Open access peer-reviewed chapter - ONLINE FIRST

Bacterial Alginate Biosynthesis and Metabolism

Written By

Rodrigo Vassoler Serrato

Submitted: 23 November 2022 Reviewed: 01 December 2022 Published: 23 December 2022

DOI: 10.5772/intechopen.109295

Alginate - Applications and Future Perspectives IntechOpen
Alginate - Applications and Future Perspectives Edited by Ihana Severo

From the Edited Volume

Alginate - Applications and Future Perspectives [Working Title]

Dr. Ihana Aguiar Severo, Dr. André Bellin Mariano and Dr. José Viriato Coelho Vargas

Chapter metrics overview

144 Chapter Downloads

View Full Metrics

Abstract

Alginate is a linear anionic heteropolysaccharide with a chemical structure consisting of 1,4-linked subunits of β-D-mannuronic acid (M) and its C-5 epimer α-L-guluronic acid (G). It is well known that the monomer composition and molecular weight of alginates affect their properties and influence their use in the food and pharmaceutical industries. Alginate is usually extracted from seaweed for commercial purposes, but can also be produced by bacteria as exopolysaccharide (EPS). Pseudomonas spp. and Azotobacter vinelandii are well-known alginate-producing microorganisms. Their biochemical machinery for alginate biosynthesis is influenced by changing culture conditions and manipulating genes/proteins, making it relatively easy to obtain customized EPS with different molecular weights, M/G compositions, and thus physicochemical properties. Although these two genera have very similar biosynthetic pathways and molecular mechanisms for alginate production, with most of the genes involved being virtually identical, their regulation has been shown to be somewhat different. In this chapter, we present the main steps of alginate biosynthesis in bacteria, including precursor synthesis, polymerization, periplasmic modifications, transport/secretion, and post-secretion modification.

Keywords

  • Exopolysaccharide
  • Pseudomonas spp.
  • Azotobacter vinelandii
  • alginic acid
  • mannuronic acid

1. Introduction

Alginate is a linear anionic biopolymer consisting of β-D-mannuronic acid units (M) and its C5-epimer α-L-guluronic acid (G) linked by 1 → 4 glycosidic bonds. It can be found as a homopolysaccharide (poly-M or poly-G) or as a heteropolysaccharide containing M/G in different ratios and sequences. Alginate has been known to be a major component of brown algae cell walls since the late nineteenth century when scientists were searching for utile products from kelp [1], and the gelatinous material isolated from these seaweeds was then named alginic acid [2]. The molecular backbone of alginate, however, was only described decades later, after the work of Fisher and Dörfel in 1955 [3], and the structural analyses of alginate have been an extensive field of research ever since due to the immense variability of M/G content and sequential organization found throughout the years in many macroalgae [4, 5]. The structural variability presented by this polymer and the ability to form gels with different physicochemical properties is nowadays reflected in its many biotechnological uses in the food sector, cosmetic industry, wastewater treatment, pharmaceutical, and biomedical applications, among others [6, 7, 8, 9].

Even though all commercial alginate produced today is extracted from seaweed, a handful group of specialized microorganisms can also produce this polymer as an exopolysaccharide [10]. Two genera of bacteria, Pseudomonas and Azotobacter, have been extensively studied over the past years on their ability to exudate alginate [11, 12, 13]. The production of alginate by Pseudomonas aeruginosa has been first reported by Linker and Jones in the mid-1960s [14, 15], and most of the knowledge on bacterial alginate production and biosynthesis emerged from the studies with this opportunistic human pathogen for patients afflicted by cystic fibrosis [16, 17], since the exopolymer represents major virulent factors during the lung infection process [18, 19]. Shortly after, alginate was described as an exopolysaccharide produced by Azotobacter vinelandii [20], a nitrogen-fixing soil bacterium found in association with plants. Over the years, other species belonging to these two genera have also been reported to produce alginate, such as P. putida and P. fluorescens [21], P. mendocina [22]; P. syringae [23] and A. chroococcum [24, 25].

The biosynthesis of bacterial alginates exhibits highly comparable mechanisms among the different producing species but differs from the seaweed polysaccharides in one particular characteristic. The microbial polymers are often O-acetylated, submitted to esterification at O-2 and/or O-3 on the D-mannuronate units [26, 27, 28], which affects their physicochemical properties in comparison to the alginate extracted from macroalgae [29, 30, 31, 32]. Furthermore, as well as different M/G compositions and sequences are found on alginates from seaweeds, the gel formation ability and viscosity of bacterial alginates are also affected by the arrangement of M-, G- and MG-blocks along their structures [7, 33]. Some mucoid strains of P. aeruginosa biosynthesize great amounts of alginate containing mainly M- and MG-blocks (Table 1), often O-acetyl-esterified, which leads to an increasing interaction of the polymer with water molecules, polymer extension, and water capacity, thus allowing the development of a thickly natured and highly structured biofilm matrix [34, 35, 36]. Azotobacter spp., on the other hand, produce alginates that are closely associated with the cells and able to capture divalent ions such as Ca2+ that form a hard and brittle gel due to the high content of L-guluronate residues (G-blocks) [37, 38], allowing the formation of desiccation-resistant cysts under adverse environmental conditions [39, 40].

Block typeDisaccharide repeating unitProducing organism
G-block→4)-α-L-GulA-(1 → 4)-α-L-GulA-(1→
G – G
Seaweed;
Azotobacter spp.
M-block*→4)-β-D-ManA-(1 → 4)-β-D-ManA-(1→
M – M
Seaweed;
Azotobacter spp.;
Pseudomonas spp.
GM-block*→4)-α-L-GulA-(1 → 4)-β-D-ManA-(1→
or
→4)-β-D-ManA-(1 → 4)-α-L-GulA-(1→
G – M or M – G
Seaweed;
Azotobacter spp.;
Pseudomonas spp.

Table 1.

Most commonly found structural disaccharides (blocks) of alginates produced by seaweeds, Pseudomonas spp. and Azotobacter spp.

ManA units are found O-acetylated on C-2 or C-3 only in bacterial alginates.


GulA (G): guluronic acid; ManA (M): mannuronic acid.

The production of tailor-made alginate in order to obtain polymers with different molecular weights, M/G ratio, and O-acetyl content to suit specific applications has been intensively investigated [41, 42, 43] as it poses as a potential commercial interest for the use-driven design of these polysaccharides in different industrial and medical applications [44, 45, 46]. Altering fermentation parameters and growth conditions in order to induce alginate production and or structural modifications during microbial cultivation has been extensively evaluated, especially for Azotobacter vinelandii [47, 48, 49, 50, 51, 52] due to its non-pathogenic characteristics, which makes this bacterium suitable for biotechnological processes. However, the genome sequencing of both P. aeruginosa [53] and A. vinelandii [54], together with the use of molecular tools for gene cloning, have expanded the possibilities for in vivo fine structural adjustments of alginate, and the isolation of alginate-modifying enzymes is allowing the on-demand chemical tailoring of these bacterial polymers [55, 56, 57, 58], including the suppression or degradation of alginate produced by Pseudomonas spp. for commercial and medical purposes [59, 60, 61, 62].

Understanding the molecular basis involved in bacterial alginate biosynthesis, transport, and metabolism, which are usually under strict regulatory control [63, 64, 65, 66], is paramount in order to develop new strains and mutants guided to the production of the idealized alginate structure in large-scale bioreactors or to be used therapeutically on clinical patients in opposition to alginate-overproducing pathogens [67, 68]. In this chapter, we are going to explore the main steps involved in the biosynthesis of microbial alginate, from the assembly of its precursor units to the polymerization, chemical and structural modifications, and transport through the cytoplasmic and outer membranes until the secretion of the mature polymer by the bacterial cells and the extracellular modifications that may follow. Gene and enzyme regulation will also be briefly addressed when relevant to each of the biosynthetic phases.

Advertisement

2. Overview of the genetics and regulation of bacterial alginate biosynthesis

The genetic organization of alginate-producing bacteria has been extensively analyzed and the regulation of the biosynthetic process is relatively well known, from the biosynthesis of precursor molecules, their polymerization into poly-D-mannuronic acid, chemical modifications/epimerization in the periplasmic space, and secretion to the extracellular environment. All of these steps have been previously identified and characterized in many reviews [42, 69, 70] and the biosynthesis of alginate has been shown to be extremely similar for both Pseudomonas aeruginosa and Azotobacter vinelandii, resulting from a complex regulatory network of proteins [41, 51, 59, 65]. In P. aeruginosa, a 12-gene operon is strictly controlled by algD transcription, the first gene of the alginate biosynthetic operon [53, 71], and two additional internal promoters upstream of algG and algI [72] (Figure 1). Three transcription promoter regions have been described for the algD operon, namely algDp1D promoter, algDp2 (AlgU-δE promoter), and algDp3 [65]. Genes algD and algA, located downstream on the operon, are involved in the biosynthesis of the precursor guanosine diphosphate D-mannuronic acid (GDP-ManA) together with algC, which is found elsewhere in the bacterial chromosome, has its own promoter regions [73], and encodes for a phosphomannomutase [74]. The gene algG encodes for a C5-mannuronate epimerase, which converts D-ManA units into L-GulA in the periplasm [75]. Further structural modification of the nascent alginate polymer is promoted by the products of genes algI, algJ, and algF, all of which are involved in the O-acetylation of ManA units [76, 77]. Periplasmic proteins encoded by genes algXand algK, also present as part of the operon, are supposedly involved in the guiding of alginate through the outer membrane porin (formed by algE) and protection of alginate against degradation of an alginate lyase encoded by algL [77, 78]. The polymerization process to form poly-D-ManA is promoted by two transmembrane proteins (localized on the cytoplasmic membrane) encoded by genes alg8 and alg44 [79, 80].

Figure 1.

Genetic organization of alginate biosynthesis/regulation related genes for Pseudomonas aeruginosa (blue) and Azotobacter vinelandii (green). Gene sizes are not to scale. Numbers on top of each gene indicate bp as available from https://www.ncbi.nlm.nih.gov/ for the genome of P. aeruginosa PAO1 (Ref. NC_002516.2) and A. vinelandii DJ (Ref. NC_012560.1). The direction of transcription and bp for alyB of A. vinelandii is still undetermined.

The alginate biosynthetic genes on Azotobacter vinelandii are also clustered and share high similarity with their previously described P. aeruginosa counterparts. However, despite their essentially identical functions, Pseudomonads algE and algJ genes have been designated, respectively, as algJ and algV in A. vinelandii (Figure 1). Also, A. vinelandii alginate gene cluster has been shown to possess two promoter regions upstream of algD, one of which is an AlgU (δE)-dependent promoter, as in P. aeruginosa, and another RpoS (δs)-dependent [64, 65, 81]. Additionally, other three internal promoters are found on the A. vinelandii operon upstream of alg8s-dependent), algG70-dependent), and algA (regulation unknown) [64, 65]. The levels of algA transcription exclusively from its upstream promoter, however, are not sufficient to sustain alginate production [82]. Only in P. aeruginosa, the transcription of the algA and algL genes has been demonstrated to be controlled by AlgU [51].

