Next Article in Journal
Enhanced Biodegradation Rate of Poly(butylene adipate-co-terephthalate) Composites Using Reed Fiber
Previous Article in Journal
Innovative Electrospun Nanofiber Mats Based on Polylactic Acid Composited with Silver Nanoparticles for Medical Applications
Previous Article in Special Issue
Biodegradation of Choline NTF2 by Pantoea agglomerans in Different Osmolarity. Characterization and Environmental Implications of the Produced Exopolysaccharide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Manipulating Microbial Cell Morphology for the Sustainable Production of Biopolymers

by
Vipin C. Kalia
1,
Sanjay K. S. Patel
1,
Kugalur K. Karthikeyan
1,
Marimuthu Jeya
2,
In-Won Kim
1 and
Jung-Kul Lee
1,*
1
Department of Chemical Engineering, Konkuk University, 120 Neungdong-ro, Gwangjin-gu, Seoul 05029, Republic of Korea
2
Marine Biotechnology Division, National Institute of Ocean Technology, Chennai 600100, India
*
Author to whom correspondence should be addressed.
Polymers 2024, 16(3), 410; https://doi.org/10.3390/polym16030410
Submission received: 11 January 2024 / Revised: 25 January 2024 / Accepted: 30 January 2024 / Published: 1 February 2024

Abstract

:
The total rate of plastic production is anticipated to surpass 1.1 billion tons per year by 2050. Plastic waste is non-biodegradable and accumulates in natural ecosystems. In 2020, the total amount of plastic waste was estimated to be 367 million metric tons, leading to unmanageable waste disposal and environmental pollution issues. Plastics are produced from petroleum and natural gases. Given the limited fossil fuel reserves and the need to circumvent pollution problems, the focus has shifted to biodegradable biopolymers, such as polyhydroxyalkanoates (PHAs), polylactic acid, and polycaprolactone. PHAs are gaining importance because diverse bacteria can produce them as intracellular inclusion bodies using biowastes as feed. A critical component in PHA production is the downstream processing procedures of recovery and purification. In this review, different bioengineering approaches targeted at modifying the cell morphology and synchronizing cell lysis with the biosynthetic cycle are presented for product separation and extraction. Complementing genetic engineering strategies with conventional downstream processes, these approaches are expected to produce PHA sustainably.

1. Introduction

Plastic products are used in almost all aspects of life and have become an integral part of our lives. Because of their unique thermochemical properties, such as their plasticity, adaptability, durability, and flexibility, they are convenient to use. Conventional plastics are derived from petroleum and natural gases. The high consumption rate of fossil fuels for their production has aggravated the energy crisis [1]. Being non-biodegradable, plastic waste accumulates in nature at a staggering rate, posing major management and environmental concerns [2]. Therefore, it poses a major threat to ecosystems and the environment. In addition, its negative impacts on human health cannot be ignored [3,4]. In 2020, the total amount of plastic waste was estimated to be 367 million metric tons. This number is anticipated to increase exponentially over the next few decades [5,6]. In principle, plastic waste can be managed through (a) recycling, which is not economically feasible, and (b) an energy recovery process, which emits greenhouse gases and toxic compounds, in addition to environmental pollution [7,8]. Therefore, strategies are needed to find alternatives to circumvent these issues.
The potential candidates envisaged to replace plastics are naturally biodegradable polymers. Based on biodegradability as the major criteria, the strong contenders are polycaprolactone (PCL), poly(butylene succinate), poly(ethylene succinate), poly(butylene adipate terephthalate, polyhydroxyalkanoates (PHAs), polylactic acid (PLA), starch, cellulosic esters, and proteins [9]. Of these, PHAs, PLAs, and starch are bio-based and -degradable polymers, which can be used for a wide range of biotechnological applications [10]. The various strategies for enhancing PHA production, include the genetic engineering of microbial cells to manipulate the cell morphology and size, modify metabolic pathways associated with PHA biosynthesis, downregulate unnecessary byproducts, and ease the recovery process [11,12,13,14,15,16]. In this article, the bioengineering of cell morphology and cell division machinery has been shown to have the potential to simplify downstream processing procedures and reduce production costs. In contrast to most of the published literature targeting sustainable PHA production strategies, the present article provides information on enhanced downstream process efficiency via exploiting genetic engineering to synchronize the processes of cell division, PHA biosynthesis, cell lysis, and product separation. The major emphasis is placed on developing methods for enhanced production and economic recovery for the sustainable production of biopolymers.
A bibliographic search was performed for research and review articles published in scientific journals and book chapters listed primarily on Scopus, PubMed, and Google Scholar, based on the following keywords: biopolymers, polyhydroxyalkanoates, inclusion bodies, cell morphology, cell division, genetic engineering, PHA biosynthesis, sustainability, down streaming, depolymerases, and recovery. The final text matter was written after critically reading and evaluating around 700 articles published primarily during the last ten years.

2. Biopolymers

Past research studies have focused on developing bioplastics based on renewable resources. Bioplastics and biopolymers can be produced from biomaterials of diverse origins [17,18]. The most used bioplastics are based on PHAs, PLA, and PCL. Other materials suitable for bioplastic production include starch, cellulosic esters, and proteins [9].

2.1. Polyhydroxyalkanoates

Polyhydroxyalkanoates are a large group of biodegradable thermoplastic polymers [19]. Many microbes that can produce PHA have been identified, including Bacillus, Cupriavidus, Pseudomonas, Aeromonas, engineered Escherichia coli, and Halomonas spp. [9,17,20,21,22,23]. Polyhydroxybutyrate (PHB) is a naturally produced biopolymer. PHB is also a homopolymer that is highly crystalline and brittle. These properties limit its range of biotechnological applications. In contrast, PHAs comprise more than 160 monomeric units [24]. PHAs are produced from pure chemicals and biowaste through fermentation under environmental stress conditions, primarily owing to the limitations of nitrogen, oxygen, phosphorus, and potassium. Pure microbial cultures and their consortia have been used to ferment diverse biowastes to produce copolymers of PHAs [22]. The diversity of monomeric units in copolymers of PHA confers them with properties like those of conventional plastics and has been demonstrated to have thermochemical properties suitable for applications in medicine, agriculture, and tackling diverse environmental issues [25,26,27,28]. A detailed description of these PHAs is presented in the following sections.

2.2. Polylactides

Polylactides are lactic acid (2-hydroxy propionic acid) polymers. Bacteria ferment carbohydrates to produce lactic acid. The biosynthesis of lactic acid is preferred because of the presence of L-stereoisomers in greater yields, leading to cheaper PLA production. These can be produced with diverse crystallinities, microstructures, and molar masses. However, they can also be synthesized chemically. Their synthesis can be completed through the polycondensation of lactic acid and polymerization of lactides via a ring-opening mechanism [29]. The major limitation is in the preparation and purification of pure lactic acid. During polymerization, certain undesirable side reactions take place. The presently available PLAs are based on linear macromolecules. For better physical features, there is a need to produce branched, dendrigrafic, or dendritic PLAs. It is difficult to produce high-molecular-weight PLAs with desirable mechanical properties, and the maximum molar masses have been limited to 6 × 104 [30]. PLA is used for packaging and producing fibers and fabrics. As a blend, its use can be extended to implants, screws, nails, and plates for applications in the medical sector [28,31].

2.3. Polycaprolactones

Polycaprolactones are synthesized from crude oils. Although they are synthetic polymers, they are biodegradable, hydrophobic, and semicrystalline. Their crystallinity can be manipulated by regulating the low-molecular-weight alcohols. These excellent blends have low melting points and high rates of solubility, absorbability, and compatibility. Their viscoelastic and rheological properties are superior to those of other polymers. These characteristics enable their application in tissue engineering. However, their use in medical devices is limited because only a few fungal and bacterial species contain enzymes that degrade them [32]. These materials have excellent long-term degradation applications, such as in drug delivery through encapsulation. Efficient drug delivery systems can be complexed with polylactic acid-co-glycolic acid, PLA, cellulose acetate, butyrate, and propionate [33].

