Next Article in Journal
Encapsulation of Lactobacillus gasseri: Characterization, Probiotic Survival, In Vitro Evaluation and Viability in Apple Juice
Next Article in Special Issue
Crucial Residues of C-Terminal Oligopeptide C60 to Improve the Yield of Prebiotic Xylooligosaccharides by Truncated Mutation
Previous Article in Journal
Preparation of Highly Substituted Sulfated Alfalfa Polysaccharides and Evaluation of Their Biological Activity
Previous Article in Special Issue
Probiotic Effects and Metabolic Products of Enterococcus faecalis LD33 with Respiration Capacity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Microbial Synthesis of Poly-γ-Glutamic Acid: A Review

1
Institute of Food Science and Technology, Chinese Academy of Agricultural Sciences, No. 2 Yuan Ming Yuan West Road, Beijing 100193, China
2
Key Laboratory of Agro-Products Quality and Safety Control in Storage and Transport Process, Ministry of Agriculture and Rural Affairs, Chinese Academy of Agricultural Sciences, Beijing 100193, China
3
Key Laboratory of Agro-Products Processing, Ministry of Agriculture and Rural Affairs, Chinese Academy of Agricultural Sciences, Beijing 100193, China
*
Authors to whom correspondence should be addressed.
Foods 2022, 11(5), 739; https://doi.org/10.3390/foods11050739
Submission received: 25 January 2022 / Revised: 12 February 2022 / Accepted: 26 February 2022 / Published: 2 March 2022

Abstract

:
Poly-γ-glutamic acid (γ-PGA) is a natural, safe, non-immunogenic, biodegradable, and environmentally friendly glutamic biopolymer. γ-PGA has been regarded as a promising bio-based materials in the food field, medical field, even in environmental engineering field, and other industrial fields. Microbial synthesis is an economical and effective way to synthesize γ-PGA. Bacillus species are the most widely studied producing strains. γ-PGA biosynthesis involves metabolic pathway of racemization, polymerization, transfer, and catabolism. Although microbial synthesis of γ-PGA has already been used extensively, productivity and yield remain the major constraints for its industrial application. Metabolic regulation is an attempt to solve the above bottleneck problems and meet the demands of commercialization. Therefore, it is important to understand critical factors that influence γ-PGA microbial synthesis in depth. This review focuses on production strains, biosynthetic pathway, and metabolic regulation. Moreover, it systematically summarizes the functional properties, purification procedure, and industrial application of γ-PGA.

1. Introduction

γ-PGA is a biopolymer of D- and/or L-glutamic acid monomers by peptide bonds [1,2,3]. Commonly, γ-PGA is natural, water-soluble, edible, non-immunogenic, and biodegradable. Nowadays, due to specific biological properties, γ-PGA has already gained significant attention and has been widely applied into the fields of food processing, agricultural production, emerging medical treatment, cosmetics, and other fields [4,5,6,7,8,9,10,11,12,13].
Microbial synthesis is an extremely economical and effective way to produce γ-PGA [11]. Bacteria, archaea, and eukaryotes are all γ-PGA producers, but Bacillus species are the natural, dominant, and safely produced strains [7,8,14,15,16]. Commonly, the production strains are divided into glutamic acid-dependent and glutamic acid-independent strains according to whether there is a need to add glutamate acids externally [17]. Glutamic acid dependent strains are regarded as the better producers, as adding glutamic acids can significantly increase the yield [1,9,18]. Biosynthesis of γ-PGA is well established, but both productivity and yield are still the decisive factors for limiting its industrial application [11].
The molecular composition, molecular weight, yield, and synthesis productivity of γ-PGA directly affect its biological properties and industrial applications. Most recent research have been paid attention to metabolism regulation of γ-PGA synthesis. Some of the strategies may be promising in the future, including genetic manipulation, culture medium, and culture condition optimization, which are all expected to improve the yield and productivity [11].
This review mainly summarizes the recent advances about the molecular components, functional properties, microbial sources, synthetic pathway, metabolic regulation, purification processes, and industrial application of γ-PGA. Meanwhile, it reveals tough challenges of γ-PGA biosynthesis ahead and strategies for challenges.

2. The Functional Properties of γ-PGA

γ-PGA is polymerized from D-/L-glutamic acid monomers alone or both enantiomers by peptide bonds, which are synthesized by the α-amino of one glutamic acid and the γ-carboxylic acid of another one (Figure 1) [19,20,21]. The peptide bonds of γ-PGA are significantly different from the amide bonds of ordinary proteins, which are formed by one amino acid with α-amino group and another one with α-carboxylic acid group [19].
The molecular structure of γ-PGA is directly affected by the molecular composition, molecular weight, and order of the isomers. Based on the monomer types, γ-PGA consists of γ-L-PGA (homopolymers containing only L-glutamic acid), γ-D-PGA (homopolymers containing only D-glutamic acid), and γ-DL-PGA (random copolymer containing D- and L-glutamic acid) [8,12]. Generally, γ-PGA covers five types of structures: α-helix, β-sheet, random coil, helix-to-random coil transition, and enveloped aggregate [12,22,23]. The molecular structures are always influenced by the extraction process [1]. Degree of monomer polymerization reflects the molecular weight of γ-PGA, which is determined by the producers, media, and cultivation conditions [24,25]. The molecular weight of γ-PGA is generally in the range of 10 kDa to 2000 kDa [19]. The higher molecular weight, the higher viscosity and the more difficult it is for further purification preparation [3,5,11,12,26,27].
The structure and molecular component of γ-PGA decisively influence its functional properties. As an anionic biopolymer with both carbonyl and amide groups, γ-PGA possesses an array of specific properties: edible, strong water solubility and water retention, non-toxicity, high biodegradability, high cation exchange capacity, metal-chelating ability, strong antioxidant, antimicrobial peptides activities, and high resistance to thermal decomposition [21,28]. However, some functional properties of γ-PGA are not completely understood. Depending on the biological properties, γ-PGA might be used as food additives, e.g., taste-masking agents, texture modifier, thickeners, stabilizers, moisturizers, and probiotics cryoprotectant [12,22,29]. It may be an essential material for maintaining sensitive activities and is used as carriers, fillers, and adhesives in complex matrices, such as for drugs in the medical industry and tissue engineering [30,31,32,33,34,35,36,37,38]. Besides, in agriculture, it can be an ideal stabilizer, metal biosorbent in soil washing, an environment-friendly fertilizer synergist, and plant growth accelerator [24,38,39,40]. Additionally, it can be used in moisturizers in the cosmetics industry [41].

3. Microbial Sources of γ-PGA

Since γ-PGA was first discovered in a capsule of Bacillus anthracis by Bruckner and his co-workers in 1937, it has been discovered successively in archaea, bacteria, and eukaryotes, such as Natrialba aegyptiaca, Natronococcus occultus, and Fusobacterium nucleatum [14,15,42]. However, natural sources of γ-PGA have been occurring in the fermented products of Bacillus species for a long time, which contains B. subtilis, B. amyloliquefaciens, B. licheniformis, B. velezensis, and others [16]. B. subtilis and B. licheniformis are high potential microbial sources of γ-PGA [3,11]. Bacillus species, such as γ-PGA producers, can be separated into two types: glutamic acid-dependent and glutamic acid-independent strains; the former need an external supply of glutamate to produce γ-PGA, and the latter can produce γ-PGA without extra adding glutamate [1,9]. Most γ-PGA producers are known as glutamic acid-dependent strains, such as B. subtilis F-2-01, B. subtilis MR-141, B. subtilis C10, B. subtilis chungkookjang, B. subtilis GXA-28, B. licheniformis ATCC 9945a, B. licheniformis A35, and B. licheniformis WX-02 [24,43,44,45,46,47,48]. Relatively few strains, such as B. subtilis TAM-4, B. subtilis C1, B. subtilis 5E, B. licheniformis A35, B. licheniformis S173, and B. amyloliquefaciens LL3, are glutamic acid-independent strains [11,42,49,50,51]. Although, compared to glutamic acid-independent strains, glutamic acid-dependent strains are generally better producers and have attracted much attention to relatively high yield of γ-PGA, the current problem is high cost of production [1,9,10]. Instead, glutamic acid-independent strains are more desirable for realistic demands of yield and productivity. They mean not only low cost but also simple fermentation processes. However, lower γ-PGA productivity is the primary limitation [3].
Nowadays, most commercial γ-PGA is mainly synthesized by Bacillus species and safely used in various well-established processing fields [1,4,5,8,10]. B. subtilis, as a traditional model organism, is a class of gram-positive, endospore-forming, and rod-shaped bacteria. It has applications in such fields as food and medicine as a Generally Recognized as Safe (GRAS) microbial producer. Furthermore, most importantly, it possesses the best-characterized γ-PGA production abilities [1,11]. Both the genome and molecular genetics relationship of B. subtilis have been clarified, and the manipulation is well established. B. licheniformis is another type of gram-positive, endospore-forming bacteria. At present, it has been exploited for γ-PGA production [1]. In addition, B. methylotrophicus, B. anthracis, and B. thuringiensis also have the great productivity of γ-PGA [26,52]. However, these strains have less relevant studies about γ-PGA production [53]. They all cannot be the great producers because of some natural flaws; e.g., B. anthracis is a pathogenic bacterium, and it is not viable for industrial applications.
In fact, engineered microorganisms have also been used as a suitable device for γ-PGA biosynthesis [1,11,54,55]. Escherichia coli are the most used host for heterologous expression. C. glutamicum is a good native L-glutamic acid producer [5,19]. Therefore, it is regarded as a host for producing recombinant γ-PGA. Similarly, B. subtilis can also be the homologous host engineered to produce γ-PGA [5,19]. However, the final yield of γ-PGA is and remains far below that of native strains.

4. Synthetic Pathway of γ-PGA

The microbial synthetic pathway of γ-PGA principally covers three distinct stages: racemization, polymerization and transfer, and catabolism (Figure 2) [10,11,22]. Glycolysis, the Pentose Phosphate Pathway (PPP), Tricarboxylic Acid (TCA) cycle, amino acid metabolism, and glutamate synthesis all participate in γ-PGA biosynthesis [56].

4.1. The Racemization Pathway of γ-PGA

Racemization is the first step of anabolism. Under the catalytic action of glutamate racemase (RacE), exogenous or endogenous L-glutamate are converted directly to D-glutamate [11]. Then these glutamic acids monomers incorporate into the growing L-chain [15,21]. In B. subtilis, racE/glr and yrpC have been identified as two homologs of RacE coding genes, and racE/glr has proven to be essential for the direct conversion between L-glutamate and D-glutamate [57]. For indirect conversion pathway, with pyruvate as the precursor, three enzymes catalyze the transformation of pyruvate to D-glutamic acid (L-glutamic acid/pyruvic acid aminotransferase, D-glutamic acid/pyruvic acid aminotransferase, and alanine racemase) [1,20].
In the presence of large amounts of L-glutamic acid, Bacillus species will initiate the transformation pathway to produce D-glutamic acid. In L-glutamic acid-dependent strains, once the L-glutamate is imported from the medium into the cytoplasm, part of them is immediately converted to D-glutamic acid and involved in the downstream synthesis pathway [3]. In L-glutamic acid-independent strains, all L-glutamic acid required is generated from external carbon sources [19]. With catalytic activity of glutamate dehydrogenase (GDH), these strains produce L-glutamic acid by transforming citric acid into both isocitric acid and α-ketoglutaric acid in TCA cycle [19]. Or L-glutamic acid production relies on glutamine synthetase-glutamate synthase (GS-GOGAT) pathway. Pyruvic acid and α-ketoglutaric acid are mutually transmitted by aminotransferase [58]. C. glutamicum has also been well-established for γ-PGA biosynthesis. Clearly, it has two completely different pathways: the GS-GOGAT pathway regulated by GlnA, GltB, and GltA and the NADPH-dependent pathway regulated by RocR, RocG, and GudB (RocR, a transcriptional regulator of glutamate; RocG, expressed by rocR; GudB, glutamate dehydrogenase) [3,19].

