You are on page 1of 15

Bioresource Technology 344 (2022) 126307

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Strategies for Biosynthesis of C1 Gas-derived Polyhydroxyalkanoates:


A review
Jihee Yoon , Min-Kyu Oh *
Department of Chemical and Biological Engineering, Korea University, Seongbuk-gu, Seoul 02841, Republic of Korea

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Bioplastics production using cheaper


carbon source is crucial for cost
reduction.
• Various strategies to improve PHA pro­
duction from C1 gases were
accomplished.
• Autotrophs, acetogens, and methano­
trophs were optimized and engineered
for PHA.
• Process optimization and PHA copol­
ymer production from C1 gases were
carried out.
• C1 refinery to bioplastics is prospective
methodology to ease environmental
issues.

A R T I C L E I N F O A B S T R A C T

Keywords: Biosynthesis of polyhydroxyalkanoates (PHAs) from C1 gases is highly desirable in solving problems such as
Polyhydroxyalkanoates climate change and microplastic pollution. PHAs are biopolymers synthesized in microbial cells and can be used
Bioplastic as alternatives to petroleum-based plastics because of their biodegradability. Because 50% of the cost of PHA
Carbon dioxide
production is due to organic carbon sources and salts, the utilization of costless C1 gases as carbon sources is
Carbon monoxide
Methane
expected to be a promising approach for PHA production. In this review, strategies for PHA production using C1
gases through fermentation and metabolic engineering are discussed. In particular, autotrophs, acetogens, and
methanotrophs are strains that can produce PHA from CO2, CO, and CH4. In addition, integrated bioprocesses for
the efficient utilization of C1 gases are introduced. Biorefinery processes from C1 gas into bioplastics are pro­
spective strategies with promising potential and feasibility to alleviate environmental issues.

1. Introduction (styrene) (PS), have desirable physical properties that include lightness,
resistance, rigidity, and inexpensive manufacture. The convenience and
Petroleum-based plastics, including poly(ethylene terephthalate) ease have spurred the increased demand for plastics. In 2019, the global
(PET), high-density poly(ethylene) (HDPE), low-density poly(ethylene) production of plastics reached 368 million tons (da Costa, 2021). Plastic
(LDPE), poly(vinyl chloride) (PVC), poly(propylene) (PP), and poly waste can be handled by recycling, incineration, and landfill deposition,

* Corresponding author.
E-mail address: mkoh@korea.ac.kr (M.-K. Oh).

https://doi.org/10.1016/j.biortech.2021.126307
Received 26 September 2021; Received in revised form 4 November 2021; Accepted 5 November 2021
Available online 9 November 2021
0960-8524/© 2021 Elsevier Ltd. All rights reserved.
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

or being leaked. Globally in 2016, 16% of plastic waste was collected for PHA is the manufacturing expense. Approximately 50% of the PHA
recycling, 25% incinerated, 40% reclaimed, and 19% dumped or leaked production cost is spent on raw materials, such as carbon sources, other
unmanageably (Hundertmark et al., 2018). While the incineration of nutrients, and salts (Li & Wilkins, 2020). To reduce this cost, the use of
plastics can be utilized to produce energy, it emits harmful byproducts in cheaper and discarded resources, such as lignocellulose, wastewater, or
the form of carbon dioxide and toxic pollutants such as dioxins, furans, C1 gases, has been explored (Favaro et al., 2019; Morgan-Sagastume
polychlorinated biphenyls and hydrochloric acid (Meereboer et al., et al., 2014; Tamis et al., 2014; Vigneswari et al., 2021). In particular,
2020; Verma et al., 2016). In addition, only half of the incinerated C1 gases can be obtained at essentially no cost from industrial and
plastic can be converted into energy in South Korea (Jang et al., 2020). agricultural sources, and from household waste. Furthermore, since C1
Thus, much of the plastic waste was not recycled. Even worse, plastic gases are the main causes of global warming, their use as microbial
waste that is not disposed of properly can enter watercourses. One feedstocks for PHA production could help alleviate the environmental
consequence is the accumulation of garbage in the ocean, such as the crises of global warming and plastic waste.
Great Pacific garbage patch (Lebreton et al., 2018). These threaten the In this review, recent advances in PHA production strategies from
marine ecosystem and return to humans through the food chain in a carbon dioxide (CO2), carbon monoxide (CO), and methane (CH4) are
microplastic form (Auta et al., 2017). Microplastics can accumulate in discussed. The metabolic characteristics of autotrophs, acetogens, and
the human body, especially in the liver, as demonstrated by in vivo ex­ methanotrophs for the utilization of C1 gases are described. Strategies to
periments (Im et al., 2021; Shi et al., 2021). This has spurred research facilitate PHA production, such as fermentation or genetic engineering,
concerning the use of bioplastics as alternatives to petroleum-based are discussed. These methodologies allow the establishment of a com­
plastics. plete closed-loop PHA production from C1 gases and its degradation to
Bioplastics are plastics that are biologically-based and/or, biode­ C1 gases (Fig. 1).
gradable. Bioplastics having both characteristics include poly­
hydroxyalkanoates (PHAs), poly(lactic acid) (PLA), poly(butylene 2. Metabolic conversion of C1 gases to feedstocks
succinate) (PBS), and starch-based biopolymers (Abioye et al., 2018).
PHA is regarded as a promising alternative to petroleum-based plastics Fossil fuel-based industries have been developed, resulting in air
since it rapidly and completely degrades in terrestrial and marine en­ pollution and global warming. The global average temperature in 2020
vironments. The market size of PHAs in 2020 was estimated to be $62 has increased by 1.3℃ compared to that at pre-industrial temperatures
million USD and is projected to reach $121 million USD by 2025 (Hansena et al., 2021). Various attempts have been made to alleviate
(Market, 2019). Especially, it is noteworthy that PHA production using atmospheric environmental issues. These include CO2 capture, renew­
greenhouse gases (GHGs) is being commercialized by Newlight Tech­ able energy development, transportation improvements, and biorefinery
nologies (Huntington Beach, CA, USA). of C1 gases. Biorefinery using C1 gases as microbial feedstocks for the
PHA is polymerized in vivo using various hydroxyalkanoyl-CoAs. production of value-added compounds has been accomplished using
More than 160 different monomers have been used for PHA polymeri­ autotrophs, acetogens, and methanotrophs. To utilize C1 gases as carbon
zation. The composition of the monomers determines the polymer sources by those strains, stirred-tank reactor (STR) has been commonly
properties. Thus, polymers with desired properties can be produced used with mixed gases (CO2/CO/CH4/H2/N2) injected into the culture
(Bhatia et al., 2021; Choi et al., 2020). A major limitation to the use of broth. Because C1 gases have very low solubility to water, additional

Fig. 1. Closed-loop bioplastic cycle from C1 gases to PHAs.

2
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Fig. 2. Metabolic pathways for C1 gas-utilizing microorganisms. (A) Metabolism of autotrophic bacteria, such as cyanobacteria and the non-photoautotrophic
bacterium, C. necator, which utilize the Calvin-Benson-Bassham (CBB) cycle. (B) Metabolism of autotrophic bacteria, such as A. woodii, which utilize the Wood-
Ljungdahl pathway (WLP). This pathway also fixes CO, as denoted in red letters. (C) Metabolism of type II methanotrophic bacteria, which utilize serine cycle
and EMCP for methane assimilation. Abbreviations; PS (photosystem), 3PGA (3-phosphoglyceric acid), 1,3-BPG (1,3-bisphosphoglyceric acid), G3P (glyceraldehyde
3-phosphate), F1,6P (fructose 1,6-bisphosphate), F6P (fructose 6-phosphate), Ru5P (ribulose 5-phosphate), RuBP (ribulose 1,5-diphosphate), PEP (phosphoenol­
pyruvate), Pyr (pyruvate), PHB (poly-3-hydroxybutyrate), OAA (oxaloacetate), EMCP (ethylmalonyl-CoA pathway).

3
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Fig. 2. (continued).

process modifications have been attempted, such as high rotational C. necator obtains energy for CO2 fixation by oxidizing H2 to two H+
speed of impeller, microbubble sparger, or increased partial pressure of atoms and two electrons catalyzed by several membrane-bound [Ni-Fe]
gases for enhancing gas–liquid mass transfer (Bae et al., 2021). Addi­ hydrogenases (Burgdorf et al., 2005; Schäfer et al., 2013). Electrons are
tional energy sources other than C1 gases, such as light and/or H2 are transported through [Fe-S] clusters in the respiratory chain, and NADPH
provided to the reactor when autotrophs or acetogens are cultured. This or NADH is produced (Panich et al., 2021). Both cyanobacteria and
section mainly discusses the metabolic features of each microorganism C. necator naturally produce PHA.
when utilizing C1 gas. Acetogens are strictly anaerobic bacteria that produce acetate as the
major product when grown on CO2 through the WLP. Clostridium sp. and
Acetobacterium sp. belong to acetogen (Vidra & Németh, 2018). WLP is
2.1. CO2 as a microbial feedstock
the most energy efficient pathway for CO2 fixation (Jiang et al., 2021)
(Fig. 2B). The WLP consists of a methyl branch and a carbonyl branch. In
CO2 is a major contributor to global warming, and the amount of
the methyl branch, CO2 is reduced to formate by formate dehydrogenase
emission is steadily increasing. Several microbial metabolic pathways
(FDH). In turn, formate is converted to methyl-corrinoid iron-sulfur
can fix CO2. These include the Calvin-Benson-Bassham (CBB) cycle,
protein (methyl-CoFeSP) through a series of reactions involving tetra­
Wood-Ljungdahl pathway (WLP), reductive tricarboxylic acid (TCA)
hydrofolate (THF). In the carbonyl branch, CO2 is converted to CO in a
cycle, reductive glycine pathway, dicarboxylate/4-hydroxybutyrate
reaction catalyzed by CO dehydrogenase/acetyl-CoA synthase (CODH/
cycle, 3-hydroxypropionate-4-hydroxybutyrate cycle, and 3-hydroxy­
Acs). The reactions in the two branches convert methyl-CoFeSP and CO
propionate cycle (Jiang et al., 2021). Since PHA-producing autotrophs
are to acetyl-CoA as catalyzed by the bifunctional enzyme CODH/Acs.
usually utilize the CBB cycle or WLP, these pathways are the focus in this
Membrane-bound Rnf complex and flavin-based electron bifurcation
section.
provide NAD(P)H and reduced ferredoxin (Fd2-) as reducing power by
Owing to the stable molecular structure of CO2, autotrophs that
oxidizing H2 when fixing CO2 (Schoelmerich & Müller, 2019; Schuch­
utilize CO2 as a carbon source receive energy from light (photoauto­
mann & Müller, 2014). Instead of hydrogen, electrotrophic metabolism
trophs) or other compounds, such as H2 or formate (lithoautotrophs)
has been verified by supplying electrons through the electrode (Lovley,
(Duerre & Eikmanns, 2015; Gleizer et al., 2019; Lau et al., 2015). Cya­
2011).
nobacteria are typical photoautotrophic bacteria that utilize CBB cycle
(Fig. 2A). These bacteria fix CO2 through the carboxylation reaction
with ribulose 1,5-biosphosphate (RuBP) catalyzed by ribulose 1,5- 2.2. CO as a microbial feedstock
bisphosphate carboxylase/oxygenase (RuBisCO), forming 3-phospho­
glycerate (3-PGA). 3-PGA can be converted to phosphoenolpyruvate An enormous amount of CO is released from industrial processes,
(PEP), which enters the central metabolism and is converted to biomass, such as thermal power generation and steel manufacture. The CO can be
and also be converted to glyceraldehyde 3-phosphate (G3P), which combusted to CO2 for energy generation. However, the efficiency of this
drives the CBB cycle (Humphreys & Minton, 2018; Sarma et al., 2016). process is low due to inert gases. Also, CO can increase the risk of ex­
When light is provided, the energy for the CBB cycle is supplemented by plosion (Choi et al., 2017). Thus, the utilization of CO as a microbial
the oxidation of water catalyzed by photosystem II. Electrons pass feedstock is a sustainable and environmentally-friendly strategy. Recent
through the electron transport chain composed of the cytochrome studies have examined CO utilization for production of biochemicals
complex and photosystems and generate NADPH (Lai & Lan, 2015; (Fernández-Naveira et al., 2016; Moreira et al., 2021; Wilkins & Atiyeh,
Vermaas, 2001). Inorganic compounds, such as H2, can also be utilized 2011). Utilization of CO can damage biological functions that include
as an energy source. A typical CBB cycle-metabolizing non-photoauto­ oxygen transport and is toxic to most organisms (Ernst & Zibrak, 1998).
trophic bacterium is Cupriavidus necator (formerly Ralstonia eutropha). Only some bacteria, such as acetogens, can utilize CO as a carbon source.

