of substrate at 10% of CH4 feed [64]. Similarly, the 0.24–0.34 g yield of PHB/g of CH4 with a concentration and contents of 0.14–0.28 mg/mL and 30.5–50.3% of the cdw by M. parvus OBBP was reported [65]. In the methanotrophic metabolism for PHA production from CH4, diverse C assimilation pathways and reduced power requirements influence the final storage of PHAs. In addition, formate supplementation of P(3HB)-rich cells delayed PHB consumption. In a previous study, a mixed culture (M. parvus OBBP and Methylosinus sp. LW4, Methylocystis 42/22, and Methylocystis KS30) exhibited PHA production of 25% of the cdw [66]. Potential methanotrophs types I (Methylocaldum O11a, Methylomicrobium album BG8, and Methylomonas LW13) and II (Methylocapsa acidiphila, Methylocystis 42/22, M. hirsuta CSC1, M. parvus OBBP, Methylocystis strain M, Methylocystis rosea SV99, Methylosinus sp. LW4, Methylocystis SC2, M. trichosporium OB3b, and Methylosinus sporium) were evaluated by Pieja et al. [67]. Among these cultures, the type II methanotrophs exhibited PHA production rates of 7.0–36% of the cdw, which was evident with phaC. In addition, the methanotrophic-enriched cultures showed a PHA production rate of up to 46% of the cdw [67].
Various extraction procedures, including solvent extraction (1,3-dioxolane), cell lysis followed by solvent extraction, and cell lysis without solvents, have been evaluated for efficient recovery of intracellular PHAs [68]. Among these methods, the integrative process of cell lysis followed by solvent extraction proved beneficial for achieving 91% PHA recovery with 93% purity. Methylocystis sp. GB 25 synthesized PHAs from CH4 and CO2 with a Mw of 1.08 × 106 and a melting temperature of 175 °C [69]. The biodegradable PHAs produced by the type II methanotroph Methylocystis sp. MJC1 exhibited a high biomass of 21.3 g cdw/L through a high-cell density cultivation approach [70]. The maximum PHAs produced was reported as 8.9 g/L (41.9% of the cdw) with 28 mol% 3HV. Waste gases and C sources, especially C1, such as CH4, CO2, and methanol, originating from anthropogenic and industrial activities, can be directly utilized to produce value-added biomolecules, such as PHAs, at up to 88% of the cdw by methanotrophs [41]. Biomass yields, PHA production, and microbial populations under high-resource (CH4, 20% and NH4+, 10 mM) or low-resource (CH4, 0.2% and NH4+, 0.1 mM) conditions influence PHA production by Methylocystis [39]. PHA production by Methylocystis was observed under high-resource conditions, with a yield of 12.6% cdw. In the case of low resources, the hindrance of PHA production by mixed inocula is associated with a lack of dominant methanotrophs [39]. Simulation analysis of PHA production suggested that M. hirsuta can produce up to 51.6% of the cdw in a forced-liquid vertical loop bioreactor [71]. In another simulation study, methanotrophs could efficiently produce PHAs with a yield of 0.32 kg/m3/d with a low valerate concentration using CH4-abundant gas streams [29]. Process simulation and design, such as superficial gas velocity, aspect ratio, reactor volume, and diameter, positively influenced PHA production by methanotrophs [72]. PHA production using CH4 in bubble column bioreactors showed an accumulation of 37.5% of the cdw by Methylocystis after 12 d. In a 14 L fermenter for PHA productions under nitrate-limiting conditions, the mixed methanotrophic culture dominated by Methylocystis accumulated PHAs up to 22.2% of the cdw. Analysis of the physical characteristics suggested that synthesized PHAs had a Mw 2.2 × 105 g/mol with a PDI of 1.82 [73]. The PHAs produced were up to 3.0 and 7.4% of the cdw by the methanotrophic bacterium strain MTS using CH4 and methanol as feed, respectively [74]. Furthermore, intracellular PHA degradation in type II methanotrophs was confirmed using 13C nuclear magnetic resonance under anoxic conditions. Wendlandt et al. [75] evaluated PHAs produced by Methylocystis sp. GB of 25 DSM7674 under non-sterile conditions. Under the continuous mode and nutrient phosphate limitation, PHA contents of up to 51% of the cdw were produced, with a high Mw of 2.5 × 106. Xin et al. [76] determined the Mw of PHAs produced by M. trichosporium IMV 3011, M. trichosporium OB3b, Methylococcus capsulatus HD6T, and Methylomonas sp. GYJ3 under different conditions, and substrates such as CH4 and methanol were variable in the range of 0.30–1.4 × 106. The Mw of PHAs is highly dependent on the type of microorganism and the critical enzyme activity needed for PHA metabolism. In another study, the co-feeding of CH4 and methanol resulted in PHA accumulation of up to 40% of the cdw, with a high Mw of 1.28–1.48 × 106 by M. trichosporium IMV3011 [77]. High methanol production by PHAs containing M. trichosporium IMV3011 cells was observed [37]. After 168 h of incubation, a maximum PHA production of 41.0% of the cdw was observed in a mineral salt medium containing copper (2 mg/L). A two-phase partitioning bioreactor and silicon oil-based cultivation of Methylobacterium organophilum CZ-2 showed PHA production of 38.0% of the cdw, compared to mixed consortium (34% of the cdw with productivity of 1.61 mg of PHAs/g·h) [78]. Co-feeding with citrate or propionate resulted in PHA production of 5.0–617 mg/L by M. organophilum CZ-2 [79].
Malvar et al. [80] developed a rheological method to identify PHA-producing bacteria. Under N depletion conditions, methanotrophs consortia showed PHA production up to 0.19 g/L under incubation for 120 h. This finding suggests a correlation between bacterial motion and PHA production. A mechanistic model was evaluated to enhance PHA production in M. hirsuta using biogas as a feed under nutrient-limiting conditions [81]. The biomass growth and PHA production on CH4 feed (0.14 and 0.25 g of chemical oxygen demand (COD)/g of COD), a CH4 affinity constant of 5.1 g of COD/m3), and an O2 affinity constant of 4.1 g of O2/m3 showed a PHA production rate of 0.39 g of COD/g of COD/d on CH4. In another study, M. hirsuta CSC1 genome sequence analysis suggested that it could be a potential PHA producer [35]. Methanotrophs enhance biomass growth and survival under variable and nutrient-limiting conditions owing to their metabolic flexibility. The H2 metabolism in Methylacidiphilum sp. RTK17.1 Verrucomicrobial methanotrophs is principally linked to energy management, and its consumption offers a competitive growth advantage within hypoxic habitats [82]. Myung et al. [83] established the synthesis of various PHA co-polymers by M. parvus OBBP with supplementation of the corresponding co-substrates, such as poly(3-hydroxybutyrate-co-4-hydroxybutyrate) (P(3HB-co-4HB)), poly(3-hydroxybutyrate-co-5-hydroxyvalerate-co-3-hydroxyvalerate) (P(3HB-co-5HV-co-3HV)), and poly(3-hydroxybutyrate-co-6-hydroxyhexanoate-co-4-hydroxybutyrate) (P(3HB-co-6HHx-co-4HB)) using butyrate or 4-hydroxybutyrate (4HB), valerate or 5-hydroxyvalerate (5HV), and hexanoate or 6-hydroxyhexanoate (6HHx), respectively. The maximum PHAs produced were up to 59% of the cdw, with corresponding 3HV (25 mol%), 4HB (9.5 mol%), 5HV (3.6 mol%), and 6HHx (1.4 mol%). These co-polymers’ Mw, melting temperature, and Tg varied in ranges of 1.22 × 106–1.33 × 106, 135–150 °C, and −1–−5 °C, respectively [83].
