Next Article in Journal
Usage of Cell-Free Protein Synthesis in Post-Translational Modification of μ-Conopeptide PIIIA
Next Article in Special Issue
Structure and Metabolically Oriented Efficacy of Fucoidan from Brown Alga Sargassum muticum in the Model of Colony Formation of Melanoma and Breast Cancer Cells
Previous Article in Journal
Exopolysaccharide Production from Marine-Derived Brevundimonas huaxiensis Obtained from Estremadura Spur Pockmarks Sediments Revealing Potential for Circular Economy
Previous Article in Special Issue
Porphyran from Porphyra haitanensis Enhances Intestinal Barrier Function and Regulates Gut Microbiota Composition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Marine Microbial Polysaccharides: An Untapped Resource for Biotechnological Applications

by
Rajesh Jeewon
1,2,*,
Aadil Ahmad Aullybux
3,
Daneshwar Puchooa
3,
Nadeem Nazurally
3,
Abdulwahed Fahad Alrefaei
2 and
Ying Zhang
4,*
1
Department of Health Sciences, Faculty of Medicine and Health Sciences, University of Mauritius, Réduit 80837, Mauritius
2
Department of Zoology, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
3
Department of Agricultural and Food Science, Faculty of Agriculture, University of Mauritius, Réduit 80837, Mauritius
4
School of Ecology and Natural Conservation, Beijing Forestry University, 35 East Qinghua Road, Haidian District, Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2023, 21(7), 420; https://doi.org/10.3390/md21070420
Submission received: 11 May 2023 / Revised: 18 July 2023 / Accepted: 19 July 2023 / Published: 24 July 2023
(This article belongs to the Special Issue Poly- and Oligosaccharides from Marine Origins)

Abstract

:
As the largest habitat on Earth, the marine environment harbors various microorganisms of biotechnological potential. Indeed, microbial compounds, especially polysaccharides from marine species, have been attracting much attention for their applications within the medical, pharmaceutical, food, and other industries, with such interest largely stemming from the extensive structural and functional diversity displayed by these natural polymers. At the same time, the extreme conditions within the aquatic ecosystem (e.g., temperature, pH, salinity) may not only induce microorganisms to develop a unique metabolism but may also increase the likelihood of isolating novel polysaccharides with previously unreported characteristics. However, despite their potential, only a few microbial polysaccharides have actually reached the market, with even fewer being of marine origin. Through a synthesis of relevant literature, this review seeks to provide an overview of marine microbial polysaccharides, including their unique characteristics. In particular, their suitability for specific biotechnological applications and recent progress made will be highlighted before discussing the challenges that currently limit their study as well as their potential for wider applications. It is expected that this review will help to guide future research in the field of microbial polysaccharides, especially those of marine origin.