2.4. Other Polymers

Starch-based polymeric films are composed of 5–30% starch along with water or plasticizers (sorbitol and glycerol). Such starch-based biodegradable thermoplastics are both crystalline and amorphous. They can be produced easily in large quantities; however, they require high temperatures (91–180 °C) to melt and their biodegradation rate is limited to 30%. The starch-based bioplastic products include grocery bags, food and fruit packing trays, paper foam, egg boxes, and packaging electronic devices [34,35].
Cellulose is the most abundant naturally produced biodegradable polymer. Cotton linters and wood pulp are the major contributors to polysaccharide cellulose. Cellulose-based films are easy to produce, whereas bioplastics are difficult to make [36,37,38]. Cellulosic bioplastics are produced on an industrial scale as esters or ether derivatives [39]. The generally produced derivatives include nanofiber cellulose, cellulose nanocrystals, cellulose acetate, cellulose acetate butyrate, and biopolyethylene. There is a critical requirement for additives in the production of thermoplastics [36,38]. Cellophane, either transparent or pigmented, is commonly used to wrap candies and flowers, for lamination, and to pack food products such as coffee, cheese, and chocolate [28].
Protein-based polymers are heteropolymers of amino acids with unique characteristics, such as high mechanical strength and the ability to resist the diffusion of gases and aromatic compounds. Blending these polymers with keratin results in a mechanically strong, highly thermostable, and flame-resistant material. Biopolymers can be produced via physicochemical and thermoplastic processes [40,41]. Their availability in large quantities, high nutritional value, biodegradability, and use to prepare films make them highly desirable for the packaging industry [41]. Their use as matrices for the well-regulated release and immobilization of enzymes and their high-water retention capacity widen the scope of their application in agriculture, horticulture, and health. The time-resolved small-angle X-ray scattering technique allows in situ observations, particularly of molecular rearrangements and orientations [42].
Polyamide 11 is a biopolymer produced from renewable sources. This polymer has greater longevity than other biopolymers. Its unique thermochemical properties, such as its high stability and melting point (200 °C), make it suitable for manufacturing in industrial sectors. Its other desirable properties include its resistance to chemicals, water, oil, salts, fuels, and radiation. It has strong resistance to cracks and abrasion. Other features of this polymer include its low rigidity and poor resistance to volatile fatty acids and phenols, radiation, and heat. Its applications range from the aerospace and automobile sectors (electrical cables, water tubing, and natural gas piping) to other industries, such as footwear, metal badminton racket strings, coatings, and shuttlecocks. Although it is non-biodegradable, its recycling efficiency rate is high. Compared with polyamides, it is costly [41,43].
Spider silk is fabricated from renewable resources using spiders. It has unique mechanical properties such as a strength of up to 1.7 GPa (gigapascal) and an extensibility rate of up to 500%. It has a natural capacity to self-assemble into hydrogels, films, capsules, and spheres but has several limitations, including its poor solubility and storage and assembly abilities [44]. However, it is still suitable for manufacturing sports goods, ropes, textiles, robotic components, and composite materials. In the medical sector, its applications range from wound healing and the bridging of nerve defects to fascial replacements [39,45,46]. At present, the commercial-scale production of this biopolymer is limited by the following factors: (i) the cannibalistic nature of spiders, (ii) the farming of the spiders, (iii) the time-consuming nature of silk harvesting, and (iv) low yields. The silk protein spidroin is genetically expressed in microbes, such as bacteria, yeast, plants, and animals [36,46,47].
The demand for PHAs is increasing because of their diverse potential applications in agriculture, medicine, and the environment. The primary issues with PHAs are their production costs and the difficulties in recovering intracellularly produced biopolymers. These targets can be achieved through (i) bioengineering the morphology and size of microbes and (ii) downstream recovery. Different strategies are anticipated to contribute to the sustainable production of PHAs.

3. Sustainable Production of Biopolymers

The scope of the biotechnological applications of PHAs has been reported in the fields of medicine, horticulture, agriculture, and food, which has increased the demand for PHAs. However, their production on an industrial scale is restricted by their poor quality and the high cost of production using natural PHA-producing bacteria. In addition, there are issues with the polymers’ instability and the high variability in their thermomechanical properties [20,47,48]. Most research efforts have targeted the screening of naturally occurring microbes with high PHA yields, exploiting biowastes to produce copolymers [24,49,50]. Recently, researchers have shifted their focus to genetic engineering, genome mining, genome reduction, and genome editing [11]. For sustainable PHA production, there are a few possibilities that can be adopted. These proposals include modifying the cell morphology, downregulating the associated biosynthetic pathways, and focusing on the downstream process [12,13,14,15,16]. Here, we focused on genetic modifications to manipulate the cell morphology, size, and extraction of intracellularly produced PHAs.
Bioprocesses can be divided into two major groups—upstream and downstream [13]. Upstream metabolic processes are limited by the cost of the substrate, pretreatment, and energy input. Other important components are the microbial cell factories. The screening and selection of robust organisms that can resist environmental stress can be complemented by the genetic modification of the microbial strain [51]. Equally important are the costs and parameters of the downstream processes [52,53]. The recovery and purification of intracellular biomolecules through various physiochemical processes are more expensive than for extracellularly produced bioproducts [54]. Overall, for the sustainable production of polymers, it is important to develop methods for enhanced production and economic recovery [49].

3.1. Bioengineering of Microbial Cells

The manipulation of microbes using genetic engineering and systems biology has been envisaged to be economical. Efforts have been focused on modifying the cell morphology and size (Figure 1) [55,56,57,58,59]. This strategy enables the greater accumulation of intracellular molecules, making the separation process convenient and economical [52].

3.1.1. Cell Morphology

The small size of bacterial cells makes their separation from the broth challenging. Furthermore, the rigid cell wall is a major limitation for the greater accumulation of inclusion bodies. These factors substantially increase the separation costs. A weak cell wall structure is likely to allow cells to expand, thereby providing more space for the storage of PHA granules. Large-sized cells can reduce recovery costs. Furthermore, to allow for the greater accumulation of intracellular biomolecules, enlarged cells can prove helpful. The primary focus of cell morphology is the modification of the cell shape and size [49]. This cell morphology engineering strategy showed that PHA production using enlarged bacteria is an effective technique for enhancing the product yield and easing downstream activities.
Many genes are involved in cell wall synthesis and division, and those responsible for maintaining the cell shape are critical for the overall morphology (Table 1) (Figure 1) [60,61,62,63,64,65]. Several proteins are associated with these cellular characteristics. The cell wall component, peptidoglycan, is a network of cross-linked glycan chains [66]. It provides mechanical strength to the microbial cells, especially for withstanding environmental stresses. Many genes are involved in the biosynthesis of the cell wall (murA (encodes for UDP-N-acetylglucosamine enolpyruvoyl transferase), murC (encodes for UDP-N-acetylmuramate-alanine ligase), murD (encodes for UDP-N-acetylmuramoyl-L-alanine:D-glutamate ligase), murE (encodes for a ligase), mraY (encodes for phosphor-N-acetylmuramoyl-pentapeptide transferase), dxs (encodes for 1-deoxyxylulose-5-phosphate synthase), and glmU (encodes for N-acetylglucosamine-1-phosphate uridyltransferase)) [65], and the associated proteins include (i) the divisome (ftsZ (encodes for tubulin-like protein), ftsA, ftsW (encodes for a lipid II flippase), minC, minD, pbp1, pbp3, sulA, slmA, and zipA, (actin-related proteins)), (ii) hydrolases (pbp5, pbp7, ampD, amiA, and amiB), and (iii) the rod complex (mreB (the cytoskeletal protein), rodZ (transmembrane protein), rodA (the transglycosylase), pbp2 (penicillin-binding protein), and transpeptidase) [67,68,69]. The major advantages of these genetic manipulations are enhanced product accumulation and the ease of the downstream processes [55,70].
As the cell wall rigidity continues to limit PHB accumulation, manipulating the cell wall’s biosynthetic pathway remains a key goal. The cell rigidity can be manipulated by inserting the PgltA constitutive promoter for the gltA gene (which encodes for citrate synthase) before the genes responsible for cell wall synthesis. Using CRISPRi technology, the expressions of the following 10 genes involved in cell wall synthesis were downregulated: (i) ftsW, (ii) dxs, (iii) glmU, and (iv) idi, which encode for isopentenyl diphosphate isomerase; (v) pgi (encodes for phosphoglucose isomerase); (vi) murA, (vii) murC, (viii) murD, (ix) murE, and (x) mraY. In E. coli cells, the overexpression of the sulA gene resulted in Young’s modulus increases of 1.32- to 1.60-fold compared to the parent cells. Depending on the cell wall rigidity and thickness, the PHB accumulation rate was almost 4-fold (93%) in weakened cell walls compared to 25% in thickened cell walls [65].