The cost-effective production of PHAs by methanotrophs to manufacture biodegradable plastics was examined. Based on an analysis of CH4 emissions from Australian landfills, the outcomes suggest that economically viable bioplastic production is feasible considering the competitive price of synthetic plastics. The waste of small (5000 t/y), medium (35,000 t/y), and large (230,000 t/y) landfills can recover 162–7480 t of CH4 to produce 71–3252 t of PHAs, with a cost of 1.5–2.0 AUD/kg of PHAs [84]. The PHAs (P(3HB-co-3HV)) were synthesized using CH4 and alkenoates (produced from recycled PHAs via catalytic pyrolytic depolymerization) by M. parvus OBBP with a Mw in the range of 1.18–1.47 × 106 and a purified yield of 33–47 mg of PHAs [85]. The potential of type I and type II methanotrophs to convert biogas containing H2S into PHAs was validated by transcriptomic analysis. Metabolic analysis suggested that the native pathway of H2S utilization involves type II methanotrophs. Under fed-batch conditions, the Methylocystis sp. MJC1 and Methylocystis sp. OK1 showed biomass growth of 4.0, and 4.5 g cdw/L, respectively. Methylocystis sp. showed a maximum PHA production rate of 2.9 g/L using biogas [86]. The pH range of 5.5 to 7.0 highly supports the growth of Methylocystis. A maximum PHA production rate of 43.7% of the cdw was reported at pH 5.5 in mixed methanotrophic cultures dominated by Methylocystis under N-deprived conditions [87]. At pH 8.5, the abundance of Methylocystis declined from 14% to 85–90% at a pH of 5.5–7.0.
Methylocystis sp. MJC1 cultured in 3:7 CH4 and air (v/v) resulted in biomass growth and PHA production of 52.9 and 28.4 g/L at an incubation time of 120 h, respectively [88]. A low PHA production (0.20 g/L·h) was reported because of the long incubation period. Interestingly, increasing the O2-to-CH4 supply to a ratio of 1.5 showed a 1.5-fold increase in biomass productivity and delayed PHA production. At an equal ratio of CH4 and O2, the maximum biomass achieved was 55.9 g/L, with PHA production of 34.5 g/L (61.7%) showing a productivity of 0.36 g/L·h after 164 h of incubation [88]. Methylomonas sp. DH-1 and M. trichosporium OB3b co-cultures have the potential to produce PHAs from CH4 and CO2 under C-rich or C-lean conditions without an external supply of O2 [32]. Using simulated biogas, the syntrophic association of these methanotrophs resulted in biomass and PHA production of 1130 and 83.0 mg/L, respectively, in the presence of a 2.9 mM N source. Under N-limiting conditions, Methylocystis spp. MJC1 produced PHAs up to 44.5% of the cdw [89]. Genomic investigations have suggested that Methylocystis sp. MJC1 contains pMMO and sMMO particulates that are not commonly reported in Methylocystis spp.
A numerical simulation analysis was conducted to describe the growth and PHA production by methanotrophs based on the roles of the gas flow rate, metabolic heat release, reactor pressure, and pH [90]. This analysis suggested that a high mass-transfer rate and high-pressure operational conditions are undesirable for achieving high PHA productivity. For the first time, Chau et al. [40], confirmed PHA production by the type I methanotroph Methylotuvimicrobium alcaliphilum 20Z. It exhibited production of 3HB and P(3HB) of 334 mg/L and 1.3% of the cdw upon co-feeding with CH4 and xylose. Methylosinus-dominant (54%) consortia produced PHAs of 183 mg/L from 12.3% (v/v) of CH4 and Cu (10 µM) in the media [91]. Remarkably, this consortium tolerated a high CH4 content of up to 70% (v/v). PHB, as an intracellular reducing agent, helps produce high levels of methanol 400 mg/L by M. hirsuta via a 75% reduction in formate requirements [92]. M. hirsuta CSC1 exhibited better PHA production at room temperature over 15 and 37 °C. A maximum PHA content of 45% of the cdw was reported at a 2:1 ratio of O2 to CH4 [93]. Under N-limiting conditions, M. parvus, a type II methanotroph, can produce PHAs up to 50% of the cdw [94]. The AD sludge-based enriched CH4-utilizing mixed cultures produced P(3HB) of approximately 51.0% of the cdw from CH4 [95]. Furthermore, supplementation with valerate enhanced the PHAs up to 52.0%, with a 3HV content of 33 mol%. Response surface methodology was used to optimize PHA production by M. trichosporium OB3b by co-feeding with methanol and CH4 [96]. The maximum PHAs produced were reported to be up to 48.7 mg/L with a cdw content of 52.5%. The methanotroph population enriched from compost soils and landfill top cover showed PHA production of up to 25 mg/g of the cdw using 40% CH4 as feed [97]. Details of PHA production by methanotrophs are presented in Table 1.

4. Genetic Engineering in Methanotrophs for Producing Polyhydroxyalkanoates (PHAs)

Polyhydroxyalkanoates (PHAs) are biopolymers with enormous potential for numerous applications, including biodegradable plastics. Methanotrophs exploit CH4 as their carbon and energy source and are promising producers of PHAs [6,94]. One key advantage of using methanotrophs for producing PHAs is the availability of CH4 feedstock as a GHG source. Methane (CH4) is the chief constituent of natural gas and can be obtained through the AD of organic wastes of agricultural, municipal, industrial, or synthetic origins. Methanotrophs convert CH4 into PHAs, which can then be extracted and processed in a manner similar to traditional petroleum-derived plastics [1,8]. Current research is focused on expanding the range of PHAs synthesized by methanotrophic bacteria [6,40]. Although early investigations have demonstrated the production of PHAs, their ability to produce other PHAs has been limited because P(3HB) is a major product [6,67]. Genetic engineering approaches have shown promise in broadening the monomer composition of PHAs formed by methanotrophs. Researchers have demonstrated PHA production with varying chain lengths and functional groups by manipulating the metabolic pathways and enzymatic activities within methanotrophic bacteria [18,37,83]. This allows for tailoring the material properties of biopolymers, such as their mechanical strength, flexibility, and biodegradability, to suit specific biotechnological applications. For the first time, Bordel et al. [98] demonstrated a type II methanotroph genome-scale metabolic model construction in Methylocystis hirsuta. Directing type II methanotrophs through metabolic engineering for high C fluxes via acetoacetyl-CoA under N-limited conditions could be a valuable platform for producing PHAs [23,98]. Engineered M. trichosporium OB3b through reconstructing 4HB biosynthetic pathways, such as NADPH-dependent succinate semialdehyde-reductase and CoA-dependent succinate semialdehyde-dehydrogenase, showed 3.08 mol% of 4HB in the produced PHAs (P(3HB-co-4HB)) from CH4 [99]. The Bacillus subtillis 168 lipase A gene was successfully expressed in M. trichosporium Ob3b to produce lipase and PHAs simultaneously [100]). The engineered methanotrophs exhibited maximum PHA production rates of 191 mg/L and 71.5 U/mg of lipase activity. In vivo quantification of PHAs by rapid and efficient approaches in a Methylocystis sp. A Rockwell model was demonstrated by Lazic et al. [101]. In contrast, PHA polymerase (phaC) deletion in M. trichosporium OB3b showed a high production of 2-hydroxyisobutyric acid (30.0 mg/L) and 1,3-butanediol (60.5 mg/L) from CH4 as feed [38]). In addition, few simulation studies have been demonstrated producing P(3HB) by methanotrophs using natural gas [102,103]. Overall, the use of genetically engineered methanotrophic bacteria for PHA production represents a promising approach to address the challenges associated with environmental pollution and traditional petroleum-based plastics.