4
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

As mentioned above, acetogen metabolism of CO is accomplished PHA synthesis. The operon harbors genes that encode PHA synthase,
through the WLP. While CO can be utilized directly in the carbonyl acetoacetyl-CoA reductase, and 3-ketothiolase.
branch, CO is oxidized to CO2 by CODH in the methyl branch, providing An important advantage of PHAs is that various monomers can be
reducing power (Fig. 2B). Rhodospirillum rubrum is a typical producer of copolymerized with 3-hydroxybutyrate (3HB, C4) with the shortest R
PHA from CO (Do et al., 2007). Also, it is accessible to convert CO into group in its basic structure. Depending on the number of carbon atoms in
soluble substrates, such as acetate or formate, through acetogens for the copolymerized monomer, it comprises short chain length (SCL; C3-
synthesis of PHA (Fei et al., 2021; Hwang et al., 2020). C5) and medium chain length (MCL; C6-C14) PHAs. PHB is the most
well-known and commercially available SCL-PHA. Two acetyl-CoA
2.3. CH4 as a microbial feedstock molecules condense to acetoacetyl-CoA, which is involved in 3-ketothio­
lase encoded by phaA. Acetoacetyl-CoA is then hydrated to 3-hydroxy­
CH4 is a GHG with a global warming potential 25 times greater than butyryl-CoA by NADPH-dependent acetoacetyl-CoA reductase encoded
that of CO2 (Ford et al., 2012). More than 60% of emitted CH4 is from by phaB. 3-Hydroxybutyryl-CoA is polymerized as catalyzed by PHA
anthropogenic sources that include animal husbandry, wastewater synthase encoded by phaC. The phaCAB operon was initially demon­
treatment, natural gas refining, and composting (La et al., 2018). In strated in C. necator (Choi et al., 2020; Schubert et al., 1988). PHB is
addition, natural gas is composed of CH4, which makes it easy to secure arranged in an (R)-configuration with a right-handed double helix,
(Liu et al., 2020). Utilization of CH4 as a microbial feedstock is an resulting in a high degree of crystallinity, stiffness, and melting tem­
optimistic strategy because of its’ pronounced reduction potential, perature (Anjum et al., 2016; D’Amico et al., 2012; Marlina et al., 2018;
which is higher than that of glucose, serving improvement of product Reichardt & Rieger, 2011). To improve the properties of PHB, copoly­
yield (Lee et al., 2016). Utilization of CH4 as microbial feedstock is merization or blends have been attempted. The typical SCL-PHA
available both anaerobically and aerobically by methanotrophs (Duerre copolymer is poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV).
& Eikmanns, 2015). Mostly, in anerobic conditions, enrichment cultures To produce PHBV, precursors that include propionate or valerate are
are utilized, while pure cultures are used in aerobic conditions (Gambelli usually supplemented to support 3-hydroxyvalerate (3HV) (Berezina &
et al., 2018; Zhu et al., 2012). Yada, 2016; Matsumoto et al., 2011). Otherwise, some wild type bac­
Methanotrophs utilize CH4 by oxidizing it to methanol as catalyzed teria including archaea can produce PHBV without any feeding pre­
by methane monooxygenase (MMO). A soluble form (sMMO) and a cursors (Koller, 2019). Genetic engineering for PHBV has also been
particulate form (pMMO) existed (Fig. 2C). Methanol is then converted performed in the absence of precursors (Orita et al., 2014; Yoon et al.,
into formaldehyde by periplasmic pyrroloquinoline quinone (PQQ)- 2021). For example, Methylorubrum extorquens is a model strain of
dependent methanol dehydrogenase. Formaldehyde can be assimilated methylotroph that is capable of serine cycle and EMCP metabolism using
through two pathways that divide methanotrophs into two groups. Type propionyl-CoA as a major metabolite. To utilize propionyl-CoA, which is
I methanotrophs belonging to γ-proteobacteria utilize the RuMP a precursor of 3HV, for copolymerization, the essential gene phaA for
pathway. Type II methanotrophs belonging to α-proteobacteria have a EMCP was deleted and the alternative gene bktB from C. necator was
serine cycle and ethylmalonyl-CoA pathway (EMCP) for formaldehyde introduced (Yoon et al., 2021).
assimilation (Jiang et al., 2021). Because type II methanotrophs natu­ MCL-PHAs are typically produced from Pseudomonas sp. The poly­
rally produce PHA, many studies have focused on the type II microor­ merization of MCL-PHAs is dependent on PHA synthases (Ray & Kalia,
ganisms in the production of PHA from methane. In type II 2017). There are four groups of PHA synthases (Class I–IV) based on
methanotrophs, formaldehyde is passed through the assimilative or their primary structure, substrate specificity, and subunit composition.
dissimilative pathways. It is first oxidized to formate and finally into CO2 Class I has a homodimer structure with a PhaC subunit and is specific to
by FDH, providing NADH for energy. Formaldehyde can also be con­ SCL-hydroxyalkanoyl-CoAs (Sagong et al., 2018). C. necator has a class I
verted into methylene-H4F, which enters the serine cycle in the assim­ PHA synthase and mainly produces PHB and copolyesters with 3HV and
ilative pathway. Serine is then converted into PEP in a series of 4-hydroxybutyrate (4HB). Class II PHA synthase also has a homodimer
sequential reactions and is used for biomass production. Glyoxylate is structure with the PhaC1 or PhaC2 subunits. It has broad substrate
required for the continuous operation of the serine cycle. To regenerate specificity and is capable of polymerizing MCL-hydroxyalkanoyl-CoAs.
glyoxylate, type II methanotrophs use the EMCP anaplerotic reaction P. aeruginosa, P. putida, P. resinovorans, and some other bacteria have
that involves various CoA esters. Methanotrophs can even produce poly Class II PHA synthases. When volatile fatty acids are supplemented, they
(3-hydroxybutyrate) (PHB) through EMCP intermediates, acetoacetyl- produce MCL-PHAs, such as poly(3-hydroxybutyrate-co-3-hydroxyhex­
CoA, and 3-hydroxybutyryl-CoA. anoate) (PHBHHx), or other PHAs that include 3-hydroxyheptanoate
(3HHp), 3-hydroxyoctanoate (3HO), 3-hydroxynonanoate (3HN), or 3-
3. Strategies for production of PHAs from C1 gases hydroxydecanoate (3HD) (Cerrone et al., 2014; Mezzolla et al., 2018).
Class III and IV are heterodimers consisting of PhaC/PhaE and PhaC/
3.1. Biosynthetic pathway of PHAs PhaR, respectively. Both are specific for SCL-hydroxyalkanoyl-CoAs
(Pradani et al., 2020). To produce MCL-PHAs, recombinant strains
PHAs are biodegradable polyesters synthesized inside cells using were constructed by introducing a Class II PHA synthase. For example,
various hydroxyalkanoyl-CoAs as monomers. The monomers determine phaC for Class I PHA synthase of C. necator was replaced by phaC1 and
properties of PHAs, such as stiffness, tensile strength, and melting phaC2 from Rhodococcus aetherivorans. Recombinant C. necator can
temperature. PHAs are polymerized via the PHA synthetic pathway produce PHBHHx from plant oil. Furthermore, phaJ involved in
starting with acetyl-CoA as catalyzed by 3-ketothiolase, acetoacetyl-CoA β-oxidation was overexpressed, resulting in increased 3-hydroxyhexa­
reductase, and PHA synthase (Fig. 2). The major physiological role of noate (3HHx) portion in PHA (Budde et al., 2011).
PHAs is storage of carbon and energy when carbon is in excess and when
other nutrients, such as nitrogen or phosphorus, are limited (Tan et al., 3.2. Strategies for PHA synthesis from CO2
2014). When nutrients are abundant, acetyl-CoA is used to operate the
TCA cycle to obtain energy and grow. Coenzyme A (CoA) produced by 3.2.1. Photoautotrophic synthesis of PHB
the TCA cycle interferes with PHA synthesis. Conversely, when nutrients Diverse cyanobacteria that grow photoautotrophically have been
are lacking, the low CoA level does not interfere with PHA synthesis, and tested for their PHB-producing ability using CO2 and light. Synecho­
acetyl-CoA is converted to PHA. The best-known PHA producer is coccus sp., Nostoc muscorum, Aulosira fertilissima, Scytonema sp., and
C. necator under both autotrophic and heterotrophic conditions others produce PHB within 10% per dry cell weight (DCW) (Bhati et al.,
(Vigneswari et al., 2021). The phaCAB operon in C. necator is involved in 2010; Tarawat et al., 2020) (Table 1). To improve PHB production,

5
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Table 1
Strategies of PHB production from CO2.
Strain Strategies Cultivation Cell PHA PHA PHA Reference
period mass titer content productivity

Photoautotrophic bacteria in autotrophic condition


Anacystis nidulans Cultivation of diverse cyanobacteria under 7 days n.d. n.d. 4.4% (w/ n.d. (Bhati et al., 2010)
photoautotrophy w)
Aulosira fertilissima 14 days n.d. n.d. 6.5% (w/ n.d.
w)
Calothrix sp. 28 days n.d. n.d. 6.8% (w/ n.d.
w)
Nostoc muscorum 21 days n.d. n.d. 8.5% (w/ n.d.
w)
Csytonema sp. 21 days n.d. n.d. 7.4% (w/ n.d.
w)
Tolypothrix distorta TISTR Cultivation of diverse cyanobacteria under 20 days n.d. n.d. n.d. 2.2 mg/ (Tarawat et al.,
8985 photoautotrophy (gDCW∙day) 2020)
Nostoc muscorum TISTR 20 days n.d. n.d. n.d. 0.8 mg/
8871 (gDCW∙day)
Synechocystis sp. PCC Introduction of heterologous phaABEC genes from 8 days 0.153 0.0106 7% (w/ 0.0000552 g/ (Hondo et al., 2015)
6803GT P. aeruginosa and cultivation with nitrogen g/L g/L w) (L∙h)
limitation
Synechocystis sp. PCC Deletion of pta and ach and introduction of xfpK 40 days ~ 2 g/L 0.232 12% (w/ 0.000304 g/ (Carpine et al., 2017)
6803 from Bifidobacterium breve g/L w) (L∙h)
Synechocystis sp. PCC Random mutation using UV lightCultivation 500 h 2.11 g/ 0.78 g/L 37% (w/ 0.00154 g/ (Kamravamanesh
6714 without N and P L w) (L∙h) et al., 2018)
Synechocystis sp. PCC Knock-out of exoD 14 days 2.06 g/ ~ 0.33 ~ 16% n.d. (Mittermair et al.,
6714 mutant Mt_a24 L g/L (w/w) 2021)
S. elongatus UTEX2973 Introduction of heterologous phaCAB genes from 10 days 1.22 g/ 0.278 21.1% 0.00116 g/ (Roh et al., 2021)
C. necator and cultivation with nitrogen limitation L g/L (w/w) (L∙h)
Mixed wastewater-borne Cultivation depending on light cycle and nutrient 8 days n.d. 0.104 5.7% (w/ 0.000542 g/ (Arias et al., 2018)
cyanobacteria deficiency (N/P) g/L w) (L∙h)
Mixed culture of Separation of Bio-module I (CO2 → sucrose) and 16 days n.d. 0.156 n.d. 0.000992 g/ (Löwe et al., 2017)
Synechococcus Bio-module II (sucrose → PHA) g/L (L∙h)
elongatus cscB and
Pseudomonas putida
cscAB
Mixed culture of Separation of Bio-module I (CO2 → sucrose) and Observation n.d. n.d. 31% (w/ 0.00118 g/ (Weiss et al., 2017)
Synechococcus Bio-module II (sucrose → PHB) for 5 months w) (L∙h)
elongatus PCC 7942
and Halomonas
boliviensis
Photoautotrophic bacteria in mixotrophic condition
Synechocystis sp. PCC Deletion of pirC (regulatory protein of intracellular 28 days n.d. n.d. 81% (w/ n.d. (Koch et al., 2020)
6803 glycogen and PHB pool) and introduction of phaAB w)
from C. necatorCultivation in CO2 + light with
supplementation of acetate
Synechocystis sp. PCC Deficiency of P with supplement of acetate and 10 days n.d. n.d. 38% (w/ n.d. (Panda & Mallick,
6803 fructose w) 2007)
Synechocystis sp. PCC Adjustment of nutrient and light 12 days n.d. n.d. 21.8% n.d. (Monshupanee &
6803 intensityCultivation with deficiency of N and P and (w/w) Incharoensakdi,
supplement of glucose 2014)
Scytonema geitleri Adjustment of pH, temperature and carbon sources 28 days n.d. n.d. 7.12% n.d. (Singh et al., 2019)
Bharadwaja such as sucrose and acetate (w/w)
Non-photoautotrophic bacteria in autotrophic condition
C. necator Development of a bench-plant scale, recycled-gas, 40 h 91.3 g/ 61.9 g/L 67.8% 1.55 g/(L∙h) (Tanaka et al., 1995)
closed-circuit culture system by maintaining L (w/w)
oxygen concentration in the gas phase below the
limit for a gas explosion (6.9%)
C. necator Cultivation in air-lift fermenter by addition of 96 h 69.3 g/ 56.4 g/L 81.4% 1.02 g/(L∙h) (Taga et al., 1997)
sodium carboxymethylcellulose to culture medium L (w/w)
C. necator Set up of mathematical model based on mass 120 h 60 g/L 49.2 g/L 82% (w/ 0.41 g/(L∙h) (Mozumder et al.,
balance and cultivation in continuous stirred-tank w) 2015)
reactor (CSTR) with an air-lift fermenter
C. necator H16 Cultivation using low content of H2 and nitrogen 144 h 0.39 g/ 0.27 g/L 70% (w/ 0.00188 g/ (Miyahara et al.,
limitation L w) (L∙h) 2020)
C. eutrophus B-10646 Chemostat batch culture using third stage, which 80 h 45 ± 5 36 g/L 80 ± 5% 0.45 g/(L∙h) (Volova et al., 2013)
stages are depending on amount of nitrogen supply g/L (w/w)
C. necator DSM545 Two-stage cultivation system composed of phase About 63 h 21 g/L 16 g/L 74% (w/ 0.252 g/(L∙h) (Garcia-Gonzalez
for heterotrophic cell mass growth and phase for w) et al., 2015)
autotrophic PHB production below explosion limit
of oxygen concentration 5%(vol)
Ideonella sp. O-1 Gas component: H2:O2:CO2 = 7:1:1Cultivation in 24 h 6.75 g/ 5.26 g/L 77.9% 0.219 g/(L∙h) (Tanaka et al., 2011)
minimal medium containing 1.0 g/L (NH4)2SO4 L (w/w)
P. denitrificans High-cell density fermentation with nitrogen About 132 h About About 4 57.3% About 0.0304 (Tanaka et al., 2016)
NBRC13301 limitation 7 g/L g/L (w/w) g/(L∙h)