3.1.2. Cell Division

Cell division via binary fission involves mass duplication and partitioning into two daughter cells via cytokinesis [71]. Inhibiting cytokinesis results in the formation of enlarged filamentous-shaped bacterial cells [72]. The most critical protein is FtsZ, which is responsible for forming a Z-shaped ring structure in the middle of the bifurcating cell. It interacts with other proteins to form a divisome [64]. Deleting the ftsZ gene inhibits FtsZ activity, leading to the abortion of the cell division process [73,74]. In addition, FtsZ protein interactions with other proteins encoded by sulA, minC, minD, slmA, and EzrA can help achieve higher PHA yields. The overexpression of sulA blocks FtsZ ring assembly and transforms the rod-shaped E. coli cells into filamentous cells shapes. This leads to the availability of a larger intracellular space [72,75]. This approach has been exploited to enhance the PHA yield in diverse bacteria (Table 2). In E. coli, filamentous cells showed 100% increases in PHB accumulation and total cell dry weight. An interesting feature of the cell enlargement strategy was the production of copolymers of PHA P(3HB-co-4HB) when using engineered E. coli harboring the phaCAB operon and the knockout of genes sad and gabD, as well as the essential genes folK and ispH. The resultant P(3HB-co-4HB) copolymer’s accumulation rate was greater than 78% of the cell dry weight. Thus, PHA facilitates recovery from the broth [75]. The overexpression of minC and minD is also instrumental in inhibiting FtsZ’s function [76]. Cell division enzymes (hydrolases) are necessary for daughter cell separation. The separation process is regulated by EnvC and NlpD, which regulate the amidase activity [77]. E. coli cells with deletions of envC and nlpD were found to inactivate amidase, resulting in their inability to separate daughter cells, leading to their filamentous shape [74]. This approach was further improved to obtain longer filamentous cells by overexpressing sulA, which inhibited the functioning of FtsZ. A switch from binary to multiple fission modes was observed in E. coli with a minCD deletion complemented by the overexpression of sulA. The mutants had 70% PHB storage capacity, which was higher than the 51% recorded with the wild type, along with a rate of 64% in filamentous cells growing in the binary fission mode. Thus, it supports the easy separation and greater recovery of PHB [74]. Halomonas bluephagenesis TD08 was engineered to overexpress minC and minD during the stationary phase of the cell cycle. The higher quantities of these proteins led to 1.4-fold longer cells, with the PHB content increasing from 69% to 82%, while the cells were 1.4-fold larger than those of the parent, attaining a cell size of a few hundred micrometers. Filamentous cells aggregate and settle in the fermenters within 12 h, facilitating cell separation without centrifugation or filtration [57].
Pseudomonas species are also well-known for their mcl-PHA-producing abilities. Initial efforts to manipulate the cell morphology did not enhance the mcl-PHA yield. Taking cues from other studies, many genes known to regulate the cell morphology, such as nlpD (peptidoglycan degradation) and mreB (cytoskeleton protein), as well as z-ring formation (ftsZ) and inhibition (sulA), were overexpressed in a minCD (regulates the z-ring location) knockout mutant of Pseudomonas mendocina NK-01, which increased the mcl-PHA yield by 45.62% (up from 0.28 to 0.41 g/L). This rate can reach up to 60.87% with the overexpressed mreB [78]. Apart from most studies focusing on the cell morphology and size, an additional feature that can complement these strategies and contribute to enhanced PHA yields is the manipulation of the PHA granule size. Phasin (PhaP), a protein located on the surfaces of PHA granules, regulates the granule number and size. The overexpression of minCD genes and deletion of phaP resulted in a lower number of PHA granules but enhanced the granular size. PHA granules of up to 10 μm were reported for the first time in Halomonas bluephagenesis TDH4-minCD-ΔphaP1. Thus, genetic engineering techniques have enabled the production of larger cells possessing larger PHA granules. In addition, the 4HB mol% in the PHA copolymer was 14% higher than that of the wild-type strain (Table 2) [16].
Table 1. Potential bacterial cell morphology and divisiome genes affecting polymer production.
Table 1. Potential bacterial cell morphology and divisiome genes affecting polymer production.
GeneGene ProductCell Function/ActivityReference
dxs 1-deoxyxylulose-5-phosphate synthase Cell wall synthesis[65]
glmU N-acetylglucosamine-1-phosphate uridyltransferaseCell wall synthesis
murA UDP- N -acetylglucosamine enolpyruvoyl transferase Cell wall synthesis
murC UDP-N-acetylmuramate-alanine ligaseCell wall synthesis
murD UDP-N-acetylmuramoyl-L-alanine:D-glutamate ligaseCell wall synthesis
murE Ligase Cell wall synthesis
murJ Putative lipid II flippase Regulates peptidoglycan incorporation to the septum[67]
ftsZ Bacterial fission ring formation protein Recruiting divisiome proteins and z-ring stabilization[64,73,74]
ftsACell division proteinDivisiome[69]
ftsWPeptidoglycan glycosyltransferase, lipid II flippaseDivisiome[65]
ftsL, ftsN, ftsQCell division proteinsDivisiome[67]
sulACell division inhibitor proteinDivisiome, induces FtsZ inhibition[72,75,78]
slmANucleoid-associated FtsZ binding proteinDivisiome[60]
minCZ-ring positioning proteinDivisiome: actin-related proteins, inhibits FtsZ polymerization[57,76]
minDZ-ring positioning proteinDivisiome: actin-related proteins, recruits MinC[57,76]
envCMurein hydrolase activatorDivisiome[74]
zipAIntegral inner membrane proteinDivisiome[62]
PBP1, PBP3Penicillin binding proteinsDivisiome[68]
envCRegulate amidase activityCell division[68,77]
nlpDMurein hydrolase activator, peptidoglycan degradationCell division[74,77,78]
mreBDynamic cytoskeletal proteinRod complex and cell division[55,59,72]
RodZTransmembrane proteinRod complex[69]
RodATransglycosylase, lipid II flippaseRod complex[63]
PBP2Penicillin binding protein, murein DD-transpeptidaseRod complex, cell elongation[68]
gltACitrate synthaseManipulate cell rigidity[65]
idi Isopentenyl diphosphate isomerase Cell wall synthesis[65]
mraY Translocase 1, phosphor-N-acetylmuramoyl-pentapeptide transferaseCell wall synthesis[65]
pgi Phosphoglucose isomerase Cell wall synthesis[65]
PBP5, PBP7DD-carboxypeptidases and DD-endopeptidasesHydrolases[61]
ampD, amiA, amiB MurNAc-L-Ala amidases Hydrolases[61]
Table 2. Genetic manipulation of cell morphology for enhanced biopolymer production.
Table 2. Genetic manipulation of cell morphology for enhanced biopolymer production.
OrganismGene Edited bCharacteristics AffectedImpact on Polyhydroxyalkanoate c ProductionReference
Escherichia coli JM109SGIKsulATransformation of rod to filamentous cell with larger internal spacePHB accumulation showed 100% increase[75]
E. coli JM109SGIKsad, gabD, ispH folk, and sulATransformation of rod to filamentous cell with larger internal spaceCopolymers of PHA [P(3HB-co-4HB)] were 10% higher (78% in cell dry weight, CDW).[75]
E. coli JM109SG (ΔmreB/pTK-mreB/pBHR68) aftsZ, mreB, and sulAEnlarged cell space due to reduced restriction on space. Larger volume to size ratio.PHB d production was observed to increase from 5.72 g/L to (9.29 g/L, with a yield of 73.53% of CDW) e in a shake flask[59]
E. colienvC and nlpDSwitch from binary to multiple fission modePHB storing capacity enhanced from 51 to 70%[74]
E. coli JM109ftsZ and mreBEnlarged cell volumeEnhanced PHB accumulations (up to 80%)[73]
E. coli JM109ftsW, dxs, glmU, idi, pgi, murA, murC, murD, murE, and mraYCell wall thickeningPHB accumulation of 93% in weakened cells and 25% in thickened cell walls[65]
Pseudomonas mendocina NK-01ftsZ, mreB, sulA, minCD, and mreBModified bacterial shape and growth patternIncreased mcl-PHA f yield by 45.62% and up to 60.87%[78]
Halomonas bluephagenesis TD08minCDEnlarged cells (1.4-fold longer than the parent)PHB content enhanced from 69 to 82%[57]
Halomonas campaniensis LS21ftsZ and mreBEnlarged cell morphologyIncrease in PHB yield accompanied by normal growth[72]
H. bluephagenesis TDH4-minCD-ΔphaP1phaP1, phaP2, phaP3, and minCDBigger PHA granules and larger cell sizePHA granules up to 10 μm. PHA copolymer with 14% higher 4HB mol%[16]
Note: a: E. coli strains were engineered using the phbCAB operon encoding PHA synthase, beta-ketothiolase, and acetoacetyl-CoA-reductase; b: genes (encoding enzyme): dxs, (1-deoxyxylulose-5-phosphate synthase); envC, (murein hydrolase activator); ftsL, ftsN, and ftsQ, (cell division proteins); ftsW, (peptidoglycan glycosyltransferase); ftsZ, (bacterial fission ring formation protein); glmU, (N-acetylglucosamine-1-phosphate uridyltransferase); gltA, (citrate synthase); idi, (isopentenyl diphosphate isomerase); minC and minD (Z-ring positioning protein); mraY, (translocase 1); murA, (UDP-N-acetylglucosamine enolpyruvoyl transferase); murC, (UDP-N-acetylmuramate-alanine ligase); murD, (UDP-N-acetylmuramoyl-L-alanine:D-glutamate ligase); murE (ligase); mreB (dynamic cytoskeletal protein); nlpD (murein hydrolase activator); pgi,(phosphoglucose isomerase); sulA (cell division inhibitor protein); c: polyhydroxyalkanoate; d: polyhydroxybutyrate; e: cell dry weight; f: medium chain length PHA.