5. Perspectives and Conclusions

The production of PHAs is an emerging alternative to conventional petroleum-derived synthetic plastics owing to its biodegradability and sustainability. These biopolymers can be formed using various renewable sources, including agricultural waste, which can help diminish the environmental pollution caused by plastic waste. Methanotrophs have garnered considerable attention because of their broad applications in numerous areas, such as bioremediation, biotechnology, and biofuel production. One particularly promising aspect of methanotrophs is their ability to synthesize PHAs, which are a class of biodegradable and renewable bioplastics. Methanotrophs produce PHAs under certain growth conditions. However, the commercial-scale production of PHAs by methanotrophs faces several challenges and limitations. A key challenge is the low yield and productivity of PHAs from methanotrophs compared to other microbial sources. To address these challenges, researchers have explored alternative substrates and cultivation strategies for PHA production using methanotrophs. For instance, using agricultural waste-derived biogas as a source of CH4 feedstock could offer a more sustainable and lucrative approach while reducing the environmental impact of waste. Therefore, integrating PHA production with other bioprocesses, such as wastewater treatment and biogas generation, could enhance the overall sustainability and economic viability of the process. PHAs have been extensively investigated for biomedical applications, such as implants and drug delivery systems, owing to their inherent biocompatibility and biodegradability. The use of additive manufacturing techniques, such as 3D printing, has further expanded the potential applications of PHB in the biomedical field, allowing for the fabrication of customized complex structures. In addition, PHA properties (physical and chemical), such as their mechanical strength, biodegradability, and biocompatibility, can be tailored by copolymerizing different hydroxyalkanoate monomers. This versatility makes PHAs produced by methanotrophs valuable materials for biomedical applications, such as tissue engineering, drug delivery, and medical implants. Recent legislation, such as the Circular Economy Action Plan 2020, has driven the development of more sustainable production processes and better end-of-life solutions for bio-derived plastics. By addressing the challenges and limitations of PHA production by methanotrophs, researchers can contribute to the development of a more circular and sustainable plastics economy. Further improvements in fermentation technology, metabolic engineering, and process optimization are required to increase the production efficiency of PHAs. In conclusion, the production of PHAs by methanotrophic bacteria represents a promising approach for addressing environmental concerns associated with traditional plastic production. However, further research and development is essential to overcome the technological and ecological challenges that currently limit the large-scale adoption of this technology.

Author Contributions

Conceptualization, S.K.S.P. and J.-K.L.; data curation, D.S., D.P., R.K.G., S.B., R.V.S. and S.K.S.P.; writing—original draft preparation, S.K.S.P.; writing—review and editing, S.K.S.P. and J.-K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Science, ICT, and Future Planning (NRF-2022M3A9I5015091, RS-2024-00351665, and RS-2023-00222078).

Acknowledgments

The author would like to thank Vice Chancellor, Hemvati Nandan Bahuguna Garhwal University (A Central University), Srinagar, for providing the necessary support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ge, X.; Yang, L.; Sheets, J.P.; Yu, Z.; Li, Y. Biological conversion of methane to liquid fuels: Status and opportunities. Biotechnol. Adv. 2014, 32, 1460–1475. [Google Scholar] [CrossRef] [PubMed]
  2. Patel, S.K.S.; Das, D.; Kim, S.C.; Cho, B.-K.; Kalia, V.C.; Lee, J.-K. Integrating strategies for sustainable conversion of waste biomass into dark-fermentative hydrogen and value-added products. Renew. Sustain. Energ. Rev. 2021, 150, 111491. [Google Scholar] [CrossRef]
  3. Singh, R.; Ryu, J.; Kim, S.W. An overview on methanotrophs and the role of Methylosinus trichosporium OB3b for biotechnological applications. Biotechnol. Bioprocess Eng. 2022, 27, 468–481. [Google Scholar] [CrossRef]
  4. Bachleitner, S.; Ata, O.; Mattanovich, D. The potential of CO2-based production cycles in biotechnology to fight the climate crisis. Nat. Commun. 2023, 14, 6978. [Google Scholar] [CrossRef] [PubMed]
  5. Vasudevan, U.M.; Mai, D.H.A.; Krishna, S.; Lee, E.Y. Methanotrophs as a reservoir for bioactive secondary metabolites: Pitfalls, insights and promises. Biotechnol. Adv. 2023, 63, 108097. [Google Scholar]
  6. Fei, Q.; Guarnieri, M.T.; Tao, L.; Laurens, L.M.L.; Dowe, N.; Pienkos, P.T. Bioconversion of natural gas to liquid fuel: Opportunities and challenges. Biotechnol. Adv. 2014, 32, 596–614. [Google Scholar] [CrossRef]
  7. Patel, S.K.S.; Shanmugam, R.; Lee, J.-K.; Kalia, V.C.; Kim, I.-W. Biomolecules production from greenhouse gases by methanotrophs. Indian J. Microbiol. 2021, 61, 449–457. [Google Scholar] [CrossRef]
  8. Chen, R.; Weng, G.-M. Sustainable energy resources for driving methane conversion. Adv. Energy Mater. 2023, 13, 2301734. [Google Scholar] [CrossRef]
  9. Patel, S.K.S.; Jeon, M.S.; Gupta, R.K.; Jeon, Y.; Kalia, V.C.; Kim, S.C.; Cho, B.K.; Kim, D.R.; Lee, J.-K. Hierarchical macroporous particles for efficient whole-cell immobilization: Application in bioconversion of greenhouse gases to methanol. ACS Appl. Mater. Interfaces 2019, 11, 18968–18977. [Google Scholar] [CrossRef]
  10. Patel, S.K.S.; Gupta, R.K.; Kalia, V.C.; Lee, J.-K. Integrating anaerobic digestion of potato peels to methanol production by methanotrophs immobilized on banana leaves. Bioresour. Technol. 2021, 323, 124550. [Google Scholar] [CrossRef]