6
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

environmental conditions of pH and temperature were adjusted (Singh 3.2.2. Non-photoautotrophic synthesis of PHB from CO2
et al., 2019), as were the intensity and cycle of lighting (Arias et al., Several studies have addressed PHB production from CO2 using non-
2018; Monshupanee & Incharoensakdi, 2014). In addition, because PHB photoautotrophs that use inorganic compounds, such as H2, as an energy
tends to accumulate when nutrients, such as nitrogen (N) and phos­ source. C. necator is a typical chemolithotroph and PHA producer. PHA
phorus (P), are limited, two-stage cultivation methods have often been production by C. necator in both autotrophic and heterotrophic condi­
used. The first stage promoted cell growth with sufficient supplemen­ tions has been studied (Cavalheiro et al., 2009; Dalsasso et al., 2019).
tation of N and P. The second stage induced PHB production in nutrient- Since C. necator also tends to accumulate PHB when mineral elements,
deficient medium. (Arias et al., 2018; Kamravamanesh et al., 2018; such as N or P, are limited, two-stage cultivation methods and further
Miyahara et al., 2020; Roh et al., 2021; Tanaka et al., 2016; Volova et al., three-stage cultivation methods that control the supply of N or P were
2013). suggested to induce PHB production using CO2. For example, a three-
To improve photoautotrophic PHB production, the use of mixotrophs stage chemostat batch culture was set up for C. necator. In the first
or mixed cultures have also been attempted. A mixotrophic culture in­ stage, the amount of N for sufficient growth was 100 mg/g cells. The
volves cultivation with other heterotrophic substrates and autotrophic PHB content per DCW was 8% at that stage. In the second stage, nitrogen
substrates. This strategy can provide sufficient energy and carbon, supplementation was limited to 60 mg/g cells. The content of accumu­
leading to improved cell growth and PHB production. Compared to lated PHB was 47%. Finally, in the third stage, N was not supplied. The
glucose, acetate, and fructose as heterotrophic substrates, acetate is a PHB content reached 80% (Volova et al., 2013). A mathematical model
good nutrient for increase of PHB production (Koch et al., 2020; Mon­ based on mass balance was applied to predict PHB production under
shupanee & Incharoensakdi, 2014; Panda & Mallick, 2007; Singh et al., various conditions, such as oxygen, nitrogen stress, and gas composition
2019). Acetate may be easily converted to acetyl-CoA, which is the (O2/H2) (Mozumder et al., 2015). H2 is an energy source for CO2 fixation
precursor of PHB. A mixed culture of cyanobacteria with other bacteria by oxidization catalyzed by membrane-bound [Ni-Fe] hydrogenases.
is another strategy for PHB production from CO2 (Arias et al., 2018). The However, it has a risk of explosion. To reduce the risk, a low H2 content
cyanobacterium Synechococcus elongatus exports sucrose using fixed CO2 of 3.6% was adopted to produce PHB in C. neactor (Miyahara et al.,
with a high efficiency of up to 85%. Co-culture of heterotrophic bacteria 2020). Similar efforts were made to lower the risk of explosion when
for PHB production using exported sucrose can increase PHB produc­ culturing using O2, another combustible element, by cultivating
tivity. For example, in mixed cultures S. elongatus cscB produced sucrose C. necator with a lowered oxygen concentration in gaseous substrates
from CO2 photoautotrophically and P. putida cscAB produced PHA from below 6.9% (Garcia-Gonzalez et al., 2015; Tanaka et al., 1995).
sucrose. By applying this system, 156 mg/L of PHB was produced (Löwe Chemolithotrophs other than C. necator have also been used for PHA
et al., 2017). Likewise, the co-culture of S. elongatus PCC 7942 and production. Paracoccus denitrificans reportedly produced 57.3% PHB per
Halomonas boliviensis enabled PHB production from CO2 by separating DCW from CO2 and H2 under N-limited conditions (Tanaka et al., 2016).
the reaction of CO2 with sucrose and that of sucrose to PHB (Weiss et al., Ideonella sp. O-1 also produced 77.9% PHB per DCW from CO2 and H2,
2017). and displayed higher tolerance to CO compared to that of C. necator,
Genetic engineering has been used to improve PHB productivity in leading to the expectation that industrial exhaust gas containing CO
cyanobacteria. A typical way to improve PHB production is to introduce would be utilized directly (Tanaka et al., 2011). The strategy of con­
heterologous PHB synthesis pathways to strengthen PHB synthesis verting CO2 into PHB using hydrogen-oxidizing bacteria is one of the
fluxes. Thus, phaCAB and phaABEC from C. necator and Microcystis aer­ most promising and feasible production platforms, given that it achieved
uginosa NIES-843, respectively, were introduced into Synechococcus sp., high concentration of PHB with competitive productivity (Table 1).
resulting in an increase in PHB content (PHB per DCW (w/w)) by 12-fold
compared to the control (Hondo et al., 2015; Roh et al., 2021). Another 3.2.3. Synthesis of PHA copolymers using CO2
example is the enhancement of the acetyl-CoA pool in Synechocystis sp. PHA copolymers have various properties depending on the monomer
PCC 6803, given that acetyl-CoA is a precursor of PHB. For this, phos­ composition. Several strategies, such as feeding precursors or genetic
photransacetylase encoded by pta and acetyl-CoA hydrolase encoded by engineering, have been explored for the production of PHA copolymers
ach were knocked out, and xfpK from Bifidobacterium breve was intro­ using CO2. Table 2 lists native producers of PHA copolymers under
duced (Carpine et al., 2017). Furthermore, the regulatory protein of PHB autotrophic conditions without any supplement of precursors (Table 2).
synthesis was engineered in Synechocystis sp. PCC 6803. PirC is a regu­ Oscillatoria okeni can produce MCL-PHA PHBHHx with 3HHx 5.5%(mol)
lator involved in the conversion of intracellular glycogen and PHB pools in PHA under N-limited conditions (Taepucharoen et al., 2017) and
under nutrient-limiting conditions. When pirC was deleted and phaAB Anabaena spiroides TISTR 8075 produces PHBV (Tarawat et al., 2020). In
from C. necator was introduced, the PHB content per DCW reached 81% addition, PHA copolymers can be produced by adding precursors
with acetate supplement (Koch et al., 2020). In cyanobacteria, over­ (Table 2). γ-Butyrolactone, valerate, and hexanoate are utilized to pro­
expression of a high copy number is achieved via plasmid or integration. duce 4HB, 3HV, and 3HHx, respectively, in C. necator (Park et al., 2014;
However, the approach is not cost-effective in large-scale reactors Volova et al., 2013).
because it requires the addition of antibiotics. To solve this problem, an In the absence of supplementation of precursors, metabolic engi­
antibiotic-free stable expression plasmid system was developed by neering can enable the synthesis of PHA copolymers. In C. necator,
constructing a cyanobacterial lethal recA null mutant and recA- heterologous expression of phaCs from Pseudomonas sp. and supple­
harboring plasmid. Using this system, PHB accumulation was improved mentation with acrylic acid, which inhibits β-oxidation, caused the
by stably expressing phaCAB (Akiyama et al., 2011). production of MCL-PHAs with C4-C6 monomers (Nangle et al., 2020). In
Random mutations can be an effective strategy to improve produc­ addition, overexpression of phaAB and bktB and introduction of phaC1
tivity. Random mutations in Synechocystis sp. PCC 6714 using ultraviolet from Pseudomonas sp. 61–3 yielded high levels of PHA per DCW with
light resulted in PHB accumulation of 37% per DCW (Kamravamanesh monomers, 3HV 1.2%(mol) and 3-hydroxy-4-methylvalerate (3H4MV)
et al., 2018). In this mutant, further rational engineering with the 1.2%(mol) in autotrophic conditions (Miyahara et al., 2020).
deletion of exoD, which is involved in exopolysaccharide (EPS) pro­
duction, reduced EPS production and increased PHB production (Mit­ 3.3. PHA synthesis from CO
termair et al., 2021). Techno-economic analysis for PHB production
using CO2 has not yet been reported. However, the synthesis of cyano­ 3.3.1. Native CO-utilizing bacteria
bacterial PHB is expected to reduce cost by up to approximately 50% Synthetic gas (syngas) is produced through the thermochemical
when compared to the synthesis of heterotrophic bacterial PHB (Costa process of gasification. In this process, the materials are not completely
et al., 2018). oxidized and create gases that contain a large amount of CO (Drzyzga

7
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Table 2
Strategies of PHA copolymers from CO2.
Strain Strategies for PHA copolymer production Substrates PHA composition Reference

Oscillatoria okeni Cultivation with nitrogen limitation CO2 + light PHBV (3HV 5.5% [mol]) (Taepucharoen et al.,
TISTR 8549 2017)
Anabaena spiroides Cultivation with various carbon sources CO2 + light PHBV (3HV 43.2% [mol]) (Tarawat et al.,
TISTR 8075 CO2 + light + acetate PHBV (3HV 24.9% [mol]) 2020)
CO2 + light + PHBV (3HV 17.6% [mol])
propionate
CO2 + light + valerate PHBV (3HV 51.1% [mol])
Supplement of precursors
C. necator ATCC Supplement of valeric acid CO2 + H2 + valerate PHBV (3HV 46% [mol]) (Park et al., 2014)
17,697
C. necator DSM 545 Mixotrophic cultivation and supplement of valeric acid Glucose + CO2 + H2 + PHBV (3HV 50% [mol]) (Ghysels et al., 2018)
valerate
C. eutrophus B-10646 Autotrophic Supplement of PHA precursors such as valerate, CO2 + H2 + valerate PHBV (3HV 85.1% [mol]) (Volova et al., 2013)
hexanoate, and γ-butyrolactone after cell growth CO2 + H2 + hexanoate PHBHHx (3HHx 20.0% [mol])
CO2 + H2 + P(3HB-co-4HB) (4HB 51.3% [mol])
γ-butyrolactone
Metabolic engineering
C. necator H 16 Expression of heterologous phaCs and the addition of acrylic CO2 + H2 MCL-PHAs (4 ≤ chain length ≤ 14) (Nangle et al., 2020)
acid (β-oxidation inhibitor)
C. necator H 16 Introduction of phaC1 from Pseudomonas sp. 61–3 and CO2 + H2 PHA terpolymer with 3HV 1.2% (Miyahara et al.,
overexpression of phaAB and bktB and 3H4MV 1.2% [mol] 2020)

et al., 2015). Gas composition is usually 40% CO, 40% H2, 10% CO2, and grown on CO, CO dehydrogenase and CO-insensitive hydrogenase
10% N2 (Karmann et al., 2019). R. rubrum produces PHB from CO as a convert CO to CO2 and H2. Although R. rubrum harbors a CBB cycle, 13C
carbon and energy source under anaerobic conditions (Table 3). When analysis and gene expression analysis have demonstrated that

Table 3
Strategies for PHA production from CO.
Strain Strategies Cultivation Cell mass PHA titer PHA content PHA Reference
period productivity