Exclusively in Azotobacter spp., seven genes encoding C5-ManA epimerases have been identified, namely, algE1-algE7 [58, 83, 84]. The translational products of these (AlgE1-AlgE7) are exported to the cell surface and released into the extracellular environment. AlgE1–7 catalyze the Ca2+-dependent epimerization of D-ManA units into L-GulA units in the alginate polymer after its secretion through the outer membrane by the AlgJ porin [51, 83]. These extracellular epimerases are structurally unrelated to AlgG, a Ca2+-independent C5-mannuronate epimerase found in the periplasm, and all seven algE1–7 genes have been reported to be regulated by the sigma factor RpoS [85]. AlgE7 has been reported to also have lyase activity, which leads to depolymerization of alginate via β-elimination at the 4-O-glycosidic bonds [86]. In addition, four other alginate lyase genes (alyA1–3 and alyB) different from algL, which is part of the biosynthetic operon, have been described for A. vinelandii [42].

AlgU, previously called AlgT, δE, or δ22 [87], is homologous to the alternative sigma factor of the stress response regulator RpoE from Escherichia coli [88] and plays a key role as the main regulator of bacterial alginate biosynthesis. AlgU ist is encoded together with the mucABCD operon, which contains four genes (mucA, mucB, mucC, and mucD). In both, A. vinelandii and Pseudomonas spp., the muc genes are located downstream of the algU gene, forming the algUmucABCD regulatory gene cluster [64, 65, 89]. Transcription of the biosynthetic algD operon is activated by the presence of AlgU, which in turn requires the activation of two other promoters located upstream of algC (algCp1) and an AlgU (δE)-dependent promoter (algDp2) [65, 90]. In A. vinelandii, however, AlgU does not participate in the regulation of algL and algA genes as it has been reported for P. aeruginosa [51, 91]. Together, algU and the mucABCD genes form the principal regulatory switch that controls the conversion between mucoid and non-mucoid phenotypes of P. aeruginosa [64, 65, 89], and in other Pseudomonads, such as P. fluorescens and P. syringae, the mucC gene is absent and the transcription mucD is not dependent on AlgU [92, 93].

The regulatory proteins encoded by the mucABCD operon are found either embedded in the inner cell membrane or soluble in the periplasmic space. MucA is a transmembrane protein that has anti-δE activity and acts as a negative regulator of AlgU and is required for maintaining cell viability in P. aeruginosa [94]. MucB has a hydrophobic cavity that interacts with MucA forming a stable complex (MucA-MucB) and serving as a fine-tune controlling mechanism that protects MucA from cleavage of its periplasmic domains by proteases [95, 96]. The role of MucC still remains to be fully determined, however, it has been hypothesized that this is an inner membrane protein that might act synergistically with MucA and MucB to negatively regulate AlgU in A. vinelandii, but not in P. aeruginosa [89, 97]. On the other hand, the role of MucD has been thoroughly determined as being a serine protease/chaperone-like protein with an important regulatory function that acts on the alginate-biosynthetic complex through AlgX [10], and also plays a central role in AlgU activation [89] functioning as a negative regulator localized in the periplasm (Figure 2).

Figure 2.

Schematic representation for the genetic regulation of bacterial alginate biosynthesis. Transcription regulation (full lines) or protein-protein modulation (dashed lines) are indicated in blue (positive regulation/modulation) or red (negative modulation).

Another set of genes and their respective translational products are part of a two-component signal transduction system and directly participate in the regulatory control network of alginate biosynthesis [89]. Four out of five chromosomal regulatory genes (algB, algP, algQ, algR, and algZ/amrZ) found in Pseudomonads and A. vinelandii transcribe in the opposite direction from the previously described algD- and algUmucABCD-operons, except for algB, which has the same transcriptional orientation of the biosynthetic cluster [82]. The promoter regions of algR and algB have significant homology to other promoters recognized by δE, and these genes are positively regulated by AlgU-(δE) [89]. Their translational products, AlgR and AlgB, act directly on different sites of the algD-operon promoter in P. aeruginosa [9899]. In contrast, AlgB and AlgR are not required for activation of algD transcription in A. vinelandii [82], and mutations in algB did not affect alginate production for this bacterium, as recently reported [100]. The kinase AlgQ transcribed from algQ undergoes autophosphorylation using ATP or GTP, and transfers a phosphate group to algR [82]. Since AlgR binds to specific sequences of the algD-promoter region, it up-regulates the transcription of the biosynthesis operon in P. aeruginosa, together with AlgU. Although algC transcription is independent of other alginate genes, it has been demonstrated that algC is up-regulated by the presence of functional algR [73]. A ribbon-helix-helix DNA binding protein, AlgZ (also called AmrZalginate and motility regulator) is the product of gene algZ and is a strictly δ22-dependent kinase that binds to sequences localized upstream of the algD promoter, activating algD transcription [82]. It has been reported that algZ mutants of A. vinelandii are unable to produce alginate, thus, it is proposed that AlgZ is required for the transcription of the biosynthetic algD operon [100], however, the molecular mechanism of the algD regulation by AlgZ is yet to be elucidated. AlgP was determined to be a histone-like protein that binds to the algD promoter and aids DNA looping to positively regulate this gene and activate translation [89].

Other regulatory systems are also present during alginate biosynthesis, including post-translational and post-transcriptional regulations [99]. A two-component system (GacS/GacA) acts as a global regulator affecting the transcription of algD and indirectly controlling alginate biosynthesis in A. vinelandii through a signaling cascade involving the post-translational system RsmA/RsmB [65]. It has been described that GacS controls the expression of algD from its three promoters and that gacS and gacA mutants of A. vinelandii have significantly lower levels of algD transcripts [5181]. Despite being a well-conserved regulatory system among Gram-negative bacteria, the GacS-RsmA pathway has not been reported to regulate alginate biosynthesis in P. aeruginosa [64, 99, 101]. It has been recently announced that the secondary messenger bis-(3′-5′) cyclic dimeric-GMP (c-di-GMP) directly regulates the transcription of the alginate operon in P. aeruginosa [59]. c-di-GMP also acts as a post-translational regulator binding to PilZ domains on proteins such as the one located at the C-terminal end of Alg44 [102, 103]. The binding of c-di-GMP to the Alg44-PilZ domain is essential to activate the alginate polymerase complex Alg8-Alg44 [104].

Although the genetic organization and regulation of alginate biosynthesis seem to be similar in Pseudomonas spp. and Azotobacter vinelandii, these two genera of bacteria have strikingly different ecologies and are present in very specific habitats, which, not surprisingly, reflect the different roles alginate plays in these microorganisms. As shown above, some marked differences between these bacteria regarding alginate biosynthesis-genes distribution and regulation are well reported in the scientific literature [51, 65, 82, 99]. However, there is still much to be unveiled on the genetics and enzymatic regulatory complex for these bacteria, especially for A. vinelandii, which is considered a generally recognized as safe (GRAS) microorganism, with much potential for the biotechnological industry. Therefore, a broad understanding of all the biochemical and regulatory aspects of alginate biosynthesis, combined with the testing of culture conditions and molecular biology tools is paramount for the development of optimal growth methods and strains able to produce this exopolysaccharide in vivo or with post-production in vitro modifications to achieve defined G/M composition, molecular weight and ideal tailor-made physicochemical properties [46, 105, 106, 107, 108].

Advertisement

3. Phases of bacterial alginate biosynthesis

3.1 GDP-ManA biosynthesis pathway

It is known that bacterial alginate is initially synthesized as a linear poly-D-mannuronic acid homopolysaccharide before undergoing chemical modifications and epimerization in the periplasmic space, which implies in mannuronic acid (ManA) being the main precursor molecule specifically involved in the biosynthesis of bacterial alginate. However, prior to its utilization to form the poly-D-ManA chain that initiates alginate biosynthesis, ManA must first be synthesized from fructose-6-phosphate (F6P), which is derived from the cell’s central carbon metabolism, and then activated with a high-energy bond to yield the sugar nucleotide GDP-ManA [109]. All steps leading to the conversion of F6P to GDP-ManA, the basic building block of alginate, occur in the cytosol and have been extensively analyzed and reported in the past, and a summary of this process is shown in Figure 3.

Figure 3.

Biosynthesis of guanosine diphosphate-mannuronic acid (GDP-ManA), the precursor molecule for bacterial alginate biosynthesis. AlgA: bifunctional phosphomannose isomerase/GDP-Mannose pyrophosphorylase; AlgC: phosphomannomutase; AlgD: GDP-mannose dehydrogenase; PMI: phosphomannose isomerase activity; GMP: guanosine-diphosphomannose pyrophosphorylase activity; F6P: D-fructose-6-phosphate; M6P: D-mannose-6-phosphate; M1P: D-mannose-1-phosphate; GDP-Man: guanosine diphosphate D-mannose; GDP-ManA: guanosine diphosphate D-mannuronic acid.

Three enzymes are directly involved in GDP-ManA biosynthesis, namely, AlgA (isomerase/pyrophosphorylase, AlgC (phosphomutase), and AlgD (dehydrogenase). GDP-ManA biosynthesis begins with the conversion of F6P to mannose-6-phosphate (M6P). This reaction is catalyzed by the bifunctional enzyme AlgA through its phosphomannose isomerase (PMI) activity [110]. Through the formation of M6P, F6P-derived carbon molecules are committed to the alginate biosynthetic pathway, and M6P is rapidly converted to mannose-1-phosphate (M1P) by the Mg2+-dependent phosphomannomutase AlgC [111], a crucial enzyme not only for the biosynthesis of alginate but also required for the production other exopolysaccharides [112], lipopolysaccharides, and rhamnolipids [113, 114]. AlgA is required once again in the next reaction step, however, it will now employ its guanosine-diphosphomannose pyrophosphorylase (GMP) activity to convert M1P to GDP-mannose, simultaneously hydrolyzing GTP and pyrophosphate (PPi). GDP-ManA is finally obtained by the irreversible oxidation of GDP-mannose by AlgD, an NAD+-dependent guanosine-diphospho-D-mannose dehydrogenase that catalyzes the limiting step and major control point of alginate biosynthesis [41, 115, 116].