  11. Samanta, D.; Sani, R.K. Methane oxidation via chemical and biological methods: Challenges and solutions. Methane 2023, 2, 279–303. [Google Scholar] [CrossRef]
  12. Patel, S.K.S.; Mardina, P.; Kim, D.; Kim, S.-Y.; Kalia, V.C.; Kim, I.-W.; Lee, J.-K. Improvement in methanol production by regulating the composition of synthetic gas mixture and raw biogas. Bioresour. Technol. 2016, 218, 202–208. [Google Scholar] [CrossRef] [PubMed]
  13. Lv, X.; Yu, W.; Zhang, C.; Ning, P.; Li, J.; Liu, Y.; Du, G.; Liu, L. C1-based biomanufacturing: Advances, challenges and perspectives. Bioresour. Technol. 2023, 367, 128259. [Google Scholar] [CrossRef] [PubMed]
  14. Patel, S.K.S.; Kalia, V.C.; Lee, J.-K. Integration of biogas derived from dark fermentation and anaerobic digestion of biowaste to enhance methanol production by methanotrophs. Bioresour. Technol. 2023, 367, 128427. [Google Scholar] [CrossRef] [PubMed]
  15. Patel, S.K.S.; Gupta, R.K.; Kondaveeti, S.; Otari, S.V.; Kumar, A.; Kalia, V.C.; Lee, J.-K. Conversion of biogas to methanol by methanotrophs immobilized on chemically modified chitosan. Bioresour. Technol. 2020, 315, 123791. [Google Scholar] [CrossRef]
  16. Patel, S.K.S.; Kalia, V.C.; Joo, J.B.; Kang, Y.C.; Lee, J.-K. Biotransformation of methane into methanol by methanotrophs immobilized on coconut coir. Bioresour. Technol. 2020, 297, 122433. [Google Scholar] [CrossRef]
  17. Priyadarsini, A.; Singh, R.; Barbora, L.; Maitra, S.S.; Moholkar, V.S. Methanotroph detection and bioconversion of methane to methanol by enriched microbial consortium from rice field soil. Bioresour. Biotechnol. Rep. 2023, 22, 101410. [Google Scholar] [CrossRef]
  18. Zheng, X.-C.; Li, H.-S.; Wang, Z.-H.; Sun, Z.-F.; Zhao, L. Intermediates production in methane oxidation coupled with denitrification: Current status, challenges, and future opportunities. Fermentation 2023, 9, 645. [Google Scholar] [CrossRef]
  19. Kalia, V.C.; Patel, S.K.S.; Shanmugam, R.; Lee, J.-K. Polyhydroxyalkanoates: Trends and advances towards biotechnological applications. Bioresour. Technol. 2021, 326, 124737. [Google Scholar] [CrossRef]
  20. Kalia, V.C.; Patel, S.K.S.; Lee, J.-K. Exploiting polyhydroxyalkanoates for biomedical applications. Polymers 2023, 15, 1937. [Google Scholar] [CrossRef] [PubMed]
  21. Kalia, V.C.; Patel, S.K.S.; Karthikeyan, K.K.; Jeya, M.; Kim, I.-W.; Lee, J.-K. Manipulating microbial cell morphology for the sustainable production of biopolymers. Polymers 2024, 16, 410. [Google Scholar] [CrossRef] [PubMed]
  22. Patel, S.K.S.; Gupta, R.K.; Kalia, V.C.; Lee, J.-K. Synthetic design of methanotroph co-cultures and their immobilization within polymers containing magnetic nanoparticles to enhance methanol production from wheat straw-based biogas. Bioresour. Technol. 2022, 364, 128032. [Google Scholar] [CrossRef] [PubMed]
  23. Pieja, A.J.; Morse, M.C.; Cal, A.J. Methane to bioproducts: The future of the bioeconomy? Curr. Opin. Chem. Biol. 2017, 41, 123–131. [Google Scholar] [CrossRef] [PubMed]
  24. Fenibo, E.O.; Selvarajan, R.; Wang, H.; Wang, Y.; Abia, A.L.K. Untapped talents, insight into the ecological significance of methanotrophs and its prospects. Sci. Total Environ. 2023, 903, 166145. [Google Scholar] [CrossRef]
  25. Yoon, J.; Oh, M.-K. Strategies for biosynthesis of C1 gas-derived polyhydroxyalkanoates: A review. Bioresour. Technol. 2022, 344, 126307. [Google Scholar] [CrossRef]
  26. Patel, S.K.S.; Lee, J.-K. Plastic eating enzymes: A step towards sustainability. Indian J. Microbiol. 2022, 62, 658–661. [Google Scholar] [CrossRef]
  27. Liu, L.-Y.; Xie, G.-J.; Xing, D.-F.; Liu, B.-F.; Ding, J.; Ren, N.-Q. Biological conversion of methane to polyhydroxyalkanoates, Current advances, challenges, and perspectives. Environ. Sci. Ecotechnol. 2021, 2, 100029. [Google Scholar] [CrossRef]
  28. Bagi, Z.; Ács, N.; Böjti, T.; Kakuk, B.; Rakhely, G.; Strang, O.; Szuhal, M.; Wirth, R.; Kovacs, K.L. Biomethane: The energy storage, platform chemical and greenhouse gas mitigation target. Anaerobe 2017, 46, 13–22. [Google Scholar] [CrossRef]
  29. Amabile, C.; Abate, T.; De Crescenzo, C.; Muñoz, R.; Sabbaresse, S.; Chianese, S.; Musmarra, D. Sustainable process for the production of poly(3-hydroxybutyrate- co-3-hydroxyvalerate) from renewable resources: A simulation study. ACS Sustain. Chem. Eng. 2022, 10, 14230–14239. [Google Scholar] [CrossRef]
  30. Salem, R.; Eldyasti, A.; Audette, G.F. Biomedical applications of biomolecules isolated from methanotrophic bacteria in wastewater treatment systems. Biomolecules 2021, 11, 1217. [Google Scholar] [CrossRef]
  31. Amabile, C.; Abate, T.; Muñoz, R.; Chianese, S.; Musmarra, D. Production of poly(3-hydroxybutyrate) and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) from methane and volatile fatty acids, properties, metabolic routes and current trend. Sci. Total Environ. 2024, 927, 172138. [Google Scholar] [CrossRef] [PubMed]
  32. Ashoor, S.; Jun, S.-H.; Ko, H.D.; Lee, J.; Hamelin, J.; Milferstedt, K.; Na, J.-G. Polyhydroxybutyrate production from methane and carbon dioxide by a syntrophic consortium of methanotrophs with oxygenic photogranules without an external oxygen supply. Microorganisms 2023, 11, 1110. [Google Scholar] [CrossRef] [PubMed]
  33. Patel, S.K.S.; Gupta, R.K.; Kim, I.-W.; Lee, J.-K. Encapsulation of methanotrophs within a polymeric matrix containing copper-and iron-based nanoparticles to enhance methanol production from a simulated biogas. Polymers 2023, 15, 3667. [Google Scholar] [CrossRef] [PubMed]
  34. Patel, S.K.S.; Kumar, V.; Mardina, P.; Li, J.; Lestari, R.; Kalia, V.C.; Lee, J.-K. Methanol production from simulated biogas mixtures by co-immobilized Methylomonas methanica and Methylocella tundrae. Bioresour. Technol. 2018, 263, 25–32. [Google Scholar] [CrossRef] [PubMed]
  35. Bordel, S.; Rodríguez, E.; Muñoz, R. Genome sequence of Methylocystis hirsuta CSC1, a polyhydroxyalkanoate producing methanotroph. MicrobiologyOpen 2019, 8, e00771. [Google Scholar] [CrossRef]