Native CO-utilizing bacteria


Rhodospirillum rubrum Fed-batch cultivation with three About 180 h 5.32 g/L PHB 1.6 g/ PHB 30.0% (w/w) About PHB (Karmann
different growth phases under nutrient L 0.0089 g/(L∙h) et al., 2019)
(C, P)-limited conditions with
supplement of acetate
Rhodospirillum rubrum Optimization of culture method with 5 days 0.45 g/L PHB 0.036 PHB 8% (w/w) PHB 0.0003 g/ (Mongili &
supplement of acetate and fructose g/L (L∙h) Fino, 2021)
Rhodospirillum rubrum Cultivation with supplement of acetate 5 days n.d. n.d. PHB 28% (w/w) n.d. (Revelles
and analysis of central carbon et al., 2016)
metabolites using 13C-labelling acetate
in co-substrates condition
Rhodospirillum rubrum Optimization of gasifier and supplement n.d. 0.00708 n.d. PHBV 38% (w/w) with PHBV 0.00247 (Do et al.,
of acetate g/(L∙h) monomer composition of g/(L∙h) 2007)
86%(mol) 3HB and 14%
(mol) 3HV
Rhodospirillum rubrum Deletion of endogenous phaC1 and 5 days 1 g/L P (3HD-co- P (3HD-co-3HO) 6.7% P (3HD-co-3HO) (Heinrich
phaC2 and introduction of phaC1, phaG, 3HO) (w/w) with monomer 0.000558 g/ et al., 2016)
and gene number PP_0763 encoding 0.067 g/L composition of 42.8% (L∙h)
CoA ligase from P. putida (mol) 3HO and 57.0%
(mol) 3HD
Seliberia Optimization of CO concentration and 56 h 20.2 g/L 12.3 g/L PHB 61–63% (w/w) 0.22 g/(L∙h) (Volova
carboxydohydrogena limitation of N and sulfur et al., 2015)
Z-1062
Clostridium Construction of PHB-producing 24 h 0.482 g/L PHB PHB 5.61% (w/w) PHB 0.00113 g/ (Lemgruber
autoethanogenum recombinatn by introducing phaCAB 0.02704 g/ (L∙h) et al., 2019)
from C. necator and comparison of PHB L
content depending on H2 supply
Clostridium coskatii Construction of PHB-producing 48 h OD 0.43 n.d. PHB 1.12% (w/w) n.d. (Flüchter
recombinant by introducing thlA, hbd, et al., 2019)
crt, phaJ and phaEC from Clostridium
species
Non-native CO-utilizing bacteria
C. necator Introduction of cox subcluster (genes for 21 days 2.62 g/L PHB 1.30 PHB 49.7% (w/w) PHB 0.00258 g/ (Heinrich
carbon monoxide dehydrogenase and g/L (L∙h) et al., 2018)
the maturation of CODH) from
Oligotropha carboxidovorans
C. necator In vitro conversion of CO into CO2 72 h 33.8 g/L PHB 14.2 PHB 42% (w/w) 0.197 g/(L∙h) (Shin et al.,
through nanoscaled cellulose particles g/L 2021)
with enzyme complex (carbon
monoxide dehydrogenase, carbon
monoxide binding unit, and carbonic
anhydrase)

8
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

carboxylases other than RuBisCO incorporate CO2 from syngas to


biomass (Revelles et al., 2017; Revelles et al., 2016). Acetate supple­

(Lagoa-Costa
et al., 2017)

et al., 2020)

et al., 2018)
(Al Rowaihi
Reference

(Fei et al.,
mentation can permit sufficient production of PHB from CO using

(Hwang
R. rubrum (Do et al., 2007; Mongili & Fino, 2021). Limiting nutrients,

2021)
such as P, can lead to a PHB content per DCW of up to 30% (Karmann
et al., 2019). Another study engineered R. rubrum to produce MCL-PHA

6.5% (w/w)PHB 1.57 g/L with

PHB 0.5 g/L with content per


3HB 1.02 g/L with yield 0.26
copolymers from CO. To incorporate MCL monomers, endogenous

content 46.6% (w/w) when


PHB content per DCW 24%

PHB 0.1 g/L with content


phaC1 and phaC2 were knocked out, phaG (3-hydroxyacyl carrier pro­

using purchased formate


tein thioesterase), phaC1 (MCL-PHA synthase), and gene number

DCW 33.3% (w/w)


PP_0763 (MCL-fatty acid coenzyme A ligase) from P. putida were
expressed under the control of a CO-inducible promoter. This recombi­
nant R. rubrum accumulated up to 7.1% PHA per DCW with a monomer

Product
composition of 42.8%(mol) 3HO and 57.0%(mol) 3HD (Heinrich et al.,

(w/w)

g/g
2016). Seliberia carboxydohydrogena Z-1062 also produces PHB from CO.
Adjustment of the CO concentration and limiting N and sulfur reportedly
led to a PHB content of 62.8% with a productivity rate of 0.22 g/(L∙h)

Cultivation using acetate solution from Process I


Fed-batch cultivation inducing consumption of

expressing phaAB from C. necator and pct from


acetate and ethanol in N limitation condition
(Volova et al., 2015).

Clostridium beijerinckiiCultivation in complex

productionFed-batch cultivation in minimal

with supplement of fructose 30 g/L for cell


Introduction of 3HB synthesis pathway by
Engineering of Clostridium sp. to produce PHB is an alternative

Overexpression of ftfL and phaCAB for


improvements of cell growth and PHB
strategy to utilize CO (Table 3). Clostridium species can grow on CO
through the WLP. However, the bacteria are non-native producers of
PHB. Therefore, some Clostridium species with PHB production capacity

medium with yeast extract


were constructed by heterologous expression of the PHB synthesis
pathway from C. necator and Pseudomonas sp. (Flüchter et al., 2019;
Lemgruber et al., 2019). In particular, metabolomics, transcriptomics,
and genome-scale metabolic models have shown that cellular redox
states, such as ATP availability, limit PHB production, rather than the

Strategies

medium
acetyl-CoA pool, in C. autoethanogenum capable of producing PHB pro­

harvest
duction (Lemgruber et al., 2019).
A techno-economic cost analysis was conducted to determine
whether the production of PHB by R. rubrum using syngas is economi­

Engineered E. coli
cally viable. The estimated cost was $1.65/kg (Choi et al., 2010), which

Methylorubrum
Enriched PHA
accumulating
is economically feasible. Further research on the utilization of syngas is

consortium

Engineered
Process II

extorquens

C. necator
anticipated.
Strain

3.3.2. Synthetic CO-utilizing bacteria


Synthetic CO-utilizing C. necator was developed for PHB production
from syngas (Table 3). CO oxidation was facilitated by expressing the

Formate 53.1 mM
LEthanol 2.22 g/
Acetate 4.976 g/

cox subcluster consisting of CODH and accessory proteins for CODH


L2,3-butanediol

Acetate 3.2 g/L


Acetate 4 g/L

maturation from Oligotropha carboxidovorans. This allowed the conver­


1.632 g/L

sion of CO to CO2 and growth in the presence of syngas. Engineered


Product

C. necator produced up to 49.7% PHB per DCW (Heinrich et al., 2018). In


another study, nanoscale cellulose particles which were attached with
CODH, carbon monoxide binding unit, and carbonic anhydrase were
conditions such as pH, CO concentrations and

Cultivation of gas mixture of CO2:H2 = 15:85


(acetate production) and solventogenic phase
Consisting of 2 phases with acidogenic phase

immobilized on the surface of C. necator. The complex catalyzed the


under elevated pressure for H2 solubility
(ethanol and 2,3-butanediol production)

conversion of CO to CO2 in vitro. C. necator utilizing CO2 as carbon


biocatalystOptimization of catalytic

source produced 14.2 g/L PHB with PHB content of 42% per DCW (Shin
Conversion process with whole-cell

et al., 2021). The findings indicated the potential for PHB production
Combined processes for Syngas into PHAs through intermediates

from syngas.
Cultivation using syngas

3.3.3. Process development of syngas for PHA production: Combining two


bioconversion processes
Among syngas components, CO is toxic and the production of PHA
Strategies

from CO is usually poor. Processes combining the two bioconversion


buffer

processes have been developed to overcome these limitations. Inte­


grated bioconversion processes consisting of Process I (CO → organic
substrate) and Process II (organic substrate → PHAs) are effective in
terms of CO fixation, H2 consumption, substrate cost, and safety (Garcia-
autoethanogenum

Gonzalez & De Wever, 2018) (Table 4). For conversion of syngas in


Acetobacterium

Process I, Clostridium species were cultivated to produce acetate, which


Clostridium

Clostridium

biocatalyst
Whole cell
liungdahlii
Process I

A. woodii

was used to synthesize PHB or monomer 3HB through a PHB-


Strain

woodii

accumulating consortium or engineered Escherichia coli, respectively,


in Process II (Fei et al., 2021; Lagoa-Costa et al., 2017). Furthermore,
combined processes using formate as an intermediate were developed.
Formate
Syngas →

Syngas →

Syngas →
Acetate

Acetate

Acetate
→ PHB

→ PHB

→ PHB
→ 3HB

In Process I, syngas was converted into formate using a whole-cell bio­


product
Table 4

CO2 →
Final

catalyst, and in Process II, engineered M. extorquens produced PHB


(Hwang et al., 2020). Similarly, two bioconversion processes were

9
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

applied to convert CO2 to PHB through acetate (Table 4). A. woodii and sludge reportedly produced PHB from waste CH4, with the accumulation
C. necator were cultivated to convert CO2 into acetate and convert ac­ of 47.8–51% (w/w) PHB per DCW (Chidambarampadmavathy et al.,
etate into PHB (Al Rowaihi et al., 2018). 2015; Fergala et al., 2018). While most methanotrophs prefer low
temperatures (<30 ◦ C), thermophilic methanotrophs produce PHBs at
55–58 ◦ C. For example, cultivation of methanotrophic consortia domi­
3.4. PHA synthesis from CH4
nated by thermophilic methanotroph Methylcaldum szegediense OR2
resulted PHB production (Luangthongkam et al., 2019a). When using
3.4.1. Enrichment culture to utilize waste CH4 and produce PHB
methanotrophic enrichment culture, a two-stage cultivation method
Enrichment cultures provide diverse metabolic activities and make
with growth phase (sufficient nutrients) and PHB production phase
particular microorganisms to favor growth in certain surroundings. Such
(nutrient limitation) was also used to induce PHB accumulation in cells
cultures are usually used for biorefining of waste CH4 emitted from
(Helm et al., 2008). In addition to the limited supply of N, P, sulfur, or
landfills and in anaerobic digestion processes (Table 5). Methanotrophic
iron, diverse factors, including CH4/O2 ratio, CH4 mass transfer,
consortia from landfill biomass, compost soils, or anaerobic digester

Table 5
Strategies for PHB production from CH4.
Strain Strategies Cultivation Cell mass PHB PHB PHB Reference
period titer content productivity

Enriched culture
Methylocystis sp.- Two-stage cultivation method separating 48 h n.d. n.d. 33.6% n.d. (Helm et al., 2008))
dominated growth phase and PHB accumulation (w/w)
methanotrophic phase with potassium deficiency
enrichment
Methylocystis-dominated Influence of temperature during culture 19 days Specific n.d. 35.1% n.d. (Pérez et al., 2019)
methanotrophic enrichment and cultivation with nitrogen growth (w/w)
enrichment limitation rate 0.05
h− 1
Methylocystis, Methyldopa, Effect of CH4/O2 ratios on PHB n.d. About 0.8 About 8.5% n.d. (Karthikeyan et al., 2015)
and production g/L 0.068 g/ (w/w)
Pseudoxanthomonas- L
dominated consortium
Methylocystis-dominated Demonstration of stability for 29 months 29 months n.d. n.d. 46.2% 1.13 g/(L∙h) (Helm et al., 2006)
enriched culture in an open system with methane (w/w)
Methylophilus, New setup for PHB production by n.d. n.d. n.d. 59.4% n.d. (Salem et al., 2021)
Methylocystis- recycling the biomass after PHB phase (w/w)
dominated enriched
culture
Methanotrophic Comparison of PHB capacity between 10 days OD 0.316 n.d. 47.88% n.d. (Chidambarampadmavathy
consortium from two methanotrophic consortium (w/w) et al., 2015)
landfill top-cover and
compost soils
Methanotrophs Development of accelerating CH4 mass 7 days OD 0.7 n.d. PHB n.d. (Myung et al., 2016a)
transfer without agitation (emulsion- 32% (w/
based fermentation) w)
Methanotrophic Production from anaerobic digester 70 days Specific n.d. 51% 0.0095 g/ (Fergala et al., 2018)
enrichment culture sludge growth (L∙h)
rate 0.078
h− 1
Methanotrophic- Optimization of Cu concentration and 72 h n.d. n.d. 48.7% n.d. (Zhang et al., 2018)
heterotrophic culture CH4:O2 ratio and cultures derived from (w/w)
seed sludge after about 320 days of
enrichment
Thermophilic Optimization of temperature (55℃ or n.d. Specific n.d. 2.4% n.d. (Luangthongkam et al.,
methanotrophic 58℃) and supplement of acetate growth (w/w) 2019a)
enrichment rate
1
0.0104 h−
Pure culture
Methylosinus trichosporium Optimization of medium, air inlet, and 120 h 18 g/L 9 g/L 50% (w/ 0.075 g/(L∙h) (Shah et al., 1996)
OB3b agitation speed w)
Methylosinus trichosporium Optimization of carbon source (methane n.d. 0.0928 g/L 0.0487 52.5% n.d. (Zaldívar Carrillo et al.,
OB3b or methanol), nitrogen source g/L (w/w) 2018)
(ammonium or nitrate), and nitrogen-to-
carbon ratioCultivation in mixotrophic
condition with methanol
Methylocystis sp. GB 25 Separation of growth phase and PHB 24 h n.d. n.d. 51% n.d. (Wendlandt et al., 2001)
DSMZ 7674 production phase
Methylocystis hirsuta Effect of magnesium and 144 h 8.0 g/L 5.87 g/L 73.4% 0.0408 g/ (Ghoddosi et al., 2019)
phosphorusCultivation in mixotrophic (L∙h)
condition with methanol and ethanol
Methylocystis hirsuta CSC1 Effect of O2:CH4 ratio, temperature, and 16 days n.d. n.d. 45% (w/ n.d. (Rodríguez et al., 2020a)
nitrogen source w)
Methylocystis hirsuta Application in a novel gas-recycling About 68 About 5 g/ n.d. 34.6% n.d. (García-Pérez et al., 2018)
bubble column bioreactor with limitation days L (w/w)
of potassium, manganese, and nitrogen
Methylocystis hirsuta Optimal operation of bubble column 65 days 3.0 g/L n.d. 14.5% n.d. (Rodríguez et al., 2020b)
bioreactor (w/w)