3.1.3. Cytoskeletal Protein

The bacterial cell morphology is also influenced by the cytoskeletal proteins encoded by the genes mreB, mreC, mreD, and rodZ [79]. The binding of MreB to the cytoplasmic membrane is regulated by PBP2, a cell wall biosynthesis enzyme [80]. The overexpression, deletion, and disruption of MreB and its associated proteins result in an abnormal cell morphology [81]. MreB deletion transforms rods into spherical cells. This transformation results in an enhanced volume-to-size ratio, enabling a larger space to accommodate intracellular molecules. Despite this benefit, the cell growth is drastically reduced, leading to a lower PHA yield. However, engineering the PHB biosynthetic pathway and expressing mreB in E. coli JM109SG at a low level helped restore the cell shape, improve the rigidity, and enhance the PHA yield by 60% (Table 2) [55,59]. These improved features can be exploited for the easy recovery of the cell biomass.
Growth retardation is a major limiting feature that is frequently encountered when engineering cell shapes. Thus, the expression of the genes ftsZ and mreB may play a vital role in optimizing these parameters [69]. Halomonas campaniensis strain LS21 with deletions of ftsZ or mreB genes was complemented with a plasmid expression system for these two genes. This enabled the mutant bacteria to grow even at 30 °C. Switching the growth temperature to 37 °C restricted the expression of the genes carried by the plasmid. This strategy resulted in greater PHB accumulation. The basic advantage of morphologically engineered cells is their rapid ability to settle to the bottom of the bioreactor and facilitate cell separation [72]. Synthetic biology techniques using clustered regularly interspaced short palindromic repeats (CRISPR) and their interference (CRISPRi) are efficient approaches for genome editing. The CRISPR system comprises the Cas9 protein and a single guide RNA (sgRNA) [82]. CRISPRi mutates the Cas9 protein, thereby allowing DNA binding to interfere with the transcription process [83]. CRISPR-based regulation was used to interfere with the expression of ftsZ and mreB in E. coli. This resulted in the production of long fat cells. The cell length and width were controlled using different sgRNAs. A wide range of morphologically diverse cells with enlarged cell volumes enabled the accumulation of PHB by up to 80% [73].

3.2. Complementary Extraction Processes

The extraction of intracellular molecules is an energy-intensive and expensive process. The major bottleneck is the breakdown of the cell walls, particularly in Gram-positive microbes [20,84]. Several studies have attempted to manipulate cell lysis and synchronize it with the substrate metabolism [20,85]. E. coli cells are susceptible to lysis because of their high concentrations of intracellular bioproducts. The PHA biosynthesis operon from Cupriavidus necator was expressed in E. coli, whereby the PHA yield increased to 70% of the total dry cell mass. The presence of E. coli biomass in a treatment with 0.2 N NaOH at 30 °C for one hour allowed the easy recovery of PHAs [86].
The bacteriophage (E. coli phage λ) holin operates by increasing the cell membrane’s permeability, whereas endolysin (lysozyme) metabolizes the cell wall [85]. Bacteriophage-based lysis takes place in the absence of Mg2+. PHA synthesis in engineered E. coli proceeds in the presence of Mg2+. The ion concentration is synchronized with the PHB production cycle, which at undetectable levels activates the phage lytic system and releases PHAs [85,87]. E. coli was engineered to trigger autolysis under stress by introducing a synthetic ribosome binding site and λ phage SRRz gene. The autolysis system (pSEVA331 plasmid) from E. coli was transferred into H. campaniensis LS21 cells to generate the Halomonas strain DL. This facilitated the economic recovery of PHA [88]. A similar approach was used to recover mcl-PHAs from Pseudomonas putida KT2440 [89,90]. The recovery rate of the PHAs was 94.2% [91]. As Gram-positive bacterial cells are more difficult to lyse, the Bacillus amyloliquefacines phage endolysin and holin system were expressed in Bacillus megaterium via the E. coliBacillus subtilis shuttle vector pX. Here, xylose was used for yeoB expression to synchronize cell lysis with the exhaustion of glucose and termination of PHA biosynthesis [20,92,93,94].
Bioengineered enlarged cells tend to settle due to gravitational forces, making the cell separation process easy and convenient to manage [55,75,95]. The efficiency of the cell-harvesting process can be improved using self-segregating and flocculating agents [96,97]. The bioengineering of Halomonas campaniesis LS21 through the deletion of the etf operon leads to a reduced cell surface charge and greater hydrophobicity. The net gain is attributed to self-flocculation, resulting in energy savings. In this way, the biopolymer’s productivity was enhanced by 1.8-fold to 0.33 g/L/h [98]. Solvent-based PHA extraction involves the usage of alcohols, ketones, dimethyl carbonates, esters, and cyclic carbonates [99]. The extraction of PHAs using halogenated solvents such as chloroform results in the excellent recovery of high-purity polymers from the biomass [99]. The major limitation is the large quantities required for extraction. Finally, a precipitation step is required to obtain high-purity PHAs. Here, “PHA anti-solvents” such as acetone, ethanol, heptane, hexane, or methanol are added in excess, which reduces the solubility of the PHAs in the solvent. This method is frequently used for PHA recovery from (i) bacterial co-cultures of C. necator DSM 428 (short chain length, scl-PHA producer) and Pseudomonas citronellolis NRRL B-2504 (medium chain length, mcl-PHA producer) [100] and (ii) a halophilic yeast Pichia kudriavzevii VITNN02 (scl-PHA, PHBHV) [101]. Other methods for PHA extraction use “green” solvents such as alcohols, ketones, cyclohexane, and esters, while a few of the novel solvents include acetone and non-cyclic ketones. “Green” solvents have been shown to recover (i) scl-PHA [102] and (ii) PHA-copolymers (poly(3HDD-co-3HD-co-3HO-co-3HHx) from Pseudomonas chlororaphis at room temperature [103]. Bioengineering cells to produce either scl- or mcl-PHA helps ease the solvent extraction process for scl-PHA from C. necator DSM 428 and the halophilic yeast Pichia kudriavzevii VIT-NN02 [101] and mcl-PHA from P. citronellolis NRRL B-2504 [100]. Because of its unique characteristics, such as its thermal stability, low inflammability, and low vapor pressure, PHA’s recovery rate is as high as 98% [104,105,106].

4. Perspectives

The replacement of non-biodegradable plastics with biopolymers has been envisaged as an ecofriendly and economical approach, especially for producing high-value products such as those required for the medical sector [107]. Biological processes have several major benefits over conventional chemical processes. The most critical requirement is mild environmental conditions, which saves on energy, potentially making them more economical. Although bioprocesses are highly specific and efficient, they are slower than chemical processes. The low quantity of bioactive molecules such as PHA produced per cell adds to the cost of the bioprocesses, making them uneconomical and unsustainable. Several mechanisms can be adopted to make biopolymer production sustainable. The production cost is primarily linked to the feed and the recovery of the biopolymer [20]. Several studies have focused on exploiting the bacterial ability to produce polymers, such as PHAs, from waste biomasses of diverse biological origins [17,18]. This also helps improve the thermochemical properties by producing copolymers of PHA [22,25,26,27,28]. However, the research efforts have focused on manipulating the bacterial morphology, particularly the cell size and morphology [11,55]. These modifications have proven beneficial for enhancing the efficiency of bioprocesses, resulting in accelerated growth, a higher cell density, and increased PHA accumulation [56,57,58,59]. These features simplify the downstream processes, helping to achieve higher yields and reduce costs [69]. In this article, we present an update on the genetic engineering of a few genes and their products responsible for the cell morphology [55]. Another potential approach that can facilitate PHA recovery is to use genetically modified organisms as supplements. Nuclease-producing genes from Staphylococcus aureus were engineered into C. necator and Delftia acidovorans. Nuclease enzyme reduced the viscosity of the broth containing disrupted cell biomass [108]. The C. necator PHA operon transformed the E. coli from a non-PHA producer to a producer state. The recombinant strain had poor cell integrity after accumulating PHA at up to 70% of the DCW. Simplified stirring of the 0.2 N NaOH broth at 30 °C for 1 h enabled the recovery of PHA with 95% efficiency [86]. These aspects are beneficial for separating the cells from the broth. Studies on genetic engineering to synchronize the cell lysis and substrate exhaustion process can help to avoid potential losses due to the onset of biodegradation within cells by depolymerases [84,85,109]. It must be emphasized that there other mechanisms, especially modifications of metabolic pathways such as biosynthesis, the availability of nutrients, energy generation, and the PHA granule size [12,13,14,15,16], can complement these approaches and help make the process economical and sustainable.
Despite the potential benefits associated with cell morphology engineering leading to higher and more sustainable PHA production rates, there are quite a few challenges that need to be overcome. Efforts need to be made to improve the following processes: (i) shortening the cell cycle period; (ii) achieving higher growth rates and cell densities; (iii) modifying cell membranes to overcome the issue of low osmolarity rates; (iv) separating cells from broth by reducing the viscosity of the medium; (v) genetically regulating the termination of the PHA biosynthesis for product exhaustion, auto cell lysis, and the inhibition of PHA depolymerases; (vi) genetically modifying additional cell division genes and their associated proteins, especially hydrolases; (vii) searching for phages that can lyse bacterial cells; (viii) synchronizing PHA production with various processes for other value-added intracellular and extracellular products.