4.2. The Polymerization and Regulation Pathway of γ-PGA

The next steps are polymerization and transfer. L-glutamic acid and D-glutamic acid monomers are transfered from cytoplasm to cell membrane with the membrane-embedded γ-PGA synthetase (Pgs) [10,11,18]. Pgs are responsible for assembling them into γ-PGA [5]. Pgs is encoded by an active membrane enzyme complex (four gene subunits operons: pgsB, C, A, and E) and the γ-PGA-release gene (PgsS) (Figure 3) [2,59]. PgsBCA has been identified as the unique mechanism of γ-PGA microbial synthesis in Bacillus species [5,60]. The expression level of PgsBCA has a direct effect on γ-PGA biosynthesis [19]. PgsB and PgsC make up the main components of the catalytic site. PgsA removes the long chains on the active site; it participates in the integration of γ-PGA long chains and later transfer [13,60]. PgsBCA is necessary for polymerizing and transporting γ-PGA to the compact cell membrane [12]. High concentrations of PgsB, PgsC, and PgsA can biosynthesize γ-PGA from a lack of PgsE; therefore, PgsE may be dispensable [60]. However, PgsE is essential for γ-PGA biosynthesis when Zn2+ is present because PgsBCA complex is extremely unstable and highly hydrophobic, and PgsE could affect PgsBCA disaggregation in Bacillus species [61]. The homologs of PgsBCA are YwsC and YwtAB in B. subtilis and CapBCA in B. anthracis [1,45]. γ-PGA can be anchored at the peptidoglycan of cellular surface with CapD catalysis or released into the medium with PgsS catalysis [52].
Meanwhile, γ-PGA biosynthesis is regulated by two intracellular signal transduction mechanisms: one is ComP-ComA regulator, and another is DegS-DegU regulator [62]. DegQ regulated by these two mechanisms inhibits the γ-PGA production and then effectively down-regulates the expression level of degradation enzymes [63]. SwrA is another regulator of PgsBCA. The fully activated PgsBCA operon needs the simulaneous presence of SwrA and phosphorylated DegU (DegU-P), but one of the two exists separately and does not have much impact on pgsBCA transcription [18,64]. High level of DegU-P could completely replace SwrA in directly activating PgsBCA expression [65]. Overall, these mechanisms are all involved in transcriptional regulation. Both ComP-ComA and DegS-DegU systems regulate the front-end transcription, while SwrA assists in regulation at the post-transcription [66].

4.3. The Catabolism Pathway of γ-PGA

γ-PGA is a secondary metabolite for the Bacillus species. However, production strains of γ-PGA may possibly degrade and utilize γ-PGA as substrates [24]. The catabolism pathway of γ-PGA is rigorously controlled by the intracellular regulatory mechanisms [19].
Three types of hydrolases can degrade γ-PGA: (1) native γ-PGA hydrolase (PgdS), whose coding gene is located in the downstream of PgsBCA operon, cleaves the γ-glutamyl bonds with strict regulation inside of hosts; (2) cell wall lyases (D- or L-endopeptidase), such as CwlO, CwlS, LytE, and LytF, disconnects between the glutamic acid residues peptide bonds; and (3) γ-glutamyl transferase (excision or incision enzyme, Ggt) hydrolyzes the γ-PGA from the N-terminal and releases shorter D- or L-glutamic acid residues to form di- and tripeptides of γ-glutamic acid in vitro [56,67,68,69].
The genes encoding Ggt (PgdS, YwtD, and Dep) are in plasmid and lie directly in the downstream of PgsBCA operon in the Bacillus species with the same orientation [67]. The above situation indicates that the enzymes are stably expressed under the ComP-ComA and DegS-DegU regulators.

5. Improvement of Microbial γ-PGA Synthesis

Although microbial synthesis of γ-PGA has been well developed, yield and productivity remain the core constraints for its industrial application [11]. Most cost of γ-PGA production depends on culture medium and producing strains [70]. Metabolic regulation, including genetic manipulation, culture medium, and culture condition optimization, is an effective way to solve the above bottleneck problems. However, the final yield of γ-PGA has so far not matched and even falls far below the productivity of the native producers [71].

5.1. Genetic Manipulation

Thus far, most native producers of γ-PGA are Bacillus species. Parts of them have great productivity. B. subtilis has been widely applied in industry by simple screening and improvement (Table 1 and Table 2) [1]. Production strains of γ-PGA have always been isolated from traditional fermented soybean products or some specific environments. However, the high production cost of glutamate-dependent strains and low substrate conversion rate of glutamate-independent strains should not be ignored. Genetic manipulation is a proven effective way to regulate the γ-PGA production [19]. Genetic manipulation is the key process of constructing genetically engineered strains. It mainly contains knock-out genes and construction plasmid. Genetic manipulation can promote or inhibit expression of enzymes and thus influences γ-PGA biosynthesis and degradation directly or indirectly in Bacillus species.
For native strains, genetic manipulation targeting γ-PGA synthesis pathway is a common strategy. Deletion the transcriptional regulating gene rocR or the glutamate dehydrogenase gene gudB individually can effectively increase the yield of γ-PGA in B. amyloliquefaciens LL3 [83]. In B. amyloliquefaciens, NK-A6 deleted fadR, lysC, aspB, pckA, proAB, rocG, and gudB for partial blockade of downstream metabolic pathways; B. amyloliquefaciens NK-A7 inserted a strong promoter PC2up for enhancing NADPH level by genetic manipulation; and B. amyloliquefaciens NK-A11 deleted the srf and itu operons, and these engineered bacteria synthesized 4.84, 6.46, and 7.53 g/L γ-PGA, respectively, all higher than the original strains of B. amyloliquefaciens LL3 [84]. Some strains, such as B. subtilis 168 and B. subtilis DB430, have the necessary synthesis gene cluster but cannot produce γ-PGA [59,85]. A mutant integrated PgsBCA or ywsC-ywtAB gene cluster with the regulating gene SwrA can increase intracellular level of DegU-P and SwrA and the production of γ-PGA. Besides, PgsBCA embedded into chromosome with a strong promoter can efficiently enhanced the production of γ-PGA.
Moreover, γ-PGA may be hydrolyzed and acts as a carbon or nitrogen source by production strains of γ-PGA if lacking external nutrient supply [19]. Repressing γ-PGA hydrolase by gene engineering could reduce loss of production and change molecular weight of γ-PGA [39,78]. Some research paid attention to the expression of γ-PGA hydrolases manipulated by gene knockout. Compared to wild strains, mutant deleted ggt in B. subtilis NAFM5 had higher molecular weight, but the yield of γ-PGA might be comparable [86]. In B. subtilis and B. amyloliquefaciens, deleting cwlO gene increased production even with higher molecular weight of γ-PGA relative to the wildtype strain [77], and deleting pgdS gene just increased molecular weight slightly [69]. However, the combined deletion of pgdS, cwlO, or ggt in the above strains resulted in a 93% increase in production compared to the wildtype [65,80]. Besides, deleting luxS gene in B. amyloliquefaciens NK-E10 resulted in a slightly improved production ability of γ-PGA [82].
In addition, enhancement of carbon flux conversion to γ-PGA biosynthesis, elimination of by-products or undesired precursor, and inhibition of other drain energy pathways can promote γ-PGA production and purity [42]. Carbon metabolism may be impacted positively by extra ATP, then improved by γ-PGA production [87]. Polysaccharides, major by-products and contaminates of γ-PGA, consume large amounts of the carbon sources and metabolic energy [19]. Mutants knocked out the synthetic genes of exopolysaccharides and lipopolysaccharide to obtain more and purer γ-PGA [82,88]. Compared with no removal of lipopolysaccharides in medium, the purity of γ-PGA in B. amyloliquefaciens NK-E5 increased to 95% [82]. Similarly, mutants removing the synthetic genes of both lactate and acetate, toxic compounds adversely impact cell growth, produced the same effects as the above study. However, approaches of eliminating glutamate-consuming pathways did not increase the yield of γ-PGA because the gene manipulations resulted in a metabolic imbalance with other related pathways in cells [19,82]. Just when γ-PGA is biosynthesized, energy-consuming substances synthesis also take places in cells, such as antibiotics. Repression of these energy-consuming substances synthetic pathways by gene knockout may increase production relative to the native strains but may not always be helpful [89].
Recombinant expression in homologous or heterologous hosts is another effective strategy for increasing γ-PGA production (Table 3). Genetic recombination is divided up into homologous expression and heterologous expression. Genetic manipulation recombined by E. coli and C. glutamicum is heterologous expression. Genetic manipulation recombined by Bacillus species is homologous expression. The xylose-induced plasmid pWH1520 with pgsBCA operon can be introduced into B. subtilis MA41 with the disruption of native pgsBCA gene, and then, γ-PGA synthetase can be successfully expressed [54]. Plasmid harboring the energy-saving NADPH-GDH pathway in B. amyloliquefaciens NK-1 could increase the yield of γ-PGA by 9%. Genetic expression patterns, including constitutive and inducible expressions, are the key factors affecting γ-PGA yield [78,90]. Inducible expression can accumulate γ-PGA in a short time. Besides, the operon pgsBCA of B. licheniformis NK-03 and racE of B. amyloliquefaciens LL3 were simultaneously cloned into an induced plasmid and co-expressed in E. coli JM109 for microbial synthesis of γ-PGA. Constitutive promoter (PHCE) from the D-amino acid aminotransferase (D-AAT) gene of Geobacillus toebii was recombined and expressed efficiently in E. coli/pCOpgs for γ-PGA biosynthesis. The final yield of γ-PGA is 3.7 g/L in the optimized medium [91].
Among others, C. glutamicum may be considered as a suitable host for recombinant production of γ-PGA, which is a native L-glutamate producer. When glutamate was deficient, heterologous expression of PgsBCA operon still resulted in the yield up to 0.7 g/L for C. glutamicum [19].