10
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

temperature, and Cu concentration, have been optimized (Karthikeyan estimated cost of $7.5/kg, the estimated costs were lower. This is
et al., 2015; Myung et al., 2016b; Pérez et al., 2019; Zhang et al., 2018). because the cost of raw material was estimated to be only 22% when
In particular, novel methods for PHB production have been developed. using CH4, while it was approximately 50% when using sugars (Levett
For example, water-in-oil emulsions were used to stimulate CH4 mass et al., 2016).
transfer without agitation, given that the solubility of CH4 is higher in oil
than in water (Myung et al., 2016b). Further, a new setup for meth­ 3.4.2. Pure culture of methanotrophs for PHB production
anotrophic enrichment culture for higher PHB production was estab­ Pure cultures of type II methanotrophs can be used because they are
lished by recycling the biomass after the PHB phase. In this system, the native PHB producers via the serine cycle and EMCP (Table 5). Similar to
PHB content reached 59.4% DCW (Salem et al., 2021). This methano­ other PHB production strategies, two-stage cultivation methods sepa­
trophic enrichment was operated stably in an open system for 29 rating the growth phase and PHB production phase have been applied to
months, with repeated multiple continuous growth and polymer for­ methanotrophs (Wendlandt et al., 2001). Optimization of the medium,
mation processes. The productivity of PHB from CH4 was 1.13 g/(L∙h) air supply, agitation, carbon source, nitrogen ratio, O2:CH4 ratio, and
(Helm et al., 2006). temperature was accomplished for cell growth and PHB production in
A techno-economic evaluation estimated the cost of PHB production Methylocystis hirsuta or Methylosinus trichosporium (Ghoddosi et al.,
from methane as $4.1–6.8/kg. Furthermore, thermophilic methano­ 2019; Rodríguez et al., 2020a; Shah et al., 1996; Zaldívar Carrillo et al.,
trophs could reduce costs by $3.2–5.4/kg, because refrigerants were not 2018). In addition, mixotrophic conditions by supplementing with
required. Compared to sugar-based PHB production processes, with an methanol or ethanol were applied to increase PHB production (Ghoddosi

Table 6
Strategies for PHA copolymers production from CH4.
Strain Strategies for PHA copolymer production Substrates PHA content and its Reference
composition

Supplement of precursors
Methylosinus-dominated Supplement of propionic acid and valeric acid CH4 + Propionic PHBV 3.5% (w/w) (3HV (Luangthongkam
methanotrophic acid 22.6% (mol)) et al., 2019b)
enrichment CH4 + Valeric acid PHBV 14.1% (w/w) (3HV
65.0% (mol))
Methanotrophic Supplement of valeric acid CH4 + Valeric acid PHBV 52% (w/w) (3HV 33% (Fergala et al., 2018)
enrichment culture (mol))
Thermophilic Supplement of propionic acid and valeric acid CH4 + Propionic PHBV about 3% (w/w) (3HV (Luangthongkam
methanotrophic acid 85% (mol)) et al., 2019a)
enrichment CH4 + Valeric acid PHBV about 10% (w/w) (3HV
75%(mol))
Methylocystis parvus OBBP Supplement of propionic acid and valeric acid CH4 PHB 50% (w/w) (Myung et al.,
CH4 + Propionic PHBV 32% (w/w) (3HV 8% 2016a)
acid (mol))
CH4 + Valeric acid PHBV 54% (w/w) (3HV 22%
(mol))
Methylocystis parvus OBBP Supplement of ω-hydroxyalkanoate monomers CH4 PHB 42% (w/w) (Myung et al., 2017)
CH4 + Butyrate PHB 55% (w/w)
CH4 + 3- PHB 59% (w/w)
hydroxybutyrate
CH4 + 4- P(3HB-co-4HB) 50% (w/w)
hydroxybutyrate (4HB 9.5% (mol))
CH4 + Valerate PHBV 54% (w/w) (3HV 25%
(mol))
CH4 + 5- P(3HB-co-3HV-co-5HV) 48%
hydroxyvalerate (w/w) (3HV 1.4% (mol) and
5HV 3.6% (mol))
CH4 + Hexanoate PHB 56% (w/w)
CH4 + 6- P(3HB-co-4HB-co-6HHx) 48%
hydoxyhexanoate (w/w) (4HB 1% (mol) and
6HHx 1.4% (mol))
Methylocystis sp. WRRC1 Supplement of valerate or n-pentanol CH4 PHB 30% (w/w) (Cal et al., 2016)
CH4 + Valerate PHBV 60% (w/w) (3HV 50%
(mol))
Methylocystis hirsuta Supplement of versatile fatty acids Biogas PHB 43.1% (w/w) (López et al., 2018)
Biogas + Acetic acid PHB 52.3% (w/w)
Biogas + Propionic PHBV 47.9% (w/w) (3HV 2%
acid (mol))
Biogas + Butyric PHB 52.2% (w/w)
acid
Biogas + Valeric PHBV 53.8% (w/w) (3HV 25%
acid (mol))
Methylobacterium Supplement of citrate or propionate CH4 + Citrate P(3HB-co-3HV-co-3HO) (3HV (Zuñiga et al., 2013)
organophilum CZ-2 35% (mol) and 3HO 10%
(mol))
CH4 + Propionate PHBV (3HV 25% (mol))
Metabolic engineering
Methylosinus trichosporium Introduction of 4HB synthesis pathway by expressing sucD (CoA- CH4 P(3HB-co-4HB) 7.01% (w/w) (Nguyen & Lee,
OB3b dependent succinate semialdehyde dehydrogenase), 4hbD (4- (4HB 3.08% (mol)) 2021)
hydroxybutyrate dehydrogenase) from Porphyromonas gingivalis
or yqhD (NADPH-dependent succinate semialdehyde reductase)
from E. coliTwo-stage cultivation without N deficiency

11
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

et al., 2019; Zaldívar Carrillo et al., 2018). Furthermore, for M. hirsuta, a PHA copolymer production having better physical and chemical prop­
mechanistic model was applied to determine the optimal O2:CH4 ratio erties through metabolic engineering would be one of the important
(Chen et al., 2020), and a novel gas-recycling bubble column reactor was points for commercializing PHA copolymer produced from C1 gases.
developed and optimized (García-Pérez et al., 2018; Rodríguez et al., Overall, PHA production using C1 gas is influential and significant in
2020b). terms of providing the closed-loop bioplastic system, where produced
bioplastics are degraded to produce new bioplastics. When PHAs are
3.4.3. Synthesis of PHA copolymers from CH4 processed through aerobic biodegradation or incinerated after use, CO2
The simplest way to produce PHA copolymers is to add related fatty or CO is generated. When the bioplastics are anaerobically digested, CH4
acids, such as propionate or valerate (Table 6). PHBV was synthesized is produced. These gases can be used to produce PHAs using the strains
when propionate or valerate was added to both enrichment cultures and reviewed in this manuscript. These recycling system of PHAs can be
pure cultures (Cal et al., 2016; Fergala et al., 2018; López et al., 2018; sustainable and environmentally benign. It is certain that improvement
Luangthongkam et al., 2019a; Luangthongkam et al., 2019b; Myung of PHA productivity and diversification of PHA species using C1 gas
et al., 2016a; Nguyen & Lee, 2021). Supplementation of 5-hydroxyvaler­ would make it a competitive platform.
ate (5HV) and 6-hydroxyhexanoate (6HHx) resulted in the synthesis of
PHA terpolymers P(3HB-co-3HV-co-5HV) and P(3HB-co-4HB-co-6HHx), 5. Conclusion
respectively, in Methylocystis parvus OBBP (Myung et al., 2017). Further
supplementation with citrate resulted in P(3HB-co-3HV-co-3HO) in PHAs are attractive alternatives to conventional plastics since they
Methylobacterium organophilum (Zuñiga et al., 2013). are biodegradable. However, manufacturing costs are high. Reducing
Recently, metabolic engineering of M. trichosporium OB3b success­ these costs by the use of inexpensive raw materials is desirable. C1 gases,
fully synthesized P(3HB-co-4HB) using CH4 as the sole carbon source. such as CO2, CO, and CH4, are cheap or cost-free substrates, as they can
For the intracellular synthesis of 4HB, sucD encoding CoA-dependent be obtained from industrial or agricultural processes, landfills, and
succinate semialdehyde dehydrogenase and 4hbD encoding 4-hydroxy­ household waste. This review introduces the strategies for PHA pro­
butyrate dehydrogenase from Porphyromonas gingivalis were intro­ duction using C1 gases, including cultivation optimization, metabolic
duced. Furthermore, to increase the succinyl-CoA pool, isocitrate engineering, and development of novel processes. As PHA production
dehydrogenase encoded by icd was overexpressed, resulting in the processes using GHGs are being commercialized, this review will inform
synthesis of P(3HB-co-4HB) with a 4HB portion of 3.08% (mol) using the recent progress of the PHA production platforms.
CH4 (Nguyen & Lee, 2021).
CRediT authorship contribution statement
4. Future perspectives
Jihee Yoon: Conceptualization, Formal analysis, Investigation, Data
C1 gases are the main culprits of global warming. Recently, to ach­ curation, Writing – original draft, Writing – review & editing. Min-Kyu
ieve carbon neutrality, various bioproducts involving biopolymers, Oh: Conceptualization, Writing – original draft, Writing – review &
biochemicals, and biofuels have been produced from C1 gases. Espe­ editing, Supervision, Project administration, Funding acquisition.
cially, PHAs, biodegradable polyesters, are promising products due to
their complete biodegradability in compost and marine conditions, of­
fering the way to solve the problem of petroleum-based plastics. Many Declaration of Competing Interest
strategies for utilizing CO2, CO, and CH4 have been carried out to
improve PHA production. Along with optimization of culture conditions, The authors declare that they have no known competing financial
metabolic engineering for upregulation of PHA synthesis pathway interests or personal relationships that could have appeared to influence
contributed to the increase in PHA productivity. Also, it facilitated the work reported in this paper.
synthesis of various PHA copolymers such as PHBV, P(3HB-co-4HB),
PHBHHx, and MCL-PHAs. However, there are still some drawbacks. i) Acknowledgements
PHA content per DCW is usually low. When growing on C1 gases,
assimilative fluxes are accelerated, which means that acetyl-CoA, a This research was supported by C1 Gas Refinery Program through
precursor of PHA, is mainly used for growth rather than production of the National Research Foundation of Korea (NRF) funded by the Min­
PHA. Given that mixotrophic culture with acetate supplement serves for istry of Science and ICT (2015M3D3A1A01064919).
improvement of PHA content per DCW, genetic engineering for increase
of intracellular acetyl-CoA pool can serve the solution. Adaptive labo­ References
ratory evolution or random mutation might be another way to improve
PHA content, as high-throughput screening using fluorescent dye as Abioye, O.P., Abioye, A.A., Afolalu, S.A., Ongbali, S.O., 2018. A review of biodegradable
BODIPY or Nile red is available. ii) Although growth has been improved plastics in Nigeria. International Journal of Mechanical Engineering and Technology
(IJMET) 9 (10).
through many studies on culture optimization, there is still a limit to Akiyama, H., Okuhata, H., Onizuka, T., Kanai, S., Hirano, M., Tanaka, S., Sasaki, K.,
high cell density. This may be improved through mixotrophic cultiva­ Miyasaka, H., 2011. Antibiotics-free stable polyhydroxyalkanoate (PHA) production
tion with balance of carbon sources. Further, as gene manipulation from carbon dioxide by recombinant cyanobacteria. Bioresource technology 102
(23), 11039–11042.
techniques and synthetic biology tools have been improved, strain Al Rowaihi, I.S., Kick, B., Grötzinger, S.W., Burger, C., Karan, R., Weuster-Botz, D.,
development is being performed more efficiently and systematically. Eppinger, J., Arold, S.T., 2018. A two-stage biological gas to liquid transfer process
Therefore, engineering to improve the assimilative carbon flux from C1 to convert carbon dioxide into bioplastic. Bioresource Technology Reports 1, 61–68.
Anjum, A., Zuber, M., Zia, K.M., Noreen, A., Anjum, M.N., Tabasum, S., 2016. Microbial
gases and increase PHA production would be accelerated. iii) It is production of polyhydroxyalkanoates (PHAs) and its copolymers: a review of recent
necessary to produce various PHA copolymers because the physical advancements. International journal of biological macromolecules 89, 161–174.
properties of PHB, such as stiffness and high melting point, are not Arias, D.M., Uggetti, E., García-Galán, M.J., García, J., 2018. Production of
polyhydroxybutyrates and carbohydrates in a mixed cyanobacterial culture: effect of
proper as commercial plastics. So far, most studies depend on supple­
nutrients limitation and photoperiods. New biotechnology 42, 1–11.
ment of precursors when producing PHA copolymers, which is not cost- Auta, H.S., Emenike, C., Fauziah, S., 2017. Distribution and importance of microplastics
effective. With the developed genetic engineering technologies, such as in the marine environment: a review of the sources, fate, effects, and potential
CRISPR-Cas system, it potentializes the production of non-natural solutions. Environment international 102, 165–176.
Bae, J., Song, Y., Lee, H., Shin, J., Jin, S., Kang, S., Cho, B.-K., 2021. Valorization of C1
polyesters, which contain monomers such as phenyllactate, glycolate, gases to value-added chemicals using acetogenic biocatalysts. Chemical Engineering
or 2-hydroxyisovalerate, as well as SCL-, and MCL-PHAs. Studies for Journal 131325.