5. Conclusions

Plastics are some of the most popular and widely used synthetic polymers produced from natural gas and petroleum. Their non-biodegradable nature causes the accumulation of plastic waste, leading to environmental pollution. Given the limited fossil fuel reserves and the need to circumvent pollution problems, the focus has shifted to biodegradable biopolymers such as PHAs. Because they are produced by bacteria as intracellular inclusion bodies, their recovery and purification are critical for reducing the production costs. It is anticipated that bioengineering the cell morphology and cell division machinery can help simplify the downstream processes. Enlarged cell spaces allow for the greater accumulation of intracellular bodies and reduce the bioproduction cost. The bioengineering of the cell size to provide a larger intracellular space and the biosynthesis of PHA copolymers and other biomolecules can be complemented by engineering metabolic pathways and deleting the associated pathways that burden them. It must be realized that the TCA cycle is the main energy-generating mechanism in bacteria under normal environmental conditions. However, under stressed environmental conditions, especially with an abundance of C in the milieu, bacteria are provoked to store energy by curtailing the TCA cycle and diverting it towards PHA synthesis. The two pathways compete for acetyl-CoA and associated energy-generating reactions. Thus, it is recommended to regulate the energy metabolism, especially the flux of C, and prevent pathways that limit PHA synthesis. A few other potential mechanisms for enhancing PHA production are manipulating the PHA operon and restricting the expression of depolymerases. The overall benefit at the industrial scale is the sustainable production of biopolymers.