5.2. Culture Medium Optimization

Optimized medium composition can effectively promote cell growth and thus accumulate large amounts of precursors for γ-PGA synthesis (Table 4) [71]. More importantly, engineered strains that metabolize cheap substrates are an effective strategy for cost reduction of γ-PGA production. Carbon sources as major substrates result in directly or indirectly influence in γ-PGA production. Glucose, sucrose, glutamic acid, citric acid, and glycerol are the common carbon sources with different metabolic pathways and influences for γ-PGA production [95]. Glucose and citric acid are the most utilized carbon sources by Bacillus species [96]. Glucose metabolism is related to protein metabolism, such as several stress-response proteins [97]. Carbon catabolite may affect γ-PGA synthetic enzyme activity. Glucose may suppress the transcription of degQ operon during γ-PGA synthesis. Both citric acid and glutamic acid could serve as promoters for γ-PGA production [98]. Though fructose, maltose, and xylose show a positive effect on cell growth, almost no change was shown in γ-PGA production. Adding metabolic precursors of γ-PGA production as carbon sources may obtain higher yields, possibly because of enzymes activity enhanced of metabolic pathways [42]. Culture medium with L-glutamine and α-ketoglutaric acid can effectively increase the yield of γ-PGA up to 20% for B. subtilis BL53 [99]. The molecular weight also increased to 570 kDa [100]. Citric acid reached maximum yield to 28.3 g/L when compared with other organic acids, such as succinic acid, malic acid, and oxalic acid [47]. Organic acids involved in metabolic pathways often impact γ-PGA biosynthesis by adjusting enzymes activities. Most glutamic acid-independent strains prefer glucose and glycerol. Glycerol never inhibits related enzymes for γ-PGA production [97]. Compared with glucose, cells in glycerol can be more permeable, which may help the produce and release of γ-PGA [101,102]. Other biomass materials or by-products also can efficiently convert into high-value γ-PGA. Agro-industrial wastes were exploited as raw materials, such as cane molasses, corncobs, rapeseed meal, soybean residue, monosodium glutamate for processing wastewater, rice straw, and crude glycerol and its hydrolysate [74,96,103]. Besides, some carbonaceous substances might replace common carbon sources, such as algae, chicken manure, animal feathers, and dairy products [104,105,106,107]. The yield of γ-PGA reached 65 g/L in an optimized culture medium containing glutamic acid extracts, soybean residue, and chicken manure [108].
The nitrogen-containing compounds also can affect γ-PGA production. NH4+, an inorganic nitrogen source, easily form monomers of γ-PGA with α-ketoglutarate by transamination reaction [118]. The synthetic pathway of γ-PGA needs free amino groups. Glutamate is the most common nitrogen source for B. subtilis. Most L-glutamic acid is metabolized as a nitrogen source, and only a small amount is used to synthesize γ-PGA [119]. In addition, others nitrogenous substances are also suitable alternatives of nitrogen sources. With fishmeal waste as the nitrogen source in optimized culture medium, the yield of γ-PGA is increased up to 25 g/L [24].
Inorganic salts can also affect γ-PGA production. Metal ions serve as cofactors in utilization of carbon sources. Mn2+ can significantly enhance the glutamate racemase activity of γ-PGA, and then, the yield of γ-PGA can increase from 9.25 to 28.42 g/L [20,98]. The molecular weight and yield of γ-PGA can be regulated by sodium salts of culture medium for halotolerant strains as well as in other extremophilic bacteria [46]. Besides, γ-PGA derivatives with phosphorylation, esterification, or sulfonation can be directly produced by adding agents into the growth media, such as phosphoric acid. γ-PGA derivatives have greater potential than native γ-PGA and have been unprecedentedly used in various fields [7].
Exogenous additives of the culture medium are useful for γ-PGA production. Additives influence cellular membrane permeability and then regulate substrate availability. Tween 80 and DMSO can increase utilization of carbon sources [118].
It has been proven that suitable culture conditions, such as temperature, oxygen content, pH, and the inoculation amount. can effectively improve the yield of γ-PGA. Most Bacillus species are mesophilic and aerobic strains, tend to thrive well on optimum temperature range, and effect γ-PGA synthesis [48]. Some thermoacidophiles still maintain thermal stability in high temperature [120,121]. In addition, some strategies for maintaining oxygen supply, such as adding oxygen-carrying agent or one cultured in a fermenter, can enhance the γ-PGA production [99]. Another important factor is pH for γ-PGA biosynthesis because reactions involving glutamate are pH sensitive [11]. pH-controlling strategy further increases the utilization of glutamate and hence improves the γ-PGA production. The amount of inoculation is determined by the type of bacteria, culture composition, and culture conditions [122]. Inappropriate concentration of inoculum causes excessive nutrient competition or inhibition of the growth of viable cells and thus would reduce γ-PGA biosynthesis.
The optimization and integration of various conditions is the focus of current research, which take less fermentation time, has lower production costs, and has more stable productivity and yield. A series of methods, such as batch, fed-batch, continuous culture, and symbiotic fermentation (mixed strains or substrates are co-cultured), were tested at small scale to optimize the productivity of γ-PGA [8,54,123,124]. Clearly, these combined methods are highly effective. The yield of γ-PGA was increased to 2.19 g/L in a fermenter by fed-batch of glucose supply [72].

6. Purification of γ-PGA

γ-PGA is synthesized by Bacillus species outside the cells, which makes its purification relatively simple. The yield and molecular weight of γ-PGA, determined by biosynthesis, are the key factors affecting its purification [118]. Moreover, it has been confirmed that the purity of γ-PGA relates to culture mediums and conditions [112,125]. The higher the yield or/and the molecular weight of γ-PGA, the higher the viscosity and the more difficult the purification is [11]. Besides, part of γ-PGA is adhered to cell wall peptidoglycan rather than being secreted into the medium, which apparently makes purification of γ-PGA more difficult [1,14]. Precipitation and filtration are commonly used to purify γ-PGA. The first step is removal of the cells of fermented broth before purification, which can simplify the purification process. Diluting the medium or adjusting the pH value to about 3 can reduce the viscosity and help to remove the producing strains [126]. Then, the broth is removed by elementary precipitation and filtration. The normal method is organic solvents precipitation [42]. However, some of water-soluble organic solvents will degrade γ-PGA into oligomers or monomers. Ethanol remains the preferred precipitator of γ-PGA precipitation from a cell-free broth [126]. However, it results in the co-precipitation of other macromolecular substances, such as proteins and polysaccharides [11,26]. Obviously, further effective purification may be required for separation, such as filtration and buffer exchange, according to the physicochemical properties’ differences in γ-PGA and other constituents [116]. Diafiltration and ultrafiltration are commonly used and cost-effective filtration technologies [126]. Continuous fermentation integrated with a membrane-based filtration technology has been applied to industry [127]. However, membrane-based filtration technique is usually exceedingly slow. Besides, metal ions also can be used to precipitate and isolate γ-PGA, and Cu2+ has been proven to be an effective precipitator at 500-mM concentration with recovery of 95% [42]. Still, most of the metal ions are toxic, and it is also possible to contaminate the γ-PGA and environment [11,128]. Next, crude γ-PGA can be extracted by centrifugation and redissolved in distilled water for dialysis. The final step is lyophilization for obtaining purified γ-PGA [100].

7. Application of γ-PGA

Due to the various biological characteristics, such as water-solubility, thickness, moist quality, edible nature, non-toxicity, non-immunogenic nature, and antioxidant and antimicrobial activities., γ-PGA has been applied extensively in the food field, medical field, in environmental engineering field, and other industrial fields [4,5,6,7,8,9,10,11,12,13].

7.1. Food Industry

In the food industry, as an edible, non-toxic, strong water-retaining antioxidant and antimicrobial biopolymer, γ-PGA can replace the existing as a kind of emerging dietary supplement in functional food. As a texture enhancer and stabilizer, γ-PGA can improve the quality of starch, wheat gluten (WG), and their containing products [31]. For fermented protein products, γ-PGA can stabilize the viability of microorganisms, texture, aroma, and flavor of foods during the production, transportation, and sale of products, such as yoghurts [129]. It has been shown that γ-PGA can relieve bitter taste as a bitterness-masking agent in substances of amino acids, peptides, quinine, caffeine, minerals, etc., and as a thickener added to fruit juice beverages [82]. Based on the anti-freezing activity, γ-PGA is also a cryoprotectant to preserve the viability of probiotics during freeze-drying [130]. γ-PGA, with a molecular weight of lower than 20,000 Da, has higher anti-freezing activity than glucose and is more effectively protective to Lactobacillus paracasei than sucrose, trehalose, and sorbitol [29,81]. Furthermore, γ-PGA can increase the Ca2+ bioavailability and effectively prompt intestinal absorption [131]. Besides, the study of its antioxidant mechanism has demonstrated that γ-PGA can protect the gastrointestinal tract against oxidative damage [40].

7.2. Agricultural Field

γ-PGA shows great promise in agroecosystem due to its biological properties of water solubility, strong water retention, high biodegradability, high cation exchange capacity, and metal-chelating ability. For instance, γ-PGA is increasingly garnering interest in soil dynamics and plant growth as an environment-friendly fertilizer synergist [104]. γ-PGA may improve uptake of water-soluble nutrients and increase both root biomass and activity [28]. γ-PGA can be exogenously applied to protect seedlings from the adverse effects [132]. Moreover, based on biocontrol capacity and antimicrobial activities, γ-PGA has been used in abiotic or biotic stress resistances against hostile conditions [133]. As a biological chelating agent, γ-PGA is extremely beneficial to alleviate toxic accumulation caused by heavy metals of crops and farmland pollution [134].

7.3. Bio-Medical Field

γ-PGA has exhibited great potential in pharmaceutical manufacturing, which involves tissue engineering, delivery systems, and preventive, immunological, and therapeutic effects [11,135]. γ-PGA, a natural, non-toxic, and non-immunogenic biomacromolecule, can significantly decrease the toxic effects and improve the efficiency of drugs when combined with other matters [11]. Especially, biodegradable γ-PGA has a bright potential in drug-delivery platforms and as suitable carriers for gene therapy [81]. Molecular weight of γ-PGA was the decisive factor for considering the drug-delivery properties, including controlling or delaying the release [11]. γ-PGA has been used as new biological adhesives, formed by chemical cross-linking with other substances [40]. Besides, γ-PGA can induce a higher immunogenicity of vaccine antigens as a good adjuvant [136]. Furthermore, γ-PGA has showed excellent protection of Caco-2 cell and probiotics from oxidative damage [137].

7.4. Other Fields

γ-PGA is a natural, high-performance, hydrophilic humectant and dispersant in cosmetics because of its hydroscopicity, surface adhesion, antioxidant effects, and cytoprotection [41]. γ-PGA is expected to be the most useful flocculating agent due to its biodegradability and non-toxicity toward humans and the environment, which is superior to other conventional agents [19,138]. γ-PGA will be utilized not only in biomass waste and wastewater treatments in the industry but also in domestic water processing [81]. Besides, numerous chemically modified derivatives of γ-PGA have been developed so far as biopolymers utilized as novel biomaterials [12]; e.g., esterified γ-PGA can be used as an excellent thermoplastic for its capability to form biodegradable fibers and film [44,81].

8. Conclusions

γ-PGA is a promising biomacromolecule constituent of D- and/or L-glutamic acids, which has been applied worldwide in various industrial sectors, such as food processing, agricultural production, emerging medicine, environmental engineering, the cosmetic industry, and other related fields. Currently, microbial synthesis is a commonly used method to produce γ-PGA, and Bacillus species are natural, dominant, and safely produced. The molecular composition, molecular weight, synthesis productivity, and yield of γ-PGA directly affect its biological properties, which finally determined its industrial applications. At present, with the gradual increase in industrial demand, γ-PGA is of great interest for potential bio-based materials even if may impact existing industries and conventional commercialized polymers. Therefore, though microbial synthesis is the extremely economical and effective way to produce γ-PGA, cost of production and yield are the major constraints of industrial applications. The improvement of γ-PGA biosynthesis, including potential microbial producers screening, genetic manipulation, culture medium, and culture condition optimization of the producing strains, can increase yield and reduce production costs. The subsequent research will certainly be focused on novel theoretical studies and advanced technologies of γ-PGA microbial synthesis in the future.

Author Contributions

Conceptualization, F.W. and S.L.; writing—original draft preparation, D.L.; writing—review and editing, supervision, D.L. and Y.G.; investigation, L.H. and Z.T.; validation, Y.G. and S.L.; funding acquisition, F.W. and S.L.; project administration, F.W., B.F. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by China Agriculture Research System (CARS-04), Central Public-interest Scientific Institution Basal Research Fund (S2021JBKY-02), and Central Public-interest Scientific Institution Basal Research Incremental Fund (Y2020PT10).

Data Availability Statement

Not applicable.