3.2 Biosynthesis of the poly-D-ManA linear chain

Polymerization of GDP-ManA precursors to polymannuronate is carried out by two enzymes located at the inner membrane of bacteria, interfacing the cytosolic material with the other periplasmatic proteins in the alginate biosynthetic complex. Both enzymes, namely Alg8 and Alg44, are indispensable for alginate production [65, 69, 117]. Alg8 is believed to be the catalytic subunit of alginate polymerase since it shares homologies with class II glycosyltransferases [118], which are responsible for the transfer of sugar units from an activated monosaccharide donor to an acceptor molecule. Similarities in the polymerization mechanism of alginate with the biosynthesis of other polysaccharides have been suggested [119], and since the catalytic mechanism of class II glycosyltransferases requires an inversion of the anomeric configuration at the reaction center, in order to yield β-linked products such as the nascent alginate polymer (poly-β-D-ManA) the donor substrates must be and α-linked sugar nucleotide (i.e.: GDP-α-D-ManA) [69] (Figure 4). Membrane topology analysis of Alg8 has shown that this protein has at least four transmembrane helices, a short periplasmic loop, and an extended cytoplasmic loop containing the glycosyltransferase domain [118, 119, 120]. It has been reported that alg8 overexpression in both A. vinelandii and P. aeruginosa greatly increases alginate production in comparison to wild-type strains [80]. The molecular weight of alginate polymers has also been shown to be higher when alg8 expression increases in A. vinelandii [121].

Figure 4.

Proposed mechanism reaction for the polymer-level epimerization of β-D-ManA into α-L-GulA by the periplasmic C5-mannuronan epimerase AlgG.

The copolymerase Alg44 is the most controversial protein in the entire alginate biosynthesis complex. It is mostly accepted as being a transmembrane protein localized in the bacterial inner membrane and bears a c-di-GMP-binding PilZ domain at its cytosolic N-terminal portion [79, 103, 104]. However, because its complete structure has not yet been determined, and due to its homology with MexA, a periplasmic membrane fusion protein involved in the multi-drug efflux system [122], some reports have proposed that Alg44 lacks transmembrane domains and instead participates in the multi-protein scaffold formation of the alginate biosynthetic complex bridging Alg8 in the inner membrane to the porin (AlgE or AlgJ) in the outer membrane [79, 112, 117]. As previously stated, however, Alg44 is required for the polymerization of ManA together with Alg8, and its regulatory function via the second messenger c-di-GMP in the Alg8-Alg44 complex has been demonstrated [45123]. One particular c-di-GMP synthesizing protein, namely MucR, has been shown to specifically influence the alginate biosynthesis in both P. aeruginosa [124] and A. vinelandii [125], and mutations in the PilZ domain of Alg44 has resulted in the loss of alginate production [112].

3.3 Periplasmic chemical modifications of the polymannuronic acid

After polymerization by the Alg8-Alg44 complex, the nascent poly-D-ManA chain is directed across a multi-enzymatic scaffold structure situated in the periplasmic space, where ManA units may suffer epimerization or O-acetylation [10, 65, 80]. The enzymes involved in these processes are the best-analyzed proteins in the biosynthesis process of alginate [126, 127], and the understanding of their reaction mechanisms is an important factor to enable the production of tailored polymers [10, 128]. Although neither epimerization nor O-acetyl-esterification is essential for the bacterial production of alginate, these chemical modifications significantly alter the physicochemical properties of the polymer [31, 32, 38].

The O-acetylation of alginate is a process naturally occurring only in bacteria. During its transit through the periplasm, some ManA units of the polymannuronate chain are O-acetyl-esterified at positions C-2 and/or C-3 by the combined activity of four enzymes: AlgX, AlgF, AlgI, and AlgJ/AlgV [10, 65, 69, 77, 110]. The acetyl groups from the donor molecules (probably acetyl-CoA, but still undetermined), are transported from the cytosol to the periplasm by AlgI, a 7-helical transmembrane protein with homology to DltB (D-alanyltransferase) in Bacillus subtilis [76], to AlgJ (in P. aeruginosa) or AlgV (in A. vinelandii). AlgJ/AlgV is found on the periplasm, closely associated with AlgI, and also anchored in the cytoplasmic membrane presumably by its hydrophobic signal peptide [129]. It shows high homology with AlgX, another protein present in the alginate biosynthesis complex. The structure of AlgX has been shown to have an N-terminal hydrolase domain involved in the acetylation of alginate, and a C-terminal portion where a sugar-binding domain is possibly involved in the binding and orientation of the polymannuronate chain [78, 130]. AlgJ/AlgV and AlgX show about 70% analogy [76, 77], very similar topology [129], and are both required for the esterification of alginate. AlgF has no sequence homology with any other proteins involved in the O-acetylation [10], and algF mutants of A. vinelandii have been shown to produce non-acetylated alginate [131]. The order of transfer of the acetyl donor to each of the acetylation-related enzymes is uncertain, but the fact is that AlgJ/V performs is directly involved in the O-acetylation of ManA residues and that AlgF serves presumably an accessory role due to the lack of identifiable catalytic residues [129]. Moreover, it is clear that AlgX is also able to catalyze the direct O-acetylation of alginate, receiving the acetate or acetyl donor molecules either from AlgJ/V or AlgF [129]. AlgX has also been shown to interact with MucD in some sort of undetermined regulatory mechanism, which affects the formation of alginate [65, 69, 132].

The esterification of ManA units on hydroxyl groups at C-2 and/or C-3 prevents degradation of the growing alginate polymer by AlgL, an alginate lyase present in the periplasm which will not be discussed in detail herein. O-acetylation also hampers the epimerization of esterified ManA units by AlgG, which is a C5-mannuronan epimerase of the alginate biosynthesis complex and catalyzes the epimerization of β-D-mannuronate (M-blocks) to α-L-guluronate (G-blocks) at polymer level [92]. Non-acetylated ManA units may undergo epimerization by AlgG, which causes the monosaccharide unit to suffer a dramatic change in both its anomeric configuration (from β to α) and pyranose chair conformation (from 4C1 to 1C4) (Figure 4). Such considerable structural changes have very important effects on the physicochemical properties of the mature alginate depending on the proportion and sequence of M-blocks that go through epimerization [69, 112, 133]. AlgG adopts a right-handed parallel β-helix fold, and the catalytic mechanism proposed for this epimerase is based on the β-elimination (not hydrolysis) reaction of polysaccharide lyases [134], which involves the neutralization of the carboxylate group of the uronic acid, removal of the proton at the C-5 position and cleavage of the glycosidic bond with the formation of a double bond between C-4 and C-5 for the formation of a transient glycal intermediate. For the epimerization reaction, however, glycosidic linkage remains intact, and a proton is added to the opposite side of the C-5 to form the epimer [135]. Functional analysis of AlgG mutants suggests that His319 is the catalytic base and that Arg345 is responsible for neutralizing the carboxylic groups during the epimerase reaction. It has also been proposed that water is the likely catalytic acid [134].

Polymer-level epimerization of monosaccharide units is rarely found in nature, and thus far has only been described for the biosynthesis of alginate (both microbial and algal) and glycosaminoglycans (animal origin) [136]. Hence, AlgG is a very exclusive and extraordinary enzyme found in both P. aeruginosa and A. vinelandii. These enzymes show an optimum pH for activity between 6.0 and 7.5 and share about 60% sequence identity [134, 135]. However, differently from Pseudomonads, the A. vinelandii genome also encodes a family of seven extracellular Ca2+-dependent epimerases (AlgE1–AlgE7), which will be discussed later in this chapter.

3.4 Secretion of the mature alginate chain

The two remaining proteins that are part of the alginate biosynthesis complex are AlgK and AlgE/AlgJ (Pseudomonas/Azotobacter). AlgK is a lipoprotein of unclear function found anchored into the inner leaflet of the bacterial outer membrane [137]. It is hypothesized that AlgK may have a protective role, guiding the nascent alginate polymer through the biosynthetic multi-protein complex in the periplasmic space, and preventing it from being degraded by lyases, since the lack of AlgK has been shown to result in the secretion of free uronic acids and/or short-chain alginate products [78, 132], showing that AlgK an essential element for the biosynthesis of full-length alginate and the development of the mucoid phenotype in P. aeruginosa [138]. Additionally, AlgK is proposed to interact with both Alg44, stabilizing the Alg8-Alg44 polymerase complex, and the porin protein AlgE/AlgJ [112, 137]. There is some evidence to suggest that AlgK is involved in the localization of AlgE/AlgJ to the outer membrane [139]. Also, it has been shown that AlgK may form complexes with AlgX and MucD in the periplasmic space, possibly halting the regulatory role of MucD due to its sequestration by an AlgK/AlgX complex in the biosynthetic multi-protein scaffold [65]. The large number of interactions of AlgK with other proteins may be due to its structure containing several tetratricopeptide-like helical motifs which allow unspecific protein-protein interactions [139].

AlgE was first described in the early 1990s as an integral outer membrane protein strongly associated with the mucoid phenotype of P. aeruginosa [140, 141]. Its counterpart in A. vinelandii, AlgJ, shares a high degree of similarity and is thought to perform the same function as in Pseudomonads [142]. Although AlgE is classified as a member of the general diffusion porin family [143], biochemical and electrophysiological analyses suggest that this protein is specifically responsible for the passage of alginate into the extracellular environment through the outer membrane, rather than allowing the passage of small molecules and ions, and that it forms a strongly positively charged anion-selective pore that can be partially blocked by GDP-ManA [144, 145]. AlgE/AlgJ is a monomeric 18-stranded β-barrel protein homologous to the OprD family of substrate-specific porins. In addition to its alginate transport function, it has been shown to play an important role in the assembly of the multiprotein complex involved in alginate biosynthesis [145, 146].

Secretion usually indicates the end of the bacterial alginate biosynthetic pathway, especially in Pseudomonads, and a schematic representation of the entire biosynthetic/regulatory process can be found in Figure 5. After release into the environment, alginate may undergo further non-enzymatic structural modifications such as de-O-acetylation and/or depolymerization, albeit to a very modest extent, depending on the chemical conditions in the medium surrounding the cell. However, the alginate produced by Azotobacter vinelandii can undergo very significant enzymatic processing extracellularly due to the presence of C5-mannuronan epimerases and alginate lyases, which are also secreted by this bacterium.

Figure 5.

Schematic representation of bacterial alginate biosynthesis, regulation, periplasmic transportation/modification, secretion, and post-secretion modification. Protein structures are merely illustrative (https://pdb101.rcsb.org/) and the nomenclature used is that established for P. aeruginosa encoded genes. (*A. vinelandii counterparts or enzymes that are exclusively found in the Azotobacter genus are marked with an asterisk).