12
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Berezina, N., Yada, B., 2016. Improvement of the poly (3-hydroxybutyrate-co-3- hydroxybutyrate from synthesis gas by autotrophic clostridia. Biomacromolecules 20
hydroxyvalerate) (PHBV) production by dual feeding with levulinic acid and sodium (9), 3271–3282.
propionate in Cupriavidus necator. New biotechnology 33 (1), 231–236. Ford, H., Garbutt, A., Jones, L., Jones, D.L., 2012. Methane, carbon dioxide and nitrous
Bhati, R., Samantaray, S., Sharma, L., Mallick, N., 2010. Poly-β-hydroxybutyrate oxide fluxes from a temperate salt marsh: Grazing management does not alter Global
accumulation in cyanobacteria under photoautotrophy. Biotechnology journal 5 Warming Potential. Estuarine, Coastal and Shelf Science 113, 182–191.
(11), 1181–1185. Gambelli, L., Guerrero-Cruz, S., Mesman, R.J., Cremers, G., Jetten, M.S., Op den
Bhatia, S.K., Otari, S.V., Jeon, J.-M., Gurav, R., Choi, Y.-K., Bhatia, R.K., Pugazhendhi, A., Camp, H.J., Kartal, B., Lueke, C., van Niftrik, L., 2018. Community composition and
Kumar, V., Banu, J.R., Yoon, J.-J., 2021. Biowaste-to-bioplastic ultrastructure of a nitrate-dependent anaerobic methane-oxidizing enrichment
(polyhydroxyalkanoates): Conversion technologies, strategies, challenges, and culture. Applied and environmental microbiology 84 (3), e02186–17.
perspective. Bioresource Technology 124733. Garcia-Gonzalez, L., De Wever, H., 2018. Acetic acid as an indirect sink of CO2 for the
Budde, C.F., Riedel, S.L., Willis, L.B., Rha, C., Sinskey, A.J., 2011. Production of poly (3- synthesis of polyhydroxyalkanoates (PHA): Comparison with PHA production
hydroxybutyrate-co-3-hydroxyhexanoate) from plant oil by engineered Ralstonia processes directly using CO2 as feedstock. Applied Sciences 8 (9), 1416.
eutropha strains. Applied and environmental microbiology 77 (9), 2847–2854. Garcia-Gonzalez, L., Mozumder, M.S.I., Dubreuil, M., Volcke, E.I., De Wever, H., 2015.
Burgdorf, T., Lenz, O., Buhrke, T., Van Der Linden, E., Jones, A.K., Albracht, S.P., Sustainable autotrophic production of polyhydroxybutyrate (PHB) from CO2 using a
Friedrich, B., 2005. [NiFe]-hydrogenases of Ralstonia eutropha H16: modular two-stage cultivation system. Catalysis Today 257, 237–245.
enzymes for oxygen-tolerant biological hydrogen oxidation. Journal of molecular García-Pérez, T., López, J.C., Passos, F., Lebrero, R., Revah, S., Muñoz, R., 2018.
microbiology and biotechnology 10 (2–4), 181–196. Simultaneous methane abatement and PHB production by Methylocystis hirsuta in a
Cal, A.J., Sikkema, W.D., Ponce, M.I., Franqui-Villanueva, D., Riiff, T.J., Orts, W.J., novel gas-recycling bubble column bioreactor. Chemical Engineering Journal 334,
Pieja, A.J., Lee, C.C., 2016. Methanotrophic production of polyhydroxybutyrate-co- 691–697.
hydroxyvalerate with high hydroxyvalerate content. International journal of Ghoddosi, F., Golzar, H., Yazdian, F., Khosravi-Darani, K., Vasheghani-Farahani, E.,
biological macromolecules 87, 302–307. 2019. Effect of carbon sources for PHB production in bubble column bioreactor:
Carpine, R., Du, W., Olivieri, G., Pollio, A., Hellingwerf, K.J., Marzocchella, A., dos Emphasis on improvement of methane uptake. Journal Of Environmental Chemical
Santos, F.B., 2017. Genetic engineering of Synechocystis sp. PCC6803 for poly- Engineering 7 (2), 102978.
β-hydroxybutyrate overproduction. Algal research 25, 117–127. Ghysels, S., Mozumder, M.S.I., De Wever, H., Volcke, E.I., Garcia-Gonzalez, L., 2018.
Cavalheiro, J.M., de Almeida, M.C.M., Grandfils, C., Da Fonseca, M., 2009. Poly (3- Targeted poly (3-hydroxybutyrate-co-3-hydroxyvalerate) bioplastic production from
hydroxybutyrate) production by Cupriavidus necator using waste glycerol. Process carbon dioxide. Bioresource technology 249, 858–868.
biochemistry 44 (5), 509–515. Gleizer, S., Ben-Nissan, R., Bar-On, Y.M., Antonovsky, N., Noor, E., Zohar, Y., Jona, G.,
Cerrone, F., Choudhari, S.K., Davis, R., Cysneiros, D., O’Flaherty, V., Duane, G., Krieger, E., Shamshoum, M., Bar-Even, A., 2019. Conversion of Escherichia coli to
Casey, E., Guzik, M.W., Kenny, S.T., Babu, R.P., 2014. Medium chain length generate all biomass carbon from CO2. Cell 179 (6).
polyhydroxyalkanoate (mcl-PHA) production from volatile fatty acids derived from Hansena, J., Satoa, M., Ruedyb, R., Schmidtc, G., Lob, K. 2021. Global Temperature in
the anaerobic digestion of grass. Applied microbiology and biotechnology 98 (2), 2020. columbia.edu.
611–620. Heinrich, D., Raberg, M., Fricke, P., Kenny, S.T., Morales-Gamez, L., Babu, R.P.,
Chen, X., Rodríguez, Y., López, J.C., Muñoz, R., Ni, B.-J., Sin, G.r. 2020. Modeling of O’Connor, K.E., Steinbüchel, A., 2016. Synthesis gas (Syngas)-derived medium-
Polyhydroxyalkanoate Synthesis from Biogas by Methylocystis hirsuta. ACS chain-length polyhydroxyalkanoate synthesis in engineered Rhodospirillum rubrum.
Sustainable Chemistry & Engineering, 8(9), 3906-3912. Applied and environmental microbiology 82 (20), 6132–6140.
Chidambarampadmavathy, K., Karthikeyan, O.P., Heimann, K., 2015. Biopolymers made Heinrich, D., Raberg, M., Steinbüchel, A., 2018. Studies on the aerobic utilization of
from methane in bioreactors. Engineering in Life Sciences 15 (7), 689–699. synthesis gas (syngas) by wild type and recombinant strains of Ralstonia eutropha
Choi, D., Chipman, D.C., Bents, S.C., Brown, R.C., 2010. A techno-economic analysis of H16. Microbial biotechnology 11 (4), 647–656.
polyhydroxyalkanoate and hydrogen production from syngas fermentation of Helm, J., Wendlandt, K.D., Jechorek, M., Stottmeister, U., 2008. Potassium deficiency
gasified biomass. Applied biochemistry and biotechnology 160 (4), 1032–1046. results in accumulation of ultra-high molecular weight poly-β-hydroxybutyrate in a
Choi, E.S., Min, K., Kim, G.-J., Kwon, I., Kim, Y.H., 2017. Expression and characterization methane-utilizing mixed culture. Journal of applied microbiology 105 (4),
of Pantoea CO dehydrogenase to utilize CO-containing industrial waste gas for 1054–1061.
expanding the versatility of CO dehydrogenase. Scientific reports 7, 44323. Helm, J., Wendlandt, K.D., Rogge, G., Kappelmeyer, U., 2006. Characterizing a stable
Choi, So Young, Rhie, Mi Na, Kim, Hee Taek, Joo, Jeong Chan, Cho, In Jin, Son, Jina, methane-utilizing mixed culture used in the synthesis of a high-quality biopolymer
Jo, Seo Young, Sohn, Yu Jung, Baritugo, Kei-Anne, Pyo, Jiwon, Lee, Youngjoon, in an open system. Journal of applied microbiology 101 (2), 387–395.
Lee, Sang Yup, Park, Si Jae, 2020. Metabolic engineering for the synthesis of Hondo, S., Takahashi, M., Osanai, T., Matsuda, M., Hasunuma, T., Tazuke, A.,
polyesters: A 100-year journey from polyhydroxyalkanoates to non-natural Nakahira, Y., Chohnan, S., Hasegawa, M., Asayama, M., 2015. Genetic engineering
microbial polyesters. Metabolic engineering 58, 47–81. and metabolite profiling for overproduction of polyhydroxybutyrate in
Costa, J.A.V., Moreira, J.B., Lucas, B.F., Braga, V.d.S., Cassuriaga, A.P.A., Morais, M.G.d. cyanobacteria. Journal of bioscience and bioengineering 120 (5), 510–517.
2018. Recent advances and future perspectives of PHB production by cyanobacteria. Humphreys, C.M., Minton, N.P., 2018. Advances in metabolic engineering in the
Industrial Biotechnology, 14(5), 249-256. microbial production of fuels and chemicals from C1 gas. Current opinion in
D’Amico, D.A., Manfredi, L.B., Cyras, V.P., 2012. Relationship between thermal biotechnology 50, 174–181.
properties, morphology, and crystallinity of nanocomposites based on Hundertmark, T., Mayer, M., McNally, C., Simons, T.J., Witte, C., 2018. How plastics
polyhydroxybutyrate. Journal of Applied Polymer Science 123 (1), 200–208. waste recycling could transform the chemical industry. McKinsey & Company 12.
da Costa, J.P. 2021. The 2019 Global pandemic and plastic pollution prevention Hwang, H.W., Yoon, J., Min, K., Kim, M.-S., Kim, S.-J., Cho, D.H., Susila, H., Na, J.-G.,
measures: playing catch-up. Science of The Total Environment, 145806. Oh, M.-K., Kim, Y.H., 2020. Two-stage bioconversion of carbon monoxide to
Dalsasso, R.R., Pavan, F.A., Bordignon, S.E., de Aragão, G.M.F., Poletto, P., 2019. biopolymers via formate as an intermediate. Chemical Engineering Journal 389,
Polyhydroxybutyrate (PHB) production by Cupriavidus necator from sugarcane 124394.
vinasse and molasses as mixed substrate. Process Biochemistry 85, 12–18. Im, C., Kim, H., Zaheer, J., Kim, J.Y., Lee, Y.J., Kang, C.M., Kim, J.S., 2021. PET tracing
Do, Y.S., Smeenk, J., Broer, K.M., Kisting, C.J., Brown, R., Heindel, T.J., Bobik, T.A., of biodistribution for orally administered 64Cu-labeled polystyrene in mice. Journal
DiSpirito, A.A., 2007. Growth of Rhodospirillum rubrum on synthesis gas: conversion of Nuclear Medicine. https://doi.org/10.2967/jnumed.120.256982.
of CO to H2 and poly-β-hydroxyalkanoate. Biotechnology and Bioengineering 97 (2), Jang, Y.-C., Lee, G., Kwon, Y., Lim, J.-H., Jeong, J.-H., 2020. Recycling and management
279–286. practices of plastic packaging waste towards a circular economy in South Korea.
Drzyzga, O., Revelles, O., Durante-Rodríguez, G., Díaz, E., García, J.L., Prieto, A., 2015. Resources, Conservation and Recycling 158, 104798.
New challenges for syngas fermentation: towards production of biopolymers. Jiang, W., Villamor, D.H., Peng, H., Chen, J., Liu, L., Haritos, V., Ledesma-Amaro, R.,
Journal of Chemical Technology & Biotechnology 90 (10), 1735–1751. 2021. Metabolic engineering strategies to enable microbial utilization of C1
Duerre, P., Eikmanns, B.J., 2015. C1-carbon sources for chemical and fuel production by feedstocks. Nature Chemical Biology 17 (8), 845–855.
microbial gas fermentation. Current opinion in biotechnology 35, 63–72. Kamravamanesh, Donya, Kovacs, Tamas, Pflügl, Stefan, Druzhinina, Irina, Kroll, Paul,
Ernst, A., Zibrak, J.D., 1998. Carbon monoxide poisoning. New England journal of Lackner, Maximilian, Herwig, Christoph, 2018. Increased poly-β-hydroxybutyrate
medicine 339 (22), 1603–1608. production from carbon dioxide in randomly mutated cells of cyanobacterial strain
Favaro, L., Basaglia, M., Casella, S., 2019. Improving polyhydroxyalkanoate production Synechocystis sp. PCC 6714: Mutant generation and characterization. Bioresource
from inexpensive carbon sources by genetic approaches: a review. Biofuels, technology 266, 34–44.
Bioproducts and Biorefining 13 (1), 208–227. Karmann, S., Panke, S., Zinn, M., 2019. Fed-batch cultivations of Rhodospirillum rubrum
Fei, P., Luo, Y., Lai, N., Wu, H., 2021. Biosynthesis of (R)-3-hydroxybutyric acid from under multiple nutrient-limited growth conditions on syngas as a novel option to
syngas-derived acetate in engineered Escherichia coli. Bioresource Technology produce poly (3-hydroxybutyrate)(PHB). Frontiers in bioengineering and
125323. biotechnology 7, 59.
Fergala, A., AlSayed, A., Khattab, S., Ramirez, M., Eldyasti, A., 2018. Development of Karthikeyan, O.P., Chidambarampadmavathy, K., Nadarajan, S., Lee, P.K., Heimann, K.,
methane-utilizing mixed cultures for the production of polyhydroxyalkanoates 2015. Effect of CH4/O2 ratio on fatty acid profile and polyhydroxybutyrate content
(PHAs) from anaerobic digester sludge. Environmental science & technology 52 in a heterotrophic–methanotrophic consortium. Chemosphere 141, 235–242.
(21), 12376–12387. Koch, M., Bruckmoser, J., Scholl, J., Hauf, W., Rieger, B., Forchhammer, K., 2020.
Fernández-Naveira, Á., Abubackar, H.N., Veiga, M.C., Kennes, C., 2016. Efficient Maximizing PHB content in Synechocystis sp. PCC 6803: a new metabolic engineering
butanol-ethanol (BE) production from carbon monoxide fermentation by Clostridium strategy based on the regulator PirC. Microbial Cell Factories 19 (1), 1–12.
carboxidivorans. Applied microbiology and biotechnology 100 (7), 3361–3370. Koller, M., 2019. Polyhydroxyalkanoate biosynthesis at the edge of water activity-
Flüchter, S., Follonier, S., Schiel-Bengelsdorf, B., Bengelsdorf, F.R., Zinn, M., Dürre, P., Haloarchaea as biopolyester factories. Bioengineering 6 (2), 34.
2019. Anaerobic production of poly (3-hydroxybutyrate) and its precursor 3- La, H., Hettiaratchi, J.P.A., Achari, G., Dunfield, P.F., 2018. Biofiltration of methane.
Bioresource technology 268, 759–772.