Acknowledgments

The authors express gratitude for the funds and to the Knowledge Innovation Program Funding of Institute of Food Science and Technology, Chinese Academy of Agricultural Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shih, I.L.; Van, Y.T. The production of poly-(gamma-glutamic acid) from microorganisms and its various applications. Bioresour. Technol. 2001, 79, 207–225. [Google Scholar] [CrossRef]
  2. Ashiuchi, M.; Misono, H. Biochemistry and molecular genetics of poly-γ-glutamate synthesis. Appl. Microbiol. Biotechnol. 2002, 59, 9–14. [Google Scholar] [CrossRef] [PubMed]
  3. Sirisansaneeyakul, S.; Cao, M.F.; Kongklom, N.; Chuensangjun, C.; Shi, Z.P.; Chisti, Y. Microbial production of poly-γ-glutamic acid. World J. Microbiol. Biotechnol. 2017, 33, 173. [Google Scholar] [CrossRef]
  4. Richard, A.; Margaritis, A. Poly (glutamic acid) for biomedical applications. Crit. Rev. Biotechnol. 2001, 21, 219–232. [Google Scholar] [CrossRef]
  5. Sung, M.H.; Park, C.; Kim, C.J.; Poo, H.; Soda, K.; Ashiuchi, M. Natural and edible biopolymer poly-γ-glutamic acid: Synthesis, production, and applications. Chem. Rec. 2005, 5, 352–366. [Google Scholar] [CrossRef] [PubMed]
  6. Akagi, T.; Matsusaki, M.; Akashi, M. Pharmaceutical and Medical Applications of Poly-Gamma-Glutamic Acid; Microbiology Monographs, (Volume: Amino-Acid Homopolymers Occurring in Nature: 119–153); Hamano, Y., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 15. [Google Scholar] [CrossRef]
  7. Ashiuchi, M. Occurrence and Biosynthetic Mechanism of Poly-Gamma-Glutamic Acid; Microbiology Monographs, (Volume: Amino-Acid Homopolymers Occurring in Nature: 77–93); Hamano, Y., Ed.; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar] [CrossRef]
  8. Ashiuchi, M. Biochemical engineering of PGA. Microb. Biotechnol. 2013, 6, 664–674. [Google Scholar] [CrossRef] [PubMed]
  9. Bajaj, I.; Singhal, R. Poly (glutamic acid)—An emerging biopolymer of commercial interest. Bioresour. Technol. 2011, 102, 5551–5561. [Google Scholar] [CrossRef]
  10. Buescher, J.M.; Margaritis, A. Microbial biosynthesis of polyglutamic acid biopolymer and applications in the biopharmaceutical, biomedical and food industries. Crit. Rev. Biotechnol. 2007, 27, 1–19. [Google Scholar] [CrossRef]
  11. Luo, Z.; Guo, Y.; Liu, J.; Qiu, H.; Zhao, M.; Zou, W.; Li, S. Microbial synthesis of poly-γ-glutamic acid: Current progress, challenges, and future perspectives. Biotechnol. Biofuels 2016, 9, 134. [Google Scholar] [CrossRef] [Green Version]
  12. Ogunleye, A.; Bhat, A.; Irorere, V.U.; Hill, D.; Williams, C.; Radecka, I. Poly-γ-glutamic acid: Production, properties and applications. Microbiology 2014, 161, 1–17. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, Q.; Wei, X.; Chen, S. Production and Application of Poly-γ-Glutamic Acid. Biotechnology of Microbial Enzymes, (Chapter 2: Production. Isolation and Purification of Industrial Products). 2017. Available online: https://www.sciencedirect.com/book/9780128037256/biotechnology-of-microbial-enzymes#book-description (accessed on 1 February 2022).
  14. Candela, T.; Moya, M.; Haustant, M.; Fouet, A. Fusobacterium nucleatum, the first gram-negative bacterium demonstrated to produce polyglutamate. Can. J. Microbiol. 2009, 55, 627–632. [Google Scholar] [CrossRef] [PubMed]
  15. Hezayen, F.F.; Rehm, B.H.; Tindall, B.J.; Steinbuchel, A. Transfer of Natrialba asiatica B1T to Natrialba taiwanensis sp. nov. and description of Natrialba aegyptiaca sp. nov., a novel extremely halophilic, aerobic, non-pigmented member of the Archaea from Egypt that produces extracellular poly (glutamic acid). Int. J. Syst. Evol. Microbiol. 2001, 51, 1133–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Tamang, J.P.; Shin, D.H.; Jung, S.J.; Chae, S.W. Functional properties of microorganisms in fermented foods. Front. Microbiol. 2016, 7, 578. [Google Scholar] [CrossRef] [Green Version]
  17. Qiu, Y.; Zhu, Y.; Sha, Y.; Lei, P.; Xu, H. Development of a robust bacillus amyloliquefaciens cell factory for efficient poly (γ-glutamic acid) production from jerusalem artichoke. ACS Sustain. Chem. Eng. 2020, 8, 9763–9774. [Google Scholar] [CrossRef]
  18. Bajaj, I.B.; Singhal, R.S. Enhanced Production of Poly (γ-glutamic acid) from Bacillus licheniformis NCIM 2324 by Using Metabolic Precursors. Appl. Biochem. Biotechnol. 2009, 159, 133–141. [Google Scholar] [CrossRef] [PubMed]
  19. CaO, M.; Feng, J.; Sirisansaneeyakul, S.; Song, C.; Chisti, Y. Genetic and metabolic engineering for microbial production of poly-γ-glutamic acid. Biotechnol. Adv. 2018, 36, 1424–1433. [Google Scholar] [CrossRef] [PubMed]
  20. Wu, Q.; Hong, X.; Lin, X.; Ouyang, P. Biosynthesis of poly (γ-glutamic acid) in Bacillus subtilis nx-2: Regulation of stereochemical composition of poly (γ-glutamic acid). Process Biochem. 2006, 41, 1650–1655. [Google Scholar] [CrossRef]
  21. Ashiuchi, M.; Shimanouchi, K.; Nakamura, H.; Kamei, T.; Soda, K.; Park, C.; Sung, M.H.; Misono, H. Enzymatic synthesis of high-molecular-mass poly-γ-glutamate and regulation of its stereochemistry. Appl. Environ. Microbiol. 2004, 70, 4249–4255. [Google Scholar] [CrossRef] [Green Version]
  22. Ho, G.H.; Ho, T.I.; Hsieh, K.H.; Su, Y.C.; Lin, P.Y.; Yang, J.; Yang, K.H.; Yang, S.H. γ-polyglutamic acid produced by Bacillus subtilis (natto): Structural characteristics, chemical properties and biological functionalities. J. Chin. Chem. Soc. 2006, 53, 1363–1384. [Google Scholar] [CrossRef]
  23. Bhat, A.R.; Irorere, V.U.; Bartlett, T.; Hill, D.; Kedia, G.; Morris, M.R.; Charalampopoulos, D.; Radecka, I. Bacillus subtilis natto: A non-toxic source of poly-γ-glutamic acid that could be used as a cryoprotectant for probiotic bacteria. AMB Express 2013, 3, 36–44. [Google Scholar] [CrossRef] [Green Version]
  24. Birrer, G.A.; Cromwick, A.M.; Gross, R.A. γ-poly (glutamic acid) formation by Bacillus licheniformis 9945a: Physiological and biochemical studies. Int. J. Biol. Macromol. 1994, 16, 265–275. [Google Scholar] [CrossRef]
  25. Chunhachart, O.; Itoh, T.; Sukchotiratana, M.; Tanimoto, H.; Tahara, Y. Characterization of γ-glutamyl hydrolase produced by Bacillus sp. isolated from thai thua-nao. Biosci. Biotechnol. Biochem. 2006, 70, 2779–2782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Park, C.; Choi, J.C.; Choi, Y.H.; Nakamura, H.; Shimanouchi, K.; Horiuchi, T.; Misono, H.; Sewaki, T.; Soda, K.; Ashiuchi, M.; et al. Synthesis of super-high-molecular-weight poly-γ-glutamic acid by Bacillus subtilis subsp. chungkookjang. J. Mol. Catal. B Enzym. 2005, 35, 128–133. [Google Scholar] [CrossRef]
  27. Peng, Y.; Bo, J.; Tao, Z.; Mu, W.; Ming, M.; Hua, Y. High-level production of poly (γ-glutamic acid) by a newly isolated glutamate-independent strain, Bacillus methylotrophicus. Process Biochem. 2015, 50, 329–335. [Google Scholar] [CrossRef]
  28. Yin, A.; Jia, Y.; Qiu, T.; Gao, M.; Cheng, S.; Wang, X.; Sun, Y. Poly-γ-glutamic acid improves the drought resistance of maize seedlings by adjusting the soil moisture and microbial community structure. Appl. Soil Ecol. 2018, 129, 128–135. [Google Scholar] [CrossRef]
  29. Jang, W.J.; Choi, S.Y.; Lee, J.M.; Lee, G.H.; Hasan, M.T.; Kong, I.S. Viability of Lactobacillus plantarum encapsulated with poly-γ-glutamic acid produced by Bacillus sp. SJ-10 during freeze-drying and in an in vitro gastrointestinal model. LWT Food Sci. Technol. 2019, 112, 108222. [Google Scholar] [CrossRef]
  30. Lee, C.Y.; Kuo, M.I. Effect of γ-polyglutamate on the rheological properties and microstructure of tofu. Food Hydrocoll. 2011, 25, 1034–1040. [Google Scholar] [CrossRef]
  31. Xie, X.; Wu, X.; Shen, Y.; Song, M.; Xu, C.; Zhang, B.; Aziz, U.; Xu, X. Effect of poly-γ-glutamic acid on hydration and structure of wheat gluten. J. Food Sci. 2020, 85, 3214–3219. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Xiao, C.; Ding, J.; Li, M.; Chen, X.; Tang, Z.; Zhuang, X.; Chen, X. A comparative study of linear, Y-shaped and linear-dendritic methoxy poly (ethylene glycol)-block-polyamidoamine-block-poly (γ-glutamic acid) block copolymers for doxorubicin delivery in vitro and in vivo. Acta Biomater. 2016, 40, 243–253. [Google Scholar] [CrossRef]
  33. Liang, H.F.; Chen, C.T.; Chen, S.C.; Kulkarni, A.R.; Chiu, Y.L.; Chen, M.C.; Sung, H.W. Paclitaxel-loaded poly (gamma glutamic acid)-poly(lactide) nanoparticles as a targeted drug delivery system for the treatment of liver cancer. Biomaterials 2006, 27, 2051–2059. [Google Scholar] [CrossRef]
  34. Manocha, B.; Margaritis, A. Production and characterization of gamma-polyglutamic acid nanoparticles for controlled anticancer drug release. Crit. Rev. Biotechnol. 2008, 28, 83–99. [Google Scholar] [CrossRef] [PubMed]
  35. Manocha, B.; Margaritis, A. Controlled Release of Doxorubicin from Doxorubicin/gamma-Polyglutamic Acid Ionic Complex. J. Nanomater. 2010, 2010, 780171. [Google Scholar] [CrossRef] [Green Version]