Author Contributions

Conceptualization, V.C.K.; methodology, V.C.K., S.K.S.P., K.K.K., M.J. and I.-W.K.; validation, V.C.K. and S.K.S.P.; formal analysis, V.C.K. and M.J.; resources, J.-K.L.; data curation, V.C.K.; writing—original draft, V.C.K. and J.-K.L.; writing—review and editing, V.C.K.; supervision, V.C.K. and J.-K.L.; funding acquisition, I.-W.K. and J.-K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) and funded by the Ministry of Science, ICT, and Future Planning (NRF-2022M3A9I3082366, RS-2023-00222078, NRF-2022R1I1A1A01073500).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.; Liu, S.; Huang, J.; Cui, R.; Xu, Y.; Song, Z. Genetic engineering strategies for sustainable polyhydroxyalkanoate (PHA) production from carbon-rich wastes. Environ. Technol. Innov. 2023, 30, 103069. [Google Scholar] [CrossRef]
  2. Atanasova, N.; Stoitsova, S.; Paunova-Krasteva, T.; Kambourova, M. Plastic degradation by extremophilic bacteria. Int. J. Mol. Sci. 2021, 22, 5610. [Google Scholar] [CrossRef]
  3. Moses, N.E.; Erhianoh, C.; Anih, C.E. Modelling and simulation of waste plastic power plant: A theoretical framework. Am. J. Chem. Eng. 2018, 6, 94–98. [Google Scholar] [CrossRef]
  4. Mahmoud, Y.S.; Belhanche-Bensemra, N.; Safidine, Z. Impact of microcrystalline cellulose extracted from walnut and apricots shells on the biodegradability of poly(lactic acid). Front. Mater. 2022, 9, 1005387. [Google Scholar] [CrossRef]
  5. Lebreton, L.; Andrady, A. Future scenarios of global plastic waste generation and disposal. Palgrave Commun. 2019, 5, 6. [Google Scholar] [CrossRef]
  6. Plastic Waste Management | Plastics and the Environment Series. Available online: https://www.genevaenvironmentnetwork.org/resources/updates/plastic-waste-management/ (accessed on 27 December 2023).
  7. Lonca, G.; Lesage, P.; Majeau-Bettez, G.; Bernard, S.; Margni, M. Assessing scaling effects of circular economy strategies: A case study on plastic bottle closed-loop recycling in the USA PET market. Resour. Conserv. Recycl. 2020, 162, 105013. [Google Scholar] [CrossRef]
  8. Kim, M.O.; Park, J.K.; Han, T.H.; Seo, J.; Lee, H. Influence of polyethylene terephthalate powder on hydration of portland cement. Polymers 2012, 13, 2551. [Google Scholar] [CrossRef] [PubMed]
  9. Costa, A.; Encarnação, T.; Tavares, R.; Todo Bom, T.; Mateus, A. Bioplastics: Innovation for green transition. Polymers 2023, 15, 517. [Google Scholar] [CrossRef] [PubMed]
  10. Haider, T.P.; Völker, C.; Kramm, J.; Landfester, K.; Wurm, F.R. Plastics of the future? The impact of biodegradable polymers on the environment and on society. Angew. Chem. Int. Edit. 2019, 58, 50–62. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, X.; Lin, Y.; Wu, Q.; Wang, Y.; Chen, G.Q. Synthetic biology and genome-editing tools for improving PHA metabolic engineering. Trends Biotechnol. 2020, 38, 7. [Google Scholar] [CrossRef] [PubMed]
  12. Li, D.; Chen, J.-C.; Chen, G.-Q. Controlling microbial PHB synthesis via CRISPRi. Appl. Microbiol. Biot. 2017, 101, 5861–5867. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, G.Q.; Jiang, X.R. Next generation industry biotechnology based on extremophiles. Curr. Opin. Biotechnol. 2018, 50, 94–100. [Google Scholar] [CrossRef] [PubMed]
  14. Ouyang, P.; Wang, H.; Hajnal, I.; Wu, Q.; Guo, Y.; Chen, G.Q. Increasing oxygen availability for improving poly(3-hydroxybutyrate) production by Halomonas. Metab. Eng. 2018, 45, 20–31. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Y.; Landick, R.; Raman, S. A regulatory NADH/NAD+ redox biosensor for bacteria. ACS Synth. Biol. 2019, 8, 264–273. [Google Scholar] [CrossRef] [PubMed]
  16. Shen, R.; Ning, Z.Y.; Lan, Y.X.; Chen, J.C.; Chen, G.Q. Manipulation of polyhydroxyalkanoate granular sizes in Halomonas bluephagenesis. Metab. Eng. 2019, 54, 117–126. [Google Scholar] [CrossRef] [PubMed]
  17. Anjum, A.; Zuber, M.; Zia, K.M.; Noreen, A.; Anjum, M.N.; Tabasum, S. Microbial production of polyhydroxyalkanoates (PHAs) and its copolymers: A review of recent developments. Int. J. Biol. Macromol. 2016, 89, 161–174. [Google Scholar] [CrossRef]
  18. Costa, S.S.; Miranda, A.L.; de Morais, M.G.; Costa, J.A.V.; Druzian, J.I. Microalgae as source of polyhydroxyalkanoates (PHAs) —A review. Int. J. Biol. Macromol. 2019, 131, 536–547. [Google Scholar] [CrossRef]
  19. Singh, M.; Kumar, P.; Patel, S.K.S.; Kalia, V.C. Production of polyhydroxyalkanoate co-polymer by Bacillus thuringiensis. Indian J. Microbiol. 2013, 53, 77–83. [Google Scholar] [CrossRef]
  20. Singh, M.; Patel, S.K.S.; Kalia, V.C. Bacillus subtilis as potential producer for polyhydroxyalkanoates. Microb. Cell Fact. 2009, 8, 38. [Google Scholar] [CrossRef] [PubMed]
  21. Balaji, S.; Gopi, K.; Muthuvelan, B. A review on production of poly β hydroxybutyrates from cyanobacteria for the production of bioplastics. Algal Res. 2013, 2, 278–285. [Google Scholar] [CrossRef]
  22. Singh, G.; Kumari, A.; Mittal, A.; Goel, V.; Yadav, A.; Aggarwal, N.K. Cost effective production of poly-β-hydroxybutyrate by Bacillus subtilis NG05 using sugar industry waste water. J. Polym. Environ. 2013, 21, 441–449. [Google Scholar] [CrossRef]
  23. Singh, A.K.; Mallick, N. Advances in cyanobacterial polyhydroxyalkanoates production. FEMS Microbiol. Lett. 2017, 364, fnx189. [Google Scholar] [CrossRef]
  24. Singh, M.; Kumar, P.; Ray, S.; Kalia, V.C. Challenges and opportunities for customizing polyhydroxyalkanoates. Indian J. Microbiol. 2015, 55, 235–249. [Google Scholar] [CrossRef] [PubMed]
  25. Peelman, N.; Ragaert, R.; De Meulenaer, B.; Adons, D.; Peeters, R.; Cardon, L.; Van Impe, F.; Devlieghere, F. Application of bioplastics for food packaging. Trends Fd. Sci. Technol. 2013, 32, 128–141. [Google Scholar] [CrossRef]
  26. Ray, S.; Kalia, V.C. Biomedical applications of polyhydroxyalkanoates. Indian J. Microbiol. 2017, 57, 261–269. [Google Scholar] [CrossRef] [PubMed]
  27. Kalia, V.C.; Ray, S.; Patel, S.K.; Singh, M.; Singh, G.P. The dawn of novel biotechnological applications of polyhydroxyalkanoates. In Biotechnological Applications of Polyhydroxyalkanoates; Springer: Singapore, 2019; pp. 1–11. [Google Scholar] [CrossRef]
  28. Venkatachalam, H.; Palaniswamy, R. Bioplastic World: A Review. J. Adv. Sci. Res. 2020, 11, 43–53. [Google Scholar]
  29. Pretula, J.; Slomkowski, S.; Penczek, S. Polylactides-methods of synthesis and characterization. Adv. Drug Deliv. Rev. 2016, 107, 3–16. [Google Scholar] [CrossRef] [PubMed]
  30. Chang, Y.A.; Waymouth, R.M. Ion pairing effects in the zwitterionic ring opening polymerization of δ-valerolactone. Polym. Chem. 2015, 6, 5212–5218. [Google Scholar] [CrossRef]
  31. Inkinen, S.; Hakkarainen, M.; Albertsson, A.C.; Södergård, A. From lactic acid to poly(lactic acid) (PLA): Characterization and analysis of PLA and Its precursors. Biomacromolecules 2011, 12, 523–532. [Google Scholar] [CrossRef]
  32. Vert, M. Degradable and bioresorbable polymers in surgery and in pharmacology: Beliefs and facts. J. Mater. Sci. Mater. Med. 2009, 20, 437–446. [Google Scholar] [CrossRef]
  33. Woodruff, M.A.; Hutmacher, D.W. The return of a forgotten polymer—Polycaprolactone in the 21st century. Prog. Polym. Sci. 2010, 35, 1217–1256. [Google Scholar] [CrossRef]
  34. Isroi, I.; Cifriadi, A.; Panji, T.; Wibowo, N.A.; Syamsu, K. Bioplastic production from cellulose of oil palm empty fruit bunch. IOP Conf. Ser. Earth Environ. Sci. 2017, 65, 012011. [Google Scholar] [CrossRef]
  35. El-malek, F.A.; Khairy, H.; Farag, A.; Omar, S. The sustainability of microbial bioplastics, production and applications. Int. J. Biol. Macromol. 2020, 157, 319–328. [Google Scholar] [CrossRef] [PubMed]
  36. Luengo, J.M.; García, B.; Sandoval, A.; Naharro, G.; Olivera, E.R. Bioplastics from microorganisms. Curr. Opin. Microbiol. 2003, 6, 251–260. [Google Scholar] [CrossRef]
  37. Karpušenkaite, A.; Varžinskas, V. Bioplastics: Development, possibilities and difficulties. Environ. Res. Eng. Manag. 2014, 68, 69–78. [Google Scholar] [CrossRef]
  38. Lackner, M. Bioplastics. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 1–41. [Google Scholar] [CrossRef]
  39. Atiwesh, G.; Mikhael, A.; Parrish, C.C.; Banoub, J.; Le, T.-A.T. Environmental impact of bioplastic use: A review. Heliyon. 2021, 7, e07918. [Google Scholar] [CrossRef]
  40. Nandakumar, A.; Chuah, J.A.; Sudesh, K. Bioplastics: A boon or bane? Renew. Sustain. Energy Rev. 2021, 147, 111237. [Google Scholar] [CrossRef]
  41. Nanda, S.; Patra, B.R.; Patel, R.; Bakos, J.; Dalai, A.K. Innovations in applications and prospects of bioplastics and biopolymers: A review. Environ. Chem. Lett. 2022, 20, 379–395. [Google Scholar] [CrossRef]