  36. Patel, S.K.S.; Shanmugam, R.; Kalia, V.C.; Lee, J.-K. Methanol production by polymer-encapsulated methanotrophs from simulated biogas in the presence of methane vector. Bioresour. Technol. 2020, 304, 123022. [Google Scholar] [CrossRef]
  37. Xin, J.Y.; Zhang, Y.X.; Zhang, S.; Xia, C.G.; Li, S.B. Methanol production from CO2 by resting cells of the methanotrophic bacterium Methylosinus trichosporium IMV3011. J. Basic Microbiol. 2007, 47, 426–435. [Google Scholar] [CrossRef]
  38. Mai, D.H.A.; Nguyen, T.T.; Lee, E.Y. The ethylmalonyl-CoA pathway for methane-based biorefineries: A case study of using Methylosinus trichosporium OB3b, an alpha-proteobacterial methanotroph, for producing 2-hydroxyisobutyric acid and 1,3-butanediol from methane. Green Chem. 2021, 23, 7712–7723. [Google Scholar] [CrossRef]
  39. Kim, Y.; Flinkstrom, Z.; Candry, P.; Winkler, M.-K.H.; Myung, J. Resource availability governs polyhydroxyalkanoate (PHA) accumulation and diversity of methanotrophic enrichments from wetlands. Front. Bioeng. Biotechnol. 2023, 11, 1210392. [Google Scholar] [CrossRef]
  40. Chau, T.H.T.; Duc Nguyen, A.; Lee, E.Y. Engineering type I methanotrophic bacteria as novel platform for sustainable production of 3-hydroxybutyrate and biodegradable polyhydroxybutyrate from methane and xylose. Bioresour. Technol. 2022, 363, 127898. [Google Scholar]
  41. Ray, S.; Jin, J.-O.; Choi, I.; Kim, M. Recent trends of biotechnological production of polyhydroxyalkanoates from C1 carbon sources. Front. Bioeng. Biotechnol. 2023, 10, 907500. [Google Scholar] [CrossRef] [PubMed]
  42. Myung, J.; Galega, W.M.; Van Nostrand, J.D.; Yuan, T.; Zhou, J.; Criddle, C.S. Long-term cultivation of a stable Methylocystis-dominated methanotrophic enrichment enabling tailored production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Bioresour. Technol. 2015, 198, 811–818. [Google Scholar] [CrossRef] [PubMed]
  43. Myung, J.; Kim, M.; Pan, M.; Criddle, C.S.; Tang, S. Low energy emulsion-based fermentation enabling accelerated methane mass transfer and growth of poly(3-hydroxybutyrate)-accumulating methanotrophs. Bioresour. Technol. 2016, 207, 302–307. [Google Scholar] [CrossRef] [PubMed]
  44. Myung, J.; Flanagan, J.C.A.; Waymouth, R.M.; Criddle, C.S. Methane or methanol-oxidation dependent synthesis of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by obligate type II methanotrophs. Process Biochem. 2016, 51, 561–567. [Google Scholar] [CrossRef]
  45. López, J.C.; Quijano, G.; Pérez, R.; Muñoz, R. Assessing the influence of CH4 concentration during culture enrichment on the biodegradation kinetics and population structure. J. Environ. Manag. 2014, 146, 116–123. [Google Scholar] [CrossRef] [PubMed]
  46. López, J.C.; Quijano, G.; Pérez, R.; Muñoz, R. Effect of pollutant concentration during isolation on the CH4 biodegradation kinetics, population structure and PHB accumulation. Chem. Eng. Transact. 2014, 40, 211–216. [Google Scholar]
  47. Helm, J.; Wendlandt, K.-D.; Rogge, G.; Kappelmeyer, U. Characterizing a stable methane-utilizing mixed culture used in the synthesis of a high-quality biopolymer in an open system. J. Appl. Microbiol. 2006, 101, 387–395. [Google Scholar] [CrossRef]
  48. Helm, J.; Wendlandt, K.D.; Jechorek, M.; Stottmeister, U. Potassium deficiency results in accumulation of ultra-high molecular weight poly-beta-hydroxybutyrate in a methaneutilizing mixed culture. J. Appl. Microbiol. 2008, 105, 1054–1061. [Google Scholar] [CrossRef]
  49. Cal, A.J.; Sikkema, W.D.; Ponce, M.I.; Franqui-Villanueva, D.; Riiff, T.J.; Orts, W.J.; Pieja, A.J.; Lee, C.C. Methanotrophic production of polyhydroxybutyrate-co-hydroxyvalerate with high hydroxyvalerate content. Int. J. Biol. Macromol. 2016, 87, 302–307. [Google Scholar] [CrossRef]
  50. García-Pérez, T.; Lópeza, J.C.; Passosc, F.; Lebreroa, R.; Revah, S.; Muñoza, R. Simultaneous methane abatement and PHB production by Methylocystis hirsuta in a novel gas-recycling bubble column bioreactor. Chem. Eng. J. 2018, 334, 691–697. [Google Scholar] [CrossRef]
  51. Cantera, S.; Sánchez-Andrea, I.; Lebrero, R.; García-Encina, P.A.; Stams, A.J.M.; Muñoz, R. Multi-production of high added market value metabolites from diluted methane emissions via methanotrophic extremophiles. Bioresour. Technol. 2018, 267, 401–407. [Google Scholar] [CrossRef] [PubMed]
  52. Wendlandt, K.D.; Jechorek, M.; Helm, J.; Stottmeister, U. Producing poly-3-hydroxybutyrate with a high molecular mass from methane. J. Biotechnol. 2001, 86, 127–133. [Google Scholar] [CrossRef] [PubMed]
  53. Rostkowski, K.H.; Pfluger, A.R.; Criddle, C.S. Stoichiometry and kinetics of the PHB-producing Type II methanotrophs Methylosinus trichosporium OB3b and Methylocystis parvus OBBP. Bioresour. Technol. 2013, 132, 71–77. [Google Scholar] [CrossRef] [PubMed]
  54. Levett, I.; Birkett, G.; Davies, N.; Bell, A.; Langford, A.; Laycock, B.; Lant, P.; Pratt, S. Techno-economic assessment of poly-3-hydroxybutyrate (PHB) production from methane—The case for thermophilic bioprocessing. J. Environ. Chem. Eng. 2016, 4, 3724–3733. [Google Scholar] [CrossRef]
  55. López, J.C.; Arnáiz, E.; Merchán, L.; Lebrero, R.; Muñoz, R. Biogas-based polyhydroxyalkanoates production by Methylocystis hirsuta: A step further in anaerobic digestion biorefineries. Chem. Eng. J. 2018, 333, 529–536. [Google Scholar] [CrossRef]
  56. Pérez, R.; Cantera, S.; Bordel, S.; Garcia-Encina, P.A.; Muñoz, R. The effect of temperature during culture enrichment on methanotrophic polyhydroxyalkanoate production. Int. Biodeter. Biodegrad. 2019, 140, 144–151. [Google Scholar] [CrossRef]
  57. Luangthongkam, P.; Laycock, B.; Evans, P.; Lant, P.; Pratt, S. Thermophilic production of poly(3-hydroxybutyrate-co-3-hydrovalerate) by a mixed methane-utilizing culture. New Biotechnol. 2019, 53, 49–56. [Google Scholar] [CrossRef]