3.5 Post-secretion chemical modifications of alginate

3.5.1 Extracellular epimerases

As previously described, Azotobacter vinelandii encodes seven different C5-ManA epimerases (AlgE1–AlgE7) that are exported to the extracellular environment and are able to promote polymer-level epimerization of D-ManA units into L-GulA on alginate. Although the epimerization mechanism for these enzymes is very similar to the mechanism proposed for AlgG, extracellular epimerases have been found to be Ca2+-dependent, in contrast to the periplasmic counterpart. It has been considered that extracellular alginate epimerases are necessary for A. vinelandii cells for the creation of a diffusion barrier against O2 [84] as a protective mechanism for its oxygen-sensitive nitrogenase complex and that enzymes of the AlgE family play an important role in the encystment process induced by adverse environmental conditions [147]. Also, only alginate-producing cells may produce cysts that can germinate after storage [65]. Alginates produced by Pseudomonads show lower L-GulA content in comparison to the A. vinelandii polymers [38, 65, 70]. Furthermore, alginates isolated from Azotobacter spp. contain G-blocks (consecutive L-GulA residues), while in Pseudomonas spp. alginates have been found to contain only single L-GulA units on their backbone structure [70]. It is proposed that post-secretion epimerization of alginate by AlgE1–7 is the reason for the higher amounts of L-GulA found in the alginate produced by A. vinelandii [10, 65, 118]. As the G-block amounts increase, so does the affinity of this polymer for divalent ions, which are responsible for the formation of cross-links between alginate chains creating a more rigid gel surrounding A. vinelandii cells, being physicochemically much different from the lower viscosity alginate produced by bacteria of the genus Pseudomonas [10]. Even though consecutive GulA units have not been found in alginates produced by Pseudomonads to date, it has been reported that P. syringae pv glycinea encodes a C5-mannuronan epimerase designated PsmE, which was found to be able to introduce G-blocks on alginate in vitro [65, 148]. Differently from the A. vinelandii extracellular epimerases, PsmE is able to epimerize acetylated D-ManA units after removing O-acetyl groups, thus showing a bifunctional deacetylase/epimerase characteristic [148].

Isomerases of the AlgE family have been shown to be differentially expressed during the life cycle of A. vinelandii and to generate different patterns of G-blocks on alginate [147]. Five of the algE genes of A. vinelandii are found clustered in the genome, while algE5 and algE7 are found isolated elsewhere (see Figure 1), and each correspondingly encoded enzyme appears to have different specificities in terms of substrate preferences and non-random formation of G- and GM-blocks on the alginate backbone [149]. These extracellular alginate epimerases from A. vinelandii are about 70% identical but share less than 10% sequence identity with AlgG [134]. The characterization of AlgEs from A. vinelandii has been made at certain levels after cloning, heterogeneous expression, and purification, and three putative AlgE-like isomerases have also been described and characterized in A. chroococcum [58]. All of the AlgE isomerases are structurally organized into two types of alternating modules, each with different roles for these enzymes, namely: A-module and R-module (Table 2). It is known that the catalytic site is located in the A-modules, each comprising about 385 amino acids (with a primary amino acid sequence homology of ~85%). These are always followed by 1–4 R-modules (~155 amino acids each), which, in turn, are thought to modulate the epimerization rate by stabilizing the process with an extended alginate binding site [150, 151], and have also been shown to play a role in the epimerization pattern of the final alginate product [152]. A-modules have been calculated to bind 11 uronic acid units, while each of the R-modules can bind five units with different binding strengths, which also affects the processivity of a particular AlgE enzyme [151]. Four to seven copies of a nine amino acid long motif are located in the N-terminal region of the R-modules and are known to be Ca2+-binding sites [153]. Calcium ions are important not only for the structural integrity of the protein but also for neutralizing the charges of ManA during the epimerization reaction, in a similar manner Arg345 functions in AlgG [149, 150]. In addition, the last R-module of each extracellular AlgE usually contains an unstructured peptide involved in the secretion of the enzyme [151]. Despite the large sequence homology they share, different AlgE epimerases catalyze distinct residue sequences in the alginate product. It has been proposed that the different epimerization patterns depend on the concerted action of both the A- and R-modules, rather than only the catalytic activity of A-modules [150].

Epimerase (kDa)ModulesSubstrateProduct
AlgE1
(147.2)
A1R1R2R3A2R4Poly-M or Poly-MGPoly-G + Poly-MG
AlgE2*
(102.9)
A1R1R2R3R4Poly-M or Poly-MGPoly-G (short)
AlgE3
(195.1)
A1R1R2R3A2R4R5R6R7Poly-M or Poly-MGPoly-G + Poly-MG
AlgE4
(57.7)
A1R1Poly-MPoly-MG
AlgE5
(103.7)
A1R1R2R3R4Poly-M or Poly-MGPoly-G (short)
AlgE6
(90.2)
A1R1R2R3Poly-M or Poly-MGPoly-G (long)
AlgE7*
(90.4)
A1R1R2R3Poly-M or Poly-MGPoly-G + Poly-MG

Table 2.

Structural organization, substrate specificity, and products formed by extracellular C5-ManA epimerases of Azotobacter vinelandii.

Bifunctional extracellular epimerase/lyase enzymes. A-modules: Location of the catalytic site; R-modules: Ca2+-binding sites and substrate-binding/stabilization extended sites.


Although the extracellular C5-epimerases AlgE1–7 play an important role in the addition of L-GulA units to the alginate produced by A. vinelandii, these enzymes have also been cloned and isolated to be used for in vitro modification of macroalgal-derived alginate [128, 149]. AlgE1 and AlgE3 are the only epimerases that contain two catalytic A-modules, one of which introduces single G units that form alternating GM sequences, while the other can generate G-blocks with consecutive GulA units [65, 154]. AlgE2 and AlgE5 both consist of an A-module followed by four R-modules, and they primarily introduce short G-blocks on the alginate chain [149]. The same module organization is found in AlgE6, which, however, is able to introduce longer G-blocks on polymannuronic acid [147, 153]. AlgE4 consists of only one A-module and one R-module and is the smallest enzyme among the extracellular C5-epimerases. Although it inserts a high proportion of GulA units into the alginate chain, this enzyme is unable to attach successive G-blocks to the polymer [65, 149]. AlgE7 generally produces the same GM alternating products as AlgE1, AlgE3, and AlgE4. However, AlgE7 and AlgE2 have also been shown to have present lyase activity (breaks G-MM and G-GM bonds), which is thought to share the same active site with epimerase activity [86]. These bifunctional enzymes can degrade the alginate chain after the epimerization reaction, but at different rates. AlgE2 is thought to have low lyase activity and has been shown to perform 1 to 3 elimination reactions per thousand epimerized units [149, 155]. In contrast, AlgE7 has a much more pronounced lyase activity, with a predicted number of 3–4 glycosyl bond breaks per 26 conversions of D-ManA to L-GulA units [86].

Regarding the reaction mechanisms of Ca2+-dependent AlgE epimerases, two reaction modes have been proposed: (1) the preferred attack mode, in which the enzyme preferentially attacks M-units with G residues as the nearest neighbor and detaches from the substrate after each epimerization reaction [58, 149], and (2) the processive mode, in which there is a relative displacement of the enzyme and the alginate chain that allows the conversion of the nearest M-unit without dissociation of the enzyme-substrate complex [149, 153]. AlgE4 has been reported to have a processive mode of action by sliding along the substrate chain and epimerizing every other M residue [156], and this is thought to be due to the 180°-turn between the successive β-(1 → 4)-linked ManA units in the alginate polymer [153]. AlgE2 follows the mechanism of preferred attack mode and is related to the Ca2+ concentration and substrate type of the blocks found in alginate [149, 155]. The remaining AlgE epimerases appear to behave according to either a preferred attack mode or a combination of the two mechanisms [149].

Advertisement

4. Concluding remarks

Despite being in general a relatively well-known process, as shown throughout this chapter, several steps of the bacterial alginate biosynthesis need to be further explored, especially those involving regulatory genes and proteins, to determine their full role during the biosynthetic process and, more importantly, how this knowledge can be useful in the production of tailored alginate polymers for various industrial and commercial uses. In addition to the alginate-producing Pseudomonads and Azotobacter spp. that have been the focus herein, a vast number of organisms bear in their biochemical machinery enzymes capable of degrading alginate (alginate lyases), including algae, marine invertebrates, and fungi, some very specific bacteriophages, and several bacteria such as Bacillus circulans, Klebsiella pneumoniae, Sphingomonas spp., and Vibrio alginolyticus, among others. The role of lyases in organisms that do not produce alginate can be explained by their metabolic ability to use alginate as a sole or secondary carbon source. Interest in these alginate-degrading enzymes has increased in recent years, as it has been shown that alginate-derived oligosaccharides have interesting biotechnological potential, as widely reported in the scientific literature. However, this is a topic that would require an entirely new chapter to adequately address.

Advertisement

Acknowledgments

The author thanks the Department of Biochemistry and Molecular Biology of the Federal University of Paraná, Curitiba- PR, Brazil, and its collaborators who supported this initiative. This work was not funded by any scientific agency or representative of the Brazilian government.