  42. Jerez, A.; Partal, P.; Martínez, I.; Gallegos, C.; Guerrero, A. Protein-based bioplastics: Effect of thermo-mechanical processing. Rheol. Acta. 2007, 46, 711–720. [Google Scholar] [CrossRef]
  43. Costa, A.A.; Gameiro, F.; Potêncio, A.; Silva, D.P.; Carreira, P.; Martinez, J.C.; Pascoal-Faria, P.; Mateus, A.; Mitchell, G.R. Evaluating the injection moulding of plastic parts using in situ Time-Resolved Small-Angle X-ray Scattering Techniques. Polymers 2022, 14, 4745. [Google Scholar] [CrossRef]
  44. Eisoldt, L.; Smith, A.; Scheibel, T. Decoding the secrets of spider silk. Materialstoday 2011, 14, 80–86. [Google Scholar] [CrossRef]
  45. Qiao, X.; Qian, Z.; Li, J.; Sun, H.; Han, Y.; Xia, X.; Zhou, J.; Wang, C.; Wang, Y.; Wang, C. Synthetic engineering of spider silk fiber as implantable optical waveguides for low-loss light guiding. ACS Appl. Mater. Interfaces 2017, 9, 14665–14676. [Google Scholar] [CrossRef] [PubMed]
  46. Johansson, J.; Rising, A. Doing what spiders cannot-a road map to supreme artificial silk fibers. ACS Nano 2021, 15, 1952–1959. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, G.Q.; Patel, M.K. Plastics derived from biological sources: Present and future: A technical and environmental review. Chem. Rev. 2012, 112, 2082–2099. [Google Scholar] [CrossRef]
  48. 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]
  49. Koutinas, A.A.; Vlysidis, A.; Pleissner, D.; Kopsahelis, N.; Lopez Garcia, I.; Kookos, I.K.; Papanikolaou, S.; Kwan, T.H.; Lin, C.S. Valorization of industrial waste and by-product streams via fermentation for the production of chemicals and biopolymers. Chem. Soc. Rev. 2014, 43, 2587–2627. [Google Scholar] [CrossRef]
  50. Chen, G.-Q.; Chen, X.-Y.; Wu, F.-Q.; Chen, J.-C. Polyhydroxyalkanoates (PHA) toward cost competitiveness and functionality. Adv. Ind. Eng. Polymer Res. 2020, 3, 1–7. [Google Scholar] [CrossRef]
  51. Ademakinwa, A.N.; Ayinla, Z.A.; Agboola, F.K. Strain improvement and statistical optimization as a combined strategy for improving fructosyltransferase production by Aureobasidium pullulans NAC8. J. Genet. Eng. Biotechnol. 2017, 15, 345–358. [Google Scholar] [CrossRef]
  52. Cheng, K.K.; Zhao, X.B.; Zeng, J.; Wu, R.C.; Xu, Y.Z.; Liu, D.H.; Zhang, J.A. Downstream processing of biotechnological produced succinic acid. Appl. Microbiol. Biotechnol. 2012, 95, 841–850. [Google Scholar] [CrossRef]
  53. Jungbauer, A. Continuous downstream processing of biopharmaceuticals. Trends Biotechnol. 2013, 31, 479–492. [Google Scholar] [CrossRef]
  54. Kreyenschulte, D.; Krull, R.; Margaritis, A. Recent advances in microbial biopolymer production and purification. Crit. Rev. Biotechnol. 2014, 34, 1–15. [Google Scholar] [CrossRef]
  55. Jiang, X.R.; Chen, G.Q. Morphology engineering of bacteria for bio-production. Biotechnol. Adv. 2016, 34, 435–440. [Google Scholar] [CrossRef] [PubMed]
  56. van Teeffelen, S.; Gitai, Z. Rotate into shape: MreB and bacterial morphogenesis. EMBO J. 2011, 30, 4856–4857. [Google Scholar] [CrossRef] [PubMed]
  57. Tan, D.; Wu, Q.; Chen, J.C.; Chen, G.Q. Engineering Halomonas TD01 for the lowcost production of polyhydroxyalkanoates. Metab. Eng. 2014, 26, 34–47. [Google Scholar] [CrossRef] [PubMed]
  58. Bisson-Filho, A.W.; Discola, K.F.; Castellen, P.; Blasios, V.; Martins, A.; Sforca, M.L.; Garcia, W.; Zeri, A.C.; Erickson, H.P.; Dessen, A.; et al. FtsZ filament capping by MciZ, a developmental regulator of bacterial division. Proc. Natl. Acad. Sci. USA 2015, 112, E2130–E2138. [Google Scholar] [CrossRef]
  59. Jiang, X.R.; Wang, H.; Shen, R.; Chen, G.Q. Engineering the bacterial shapes for enhanced inclusion bodies accumulation. Metab. Eng. 2015, 29, 227–237. [Google Scholar] [CrossRef]
  60. Bernhardt, T.G.; de Boer, P.A.J. SlmA, a Nucleoid-associated, FtsZ binding protein required for blocking septal ring assembly over chromosomes in E. coli. Mol. Cell. 2005, 18, 555–564. [Google Scholar] [CrossRef]
  61. Van Heijenoort, J. Peptidoglycan hydrolases of Escherichia coli. Microbiol. Mol. Biol. Rev. 2011, 75, 636–663. [Google Scholar] [CrossRef]
  62. Shiomi, D.; Niki, H. A mutation in the promoter region of zipA, a component of the divisome, suppresses the shape defect of RodZ-deficient cells. MicrobiologyOpen 2013, 2, 798–810. [Google Scholar] [CrossRef]
  63. Shiomi, D.; Toyoda, A.; Aizu, T.; Ejima, F.; Fujiyama, A.; Shini, T.; Kohara, Y.; Niki, H. Mutations in cell elongation genes mreB, mrdA and mrdB suppress the shape defect of RodZ-deficient cells. Mol. Microbiol. 2013, 87, 1029–1044. [Google Scholar] [CrossRef]
  64. Loose, M.; Mitchison, T.J. The bacterial cell division proteins FtsA and FtsZ self-organize into dynamic cytoskeletal patterns. Nat. Cell Biol. 2014, 16, 38–46. [Google Scholar] [CrossRef]
  65. Zhang, X.C.; Guo, Y.; Liu, X.; Chen, X.G.; Wu, Q.; Chen, G.Q. Engineering cell wall synthesis mechanism for enhanced PHB accumulation in E. coli. Metab. Eng. 2018, 45, 32–42. [Google Scholar] [CrossRef]
  66. Turner, R.D.; Hurd, A.F.; Cadby, A.; Hobbs, J.K.; Foster, S.J. Cell wall elongation mode in Gram-negative bacteria is determined by peptidoglycan architecture. Nat. Commun. 2013, 4, 1496. [Google Scholar] [CrossRef] [PubMed]
  67. Monteiro, J.M.; Pereira, A.R.; Reichmann, N.T.; Saraiva, B.M.; Fernandes, P.B.; Veiga, H.; Tavares, A.C.; Santos, M.; Ferreira, M.T.; Macario, V.; et al. Peptidoglycan synthesis drives an FtsZ-treadmilling-independent step of cytokinesis. Nature 2018, 554, 528–532. [Google Scholar] [CrossRef] [PubMed]
  68. van Teeffelen, S.; Renner, L.D. Recent advances in understanding how rod-like bacteria stably maintain their cell shapes. F1000Research 2018, 7, 241. [Google Scholar] [CrossRef] [PubMed]
  69. Wang, Y.; Ling, C.; Chen, Y.; Jiang, X.; Chen, G.-Q. Microbial engineering for easy downstream processing. Biotechnol. Adv. 2019, 37, 107365. [Google Scholar] [CrossRef] [PubMed]
  70. Singh, S.K.; Parveen, S.; SaiSree, L.; Reddy, M. Regulated proteolysis of a crosslink-specific peptidoglycan hydrolase contributes to bacterial morphogenesis. Proc. Natl. Acad. Sci. USA 2015, 112, 10956–10961. [Google Scholar] [CrossRef]
  71. Egan, A.J.; Vollmer, W. The physiology of bacterial cell division. Ann. N. Y. Acad. Sci. 2013, 1277, 8–28. [Google Scholar] [CrossRef] [PubMed]
  72. Jiang, X.R.; Yao, Z.H.; Chen, G.Q. Controlling cell volume for efficient PHB production by Halomonas. Metab. Eng. 2017, 44, 30–37. [Google Scholar] [CrossRef]
  73. Elhadi, D.; Lv, L.; Jiang, X.R.; Wu, H.; Chen, G.Q. CRISPRi engineering E. coli for morphology diversification. Metab. Eng. 2016, 38, 358e369. [Google Scholar] [CrossRef]
  74. Wu, H.; Chen, J.; Chen, G.Q. Engineering the growth pattern and cell morphology for enhanced PHB production by Escherichia coli. Appl. Microbiol. Biotechnol. 2016, 100, 9907–9916. [Google Scholar] [CrossRef]
  75. Wang, Y.; Wu, H.; Jiang, X.; Chen, G.Q. Engineering Escherichia coli for enhanced production of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) in larger cellular space. Metab. Eng. 2014, 25, 183–193. [Google Scholar] [CrossRef]
  76. Ghosal, D.; Trambaiolo, D.; Amos, L.A.; Lowe, J. MinCD cell division proteins form alternating copolymeric cytomotive filaments. Nat. Commun. 2014, 5, 5341. [Google Scholar] [CrossRef] [PubMed]
  77. Uehara, T.; Dinh, T.; Bernhardt, T.G. LytM-domain factors are required for daughter cell separation and rapid ampicillin-induced lysis in Escherichia coli. J. Bacteriol. 2009, 191, 5094–5107. [Google Scholar] [CrossRef] [PubMed]
  78. Zhao, F.; Gong, T.; Liu, X.; Fan, X.; Huang, R.; Ma, T.; Wang, S.; Gao, W.; Yang, C. Morphology engineering for enhanced production of medium-chain-length polyhydroxyalkanoates in Pseudomonas mendocina NK-01. Appl. Microbiol. Biotechnol. 2019, 103, 1713–1724. [Google Scholar] [CrossRef] [PubMed]
  79. Strahl, H.; Burmann, F.; Hamoen, L.W. The actin homologue MreB organizes the bacterial cell membrane. Nat. Commun. 2014, 5, 3442. [Google Scholar] [CrossRef] [PubMed]
  80. Morgenstein, R.M.; Bratton, B.P.; Nguyen, J.P.; Ouzounov, N.; Shaevitz, J.W.; Gitai, Z. RodZ links MreB to cell wall synthesis to mediate MreB rotation and robust morphogenesis. Proc. Natl. Acad. Sci. USA 2015, 112, 12510–12515. [Google Scholar] [CrossRef] [PubMed]
  81. Colavin, A.; Shi, H.; Huang, K.C. RodZ modulates geometric localization of the bacterial actin MreB to regulate cell shape. Nat. Commun. 2018, 9, 1280. [Google Scholar] [CrossRef]
  82. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex genome engineering using CRISPR/Cas systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef] [PubMed]