  36. Li, M.; Song, W.; Tang, Z.; Lv, S.; Lin, L.; Sun, H.; Li, Q.; Yang, Y.; Hong, H.; Chen, X. Nanoscaled Poly (γ-glutamic acid)/Doxorubicin-Amphiphile Complex as pH-responsive Drug Delivery System for Effective Treatment of Nonsmall Cell Lung Cancer. ACS Appl. Mater. Interfaces 2013, 5, 1781–1792. [Google Scholar] [CrossRef]
  37. Zhang, L.; Chang, J.; Zhao, Y.; Xu, H.; Wang, T.; Li, Q.; Xing, L.; Huang, J.; Wang, Y.; Liang, Q. Fabrication of a triptolide-loaded and poly-γ-glutamic acid-based amphiphilic nanoparticle for the treatment of rheumatoid arthritis. Int. J. Nanomed. 2018, 13, 2051–2064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Yang, Z.H.; Dong, C.D.; Chen, C.W.; Shea, Y.T.; Kao, C.M. Using poly-glutamic acid as soil-washing agent to remediate heavy metal-contaminated soils. Environ. Sci. Pollut. Res. 2018, 25, 5231–5242. [Google Scholar] [CrossRef]
  39. Mitsunaga, H.; Meissner, L.; Bu¨chs, J.; Fukusaki, E. Branched chain amino acids maintain the molecular weight of poly (γ-glutamic acid) of Bacillus licheniformis ATCC 9945 during the fermentation. J. Biosci. Bioeng. 2016, 122, 400–405. [Google Scholar] [CrossRef]
  40. Lee, J.M.; Jang, W.J.; Park, S.H.; Kong, I.S. Antioxidant and gastrointestinal cytoprotective effect of edible polypeptide poly-γ-glutamic acid. Int. J. Biol. Macromol. 2020, 153, 616–624. [Google Scholar] [CrossRef]
  41. Ben-Zur, N.; Goldman, D.M. γ-Poly glutamic acid: A novel peptide for skin care. Cosmet Toilet. 2007, 122, 64–72. [Google Scholar]
  42. Nair, P.; Navale, G.R.; Dharne, M.S. Poly-gamma-glutamic acid biopolymer: A sleeping giant with diverse applications and unique opportunities for commercialization. Biomass Conv. Bioref. 2021, 1–19. [Google Scholar] [CrossRef]
  43. Cheng, C.; Asada, Y.; Aida, T. Production of γ-polyglutamic acid by Bacillus subtilis A35 under denitrifying conditions. Agric. Biol. Chem. 1989, 53, 2369–2375. [Google Scholar] [CrossRef]
  44. Kubota, H.; Matsunobu, T.; Uotani, K.; Takebe, H.; Satoh, A.; Tanaka, T.; Taniguchi, M. Production of poly (γ-glutamic acid) by Bacillus subtilis F-2-01. Biosci. Biotech. Biochem. 1993, 57, 1212–1213. [Google Scholar] [CrossRef] [PubMed]
  45. Ashiuchi, M.; Kamei, T.; Baek, D.H.; Shin, S.Y.; Sung, M.H.; Soda, K.; Yagi, T.; Mispno, H. Isolation of Bacillus subtilis (chungkookjang), a poly-γ-glutamate producer with high genetic competence. Appl. Microbiol. Biotechnol. 2001, 57, 764–769. [Google Scholar] [CrossRef] [PubMed]
  46. Wei, X.; Ji, Z.; Chen, S. Isolation of halotolerant Bacillus licheniformis WX-02 and regulatory effects of sodium chloride on yield and molecular sizes of poly-γ-glutamic acid. Appl. Microbiol. Biotechnol. 2010, 160, 1332–1340. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, H.; Zhu, J.; Zhu, X.; Cai, J.; Zhang, A.; Hong, Y.; Huang, J.; Huang, L.; Xu, Z. High-level exogenous glutamic acid independent production of poly-(γ-glutamic acid) with organic acid addition in a new isolated Bacillus subtilis C10. Bioresour. Technol. 2012, 116, 241–246. [Google Scholar] [CrossRef] [PubMed]
  48. Zeng, W.; Chen, G.; Wang, Q.; Zheng, S.; Shu, L.; Liang, Z. Metabolic studies of temperature control strategy on poly (gamma-glutamic acid) production in a thermophilic strain Bacillus subtilis GXA-28. Bioresour. Technol. 2014, 155, 104–110. [Google Scholar] [CrossRef] [PubMed]
  49. Yoshihito, I.; Takeshi, T.; Tetsuo, O.; Ito, Y.; Tanaka, T.; Ohmachi, T.; Asada, Y. Glutamic acid independent production of poly (γ-glutamic acid) by Bacillus subtilis tam-4. Biosci. Biotechnol. Biochem. 1996, 66, 1239–1242. [Google Scholar] [CrossRef] [Green Version]
  50. Shih, I.L.; Wu, P.J.; Shieh, C.J. Microbial production of a poly (gamma-glutamic acid) derivative by Bacillus subtilis. Process. Biochem. 2005, 40, 2827–2832. [Google Scholar] [CrossRef]
  51. Cao, M.; Geng, W.; Liu, L.; Song, C.; Xie, H.; Guo, W.; Jin, Y.; Wang, S. Glutamic acid independent production of poly-γ-glutamic acid by Bacillus amyloliquefaciens LL3 and cloning of pgsBCA genes. Bioresour. Technol. 2011, 102, 4251–4257. [Google Scholar] [CrossRef]
  52. Cachat, E.; Barker, M.; Read, T.D.; Priest, F.G. A Bacillus thuringiensis strain producing a polyglutamate capsule resembling that of Bacillus anthracis. FEMS Microbiol. Lett. 2008, 285, 220–226. [Google Scholar] [CrossRef] [Green Version]
  53. Chatterjee, P.M.; Datta, S.; Tiwari, D.P.; Raval, R.; Dubey, A.K. Selection of an effective indicator for rapid detection of microorganisms producing γ-polyglutamic acid and its biosynthesis under submerged fermentation conditions using Bacillus methylotrophicus. Appl. Biochem. Biotechnol. 2018, 185, 270–288. [Google Scholar] [CrossRef]
  54. Ashiuchi, M.; Shimanouchi, K.; Horiuchi, T.; Kame, T.; Misono, H. Genetically engineered poly-gamma-glutamate producer from Bacillus subtilis ISW1214. Biosci. Biotechnol. Biochem. 2006, 70, 1794–1797. [Google Scholar] [CrossRef] [Green Version]
  55. Cao, M.; Geng, W.; Zhang, W.; Sun, J.; Wang, S.; Feng, J.; Zheng, P.; Jiang, A.; Song, C. Engineering of recombinant Escherichia coli cells co-expressing poly-γ-glutamic acid (γ-PGA) synthetase and glutamate racemase for differential yielding of γ-PGA. Microb. Biotechnol. 2013, 6, 675–684. [Google Scholar] [CrossRef]
  56. Sha, Y.; Sun, T.; Qiu, Y.; Zhu, Y.; Zhan, Y.; Zhang, Y.; Xu, Z.; Li, S.; Feng, X.; Xu, H. Investigation of Glutamate Dependence Mechanism for Poly-γ-glutamic Acid Production in Bacillus subtilis on the Basis of Transcriptome Analysis. J. Agric. Food Chem. 2019, 67, 6263–6274. [Google Scholar] [CrossRef] [PubMed]
  57. Ashiuchi, M.; Soda, K.; Misono, H. Characterization of yrpC gene product of Bacillus subtilis IFO 3336 as glutamate racemase isozyme. Biosci. Biotechnol. Biochem. 1999, 63, 792–798. [Google Scholar] [CrossRef]
  58. Bormann, E.; Eikmanns, B.; Sahm, H. Molecular analysis of the Corynebacterium glutamicum gdh gene encoding glutamate dehydrogenase. Mol. Microbiol. 1992, 6, 317–326. [Google Scholar] [CrossRef] [PubMed]
  59. Urushibata, Y.; Tokuyama, S.; Tahara, Y. Difference in transcription levels of cap genes for y-polyglutamic acid production between Bacillus subtilis IFO 16449 and Marburg 168. J. Biosci. Bioeng. 2002, 93, 252–254. [Google Scholar] [CrossRef]
  60. Candela, T.; Fouet, A. Poly-gamma-glutamate in bacteria. Mol. Microbiol. 2006, 60, 1091–1098. [Google Scholar] [CrossRef]
  61. Yamashiro, D.; Yoshioka, M.; Ashiuchi, M. Bacillus subtilis pgsE (formerly ywtC) stimulates poly-γ-glutamate production in the presence of Zinc. Biotechnol. Bioeng. 2011, 108, 226–230. [Google Scholar] [CrossRef]
  62. Tran, L.S.; Nagai, T.; Itoh, Y. Divergent structure of the ComQXPA quorumsensing components: Molecular basis of strain-specific communication mechanism in Bacillus subtilis. Mol. Microbiol. 2000, 37, 1159–1171. [Google Scholar] [CrossRef]
  63. Do, T.H.; Suzuki, Y.; Abe, N.; Kaneko, J.; Itoh, Y.; Kimura, K. Mutations suppressing the loss of DegQ function in Bacillus subtilis (natto) poly-γ-glutamate synthesis. Appl. Environ. Microbiol. 2011, 77, 8249–8258. [Google Scholar] [CrossRef] [Green Version]
  64. Osera, C.; Amati, G.; Calvio, C.; Galizzi, A. SwrAA activates poly-gamma-glutamate synthesis in addition to swarming in Bacillus subtilis. Microbiology 2009, 155, 2282–2287. [Google Scholar] [CrossRef] [Green Version]
  65. Ohsawa, T.; Tsukahara, K.; Ogura, M. Bacillus subtilis response regulator DegU is a direct activator of pgsB transcription involved in gamma-poly-glutamic acid synthesis. Biosci. Biotechnol. Biochem. 2009, 73, 2096–2102. [Google Scholar] [CrossRef] [PubMed]
  66. Stanley, N.R.; Lazazzera, B.A. Defining the genetic differences between wild and domestic strains of Bacillus subtilis that affect poly-γ-DL-glutamic acid production and biofilm formation. Mol. Microbiol. 2005, 57, 1143–1158. [Google Scholar] [CrossRef] [PubMed]
  67. Ashiuchi, M.; Kamei, T.; Misono, H. Poly-γ-glutamate synthetase of Bacillus subtilis. J. Mol. Catal. B Enzym. 2003, 23, 101–106. [Google Scholar] [CrossRef]
  68. Feng, J.; Gao, W.; Gu, Y.; Zhang, W.; Cao, M.; Song, C.; Zhang, P.; Sun, M.; Yang, C.; Wang, S. Functions of poly-gamma-glutamic acid (γ-PGA)degradation genesin γ-PGA synthesis and cell morphology maintenance. Appl. Microbiol. Biotechnol. 2014, 98, 6397–6407. [Google Scholar] [CrossRef]
  69. Mitsui, N.; Murasawa, H.; Sekiguchi, J. Disruption of the cell wall lytic enzyme Cwlo affects the amount and molecular size of poly-γ-glutamic acid produced by Bacillus subtilis (natto). J. Gen. Appl. Microbiol. 2011, 57, 35–43. [Google Scholar] [CrossRef] [Green Version]
  70. Singh, V.; Haque, S.; Niwas, R.; Srivastava, A.; Pasupuleti, M.; Tripathi, C.K.M. Strategies for fermentation medium optimization: An in-depth review. Front. Microbiol. 2017, 7, 2087. [Google Scholar] [CrossRef]
  71. Gao, W.; Zhang, Z.; Feng, J.; Dang, Y.; Quan, Y.; Gu, Y.; Wang, S.; Song, C. Effects of MreB paralogs on poly-γ-glutamic acid synthesis and cell morphology in Bacillus amyloliquefaciens. FEMS Microbiol. Lett. 2016, 363, fnw187. [Google Scholar] [CrossRef] [Green Version]