Graphical Abstract

1. Introduction

Over the past few years, there has been increasing interest in the development of natural polymers, also referred to as biopolymers, for industrial applications [1], and in particular, polysaccharides have been gaining much attention in the biomedical, cosmetic, food, and pharmaceutical fields. Although polysaccharides can be produced by different types of organisms (e.g., bacteria, fungi, algae, crustaceans, and plants), those from bacteria and fungi have been highly popular as they replicate rapidly, are easier to manipulate, and are abundant producers of those polymers, with the latter also more easily separated than those from non-microbial counterparts [2,3]. Furthermore, in addition to their biological activities, they also display low toxicity, biocompatibility, biodegradability, and other physical and chemical characteristics [4,5,6,7]. Thus, without undermining the value of non-microbial sources of polysaccharides, this review mainly focuses on those obtained from bacteria and fungi in order to highlight their potential for future studies, especially in view of developing polymers for practical applications.
An overview of these microbial compounds suggests that they are usually macromolecules of high molecular weight and are made up of at least 10 monosaccharide units held together by glycosidic bonds [8]. They can also be either linear or branched in structure, while in terms of content, they may be classified as homopolysaccharides or heteropolysaccharides if they, respectively, contain one type or different types of monosaccharides [7,8,9]. In the latter case, while D-glucose is often the most common constituent, other sugars (e.g., D-xylose, D-mannose, D-galactose, L-galactose, L-arabinose, and D-fructose) may also be present alongside derivatives such as simple sugar acids (glucuronic and iduronic acids) or amino sugars (D-glucosamine and D-galactosamine) [10]. Finally, in addition to the above sugars, the presence of non-organic moieties, including sulfates, phosphates, pyruvates, and acetates, is also frequently noted [4,5]. Altogether, different combinations of these variable features contribute to the extensive compositional and structural diversity displayed by microbial polysaccharides.
Given that the structure of polysaccharides ultimately determines their functions, such diversity is actually of functional significance to microorganisms. Indeed, through variations in the monosaccharide composition, non-sugar side chains, glycosidic linkages, and other characteristics, bacteria and fungi are able to synthesize a wide range of polysaccharides that are involved in different biological processes [8]. For instance, some polysaccharides, such as glycogen, occur intracellularly, where they act as storage molecules to provide energy under starvation conditions [4,8,11]. Others are part of bacterial and fungal cell walls or capsules, and as structural polymers, they not only are responsible for maintaining cellular integrity but also assist in other functions such as conferring protection against environmental stresses, regulating membrane permeability, or even mediating interactions with the surroundings, which, in the case of pathogenic microorganisms, may include the onset of immunological responses [12,13,14]. Finally, there are polysaccharides that are secreted outside microbial cells and are thus often termed as exopolysaccharides or extracellular polysaccharides (EPSs). Being commonly synthesized by bacteria and fungi, EPSs are arguably one of the most studied ones, as reflected in the number of publications available on the subject. This can probably be attributed to the fact that, unlike intracellular or cell wall polysaccharides, EPSs are produced in relatively larger amounts within a short time while being more easily isolated and purified. As such, they are better suited for practical applications and will be the main subject of focus in this review [15,16].