  58. Pérez, R.; Casal, J.; Muñoz, R.; Lebrero, R. Polyhydroxyalkanoates production from methane emissions in Sphagnum mosses: Assessing the effect of temperature and phosphorus limitation. Sci. Total Environ. 2019, 688, 684–690. [Google Scholar] [CrossRef]
  59. Luangthongkam, P.; Strong, P.J.; Syed Mahamud, S.N.; Evans, P.; Jensen, P.; Tyson, G.; Laycock, B.; Lant, P.A.; Pratt, S. The effect of methane and odd-chain fatty acids on 3-hydroxybutyrate (3HB) and 3-hydroxyvalerate (3HV) synthesis by a Methylosinus-dominated mixed culture. Bioresour. Bioproces. 2019, 6, 50. [Google Scholar] [CrossRef]
  60. Eam, H.; Ko, D.; Lee, C.; Myung, J. Methylosinus trichosporium OB3b bioaugmentation unleashes polyhydroxybutyrate-accumulating potential in waste-activated sludge. Microb. Cell Fact. 2024, 23, 160. [Google Scholar] [CrossRef]
  61. Pham, D.N.; Mai, D.H.A.; Lee, E.Y. Biosynthesis of polyhydroxybutyrate from methane and carbon dioxide using type II methanotrophs. Bioresour. Technol. 2024, 405, 130931. [Google Scholar] [CrossRef] [PubMed]
  62. Delherbe, N.A.; Pearce, D.; But, S.Y.; Murrell, J.C.; Khmelenina, V.N.; Kalyuzhnaya, M.G. Genomic insights into moderately thermophilic methanotrophs of the genus Methylocaldum. Microorganisms 2024, 12, 469. [Google Scholar] [CrossRef] [PubMed]
  63. Myung, J.; Flanagan, J.C.A.; Waymouth, R.M.; Criddle, C.S. Ethane-dependent synthesis of polyhydroxyalkanoates by the obligate methanotroph Methylocystis parvus OBBP. Chem. Eng. J. 2024, 483, 149210. [Google Scholar] [CrossRef]
  64. Gęsicka, A.; Gutowska, N.; Palaniappan, S.; Oleskowicz-Popiel, P.; Łężyk, M. Enrichment of mixed methanotrophic cultures producing polyhydroxyalkanoates (PHAs) from various environmental sources. Sci. Total Environ. 2024, 912, 168844. [Google Scholar] [CrossRef] [PubMed]
  65. Pieja, A.J.; Sundstrom, E.R.; Criddle, C.S. Poly-3-hydroxybutyrate metabolism in the type II methanotroph Methylocystis parvus OBBP. Appl. Environ. Microbiol. 2011, 77, 6012–6019. [Google Scholar] [CrossRef] [PubMed]
  66. Pieja, A.J.; Sundstrom, E.R.; Criddle, C.S. Cyclic, alternating methane and nitrogen limitation increases PHB production in a methanotrophic community. Bioresour. Technol. 2012, 107, 385–392. [Google Scholar] [CrossRef]
  67. Pieja, A.J.; Rostkowski, K.H.; Criddle, C.S. Distribution and selection of poly-3-hydroxybutyrate production capacity in methanotrophic proteobacteria. Microb. Ecol. 2011, 62, 564–573. [Google Scholar] [CrossRef]
  68. Tran, M.H.; Choi, T.-R.; Yang, Y.-H.; Lee, O.K.; Lee, E.Y. An efficient and eco-friendly approach for the sustainable recovery and properties characterization of polyhydroxyalkanoates produced by methanotrophs. Int. J. Biol. Macromol. 2023, 257, 128687. [Google Scholar] [CrossRef]
  69. de Sousa Junior, R.R.; Cezario, F.E.M.; Antonino, L.D.; dos Santos, D.J.; Lackner, M. Characterization of poly(3-hydroxybutyrate) (P3HB) from alternative, scalable (Waste) feedstocks. Bioengineering 2023, 10, 1382. [Google Scholar] [CrossRef]
  70. Lee, O.K.; Kang, S.G.; Choi, T.-R.; Yang, Y.-H.; Lee, E.Y. Production and characterization of a biodegradable polymer, poly(3-hydroxybutyrate-co-3-hydroxyvalerate), using the type II methanotroph, Methylocystis sp. MJC1. Bioresour. Technol. 2023, 389, 129853. [Google Scholar] [CrossRef]
  71. Safaeian, P.; Yazdian, F.; Khosravi-Darani, K.; Rashedi, H.; Lackner, M. P3HB from CH4 using methanotrophs: Aspects of bioreactor, fermentation process and modelling for cost-effective biopolymer production. Front. Bioeng. Biotechnol. 2023, 11, 1137749. [Google Scholar] [CrossRef] [PubMed]
  72. Amabile, C.; Abate, T.; De Crescenzo, C.; Sabbaresse, S.; Migliaccio, A.; Chianese, S.; Musmarra, D. Poly(3-hydroxybutyrate) production from methane in bubble column bioreactors: Process simulation and design optimization. New Biotechnol. 2022, 70, 39–48. [Google Scholar] [CrossRef] [PubMed]
  73. Kulkarni, P.P.; Chavan, S.B.; Deshpande, M.S.; Sagotra, D.; Kumbhar, P.S.; Ghosalkar, A.R. Enrichment of Methylocystis dominant mixed culture from rice field for PHB production. J. Biotechnol. 2022, 343, 62–70. [Google Scholar] [CrossRef] [PubMed]
  74. Vecherskaya, M.; Dijkema, C.; Stams, A.J. Intracellular PHB conversion in a type II methanotroph studied by 13C NMR. J. Ind. Microbiol. Biotechnol. 2001, 26, 15–21. [Google Scholar] [CrossRef]
  75. Wendlandt, K.D.; Geyer, W.; Mirschel, G.; Hemidi, F.A. Possibilities for controlling a PHB accumulation process using various analytical methods. J. Biotechnol. 2005, 117, 119–129. [Google Scholar] [CrossRef]
  76. Xin, J.; Zhang, Y.; Dong, J.; Song, H.; Xia, C. An experimental study on molecular weight of poly-3-hydroxybutyrate (PHB) accumulated in Methylosinus trichosporium IMV 3011. Afr. J. Biotechnol. 2011, 10, 7078–7087. [Google Scholar]
  77. Zhang, Y.; Xin, J.; Chen, L.; Song, H.; Xia, C. Biosynthesis of poly-3-hydroxybutyrate with a high molecular weight by methanotroph from methane and methanol. J. Nat. Gas Chem. 2008, 17, 103–109. [Google Scholar] [CrossRef]
  78. Zúniga, C.; Morales, M.; Le Borgne, S.; Revah, S. Production of poly-β-hydroxybutyrate (PHB) by Methylobacterium organophilum isolated from a methanotrophic consortium in a two-phase partition bioreactor. J. Hazard. Mater. 2011, 190, 876–882. [Google Scholar] [CrossRef]
  79. Zúniga, C.; Morales, M.; Revah, S. Polyhydroxyalkanoates accumulation by Methylobacterium organophilum CZ-2 during methane degradation using citrate or propionate as cosubstrates. Bioresour. Technol. 2013, 129, 686–689. [Google Scholar] [CrossRef]
  80. Malvar, S.; Cardoso, L.O.B.; Karolski, B.; Perpetuo, E.A.; Carmo, B.S.; Meneghini, J.R. A rheological approach to identify efficient biopolymer producing bacteria. Biotechnol. Bioeng. 2021, 118, 622–632. [Google Scholar] [CrossRef]
  81. Chen, X.; Rodríguez, Y.; López, J.C.; Muñoz, R.; Ni, B.-J.; Sin, G. Modeling of polyhydroxyalkanoate synthesis from biogas by Methylocystis hirsuta. ACS Sustain. Chem. Eng. 2020, 8, 3906–3912. [Google Scholar] [CrossRef]