  72. Huang, J.; Du, Y.; Xu, G.; Zhang, H.; Zhu, F.; Huang, L.; Xu, Z. High yield and cost-effective production of poly (gammaglutamic acid) with Bacillus subtilis. Eng. Life Sci. 2011, 11, 291–297. [Google Scholar] [CrossRef]
  73. Xu, Z.; Feng, X.; Zhang, D.; Tang, B.; Lei, P.; Liang, J.; Xu, H. Enhanced poly (γ-glutamic acid) fermentation by Bacillus subtilis NX-2 immobilized in an aerobic plant fibrous-bed bioreactor. Bioresour. Technol. 2014, 155, 8–14. [Google Scholar] [CrossRef]
  74. Zhang, D.; Feng, X.; Zhou, Z.; Zhang, Y.; Xu, H. Economical production of poly (γ-glutamic acid) using untreated cane molasses and monosodium glutamate waste liquor by Bacillus subtilis NX-2. Bioresour. Technol. 2012, 114, 583–588. [Google Scholar] [CrossRef] [PubMed]
  75. Ju, W.T.; Song, Y.S.; Jung, W.J.; Park, R.D. Enhanced production of poly-γ glutamic acid by a newly-isolated Bacillus subtilis. Biotechnol. Lett. 2014, 36, 2319–2324. [Google Scholar] [CrossRef] [PubMed]
  76. Zhu, F.; Cai, J.; Zheng, Q.; Zhu, X.; Cen, P.; Xu, Z. A novel approach for poly-gamma-glutamic acid production using xylose and corncob fibres hydrolysate in Bacillus subtillis. HB-1. J. Chem. Technol. Biotechnol. 2014, 89, 616–622. [Google Scholar] [CrossRef]
  77. Zhao, C.; Zhang, Y.; Wei, X.; Hu, Z.; Zhu, F.; Xu, L.; Luo, M. Production of ultra-high molecular weight poly-γ-glutamic acid with Bacillus licheniformis P-104 and characterization of its flocculation properties. Appl. Biochem. Biotechnol. 2013, 170, 562–572. [Google Scholar] [CrossRef] [PubMed]
  78. Scoffone, V.; Dondi, D.; Biino, G.; Borghese, G.; Pasini, D.; Galizzi, A.; Calvio, C. Knockout of pgds and ggt genes improves y-PGA yield in B. subtilis. Biotechnol. Bioeng. 2013, 110, 2006–2012. [Google Scholar] [CrossRef] [PubMed]
  79. Jiang, F.; Qi, G.; Ji, Z.; Zhang, S.; Liu, J.; Ma, X.; Chen, S. Expression of glr gene encoding glutamate racemase in Bacillus licheniformis WX-02 and its regulatory effects on synthesis of poly-gamma-glutamic acid. Biotechnol. Lett. 2011, 33, 1837–1840. [Google Scholar] [CrossRef] [PubMed]
  80. Tian, G.; Fu, J.; Wei, X.; Ji, Z.; Ma, X.; Qi, G.; Chen, S. Enhanced expression of pgdS gene for high production of poly-γ-glutamic acid with lower molecular weight in Bacillus licheniformis WX-02. J. Chem. Technol. Biotechnol. 2014, 89, 1825–1832. [Google Scholar] [CrossRef]
  81. Zhan, Y.; Sheng, B.; Wang, H.; Shi, J.; Cai, D.; Yi, L.; Yang, S.; Wen, Z.; Ma, X.; Chen, S. Rewiring glycerol metabolism for enhanced production of poly-γ-glutamic acid in Bacillus licheniformis. Biotechnol. Biofuels 2018, 11, 306. [Google Scholar] [CrossRef] [Green Version]
  82. Feng, J.; Gu, Y.; Quan, Y.; Cao, M.; Gao, W.; Zhang, W.; Wang, S.; Yang, C.; Song, C. Improved poly-γ-glutamic acid production in Bacillus amyloliquefaciens bymodular pathway engineering. Metab. Eng. 2015, 32, 106–115. [Google Scholar] [CrossRef]
  83. Zhang, W.; He, Y.; Gao, W.; Feng, J.; Cao, M.; Yang, C.; Song, C.; Wang, S. Deletion of genes involved in glutamate metabolism to improve poly-gamma-glutamic acid production in Bacillus amyloliquefaciens LL3. J. Ind. Microbiol. Biotechnol. 2015, 42, 297–305. [Google Scholar] [CrossRef]
  84. Gao, W.; He, Y.; Zhang, F.; Zhao, F.; Huang, C.; Zhang, Y.; Zhao, Q.; Wang, S.; Yang, C. Metabolic engineering of Bacillus amyloliquefaciens LL3 for enhanced poly-γ-glutamic acid synthesis. Microb. Biotechnol. 2019, 12, 932–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Yeh, C.M.; Wang, J.P.; Lo, S.C.; Chan, W.C.; Lin, M.L.Y. Chromosomal integration of a synthetic expression control sequence achieves poly-y-glutamate production in a Bacillus subtilis strain. Biotechnol. Prog. 2010, 26, 1001–1007. [Google Scholar] [CrossRef] [PubMed]
  86. Kimura, K.; Tran, L.S.; Uchida, L.; Itoh, Y. Characterization of Bacillus subtilis γ-glutamyltransferase and its involvement in the degradation of capsule poly-γ-glutamate. Microbiology 2004, 150, 4115–4123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Peng, Y.; Zhang, T.; Mu, W.; Miao, M.; Jiang, B. Intracellular synthesis of glutamic acid in Bacillus methylotrophicus SK19.001, a glutamate-independent poly (γ-glutamic acid)-producing strain. J. Sci. Food Agric. 2016, 96, 66–72. [Google Scholar] [CrossRef] [PubMed]
  88. Tian, T.; Kang, J.W.; Kang, A.; Lee, T.S. Redirecting metabolic flux via combinatorial multiplex crispri-mediated repression for isopentenol production in Escherichia coli. ACS Synth Biol. 2019, 8, 391–402. [Google Scholar] [CrossRef] [PubMed]
  89. Gao, W.; Liu, F.; Zhang, W.; Quan, Y.; Dang, Y.; Feng, J.; Gu, Y.; Wang, S.; Song, C.J.; Yang, C. Mutations in genes encoding antibiotic substances increase the synthesis of poly-γ-glutamic acid in Bacillus amyloliquefaciens LI.3. Microbiologyopen 2017, 6, e00398. [Google Scholar] [CrossRef] [Green Version]
  90. Chen, S.; He, P.; Wan, N.; Cai, D.; Hu, S.; Chen, Y.; Li, S. 13C-Metabolic flux analysis reveals the metabolic flux redistribution for enhanced production of poly-γ-glutamic acid in dlt over-expressed Bacillus Licheniformis. Front. Microbiol. 2018, 10, 105. [Google Scholar] [CrossRef]
  91. Jiang, H.; Shang, L.; Yoon, S.H.; Lee, S.Y.; Yu, Z. Optimal production of poly-gamma-glutamic acid by metabolically engineered Escherichia coli. Biotechnol. Lett. 2006, 28, 1241–1246. [Google Scholar] [CrossRef]
  92. Yao, W.; Meng, G.; Zhang, W.; Chen, X.; Yin, R. China National Intellectual Property Administration. Vol CN103146630 (A), 2013. Available online: https://english.cnipa.gov.cn/col/col1584/index.html (accessed on 24 January 2022).
  93. Cao, M.; Song, C.; Jin, Y.; Liu, L.; Liu, J.; Xie, H.; Guo, W.; Wang, S. Synthesis of poly (γ-glutamic acid) and heterologous expression of pgsBCA genes. J. Mol. Catal. B Enzym. 2010, 67, 111–116. [Google Scholar] [CrossRef]
  94. Lin, B.; Li, Z.; Zhang, H.; Wu, J.; Luo, M. Cloning and expression of the γ-polyglutamic acid synthetase gene pgsBCA in Bacillus subtilis WB600. BioMed Res. Int. 2016, 2016, 3073949. [Google Scholar] [CrossRef] [Green Version]
  95. Moraes, L.P.; Alegre, R.M.; Brito, P.N. Optimisation of poly (γ-glutamic acid) production by Bacillus velezensis NRRL B-23189 in liquid fermentation with molasses as the carbon source without addition of glutamic acid. Int. Rev. Chem. Eng. 2012, 4, 618–623. [Google Scholar]
  96. Tork, S.E.; Aly, M.M.; Alakilli, S.Y.; Al-Seeni, M.N. Purification and characterization of gamma poly glutamic acid from newly Bacillus licheniformis NRC20. Int. J. Biol. Macromol. 2015, 74, 382–391. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, D.; Hwang, J.S.; Kim, D.H.; Lee, S.; Joe, M.H. A newly isolated bacillus siamensis sb1001 for mass production of poly-γ-glutamic acid. Process. Biochem. 2019, 92, 164–173. [Google Scholar] [CrossRef]
  98. Kongklom, N.; Luo, H.; Shi, Z.; Pechyen, C.; Chisti, Y.; Sirisansaneeyakul, S. Production of poly-gamma-glutamic acid by glutamic acid-independent Bacillus licheniformis TISTR 1010 using different feeding strategies. Biochem. Eng. J. 2015, 100, 67–75. [Google Scholar] [CrossRef]
  99. Da Silva, S.B.; Cantarelli, V.V.; Ayub, M.A.Z. Production and optimization of poly-γ-glutamic acid by Bacillus subtilis BL53 isolated from the Amazonian environment. Bioprocess. Biosyst. Eng. 2014, 37, 469–479. [Google Scholar] [CrossRef]
  100. Lim, S.M.; Kim, J.I.; Shim, J.Y.; Imm, B.Y.; Sung, M.H.; Imm, J.Y. Effect of poly-γ-glutamic acids (PGA) on oil uptake and sensory quality in doughnuts. Food Sci. Biotechnol. 2012, 21, 247–252. [Google Scholar] [CrossRef]
  101. Du, G.; Yang, G.; Qu, Y.; Chen, J.; Lun, S. Effects of glycerol on the production of poly (γ-glutamic acid) by Bacillus licheniformis. Process. Biochem. 2005, 40, 2143–2147. [Google Scholar] [CrossRef]
  102. Shih, I.L.; Van, Y.T.; Chang, Y.N. Application of statistical experimental methods to optimize production of poly (γ-glutamic acid) by Bacillus licheniformis CCRC 12826. Enzyme Microb. Technol. 2002, 31, 213–220. [Google Scholar] [CrossRef]
  103. Zhang, C.; Wu, D.J.; Jia, J.; Yang, H.Q. Fishmeal wastewater as a low-cost nitrogen source for γ-polyglutamic acid production using Bacillus subtilis. Waste Biomass. Valorization 2019, 10, 789–795. [Google Scholar] [CrossRef]
  104. Chen, X.; Chen, S.; Sun, M.; Yu, Z. High yield of poly-y-glutamic acid from Bacillus subtilis by solid-state fermentation using swine manure as the basis of a solid substrate. Bioresour. Technol. 2005, 96, 1872–1879. [Google Scholar] [CrossRef]
  105. Kim, J.H.; Lee, J.M.; Jang, W.J.; Park, H.D.; Kim, Y.O.; Kim, C.H.; Kong, I.S. Efficient production of poly γ-d-glutamic acid from the bloom-forming green macroalgae, Ulva sp., by Bacillus sp. SJ-10. Biotechnol. Bioeng. 2019, 116, 1594–1603. [Google Scholar] [CrossRef]
  106. Altun, M. Bioproduction of γ-Poly (glutamic acid) using feather hydrolysate as a fermentation substrate. Trak. Univ. J. Nat. Sci. 2019, 20, 27–34. [Google Scholar] [CrossRef]
  107. Scheel, R.A.; Fusi, A.D.; Min, B.C.; Thomas, C.M.; Ramarao, B.V.; Nomura, C.T. Increased production of the value-added biopolymers poly (γ-3-hydroxyalkanoate) and poly (γ-glutamic acid) from hydrolyzed paper recycling waste fines. Front. Bioeng. Biotechnol. 2019, 7, 1–8. [Google Scholar] [CrossRef] [Green Version]