2. The Case of Marine Polysaccharides

With around 70% of the Earth’s surface covered with water, the marine environment is undoubtedly an attractive source of microbial polysaccharides [17]. Indeed, aquatic habitats are known to harbor a large diversity of microorganisms, and, in a similar way to terrestrial microbes, it is likely that this also translates into significant diversity in terms of the polysaccharides that they can produce [18], as evidenced by previous reports (e.g., [5,19]). However, increasing interest in marine polysaccharides over the past few years can arguably be attributed to the potential of isolating polysaccharide-producing organisms that can be sources of novel polymers. Indeed, it is often reported that fewer than 1% of marine microorganisms are currently known or cultured, and as a result, the microbial populations in different marine ecosystems remain relatively under-explored [20,21,22]. Therefore, there is an increased likelihood of identifying novel microorganisms, which, in turn, may lead to the isolation of novel polysaccharides with unique properties, especially since some taxa can produce only specific polymers [4]. For instance, in India, Srivastava et al. [23] isolated a novel heteropolysaccharide of around 286 kDa, made up of glucose and galacturonic acid, from the marine bacterium Brevibacillus borstelensis, while in a different study, Rhodotorula mucilaginosa, a marine-derived red yeast, was found to yield a new 1200 KDa exopolysaccharide that consisted of fucose, galactose, mannose, and glucose [24]. Similarly, there are also reports on the isolation of novel exopolysaccharide-producing species or other new marine-derived polymers [25,26,27,28,29]. Table 1 provides some examples of novel polysaccharides and/or novel polysaccharides-producing marine bacteria and fungi that have been reported over the past decade. While such examples may not be exhaustive, they clearly highlight the potential of exploring aquatic species in the search of new polysaccharides.
Furthermore, as already pointed out, polysaccharides are often produced to protect bacteria and fungi from their surroundings. Hence, in response to the unique conditions prevailing in marine environments, it can be expected that aquatic species may develop specific metabolic and physiological capabilities for better adaptation, thereby resulting in the production of compounds, including polysaccharides, which may be absent from terrestrial microbes [48]. This was highlighted in the study by Abdel-Wahab et al. [37], in which a marine strain of the fairly common Bacillus subtilis species yielded a novel β-glycosidic sulfated heteropolysaccharide. This polymer, consisting of glucose, rhamnose, and arabinose, could also exhibit a wide range of biological activities (anti-oxidant, anti-inflammatory, cytotoxicity, and anti-Alzheimer activities). Such protective functions have been of particular value in studies involving microorganisms from extreme marine habitats. Indeed, some species are able to survive in specific areas characterized by very high or low temperatures (thermophiles and psychrophiles), high or low pH (acidophiles and alkalophiles), high pressures (piezophiles), or even high ionic strengths (halophiles) [49]. In these cases, the extremophiles adopt specific survival strategies, which include but are not limited to the production of polysaccharides with unique properties [50,51], with examples of such polysaccharides that have been isolated from extremophilic microorganisms during the last decade, provided in Table 2. Thus, it can be expected that the study of microbial species from extreme environments could yield polysaccharides with new or improved properties.
An overview of the above tables suggests some common features in the study of microbial polysaccharides. Firstly, the high diversity of novel bacterial and fungal polysaccharides is quite obvious, especially in terms of the molecular weight, composition, and biological activities, all of which further lend support to the potential of exploring such microbial compounds. However, although the above list is not exhaustive, there seems to be a greater focus on polysaccharides from marine bacteria as compared to fungal sources, probably due to the former’s higher diversity as free-living organisms, as well as their ease of isolation in terms of growth requirements. This is also particularly obvious as far as polymers from extremophiles are concerned. Nevertheless, fungal sources remain a major source of polysaccharides, and research on such species is still likely to yield polysaccharides of interest. In addition, in terms of the level of characterization, determining the molecular weight, functional groups, and monosaccharide composition tend to be standard practices, but interestingly, a survey of the literature suggests that structural characterizations, even partial ones, seem to also be gaining in importance. However, studies that fully establish structure–function relationships for polysaccharides in view of explaining their biological activities are still far from common, probably because a highly technical and experimental analysis is required to determine the structures of such high molecular weight polymers [37].

3. Current Research on Marine Microbial Polysaccharides

Over the years, the potential of microbial polysaccharides for biotechnological applications became increasingly recognized, as reflected by the number of studies on the subject, with Osemwegie et al. [2], as reported by Nadzir et al. [57], identifying thousands of publications focused on microbial polysaccharides between 1976 and 2018 alone. Although many of these did not specifically involve marine species, a survey of the literature suggests a similar trend as far as marine microbial polysaccharides are concerned. The following sections provide an overview of the current trends in research on marine polysaccharides from bacterial and fungal species.

3.1. Biomedical Applications

One of the most promising properties displayed by marine microbial polysaccharides is their biological activities. Indeed, it is not uncommon for studies to investigate such characteristics in view of presenting these polymers as attractive candidates for biomedical applications, with some of the most studied biological activities being anticancer, antimicrobial, anti-oxidant, and immunomodulation [18].