  82. Carere, C.R.; McDonald, B.; Peach, H.A.; Greening, C.; Gapes, D.J.; Collet, C.; Stott, M.B. Hydrogen oxidation influences glycogen accumulation in a verrucomicrobial methanotroph. Front. Microbiol. 2019, 10, 1873. [Google Scholar] [CrossRef] [PubMed]
  83. Myung, J.; Flanagan, J.C.A.; Waymouth, R.M.; Criddle, C.S. Expanding the range of polyhydroxyalkanoates synthesized by methanotrophic bacteria through the utilization of omega-hydroxyalkanoate co-substrates. AMB Express 2017, 7, 118. [Google Scholar] [CrossRef] [PubMed]
  84. Chidambarampadmavathy, K.; Karthikeyan, O.P.; Heimann, K. Sustainable bio-plastic production through landfill methane recycling. Renew. Sustain. Energy Rev. 2017, 71, 555–562. [Google Scholar] [CrossRef]
  85. Flanagan, J.C.A.; Myung, J.; Criddle, C.S.; Waymouth, R.M. Poly(hydroxyalkanoate)s from waste biomass: A combined chemical–biological approach. ChemistrySelect 2016, 1, 2327–2331. [Google Scholar] [CrossRef]
  86. Hyun, S.W.; Krishna, S.; Chau, T.H.T.; Lee, E.Y. Methanotrophs mediated biogas valorization: Sustainable route to polyhydroxybutyrate production. Bioresour. Technol. 2024, 402, 130759. [Google Scholar] [CrossRef]
  87. Pérez, V.; Lebrero, R.; Muñoz, R.; Pérez, R. The fundamental role of pH in CH4 bioconversion into polyhydroxybutyrate in mixed methanotrophic cultures. Chemosphere 2024, 355, 141832. [Google Scholar] [CrossRef]
  88. Hong, H.J.; Hyung, J.S.; Lee, J.; Na, J.-G. Effects of methane to oxygen ratio on cell growth and polyhydroxybutyrate synthesis in high cell density cultivation of Methylocystis sp. MJC1. Environ. Sci. Pollut. Res. 2024, in press. [Google Scholar] [CrossRef]
  89. Naizabekov, S.; Hyun, S.W.; Na, J.-G.; Yoon, S.; Lee, O.K.; Lee, E.Y. Comparative genomic analysis of Methylocystis sp. MJC1 as a platform strain for polyhydroxybutyrate biosynthesis. PLoS ONE 2023, 18, e0284846. [Google Scholar] [CrossRef]
  90. Meraz, J.L.; Abel, A.J.; Clark, D.S.; Criddle, C.S. Biological conversion of methane to bioplastics: Kinetics, stoichiometry, and thermodynamic considerations for process optimization. Chem. Eng. J. 2023, 454, 140166. [Google Scholar] [CrossRef]
  91. Cardoso, L.O.B.; Karolski, B.; Gracioso, L.H.; Borrego, B.B.; do Nascimento, C.A.O.; Perpetuo, E.A. Enrichment of Methylosinus-dominant consortia from mangroves for polyhydroxybutyrate (PHB) production. J. Environ. Chem. Eng. 2022, 10, 108490. [Google Scholar] [CrossRef]
  92. Fergala, A.; Alsayed, A.; Eldyasti, A. Utilization of polyhydroxybutyrate (PHB) as intracellular reducing power for methanol production to alleviate the reliance on external energy sources by Methylocystis hirsuta. J. Environ. Chem. Eng. 2021, 9, 105314. [Google Scholar] [CrossRef]
  93. Rodríguez, Y.; Firmino, P.I.M.; Arnáiz, E.; Lebrero, R.; Muñoz, R. Elucidating the influence of environmental factors on biogas-based polyhydroxybutyrate production by Methylocystis hirsuta CSC1. Sci. Total Environ. 2020, 706, 135136. [Google Scholar] [CrossRef]
  94. Bordel, S.; Rojas, A.; Muñoz, R. Reconstruction of a genome scale metabolic model of the polyhydroxybutyrate producing methanotroph Methylocystis parvus OBBP. Microb. Cell Fact. 2019, 18, 104. [Google Scholar] [CrossRef] [PubMed]
  95. Fergala, A.; Alsayed, A.; Khattab, S.; Ramirez, M.; Eldyasti, A. Development of methane-utilizing mixed cultures for the production of polyhydroxyalkanoates (PHAs) from anaerobic digester sludge. Environ. Sci. Technol. 2018, 52, 12376–12387. [Google Scholar] [CrossRef] [PubMed]
  96. Carrillo, J.A.Z.; Stein, L.Y.; Sauvageau, D. Defining nutrient combinations for optimal growth and polyhydroxybutyrate production by Methylosinus trichosporium OB3b using response surface methodology. Front. Microbiol. 2018, 9, 1513. [Google Scholar]
  97. Chidambarampadmavathy, K.; Karthikeyan, O.P.; Huerlimann, R.; Maes, G.E.; Heimann, K. Response of mixed methanotrophic consortia to different methane to oxygen ratios. Waste Manag. 2017, 61, 220–228. [Google Scholar] [CrossRef]
  98. Bordel, S.; Rodríguez, Y.; Hakobyan, A.; Rodriguez, E.; Lebrero, R.; Muñoz, R. Genome scale metabolic modeling reveals the metabolic potential of three Type II methanotrophs of the genus Methylocystis. Metab. Eng. 2019, 54, 191–199. [Google Scholar] [CrossRef]
  99. Nguyen, T.T.; Lee, E.Y. Methane-based biosynthesis of 4-hydroxybutyrate and P(3-hydroxybutyrate-co-4-hydroxybutyrate) using engineered Methylosinus trichosporium OB3b. Bioresour. Technol. 2021, 335, 125263. [Google Scholar] [CrossRef]
  100. Kulkarni, P.P.; Khonde, V.K.; Ghosalkar, A.R. Simultaneous production of lipase and PHB using methanotrophic organism Methylosinus trichosporium OB3b. Bioresour. Technol. Rep. 2024, 25, 101751. [Google Scholar] [CrossRef]
  101. Lazic, M.; Gudneppanavar, R.; Whiddon, K.; Sauvageau, D.; Stein, L.Y.; Konopka, M. In vivo quantification of polyhydroxybutyrate (PHB) in the alphaproteobacterial methanotroph, Methylocystis sp. Rockwell. Appl. Microbiol. Biotechnol. 2022, 106, 811–819. [Google Scholar] [CrossRef] [PubMed]
  102. Moradi, M.; Rashedi, H.; Mofradnia, S.R.; Khosravi-Darani, K.; Ashouri, R.; Yazdian, F. Polyhydroxybutyrate production from natural gas in a bubble column bioreactor: Simulation using COMSOL. Bioengineering 2019, 6, 84. [Google Scholar] [CrossRef] [PubMed]
  103. Darani, K.K.; Yazdian, F.; Rashedi, H.; Bozorg, N.M.; Moradi, M.; Mofradnia, S.R.; Koller, M. Simulation of bioreactors for PHB production from natural gas. Iran. J. Chem. Chem. Eng. 2020, 39, 1. [Google Scholar]
Figure 1. The polyhydroxyalkanoate (PHA) synthesis pathway in the methanotrophs: sMMOs—soluble methane (CH4) monooxygenases, pMMOs—particulate CH4 monooxygenases, NADH/NAD+—nicotinamide adenine dinucleotide, MDH—methanol dehydrogenase, PQQ/PQQH2—pyrroloquinoline quinone, FalDH—formaldehyde dehydrogenase, FDH—formate dehydrogenase, RuMP—ribulose monophosphate pathway, PEP—phosphoenolpyruvate, phaA—3-ketothiolase, phaB—acetoacetyl-CoA-reductase, phaC—PHAs synthetase, and phaZ—polyhydroxybutyrate depolymerase.