  108. Zhang, C.; Wu, D.; Ren, H. Economical production of agricultural γ-polyglutamic acid using industrial wastes by Bacillus subtilis. Biochem. Eng. J. 2019, 146, 117–123. [Google Scholar] [CrossRef]
  109. Zhang, D.; Xu, Z.; Xu, H.; Feng, X.; Li, S.; Cai, H.; Wei, Y.; Ouyang, P. Improvement of poly (gamma-glutamic acid) biosynthesis and quantitative metabolic flux analysis of a two-stage strategy for agitation speed control in the culture of Bacillus subtilis NX-2. Biotechnol. Bioproc. Eng. 2011, 16, 1144–1151. [Google Scholar] [CrossRef]
  110. Zhang, D.; Feng, X.; Li, S.; Chen, F.; Xu, H. Effects of oxygen vectors on the synthesis and molecular weight of poly (gamma-glutamic acid) and the metabolic characterization of Bacillus subtilis NX-2. Process. Biochem. 2012, 47, 2103–2109. [Google Scholar] [CrossRef]
  111. Jeong, J.H.; Kim, J.N.; Wee, Y.J.; Ryu, H.W. The statistically optimized production of poly (γ-glutamic acid) by batch fermentation of a newly isolated Bacillus subtilis RKY3. Bioresour. Technol. 2010, 101, 4533–4539. [Google Scholar] [CrossRef]
  112. Goto, A.; Kunioka, M. Biosynthesis and hydrolysis of poly (gamma-glutamic acid) from Bacillus subtilis IF03335. Biosci. Biotechnol. Biochem. 1992, 56, 1031–1035. [Google Scholar] [CrossRef] [Green Version]
  113. Silva, G.D.; Mack, M.; Contiero, J. Glycerol: A promising and abundant carbon source for industrial microbiology. Biotechnol. Adv. 2009, 27, 30–39. [Google Scholar] [CrossRef]
  114. Zeng, W.; Liang, Z.; Li, Z.; Bian, Y.; Li, Z.; Tang, Z.; Chen, G. Regulation of poly-γ-glutamic acid production in Bacillus subtilis GXA-28 by potassium. J. Taiwan Inst. Chem. Eng. 2016, 61, 83–89. [Google Scholar] [CrossRef] [Green Version]
  115. Cromwick, A.M.; Birrer, G.A.; Gross, R.A. Effects of pH and aeration on γ-poly (glutamic acid) formation by Bacillus licheniformis in controlled batch fermentor cultures. Biotechnol. Bioeng. 1996, 50, 222–227. [Google Scholar] [CrossRef]
  116. Yoon, S.H.; Hwan Do, J.; Yup Lee, S.; Nam Chang, H. Production of poly-γ-glutamic acid by fed-batch culture of Bacillus licheniformis. Biotechnol. Lett. 2000, 22, 585–588. [Google Scholar] [CrossRef]
  117. Wang, J.; Yuan, H.; Wei, X.; Chen, J.; Chen, S. Enhancement of poly-γ-glutamic acid production by alkaline pH stress treatment in Bacillus licheniformis WX-02. J. Chem. Technol. Biotechnol. 2015, 91, 2399–2403. [Google Scholar] [CrossRef]
  118. Wu, Q.; Xu, H.; Ying, H.J.; Ouyang, P.K. Kinetic analysis and pH-shift control strategy for poly (gamma-glutamic acid) production with Bacillus subtilis CGMCC 0833. Biochem. Eng. J. 2010, 50, 24–28. [Google Scholar] [CrossRef]
  119. Feng, J.; Gu, Y.; Quan, Y.; Gao, W.; Dang, Y.; Cao, M.; Lu, X.; Wang, Y.; Song, C.; Wang, S. Construction of energy-conserving sucrose utilization pathways for improvingpoly-y-glutamic acid production in Bacillus amyloliquefaciens. Microb. Cell Factoriesi. 2017, 16, 98. [Google Scholar] [CrossRef] [Green Version]
  120. Zeng, W.; Chen, G.; Guo, Y.; Zhang, B.; Dong, M.; Wu, Y.; Wang, J.; Che, Z.; Liang, Z. Production of poly-γ-glutamic acid by a thermotolerant glutamate-independent strain and comparative analysis of the glutamate dependent difference. AMB Express 2017, 7, 213. [Google Scholar] [CrossRef] [Green Version]
  121. Zeng, W.; Chen, G.; Wu, Y.; Dong, M.; Zhang, B.; Liang, Z. Nonsterilized fermentative production of poly-γ-glutamic acid from cassava starch and corn steep powder by a thermophilic Bacillus subtilis. J. Chem. Technol. Biotechnol. 2018, 93, 2917–2924. [Google Scholar] [CrossRef]
  122. Meissner, L.; Kauffmann, K.; Wengeler, T.; Mitsunaga, H.; Fukusaki, E.; Büchs, J. Influence of nitrogen source and pH value on undesired poly (γ-glutamic acid) formation of a protease producing Bacillus licheniformis strain. J. Ind. Microbiol. Biotechnol. 2015, 42, 1203–1215. [Google Scholar] [CrossRef]
  123. Hirasawa, T.; Shimizu, H. Recent advances in amino acid production by microbial cells. Curr. Opin. Biotechnol. 2016, 42, 133–146. [Google Scholar] [CrossRef]
  124. Jose, A.A.; Anusree, G.; Pandey, A.; Binod, P. Production optimization of poly-γ-glutamic acid by Bacillus amyloliquefaciens under solid-state fermentation using soy hull as substrate. Indian J. Biotechnol. 2018, 17, 44–49. [Google Scholar]
  125. Bajaj, I.B.; Singhal, R.S. Flocculation properties of poly (γ-glutamic acid) produced from Bacillus subtilis isolate. Food Bioprocess. Technol. 2011, 4, 745–752. [Google Scholar] [CrossRef]
  126. Do, J.H.; Chang, H.N.; Lee, S.Y. Efficient recovery of gamma-poly (glutamic acid) from highly viscous culture broth. Biotechnol. Bioeng. 2001, 76, 219–223. [Google Scholar] [CrossRef] [PubMed]
  127. Kumar, R.; Pal, P. Fermentative production of poly (γ-glutamic acid) from renewable carbon source and downstream purification through a continuous membrane-integrated hybrid process. Bioresour. Technol. 2015, 177, 141–148. [Google Scholar] [CrossRef]
  128. McLean, R.J.; Beauchemin, D.; Clapham, L.; Beveridge, T.J. Metal-binding characteristics of the gamma-glutamyl capsular polymer of Bacillus licheniformis ATCC 9945. Appl. Environ. Microbiol. 1990, 56, 3671–3677. [Google Scholar] [CrossRef] [Green Version]
  129. Yu, H.; Liu, H.; Wang, L.; Zhang, Y.; Tian, H.; Ma, X. Effect of poly-γ-glutamic acid on the stability of set yoghurts. J. Food Sci. Technol. 2018, 55, 4634–4641. [Google Scholar] [CrossRef]
  130. Siaterlis, A.; Deepika, G.; Charalampopoulos, D. Effect of culture medium and cryoprotectants on the growth and survival of probiotic lactobacilli during freeze drying. Lett. Appl. Microbiol. 2009, 48, 295–301. [Google Scholar] [CrossRef]
  131. Tanimoto, H.; Fox, T.; Eagles, J.; Satoh, H.; Nozawa, H.; Okiyama, A.; Morinaga, Y.; Fairweather-Tait, S. Acute effect of poly-gamma-glutamic acid on calcium absorption in post-menopausal women. J. Am. Coll. Nutr. 2007, 26, 645–649. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, Q.; Chen, S.; Zhang, J.; Sun, M.; Liu, Z.; Yu, Z. Co-producing lipopeptides and poly-gamma-glutamic acid by solid-state fermentation of Bacillus subtilis using soybean and sweet potato residues and its biocontrol and fertilizer synergistic effects. Bioresour. Technol. 2008, 99, 3318–3323. [Google Scholar] [CrossRef]
  133. Lee, W.; Kim, M.; Lee, S.H.; Jung, H.G.; Oh, J.W. Prophylactic efficacy of orally administered Bacillus poly-γ-glutamic acid, a non-LPS TLR4 ligand, against norovirus infection in mice. Sci. Rep. 2018, 8, 8667. [Google Scholar] [CrossRef]
  134. Pang, X.; Lei, P.; Feng, X.; Xu, Z.; Xu, H.; Liu, K. Poly-γ-glutamic acid, a bio-chelator, alleviates the toxicity of Cd and Pb in the soil and promotes the establishment of healthy Cucumis sativus L. seedling. Environ. Sci. Pollut. Res. 2018, 25, 19975–19988. [Google Scholar] [CrossRef]
  135. Singer, J.W. Paclitaxel poliglumex (XYOTAX (TM), CT-2103): A macromolecular taxane. J. Control. Release 2005, 109, 120–126. [Google Scholar] [CrossRef] [PubMed]
  136. Nguyen, Q.T.; Kwak, C.; Lee, W.S.; Kim, J.; Jeong, J.; Sung, M.H.; Yang, J.; Poo, H. Poly-γ-Glutamic Acid Complexed with Alum Induces Cross-Protective Immunity of Pandemic H1N1 Vaccine. Front. Immunol. 2019, 10, 1604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Csikós, Z.; Fazekas, E.; Rózsa, D.; Borbély, J.; Kerekes, K. Crosslinked poly-γ-glutamic acid based nanosystem for drug delivery. J. Drug Deliv. Sci. Technol. 2018, 48, 478–489. [Google Scholar] [CrossRef]
  138. Shih, I.L.; Van, Y.T.; Yeh, L.C.; Lin, H.G.; Chang, Y.N. Production of a biopolymer flocculant from Bacillus licheniformis and its flocculation properties. Bioresour. Technol. 2001, 78, 267–272. [Google Scholar] [CrossRef]
Figure 1. The structural formula of γ-PGA and its constituent units. The polymer of γ-PGA (n: repeating units approach at least 10,000) (A) and the L-glutamic acid monomer (B) and D-glutamic acid monomer (C) of γ-PGA.
Figure 1. The structural formula of γ-PGA and its constituent units. The polymer of γ-PGA (n: repeating units approach at least 10,000) (A) and the L-glutamic acid monomer (B) and D-glutamic acid monomer (C) of γ-PGA.
Foods 11 00739 g001
Figure 2. The synthetic pathway of γ-PGA. Main biologically synthetic pathway of γ-PGA involves three steps: racemization, polymerization and transfer, and catabolism. Various substrates (e.g., glucose, citric acid, glycerol, biomass materials, and by-products) enter by the main anabolism pathway. For glutamate-independent strains, they can utilize exogenous nutrients to product γ-PGA through glycolysis, PPP, and TCA cycle. Instead, for glutamate-dependent strains, glutamic acid or glutamate are added to the medium directly to product γ-PGA. PPP, Pentose Phosphate Pathway; G3P, glyceraldehyde 3-phosphate; TCA cycle, Tricarboxylic Acid cycle; E1, pyruvic acid aminotransferase (L-Glutamic acid); E2, alanine racemase; E3, pyruvic aminotransferase (D-Glutamic acid); RacE, glutamate racemase; Pgs, γ-PGA synthetase (four gene subunits operons: pgsB, C, A and E); GDH, glutamate dehydrogenase; GS, glutamine synthetase; GOGAT, glutamate synthase or glutamine oxoglutarate aminotransferase. PgdS, γ-PGA hydrolase; Ggt, γ-glutamyl transferase; CwlO, cell wall lyases.