References

  1. 1. Nussinovitch A. Alginates. In: Hydrocolloid Applications: Gum Technology in the Food and Other Industries. Berlin, Germany: Springer Science & Business Media; 2012. pp. 19-39
  2. 2. Draget KI, Smidsrød O, Skjåk-Bræk G. Alginates from algae. In: Polysaccharides and Polyamides in the Food Industry: Properties, Production and Patents. Hoboken, New Jersey, USA: Wiley-Blackwell; 2005. pp. 1-30
  3. 3. Fischer FG, Dörfel H. Die Polyuronsäuren der Braunalgen (Kohlenhydrate der Algen I). Hoppe-Seyler’s. Zeitschrift Für Physiologische Chemie. 1955;302:186-203. DOI: 10.1515/bchm2.1955.302.1-2.186
  4. 4. Haug A, Larsen B, Smidsrød O. Uronic acid sequence in alginate from different sources. Carbohydrate Research. 1974;32:217-225. DOI: 10.1016/s0008-6215(00)82100-x
  5. 5. Szekalska M, Puciłowska A, Szymańska E, Ciosek P, Winnicka K. Alginate: Current use and future perspectives in pharmaceutical and biomedical applications. International Journal of Polymer Science. 2016;2016:1-17. DOI: 10.1155/2016/7697031
  6. 6. Draget KI. Alginates. In: Handbook of Hydrocolloids. Whashington, DC, USA: Elsevier; 2009. pp. 807-828
  7. 7. Jiao W, Chen W, Mei Y, Yun Y, Wang B, Zhong Q, et al. Effects of molecular weight and guluronic acid/mannuronic acid ratio on the rheological behavior and stabilizing property of sodium alginate. Molecules. 2019;24:4374. DOI: 10.3390/molecules24234374
  8. 8. Giridhar RS. Alginates—A seaweed product: Its properties and applications. In: Properties and Applications of Alginates. London, UK: IntechOpen; 2022. DOI: 10.5772/intechopen.98831
  9. 9. Yarkent C, Aslanbay Guler B, Gurlek C, Sahin Y, Kose A, Oncel S, et al. Algal alginate in biotechnology: Biosynthesis and applications. In: Properties and Applications of Alginates. London, UK: IntechOpen; 2022. DOI: 10.5772/intechopen.101407
  10. 10. Hay ID, Rehman ZU, Moradali MF, Wang Y, Rehm BHA. Microbial alginate production, modification and its applications. Microbial Biotechnology. 2013;6:637-650. DOI: 10.1111/1751-7915.12076
  11. 11. May TB, Shinabarger D, Maharaj R, Kato J, Chu L, DeVault JD, et al. Alginate synthesis by Pseudomonas aeruginosa: A key pathogenic factor in chronic pulmonary infections of cystic fibrosis patients. Clinical Microbiology Reviews. 1991;4:191-206. DOI: 10.1128/cmr.4.2.191
  12. 12. Rehm BHA, Valla S. Bacterial alginates: Biosynthesis and applications. Applied Microbiology and Biotechnology. 1997;48:281-288. DOI: 10.1007/s002530051051
  13. 13. Gacesa P. Bacterial alginate biosynthesis—Recent progress and future prospects. Microbiology. 1998;144:1133-1143. DOI: 10.1099/00221287-144-5-1133
  14. 14. Linker A, Jones RS. A polysaccharide resembling alginic acid from a Pseudomonas micro-organism. Nature. 1964;204:187-188. DOI: 10.1038/204187a0
  15. 15. Linker A, Jones RS. A new polysaccharide resembling alginic acid isolated from Pseudomonads. Journal of Biological Chemistry. 1966;241:3845-3851. DOI: 10.1016/s0021-9258(18)99848-0
  16. 16. Gill JF, Deretic V, Chakrabarty AM. Alginate production by the mucoid Pseudomonas aeruginosa associated with cystic fibrosis. Microbiological Sciences. 1987;4(10):296-2999
  17. 17. May TB, Chakrabarty AM. Isolation and assay of Pseudomonas aeruginosa alginate. In: Methods in Enzymology. Cambridge, Massachusetts, USA: Elsevier; 1994. pp. 295-304. DOI: 10.1016/0076-6879(94)35148-1
  18. 18. Pedersen SS, Hoiby N, Espersen F, Koch C. Role of alginate in infection with mucoid Pseudomonas aeruginosa in cystic fibrosis. Thorax. 1992;47:6-13. DOI: 10.1136/thx.47.1.6
  19. 19. Yorgey P, Rahme LG, Tan M-W, Ausubel FM. The roles of mucD and alginate in the virulence of Pseudomonas aeruginosa in plants, nematodes and mice. Molecular Microbiology. 2008;41:1063-1076. DOI: 10.1046/j.1365-2958.2001.02580.x
  20. 20. Gorin PAJ, Spencer JFT. Exocellular alginic acid from Azotobacter vinelandii. Canadian Journal of Chemistry. 1966;44:993-998. DOI: 10.1139/v66-147
  21. 21. Conti E, Flaibani A, O'Regan M, Sutherland IW. Alginate from Pseudomonas fluorescens and P. putida: Production and properties. Microbiology. 1994;140:1125-1132. DOI: 10.1099/13500872-140-5-1125
  22. 22. Sengha SS, Anderson AJ, Hacking AJ, Dawes EA. The production of alginate by Pseudomonas mendocina in batch and continuous culture. Microbiology. 1989;135:795-804. DOI: 10.1099/00221287-135-4-795
  23. 23. Fakhr MK, Alejandro P-V, Chakrabarty AM, Bender CL. Regulation of alginate biosynthesis in Pseudomonas syringae pv. syringae. Journal of Bacteriology. 1999;181:3478-3485. DOI: 10.1128/jb.181.11.3478-3485.1999
  24. 24. Cote GL, Krull LH. Characterization of the exocellular polysaccharides from Azotobacter chroococcum. Carbohydrate Research. 1988;181:143-152. DOI: 10.1016/0008-6215(88)84030-8
  25. 25. Pecina A, Paneque A. Studies on some enzymes of alginic acid biosynthesis in mucoid and non mucoid Azotobacter chroococcum strains. Applied Biochemistry and Biotechnology. 1994;49:51-58. DOI: 10.1007/bf02888847
  26. 26. Davidson IW, Sutherland IW, Lawson CJ. Localization of O-acetyl groups of bacterial alginate. Journal of General Microbiology. 1977;98:603-606. DOI: 10.1099/00221287-98-2-603
  27. 27. Sherbrock-Cox V, Russell NJ, Gacesa P. The purification and chemical characterisation of the alginate present in extracellular material produced by mucoid strains of Pseudomonas aeruginosa. Carbohydrate Research. 1984;135:147-154. DOI: 10.1016/0008-6215(84)85012-0
  28. 28. Skjåk-Bræk G, Grasdalen H, Larsen B. Monomer sequence and acetylation pattern in some bacterial alginates. Carbohydrate Research. 1986;154:239-250. DOI: 10.1016/s0008-6215(00)90036-3
  29. 29. Skjåk-Bræk G, Zanetti F, Paoletti S. Effect of acetylation on some solution and gelling properties of alginates. Carbohydrate Research. 1989;185:131-138. DOI: 10.1016/0008-6215(89)84028-5
  30. 30. Smidsrød O, Draget KI. Alginate gelation technologies. In: Food Colloids: Proteins, Lipids and Polysaccharides. Cambridge, UK: Woodhead Publishing Limited; 2004. pp. 279-293
  31. 31. Wong TY, Preston LA, Schiller NL. Alginate lyase: Review of major sources and enzyme characteristics, structure-function analysis, biological roles, and applications. Annual Review of Microbiology. 2000;54:289-340. DOI: 10.1146/annurev.micro.54.1.289
  32. 32. Abka-khajouei R, Tounsi L, Shahabi N, Patel AK, Abdelkafi S, Michaud P. Structures, properties and applications of alginates. Marine Drugs. 2022;20:364. DOI: 10.3390/md20060364
  33. 33. Wloka M, Rehage H, Flemming H-C, Wingender J. Rheological properties of viscoelastic biofilm extracellular polymeric substances and comparison to the behavior of calcium alginate gels. Colloid and Polymer Science. 2004;282:1067-1076. DOI: 10.1007/s00396-003-1033-8
  34. 34. Boyd A, Chakrabarty AM. Pseudomonas aeruginosa biofilms: Role of the alginate exopolysaccharide. Journal of Industrial Microbiology. 1995;15:162-168. DOI: 10.1007/bf01569821
  35. 35. Hentzer M, Teitzel GM, Balzer GJ, Heydorn A, Molin S, Givskov M, et al. Alginate overproduction affects Pseudomonas aeruginosa biofilm structure and function. Journal of Bacteriology. 2001;183:5395-5401. DOI: 10.1128/jb.183.18.5395-5401.2001
  36. 36. Gloag ES, German GK, Stoodley P, Wozniak DJ. Viscoelastic properties of Pseudomonas aeruginosa variant biofilms. Scientific Reports. 2018;8:9691. DOI: 10.1038/s41598-018-28009-5
  37. 37. Clementi F. Alginate production by Azotobacter vinelandii. Critical Reviews in Biotechnology. 1997;17:327-361. DOI: 10.3109/07388559709146618
  38. 38. Díaz-Barrera A, Sanchez-Rosales F, Padilla-Córdova C, Andler R, Peña C. Molecular weight and guluronic/mannuronic ratio of alginate produced by Azotobacter vinelandii at two bioreactor scales under diazotrophic conditions. Bioprocess and Biosystems Engineering. 2021;44:1275-1287. DOI: 10.1007/s00449-021-02532-8
  39. 39. Segura D, Núñez C, Espín G. Azotobacter cysts. In: eLS. John Wiley & Sons, Ltd(Ed.). Chichester; 2014. pp. 1-8. DOI: 10.1002/9780470015902.a0000295.pub2
  40. 40. Sadoff HL. Encystment and germination in Azotobacter vinelandii. Bacteriological Reviews. 1975;39:516-539. DOI: 10.1128/br.39.4.516-539.1975
  41. 41. Galindo E, Peña C, Núñez C, et al. Molecular and bioengineering strategies to improve alginate and polydydroxyalkanoate production by Azotobacter vinelandii. Microbial Cell Factories. 2007;6:7. DOI: 10.1186/1475-2859-6-7
  42. 42. Ertesvåg H. Alginate-modifying enzymes: Biological roles and biotechnological uses. Frontiers in Microbiology. 2015;6. DOI: 10.3389/fmicb.2015.00523
  43. 43. Kirdponpattara S, Khamkeaw A, Sanchavanakit N, Pavasant P, Phisalaphong M. Structural modification and characterization of bacterial cellulose–alginate composite scaffolds for tissue engineering. Carbohydrate Polymers. 2015;132:146-155. DOI: 10.1016/j.carbpol.2015.06.059
  44. 44. Rehm BHA. Alginate production: Precursor biosynthesis, polymerization and secretion. In: Alginates: Biology and Applications. Berlin, Germany: Springer; 2009. pp. 55-71. DOI: 10.1007/978-3-540-92679-5_2
  45. 45. Moradali MF, Donati I, Sims I, Ghods S, Rehm BHA. Alginate polymerization and modification are linked in Pseudomonas aeruginosa. mBio. 2015;6(3):e00453-e00415. DOI: 10.26686/wgtn.12597782.v1