  83. Cress, B.F.; Jones, J.A. Rapid generation of CRISPR/dCas9-regulated, orthogonally repressible hybrid T7-lac promoters for modular, tuneable control of metabolic pathway fluxes in Escherichia coli. Nucleic Acids Res. 2016, 44, 4472–4485. [Google Scholar] [CrossRef]
  84. Peternel, S. Bacterial cell disruption: A crucial step in protein production. New Biotechnol. 2013, 30, 250–254. [Google Scholar] [CrossRef] [PubMed]
  85. Gao, Y.; Feng, X.; Xian, M.; Wang, Q.; Zhao, G. Inducible cell lysis systems in microbial production of bio-based chemicals. Appl. Microbiol. Biotechnol. 2013, 97, 7121–7129. [Google Scholar] [CrossRef] [PubMed]
  86. Choi, J.I.; Lee, S.Y. Efficient and economical recovery of poly(3-hy-droxybutyrate) from recombinant Escherichia coli by simple digestion with chemicals. Biotechnol. Bioeng. 1999, 62, 546–553. [Google Scholar] [CrossRef]
  87. Zhang, X.; Pan, Z.; Fang, Q.; Zheng, J.; Hu, M.; Jiao, X. An auto-inducible Escherichia coli lysis system controlled by magnesium. J. Microbiol. Methods 2009, 79, 199–204. [Google Scholar] [CrossRef]
  88. Hajnal, I.; Chen, X.; Chen, G.Q. A novel cell autolysis system for cost-competitive downstream processing. Appl. Microbiol. Biotechnol. 2016, 100, 9103–9110. [Google Scholar] [CrossRef]
  89. Martínez, V.; García, P.; García, J.L.; Prieto, M.A. Controlled autolysis facilitates the polyhydroxyalkanoate recovery in Pseudomonas putida KT2440. Microb. Biotechnol. 2011, 4, 533–547. [Google Scholar] [CrossRef]
  90. Borrero-De Acuna, J.M.; Hidalgo-Dumont, C.; Pacheco, N.; Cabrera, A.; Poblete-Castro, I. A novel programmable lysozyme-based lysis system in Pseudomonas putida for biopolymer production. Sci. Rep. 2017, 7, 4373. [Google Scholar] [CrossRef]
  91. Poblete-Castro, I.; Aravena-Carrasco, C.; Orellana-Saez, M.; Pacheco, N.; Cabrera, A.; Borrero-de Acuña, J.M. Engineering the osmotic state of Pseudomonas putida KT2440 for efficient cell disruption and downstream processing of poly(3-Hydroxyalkanoates). Front Bioeng Biotechnol. 2020, 8, 161. [Google Scholar] [CrossRef]
  92. Morita, M.; Tanji, Y.; Mizoguchi, K.; Soejima, A.; Orito, Y.; Unno, H. Antibacterial activity of Bacillus amyloliquefaciens phage endolysin without holing conjugation. J. Biosci. Bioeng. 2001, 91, 469–473. [Google Scholar] [CrossRef]
  93. Hori, K.; Kaneko, M.; Tanji, Y.; Xing, X.H.; Unno, H. Construction of self disruptive Bacillus megaterium in response to substrate exhaustion for polyhydroxybutyrate production. Appl. Microbiol. Biotechnol. 2002, 59, 211–216. [Google Scholar] [CrossRef]
  94. Salzberg, L.I.; Helmann, J.D. An antibiotic inducible cell wall associated protein that protects Bacillus subtilis from autolysis. J. Bacteriol. 2007, 189, 4671–4680. [Google Scholar] [CrossRef]
  95. Chatsungnoen, T.; Chisti, Y. Continuous flocculation-sedimentation for harvesting Nannochloropsis salina biomass. J. Biotechnol. 2016, 222, 94–103. [Google Scholar] [CrossRef] [PubMed]
  96. Liu, C.G.; Hao, X.M.; Lin, Y.H.; Bai, F.W. Redox potential driven aeration during very-high-gravity ethanol fermentation by using flocculating yeast. Sci. Rep. 2016, 6, 25763. [Google Scholar] [CrossRef]
  97. Cheng, C.; Zhang, M.M.; Xue, C.; Bai, F.W.; Zhao, X.Q. Development of stress tolerant Saccharomyces cerevisiae strains by metabolic engineering: New aspects from cell flocculation and zinc supplementation. J. Biosci. Bioeng. 2017, 123, 141–146. [Google Scholar] [CrossRef] [PubMed]
  98. Ling, C.; Qiao, G.Q.; Shuai, B.W.; Song, K.N.; Yao, W.X.; Jiang, X.R.; Chen, G.Q. Engineering self-flocculating Halomonas campaniensis for wastewaterless open and continuous fermentation. Biotechnol. Bioeng. 2018, 116, 805–815. [Google Scholar] [CrossRef] [PubMed]
  99. Koller, M. Established and advanced approaches for recovery of microbial polyhydroxyalkanoate (PHA) biopolyesters from surrounding microbial biomass. EuroBiotech. J. 2020, 4, 113–126. [Google Scholar] [CrossRef]
  100. Rebocho, A.T.; Pereira, J.R.; Neves, L.A.; Alves, V.D.; Sevrin, C.; Grandfils, C.; Freitas, F.; Reis, M.A.M. Preparation and characterization of films based on a natural P(3HB)/mcl-PHA blend obtained through the co-culture of Cupriavidus necator and Pseudomonas citronellolis in apple pulp waste. Bioengineering 2020, 7, 34. [Google Scholar] [CrossRef]
  101. Ojha, N.; Das, N. Process optimization and characterization of polyhydroxyalkanoate copolymers produced by marine Pichia kudriavzevii VIT-NN02 using banana peels and chicken feather hydrolysate. Biocat. Agri. Biotechnol. 2020, 27, 101616. [Google Scholar] [CrossRef]
  102. Koller, M.; Bona, R.; Chiellini, E.; Braunegg, G. Extraction of short-chain- length poly-(I-hydroxyalkanoates) (scl-PHA) by the “anti-solvent” acetone under elevated temperature and pressure. Biotechnol. Lett. 2013, 35, 1023–1028. [Google Scholar] [CrossRef]
  103. Cerrone, F.; Radivojevic, J.; Nikodinovic-Runic, J.; Walsh, M.; Kenny, S.T.; Babu, R.; O’Connor, K.E. Novel sodium alkyl-1, 3-disulfates, anionic biosurfactants produced from microbial polyesters. Colloid Surface B 2019, 182, 110333. [Google Scholar] [CrossRef]
  104. Tang, S.; Baker, G.A.; Zhao, H. Ether- and alcohol-functionalized task-specific ionic liquids: Attractive properties and applications. Chem. Soc. Rev. 2012, 41, 4030–4066. [Google Scholar] [CrossRef]
  105. Fujita, K.; Kobayashi, D.; Nakamura, N.; Ohno, H. Direct dissolution of wet and saliferous marine microalgae by polar ionic liquids without heating. Enzyme Microb. Tech. 2013, 52, 199–202. [Google Scholar] [CrossRef]
  106. Kobayashi, D.; Fujita, K.; Nakamura, N.; Ohno, H. A simple recovery process for biodegradable plastics accumulated in cyanobacteria treated with ionic liquids. Appl. Microbiol. Biotechnol. 2015, 99, 1647–1653. [Google Scholar] [CrossRef]
  107. Kalia, V.C.; Patel, S.K.S.; Lee, J.-K. Exploiting polyhydroxyalkanoates for biomedical applications. Polymers 2023, 15, 1937. [Google Scholar] [CrossRef] [PubMed]
  108. Rodríguez Gamero, J.E.; Favaro, L.; Pizzocchero, V.; Lomolino, G.; Basaglia, M.; Casella, S. Nuclease expression in efficient polyhydroxyalkanoates-producing bacteria could yield cost reduction during downstream processing. Bioresour. Technol. 2018, 261, 176–181. [Google Scholar] [CrossRef] [PubMed]
  109. Ray, S.; Kalia, V.C. Polyhydroxyalkanoate production and degradation patterns in Bacillus species. Indian J. Microbiol. 2017, 57, 387–392. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diagrammatic representation of the effects of bioengineering cell division genes on the cell morphology and size: (A) binary cell division; (B) enlarged cell formation through inhibition of Z-ring formation via the overexpression of sulA or deletion or downregulation of genes involved in the biosynthesis of cell walls (murA, murC, murD, murE, mraY, dxs, glmU, idi, pgi); (C) filamentous cell formation via multiple fission through the deletion and downregulation of genes involved in the divisome (ftsZ, ftsA, ftsW, minC, minD, pbp1, pbp3, sulA, slmA, and zipA), hydrolases (pbp5, pbp7, ampD, amiA, and amiB), and rod complex (mreB, rodZ, rodA, and pbp2). Cell densities (A–C) vary by size and type [55,56,57,58,59,60,61,62,63,64,65,66,67,68,69].
Figure 1. Diagrammatic representation of the effects of bioengineering cell division genes on the cell morphology and size: (A) binary cell division; (B) enlarged cell formation through inhibition of Z-ring formation via the overexpression of sulA or deletion or downregulation of genes involved in the biosynthesis of cell walls (murA, murC, murD, murE, mraY, dxs, glmU, idi, pgi); (C) filamentous cell formation via multiple fission through the deletion and downregulation of genes involved in the divisome (ftsZ, ftsA, ftsW, minC, minD, pbp1, pbp3, sulA, slmA, and zipA), hydrolases (pbp5, pbp7, ampD, amiA, and amiB), and rod complex (mreB, rodZ, rodA, and pbp2). Cell densities (A–C) vary by size and type [55,56,57,58,59,60,61,62,63,64,65,66,67,68,69].
Polymers 16 00410 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kalia, V.C.; Patel, S.K.S.; Karthikeyan, K.K.; Jeya, M.; Kim, I.-W.; Lee, J.-K. Manipulating Microbial Cell Morphology for the Sustainable Production of Biopolymers. Polymers 2024, 16, 410. https://doi.org/10.3390/polym16030410

AMA Style

Kalia VC, Patel SKS, Karthikeyan KK, Jeya M, Kim I-W, Lee J-K. Manipulating Microbial Cell Morphology for the Sustainable Production of Biopolymers. Polymers. 2024; 16(3):410. https://doi.org/10.3390/polym16030410

Chicago/Turabian Style

Kalia, Vipin C., Sanjay K. S. Patel, Kugalur K. Karthikeyan, Marimuthu Jeya, In-Won Kim, and Jung-Kul Lee. 2024. "Manipulating Microbial Cell Morphology for the Sustainable Production of Biopolymers" Polymers 16, no. 3: 410. https://doi.org/10.3390/polym16030410

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