13
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Lagoa-Costa, B., Abubackar, H.N., Fernández-Romasanta, M., Kennes, C., Veiga, M.C., Mozumder, M.S.I., Garcia-Gonzalez, L., De Wever, H., Volcke, E.I., 2015. Poly (3-
2017. Integrated bioconversion of syngas into bioethanol and biopolymers. hydroxybutyrate)(PHB) production from CO2: model development and process
Bioresource technology 239, 244–249. optimization. Biochemical Engineering Journal 98, 107–116.
Lai, M.C., Lan, E.I., 2015. Advances in metabolic engineering of cyanobacteria for Myung, J., Flanagan, J.C., Waymouth, R.M., Criddle, C.S., 2017. Expanding the range of
photosynthetic biochemical production. Metabolites 5 (4), 636–658. polyhydroxyalkanoates synthesized by methanotrophic bacteria through the
Lau, N.-S., Matsui, M., Abdullah, A.A.-A., 2015. Cyanobacteria: photoautotrophic utilization of omega-hydroxyalkanoate co-substrates. AMB Express 7 (1), 1–10.
microbial factories for the sustainable synthesis of industrial products. BioMed Myung, J., Flanagan, J.C., Waymouth, R.M., Criddle, C.S., 2016a. Methane or methanol-
research international 2015, 754934. oxidation dependent synthesis of poly (3-hydroxybutyrate-co-3-hydroxyvalerate) by
Lebreton, L., Slat, B., Ferrari, F., Sainte-Rose, B., Aitken, J., Marthouse, R., Hajbane, S., obligate type II methanotrophs. Process Biochemistry 51 (5), 561–567.
Cunsolo, S., Schwarz, A., Levivier, A., 2018. Evidence that the Great Pacific Garbage Myung, J., Kim, M., Pan, M., Criddle, C.S., Tang, S.K., 2016b. Low energy emulsion-
Patch is rapidly accumulating plastic. Scientific reports 8 (1), 1–15. based fermentation enabling accelerated methane mass transfer and growth of poly
Lee, O.K., Hur, D.H., Nguyen, D.T.N., Lee, E.Y., 2016. Metabolic engineering of (3-hydroxybutyrate)-accumulating methanotrophs. Bioresource technology 207,
methanotrophs and its application to production of chemicals and biofuels from 302–307.
methane. Biofuels, Bioproducts and Biorefining 10 (6), 848–863. Nangle, S.N., Ziesack, M., Buckley, S., Trivedi, D., Loh, D.M., Nocera, D.G., Silver, P.A.,
Lemgruber, R.d.S.P., Valgepea, K., Tappel, R., Behrendorff, J.B., Palfreyman, R.W., Plan, 2020. Valorization of CO2 through lithoautotrophic production of sustainable
M., Hodson, M.P., Simpson, S.D., Nielsen, L.K., Köpke, M. 2019. Systems-level chemicals in Cupriavidus necator. Metabolic Engineering 62, 207–220.
engineering and characterisation of Clostridium autoethanogenum through Nguyen, T.T., Lee, E.Y., 2021. Methane-based biosynthesis of 4-hydroxybutyrate and P
heterologous production of poly-3-hydroxybutyrate (PHB). Metabolic engineering, 53, (3-hydroxybutyrate-co-4-hydroxybutyrate) using engineered Methylosinus
14-23. trichosporium OB3b. Bioresource Technology 335, 125263.
Levett, I., Birkett, G., Davies, N., Bell, A., Langford, A., Laycock, B., Lant, P., Pratt, S., Orita, I., Nishikawa, K., Nakamura, S., Fukui, T., 2014. Biosynthesis of
2016. Techno-economic assessment of poly-3-hydroxybutyrate (PHB) production polyhydroxyalkanoate copolymers from methanol by Methylobacterium extorquens
from methane—The case for thermophilic bioprocessing. Journal of Environmental AM1 and the engineered strains under cobalt-deficient conditions. Applied
Chemical Engineering 4 (4), 3724–3733. microbiology and biotechnology 98 (8), 3715–3725.
Li, M., Wilkins, M.R., 2020. Recent advances in polyhydroxyalkanoate production: Panda, B., Mallick, N., 2007. Enhanced poly-β-hydroxybutyrate accumulation in a
feedstocks, strains and process developments. International journal of biological unicellular cyanobacterium, Synechocystis sp. PCC 6803. Letters in applied
macromolecules 156, 691–703. microbiology 44 (2), 194–198.
Liu, L.-Y., Xie, G.-J., Xing, D.-F., Liu, B.-F., Ding, J., Ren, N.-Q., 2020. Biological Panich, J., Fong, B., Singer, S.W., 2021. Metabolic Engineering of Cupriavidus necator
conversion of methane to polyhydroxyalkanoates: current advances, challenges, and H16 for Sustainable Biofuels from CO2. Trends in biotechnology 39 (4), 412–424.
perspectives. Environmental Science and Ecotechnology 2, 100029. Park, I., Jho, E.H., Nam, K., 2014. Optimization of carbon dioxide and valeric acid
López, J.C., Arnáiz, E., Merchán, L., Lebrero, R., Muñoz, R., 2018. Biogas-based utilization for polyhydroxyalkanoates synthesis by Cupriavidus necator. Journal of
polyhydroxyalkanoates production by Methylocystis hirsuta: a step further in Polymers and the Environment 22 (2), 244–251.
anaerobic digestion biorefineries. Chemical Engineering Journal 333, 529–536. Pérez, R., Cantera, S., Bordel, S., García-Encina, P.A., Muñoz, R., 2019. The effect of
Lovley, D.R., 2011. Powering microbes with electricity: direct electron transfer from temperature during culture enrichment on methanotrophic polyhydroxyalkanoate
electrodes to microbes. Environmental microbiology reports 3 (1), 27–35. production. International Biodeterioration & Biodegradation 140, 144–151.
Löwe, H., Hobmeier, K., Moos, M., Kremling, A., Pflüger-Grau, K., 2017. Pradani, L., Rohman, M.S., Margino, S., 2020. The structural insight of class III of
Photoautotrophic production of polyhydroxyalkanoates in a synthetic mixed culture polyhydroxyalkanoate synthase from Bacillus sp. PSA10 as revealed by in silico
of Synechococcus elongatus cscB and Pseudomonas putida cscAB. Biotechnology for analysis. Indonesian. Journal of Biotechnology 25 (1), 33–42.
biofuels 10 (1), 1–11. Ray, S., Kalia, V.C., 2017. Microbial cometabolism and polyhydroxyalkanoate co-
Luangthongkam, P., Laycock, B., Evans, P., Lant, P., Pratt, S., 2019a. Thermophilic polymers. Indian journal of microbiology 57 (1), 39–47.
production of poly (3-hydroxybutyrate-co-3-hydrovalerate) by a mixed methane- Reichardt, R., Rieger, B., 2011. Poly (3-hydroxybutyrate) from carbon monoxide.
utilizing culture. New biotechnology 53, 49–56. Synthetic Biodegradable Polymers 49–90.
Luangthongkam, P., Strong, P.J., Mahamud, S.N.S., Evans, P., Jensen, P., Tyson, G., Revelles, O., Beneroso, D., Menendez, J.A., Arenillas, A., García, J.L., Prieto, M.A., 2017.
Laycock, B., Lant, P.A., Pratt, S., 2019b. The effect of methane and odd-chain fatty Syngas obtained by microwave pyrolysis of household wastes as feedstock for
acids on 3-hydroxybutyrate (3HB) and 3-hydroxyvalerate (3HV) synthesis by a polyhydroxyalkanoate production in Rhodospirillum rubrum. Microbial biotechnology
Methylosinus-dominated mixed culture. Bioresources and Bioprocessing 6 (1), 1–10. 10 (6), 1412–1417.
Market, M., 2019. Polyhydroxyalkanoate (PHA) Market by Type (Short Chain Length, Revelles, O., Tarazona, N., García, J.L., Prieto, M.A., 2016. Carbon roadmap from syngas
Medium Chain Length), Production Method (Sugar Fermentation, Vegetable Oil to polyhydroxyalkanoates in Rhodospirillum rubrum. Environmental microbiology 18
Fermentation, Methane Fermentation). Application, and Region-Global Forecast to (2), 708–720.
2024. Markets and Markets Research Private Ltd. Rodríguez, Y., Firmino, P.I.M., Arnáiz, E., Lebrero, R., Muñoz, R., 2020a. Elucidating the
Marlina, Dian, Sato, Harumi, Hoshina, Hiromichi, Ozaki, Yukihiro, 2018. Intermolecular influence of environmental factors on biogas-based polyhydroxybutyrate production
interactions of poly (3-hydroxybutyrate-co-3-hydroxyvalerate) (P (HB-co-HV)) with by Methylocystis hirsuta CSC1. Science of The Total Environment 706, 135136.
PHB-type crystal structure and PHV-type crystal structure studied by low-frequency Rodríguez, Y., Firmino, P.I.M., Pérez, V., Lebrero, R., Muñoz, R., 2020b. Biogas
Raman and terahertz spectroscopy. Polymer 135, 331–337. valorization via continuous polyhydroxybutyrate production by Methylocystis hirsuta
Matsumoto, K.i., Kitagawa, K., Jo, S.-J., Song, Y., Taguchi, S. 2011. Production of poly in a bubble column bioreactor. Waste Management 113, 395–403.
(3-hydroxybutyrate-co-3-hydroxyvalerate) in recombinant Corynebacterium Roh, H., Lee, J.S., Choi, H.I., Sung, Y.J., Choi, S.Y., Woo, H.M., Sim, S.J., 2021. Improved
glutamicum using propionate as a precursor. Journal of biotechnology, 152(4), 144- CO2-derived polyhydroxybutyrate (PHB) production by engineering fast-growing
146. cyanobacterium Synechococcus elongatus UTEX 2973 for potential utilization of flue
Meereboer, K.W., Misra, M., Mohanty, A.K., 2020. Review of recent advances in the gas. Bioresource technology 327, 124789.
biodegradability of polyhydroxyalkanoate (PHA) bioplastics and their composites. Sagong, H.-Y., Son, H.F., Choi, S.Y., Lee, S.Y., Kim, K.-J., 2018. Structural insights into
Green Chemistry 22 (17), 5519–5558. polyhydroxyalkanoates biosynthesis. Trends in biochemical sciences 43 (10),
Mezzolla, V., D’Urso, O.F., Poltronieri, P., 2018. Role of PhaC type I and type II enzymes 790–805.
during PHA biosynthesis. Polymers 10 (8), 910. Salem, R., Soliman, M., Fergala, A., Audette, G.F., ElDyasti, A., 2021. Screening for
Mittermair, S., Richter, J., Doppler, P., Trenzinger, K., Nicoletti, C., Forsich, C., methane utilizing mixed communities with high polyhydroxybutyrate (Phb)
Spadiut, O., Herwig, C., Lackner, M., 2021. Impact of exoD gene knockout on the production capacity using different design approaches. Polymers 13 (10), 1579.
polyhydroxybutyrate overaccumulating mutant Mt_a24. International Journal of Sarma, M.K., Kaushik, S., Goswami, P., 2016. Cyanobacteria: A metabolic power house
Biobased Plastics 3 (1), 1–18. for harvesting solar energy to produce bio-electricity and biofuels. Biomass and
Miyahara, Y., Yamamoto, M., Thorbecke, R., Mizuno, S., Tsuge, T., 2020. Autotrophic Bioenergy 90, 187–201.
biosynthesis of polyhydroxyalkanoate by Ralstonia eutropha from non-combustible Schäfer, C., Friedrich, B., Lenz, O., 2013. Novel, oxygen-insensitive group 5 [NiFe]-
gas mixture with low hydrogen content. Biotechnology letters 42 (9), 1655–1662. hydrogenase in Ralstonia eutropha. Applied and environmental microbiology 79 (17),
Mongili, B., Fino, D., 2021. Carbon monoxide fermentation to bioplastic: the effect of 5137–5145.
substrate adaptation on Rhodospirillum rubrum. Biomass Conversion and Biorefinery Schoelmerich, M.C., Müller, V., 2019. Energy conservation by a hydrogenase-dependent
11 (2), 705–714. chemiosmotic mechanism in an ancient metabolic pathway. Proceedings of the
Monshupanee, T., Incharoensakdi, A., 2014. Enhanced accumulation of glycogen, lipids National Academy of Sciences 116 (13), 6329–6334.
and polyhydroxybutyrate under optimal nutrients and light intensities in the Schubert, P., Steinbüchel, A., Schlegel, H.G., 1988. Cloning of the Alcaligenes eutrophus
cyanobacterium Synechocystis sp. PCC 6803. Journal of applied microbiology 116 genes for synthesis of poly-beta-hydroxybutyric acid (PHB) and synthesis of PHB in
(4), 830–838. Escherichia coli. Journal of bacteriology 170 (12), 5837–5847.
Moreira, J.P., Diender, M., Arantes, A.L., Boeren, S., Stams, A.J., Alves, M.M., Alves, J.I., Schuchmann, K., Müller, V., 2014. Autotrophy at the thermodynamic limit of life: a
Sousa, D.Z., 2021. Propionate production from carbon monoxide by synthetic co- model for energy conservation in acetogenic bacteria. Nature Reviews Microbiology
cultures of Acetobacterium wieringae spp. and propionigenic bacteria. Applied and 12 (12), 809–821.
Environmental Microbiology. Shah, N.N., Hanna, M.L., Taylor, R.T., 1996. Batch cultivation of Methylosinus
Morgan-Sagastume, F., Valentino, F., Hjort, M., Cirne, D., Karabegovic, L., Gerardin, F., trichosporium OB3b: V. Characterization of poly-β-hydroxybutyrate production under
Johansson, P., Karlsson, A., Magnusson, P., Alexandersson, T., 2014. methane-dependent growth conditions. Biotechnology and bioengineering 49 (2),
Polyhydroxyalkanoate (PHA) production from sludge and municipal wastewater 161–171.
treatment. Water Science and Technology 69 (1), 177–184. Shi, Q., Tang, J., Liu, R., Wang, L., 2021. Toxicity in vitro reveals potential impacts of
microplastics and nanoplastics on human health: A review. Critical Reviews in
Environmental Science and Technology 1–33.