  46. 46. Dudun AA, Akoulina EA, Zhuikov VA, Makhina TK, Voinova VV, Belishev NV, et al. Competitive biosynthesis of bacterial alginate using Azotobacter vinelandii 12 for tissue engineering applications. Polymers. 2022;14:131. DOI: 10.3390/polym14010131
  47. 47. Clementi F, Fantozzi P, Mancini F, Moresi M. Optimal conditions for alginate production by Azotobacter vinelandii. Enzyme and Microbial Technology. 1995;17:983-988. DOI: 10.1016/0141-0229(95)00007-0
  48. 48. Gauri SS, Mandal SM, Mondal KC, Dey S, Pati BR. Enhanced production and partial characterization of an extracellular polysaccharide from newly isolated Azotobacter sp.. SSB81. Bioresource Technology. 2009;100:4240-4243. DOI: 10.1016/j.biortech.2009.03.064
  49. 49. Díaz-Barrera A, Gutierrez J, Martínez F, Altamirano C. Production of alginate by Azotobacter vinelandii grown at two bioreactor scales under oxygen-limited conditions. Bioprocess and Biosystems Engineering. 2014;37:1133-1140. DOI: 10.1007/s00449-013-1084-2
  50. 50. Flores C, Díaz-Barrera A, Martínez F, Galindo E, Peña C. Role of oxygen in the polymerization and de-polymerization of alginate produced byAzotobacter vinelandii. Journal of Chemical Technology & Biotechnology. 2014;90:356-365. DOI: 10.1002/jctb.4548
  51. 51. Pacheco-Leyva I, Guevara Pezoa F, Díaz-Barrera A. Alginate biosynthesis inAzotobacter vinelandii: Overview of molecular mechanisms in connection with the oxygen availability. International Journal of Polymer Science. 2016;2016:1-12. DOI: 10.1155/2016/2062360
  52. 52. Bonartseva GA, Akulina EA, Myshkina VL, Voinova VV, Makhina TK, Bonartsev AP. Alginate biosynthesis by Azotobacter bacteria. Applied Biochemistry and Microbiology. 2017;53:52-59. DOI: 10.1134/s0003683817010070
  53. 53. Stover CK, Pham XQ, Erwin AL, Mizoguchi SD, Warrener P, Hickey MJ, et al. Complete genome sequence of Pseudomonas aeruginosa pao1, an opportunistic pathogen. Nature. 2000;406:959-964. DOI: 10.1038/35023079
  54. 54. Setubal JC, dos Santos P, Goldman BS, Ertesvåg H, Espín G, Rubio LM, et al. Genome sequence of Azotobacter vinelandii, an obligate aerobe specialized to support diverse anaerobic metabolic processes. Journal of Bacteriology. 2009;191:4534-4545. DOI: 10.1128/jb.00504-09
  55. 55. Gaardløs M, Heggeset TM, Tøndervik A, Tezé D, Svensson B, Ertesvåg H, et al. Mechanistic basis for understanding the dual activities of the bifunctional Azotobacter vinelandii mannuronan C-5-epimerase and alginate lyase AlgE7. Applied and Environmental Microbiology. 2022;88(3). DOI: 10.1128/aem.01836-21
  56. 56. Peña C, Miranda L, Segura D, Núñez C, Espín G, Galindo E. Alginate production by Azotobacter vinelandii mutants altered in poly-β-hydroxybutyrate and alginate biosynthesis. Journal of Industrial Microbiology & Biotechnology. 2002;29:209-213. DOI: 10.1038/sj.jim.7000310
  57. 57. García A, Castillo T, Ramos D, Ahumada-Manuel CL, Núñez C, Galindo E, et al. Molecular weight and viscosifying power of alginates produced by mutant strains of Azotobacter vinelandii under microaerophilic conditions. Biotechnology Reports. 2020;26:e00436. DOI: 10.1016/j.btre.2020.e00436
  58. 58. Gawin A, Tietze L, Aarstad OA, Aachmann FL, Brautaset T, Ertesvåg H. Functional characterization of three Azotobacter chroococcum alginate-modifying enzymes related to the Azotobacter vinelandii AlgE Mannuronan C-5-epimerase family. Scientific Reports. 2020:10. DOI: 10.1038/s41598-020-68789-3
  59. 59. Liang Z, Rybtke M, Kragh KN, Johnson O, Schicketanz M, Zhang YE, et al. Transcription of the alginate operon in Pseudomonas aeruginosa is regulated by c-di-GMP. Microbiology Spectrum. 2022;10(4). DOI: 10.1128/spectrum.00675-22
  60. 60. Mahajan S, Sunsunwal S, Gautam V, Singh M, Ramya TNC. Biofilm inhibitory effect of alginate lyases on mucoid P. aeruginosa from a cystic fibrosis patient. Biochemistry and Biophysics Reports. 2021;26:101028. DOI: 10.1016/j.bbrep.2021.101028
  61. 61. Valentine ME, Kirby BD, Withers TR, Johnson SL, Long TE, Hao Y, et al. Generation of a highly attenuated strain of Pseudomonas aeruginosa for commercial production of alginate. Microbial Biotechnology. 2020;13:162-175. DOI: 10.1111/1751-7915.13411
  62. 62. Tavafi H, Ali AA, Ghadam P, Gharavi S. Screening, cloning and expression of a novel alginate lyase gene from P. aeruginosa tag 48 and its antibiofilm effects on P. aeruginosa biofilm. Microbial Pathogenesis. 2018;124:356-364. DOI: 10.1016/j.micpath.2018.08.018
  63. 63. Wozniak DJ, Ohman DE. Transcriptional analysis of the Pseudomonas aeruginosa genes algR, algB, and algD reveals a hierarchy of alginate gene expression which is modulated by AlgT. Journal of Bacteriology. 1994;176:6007-6014. DOI: 10.1128/jb.176.19.6007-6014.1994
  64. 64. Hay ID, Wang Y, Moradali MF, Rehman ZU, Rehm BH. Genetics and regulation of bacterial alginate production. Environmental Microbiology. 2014;16:2997-3011. DOI: 10.1111/1462-2920.12389
  65. 65. Urtuvia V, Maturana N, Acevedo F, Peña C, Díaz-Barrera A. Bacterial alginate production: An overview of its biosynthesis and potential industrial production. World Journal of Microbiology and Biotechnology. 2017;33(198). DOI: 10.1007/s11274-017-2363-x
  66. 66. García A, Ferrer P, Albiol J, Castillo T, Segura D, Peña C. Metabolic flux analysis and the NAD(P)H/NAD(P)+ ratios in chemostat cultures of Azotobacter vinelandii. Microbial Cell Factories. 2018;17(10). DOI: 10.1186/s12934-018-0860-8
  67. 67. Borges AL, Castro B, Govindarajan S, Solvik T, Escalante V, Bondy-Denomy J. Bacterial alginate regulators and phage homologs repress CRISPR–cas immunity. Nature Microbiology. 2020;5:679-687. DOI: 10.1038/s41564-020-0691-3
  68. 68. Damron FH, Goldberg JB. Proteolytic regulation of alginate overproduction in Pseudomonas aeruginosa. Molecular Microbiology. 2012;84:595-607. DOI: 10.1111/j.1365-2958.2012.08049.x
  69. 69. Remminghorst U, Rehm BH. Bacterial alginates: From biosynthesis to applications. Biotechnology Letters. 2006;28:1701-1712. DOI: 10.1007/s10529-006-9156-x
  70. 70. Moradali MF, Ghods S, Rehm BH. Alginate biosynthesis and biotechnological production. In: Biomaterials Science and Engineering. Singapore: Springer; 2017;11:1-25. DOI: 10.1007/978-981-10-6910-9_1
  71. 71. Dharshini RS, Manickam R, Curtis WR, Rathinasabapathi P, Ramya M. Genome analysis of alginate synthesizing Pseudomonas aeruginosa strain SW1 isolated from degraded seaweeds. Antonie Van Leeuwenhoek. 2021;114:2205-2217. DOI: 10.1007/s10482-021-01673-w
  72. 72. Paletta JL, Ohman DE. Evidence for two promoters internal to the alginate biosynthesis operon in Pseudomonas aeruginosa. Current Microbiology. 2012;65:770-775. DOI: 10.1007/s00284-012-0228-y
  73. 73. Zielinski NA, Maharaj R, Roychoudhury S, Danganan CE, Hendrickson W, Chakrabarty AM. Alginate synthesis in Pseudomonas aeruginosa: Environmental regulation of the ALGC promoter. Journal of Bacteriology. 1992;174:7680-7688. DOI: 10.1128/jb.174.23.7680-7688.1992
  74. 74. Ye RW, Zielinski NA, Chakrabarty AM. Purification and characterization of phosphomannomutase/phosphoglucomutase from Pseudomonas aeruginosa involved in biosynthesis of both alginate and lipopolysaccharide. Journal of Bacteriology. 1994;176:4851-4857. DOI: 10.1128/jb.176.16.4851-4857.1994
  75. 75. Franklin MJ, Chitnis CE, Gacesa P, Sonesson A, White DC, Ohman DE. Pseudomonas aeruginosa AlgG is a polymer level alginate C5-mannuronan epimerase. Journal of Bacteriology. 1994;176:1821-1830. DOI: 10.1128/jb.176.7.1821-1830.1994
  76. 76. Franklin MJ, Ohman DE. Identification of algI and algJ in the Pseudomonas aeruginosa alginate biosynthetic gene cluster which are required for alginate O-acetylation. Journal of Bacteriology. 1996;178:2186-2195. DOI: 10.1128/jb.178.8.2186-2195.1996
  77. 77. Franklin MJ, Ohman DE. Mutant analysis and cellular localization of the AlgI, AlgJ, and AlgF proteins required for o acetylation of alginate in Pseudomonas aeruginosa. Journal of Bacteriology. 2002;184:3000-3007. DOI: 10.1128/jb.184.11.3000-3007.2002
  78. 78. Robles-Price A, Wong TY, Havard S, Valla S, Schiller NL. AlgX is a periplasmic protein required for alginate biosynthesis in Pseudomonas aeruginosa. Journal of Bacteriology. 2004;186:7369-7377. DOI: 10.1128/jb.186.21.7369-7377.2004
  79. 79. Remminghorst U, Rehm BHA. Alg44, a unique protein required for alginate biosynthesis in Pseudomonas aeruginosa. FEBS Letters. 2006;580:3883-3888. DOI: 10.1016/j.febslet.2006.05.077
  80. 80. Remminghorst U, Rehm BH. In vitro alginate polymerization and the functional role of Alg8 in alginate production by Pseudomonas aeruginosa. Applied and Environmental Microbiology. 2006;72:298-305. DOI: 10.1128/aem.72.1.298-305.2006
  81. 81. Castañeda M, Sánchez J, Moreno S, Núñez C, Espín G. The global regulators GacA and δS form part of a cascade that controls alginate production in Azotobacter vinelandii. Journal of Bacteriology. 2001;183:6787-6793. DOI: 10.1128/jb.183.23.6787-6793.2001
  82. 82. Muhammadi AN. Genetics of bacterial alginate: Alginate genes distribution, organization and biosynthesis in bacteria. Current Genomics. 2007;8:191-202. DOI: 10.2174/138920207780833810
  83. 83. Svanem BI, Skjåk-Bræk G, Ertesvåg H, Valla S. Cloning and expression of three new Azotobacter vinelandii genes closely related to a previously described Gene Family encoding Mannuronan C-5-epimerases. Journal of Bacteriology. 1999;181:68-77. DOI: 10.1128/jb.181.1.68-77.1999