Figure 2. The synthetic pathway of γ-PGA. Main biologically synthetic pathway of γ-PGA involves three steps: racemization, polymerization and transfer, and catabolism. Various substrates (e.g., glucose, citric acid, glycerol, biomass materials, and by-products) enter by the main anabolism pathway. For glutamate-independent strains, they can utilize exogenous nutrients to product γ-PGA through glycolysis, PPP, and TCA cycle. Instead, for glutamate-dependent strains, glutamic acid or glutamate are added to the medium directly to product γ-PGA. PPP, Pentose Phosphate Pathway; G3P, glyceraldehyde 3-phosphate; TCA cycle, Tricarboxylic Acid cycle; E1, pyruvic acid aminotransferase (L-Glutamic acid); E2, alanine racemase; E3, pyruvic aminotransferase (D-Glutamic acid); RacE, glutamate racemase; Pgs, γ-PGA synthetase (four gene subunits operons: pgsB, C, A and E); GDH, glutamate dehydrogenase; GS, glutamine synthetase; GOGAT, glutamate synthase or glutamine oxoglutarate aminotransferase. PgdS, γ-PGA hydrolase; Ggt, γ-glutamyl transferase; CwlO, cell wall lyases.
Foods 11 00739 g002
Figure 3. The synthetase gene operons for microbial γ-PGA in Bacillus bacteria. Pgs are responsible for assembling L-glutamic acid and D-glutamic acid units into γ-PGA. Pgs is encoded by an active membrane enzyme complex (pgsB, C, A, and E) and the γ-PGA-release gene (PgsS).
Figure 3. The synthetase gene operons for microbial γ-PGA in Bacillus bacteria. Pgs are responsible for assembling L-glutamic acid and D-glutamic acid units into γ-PGA. Pgs is encoded by an active membrane enzyme complex (pgsB, C, A, and E) and the γ-PGA-release gene (PgsS).
Foods 11 00739 g003
Table 1. Sources of γ-PGA production trains.
Table 1. Sources of γ-PGA production trains.
StrainsSourceMain Medium ComponentsCultural ConditionsFinal Yield (g/L)Ref.
B. subtilis ZJU-7Isolated from fermented bean curdGlucose,
L-glutamic, yeast extract, NaCl, Ca2+, Mg2+, Mn2+
Bioreactor, pH 6.5, 37 °C101.1[72]
B. subtilis NX-2Isolated from soil samplesGlucose,
glutamic, (NH4)2SO4, yeast extract, K2HPO4, Mg2+, Mn2+
APFB (aerobic plant fibrous-bed bioreactor) immobilized cell fermentation, pH 7.0, 32 °C71.21[73]
Isolated from soil samplesCane molasses, monosodium glutamate liquid wasteBioreactor, pH 7.0, 32 °C52.1[74]
B. subtilis MJ80Isolated from soil samplesGlutamic acid,
starch, urea, citric acid, glycerol, NaCl, K2HPO4, Mg2+, Mn2+
Fermenter for immobilized cell fermentation, pH 7.0, 37 °C68.7[75]
B. subtilis HB-1Isolated from soil samplesGlutamate, xylose, corncob fibers hydrolysate,
yeast extract, NaCl
Bioreactor, pH 6.5, 37 °C28.15[76]
B. methyotrophicus SK 19.001Isolated from soil samplesGlucose,
yeast extract, K2HPO4, Mg2+, Mn2+
Flask, pH 7.2, 37 °C35.34[27]
B. licheniformis P-104Isolated from Chinese soybean pasteGlucose, glutamate, citric acid, (NH4)2SO4, K2HPO4, Mg2+, Mn2+Bioreactor, pH 7.0, 37 °C41.6[77]
Table 2. Genetic manipulations of γ-PGA.
Table 2. Genetic manipulations of γ-PGA.
StrainsEngineering MethodsMain Medium ComponentsFinal Yield (g/L)Ref.
B. subtilis ISW1214Carrying the plasmid of γ-PGA synthetic systemSucrose, xylose, NaCl, NaHPO4, KH2PO4, Mg2+9.0[54]
B. subtilis PB5249Deletion of genes (pgdS and ggt) Glucose, L-glutamic acid, citric acid, NH4Cl, K2HPO4, Mg2+, Mn2, Ca2+, Fe2+40[78]
B. licheniformis WX-02Expression of glr gene for encoding glutamate recemaseGlucose, L-glutamic acid, citric acid, NH4Cl, NaCl, K2HPO4, Mg2+, Mn2+, Ca2+, Zn2+14.38[79]
Enhanced expression of pgdS geneGlucose, glutamate, citric acid, NH4Cl, K2HPO4, Mg2+, Mn2, Ca2+, Zn2+20.16[80]
Substituted by the native glpFK promoter with the constitutive promoter (P43), ytzE promoter (PytzE), and bacABC operon promoter (PbacA) Sodium glutamate, citric acid, glycerol,17.65[81]
Over-expression of glpK, glpX, zwf, and tkt1 promoters Sodium citrate, glycerol, NaNO3, NH4Cl12.83[81]
B. amyloliquefaciensDeletion of gene (epsA-O, sac, lps, pta, pgdS, cwlO, luxS, and rocG), expression of synthetic small synthetic regulator RNAs (repressed the expression of rocG and glnA gene) Sucrose, (NH4)2SO4, K2HPO4, KH2PO4, Mg2+20.3[82]
B. amyloliquefaciens LL3Double knockout of gene (pgdS and cwlO) Sucrose, (NH4)2SO4, NaCl, K2HPO4, KH2PO47.12[68]
Gene knockout of rocR, rocG, gudB, and odhA5.68[83]
Gene knockout of fadR, lysC, aspB, pckA, proAB, rocG, and gudBTryptone, xylose, yeast extract, NaCl, ampicillin, chloramphenicol, or tetracycline4.84[84]
Enhancing NADPH level by inserting a strong promoter PC2up6.46[84]
Table 3. Genetic manipulations of γ-PGA.
Table 3. Genetic manipulations of γ-PGA.
StrainsGenetic Engineering Main Medium ComponentsFinal Yield (g/L)Ref.
C. glutamicum ATCC 13869Cloning and expressing pgsBCA of B. licheniformis TKPG011Glucose, (NH4)2SO4, soy protein hydrolysate, thiamine hydrochloride, KH2PO4, Mg2+, Mn2, Fe2+, Ca2+18[92]
C. glutamicum ATCC 13032Cloning and expressing pgsABC from B. licheniformis NK-03Glucose, tryptone, yeast extract0.7[93]
E. coli BL21Cloning and overexpressing γ-PGA biosynthesis genesGlucose, L-glutamic acid, yeast extract, (NH4)2SO43.7[91]
E. coli LRPExpressing pgsBCA and race from B. amyloliquefaciens LL3Glucose, yeast extract, NaCl,0.7[1]
E. coli JM 109Cloning pgsBCA and racE from both B. licheniformis NK-03 and B. amyloliquefaciens LL3 and co-expressionGlucose, L-glutamic acid, yeast extract, NaCl, (NH4)2SO4, K2HPO4, KH2PO4, Mg2+,0.65[55]
B. subtilis PB5249pgdSggt deletion mutantsglucose, L-glutamate, citric acid, NH4Cl, K2HPO4, MgSO4·7H2O, MnSO4·H2O, FeCl3·6H2O, CaCl2·2H2O 40[78]
B. subtilis WB600Overexpressing pWB980-pgsBCAGlucose, sodium glutamate, MgSO4, (NH4)2SO4, K2HPO41.74[94]
B. subtilis ISW1214Overexpressing pWH1520-PxylA-pgsBCASucrose, xylose, NaCl, MgSO4, KH2PO4, NaHPO4 9.0[54]
B. licheniformis WX-02Overexpressing pHY300PLK-P43-glrSucrose, (NH4)2SO4, MgSO4, KH2PO4, K2HPO414.38[79]
Overexpressing pHY300PLK-PpgdS-pgdSGlucose, sodium glutamate, sodium citrate, NH4Cl, MgSO4, ZnSO4, MnSO4, CaCl2, K2HPO420.16[80]
B. amyloliquefaciens LL3pgdScwlO deletion mutantsSucrose, (NH4)2SO4, MgSO4, KH2PO4, K2HPO47.12[68]
rocRrocGgudBodhA deletion mutantsSucrose, (NH4)2SO4, MgSO4, KH2PO4, K2HPO45.68[83]
B. amyloliquefacienscwlOepsA-Ovgb deletion mutantsSucrose, (NH4)2SO4, MgSO4, FeSO4·4H2O, CaCl2·2H2O, MnSO4·4H2O, ZnCl2, KH2PO4, K2HPO4,5.12[68]
Repressed both rocG and glnA genesSucrose, (NH4)2SO4, MgSO4, KH2PO4, K2HPO420.3[82]
Table 4. The culture methods of γ-PGA.
Table 4. The culture methods of γ-PGA.
StrainsCultural MethodsFinal Yield (g/L)Ref.
B. subtilis NX-2Using a two-stage strategy for agitation speed control40.5[109]
Adding hydrogen peroxide33.9[110]
B. subtilis CGMCC 0833Applying pH-shift control strategy27.7[111]
B. subtilis F-2-01Adding more carbon sources (L-glutamic acid and glycerol)45.5[43]
B. subtilis BL53Adding some precursors25.2[112]
B. subtilis C10Abundant supply of organic acid27.7[47]
B. subtilis GXA-28Addition of KCl25.62[113]
B. licheniformis TISTR 1010Different feeding strategies
(glucose, citric acid, NH4Cl, NaCl, Mg2+, Mn2, Ca2+, K2HPO4, Tween-80)
27.5[114]
B. licheniformis ATCC 9945ADifferent feeding strategies
(L-glutamic acid, citric acid, glycerol, NH4Cl)
23.0[98]
35.0[115]
57.5[116]
B. licheniformis WBL-3Optimized culture medium
(L-glutamic acid, citric acid, glycerol, NH4Cl)
22.8[102]
Optimized culture medium
(glutamic acid, citric acid, glycerol)
19.3[102]
B. licheniformis P-104Optimized culture medium
(glutamate, glucose, citric acid, (NH4)2SO4)
41.6[77]
B. licheniformis NCIM 2324Addition of metabolic precursors
(α-ketoglutaric acid)
35.75[18]
Optimized culture medium
(L-glutamic acid, citric acid, glycerol, (NH4)2SO4, K2HPO4, Mg2+, Mn2+)
26.12[18]
Optimization effecting factors at a time
(L-glutamic acid, sugarcane juice, citric acid, NH4Cl)
35.75[18]
B. licheniformis WX-02Adding pH stress treatment36.26[101]
B. velezensis NRRL B-23189Optimized culture medium
(molasses, citric acid, NH4Cl)
4.82[117]
B. methylotrophicus SK 19.001Optimized culture medium
(glucose, citric acid, NH4Cl)
35.3[27]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, D.; Hou, L.; Gao, Y.; Tian, Z.; Fan, B.; Wang, F.; Li, S. Recent Advances in Microbial Synthesis of Poly-γ-Glutamic Acid: A Review. Foods 2022, 11, 739. https://doi.org/10.3390/foods11050739

AMA Style

Li D, Hou L, Gao Y, Tian Z, Fan B, Wang F, Li S. Recent Advances in Microbial Synthesis of Poly-γ-Glutamic Acid: A Review. Foods. 2022; 11(5):739. https://doi.org/10.3390/foods11050739

Chicago/Turabian Style

Li, Danfeng, Lizhen Hou, Yaxin Gao, Zhiliang Tian, Bei Fan, Fengzhong Wang, and Shuying Li. 2022. "Recent Advances in Microbial Synthesis of Poly-γ-Glutamic Acid: A Review" Foods 11, no. 5: 739. https://doi.org/10.3390/foods11050739

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