14
J. Yoon and M.-K. Oh Bioresource Technology 344 (2022) 126307

Shin, S.K., Ko, Y.J., Lee, M.-E., You, S.K., Hyeon, J.E., Han, S.O., 2021. Enhanced Vidra, A., Németh, Á., 2018. Bio-produced acetic acid: a review. Periodica Polytechnica
production of polyhydroxybutyrate from syngas by using nanoscaled cellulose Chemical Engineering 62 (3), 245–256.
particles with a syngas-converting enzyme complex immobilized on Ralstonia Vigneswari, S., Noor, M.S.M., Amelia, T.S.M., Balakrishnan, K., Adnan, A., Bhubalan, K.,
eutropha. Journal of Cleaner Production 285, 124903. Amirul, A.-A.A., Ramakrishna, S., 2021. Recent Advances in the Biosynthesis of
Singh, M.K., Rai, P.K., Rai, A., Singh, S., Singh, J.S., 2019. Poly-β-hydroxybutyrate Polyhydroxyalkanoates from Lignocellulosic Feedstocks. Life 11 (8), 807.
production by the cyanobacterium Scytonema geitleri Bharadwaja under varying Volova, T., Zhila, N., Shishatskaya, E., 2015. Synthesis of poly (3-hydroxybutyrate) by
environmental conditions. Biomolecules 9 (5), 198. the autotrophic CO-oxidizing bacterium Seliberia carboxydohydrogena Z-1062.
Taepucharoen, K., Tarawat, S., Puangcharoen, M., Incharoensakdi, A., Monshupanee, T., Journal of Industrial Microbiology and Biotechnology 42 (10), 1377–1387.
2017. Production of poly (3-hydroxybutyrate-co-3-hydroxyvalerate) under Volova, T.G., Kiselev, E.G., Shishatskaya, E.I., Zhila, N.O., Boyandin, A.N.,
photoautotrophy and heterotrophy by non-heterocystous N2-fixing cyanobacterium. Syrvacheva, D.A., Vinogradova, O.N., Kalacheva, G.S., Vasiliev, A.D., Peterson, I.V.,
Bioresource technology 239, 523–527. 2013. Cell growth and accumulation of polyhydroxyalkanoates from CO2 and H2 of a
Taga, Naohiko, Tanaka, Kenji, Ishizaki, Ayaaki, 1997. Effects of rheological change by hydrogen-oxidizing bacterium, Cupriavidus eutrophus B-10646. Bioresource
addition of carboxymethylcellulose in culture media of an air-lift fermentor on poly- technology 146, 215–222.
D-3-hydroxybutyric acid productivity in autotrophic culture of hydrogen-oxidizing Weiss, T.L., Young, E.J., Ducat, D.C., 2017. A synthetic, light-driven consortium of
bacterium. Alcaligenes eutrophus. Biotechnology and bioengineering 53 (5), 529–533. cyanobacteria and heterotrophic bacteria enables stable polyhydroxybutyrate
Tamis, J., Lužkov, K., Jiang, Y., van Loosdrecht, M.C., Kleerebezem, R., 2014. production. Metabolic engineering 44, 236–245.
Enrichment of Plasticicumulans acidivorans at pilot-scale for PHA production on Wendlandt, K.-D., Jechorek, M., Helm, J., Stottmeister, U., 2001. Producing poly-3-
industrial wastewater. Journal of Biotechnology 192, 161–169. hydroxybutyrate with a high molecular mass from methane. Journal of
Tan, G.-Y.A., Chen, C.-L., Li, L., Ge, L., Wang, L., Razaad, I.M.N., Li, Y., Zhao, L., Mo, Y., biotechnology 86 (2), 127–133.
Wang, J.-Y., 2014. Start a research on biopolymer polyhydroxyalkanoate (PHA): a Wilkins, M.R., Atiyeh, H.K., 2011. Microbial production of ethanol from carbon
review. Polymers 6 (3), 706–754. monoxide. Current Opinion in Biotechnology 22 (3), 326–330.
Tanaka, K., Ishizaki, A., Kanamaru, T., Kawano, T., 1995. Production of poly (D-3- Yoon, J., Chang, W., Oh, S.-H., Choi, S.-H., Yang, Y.-H., Oh, M.-K., 2021. Metabolic
hydroxybutyrate) from CO2, H2, and O2 by high cell density autotrophic cultivation engineering of Methylorubrum extorquens AM1 for poly (3-hydroxybutyrate-co-3-
of Alcaligenes eutrophus. Biotechnology and bioengineering 45 (3), 268–275. hydroxyvalerate) production using formate. International Journal of Biological
Tanaka, K., Miyawaki, K., Yamaguchi, A., Khosravi-Darani, K., Matsusaki, H., 2011. Cell Macromolecules 177, 284–293.
growth and P (3HB) accumulation from CO2 of a carbon monoxide-tolerant Zaldívar Carrillo, J.A., Stein, L.Y., Sauvageau, D., 2018. Defining nutrient combinations
hydrogen-oxidizing bacterium, Ideonella sp. O-1. Applied microbiology and for optimal growth and polyhydroxybutyrate production by Methylosinus
biotechnology 92 (6), 1161–1169. trichosporium OB3b using response surface methodology. Frontiers in microbiology
Tanaka, K., Mori, S., Hirata, M., Matsusaki, H., 2016. Autotrophic growth of Paracoccus 9, 1513.
denitrificans in aerobic condition and the accumulation of biodegradable plastics Zhang, T., Wang, X., Zhou, J., Zhang, Y., 2018. Enrichments of
from CO2. Environ Ecol Res 4, 231–236. methanotrophic–heterotrophic cultures with high poly-β-hydroxybutyrate (PHB)
Tarawat, Somchai, Incharoensakdi, Aran, Monshupanee, Tanakarn, 2020. accumulation capacities. Journal of Environmental Sciences 65, 133–143.
Cyanobacterial production of poly (3-hydroxybutyrate-co-3-hydroxyvalerate) from Zhu, B., van Dijk, G., Fritz, C., Smolders, A.J., Pol, A., Jetten, M.S., Ettwig, K.F., 2012.
carbon dioxide or a single organic substrate: improved polymer elongation with an Anaerobic oxidization of methane in a minerotrophic peatland: enrichment of
extremely high 3-hydroxyvalerate mole proportion. Journal of Applied Phycology 32 nitrite-dependent methane-oxidizing bacteria. Applied and environmental
(2), 1095–1102. microbiology 78 (24), 8657–8665.
Verma, R., Vinoda, K., Papireddy, M., Gowda, A., 2016. Toxic pollutants from plastic Zuñiga, C., Morales, M., Revah, S., 2013. Polyhydroxyalkanoates accumulation by
waste-a review. Procedia Environmental Sciences 35, 701–708. Methylobacterium organophilum CZ-2 during methane degradation using citrate or
Vermaas, W.F. 2001. Photosynthesis and respiration in cyanobacteria. e LS, https://doi. propionate as cosubstrates. Bioresource technology 129, 686–689.
org/10.1038/npg.els.0001670.

15

You might also like