  84. 84. Steigedal M, Sletta H, Moreno S, Mærk M, Christensen BE, Bjerkan T, et al. The Azotobacter vinelandii alge Mannuronan C-5-epimerase family is essential for the in vivo control of alginate monomer composition and for functional cyst formation. Environmental Microbiology. 2008;10:1760-1770. DOI: 10.1111/j.1462-2920.2008.01597.x
  85. 85. Moreno S, Ertesvåg H, Valla S, Núñez C, Espin G, Cocotl-Yañez M. RpoS controls the expression and the transport of the alge1-7 epimerases in Azotobacter vinelandii. FEMS Microbiology Letters. 2018;365(19). DOI: 10.1093/femsle/fny210
  86. 86. Svanem BI, Strand WI, Ertesvåg H, Skjåk-Bræk G, Hartmann M, Barbeyron T, et al. The catalytic activities of the bifunctional Azotobacter vinelandii Mannuronan C-5-epimerase and alginate lyase alge7 probably originate from the same active site in the enzyme. Journal of Biological Chemistry. 2001;276:31542-31550. DOI: 10.1074/jbc.m102562200
  87. 87. Wood LF, Ohman DE. Use of cell wall stress to characterize σ22 (algt/U) activation by regulated proteolysis and its regulon in Pseudomonas aeruginosa. Molecular Microbiology. 2009;72:183-201. DOI: 10.1111/j.1365-2958.2009.06635.x
  88. 88. Wang H, Yang Z, Swingle B, Kvitko BH. Algu, a conserved sigma factor regulating abiotic stress tolerance and promoting virulence in Pseudomonas syringae. Molecular Plant-Microbe Interactions. 2021;34:326-336. DOI: 10.1094/mpmi-09-20-0254-cr
  89. 89. Cinthia N, Renato L, Josefina G, Guadalupe E, Gloria S-C. Role of Azotobacter vinelandii muca and mucc gene products in alginate production. Journal of Bacteriology. 2000;182:6550-6556. DOI: 10.1128/jb.182.23.6550-6556.2000
  90. 90. Mejía-Ruíz H, Guzmán J, Moreno S, Soberón-Chávez G, Espín G. The Azotobacter vinelandii alg8 and alg44 genes are essential for alginate synthesis and can be transcribed from an AlgD-independent promoter. Gene. 1997;199:271-277. DOI: 10.1016/s0378-1119(97)00380-6
  91. 91. Lloret L, Barreto R, León R, Moreno S, MartÍnez-Salazar J, Espín G, et al. Genetic analysis of the transcriptional arrangement of Azotobacter vinelandii alginate biosynthetic genes: Identification of two independent promoters. Molecular Microbiology. 1996;21:449-457. DOI: 10.1111/j.1365-2958.1996.tb02554.x
  92. 92. Maleki S, Mærk M, Hrudikova R, Valla S, Ertesvåg H. New insights into Pseudomonas fluorescens alginate biosynthesis relevant for the establishment of an efficient production process for microbial alginates. New Biotechnology. 2017;37:2-8. DOI: 10.1016/j.nbt.2016.08.005
  93. 93. Ertesvåg H, Sletta H, Senneset M, Sun Y-Q, Klinkenberg G, Konradsen TA, et al. Identification of genes affecting alginate biosynthesis in Pseudomonas fluorescens by screening a transposon insertion library. BMC Genomics. 2017;18(11). DOI: 10.1186/s12864-016-3467-7
  94. 94. Schofield MC, Rodriguez DQ, Kidman AA, Cassin EK, Michaels LA, Campbell EA, et al. The anti-sigma factor MucA is required for viability in Pseudomonas aeruginosa. Molecular Microbiology. 2021;116:550-563. DOI: 10.1111/mmi.14732
  95. 95. Li S, Lou X, Xu Y, Teng X, Liu R, Zhang Q, et al. Structural basis for the recognition of MucA by MucB and AlgU in Pseudomonas aeruginosa. The FEBS Journal. 2019;286:4982-4994. DOI: 10.1111/febs.14995
  96. 96. Li T, He L, Li C, Kang M, Song Y, Zhu Y, et al. Molecular basis of the lipid-induced Muca-MucB dissociation in Pseudomonas aeruginosa. Communications Biology. 2020;3(418). DOI: 10.1038/s42003-020-01147-1
  97. 97. Ramsey DM, Wozniak DJ. Understanding the control of Pseudomonas aeruginosa alginate synthesis and the prospects for management of chronic infections in cystic fibrosis. Molecular Microbiology. 2005;56:309-322. DOI: 10.1111/j.1365-2958.2005.04552.x
  98. 98. Xu B, Ju Y, Soukup RJ, Ramsey DM, Fishel R, Wysocki VH, et al. The Pseudomonas aeruginosa AmrZ c-terminal domain mediates tetramerization and is required for its activator and repressor functions. Environmental Microbiology Reports. 2016;8:85-90. DOI: 10.1111/1758-2229.12354
  99. 99. Núñez C, López-Pliego L, Ahumada-Manuel CL, Castañeda M. Genetic regulation of alginate production in Azotobacter vinelandii a bacterium of biotechnological interest: A mini-review. Frontiers in Microbiology. 2022;13. DOI: 10.3389/fmicb.2022.845473
  100. 100. Mærk M, Jakobsen ØM, Sletta H, Klinkenberg G, Tøndervik A, Ellingsen TE, et al. Identification of regulatory genes and metabolic processes important for alginate biosynthesis in Azotobacter vinelandii by screening of a transposon insertion mutant library. Frontiers in Bioengineering and Biotechnology. 2020;7. DOI: 10.3389/fbioe.2019.00475
  101. 101. Quiroz-Rocha E, Bonilla-Badía F, García-Aguilar V, López-Pliego L, Serrano-Román J, Cocotl-Yañez M, et al. Two-component system CBRA/CBRB controls alginate production in Azotobacter vinelandii. Microbiology. 2017;163:1105-1115. DOI: 10.1099/mic.0.000457
  102. 102. Chou S-H, Galperin MY. Diversity of cyclic di-GMP-binding proteins and mechanisms. Journal of Bacteriology. 2016;198:32-46. DOI: 10.1128/jb.00333-15
  103. 103. Amikam D, Galperin MY. Pilz domain is part of the bacterial c-di-GMP binding protein. Bioinformatics. 2006;22:3-6. DOI: 10.1093/bioinformatics/bti739
  104. 104. Merighi M, Lee VT, Hyodo M, Hayakawa Y, Lory S. The second messenger bis-(3′-5′)-cyclic-GMP and its PilZ Domain-containing receptor ALG44 are required for alginate biosynthesis in Pseudomonas aeruginosa. Molecular Microbiology. 2007;65:876-895. DOI: 10.1111/j.1365-2958.2007.05817.x
  105. 105. Aarstad OA, Stanisci A, Sætrom GI, Tøndervik A, Sletta H, Aachmann FL, et al. Biosynthesis and function of long guluronic acid-blocks in alginate produced by Azotobacter vinelandii. Biomacromolecules. 2019;20:1613-1622. DOI: 10.1021/acs.biomac.8b01796
  106. 106. Rosiak P, Latanska I, Paul P, Sujka W, Kolesinska B. Modification of alginates to modulate their physic-chemical properties and obtain biomaterials with different functional properties. Molecules. 2021;26:7264. DOI: 10.3390/molecules26237264
  107. 107. Sabra W. The promise and challenge of microbial alginate production: A product with novel applications. In: Polysaccharides of Microbial Origin. Cham, Switzerland: Springer; 2022. pp. 79-98. DOI: 10.1007/978-3-030-42215-8_5
  108. 108. Sarkar N, Maity A. Modified alginates in drug delivery. In: Tailor-Made Polysaccharides in Drug Delivery. India: Academic Press; 2023. pp. 291-325. DOI: 10.1016/b978-0-12-821286-8.00010-0
  109. 109. Ahmadipour S, Wahart AJC, Dolan JP, Beswick L, Hawes CS, Field RA, Miller GJ. Synthesis of C6-modified mannose 1-phosphates and evaluation of derived sugar nucleotides against GDP-mannose dehydrogenase. Beilstein Journal of Organic Chemistry. 2022;18:1379-1384. DOI: 10.3762/bjoc.18.142
  110. 110. May TB, Shinabarger D, Boyd A, Chakrabarty AM. Identification of amino acid residues involved in the activity of phosphomannose isomerase-guanosine 5′-diphospho-D-mannose pyrophosphorylase. A bifunctional enzyme in the alginate biosynthetic pathway of Pseudomonas aeruginosa. Journal of Biological Chemistry. 1994;269:4872-4877. DOI: 10.1016/s0021-9258(17)37625-1
  111. 111. Regni C, Tipton PA, Beamer LJ. Crystal structure of PMM/PGM. Structure. 2002;10:269-279. DOI: 10.1016/s0969-2126(02)00705-0
  112. 112. Franklin MJ, Nivens DE, Weadge JT, Howell PL. Biosynthesis of the Pseudomonas aeruginosa extracellular polysaccharides, alginate, Pel, and PSL. Frontiers in Microbiology. 2011;2. DOI: 10.3389/fmicb.2011.00167
  113. 113. Gaona G, Nunes C, Goldberg JB, Linford AS, Najera R, Castaneda M, et al. Characterization of theAzotobacter vinelandii algcgene involved in alginate and lipopolysaccharide production. FEMS Microbiology Letters. 2004;238:199-206. DOI: 10.1111/j.1574-6968.2004.tb09756.x
  114. 114. Liu S, Xu N, Liu H, Zhou J, Xin F, Zhang W, et al. Genome characterization of Pseudomonas aeruginosa KT1115, a high di-rhamnolipid-producing strain with strong oils metabolizing ability. Current Microbiology. 2020;77:1890-1895. DOI: 10.1007/s00284-020-02009-z
  115. 115. Tatnell PJ, Russell NJ, Gacesa P. GDP-mannose dehydrogenase is the key regulatory enzyme in alginate biosynthesis in Pseudomonas aeruginosa: Evidence from metabolite studies. Microbiology. 1994;140:1745-1754. DOI: 10.1099/13500872-140-7-1745
  116. 116. Manzo J, Cocotl-Yañez M, Tzontecomani T, Martínez VM, Bustillos R, Velásquez C, et al. Post-transcriptional regulation of the alginate biosynthetic gene algD by the Gac/Rsm system in Azotobacter vinelandii. Microbial Physiology. 2011;21:147-159. DOI: 10.1159/000334244
  117. 117. Hay ID, Ur Rehman Z, Ghafoor A, Rehm BH. Bacterial biosynthesis of alginates. Journal of Chemical Technology & Biotechnology. 2010;85:752-759. DOI: 10.1002/jctb.2372
  118. 118. Remminghorst U, Hay ID, Rehm BHA. Molecular characterization of Alg8, a putative glycosyltransferase, involved in alginate polymerisation. Journal of Biotechnology. 2009;140:176-183. DOI: 10.1016/j.jbiotec.2009.02.006
  119. 119. Saxena IM, Brown RM Jr, Dandekar T. Structure–function characterization of cellulose synthase: Relationship to other glycosyltransferases. Phytochemistry. 2001;57:1135-1148. DOI: 10.1016/s0031-9422(01)00048-6
  120. 120. Oglesby LL, Jain S, Ohman DE. Membrane topology and roles of Pseudomonas aeruginosa Alg8 and Alg44 in alginate polymerization. Microbiology. 2008;154:1605-1615. DOI: 10.1099/mic.0.2007/015305-0
  121. 121. Díaz-Barrera A, Soto E, Altamirano C. Alginate production and alg8 gene expression by Azotobacter vinelandii in continuous cultures. Journal of Industrial Microbiology and Biotechnology. 2012;39:613-621. DOI: 10.1007/s10295-011-1055-z
  122. 122. Akama H, Matsuura T, Kashiwagi S, Yoneyama H, S-ichiro N, Tsukihara T, et al. Crystal structure of the membrane fusion protein, MexA, of the multidrug transporter in Pseudomonas aeruginosa. Journal of Biological Chemistry. 2004;279:25939-25942. DOI: 10.1074/jbc.c400164200
  123. 123. Ahumada-Manuel CL, Martínez-Ortiz IC, Hsueh BY, Guzmán J, Waters CM, Zamorano-Sánchez D, et al. Increased c-di-GMP levels lead to the production of alginates of high molecular mass in Azotobacter vinelandii. Journal of Bacteriology. 2020;202(24). DOI: 10.1128/jb.00134-20
  124. 124. Hay ID, Remminghorst U, Rehm BH. MucR, a novel membrane-associated regulator of alginate biosynthesis in Pseudomonas aeruginosa. Applied and Environmental Microbiology. 2009;75:1110-1120. DOI: 10.1128/aem.02416-08
  125. 125. Martínez-Ortiz IC, Ahumada-Manuel CL, Hsueh BY, Guzmán J, Moreno S, Cocotl-Yañez M, et al. Cyclic di-GMP-mediated regulation of extracellular Mannuronan C-5 epimerases is essential for cyst formation in Azotobacter vinelandii. Journal of Bacteriology. 2020;202(24). DOI: 10.1128/jb.00135-20
  126. 126. Hay ID, Schmidt O, Filitcheva J, Rehm BH. Identification of a periplasmic AlgK–AlgX–MucD multiprotein complex in Pseudomonas aeruginosa involved in biosynthesis and regulation of alginate. Applied Microbiology and Biotechnology. 2012;93:215-227. DOI: 10.1007/s00253-011-3430-0
  127. 127. Jain S, Franklin MJ, Ertesvåg H, Valla S, Ohman DE. The dual roles of algg in C-5-epimerization and secretion of alginate polymers in Pseudomonas aeruginosa. Molecular Microbiology. 2003;47:1123-1133. DOI: 10.1046/j.1365-2958.2003.03361.x
  128. 128. Pawar SN, Edgar KJ. Alginate derivatization: A review of chemistry, properties and applications. Biomaterials. 2012;33:3279-3305. DOI: 10.1016/j.biomaterials.2012.01.007
  129. 129. Baker P, Ricer T, Moynihan PJ, Kitova EN, Walvoort MT, Little DJ, et al. P. aeruginosa SGNH hydrolase-like proteins algj and algx have similar topology but separate and distinct roles in alginate acetylation. PLoS Pathogens. 2014;10(8):e1004334. DOI: 10.1371/journal.ppat.1004334
  130. 130. Riley LM, Weadge JT, Baker P, Robinson H, Codée JDC, Tipton PA, et al. Structural and functional characterization of Pseudomonas aeruginosa AlgX. Journal of Biological Chemistry. 2013;288:22299-22314. DOI: 10.1074/jbc.m113.484931
  131. 131. Vazquez A, Moreno S, Guzmán J, Alvarado A, Espín G. Transcriptional organization of the Azotobacter vinelandii AlgGXLVIFA genes: Characterization of algF mutants. Gene. 1999;232:217-222. DOI: 10.1016/s0378-1119(99)00119-5
  132. 132. Gutsche J, Remminghorst U, Rehm B. Biochemical analysis of alginate biosynthesis protein AlgX from Pseudomonas aeruginosa: Purification of an AlgX-MucD (AlgY) protein complex. Biochimie. 2006;88:245-251. DOI: 10.1016/j.biochi.2005.06.003
  133. 133. Ertesvåg H, Høidal HK, Schjerven H, Svanem BIG, Valla S. Mannuronan C-5-epimerases and their application for in vitro and in vivo design of new alginates useful in biotechnology. Metabolic Engineering. 1999;1:262-269. DOI: 10.1006/mben.1999.0130
  134. 134. Wolfram F, Kitova EN, Robinson H, Walvoort MTC, Codée JDC, Klassen JS, et al. Catalytic mechanism and mode of action of the periplasmic alginate epimerase AlgG. Journal of Biological Chemistry. 2014;289:6006-6019. DOI: 10.1074/jbc.m113.533158
  135. 135. Jerga A, Stanley MD, Tipton PA. Chemical mechanism and specificity of the C5-mannuronan epimerase reaction. Biochemistry. 2006;45:9138-9144. DOI: 10.1021/bi060748f
  136. 136. Valla S, J-ping L, Ertesvåg H, Barbeyron T, Lindahl U. Hexuronyl c5-epimerases in alginate and glycosaminoglycan biosynthesis. Biochimie. 2001;83:819-830. DOI: 10.1016/s0300-9084(01)01313-x
  137. 137. Keiski C-L, Yip P, Robinson H, Burrows LL, Howell PL. Expression, purification, crystallization and preliminary x-ray analysis of Pseudomonas fluorescensalgk. Acta Crystallographica Section F Structural Biology and Crystallization Communications. 2007;63:415-418. DOI: 10.1107/s1744309107016880
  138. 138. Jain S, Ohman DE. Deletion of algk in mucoid Pseudomonas aeruginosa blocks alginate polymer formation and results in uronic acid secretion. Journal of Bacteriology. 1998;180:634-641. DOI: 10.1128/jb.180.3.634-641.1998
  139. 139. Keiski C-L, Harwich M, Jain S, Neculai AM, Yip P, Robinson H, et al. AlgK is a TPR-containing protein and the periplasmic component of a novel exopolysaccharide secretin. Structure. 2010;18:265-273. DOI: 10.1016/j.str.2009.11.015
  140. 140. Grabert E, Wingeder J, Winkler UK. An outer membrane protein characteristic of mucoid strains of Pseudomonas aeruginosa. FEMS Microbiology Letters. 1990;68:83-87. DOI: 10.1111/j.1574-6968.1990.tb04127.x
  141. 141. Rehm BHA, Grabert E, Hein J, Winkler UK. Antibody response of rabbits and cystic fibrosis patients to an alginate-specific outer membrane protein of a mucoid strain of Pseudomonas aeruginosa. Microbial Pathogenesis. 1994;16:43-51. DOI: 10.1006/mpat.1994.1004
  142. 142. Rehm BH. The Azotobacter vinelandii gene algJ encodes an outer-membrane protein presumably involved in export of alginate. Microbiology. 1996;142:873-880. DOI: 10.1099/00221287-142-4-873
  143. 143. Hancock RE, Brinkman FS. Function of Pseudomonas porins in uptake and efflux. Annual Review of Microbiology. 2002;56:17-38. DOI: 10.1146/annurev.micro.56.012302.160310
  144. 144. Rehm BH, Boheim G, Tommassen J, Winkler UK. Overexpression of algE in Escherichia coli: Subcellular localization, purification, and ion channel properties. Journal of Bacteriology. 1994;176:5639-5647. DOI: 10.1128/jb.176.18.5639-5647.1994
  145. 145. Whitney JC, Hay ID, Li C, Eckford PD, Robinson H, Amaya MF, et al. Structural basis for alginate secretion across the bacterial outer membrane. Proceedings of the National Academy of Sciences. 2011;108:13083-13088. DOI: 10.1073/pnas.1104984108
  146. 146. Rehman ZU, Rehm BH. Dual roles of Pseudomonas aeruginosa alge in secretion of the virulence factor alginate and formation of the secretion complex. Applied and Environmental Microbiology. 2013;79:2002-2011. DOI: 10.1128/aem.03960-12
  147. 147. Hoidal HK, Svanem BI, Gimmestad M, Valla S. Mannuronan C-5 epimerases and cellular differentiation of Azotobacter vinelandii. Environmental Microbiology. 2000;2:27-38. DOI: 10.1046/j.1462-2920.2000.00074.x
  148. 148. Bjerkan TM, Bender CL, Ertesvåg H, Drabløs F, Fakhr MK, Preston LA, et al. The Pseudomonas syringae genome encodes a combined mannuronan C-5-epimerase and O-acetylhydrolase, which strongly enhances the predicted gel-forming properties of alginates. Journal of Biological Chemistry. 2004;279:28920-28929. DOI: 10.1074/jbc.m313293200
  149. 149. Ci F, Jiang H, Zhang Z, Mao X. Properties and potential applications of mannuronan C5-epimerase: A biotechnological tool for modifying alginate. International Journal of Biological Macromolecules. 2021;168:663-675. DOI: 10.1016/j.ijbiomac.2020.11.123
  150. 150. Håti AG, Aachmann FL, Stokke BT, Skjåk-Bræk G, Sletmoen M. Energy landscape of alginate-epimerase interactions assessed by optical tweezers and atomic force microscopy. PLoS One. 2015;10(10):e0141237. DOI: 10.1371/journal.pone.0141237
  151. 151. Buchinger E, Knudsen DH, Behrens MA, Pedersen JS, Aarstad OA, Tøndervik A, et al. Structural and functional characterization of the R-modules in alginate C-5 epimerases AlgE4 and AlgE6 from Azotobacter vinelandii. Journal of Biological Chemistry. 2014;289:31382-31396. DOI: 10.1074/jbc.m114.567008
  152. 152. Tøndervik A, Klinkenberg G, Aachmann FL, Svanem BI, Ertesvåg H, Ellingsen TE, et al. Mannuronan C-5 epimerases suited for tailoring of specific alginate structures obtained by high-throughput screening of an epimerase mutant library. Biomacromolecules. 2013;14:2657-2666. DOI: 10.1021/bm4005194
  153. 153. Holtan S, Bruheim P, Skjåk-Bræk G. Mode of action and subsite studies of the guluronan block-forming mannuronan C-5 epimerases AlgE1 and AlgE6. Biochemical Journal. 2006;395:319-329. DOI: 10.1042/bj20051804
  154. 154. Ertesvåg H, Valla S. The A modules of the Azotobacter vinelandii mannuronan-C-5-epimerase AlgE1 are sufficient for both epimerization and binding of Ca2+. Journal of Bacteriology. 1999;181:3033-3038. DOI: 10.1128/jb.181.10.3033-3038.1999
  155. 155. Ramstad MV, Ellingsen TE, Josefsen KD, Høidal HK, Valla S, Skjåk-Bræk G, et al. Properties and action pattern of the recombinant mannuronan C-5-epimerase alge2. Enzyme and Microbial Technology. 1999;24:636-646. DOI: 10.1016/s0141-0229(98)00148-3
  156. 156. Campa C, Holtan S, Nilsen N, Bjerkan TM, Stokke BT, Skjåk-Bræk G. Biochemical analysis of the processive mechanism for epimerization of alginate by mannuronan C-5 epimerase AlgE4. Biochemical Journal. 2004;381:155-164. DOI: 10.1042/bj20031265

Written By

Rodrigo Vassoler Serrato

Submitted: 23 November 2022 Reviewed: 01 December 2022 Published: 23 December 2022