Figure 1. The polyhydroxyalkanoate (PHA) synthesis pathway in the methanotrophs: sMMOs—soluble methane (CH4) monooxygenases, pMMOs—particulate CH4 monooxygenases, NADH/NAD+—nicotinamide adenine dinucleotide, MDH—methanol dehydrogenase, PQQ/PQQH2—pyrroloquinoline quinone, FalDH—formaldehyde dehydrogenase, FDH—formate dehydrogenase, RuMP—ribulose monophosphate pathway, PEP—phosphoenolpyruvate, phaA—3-ketothiolase, phaB—acetoacetyl-CoA-reductase, phaC—PHAs synthetase, and phaZ—polyhydroxybutyrate depolymerase.
Polymers 16 02570 g001
Table 1. Polyhydroxyalkanoate (PHA) production potentials of methanotrophs from methane (CH4).
Table 1. Polyhydroxyalkanoate (PHA) production potentials of methanotrophs from methane (CH4).
MethanotrophsCH4 (%)Fermentation Conditions
(Mode/Working Capacity (L)//Incubation Period (h))
PHAsReference
% in cdwMw (×106)
Enriched methanotrophs/consortia25Batch/70.0/2446.22.41[47]
25Batch/70.0/2410.4–33.61.81–3.10[48]
50Batch/0.05/2446.0-[67]
80 aContinuous/2.00/38434.0-[78]
150 bContinuous/4.00/2425.0-[66]
5Continuous/0.40/43212.6-[45]
5Continuous/0.40/3101.0–12.6-[46]
50Batch/0.05/4839.0–45.00.93–1.20[42]
40Batch/0.20/48025.0 d-[97]
46 aContinuous/2.00/1500.01 e-[51]
50Batch/0.05/4851.0-[95]
50Semi-continuous/0.24/728.6–14.1-[59]
50Batch/0.03/488.0–10.0-[57]
177 bBatch/0.20/-34.1–35.1-[56]
161 bBatch/0.20/38413.6-[58]
- cBatch/0.05/1920.19 e-[80]
12.3Batch/-/2680.18 e-[91]
50Continuous/8.00/12022.22.20[72]
20Batch/0.05/14412.6-[39]
30Batch/0.04/16812.9-[64]
9Batch/2.50/19243.7-[87]
Activated sludge and Methylosinus trichosporium OB3b50Batch/0.04/7237.1-[60]
Methanotrophic bacterium MTS25Batch/0.30/-3.00-[74]
Methylobacterium organophilus CZ-280 aContinuous/2.00/38438.0–39.0-[78]
42 aContinuous/2.00/24088.0-[79]
Methylococcus capsulatus HD6T50Batch/0.10/120-0.95[76]
Methylocystis4Semi-continuous/400/-37.5-[72]
Methylocystis 42/2250Batch/0.05/2425.0-[67]
Methylocystis SC250Batch/0.05/2430.0-[67]
Methylocystis hirsuta CSC150Batch/0.05/247.0-[67]
29.2Batch/0.40/50445.0-[93]
M. hirsuta DSMZ 185004Batch/2.50/163228–34.6-[50]
35Batch/0.05/16845.0–54.0-[55]
M. hirsuta50Continuous/10.0/12051.6-[71]
Methylocystis parvus OBBP50Batch/0.05/6630.5–50.3-[65]
50Batch/0.05/2436.0-[67]
30Batch/0.05/2260.0-[53]
40Batch/0.05/2448.0–64.0 d1.18–1.47[85]
40Batch/-/16832.0-[43]
40Batch/0.05/4832.0–60.0-[44]
40Batch/0.05/4859.01.22–1.33[83]
40Batch/0.05/2435.0–48.0-[63]
M. parvus74 bBatch/0.02/-50.0-[94]
Methylocystis rosea SV9950Batch/0.05/249.00-[67]
Methylocystis sp. MJC130Batch/0.05/9641.9-[70]
30Batch/3.00/9644.5-[89]
30Batch/2.50/20861.7-[88]
30Batch/1.20/1402.90 e-[86]
20Batch/0.10/2438.0-[61]
Methylocystis sp. GB2520Batch/70.0/2428.3–51.3-[52]
15Batch/70.0/2445.0–51.02.50[75]
-Batch/30.0/504-1.08[69]
Methylocystis strain M50Batch/0.05/2414.0-[67]
Methylocystis sp. WRRC150Batch/0.02/720.20–0.57 e-[49]
Methylomonas sp. GYJB50Batch/0.10/120-0.30[76]
Methylosinus sp. LW450Batch/0.05/24100-[67]
Methylosinus sporium50Batch/0.05/249.00-[67]
M. trichosporium IMV301150Batch/0.10/120-1.20[76]
50Batch/0.05/16841.0 [37]
50Batch/0.10/14438.11.50[77]
M. trichosporium OB3b50Batch/0.05/2438.0 [67]
50Batch/0.10/120-0.95[76]
80 aContinuous/2.00/38457.0-[78]
30Batch/0.05/2829.0–60.0-[53]
50Batch/0.10/12052.5-[95]
Methylotuvimicrobium alcaliphilum 20Z-Batch/0.10/1681.30-[40]
Methylomonas sp. DH-1 and M. trichosporium OB3b30Batch/0.10/1680.08 e-[32]
a CH4 in mg/L·h, b CH4 in mg/L; c not available or reported; d PHAs in mg or mg/g of cdw; e PHAs in g/L.
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

Patel, S.K.S.; Singh, D.; Pant, D.; Gupta, R.K.; Busi, S.; Singh, R.V.; Lee, J.-K. Polyhydroxyalkanoate Production by Methanotrophs: Recent Updates and Perspectives. Polymers 2024, 16, 2570. https://doi.org/10.3390/polym16182570

AMA Style

Patel SKS, Singh D, Pant D, Gupta RK, Busi S, Singh RV, Lee J-K. Polyhydroxyalkanoate Production by Methanotrophs: Recent Updates and Perspectives. Polymers. 2024; 16(18):2570. https://doi.org/10.3390/polym16182570

Chicago/Turabian Style

Patel, Sanjay K. S., Deepshikha Singh, Diksha Pant, Rahul K. Gupta, Siddhardha Busi, Rahul V. Singh, and Jung-Kul Lee. 2024. "Polyhydroxyalkanoate Production by Methanotrophs: Recent Updates and Perspectives" Polymers 16, no. 18: 2570. https://doi.org/10.3390/polym16182570

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop