Next Article in Journal
Lignin-Based Materials for Sustainable Rechargeable Batteries
Next Article in Special Issue
Functionality of Yeast β-Glucan Recovered from Kluyveromyces marxianus by Alkaline and Enzymatic Processes
Previous Article in Journal
Development of Porous Polyvinyl Acetate/Polypyrrole/Gallic Acid Scaffolds Using Supercritical CO2 as Tissue Regenerative Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Critical Review on the Economically Feasible and Sustainable Poly(3-Hydroxybutyrate-co-3-hydroxyvalerate) Production from Alkyl Alcohols

by
Hau Seung Jeremy Wong
1,2,
Kesaven Bhubalan
3,4 and
Al-Ashraf Abdullah Amirul
1,2,*
1
School of Biological Sciences, Universiti Sains Malaysia, Gelugor 11800, Penang, Malaysia
2
Centre for Chemical Biology, Universiti Sains Malaysia, Bayan Lepas 11900, Penang, Malaysia
3
Eco-Innovation Research Interest Group, Faculty of Science and Marine Environment, Universiti Malaysia Terengganu, Kuala Nerus 21030, Terengganu, Malaysia
4
Institute of Marine Biotechnology, Universiti Malaysia Terengganu, Kuala Nerus 21030, Terengganu, Malaysia
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(4), 670; https://doi.org/10.3390/polym14040670
Submission received: 17 January 2022 / Revised: 3 February 2022 / Accepted: 5 February 2022 / Published: 10 February 2022
(This article belongs to the Special Issue Sustainable Biochemicals and Biopolymers from Bioresources)

Abstract

:
Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (P(3HB-co-3HV)) is the most studied short-chain-length polyhydroxyalkanoates (PHA) with high application importance in various fields. The domination of high-cost propionate and valerate over other 3-hydroxyvalerate (3HV) precursors owing to their wide preference among PHA-producing bacteria has hindered the development of diverse production processes. As alkyl alcohols are mainly produced from inexpensive starting materials through oxo synthesis, they contribute a cost-effective advantage over propionate and valerate. Moreover, alkyl alcohols can be biosynthesized from natural substrates and organic wastes. Despite their great potential, their toxicity to most PHA-producing bacteria has been the major drawback for their wide implementation as 3HV precursors for decades. Although the standard PHA-producing bacteria Cupriavidus necator showed promising alcohol tolerance, the 3HV yield was discouraging. Continuous discovery of alkyl alcohols-utilizing PHA-producing bacteria has enabled broader choices in 3HV precursor selection for diverse P(3HB-co-3HV) production processes with higher economic feasibility. Besides continuous effort in searching for promising wild-type strains, genetic engineering to construct promising recombinant strains based on the understanding of the mechanisms involved in alkyl alcohols toxicity and tolerance is an alternative approach. However, more studies are required for techno-economic assessment to analyze the economic performance of alkyl alcohol-based production compared to that of organic acids.

Graphical Abstract

1. General Overview

Polyhydroxyalkanoates (PHA) are emerging as the next generation plastics owing to their plastic-like properties, renewability, biodegradability, and biocompatibility [1]. PHA are accumulated by bacteria under carbon excess but nitrogen-limiting conditions and stored as a reserved energy source in the form of single or multiple granules in the cytoplasm [2]. PHA have gained much industrial interest in the last few decades due to their potential as substitutes for conventional plastics, and various fermentation strategies have been developed to establish microbial PHA production for commercialization. Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (P(3HB-co-3HV)) is the most studied PHA copolymer with mechanical properties comparable to that of polypropylene. The 3-hydroxyvalerate (3HV) monomer provides elastomeric property to the copolymer, enabling broader application compared to the homopolymer poly(3-hydroxybutyrate) (P(3HB)) [3]. The improvement in mechanical properties has paved the way for it to be established for medical, tissue engineering, aquacultural, agricultural, and commodity applications. The commercialization of P(3HB) and P(3HB-co-3HV) started in the 1970s by Imperial Chemical Industries, U.K., and Chemie Linz AG, Austria [4]. Currently, P(3HB) and P(3HB-co-3HV) are commercialized by TianAn Biopolymer, China, and Sigma-Aldrich, USA.
Commercialization of PHA is hampered by its high production cost, majorly due to the cost of the carbon feedstock used in microbial fermentation. Over recent decades, various industrial wastes were explored as alternative carbon sources, and numerous mitigation strategies were taken to establish microbial production of P(3HB-co-3HV) with high economic feasibility at a commercial scale. Bioconversion of unrelated carbon sources into P(3HB-co-3HV) was attempted, but metabolic engineering strategies are generally required to promote precursor-independent pathways to synthesis P(3HB-co-3HV), with exceptions for wild types Nocardia or Rhodococcus that can generate propionyl-CoA endogenously from a single carbon source [5,6,7,8,9]. Owing to the relatively simpler practical requirement, P(3HB-co-3HV) production from related carbon source(s) remains competitive. Although the employment of wastes contributes to higher economic feasibility, P(3HB-co-3HV) production from a single carbon source has low practicability due to the composition inconsistency of raw components for 3HV formation [3].
The most common way to incorporate 3HV monomers is by employing a precursor carbon source as a co-substrate along with the main carbon source that contributes to the 3-hydroxybutyrate (3HB) monomer. Precursor carbon sources such as organic acids, alcohols, or some amino acids were studied thoroughly to clarify the metabolic pathways involved and to search for promising precursors of greater potential. Organic acids, especially propionic acid, valeric acid, and their respective salts, are the standard 3HV precursors owing to their wide acceptance among PHA-producing bacteria. However, organic acids can only be added in low concentrations due to their high toxicity to the bacteria, and their high substrate cost causes lower profitability. Although levulinic acid is way more cost-effective than propionic acid and valeric acid, it seems to be a privilege for Cupriavidus necator, and the production mechanism is yet to be clarified [10,11]. Although some amino acids such as threonine, valine, and isoleucine could be employed as 3HV precursors, metabolic engineering of the amino acid biosynthetic pathways is required to convert amino acids into propionyl-CoA, which is essential for 3HV formation. The rare occurrence of alcohols-utilizing ability among PHA-producing bacteria hinders the employment of alkyl alcohols as 3HV precursors despite their potential as cost-effective substitutes for organic acids [3]. In addition to the merit in lowering the substrate cost, naturally occurring carbon sources such as glucose and glycerol or organic wastes can be converted by microorganisms into alkyl alcohols, thus are promising as cost-effective and sustainable bioresources for P(3HB-co-3HV) production [12].
C. necator is the standard PHA-producing bacterium well-known with its wide substrate acceptance range, including alcohols and mercury. Nevertheless, its capability to convert alcohols into PHA is substandard. The 3HV yield from 1-propanol is low despite its high tolerance toward 1-propanol, and the employment of 1-pentanol results in a remarkably high reduction in C. necator cell biomass and PHA content [10,13]. Owing to the economic advantage over organic acids, the employment of alcohols as the 3HV precursors for P(3HB-co-3HV) production was attempted for various bacteria. Interestingly, P(3HB-co-3HV)-producing bacteria favoring alcohols as 3HV precursors are emerging since the last decade. Since the discovery of Paracoccus denitrificans ATCC 17741 with the capability to convert 1-pentanol into 3HV in 1996, various alkyl alcohol-tolerant PHA-producing bacteria were discovered continually whereby several of them depicted promising 3HV yield [14].
This critical review condenses the production of P(3HB-co-3HV) from alkyl alcohols and the promising potential of alkyl alcohols as cost-effective 3HV precursors to go beyond the bottleneck in precursors selection that is limited to organic acids. The properties and applications of P(3HB-co-3HV) are also discussed. The bioconversion pathways of 1-propanol and 1-pentanol into 3HV with respect to propionic acid and valeric acid are visualized, and the performance of discovered alkyl alcohol-tolerant PHA-producing bacteria is highlighted. Oxo synthesis and biosynthesis of 1-propanol and 1-pentanol from natural substrates as well as organic wastes were described. Furthermore, the mode of action of alkyl alcohols on bacterial proteins and the bacterial mechanisms involved in response to alcoholic stress are also discussed. The strategies for wide implementation of alkyl alcohols for P(3HB-co-3HV) production and the challenges ahead are highlighted as well to comment on the potential of alkyl alcohols as the next generation 3HV precursors.

2. P(3HB-co-3HV) Properties and Applications

P(3HB) is a relatively stiff and brittle polyester with poor elongation at break [15]. It is a fragile material, and its mechanical properties deteriorate with time due to secondary crystallization accompanied by aging at room temperature, which is the major cause of its brittleness [16]. Although the lack of elasticity causes a drawback in its application as packaging materials, its high mechanical properties are applicable as bone tissues aid in supporting body weight. P(3HB) facilitates reconstructive osteogenesis. P(3HB) and its biocomposite incorporated with 20 wt% hydroxyapatite, which makes up 65–70% of the bone matrix, show pronounced osteoplastic properties owing to their slow degradation that corresponds to the growth of new bones. Powdered P(3HB) and P(3HB)/tienam are excellent antibacterial bone filling materials that contribute to 1-fold lower growth and complete growth inhibition of Staphylococcus aureus post surgery, respectively [17].
The incorporation of the C5 3HV monomer into P(3HB) results in P(3HB-co-3HV) with decreased crystallinity, thus leading to decreased stiffness, decreased brittleness, and enhanced biodegradability compared to that of P(3HB) [18]. The properties of P(3HB-co-3HV) are dependent on the ratio of the two monomers where the 3HB monomer contributes stiffness, and the 3HV monomer contributes flexibility to the copolymer. The composition of the 3HV monomer determines the defection of the P(3HB) lamellae crystals, leading to the disruption of its crystallinity and resulting in improved polymer flexibility (Figure 1) [19]. The lower degree of crystallinity and melting point of P(3HB-co-3HV) lead to a higher degradation rate that is directly proportional to the molar fraction of 3HV of the copolymer compared to that of P(3HB) [18]. The 3HV fraction contributes to a greater amorphous region for enzymatic attacks that leads to enhanced and adjustable biodegradability for applications such as implants for bone support, stents for artery support in angioplasty, and drug delivery carriers. Although P(3HB-co-3HV) has a 2-fold lower maximum water permeability than poly(lactic acid) which is another biodegradable aliphatic polyester of great biotechnological importance, causing lower hydrolytic degradation due to lower water uptake, the degradation rate of P(3HB-co-3HV)-based biomedical devices are adjustable with molar fraction of 3HV [18,20,21]. Hydrophilic poly(ethylene glycol) and monomethoxy poly(ethylene glycol) can also be incorporated into P(3HB-co-3HV) to form nanoparticles with a hydrophilic outer layer and a hydrophobic inner layer for improved chemical functionalization and compatibility with therapeutic drugs besides benefiting drug release control [22,23,24]. Incorporation of other desired properties for biomedical applications can also be achieved (Table 1).
P(3HB-co-3HV) is a potential substitute for petroleum-based plastic packaging material as it possesses high water and aroma (limonene and linalool) barrier properties while having comparable thermal and mechanical properties to that of polypropylene (PP) and low-density polyethylene (LDPE) [15]. As PP and LDPE are applied extensively for packaging and consumables, which are highly disposable, the substitution with P(3HB-co-3HV) can contribute to reduced stable solid waste creation of petroleum-based plastics [47,48]. Unlike the augmented cytotoxicity by higher 3HV molar fraction, lower 3HV molar fraction causes high stereoregularity, slow crystallization rate, formation of large size spherulites, and secondary crystallization that are discouraging for packaging purposes [24,47,48]. Poly(butylene succinate), poly(butylene adipate-co-terephthalate), natural rubber, or other polymers with plasticizer or toughness properties can be incorporated to overcome the limitations and extend its application as packaging materials (Table 1).
Moreover, PHA-based mulch films are potential substitutes for conventional plastic mulch films. Mulching increases crops productivity, increases horticulture products, prevents water evaporation from the soil, prevents soil erosion, reduces water consumption, and controls weeds [49]. PHA-based mulch films overcome the environmental problems caused by the post-consumption of plastic mulch films made from LDPE, linear low-density polyethylene (LLDPE), and high-density polyethylene (HDPE) due to their poor degradability [50]. Moreover, the physicochemical properties of P(3HB-co-3HV) enable the controlled release of herbicides and insecticides. Herbicides and insecticides can be integrated into P(3HB-co-3HV)-containing pellets and sown along the plantation to be released upon degradation from the pellets depending on the level of pest activity [51,52].
On the other hand, endogenous P(3HB-co-3HV) acts as the electron donor for the denitrification of wastewater in the aquaculture industry. Biomass with PHA-accumulating ability, generally P(3HB) and poly(3-hydroxyvalerate) (P(3HV)), from activated sludge, is employed to remove resulting ammonia from fish excretion and dead animal bodies in circulating water. Unlike the conventional techniques that involve the addition of acetate and ethanol to promote microbial activity, the biomass is precultured for PHA accumulation. The endogenous PHA is used for denitrification that accurately couples with slow metabolic activity in the absence of exogenous carbon source and in the presence of nitrogen [53,54]. The exclusion of volatile fatty acids feeding during the denitrification process prevents the contamination with the dissolved organic carbon that lowers the effluent water quality, and the employment of endogenous PHA is more cost-effective compared to feeding extracted PHA to denitrifying bacteria [55].

3. Bioconversion of Alkyl Alcohols and Organic Acids into P(3HB-co-3HV)

The conversion of organic acid into 3HV starts with β-oxidation, where propionic acid (C3) is converted into propionyl-CoA, whereas valeric acid (C5) is converted into propionyl-CoA and acetyl-CoA, respectively [56]. The 3HV monomer is formed from the resulting propionyl-CoA couples with acetyl-CoA and is polymerized to P(3HB-co-3HV) copolymer with the 3HB monomer. The 3HB monomer is formed from the resulting acetyl-CoA provided majorly by the main carbon source such as oils or sugars (Figure 2) [14,57,58,59,60].
The employment of alkyl alcohols as 3HV precursors is limited to odd carbon number primary alcohols. Primary alcohols are oxidized to aldehydes that can be further oxidized more easily to their respective carboxylic acids. The oxidation processes can occur chemically with the presence of oxidizing agents or biologically with the presence of alcohol dehydrogenase and aldehyde dehydrogenase [61]. Oxidation of secondary alcohols liberates ketones with no further oxidation due to the oxidatively stable nature of ketones [62,63]. Odd carbon number primary alcohols such as 1-propanol or 1-pentanol are oxidized to 1-propanal and 1-pentanal that further oxidized to propanoic acid and valeric acid, respectively. The resulting propionic acid or valeric acid enters β-oxidation to liberate propionyl-CoA for 3HV formation (Figure 2) [14,57,58,59,60].
Although levulinic acid is a cost-effective 3HV precursor, the catabolic pathway involved is undetermined. Generally, levulinic acid catabolism releases intermediates that are converted via β-oxidation to release acetyl-CoA and propionyl-CoA for P(3HB-co-3HV) biosynthesis [64]. Bacteria capable of using levulinic acid as the 3HV precursor are rare and are mainly C. necator, with the exception of Burkholderis sp. IS-01 and Hydrogenophaga pseudoflava DSM 1034 [10,11,65,66,67,68]. C. necator KHB-8862 and H. pseudoflava DSM 1034 are two promising strains reported with a high 3HV yield of 0.50 and 1.00 g/g, respectively. However, other studies reported low PHA content and 3HV yield (Table 2).

4. Techno-Economic and Sustainability Assessment

The annual operating costs in PHA production generally include the direct fixed capital-dependent items, labor-dependent items, administration, and overhead expenses, raw materials, utilities, and downstream processing such as waste management. According to the techno-economic analysis conducted by Choi and Lee (1999) for various pure carbon sources, the substrate cost accounted for 48%–60% of the total costs (Figure 3) [93]. After excluding the trace elements, which are essentials, pure carbon sources that possess high nutritional value such as glucose, glycerol, starch, methane, oils, and volatile fatty acids are commercial products, and their employment leads to higher substrate cost compared to that of industrial or domestic wastes. Due to higher economic advantage and increasing emphasis on sustainability, the employment of wastes as carbon sources is widely attempted. Theoretically, substituting pure substrates with wastes contributes to a huge reduction in raw material expenses. However, pretreatments are needed for certain wastes to remove impurities and toxins or to adjust pH [94]. Pretreatments impose additional costs whereby extra chemicals or equipment are necessary with possible individual optimization. Bhattacharyya and co-workers (2015) reported decreased raw material cost to 39% with the employment of wheat stillage, but the utilities cost increased to 21% as compared to that reported by Choi and Lee (1999) (Figure 3) [93,95].
As opposed to main carbon sources, where numerous studies have been conducted on various wastes, employing wastes as 3HV precursors is not practical due to the composition inconsistency [95]. Due to the necessity of propionyl-CoA for 3HV formation, sole reliance on wastes results in the narrow choice to those with propionate or valerate related components; thus, in most cases, a 3HV precursor is still required to achieve sufficient 3HV fraction for the copolymer to be practically useful [64,95]. This leads to increased raw material cost as propionic acid and valeric acid, which are widely preferred by PHA-producing bacteria, are high-cost precursors (Table 2). The potential of 1-propanol and 1-pentanol as alternatives for propionic acid and valeric acid is well-known but lack practicality due to its high toxicity to the majority of bacteria. Since 1996, several PHA-producing bacteria from different genera have been reported to use 1-propanol or/and 1-pentanol as 3HV precursors (Table 2). The emergence of these bacteria bypasses the bottleneck of precursor dominance by organic acids and enables further innovation in fermentation strategies to develop economically feasible and sustainable production processes. Furthermore, 1-propanol and 1-pentanol are manufactured through well-established oxo synthesis and can be biosynthesized by bacteria from sustainable carbon sources such as glucose, glycerol, and organic wastes, which are abundant in nature.

5. Oxo Synthesis of Alkyl Alcohols

Oxo synthesis is an established process for the manufacture of alkyl alcohols at an industrial scale with simple operational requirements and low specificity in raw materials, including branched-chain, long-chain, and cyclic olefins [96,97]. It is thoroughly investigated for the production of a wide variety of industrial chemicals. The synthesis involves hydroformylation to convert olefins (also known as alkenes) into aldehydes to be further converted into alcohols through hydrogenation. Homogeneous catalysts are employed in hydroformylation, while heterogenous catalysts are employed in hydrogenation for reaction induction. Generally, these reactions are carried out in separate reactors where the resulting aldehydes from the primary reactor are transferred into the second reactor to be hydrogenated. Catalysts and carbon monoxide in the primary reactor are removed either by decobalting or been recycled back to the primary reactor to prevent entry into the second reactor as a precautious measure to extend the shelf life of hydrogenation catalysts. Recycling the catalysts contributes to high economic feasibility as high-cost catalysts such as rhodium-based catalysts can be reused for subsequent batches. However, an 8–55% decrease in catalyst yield after repeated recycling is expected [98]. The resulting alcohols are purified from the mixture via distillation (Figure 4) [97].

5.1. 1-Propanol

Oxo synthesis of 1-propanol begins with the rhodium-catalyzed hydroformylation of ethylene (also known as ethene) to propanal with the aid of rhodium–triphenylphosphine catalysts. The resulting 1-propanal is distilled from the catalyst-containing solution, and carbon monoxide is removed. Hydrogenation can be carried out in either the heterogenous vapor phase or the heterogenous liquid phase. Heterogeneous vapor phase hydrogenation takes place at 110–150 °C and 0.14–1.00 MPa with the aid of copper, zinc, nickel, and chromium catalysts supported on alumina (CAS:1344-28-1) or kieselguhr (CAS:91053-39-3) [62]. Heat is removed either by an external heat exchange device or an internal cooler [100]. This process produces impurities such as dipropyl ether, ethane, and propyl propionate. Selectivity enhancers such as alkali or transition metals are added to reduce the formation of esters, while an additional 1%–10% water could suppress the formation of ether [62,101]. Propyl propionate is separated from the product mixture and hydrogenolyzed with the aid of reduced CuO–ZnO catalysts at 75–300 °C and 9.8 kPa–9.8 MPa to produce 1-propanol as the major product [62]. Heterologous liquid phase hydrogenation involved nickel or copper catalysts at a lower temperature of 95–120 °C and a higher pressure of 3.5 MPa. Crude 1-propanol is purified via distillation with the aid of an azeotroping agent such as dipropyl ether or cyclo-hexane to remove water for highly pure 1-propanol yield (>99%) (Figure 4) [62,102].

5.2. 1-Pentanol

Oxo synthesis of 1-pentanol begins with hydroformylation of 1-butene. Subsequent hydrogenation yields two C5 products that are 1-pentanol and 2-methyl-1-butanol. For cobalt-catalyzed hydroformylation, the ratio of the product is 7:3 (1-pentanol:2-methyl-1-butanol) after subsequent hydrogenation. When rhodium–triphenylphosphane is employed instead, a higher yield of 1-pentanol is achieved with a 9:1 (1-pentanol:2-methyl-1-butanol) ratio (Figure 4) [99].

6. Biosynthesis of 1-Propanol and 1-Pentanol by Wild-Type Bacteria

6.1. The Wood–Werkman Pathway in Propionibacteria

Biosynthesis of 1-propanol by wild-type bacteria is inefficient as 1-propanol is synthesized as a byproduct through propionic acid synthesis processes. Propionibacteria such as Propionibacterium acidipropionici and Propionibacterium freudenreichii are able to produce 1-propanol through the Wood–Werkman pathway (also known as the dicarboxylic pathway, or the methylmalonyl-CoA pathway). The synthesis process requires an anaerobic condition where the carbon source is converted into pyruvate and enters the Wood–Werkman pathway to produce propionic acid as the main product [103,104]. The 1-propanol yield reported was in the range of 0.04–0.14 mol/mol, equivalent to 0.6–1.8 g/L. The 1-propanol production was found to be higher when glycerol was employed, compared to glucose [105,106]. The precise processes involved in 1-propanol formation are undetermined but could probably be by two-step reduction from propionyl-CoA to 1-propanol aided by acylating propionaldehyde dehydrogenase and propanol dehydrogenase (Figure 5) [107].

6.2. The Acrylate Pathway in Clostridium

Clostridium propionicumable and Clostridium neopropionicumable are able to use amino acids (alanine and serine), lactate, and ethanol as growth-promoting substances under anaerobic conditions [108,109]. C. neopropionicumable synthesizes a small amount of 1-propanol (0.06 g/L, 0.03 mol/mol) from ethanol with propionate and acetate as the main products [109]. By employing the bacterial mixture dominated by Alkalibaculum bacchi (34%) and C. propionicumable (54%), C. propionicumable produced 6.0 g/L 1-propanol and 1.0 g/L 1-butanol, whereas A. bacchi produced 8.0 g/L ethanol from syngas (the carbon source) and corn-steep liquor (the source of amino acids and minerals) [110]. The resulting 1-propanol was proposed to be the product from a two-step reduction in propionyl-CoA produced through the acrylate pathway by using the lactoyl-CoA that is not used for propionic acid synthesis (Figure 6) [109]. However, further experimentations are needed to provide essential information for a complete view of the biosynthesis pathway.

6.3. The Carboxylate Reduction Pathway in Clostridium

Anaerobic digestion by microbial consortia is a promising hydrogen production process where the members in the microbial community play different roles to convert raw materials into hydrogen under anaerobic conditions. As sterilization is commonly excluded from anaerobic digestion, organic acids produced by acetogens in the consortia cause decreased pH that disrupts the metabolic activity of hydrogen-producing bacteria [111]. Clostridium ragsdalei (ATCC BAA-622, DSM 15248) is an acetogen capable of synthesizing alcohols by ferredoxin-mediated carboxylate reduction. With the involvement of exogenous CO and ferredoxin, n-fatty acids up to six carbons in length can be reduced to corresponding alcohols (Figure 7). The concentration of produced 1-propanol reported was 1.7 g/L 1-propanol from propionic acid, with a conversion efficiency of 97%. However, the concentration of 1-pentanol obtained was merely 0.2 g/L, with a conversion efficiency of 82% [112].

7. Biosynthesis of 1-Propanol and 1-Pentanol by Genetic-Engineered E. coli

7.1. Co-Expression of the Citramalate and Threonine Pathway

Numerous genetic engineering attempts were carried out for alkyl alcohols biosynthesis through the individual threonine or citramalate pathway and showed successful biosynthesis of 1-propanol from the intermediate 2-ketobutyrate in the pathways [113,114]. For greater industrial applicability, co-expression of both pathways was attempted in E. coli BW25113. The simultaneous operation of the pathways in a single host showed a synergic effect on 1-propanol production. The co-expression provided a larger 2-ketobutyrate pool for decarboxylation and reduction to 1-propanol (Figure 8). A high 1-propanol concentration of 8.0 g/L was reported with a 1-propanol yield of 0.15 g/g from glucose, which was higher than 0.09 and 0.11 g/g for individual threonine and citramalate pathway, respectively [115].

7.2. Interactive Elongation Cycle of 2-Ketoacids

Biosynthesis of 1-pentanol was made possible by introducing Lactococcus lactis ketoisovalerate decarboxylase (Kivd) modified via saturated mutagenesis of the V461 key residue of the enzyme with glycine and serine into E. coli BW25113 to promote its selectivity toward 2-ketocaproate, which is the precursor for 1-pentanol. Besides lowered catalytic efficiency of the modified Kivd toward 2-ketoacids upstream of 2-ketocaproante, the increased supply of acetyl-CoA by acetate feeding encouraged 2-ketoacid elongation cycle for enhanced 1-pentanol production (Figure 9). The high specificity of this approach was implied by 90% 1-pentanol in the alcohol product mixture, equivalent to 2.2–2.4 g/L upon production harvest. The synthesis of alcohols with a longer alkyl chain was found to be minimized as further elongation of the 2-ketoacid was discouraged due to the active use of 2-kerocaproate for 1-pentanol synthesis [116].

7.3. Extended Dissimilation of Succinate

The sleeping beauty mutase (SBM) operon in E. coli is a four-gene operon (sbm-ygfD-ygfG-ygfH) that encodes various enzymes required in a cobalamin-dependent metabolic pathway for decarboxylation of succinate into propionate [117]. An activated chromosomal SBM operon encodes methylmalonyl-mutase (by sbm), methylmalonyl-CoA decarboxylase (by ygfG), and propionyl-CoA:succinate CoA transferase (by ygfH) in plasmid-free propanogenic E. coli BW25113 enabled extended dissimilation of succinate to synthesis 1-propanol (Figure 10). Glycerol favored solventogenesis over glucose due to the necessity of a solventogenic pathway as an auxiliary channel for redox balance upon glycerol dissimilation under anaerobic conditions. An anaerobic fed-batch strategy established by using the engineered E. coli strain produced high titers of 7.0 g/L 1-propanol, thus implying its high industrial applicability [118].

7.4. Acquired Carboxylate Reduction Pathway

Conversion of organic acids produced by acetogens during anaerobic digestion into other useful products is suggested to be beneficial as a solution to maintain the stability of the biogas production process. An E. coli BL21(DE3) strain harboring Clostridium acetobutylicum alcohol dehydrogenase (AdhE2) and Megasphaera hexanoica acyl-CoA transferase (ACT01_02765) was developed for conversion of the C2-C8 organic acids commonly found in anaerobic digestion into corresponding primary alcohols. The metabolic pathway is relatively simpler as it only involves two steps aided by two enzymes (Figure 11). Following the conversion rate of 1.1 for C4 acid into 1-butanol, the functional alcohol dehydrogenase and acyl-CoA transferase resulted in a promising conversion rate of 0.8 for both 1-propanol and 1-pentanol [119].

8. Alkyl Alcohol-Tolerant P(3HB-co-3HV)-Producing Bacteria

Alcohols are unsuitable to be employed as 3HV precursors for C. necator (also known as Ralstonia eutropha, Alcaligenes eutrophus, or Wautersia eutropha), which is the standard PHA-producing bacteria. Although C. necator H16 is capable of surviving methanol, ethanol, and propanol, extensive exposure to these alcohols is detrimental to PHA accumulation, thus resulting in lower biomass. The employment of 8.0 g/L 1-propanol, which is convertible into propionyl-CoA, contributed to merely 3 mol% 3HV with a 3HV yield of 0.14 g/g [13,120] (Table 2). The individual employment of 1-propanol and 1-pentanol also caused a remarkably high reduction in biomass and PHA content of C. necator DSM 545. The employment of 1-pentanol caused C. necator DSM 545 biomass and PHA content to decrease by 40% and 20%, respectively. Comparatively,1-propanol exerted a lower adverse effect compared to 1-pentanol, whereby its employment decreased C. necator DSM 545 PHA content by 10% with no negative influence on bacterial biomass [10] (Table 2). To overcome the limitation in 3HV precursor selection, isolation of alkyl alcohol-tolerant P(3HB-co-3HV)-producing bacteria is continuously attempted and has led to the discovery of various promising bacteria with the capability to use alkyl alcohols as 3HV precursors (Figure 12).
P. denitrificans ATCC 17741 was the first bacteria reported in 1996 for the use of alkyl alcohol as the 3HV precursor. P. denitrificans ATCC 17741 is a mixotrophic colorless sulfur bacterium capable of using 1-pentanol as the sole carbon source for growth and P(3HV) accumulation [14,121]. The study was conducted by maintaining the concentration of 1-pentanol at 1.6 g/L for 24 h. Approximately 6.8 g/L biomass with 1.2 g/L P(3HV) homopolymer was achieved [14] (Table 2).
Erwinia sp. USMI-20 was reported with its preference for alkyl alcohols instead of organic acids as 3HV precursors. Erwinia sp. USMI-20 achieved higher biomass with the co-employment of 1-propanol and 1-pentanol compared to that when palm oil was employed solely. A higher PHA content of 50 and 62 wt% was also achieved for 1-propanol and 1-pentanol, respectively, compared to 40 wt% and 34 wt% for propionic acid and valeric acid. 1-pentanol was more promising compared to 1-propanol as Erwinia sp. USMI-20 accumulated a higher 3HV fraction of 20 mol% from 1-pentanol compared to 6 mol% from 1-propanol. 1-pentanol can be employed as a substitute for valeric for Erwinia sp. USMI-20 owing to the higher 3HV yield of 0.43 g/g for 1-pentanol, which was 2-fold higher than that for valeric acid [60]. The production was scaled up to 10 L by employing 4.6 g/L palm oil and 1.4 g/L 1-pentanol, where 1-pentanol was added at 20 h post incubation. The 3HV fraction achieved was 20 mol% in 56 wt% PHA content of 5.4 g/L biomass, with 0.43 g/g 3HV yield [122] (Table 2).
Despite the negative influence observed for C. necator, there are several Cupriavidus sp. that are capable of using alkyl alcohols as 3HV precursors with no adverse effect on either bacterial biomass or PHA accumulation. C. malaysiensis USMAA2-4, C. malaysiensis USMAA1020, and C. malaysiensis USMAA9-39 are three PHA-producing bacteria favoring alkyl alcohols over organic acids for 3HV formation [123]. C. malaysiensis USMAA2-4 and C. malaysiensis USMAA1020 were able to accumulate 7–10 mol% 3HV from 1-pentanol [87,88,123,124,125] (Table 1). The 3HV yield of C. malaysiensis USMAA2-4 and its transformant strain harboring C. necator H16 lipAB genes was 0.22 g/g and 0.33 g/g, respectively, which were both higher than 0.14 g/g for C. necator H16 [13,87,88]. The sole employment of 1-pentanol resulted in a higher C. malaysiensis USMAA9-39 PHA content of 46 wt% compared to 37 wt% for valeric acid. Despite the 1-fold lower C. malaysiensis USMAA9-39 biomass resulting from the co-employment of 1-pentanol with oleic acid, the 3HV yield of 0.44 g/g from 1-pentanol was comparable to 0.42 g/g from valeric acid and a high 3HV composition of 24 mol% was achieved [76] (Table 2).
M. extorquens G10 demonstrated the production of P(3HB-co-3HV) from an alkyl alcohol mixture of C1 and C5 alcohol. A 4 L production of P(3HB-co-3HV) from a methanol-pentanol mixture by M. extorquens G10 showed a promisingly high PHA concentration of 7.5–18.0 g/L. The carbon mixture was supplemented fractionally based on the dissolved oxygen peaks observed. With an increased portion of 1-pentanol from 2 to 20 mol%, the biomass decreased with association to reduction in PHA content from 40.0 to 25.0 g/L and 45 to 30 wt%, respectively. Despite the negative influence on biomass and PHA content, 3HV composition of 14–50 mol% was achieved [90] (Table 2).
M. halotolerans C2 demonstrated P(3HB-co-3HV) production from C1, C2, and C5 alkyl alcohol. P(3HB-co-3HV) production by M. halotolerans C2 through fractional feeding of methanol-ethanol mixture resulted in increased 3HV composition from 2 to 51 mol% parallel to increased 1-pentanol supply from 5 to 15 % v/v methanol. A considerably high PHA content of 73–98 wt% was accumulated by the bacterium [91] (Table 2).
P(3HB-co-3HV) production by Methylocystis sp. WRRC1 from methane and 1-pentanol demonstrated a 0.17 g/g 3HV yield from 1.0 g/L 1-pentanol. The 6-fold lower consumption of methane by the bacteria with the co-employment of 1-pentanol compared to that of sole employment of methane denoted the preference of the bacteria for 1-pentanol over methane. However, 1-pentanol is non-competitive against valerate where Methylocystis sp. WRRC1 achieved a 1-fold higher 3HV concentration with the co-employment of sodium valerate compared to that of 1-pentanol. On the other hand, the co-employment of sodium valerate did not cause reduced methane consumption and contributed to a higher 3HB concentration [86] (Table 2).
M. haematophila (also known as Naxibacter haematophila) UMTKB-2, a slow-growing bacterium, was also reported with the capability to use 1-pentanol for 3HV accumulation with a preference for 1-pentanol over valeric acid and sodium valerate. The co-employment of 1-pentanol resulted in 2-fold and 11-fold higher biomass and PHA content compared to that of valeric acid and sodium valerate, respectively. Upon optimization by using response surface methodology, M. haematophila UMTKB-2 achieved 7 mol% 3HV with 0.40 g/g 3HV yield. Unlike the PHA accumulation process of Cupriavidus sp. that ends within 48–72 h, 122 h was needed for optimum P(3HB-co-3HV) accumulation by M. haematophila UMTKB-2 [89] (Table 2).

9. Mode of Action of 1-Propanol and 1-Pentanol on Proteins

Short-chain alcohols exert a hydrophobic effect by interacting with proteins and lead to the structural unfolding of the protein [126]. Changes in membrane fluidity ensue due to the direct insertion of lipophilic agents into the cellular membrane after direct physicochemical interaction with alcohols. This induces adaptive membrane alteration by changing the fatty acid composition of the membrane [127]. Impaired inner membrane integrity associated with depletion in proton motive force due to the increased proton motive force demand for chemical, osmotic and mechanical adjustment induces the psp operon to prevent proton loss. As a result, the cells experience a metabolic shift to anaerobic respiration together with downregulation of motility for adjustment and maintenance of energy as well as for proton motive force usage [128]. The extent of water exclusion is greater with increasing alkyl groups of the alcohol, which is non-polar. By considering the hydrophobic effect of methanol < ethanol < propanol = butanol, pentanol may exert a similar hydrophobic effect on protein and result in pentanol-induced protein unfolding [126]. Furthermore, pentanol is capable of inactivating membrane proteins such as transporters but rarely causes structural changes to the cell membrane [129].

10. Mechanisms Involved in Alcohols Tolerance

Aliphatic alcohols, aromatic compounds, or other organic solvents are toxic to bacteria when present in high concentrations. Nevertheless, certain bacteria are able to thrive in the high concentration of such toxic organic chemicals. Bacterial solvent tolerance is a multifactorial process that involves gene expression and subsequent physiological changes to respond to stress conditions. Extrusion of the toxic compounds from the cell to the external environment and reduced cell membrane permeability to prevent further influx of toxic compounds are the relevant mechanisms to survive alcohol stress.

10.1. Changes in the Cell Membrane

Alcohol-induced cell leakage of magnesium and nucleotides is the primary damaging action that affects bacterial viability in alcohols [130]. As alcohols interact with the cell membrane and decrease the degree of membrane organization, proteins that participate in membrane structure organization and surface stabilization are critical in alcohol tolerance [131]. Isomerase incorporates fatty acids into the phospholipid headgroups of the phospholipid bilayer and causes isomerization of cis unsaturated fatty acids to trans unsaturated fatty acids to form a denser membrane, as demonstrated by Pseudomonas and Vibiro [132]. Changes in cell membrane composition that attributed to increased cis-11 vaccenic acid (18:1) or cis-9 oleic acid (18:1) with a corresponding decrease in palmitic acid (16:0) were demonstrated for E. coli, Lactobacillus homohiochii, and Saccharomyces cerevisiae [131,133,134]. The synthesis of phosphatidylethanolamine by Zymomonas mobilis was partially inhibited in the presence of alcohols. As a result, a membrane with an elevated proportion of acidic phospholipids (phosphatidylglycerol and cardiolipin) and an overall reduction in the phospholipid:protein ratio is synthesized, thus increasing the efficiency of efflux pumps in alcohol extrusion [135,136].

10.2. Stress Response System

Exposure to alkyl alcohols leads to changes in the level of expression of certain genes as responses to stress for adaptation. As demonstrated in E. coli, exposure to 1-butanol causes downregulation of several genes related to histidine, leucine, arginine, tryptophan, and methionine biosynthesis and transport, thus leading to a significantly lower level of related proteins. Downregulation of genes related to amino acids metabolism is an indicator for bacterial growth inhibition in alcohols. As opposed to that, opp operon (oppABCDF) that encodes the components in a polyamine-induced oligopeptide ABC transport system is upregulated for the transport of hydrophilic substances to compensate for the hydrophobic pressure exerted by alcohols [137,138]. Genes responsible for response to heat shock and extracytoplasmic stress (cpx regulon) are upregulated, and periplasmic chaperone Spy is encoded to respond to protein misfolding activity [139,140,141,142]. Increased isobutanol tolerance of C. acetobutylicum is also conferred to overexpression of genes related to heat shock [139,143]. Genes related to the membrane and periplasmic space carbohydrate transport and metabolisms are upregulated to transport and phosphorylate hexoses and release the phosphate esters into the cytoplasm, probably as a repair mechanism for damaged bilayer [139,144]. Furthermore, the upregulation of genes from the 13-member nuo operon and 5-member cyo operon is also an indicator for the increased requirement of energy or disruption of respiratory efficiency upon exposure to 1-butanol [139]. However, the operons are downregulated when exposed to isobutanol [137]. Exposure to ethanol causes induced expression of psp operon to restore proton motive force lost due to disruption of the cell membrane by ethanol, but the expression level remains unchanged for isobutanol [128,137].

11. Challenges in Wide Implementation of Alkyl Alcohols as 3HV Precursors

Low alcohol tolerance due to alcohol toxicity is the major drawback for the employment of alkyl alcohols as 3HV precursors. Isolation of novel PHA-producing bacteria with substantial alcohol tolerance is a continuous effort in developing production processes with higher economic feasibility. With established primary alkyl alcohol bioproduction processes, the employment of alkyl alcohols also contributes to sustainability. Alcohol tolerance involves complex regulatory systems, and knowledge from cell-wide stress response is still in demand. Theoretically, genetic engineering can be adopted to create an alkyl alcohol-tolerant PHA-producing bacteria by either introducing pha genes into an alkyl alcohol-tolerant host or modulating alcohol tolerance of a non-alkyl alcohol-tolerant PHA-producing bacteria. Comparatively, the former approach is more rational as alcohol tolerance involves complex systems and is not economically feasible for commercial importance.
Although genetic-engineered E. coli with mutated rpoA gene was constructed successfully to produce products with commercial importance such as 1-butanol, the attempt was based on extensive studies on the rpoA gene and its roles in phenotypic changes of E. coli [145,146,147,148,149]. Owing to numerous studies on the incorporation of pha genes into E. coli, which demonstrated successful production of various PHA, such approaches can be adopted for the construction of alkyl alcohol-tolerant strains with acquired PHA-producing ability [150,151,152]. However, a candidate strain with broad substrate preference is preferred for production process establishment with different substrates and fermentation strategies. The capability to use wastes with high carbon content will be an added value for higher industrial applicability owing to its sustainability and higher economic feasibility compared to pure carbon sources [87].
Despite the promising potential shown by the known alkyl alcohol-tolerant P(3HB-co-3HV)-producing bacteria, scaling up the production remains challenging. As low 3HV compositions are commonly reported for shake flask scale production, various production strategies have to be adopted to increase the molar fraction of 3HV. Fed-batch production strategies that enable the addition of alkyl alcohols eventually are practically preferred to achieve high 3HV composition of P(3HB-co-3HV) and at the same time minimize the negative influences caused by the relative toxicity of alkyl alcohol. However, some of the bacteria that depicted decreased biomass and PHA content with the employment of alkyl alcohol at low concentration or with a preference for organic acid sodium salt over alkyl alcohol have low applicability as candidate P(3HB-co-3HV) producers when alkyl alcohols are to be employed. In addition, more studies on large-scale P(3HB-co-3HV) production involving alkyl alcohols are still in demand to compare their industrial practicality as alternative 3HV precursors for organic acids in terms of sustainability and economic feasibility.

12. Concluding Remark

The high sale price of P(3HB-co-3HV) has been the major obstacle to commercialization. Although various carbon sources have been explored, limited precursor choice due to the domination by propionate and valerate has caused the development of diverse P(3HB-co-3HV) production to reach a bottleneck. With increasing studies reporting the discovery of alkyl alcohol-utilizing PHA-producing bacteria with promising bioconversion efficiency of 1-propanol and 1-pentanol into 3HV, the toxicity of alkyl alcohols and low 3HV yield are no longer the major concern. Future attempts should focus on continuous searching of alkyl alcohols tolerant PHA-producing bacteria to discover more promising wild-type strains. Moreover, genetic engineering of bacterial metabolic pathways to achieve successful or higher bioconversion rate of alkyl alcohols into 3HV is also important to overcome low bacterial viability and alcohol-3HV bioconversion efficiency. However, more studies are required for techno-economic assessment to compare to what extent 1-propanol and 1-pentanol could contribute to higher economic feasibility than propionate and valerate.

Author Contributions

Conceptualization, investigation, data curation, formal analysis, visualization, writing—original draft, review and editing, H.S.J.W.; writing—review and editing, K.B.; supervision, writing—review and editing, A.-A.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Grant Scheme (FRGS) (ref: FRGS/1/2021/STG01/USM/01/2) from the Ministry of Higher Education, Malaysia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest to disclose.

Abbreviations

3HB3-hydroxybutyrate
3HV3-hydroxyvalerate
α-P(3HB)Synthetic atactic poly(3-hydroxybutyrate)
Ag/BSABovine serum albumin capped silver
ASAscorbic acid
CNCCellulose nanocrystals
CNTCarbon nanotubes
DCPDicumyl peroxide
HAHydroxyapatite
HDPEHigh-density polyethylene
LDPELow-density polyethylene
LLDPELinear low-density polyethylene
MATOrganophilic attapulgite
mPEGMonomethoxy poly(ethylene glycol)
NANot available
P(3HB)Poly(3-hydroxybutyrate)
P(3HV)Poly(3-hydroxyvalerate)
P(3HB-co-3HV)Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
PBATPoly(butylene adipate-co-terephthalate)
PBSPoly(butylene succinate)
PCLPoly(ε-caprolactone)
PDLLAPoly(d,l-lactide)
PEGPoly(ethylene glycol)
PHAPolyhydroxyalkanoates
PHEMAPoly(2-hydroxyl ethyl methacrylate)
PLAPoly(lactic acid)
PPPolypropylene
PPCPoly(propylene carbonate)
Ref.References
SBMSleeping beauty mutase

References

  1. Pellis, A.; Malinconico, M.; Guarneri, A.; Gardossi, L. Renewable polymers and plastics: Performance beyond the green. New Biotechnol. 2020, 60, 146–158. [Google Scholar] [CrossRef] [PubMed]
  2. Braunegg, G.; Lefebvre, G.; Genser, K.F. Polyhydroxyalkanoates, biopolyesters from renewable resources: Physiological and engineering aspects. J. Biotechnol. 1998, 65, 127–161. [Google Scholar] [CrossRef]
  3. Policastro, G.; Panico, A.; Fabbricino, M. Improving biological production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) co-polymer: A critical review. Rev. Environ. Sci. Biotechnol. 2021, 1–35. [Google Scholar] [CrossRef]
  4. Taguchi, S.; Iwata, T.; Abe, H.; Doi, Y. 9.09-Poly(hydroxyalkanoate)s. In Polymer Science: A Comprehensive Reference; Matyjaszewski, K., Möller, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; pp. 157–182. [Google Scholar] [CrossRef]
  5. Alvarez, H.M.; Kalscheuer, R.; Steinbüchel, A. Accumulation of storage lipids in species of Rhodococcus and Nocardia and effect of inhibitors and polyethylene glycol. Lipid/Fett 1997, 99, 239–246. [Google Scholar] [CrossRef]
  6. Anderson, A.J.; Williams, D.R.; Dawes, E.A.; Ewing, D.F. Biosynthesis of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) in Rhodococcus ruber. Can. J. Microbiol. 1995, 41, 4–13. [Google Scholar] [CrossRef]
  7. Haywood, G.W.; Anderson, A.J.; Williams, D.R.; Dawes, E.A.; Ewing, D.F. Accumulation of a poly(hydroxyalkanoate) copolymer containing primarily 3-hydroxyvalerate from simple carbohydrate substrates by Rhodococcus sp. NCIMB 40126. Int. J. Biol. Macromol. 1991, 13, 83–88. [Google Scholar] [CrossRef]
  8. Valentin, H.F.; Dennis, D. Metabolic pathway for poly(3-hydroxybutyrate-co-3-hydroxyvalerate) formation in Nocardia corallina: Inactivation of mutB by chromosomal integration of a kanamycin resistance gene. Appl. Environ. Microbiol. 1996, 62, 372–379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Williams, D.R.; Anderson, A.J.; Dawes, E.A.; Ewing, D.F. Production of a co-polyester of 3-hydroxybutyric acid and 3-hydroxyvaleric acid from succinic acid by Rhodococcus ruber: Biosynthetic considerations. Appl. Microbiol. Biotechnol. 1994, 40, 717–723. [Google Scholar] [CrossRef]
  10. Berezina, N. Enhancing the 3-hydroxyvalerate component in bioplastic PHBV production by Cupriavidus necator. Biotechnol. J. 2012, 7, 304–309. [Google Scholar] [CrossRef]
  11. Novackova, I.; Kucera, D.; Porizka, J.; Pernicova, I.; Sedlacek, P.; Koller, M.; Kovalcik, A.; Obruca, S. Adaptation of Cupriavidus necator to levulinic acid for enhanced production of P(3HB-co-3HV) copolyesters. Biochem. Eng. J. 2019, 151, 107350–107360. [Google Scholar] [CrossRef]
  12. Scully, S.M.; Orlygsson, J. Biological production of alcohols. In Advanced Bioprocessing for Alternative Fuels, Biobased Chemicals, and Bioproducts; Woodhead Publishing: Cambridge, MA, USA, 2019; pp. 83–108. [Google Scholar]
  13. Obruca, S.; Marova, I.; Snajdar, O.; Mravcova, L.; Svoboda, Z. Production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Cupriavidus necator from waste rapeseed oil using propanol as a precursor of 3-hydroxyvalerate. Biotechnol. Lett. 2010, 32, 1925–1932. [Google Scholar] [CrossRef]
  14. Yamane, T.; Chen, X.; Ueda, S. Growth-associated production of poly(3-hydroxyvalerate) from n-pentanol by a methylotrophic bacterium, Paracoccus denitrificans. Appl. Environ. Microbiol. 1996, 62, 380–384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Strong, P.J.; Laycock, B.; Mahamud, S.N.S.; Jensen, P.D.; Lant, P.A.; Tyson, G.; Pratt, S. The opportunity for high-performance biomaterials from methane. Microorganisms 2016, 4, 11. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Q.; Zhang, H.; Deng, B.; Zhao, X. Poly(3-hydroxybutyrate) and poly(3-hydroxybutyrate-co-3-hydroxyvalerate): Structure, property, and fiber. Int. J. Polym. Sci. 2014, 2014, 1–11. [Google Scholar] [CrossRef]
  17. Shishatskaya, E.I.; Kamendov, I.V.; Starosvetsky, S.I.; Vinnik, Y.S.; Markelova, N.N.; Shageev, A.A.; Khorzhevsky, V.A.; Peryanova, O.V.; Shumilova, A.A. An in vivo study of osteoplastic properties of resorbable poly-3-hydroxybutyrate in models of segmental osteotomy and chronic osteomyelitis. Artif. Cells, Nanomed. Biotechnol. 2014, 42, 344–355. [Google Scholar] [CrossRef] [PubMed]
  18. Naser, A.Z.; Deiab, I.; Darras, B.M. Poly(lactic acid)(PLA) and polyhydroxyalkanoates (PHAs), green alternatives to petroleum-based plastics: A review. RSC Adv. 2021, 11, 17151–17196. [Google Scholar] [CrossRef]
  19. Laycock, B.; Halley, P.; Pratt, S.; Werker, A.; Lant, P. The chemomechanical properties of microbial polyhydroxyalkanoates. Prog. Polym. Sci. 2013, 38, 536–583. [Google Scholar] [CrossRef]
  20. Ali, I.; Jamil, N. Polyhydroxyalkanoates: Current applications in the medical field. Front. Biol. 2016, 11, 19–27. [Google Scholar] [CrossRef]
  21. Wu, L.P.; Wang, D.; Parhamifar, L.; Hall, A.; Chen, G.Q.; Moghimi, S.M. Poly(3-hydroxybutyrate-co-R-3-hydroxyhexanoate) nanoparticles with polyethylenimine coat as simple, safe, and versatile vehicles for cell targeting: Population characteristics, cell uptake, and intracellular trafficking. Adv. Healthc. Mater. 2014, 3, 817–824. [Google Scholar] [CrossRef]
  22. Catoni, S.E.; Trindade, K.N.; Gomes, C.A.; Schneider, A.L.; Pezzin, A.; Soldi, V. Influence of poly(ethylene grycol)-(PEG) on the properties of influence of poly(3-hydroxybutyrate-co-3-hydroxyvalerate)-PHBV. Polímeros 2013, 23, 320–325. [Google Scholar] [CrossRef] [Green Version]
  23. Li, Z.; Loh, X.J. Water soluble polyhydroxyalkanoates: Future materials for therapeutic applications. Chem. Soc. Rev. 2015, 44, 2865–2879. [Google Scholar] [CrossRef]
  24. Shah, M.; Naseer, M.I.; Choi, M.H.; Kim, M.O.; Yoon, S.C. Amphiphilic PHA–mPEG copolymeric nanocontainers for drug delivery: Preparation, characterization and in vitro evaluation. Int. J. Pharm. 2010, 400, 165–175. [Google Scholar] [CrossRef] [PubMed]
  25. Scandola, M.; Focarete, M.L.; Adamus, G.; Sikorska, W.; Baranowska, I.; Świerczek, S.; Gnatowski, M.; Kowalczuk, M.; Jedliński, Z. Polymer blends of natural poly(3-hydroxybutyrate-co-3-hydroxyvalerate) and a synthetic atactic poly(3-hydroxybutyrate). Characterization and biodegradation studies. Macromolecules 1997, 30, 2568–2574. [Google Scholar] [CrossRef]
  26. Bhatia, S.K.; Wadhwa, P.; Hong, J.W.; Hong, Y.G.; Jeon, J.M.; Lee, E.S.; Yang, Y.H. Lipase mediated functionalization of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) with ascorbic acid into an antioxidant active biomaterial. Int. J. Biol. Macromol. 2019, 123, 117–123. [Google Scholar] [CrossRef]
  27. Malmir, S.; Montero, B.; Rico, M.; Barral, L.; Bouza, R. Morphology, thermal and barrier properties of biodegradable films of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) containing cellulose nanocrystals. Compos. A Appl. Sci. Manuf. 2017, 93, 41–48. [Google Scholar] [CrossRef]
  28. Meereboer, K.W.; Pal, A.K.; Cisneros-López, E.O.; Misra, M.; Mohanty, A.K. The effect of natural fillers on the marine biodegradation behaviour of poly(3-hydroxybutyrate-co-3-hydroxyvalerate)(PHBV). Sci. Rep. 2021, 11, 1–11. [Google Scholar] [CrossRef]
  29. Figueroa-Lopez, K.J.; Cabedo, L.; Lagaron, J.M.; Torres-Giner, S. Development of electrospun poly(3-hydroxybutyrate-co-3-hydroxyvalerate) monolayers containing eugenol and their application in multilayer antimicrobial food packaging. Front. Nutr. 2020, 7, 140–155. [Google Scholar] [CrossRef]
  30. Galego, N.; Rozsa, C.; Sánchez, R.; Fung, J.; Vázquez, A.; Santo Tomás, J. Characterization and application of poly(β-hydroxyalkanoates) family as composite biomaterials. Polym. Test. 2000, 19, 485–492. [Google Scholar] [CrossRef]
  31. Thiré, R.M.D.S.M.; Arruda, L.C.; Barreto, L.S. Morphology and thermal properties of poly(3-hydroxybutyrate-co-3-hydroxyvalerate)/attapulgite nanocomposites. Mater. Res. 2011, 14, 340–344. [Google Scholar] [CrossRef] [Green Version]
  32. Kwiecien, I.; Adamus, G.; Jiang, G.; Radecka, I.; Baldwin, T.C.; Khan, H.R.; Johnston, B.; Pennetta, V.; Hill, D.; Bretz, I.; et al. Biodegradable PBAT/PLA blend with bioactive MCPA-PHBV conjugate suppresses weed growth. Biomacromolecules 2018, 19, 511–520. [Google Scholar] [CrossRef]
  33. Bakare, R.A.; Bhan, C.; Raghavan, D. Synthesis and characterization of collagen grafted poly(hydroxybutyrate–valerate)(PHBV) scaffold for loading of bovine serum albumin capped silver (Ag/BSA) nanoparticles in the potential use of tissue engineering application. Biomacromolecules 2014, 15, 423–435. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, X.; Cornish, K.; Vodovotz, Y. Synergistic mechanisms underlie the peroxide and coagent improvement of natural-rubber-toughened poly(3-hydroxybutyrate-co-3-hydroxyvalerate) mechanical performance. Polymers 2019, 11, 565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Javadi, A.; Kramschuster, A.J.; Pilla, S.; Lee, J.; Gong, S.; Turng, L.S. Processing and characterization of microcellular PHBV/PBAT blends. Poly. Eng. Sci. 2010, 50, 1440–1448. [Google Scholar] [CrossRef]
  36. Qiu, Z.; Ikehara, T.; Nishi, T. Miscibility and crystallization behaviour of biodegradable blends of two aliphatic polyesters. Poly(3-hydroxybutyrate-co-hydroxyvalerate) and poly(butylene succinate) blends. Polymer 2003, 44, 7519–7527. [Google Scholar] [CrossRef]
  37. Ma, P.; Hristova-Bogaerds, D.G.; Lemstra, P.J.; Zhang, Y.; Wang, S. Toughening of PHBV/PBS and PHB/PBS blends via in situ compatibilization using dicumyl peroxide as a free-radical grafting initiator. Macromol. Mater. Eng. 2012, 297, 402–410. [Google Scholar] [CrossRef] [Green Version]
  38. Chun, Y.S.; Kim, W.N. Thermal properties of poly(hydroxybutyrate-co-hydroxyvalerate) and poly(ε-caprolactone) blends. Polymer 2000, 41, 2305–2308. [Google Scholar] [CrossRef]
  39. Wang, S.; Ma, P.; Wang, R.; Wang, S.; Zhang, Y.; Zhang, Y. Mechanical, thermal and degradation properties of poly(d,l-lactide)/poly(hydroxybutyrate-co-hydroxyvalerate)/poly(ethylene glycol) blend. Polym. Degrad. Stab. 2008, 93, 1364–1369. [Google Scholar] [CrossRef]
  40. Pillai, A.B.; Kumar, A.J.; Kumarapillai, H. Biosynthesis of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) in Bacillus aryabhattai and cytotoxicity evaluation of PHBV/poly(ethylene glycol) blends. 3 Biotech 2020, 10, 1–10. [Google Scholar] [CrossRef]
  41. Silva, A.P.B.; Montagna, L.S.; Passador, F.R.; Rezende, M.C.; Lemes, A.P. Biodegradable nanocomposites based on PLA/PHBV blend reinforced with carbon nanotubes with potential for electrical and electromagnetic applications. Express Polym. Lett. 2021, 15, 1–12. [Google Scholar] [CrossRef]
  42. Zhao, H.; Cui, Z.; Wang, X.; Turng, L.S.; Peng, X. Processing and characterization of solid and microcellular poly(lactic acid)/polyhydroxybutyrate-valerate (PLA/PHBV) blends and PLA/PHBV/Clay nanocomposites. Compos. B Eng. 2013, 51, 79–91. [Google Scholar] [CrossRef]
  43. Tao, J.; Song, C.; Cao, M.; Hu, D.; Liu, L.; Liu, N.; Wang, S. Thermal properties and degradability of poly(propylene carbonate)/poly(β-hydroxybutyrate-co-β-hydroxyvalerate) (PPC/PHBV) blends. Polym. Degrad. Stab. 2009, 94, 575–583. [Google Scholar] [CrossRef]
  44. Syahirah, W.N.; Azami, N.A.; Huong, K.H.; Amirul, A.A. Preparation, characterization and biodegradation of blend films of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) with natural biopolymers. Polym. Bull. 2020, 78, 3973–3993. [Google Scholar] [CrossRef]
  45. Castro-Mayorga, J.L.; Fabra, M.J.; Pourrahimi, A.M.; Olsson, R.T.; Lagaron, J.M. The impact of zinc oxide particle morphology as an antimicrobial and when incorporated in poly(3-hydroxybutyrate-co-3-hydroxyvalerate) films for food packaging and food contact surfaces applications. Food Bioprod. Process. 2017, 101, 32–44. [Google Scholar] [CrossRef] [Green Version]
  46. Shuai, C.; Wang, C.; Qi, F.; Peng, S.; Yang, W.; He, C.; Wang, G.; Qian, G. Enhanced crystallinity and antibacterial of PHBV scaffolds incorporated with zinc oxide. J. Nanomater. 2020, 2020, 1–12. [Google Scholar] [CrossRef]
  47. Alsabri, A.; Tahir, F.; Al-Ghamdi, S.G. Environmental impacts of polypropylene (PP) production and prospects of its recycling in the GCC region. Mat. Today Proc. 2021; in press. [Google Scholar] [CrossRef]
  48. Sen, S.K.; Raut, S. Microbial degradation of low density polyethylene (LDPE): A review. J. Environ. Chem. Eng. 2015, 3, 462–473. [Google Scholar] [CrossRef]
  49. Kader, M.A.; Senge, M.; Mojid, M.A.; Ito, K. Recent advances in mulching materials and methods for modifying soil environment. Soil Tillage Res. 2017, 168, 155–166. [Google Scholar] [CrossRef]
  50. Sarkar, D.J.; Barman, M.; Bera, T.; De, M.; Chatterjee, D. Agriculture: Polymers in crop production mulch and fertilizer. In Encyclopedia of Polymer Applications; Routledge: England, UK, 2018; Volume 1, pp. 1–20. [Google Scholar]
  51. Philip, S.; Keshavarz, T.; Roy, I. Polyhydroxyalkanoates: Biodegradable polymers with a range of applications. J. Chem. Technol. Biotechnol. 2007, 82, 233–247. [Google Scholar] [CrossRef]
  52. Yogesh, C.; Pathak, B.; Fulekar, M.H. PHA-production application and its bioremediation in environment. Res. J. Environ. Sci. 2012, 1, 46–52. Available online: http://www.isca.in/IJENS/Archive/v1/i2/9.ISCA-JEvsS-2012-009.pdf (accessed on 15 January 2022).
  53. Krasnits, E.; Beliavsky, M.; Tarre, S.; Green, M. PHA based denitrification: Municipal wastewater vs. acetate. Bioresour. Technol. 2013, 132, 28–37. [Google Scholar] [CrossRef]
  54. Santorio, S.; Fra-Vázquez, A.; Del Rio, A.V.; Mosquera-Corral, A. Potential of endogenous PHA as electron donor for denitrification. Sci. Total Environ. 2019, 695, 133747–133753. [Google Scholar] [CrossRef] [PubMed]
  55. Hiraishi, A.; Khan, S.T. Application of polyhydroxyalkanoates for denitrification in water and wastewater treatment. Appl. Microbiol. Biotechnol. 2003, 61, 103–109. [Google Scholar] [CrossRef] [PubMed]
  56. Schulz, H. Beta oxidation of fatty acids. Biochim. Biophys. Acta Lipids Lipid Metab. 1991, 1081, 109–120. [Google Scholar] [CrossRef]
  57. Azira, T.F.; Nursolehah, A.A.; Norhayati, Y.; Majid, M.I.A.; Amirul, A.A. Biosynthesis of poly(3-hydroxybutyrate-co-3-hydroxyvalerate-co-4-hydroxybutyrate) terpolymer by Cupriavidus sp. USMAA2-4 through two-step cultivation process. World J. Microbiol. Biotechnol. 2011, 27, 2287–2295. [Google Scholar] [CrossRef]
  58. Chen, Q.; Wang, Q.; Wei, G.; Liang, Q.; Qi, Q. Production in Escherichia coli of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) with differing monomer compositions from unrelated carbon sources. Appl. Environ. Microbiol. 2011, 77, 4886–4893. [Google Scholar] [CrossRef] [Green Version]
  59. Huijberts, G.N.; Eggink, G.; De Waard, P.; Huisman, G.W.; Witholt, B. Pseudomonas putida KT2442 cultivated on glucose accumulates poly(3-hydroxyalkanoates) consisting of saturated and unsaturated monomers. Appl. Environ. Microbiol. 1992, 58, 536–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Majid, M.I.A.; Akmal, D.H.; Few, L.L.; Agustien, A.; Toh, M.S.; Samian, M.R.; Najimudin, N.; Azizan, M.N. Production of poly(3-hydroxybutyrate) and its copolymer poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Erwinia sp. USMI-20. Int. J. Biol. Macromol. 1999, 25, 95–104. [Google Scholar] [CrossRef]
  61. Bingham, E.; Cohrssen, B.; Powell, C.H. Patty’s Toxicology Volume 1–6, 6th ed.; John Wiley & Sons: New York, NY, USA, 2012. [Google Scholar]
  62. Klabunde, J.; Bischoff, C.; Papa, A.J. Propanols. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley: Hoboken, NJ, USA, 2018; pp. 1–14. Available online: https://onlinelibrary.wiley.com/doi/10.1002/14356007.a22_173.pub3 (accessed on 15 January 2022).
  63. Mallat, T.; Baiker, A. Oxidation of alcohols with molecular oxygen on solid catalysts. Chem. Rev. 2004, 104, 3037–3058. [Google Scholar] [CrossRef]
  64. Steinbüchel, A.; Lütke-Eversloh, T. Metabolic engineering and pathway construction for biotechnological production of relevant polyhydroxyalkanoates in microorganisms. Biochem. Eng. J. 2003, 16, 81–96. [Google Scholar] [CrossRef]
  65. Berezina, N.; Yada, B. Improvement of the poly(3-hydroxybutyrate-co-3-hydroxyvalerate)(PHBV) production by dual feeding with levulinic acid and sodium propionate in Cupriavidus necator. New Biotechnol. 2016, 33, 231–236. [Google Scholar] [CrossRef]
  66. Chung, S.H.; Park, G.G.; Kim, H.W.; Rhee, Y.H. Effect of levulinic acid on the production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Ralstonia eutropha KHB-8862. J. Microbiol. 2001, 39, 79–82. [Google Scholar]
  67. Kim, D.Y.; Park, D.S.; Kwon, S.B.; Chung, M.G.; Bae, K.S.; Park, H.Y.; Rhee, Y.H. Biosynthesis of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) copolyesters with a high molar fraction of 3-hydroxyvalerate by an insect-symbiotic Burkholderia sp. IS-01. J. Microbiol. 2009, 47, 651–656. [Google Scholar] [CrossRef] [PubMed]
  68. Koller, M.; Hesse, P.; Fasl, H.; Stelzer, F.; Braunegg, G. Study on the effect of levulinic acid on whey-based biosynthesis of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Hydrogenophaga pseudoflava. Appl. Food Biotechnol. 2017, 4, 65–78. [Google Scholar] [CrossRef]
  69. Park, S.K.; Lee, K.T.; Kim, Y.B.; Rhee, Y.H. Biosynthesis of polyhydroxybutyrate and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) by Bacillus thuringiensis R-510. J. Microbiol. 1997, 35, 127–133. [Google Scholar]
  70. Gahlawat, G.; Soni, S.K. Valorization of waste glycerol for the production of poly(3-hydroxybutyrate) and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) copolymer by Cupriavidus necator and extraction in a sustainable manner. Bioresour. Technol. 2017, 243, 492–501. [Google Scholar] [CrossRef]
  71. Grousseau, E.; Blanchet, E.; Déléris, S.; Albuquerque, M.G.; Paul, E.; Uribelarrea, J.L. Phosphorus limitation strategy to increase propionic acid flux towards 3-hydroxyvaleric acid monomers in Cupriavidus necator. Bioresour. Technol. 2014, 153, 206–215. [Google Scholar] [CrossRef]
  72. Khanna, S.; Srivastava, A.K. Production of poly(3-hydroxybutyric-co-3-hydroxyvaleric acid) having a high hydroxyvalerate content with valeric acid feeding. J. Ind. Microbiol. Biotechnol. 2007, 34, 457–461. [Google Scholar] [CrossRef]
  73. Kim, B.S.; Lee, S.C.; Lee, S.Y.; Chang, H.N.; Chang, Y.K.; Woo, S.I. Production of poly(3-hydroxybutyric-co-3-hydroxyvaleric acid) by fed-batch culture of Alcaligenes eutrophus with substrate control using on-line glucose analyzer. Enzym. Microb. Technol. 1994, 16, 556–561. [Google Scholar] [CrossRef]
  74. Dionisi, D.; Majone, M.; Papa, V.; Beccari, M. Biodegradable polymers from organic acids by using activated sludge enriched by aerobic periodic feeding. Biotechnol. Bioeng. 2004, 85, 569–579. [Google Scholar] [CrossRef] [PubMed]
  75. Das, R.; Saha, N.R.; Pal, A.; Chattopadhyay, D.; Paul, A.K. Comparative evaluation of physicochemical characteristics of biopolyesters P(3HB) and P(3HB-co-3HV) produced by endophytic Bacillus cereus RCL 02. Front. Biol. 2018, 13, 297–308. [Google Scholar] [CrossRef]
  76. Amirul, A.A.; Syairah, S.N.; Yahya, A.R.; Azizan, M.N.M.; Majid, M.I.A. Synthesis of biodegradable polyesters by Gram-negative bacterium isolated from Malaysian environment. World J. Microbiol. Biotechnol. 2008, 24, 1327–1332. [Google Scholar] [CrossRef]
  77. Kim, S.W.; Kim, P.; Kim, J.H. Production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) from Methylobacterium organophilum by potassium-limited fed-batch culture. Enzym. Microb. Technol. 1999, 24, 555–560. [Google Scholar] [CrossRef]
  78. Sheu, D.S.; Chen, W.M.; Yang, J.Y.; Chang, R.C. Thermophilic bacterium Caldimonas taiwanensis produces poly(3-hydroxybutyrate-co-3-hydroxyvalerate) from starch and valerate as carbon sources. Enzym. Microb. Technol. 2009, 44, 289–294. [Google Scholar] [CrossRef]
  79. Myung, J.; Galega, W.M.; Van Nostrand, J.D.; Yuan, T.; Zhou, J.; Criddle, C.S. Long-term cultivation of a stable Methylocystis-dominated methanotrophic enrichment enabling tailored production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Bioresour. Technol. 2015, 198, 811–818. [Google Scholar] [CrossRef]
  80. Amini, M.; Sobhani, S.; Younesi, H.; Abyar, H.; Salamatinia, B.; Mohammadi, M. Evaluating the feasibility of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) co-biopolymer production from rice wastewater by Azohydromonas lata. Appl. Food Biotechnol. 2020, 7, 73–83. [Google Scholar] [CrossRef]
  81. Matsumoto, K.I.; Kitagawa, K.; Jo, S.J.; Song, Y.; Taguchi, S. Production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) in recombinant Corynebacterium glutamicum using propionate as a precursor. J. Biotechnol. 2011, 152, 144–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Doi, Y.; Kunioka, M.; Nakamura, Y.; Soga, K. Biosynthesis of copolyesters in Alcaligenes eutrophus H16 from carbon-13 labeled acetate and propionate. Macromolecules 1987, 20, 2988–2991. [Google Scholar] [CrossRef]
  83. Bhubalan, K.; Lee, W.H.; Loo, C.Y.; Yamamoto, T.; Tsuge, T.; Doi, Y.; Sudesh, K. Controlled biosynthesis and characterization of poly(3-hydroxybutyrate-co-3-hydroxyvalerate-co-3-hydroxyhexanoate) from mixtures of palm kernel oil and 3HV-precursors. Polym. Degrad. Stab. 2008, 93, 17–23. [Google Scholar] [CrossRef]
  84. Catalán, A.I.; Malan, A.K.; Ferreira, F.; Gill, P.R.; Batista, S. Propionic acid metabolism and poly-3-hydroxybutyrate-co-3-hydroxyvalerate production by a prpC mutant of Herbaspirillum seropedicae Z69. J. Biotechnol. 2018, 286, 36–44. [Google Scholar] [CrossRef]
  85. Lee, W.H.; Loo, C.Y.; Nomura, C.T.; Sudesh, K. Biosynthesis of polyhydroxyalkanoate copolymers from mixtures of plant oils and 3-hydroxyvalerate precursors. Bioresour. Technol. 2008, 99, 6844–6851. [Google Scholar] [CrossRef]
  86. Cal, A.J.; Sikkema, W.D.; Ponce, M.I.; Franqui-Villanueva, D.; Riiff, T.J.; Orts, W.J.; Pieja, A.J.; Lee, C.C. Methanotrophic production of polyhydroxybutyrate-co-hydroxyvalerate with high hydroxyvalerate content. Int. J. Biol. Macromol. 2016, 87, 302–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Wong, H.S.J.; Huong, K.H.; Hani, S.N.A.; Amirul, A.A.A. Genetic incorporation of oil-utilizing ability in Cupriavidus malaysiensis USMAA2-4 for sustainable polyhydroxyalkanoates production from palm olein and 1-pentanol. J. Biotechnol. 2021, 337, 71–79. [Google Scholar] [CrossRef] [PubMed]
  88. Huong, K.H.; Shantini, K.; Sharmini, R.; Amirul, A.A. Exploring the potential of 1-pentanol and oleic acid for optimizing the production of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) copolymer by Cupriavidus sp. USMAA1020. Arab. J. Sci. Eng. 2017, 42, 2313–2320. [Google Scholar] [CrossRef]
  89. Kiun, J.T.; Amelia, T.S.M.; Huong, K.H.; Amirul, A.A.; Bhubalan, K. Optimizing the biosynthesis of renewable polyhydroxyalkanoate copolymer containing 3-hydroxyvalerate by Massilia haematophila using statistical modeling. BioTechnologia 2019, 100, 359–371. [Google Scholar] [CrossRef]
  90. Ezhov, V.A.; Doronina, N.V.; Trotsenko, Y.A. Biosynthesis of polyhydroxybutyrate/valerate with different molecular weights during the growth of Methylobacterium extorquens G-10 on a methanol-pentanol mixture. Appl. Biochem. Microbiol. 2013, 49, 150–153. [Google Scholar] [CrossRef]
  91. Galuzina, T.V.; Gerasin, V.A.; Doronina, N.V.; Ezhov, V.A.; Trotsenko, Y.A.; Kiprianov, S.V.; Ivanov, A.O.; Filatova, M.P.; Shklyaruk, B.F. Structures and properties of polyhydroxyalkanoates synthesized by Methyloligella halotolerans C2 and Methylobacterium extorquens G10 from a methanol–pentanol mixture. Polym. Sci. Ser. A 2015, 57, 729–737. [Google Scholar] [CrossRef]
  92. Raberg, M.; Voigt, B.; Hecker, M.; Steinbüchel, A. A closer look on the polyhydroxybutyrate-(PHB)-negative phenotype of Ralstonia eutropha PHB-4. PLoS ONE 2014, 9, e95907. [Google Scholar] [CrossRef] [Green Version]
  93. Choi, J.; Lee, S.Y. Factors affecting the economics of polyhydroxyalkanoate production by bacterial fermentation. Appl. Microbiol. Biotechnol. 1999, 51, 13–21. [Google Scholar] [CrossRef]
  94. Saratale, R.G.; Cho, S.K.; Saratale, G.D.; Kadam, A.A.; Ghodake, G.S.; Kumar, M.; Bharagava, R.N.; Kumar, G.; Kim, D.S.; Mulla, S.I.; et al. A comprehensive overview and recent advances on polyhydroxyalkanoates (PHA) production using various organic waste streams. Bioresour. Technol. 2021, 325, 124685–124699. [Google Scholar] [CrossRef] [PubMed]
  95. Bhattacharyya, A.; Jana, K.; Haldar, S.; Bhowmic, A.; Mukhopadhyay, U.K.; De, S.; Mukherjee, J. Integration of poly-3-(hydroxybutyrate-co-hydroxyvalerate) production by Haloferax mediterranei through utilization of stillage from rice-based ethanol manufacture in India and its techno-economic analysis. World J. Microbiol. Biotechnol. 2015, 31, 717–727. [Google Scholar] [CrossRef] [PubMed]
  96. Trzeciak, A.M. 6.02–Hydroformylation. In Comprehensive Inorganic Chemistry II, 2nd ed.; Reedijk, J., Poeppelmeier, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 25–46. [Google Scholar] [CrossRef]
  97. Torres, G.M.; Frauenlob, R.; Franke, R.; Börner, A. Production of alcohols via hydroformylation. Catal. Sci. Technol. 2015, 5, 34–54. [Google Scholar] [CrossRef]
  98. Molnár, Á.; Papp, A. Catalyst recycling—a survey of recent progress and current status. Coord. Chem. Rev. 2017, 349, 1–65. [Google Scholar] [CrossRef]
  99. Lappe, P.; Hofmann, T. Pentanols. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley: Hoboken, NJ, USA, 2012; pp. 245–258. Available online: https://onlinelibrary.wiley.com/doi/10.1002/14356007.a19_049.pub2 (accessed on 15 January 2022).
  100. Cropley, J.B.; Burgess, L.M.; Loke, R.A. Butyraldehyde hydrogenation-a case-study in process design. Chemtech 1984, 14, 374–380. [Google Scholar]
  101. Faith, W.L.; Keyes, D.B.; Clark, R.L. Industrial Chemicals, 3rd ed.; Wiley & Sons: New York, NY, USA, 1965. [Google Scholar]
  102. Fuchs, D.; Rousseau, G.; Diab, L.; Gellrich, U.; Breit, B. Tandem rhodium-catalyzed hydroformylation–hydrogenation of alkenes by employing a cooperative ligand system. Angew. Chem. Int. Ed. 2012, 51, 2178–2182. [Google Scholar] [CrossRef] [PubMed]
  103. Allen, S.H.G.; Kellermeyer, R.W.; Stjernholm, R.L.; Wood, H.G. Purification and properties of enzymes involved in the propionic acid fermentation. J. Bacteriol. 1964, 87, 171–187. [Google Scholar] [CrossRef] [Green Version]
  104. Wood, H.G. Metabolic cycles in the fermentation by propionic acid bacteria. In Current Topics in Cellular Regulation; Academic Press: Cambridge, MA, USA, 1981; Volume 18, pp. 255–287. [Google Scholar] [CrossRef]
  105. Himmi, E.H.; Bories, A.; Boussaid, A.; Hassani, L. Propionic acid fermentation of glycerol and glucose by Propionibacterium acidipropionici and Propionibacterium freudenreichii ssp. shermanii. Appl. Microbiol. Biotechnol. 2000, 53, 435–440. [Google Scholar] [CrossRef] [PubMed]
  106. Seeliger, S.; Janssen, P.H.; Schink, B. Energetics and kinetics of lactate fermentation to acetate and propionate via methylmalonyl-CoA or acrylyl-CoA. FEMS Microbiol. Lett. 2002, 211, 65–70. [Google Scholar] [CrossRef] [PubMed]
  107. Walther, T.; François, J.M. Microbial production of propanol. Biotechnol. Adv. 2016, 34, 984–996. [Google Scholar] [CrossRef]
  108. Cardon, B.P.; Barker, H.A. Two new amino-acid-fermenting bacteria, Clostridium propionicum and Diplococcus glycinophilus. J. Bacterial. 1946, 52, 629–634. [Google Scholar] [CrossRef] [Green Version]
  109. Tholozan, J.L.; Touzel, J.P.; Samain, E.; Grivet, J.P.; Prensier, G.; Albagnac, G. Clostridium neopropionicum sp. nov., a strict anaerobic bacterium fermenting ethanol to propionate through acrylate pathway. Arch. Microbiol. 1992, 157, 249–257. [Google Scholar] [CrossRef]
  110. Liu, K.; Atiyeh, H.K.; Stevenson, B.S.; Tanner, R.S.; Wilkins, M.R.; Huhnke, R.L. Continuous syngas fermentation for the production of ethanol, n-propanol and n-butanol. Bioresour. Technol. 2014, 151, 69–77. [Google Scholar] [CrossRef] [PubMed]
  111. Millat, T.; Janssen, H.; Thorn, G.J.; King, J.R.; Bahl, H.; Fischer, R.J.; Wolkenhauer, O. A shift in the dominant phenotype governs the pH-induced metabolic switch of Clostridium acetobutylicum in phosphate-limited continuous cultures. Appl. Microbiol. Biotechnol. 2013, 97, 6451–6466. [Google Scholar] [CrossRef] [PubMed]
  112. Isom, C.E.; Nanny, M.A.; Tanner, R.S. Improved conversion efficiencies for n-fatty acid reduction to primary alcohols by the solventogenic acetogen “Clostridium ragsdalei”. J. Ind. Microbiol. Biotechnol. 2015, 42, 29–38. [Google Scholar] [CrossRef] [PubMed]
  113. Atsumi, S.; Liao, J.C. Directed evolution of Methanococcus jannaschii citramalate synthase for biosynthesis of 1-propanol and 1-butanol by Escherichia coli. Appl. Environ. Microbiol. 2008, 74, 7802–7808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Shen, C.R.; Liao, J.C. Metabolic engineering of Escherichia coli for 1-butanol and 1-propanol production via the keto-acid pathways. Metab. Eng. 2008, 10, 312–320. [Google Scholar] [CrossRef]
  115. Shen, C.R.; Liao, J.C. Synergy as design principle for metabolic engineering of 1-propanol production in Escherichia coli. Metab. Eng. 2013, 17, 12–22. [Google Scholar] [CrossRef]
  116. Chen, G.S.; Siao, S.W.; Shen, C.R. Saturated mutagenesis of ketoisovalerate decarboxylase V461 enabled specific synthesis of 1-pentanol via the ketoacid elongation cycle. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  117. Haller, T.; Buckel, T.; Rétey, J.; Gerlt, J.A. Discovering new enzymes and metabolic pathways: Conversion of succinate to propionate by Escherichia coli. Biochemistry 2000, 39, 4622–4629. [Google Scholar] [CrossRef]
  118. Srirangan, K.; Liu, X.; Westbrook, A.; Akawi, L.; Pyne, M.E.; Moo-Young, M.; Chou, C.P. Biochemical, genetic, and metabolic engineering strategies to enhance coproduction of 1-propanol and ethanol in engineered Escherichia coli. Appl. Microbiol. Biotechnol. 2015, 98, 9499–9515. [Google Scholar] [CrossRef]
  119. Park, H.; Jeon, B.S.; Sang, B.I. efficient, simple production of corresponding alcohols from supplemented C2-C8 carboxylic acids in Escherichia coli using acyl-CoA transferase from Megasphaera hexanoica. Biotechnol. Bioprocess. Eng. 2020, 25, 599–606. [Google Scholar] [CrossRef]
  120. Steinbüchel, A.; Schlegel, H.G. Physiology and molecular genetics of poly(β-hydroxyalkanoic acid) synthesis in Alcaligenes eutrophus. Mol. Microbiol. 1991, 5, 535–542. [Google Scholar] [CrossRef] [PubMed]
  121. Kelly, D.P.; Rainey, F.A.; Wood, A.P. The genus Paracoccus. In The Prokaryotes; Dworkin, M., Falkow, S., Rosenberg, E., Schleifer, K.H., Stackebrandt, E., Eds.; Springer: New York, NY, USA, 2006; pp. 232–249. [Google Scholar] [CrossRef]
  122. Akmal, D.; Asiska, P.D.; Wangi, Q.A.; Rivai, H.; Agustien, A. Biosynthesis of copolymer poly(3-hydroxybutyrate-co-3-hydroxyvalerate) from palm oil and n-pentanol in a 10 L bioreactor. Rasayan J. Chem. 2015, 8, 389–395. [Google Scholar]
  123. Amirul, A.A.; Yahya, A.R.; Sudesh, K.; Azizan, M.N.M.; Majid, M.I.A. Isolation of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) producer from Malaysian environment using γ-butyrolactone as carbon source. World J. Microbiol. Biotechnol. 2009, 25, 1199–1206. [Google Scholar] [CrossRef]
  124. Ramachandran, H.; Shafie, N.A.H.; Sudesh, K.; Azizan, M.N.; Majid, M.I.A.; Amirul, A.A.A. Cupriavidus malaysiensis sp. nov.; a novel poly(3-hydroxybutyrate-co-4-hydroxybutyrate) accumulating bacterium isolated from the Malaysian environment. Antonie Leeuwenhoek 2018, 111, 361–372. [Google Scholar] [CrossRef] [PubMed]
  125. Shantini, K.; Bhubalan, K.; Yahya, A.R.M.; Amirul, A.A. Productivity increment of biodegradable and biorenewable copolymer containing 3-hydroxyvalerate monomer initiated by alcohols as precursor substrates. J. Chem. Technol. Biotechnol. 2013, 88, 1364–1370. [Google Scholar] [CrossRef]
  126. Sashi, P.; Yasin, U.M.; Bhuyan, A.K. Unfolding action of alcohols on a highly negatively charged state of cytochrome c. Biochemistry 2012, 51, 3273–3283. [Google Scholar] [CrossRef]
  127. Ingram, L.O. Adaptation of membrane lipids to alcohols. J. Bacterial. 1976, 125, 670–678. [Google Scholar] [CrossRef] [Green Version]
  128. Jovanovic, G.; Lloyd, L.J.; Stumpf, M.P.; Mayhew, A.J.; Buck, M. Induction and function of the phage shock protein extracytoplasmic stress response in Escherichia coli. J. Biol. Chem. 2006, 281, 21147–21161. [Google Scholar] [CrossRef] [Green Version]
  129. Yano, T.; Miyahara, Y.; Morii, N.; Okano, T.; Kubota, H. Pentanol and benzyl alcohol attack bacterial surface structures differently. Appl. Environ. Microbiol. 2016, 82, 402–408. [Google Scholar] [CrossRef] [Green Version]
  130. Osman, Y.A.; Ingram, L.O. Mechanism of ethanol inhibition of fermentation in Zymomonas mobilis CP4. J. Bacterial. 1985, 164, 173–180. [Google Scholar] [CrossRef] [Green Version]
  131. Ingram, L.O.N.; Buttke, T.M. Effects of alcohols on micro-organisms. Adv. Microb. Physiol. 1985, 25, 253–300. [Google Scholar] [CrossRef]
  132. Heipieper, H.J.; Meinhardt, F.; Segura, A. The cis–trans isomerase of unsaturated fatty acids in Pseudomonas and Vibrio: Biochemistry, molecular biology and physiological function of a unique stress adaptive mechanism. FEMS Microbiol. Lett. 2003, 229, 1–7. [Google Scholar] [CrossRef] [Green Version]
  133. Beaven, M.J.; Charpentier, C.; Rose, A.H. Production and tolerance of ethanol in relation to phospholipid fatty-acyl composition in Saccharomyces cerevisiae NCYC 431. Microbiology 1982, 128, 1447–1455. [Google Scholar] [CrossRef] [Green Version]
  134. Kitahara, K.; Takichi, K.; Osamu, G. Taxonomic studies on the hiochi-bacteria, specific saprophytes of sake II. Identification and classification of hiochi-bacteria. J. Gen. Appl. Microbiol. 1957, 3, 111–120. [Google Scholar] [CrossRef] [Green Version]
  135. Bernal, P.; Muñoz-Rojas, J.; Hurtado, A.; Ramos, J.L.; Segura, A. A Pseudomonas putida cardiolipin synthesis mutant exhibits increased sensitivity to drugs related to transport functionality. Environ. Microbiol. 2007, 9, 1135–1145. [Google Scholar] [CrossRef] [PubMed]
  136. Marr, A.G.; Ingraham, J.L. Effect of temperature on the composition of fatty acids in Escherichia coli. J. Bacteriol. 1962, 84, 1260–1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Brynildsen, M.P.; Liao, J.C. An integrated network approach identifies the isobutanol response network of Escherichia coli. Mol. Syst. Biol. 2009, 5, 277–279. [Google Scholar] [CrossRef] [PubMed]
  138. Davidson, A.L.; Dassa, E.; Orelle, C.; Chen, J. Structure, function, and evolution of bacterial ATP-binding cassette systems. Microbiol. Mol. Biol. Rev. 2008, 72, 317–364. [Google Scholar] [CrossRef] [Green Version]
  139. Rutherford, B.J.; Dahl, R.H.; Price, R.E.; Szmidt, H.L.; Benke, P.I.; Mukhopadhyay, A.; Keasling, J.D. Functional genomic study of exogenous n-butanol stress in Escherichia coli. Appl. Environ. Microbiol. 2010, 76, 1935–1945. [Google Scholar] [CrossRef] [Green Version]
  140. Hews, C.L.; Cho, T.; Rowley, G.; Raivio, T.L. Maintaining integrity under stress: Envelope stress response regulation of pathogenesis in Gram-negative bacteria. Front. Cell. Infect. Microbiol. 2019, 9, 313–337. [Google Scholar] [CrossRef] [Green Version]
  141. Isaac, D.D.; Pinkner, J.S.; Hultgren, S.J.; Silhavy, T.J. The extracytoplasmic adaptor protein CpxP is degraded with substrate by DegP. Proc. Natl. Acad. Sci. USA 2005, 102, 17775–17779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Mitra, R.; Gadkari, V.V.; Meinen, B.A.; van Mierlo, C.P.; Ruotolo, B.T.; Bardwell, J.C. Mechanism of the small ATP-independent chaperone Spy is substrate specific. Nat. Commun. 2021, 12, 1–13. [Google Scholar] [CrossRef] [PubMed]
  143. Tomas, C.A.; Welker, N.E.; Papoutsakis, E.T. Overexpression of groESL in Clostridium acetobutylicum results in increased solvent production and tolerance, prolonged metabolism, and changes in the cell’s transcriptional program. Appl. Environ. Microbiol. 2003, 69, 4951–4965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Postma, P.W.; Lengeler, J.W.; Jacobson, G.R. Phosphoenolpyruvate: Carbohydrate phosphotransferase systems of bacteria. Microbiol Rev. 1993, 57, 543–594. [Google Scholar] [CrossRef] [PubMed]
  145. Igarashi, K.; Fujita, N.; Ishihama, A. Sequence analysis of two temperature-sensitive mutations in the alpha subunit gene (rpoA) of Escherichia coli RNA polymerase. Nucleic Acids Res. 1990, 18, 5945–5948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Jafri, S.; Urbanowski, M.L.; Stauffer, G.V. A mutation in the rpoA gene encoding the alpha subunit of RNA polymerase that affects metE-metR transcription in Escherichia coli. J. Bacteriol. 1995, 177, 524–529. [Google Scholar] [CrossRef] [Green Version]
  147. Klein-Marcuschamer, D.; Santos, C.N.S.; Yu, H.; Stephanopoulos, G. Mutagenesis of the bacterial RNA polymerase alpha subunit for improvement of complex phenotypes. Appl. Environ. Microbiol. 2009, 75, 2705–2711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Peck, M.C.; Gaal, T.; Fisher, R.F.; Gourse, R.L.; Long, S.R. The RNA polymerase α subunit from Sinorhizobium meliloti can assemble with RNA polymerase subunits from Escherichia coli and function in basal and activated transcription both in vivo and in vitro. J. Bacteriol. 2002, 184, 3808–3814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Thomas, M.S.; Glass, R.E. Escherichia coli rpoA mutation which impairs transcription of positively regulated systems. Mol. Microbiol. 1991, 5, 2719–2725. [Google Scholar] [CrossRef] [PubMed]
  150. Kidwell, J.; Valentin, H.E.; Dennis, D. Regulated expression of the Alcaligenes eutrophus pha biosynthesis genes in Escherichia coli. Appl. Environ. Microbiol. 1995, 61, 1391–1398. [Google Scholar] [CrossRef] [Green Version]
  151. Nomura, C.T.; Taguchi, K.; Taguchi, S.; Doi, Y. Coexpression of genetically engineered 3-ketoacyl-ACP synthase III (fabH) and polyhydroxyalkanoate synthase (phaC) genes leads to short-chain-length-medium-chain-length polyhydroxyalkanoate copolymer production from glucose in Escherichia coli JM109. Appl. Environ. Microbiol. 2004, 70, 999–1007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Taguchi, K.; Aoyagi, Y.; Matsusaki, H.; Fukui, T.; Doi, Y. Over-expression of 3-ketoacyl-ACP synthase III or malonyl-CoA-ACP transacylase gene induces monomer supply for polyhydroxybutyrate production in Escherichia coli HB101. Biotechnol. Lett. 1999, 21, 579–584. [Google Scholar] [CrossRef]
Figure 1. Microbial PHA granule, P(3HB-co-3HV) structure, and applications.
Figure 1. Microbial PHA granule, P(3HB-co-3HV) structure, and applications.
Polymers 14 00670 g001
Figure 2. Schematic bioconversion pathway of organic acids and alkyl alcohols into 3HV [14,57,58,59,60].
Figure 2. Schematic bioconversion pathway of organic acids and alkyl alcohols into 3HV [14,57,58,59,60].
Polymers 14 00670 g002
Figure 3. Techno-economic analysis on PHA production annual costs using pure carbon sources and wastes [93,95].
Figure 3. Techno-economic analysis on PHA production annual costs using pure carbon sources and wastes [93,95].
Polymers 14 00670 g003
Figure 4. Oxo synthesis of alkyl alcohols [62,97,99].
Figure 4. Oxo synthesis of alkyl alcohols [62,97,99].
Polymers 14 00670 g004
Figure 5. Biosynthesis of 1-propanol by wild-type Propionibacteria through the Wood–Werkman pathway [103,104].
Figure 5. Biosynthesis of 1-propanol by wild-type Propionibacteria through the Wood–Werkman pathway [103,104].
Polymers 14 00670 g005
Figure 6. Biosynthesis of 1-propanol by wild-type C. propionicumable and C. neopropionicumable through the acrylate pathway [109,110].
Figure 6. Biosynthesis of 1-propanol by wild-type C. propionicumable and C. neopropionicumable through the acrylate pathway [109,110].
Polymers 14 00670 g006
Figure 7. Biosynthesis of 1-propanol and 1-pentanol by wild-type C. ragsdalei through the carboxylate reduction pathway [112].
Figure 7. Biosynthesis of 1-propanol and 1-pentanol by wild-type C. ragsdalei through the carboxylate reduction pathway [112].
Polymers 14 00670 g007
Figure 8. Biosynthesis of 1-propanol by genetically engineered E. coli BW25113 through co-expression of the citramalate and threonine pathway [115].
Figure 8. Biosynthesis of 1-propanol by genetically engineered E. coli BW25113 through co-expression of the citramalate and threonine pathway [115].
Polymers 14 00670 g008
Figure 9. Biosynthesis of 1-pentanol by genetically engineered E. coli BW25113 through interactive elongation cycles of 2-letoacids [116]. Dash arrow indicates lower selectivity toward the reaction.
Figure 9. Biosynthesis of 1-pentanol by genetically engineered E. coli BW25113 through interactive elongation cycles of 2-letoacids [116]. Dash arrow indicates lower selectivity toward the reaction.
Polymers 14 00670 g009
Figure 10. Biosynthesis of 1-propanol by genetically engineered E. coli BW25113 with activated SBM operon for extended dissimilation of succinate [118].
Figure 10. Biosynthesis of 1-propanol by genetically engineered E. coli BW25113 with activated SBM operon for extended dissimilation of succinate [118].
Polymers 14 00670 g010
Figure 11. Biosynthesis of 1-pentanol by genetically engineered E. coli BL21(DE3 with acquired carboxylate reduction pathway [119].
Figure 11. Biosynthesis of 1-pentanol by genetically engineered E. coli BL21(DE3 with acquired carboxylate reduction pathway [119].
Polymers 14 00670 g011
Figure 12. Timeline of the emergence of alkyl alcohol-tolerant P(3HB-co-3HV)-producing bacteria.
Figure 12. Timeline of the emergence of alkyl alcohol-tolerant P(3HB-co-3HV)-producing bacteria.
Polymers 14 00670 g012
Table 1. Properties improvement after the incorporation of a secondary (and tertiary) component into P(3HB-co-3HV) and their potential applications.
Table 1. Properties improvement after the incorporation of a secondary (and tertiary) component into P(3HB-co-3HV) and their potential applications.
Incorporated Components AChanges in the PropertiesPotential
Applications
Ref.
α-P(3HB)
Incorporation method:
Solvent casting
3HV fraction: 10 mol%
P(3HB-co-3HV):α-P(3HB) (100:0 → 50:50) 
Melting temperature: 145 → 133 °C
Degree of crystallinity: 61% → 30%
Tensile strength: 27 → 7 MPa
Elongation at break: 1% → 29%
Young’s modulus: 1500 → 240 MPa
Enzymatic degradation: 85% → 94%
Packaging
material
[25]
AS
Incorporation method:
Solvent casting
3HV fraction: 59 mol%
P(3HB-co-3HV)/P(3HB-co-3HV):AS
Melting temperature: 275.84 °C/294.97 °C
Degree of crystallinity: 98.96%/98.23%
Free radical scavenging activity (24 h): 1%/14%
Incubation biodegradation (day 6):
smooth surface/small pits
Therapeutic
implant
[26]
CNC
Incorporation method:
Solvent casting
3HV fraction: 12 mol%
P(3HB-co-3HV):CNC (100:0 → 94:6)
Melting temperature: 136.8 → 151.1 °C
Crystallization temperature: 96.5 → 101.2 °C
Degree of crystallinity: 49.9% → 57.5%
Water vapor transmission rate: 308 → 115 g m−2 day−1
Oxygen transfer rate: 425 → 113 cm m−2 day−1
Packaging
material
[27]
DDGS or Misc
Incorporation method:
Twin screw extrusion
3HV fraction: 5 mol%
P(3HB-co-3HV):DDGS (100:0/85:15/75:25)
Tensile strength: 8.5 MPa/6.0 MPa/4.8 MPa
Young’s modulus: 3.9 GPa/3.9 GPa/3.8 GPa
Flexural strength: 7.0 MPa/5.8 MPa/4.7 MPa
Flexural modulus: 4.8 GPa/4.6 GPa/4.4 GPa
CO2 evolution (day 320): 155 mg/175 mg/200 mg
Marine biodegradation (day 320): 73%/90%/100%
P(3HB-co-3HV):Misc (85:15/75:25)
Tensile strength: 8.8 MPa/8.9 MPa
Young’s modulus: 5.9 GPa/7.7 GPa
Flexural strength: 7.8 MPa/7.4 MPa
Flexural modulus:5.6 GPa/6.6 GPa
CO2 evolution (day 320):175 mg/180 mg
Marine biodegradation (day 320): 84%/88%
Packaging
material
[28]
Eugenol
Incorporation method:
Electrospinning
3HV fraction: 3 mol%
P(3HB-co-3HV):Eugenol (100:0 → 85:15)
Temperature of 5% weight loss: 276.6 → 160.8 °C
Degradation temperature: 304.7 → 293.3 °C
Mass loss at degradation temperature: 61.01% → 76.36%
Water vapor permeability:
4.05 × 1014→ 0.95 × 1014 Kg m m−2 s−1 Pa−1
Limonene vapor permeability: 3.75 → 0.81 Kg m m−2 s−1 Pa−1
Water vapor permeance: 5.87 → 1.33 Kg m m−2 s−1 Pa−1
Limonene vapor permeance: 5.44 → 1.14 Kg m m−2 s−1 Pa−1
Tensile strength: 1252 → 1897 MPa
Elongation at break: 2.0% → 2.5%
Young’s modulus: 18.1 → 26.5 MPa
S. aureus growth: 5.16 → 3.45 log(CFU mL−1)
Escherichia coli growth: 5.79 → 3.88 log(CFU mL−1)
Antimicrobial
food
packaging
[29]
HA
Incorporation method:
Melt-pressing
3HV fraction: 8–24 mol%
P(3HB-co-3HV), 0→24 mol% 3HV
Melting temperature: 170 → 129 °C
Degree of crystallinity: 69% → 55%
P(3HB-co-3HV):HA (30:70), 0 → 24 mol% 3HV
Tensile strength: 67 → 23 MPa
Elongation at break: 2.65% → 3.84%
Young’s modulus: 2.52 → 0.47 GPa
Bone
implant
[30]
MAT
Incorporation method:
Solvent casting
3HV fraction: 4 mol%
P(3HB-co-3HV):MAT (100:0 → 95:5)
Melting temperature: 168.58 → 130.91 °C
Glass transition temperature: −2.03 → −6.61 °C
Crystallization temperature: 46.15 → 46.98 °C
Degree of crystallinity: 53.7% → 36.8%
Packaging
material
[31]
MCPA
Incorporation method:
Melt-blending and
hot-pressing
3HV fraction: 3 mol%
P(3HB-co-3HV)-MCPA (95:5/90:10/85:15)
Melting temperature 1: 123.2 °C/124.1 °C /NA
Melting temperature 2: 150.7 °C/150.7 °C/140.9 °C
Enthalpy of fusion 1:1944 J g−1/2482 J g−1/NA
Enthalpy of fusion 2:1745 J g−1/1745 J g−1/1509 J g−1
Glass transition temperature 1: −28.2 °C/−28.0 °C/−27.4 °C
Glass transition temperature 2: 48.6 °C/47.9 °C/36.9 °C
Crystallization temperature: 102.4 °C/102.2 °C/99.0 °C
Chlorine loss: 0.3%/1.3%/1.7%
MCPA loss: 5.1%/7.4%/9.7%
P(3HB-co-3HV) loss before bond scission: 20.6%/29.7%/38.8%
P(3HB-co-3HV) loss after bond scission: 2.8%/2.4%/2.4%
Mulch[32]
mPEG
Incorporation method:
Transesterification
3HV fraction: 12 and 33 mol%
P(3HB-co-3HV):mPEG, 12 mol%/33 mol% 3HV
Number average molecular weight: 8980/4980
Weight average molecular weight: 6200/2650
Polydispersity index: 1.44/1.84
Melting temperature of P(3HB-co-3HV) block: 140.5 °C/133.6 °C
Melting temperature of mPEG block: 49.1 °C/49.3 °C
Particle size: 162 nm/125 nm
Encapsulation efficiency: 43%/57%
Cytotoxicity (100 → 500 µg/mL nanoparticles): 94% → 80%/88% → 78%
Drug
delivery
carrier
[24]
NH2-g-collagen or
PHEMA-g-collagen
Incorporation method:
Solvent
casting
followed by
solute leaching technique
3HV fraction: 12 mol%
Porous P(3HB-co-3HV)
Decomposition temperature at 10% weight loss: 263.15 °C
Collagen concentration: NA
Ag/BSA load: 0.037 µg cm−2
Surface roughness: 0.1983 µm
P(3HB-co-3HV)--g-PHEMA-g-collagen
Decomposition temperature at 10% weight loss: 264.60 °C
Collagen concentration: 29.93 µg cm−2
Ag/BSA load: 0.29 µg cm−2
Surface roughness: NA
P(3HB-co-3HV)-g-NH2-g-collagen
Decomposition temperature at 10% weight loss: 256.15 °C
Collagen concentration: 55.16 µg cm−2
Ag/BSA load: 0.26 µg cm−2
Surface roughness: 0.2643 µm
Bone
implant
[33]
NR
Incorporation method:
Twin screw extrusion
3HV fraction: 3 mol%
P(3HB-co-3HV):NR (100:0/85:15)
Melting temperature: 172.05 °C/171.95 °C
P(3HB-co-3HV) glass transition temperature: 5.65 °C/6.05 °C
NR glass transition temperature: NA/−64.5 °C
Crystallization temperature: 120.85 °C/119.45 °C
Degree of crystallinity: 74.7%/61.6%
Tensile strength: 43 MPa/26 MPa
Elongation at break: 8%/16%
Notched impact strength: 15 J m−1/14 J m−1
Secant modulus: 12 GPa/0.9 GPa
Packaging
material
[34]
PBAT
Incorporation method:
Conventional injection
molding or microcellular
injection molding
3HV fraction: NA
Solid P(3HB-co-3HV):PBAT (98.5:1.5 → 30:70)
Melting temperature: 166.2 → 170.4 °C
Cold crystallization temperature: NA → 44.7 °C
Degree of crystallinity: 78% → 29%
Specific toughness: 5.3 x 10−4 → 7.1 x 10−2 MPa kg−1 m−3
Elongation at break: 2.7% → 555.7%
Specific tensile strength:
3.2 × 10−2 → 1.5 x 10−2 MPa kg−1 m−3
Specific Young’s modulus: 2.2 → 0.5 MPa kg−1 m−3
Microcellular P(3HB-co-3HV):PBAT (98.5:1.5 → 30:70)
Melting temperature: 167.1 → 169.6 °C
Cold crystallization temperature: NA → 45.7 °C
Degree of crystallinity: 80% → 25%
Specific toughness: 3.8 x 10−4 → 5.8 x 10−2 MPa kg−1 m−3
Elongation at break: 2.2% → 493.9%
Specific tensile strength:
2.7 × 10−2 → 1.3 x 10−2 MPa kg−1 m−3
Specific Young’s modulus: 2.1 → 0.5 MPa kg−1 m−3
Packaging
material
[35]
PBS
Incorporation method:
Solvent casting
3HV fraction: 14 mol%
P(3HB-co-3HV):PBS (100:0 → 40:60)
Crystallization time at 60 °C: 8 → 14.5 min
Overall crystallization constant:
3.13 × 10−2 → 2.22 × 10−3 min−n
Avrami index: 2.57 → 2.67
Packaging
material
[36]
PBS-DCP
Incorporation method:
Compression molding
3HV fraction: 13 mol%
P(3HB-co-3HV):PBS (100:0 → 70:30)
Tensile strength: 22 → 23 MPa
Elongation at break: 4.5% → 6.5%
80 wt%P(3HB-co-3HV)–20 wt%PBS:DCP (100:0 → 99:1)
Tensile strength: 25 → 27 MPa
Elongation at break: 8% → 350%
Notched Izod impact toughness: 2.8 → 5.5 kJ m−2
Flexural strength: 39 → 30 MPa
Flexural modulus: 1.2 → 0.6 GPa
Packaging
material
[37]
PCL
Incorporation method:
Solvent casting
3HV fraction: 7 mol%
P(3HB-co-3HV)/PCL
Number average molecular weight: 127,000/56,400
Weight average molecular weight: 470,000/163,300
Melting temperature: 151.2 °C/64.0 °C
Glass transition temperature: 5.2 °C/−61.0 °C
Crystallization temperature: 97.0 °C/22.2 °C
P(3HB-co-3HV):PCL (100:0 → 50:50)
Isothermal crystallization temperature: 120 → 120 °C
Overall crystallization constant: 2.20 × 10−7 → 1.00 × 10−8 s−n
Avrami index: 2.80 → 2.66
Packaging
material
[38]
PDLLA-PEG
Incorporation method:
Compression molding
3HV fraction: 1 mol%
P(3HB-co-3HV):PDLLA (100:0 → 30:70)
Melting temperature: 157.8 → 169.8 °C
Degree of crystallinity: 53.6 → 9.9 °C
Tensile strength: 19.7 → 49.7 MPa
Elongation at break: 0.17% → 2.07%
Flexural strength: 39.1 → 75.0 MPa
Flexural modulus:3646 → 3507 MPa
Burial biodegradation (day 30): 0% → 1%
30 wt%P(3HB-co-3HV)–70 wt%PDLLA:PEG (90:10 → 80:20)
Melting temperature: 171.2 → 170.8 °C
Degree of crystallinity: 10.5 → 13.0 °C
Tensile strength: 29.7 → 24.1 MPa
Elongation at break: 28.7% → 237.0%
Flexural strength: 36.1 → 5.48 MPa
Flexural modulus: 1127 → 220 MPa
Burial biodegradation (day 30): 3% → 11%
Biomedical,
agricultural
and packaging
material
[39]
PEG
Incorporation method:
Solvent casting
3HV fraction: 4 mol%
P(3HB-co-3HV):PEG (100:0 → 20:80)
Melting temperature: 163.2 → 145.0 °C
Enthalpy of fusion: 89.62 → 1.63 J g−1
Drug
delivery
carrier
[22]
PEG
Incorporation method:
Solvent casting
3HV fraction: NA
P(3HB-co-3HV)
Melting temperature: 90 °C
Initial thermal degradation temperature: 220 °C
Final thermal degradation temperature: 255 °C
Tensile strength: 10.3 MPa
Elongation at break: 13.3%
Cytotoxicity: 20%
P(3HB-co-3HV):PEG (4:1)
Cytotoxicity: 0%–10%
Skin
grafting
[40]
PLA-CNT
Incorporation method:
High-speed spinning
3HV fraction: 2 mol%
P(3HB-co-3HV)/PLA
Melting temperature: 172 °C/170 °C
Glass transition temperature: 5 °C/64 °C
Enthalpy of fusion: 92.8 J g−1/44.2 J g−1
Crystallization temperature: 122 °C/112 °C
Decomposition temperature: 303 °C/382 °C
Izod impact strength: 1.99 kJ m−2/2.14 kJ m−2
Flexural strength: 47.70 MPa/58.07 MPa
Flexural modulus: 3.48 GPa/2.94 GPa
80 wt%P(3HB-co-3HV)–20 wt%PLA:CNT (100:0 → 99:1)
Melting temperature: 169 → 168 °C
Glass transition temperature: −2 → −2 °C
Enthalpy of fusion: 44.11 → 48.10 J g−1
Crystallization temperature: 112 → 122 °C
Decomposition temperature: 379 → 380 °C
Izod impact strength: 4.10 → 2.46 kJ m−2
Flexural strength: 51.60 → 61.01 MPa
Flexural modulus: 3.10 → 3.25 GPa
Electrical conductivity: 8.67 × 10−14 → 2.79 × 10−2 S m−1
Reflectivity (frequency): 0 dB (NA) → −15 dB (11 GHz)
Electrical
and
electromagnetic
[41]
PLA-nanoclay
Incorporation method:
Twin screw extrusion
3HV fraction: NA
P(3HB-co-3HV):PLA (15:85 → 30:70)
Melting temperature: 154.75 → 156.40 °C
Cold crystallization temperature: 133.45 → 121.89 °C
Degree of crystallinity: 1.98% → 4.33%
Tensile strength: 52.5 → 47.5 MPa
Elongation at break: 9.0% → 6.0%
Young’s modulus: 1700 → 1750 MPa
P(3HB-co-3HV)-PLA:nanoclay (15:85 → 30:70)
Melting temperature: 156.52 → 157.43 °C
Cold crystallization temperature: 129.09 → 111.04 °C
Degree of crystallinity: 13.05% → 18.40%
Tensile strength: 49.2 → 48.0 MPa
Elongation at break: 8.5% → 4.0%
Young’s modulus: 2060 → 2060 MPa
Packaging
material
[42]
PPC
Incorporation method:
Solvent casting
3HV fraction: 5 mol%
P(3HB-co-3HV):PPC (100:0 → 20:80)
Melting temperature: 163 → 162 °C
Thermal decomposition temperature: 199 → 190 °C
Maximum mass loss rate temperature: 286 → 267 °C
Burial biodegradation: 100% (day 12) → 85% (day 30)
Packaging
material
[43]
starch, cellulose or alginate
Incorporation method:
Solvent casting
3HV fraction: 6 mol%
P(3HB-co-3HV)-starch (100:0 → 30:70)
Tensile strength: 25 → 1 MPa
Elongation at break: 8% → 4%
Young’s modulus: 181 → 4 MPa
Density: 0.974 → 1.243 g cm−3
Solubility: 0% → 6.0%
Water absorption capacity: 0% → 21.0%
Burial biodegradation (day 30): 10% → 100%
Immersion biodegradation (day 30): 23% → 100%
P(3HB-co-3HV)-cellulose (100:0 → 30:70)
Tensile strength: 25 → 1 MPa
Elongation at break: 8% → 3%
Young’s modulus: 181 → 7 MPa
Density: 0.974 → 1.212 g cm−3
Solubility: 0% → 1.7%
Water absorption capacity: 0% → 4.7%
Burial biodegradation (day 30): 10% → 70%
Immersion biodegradation (day 30): 23% → 100%
P(3HB-co-3HV)-arginate (100:0 → 30:70)
Tensile strength: 25 → 1 MPa
Elongation at break: 8% → 2%
Young’s modulus: 181 → 3 MPa
Density: 0.974 → 1.053 g cm−3
Solubility: 0% → 19.0%
Water absorption capacity: 0% → 33.0%
Burial biodegradation (day 30): 10% → 80%
Immersion biodegradation (day 30): 21% → 100%
Mulch[44]
ZnO
Incorporation method:
Melt-mixed compression molding, electrospinning or coating
3HV fraction: 3 and 18 mol%
P(3HB-co-3 mol%3HV)/ P(3HB-co-18 mol%3HV)
Melting temperature: 168.7 °C/170.9 °C
Decomposition temperature: 290.8 °C/283.1 °C
Crystallization temperature: 114.7 °C/101.0 °C
Degree of crystallinity: 66%/63%
Tensile strength: 33.9 MPa/18.5 MPa
Elongation at break: 1.5%/1.3%
Young’s modulus: 2.6 GPa/2.2 GPa
L*, a*, b*: 82.3, 1.4, 17.7/32.7, 6.7, 10.2
PHBVs-D/PHBVs-P/PHBVs-C B
Melting temperature: 166.9 °C/166.5 °C/169.0 °C
Decomposition temperature: 271.3 °C/270.3 °C/270.8 °C
Crystallization temperature: 112.1 °C/111.6 °C/118.0 °C
Degree of crystallinity: 50%/51%/35%
Tensile strength: 12.5 MPa/34.8 MPa/22.6 MPa
Elongation at break: 6.5%/2.3%/6.2%
Young’s modulus: 1.5 GPa/2.1 GPa/1.4 GPa
L*, a*, b*: 56.9, 9.0, 25.3/58.4, 8.5, 25.1/72.5, 3.8, 24.7
Active food
packaging and
food contact
surface
applications
[45]
ZnO
Incorporation method:
Laser 3D molding
3HV fraction: NA
P(3HB-co-3HV)-ZnO (100:0→95:5)
Melting temperature: 171 → 158 °C
Decomposition temperature: 261.2 → 288.7 °C
Strain:14.0% → 9.5%
Stress: 3.5 → 4.5 MPa
Compression strength: 4 → 5 MPa
Compression modulus: 60 → 80 MPa
Bacterial inhibition rate (day 5): 2.5% → 79.0%
Zn2+ release in deionized water (day 7): 0.19 → 0.34 mg L−1
Bone
repair
[46]
A Synthetic atactic poly(3-hydroxybutyrate) (α-P(3HB)), bovine serum albumin capped silver (Ag/BSA), ascorbic acid (AS), cellulose nanocrystals (CNC), carbon nanotubes (CNT), dicumyl peroxide (DCP), distillers’ dried grains with solubles (DDGS), hydroxyapatite (HA), organophilic attapulgite (MAT), Miscanthus (Misc), 2-methyl-4-chlorophenoxyacetic acid (MCPA), monomethoxy poly(ethylene glycol) (mPEG), natural rubber (NR), poly(ε-caprolactone) (PCL), poly(d,l-lactide) (PDLLA), poly(butylene succinate) (PBS), poly(butylene adipate-co-terephthalate) (PBAT), poly(ethylene glycol) (PEG), poly(2-hydroxyl ethyl methacrylate) (PHEMA), poly(lactic acid) (PLA), poly(propylene carbonate) (PPC), not available (NA). B Melt-mixed compress molded P(3HB-co-3 mol%3HV):P(3HB-co-18 mol%3HV):ZnO (70:24:6) (PHBVs-D), electrospun P(3HB-co-18 mol%3HV):ZnO (50:50) (PHBVs-P), P(3HB-co-18 mol%3HV):ZnO (50:50) coating on compressed molded P(3HB-co-3 mol%3HV) (PHBVs-C).
Table 2. P(3HB-co-3HV) production by bacteria from various 3HV precursors.
Table 2. P(3HB-co-3HV) production by bacteria from various 3HV precursors.
Microorganisms and
Carbon Sources
Biomass
(g/L)
PHA Content3HV Composition3HV Yield
(g/g)
Ref.
(wt%)(g/L)(mol%)(g/L)
Organic acids
Bacillus aryabhattai PHB10
Glucose (20.0 g/L)
Propionic acid (0.7 g/L)
3.9722.8---[28]
Bacillus thuringiensis R-510
Glucose (23.5 g/L)
Propionic acid (1.0 g/L)
2.9210.6410.20.25[69]
C. necator DSM 545
Waste glycerol (20.0 g/L)
Propionic acid (4.0 g/L)
4.5572.6250.70.16[70]
C. necator DSM 545
Butyric acid (246.0 g/L)
Propionic acid (186.0 g/L)
65.988583620.80.11[71]
C. necator NRRL B 14690
Fructose (40.0 g/L)
Propionic acid (4.0 g/L)
8.2736.0231.40.35[72]
C. necator NCIMB 11599
Glucose (maintained at 10.0–20.0 g/L)
Propionic acid (0.52 mol/mol glucose)
112.35764.01415.7-[73]
Erwinia sp. USMI-20
Palm oil (4.6 g/L)
Propionic acid (1.9 g/L)
4.2401.7340.60.30[60]
Activated sludge mixed culture
Acetic acid, lactic acid, propionic acid
---31–66--[74]
Bacillus cereus RCL 02
Glucose (25.0 g/L)
Valeric acid (1.9 g/L)
8.1725.8150.90.46[75]
C. malaysiensis USMAA9-39
Oleic acid (6.5 g/L)
Valeric acid (0.9 g/L)
5.2432.2170.40.42[76]
C. necator DSM 545
Waste glycerol (20.0 g/L)
Valeric acid (4.0 g/L)
5.3643.4311.10.26[70]
C. necator NRRL B 14690
Fructose (40.0 g/L)
Valeric acid (4.0 g/L)
7.2402.9621.80.45[72]
Erwinia sp. USMI-20
Palm oil (4.6 g/L)
Valeric acid (2.0 g/L)
4.8341.6470.30.14[60]
Methylobacterium organophilum NCIB 11278
Methanol (4.0 g/L)
Valeric aid (0.5 g/L)
2.5501.3100.10.25[77]
Burkholderis sp. IS-01
Gluconate (20.0 g/L)
Levulinic acid (12.5 g/L)
5.9623.7873.20.25[67]
C. necator KHB-8862
Fructose syrup (20.0 g/L)
Levulinic acid
(1.0 g/L, initial and 3 times feeding)
8.6847.2282.00.50[66]
C. necator H16
Fructose (20.0 g/L)
Levulinic acid (3.5 g/L)
7.3483.5160.60.16[11]
Hydrogenophaga pseudoflava DSM 1034
Whey permeate (47 mL/L)
Levulinic acid (1.0 g/L)
4.5492.2451.01.00[68]
Conjugate bases of organic acids
Caldimonas taiwanensis
Sugars (1.5%)
Valerate (0.5 g/L)
1.6–4.142–670.8–2.110–130.1–0.20.16–0.49[78]
Methylocystis dominated mixed culture
Methane gas
(repeating 48 h fed-batch cycle)
Valerate (0.4 g/L)
1.5300.5390.20.45[79]
Sodium salts of organic acids
Azohydromonas lata
Rice wastewater (21 g/L)
Sodium acetate (10 g/L)
5.0321.660.10.01[80]
Corynebacterium glutamicum ATCC13869 transformant A
Sodium propionate (1.0 g/L)
-31-28--[81]
C. necator H16
Sodium acetate (0–20 g/L)
Sodium propionate (0–20 g/L)
0.3–0.712–56Trace0–45Trace-[82]
C. necator PHB4 C
Palm kernel oil (5.0 g/L)
Sodium propionate (5.0 g/L)
3.0300.9120.10.02[83]
Herbaspirillum seropedicae Z69Prp D
Glucose (7.0 g/L)
Sodium propionate (0.5 g/L)
2.4370.9140.10.25[84]
C. necator H16
Plant oils (5.0 g/L)
Sodium valerate (5.0 g/L)
4.1–6.164–892.1–5.43–140.1–0.90.03–0.17[85]
C. necator PHB4 C
Palm kernel oil (5.0 g/L)
Sodium valerate (1.0 g/L)
4.2522.260.10.13[83]
Methylocystis parvus OBB3
Methane gas (75 mL)
Sodium valerate (1.0 g/L)
--0.3-0.20.20[86]
Alkyl alcohols
C. necator H16
Waste rapeseed oil (20.0 g/L)
1-propanol (8.0 g/L)
14.78011.791.10.14[13]
Erwinia sp. USMI-20
Palm oil (4.6 g/L)
1-propanol (2.3 g/L)
5.4502.760.20.07[60]
C. malaysiensis USMAA2-4
Oleic acid (6.5 g/L)
1-pentanol (0.9 g/L)
5.1402.180.20.22[87]
C. malaysiensis USMAA2-4ABH16 B
Palm olein (6.5 g/L)
1-pentanol (0.9 g/L)
5.4693.770.30.33[87]
C. malaysiensis USMAA1020
Oleic acid (6.5 g/L)
1-pentanol (1.3 g/L)
-76-10--[88]
Erwinia sp. USMI-20
Palm oil (4.6 g/L)
1-pentanol (1.4 g/L)
4.8623.0200.60.43[60]
Massilia haematophila UMTKB-2
Glucose (16.0 g/L)
1-pentanol (1 g/L)
--5.070.40.40[89]
Methylobacterium extorquens G10
Methanol (fractional supply by 5–20 mL)
1-pentanol (fractional supply by 2%–20% v/v methanol)
25–4030–457.5–18.014–502.5–4.5-[90]
Methylocystis sp. WRRC1
Methane gas (75 mL)
1-pentanol (1.0 g/L)
--0.3-0.20.17[86]
Methyloligella halotolerans C2
Methanol (5–20 mL fractional supply)
1-pentanol (fractional supply by 5–15% v/v methanol)
-49–98-2–51--[91]
P. denitrificans ATCC 17741
1-pentanol (maintained at 1.6 g/L)
6.8181.21001.2-[14]
Mixed precursors
C. necator DSM 545
Levulinic acid (1.0 g/L)
Sodium propionate (2.5 g/L)
1.0330.3730.20.24[65]
C. necator DSM 545
Levulinic acid (1.0 g/L)
Sodium propionate (1.0 g/L)
0.5190.178Trace-[10]
H. pseudoflava DSM 1034
Whey permeate (47.0 mL/L)
Levulinic acid
(0.5 g/L, initial and 3 times feeding)
Sodium valerate
(1.0 g/L, initial and 3 times feeding)
6.6674.4552.40.43[68]
Only the most promising condition was included for studies involving multiple cultivation conditions. Trace (concentration < 0.1 g/L). A C. glutamicum ATCC13869 transformant harboring C. necator phaCABRe genes. B C. malaysiensis USMAA2-4 transformant harboring C. necator H16 lipAB genes. C C. necator mutant with P(3HB)-negative phenotype [92]. D H. seropedicae Z69 with the 2-methylcitrate synthase (PrpC) gene eliminated.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wong, H.S.J.; Bhubalan, K.; Amirul, A.-A.A. A Critical Review on the Economically Feasible and Sustainable Poly(3-Hydroxybutyrate-co-3-hydroxyvalerate) Production from Alkyl Alcohols. Polymers 2022, 14, 670. https://doi.org/10.3390/polym14040670

AMA Style

Wong HSJ, Bhubalan K, Amirul A-AA. A Critical Review on the Economically Feasible and Sustainable Poly(3-Hydroxybutyrate-co-3-hydroxyvalerate) Production from Alkyl Alcohols. Polymers. 2022; 14(4):670. https://doi.org/10.3390/polym14040670

Chicago/Turabian Style

Wong, Hau Seung Jeremy, Kesaven Bhubalan, and Al-Ashraf Abdullah Amirul. 2022. "A Critical Review on the Economically Feasible and Sustainable Poly(3-Hydroxybutyrate-co-3-hydroxyvalerate) Production from Alkyl Alcohols" Polymers 14, no. 4: 670. https://doi.org/10.3390/polym14040670

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

Article Metrics

Back to TopTop