3.1.1. Anticancer Activity

Cancer, characterised by an uncontrolled proliferation of cells, is currently one of the major diseases affecting human health, as well as the second cause of death in the world, with an estimated 18.1 million people diagnosed with the condition in 2018 [58,59]. So far, surgery, chemotherapy, and radiotherapy remain the main forms of treatment for different types of cancers, but these approaches are not without side effects, which include a number of health complications as well as toxicity and/or injury to non-targeted organs and cells [58,60,61]. Consequently, the search for alternative forms of treatment has prompted interest in natural compounds, with results of studies often highlighting the potential of marine microbial polysaccharides for such applications.
Indeed, the anticancer activities of polysaccharides from marine bacteria and fungi are already well established, with cytotoxic effects often reported against lung [32], liver [29,31,62,63], breast [59,62,64,65,66], cervical [62], and colorectal cancers [59]. In these cases, the apoptosis of cancer cells seems to be a common mechanism, although programmed cell death can be mediated through different pathways. These different mechanisms of action are particularly obvious from studies in which different polysaccharides were tested against different cell lines. For example, Tukenmez et al. [67] noted changes in the gene and protein expression of Bax, Bcl-2, Caspase 3, Caspase 9, and Survivin when EPSs of L. delbrueckii ssp. Bulgaricus were tested against colon cancer cells at a concentration of 400 µg/mL for 24 h or 48 h. The EPS consisted of glucose, mannose, fructose, sucrose, maltose, and N-acetylglucosamine, with the observed effects attributed to the glucose and mannose content. In contrast, within a concentration range of 5–80 μg/mL, marine polysaccharides from Bacillus velezensis activated caspase-3 while increasing levels of cytochrome C to induce apoptosis in breast cancer cells [65]. Such differences in the mechanism of action could arguably be attributed to differences in the composition and/or structure of the two polysaccharides, as suggested by Tukenmez et al. [67], although the influence of other factors is not excluded. In addition, within similar cell lines, anticancer effects may also occur through different mechanisms. For instance, Cao et al. [32] reported the isolation of EPS11, a 22.3 kDa polysaccharide fraction made up of glucose, mannose, xylose, glucosamine, and galacturonic acid, from a marine Bacillus species. At varying concentrations up to 90 nM, the authors showed that this polymer could not only affect cell proliferation and adhesion of lung (A549) and liver (Huh7.5) cells, but also induce apoptosis by preventing the expression of βIII-tubulin, as well as reducing the phosphorylation of protein kinase B (PKB or AKT). However, the same polysaccharide also downregulated proteins related to the extracellular matrix–receptor interaction signaling pathway and targeted collagen I through the β1-integrin-mediated signaling pathway to prevent cell adhesion, migration, and invasion [31]. This feature of polysaccharides is, in fact, of practical significance, as the ability to induce anticancer effects through different mechanisms may reduce the likelihood that a particular cell line develops resistance to therapy, as is currently the case for a number of chemotherapeutic drugs [68].

3.1.2. Antimicrobial Activity

Another biological activity of microbial polysaccharides that is commonly investigated is their ability to inhibit the growth and/or proliferation of pathogenic microorganisms. This property is particularly relevant nowadays due to the emergence of drug-resistant pathogens that constantly threaten public health, thereby prompting the need to develop new and more potent antibiotics [69,70]. In this context, Aullybux et al. [71] isolated two sulfated EPSs from a marine Alcaligenes and Halomonas sp., which could, in addition to different pathogens, also inhibit the growth of methicillin-resistant Staphylococcus aureus (MRSA) at concentrations between 0.25 and 2 mg/mL. Although not specifically attributed to any particular features of the polymers, the authors argued that the antibacterial activities could be linked to the presence of certain functional groups that are known to act as metal chelators. Similarly, polysaccharides (1 mg/mL) from a haloalkalitolerant Alkalibacillus sp., recovered from a salt lake, displayed antibacterial effects against Candida albicans, as well as a number of Gram-positive and Gram-negative bacteria [72].
Although these studies did not determine the underlying mechanism of the antimicrobial properties, this can be inferred based on existing reports from other non-marine polysaccharides. For example, electrostatic interactions between oppositely charged polysaccharides and pathogens’ cell walls, as well as the latter’s subsequent hydrolysis to leak cell content, have been suggested as one of the mechanisms responsible for the observed antimicrobial activities [71]. Similarly, Zhou et al. [73] proposed that interactions between polysaccharides and biofilm-related signal molecules or cell-surface receptors of pathogens could disrupt cell communication and biofilm formation, while Rajoka et al. [74] suggested that metal chelation, as well as nutrient suppression through the formation of an external barrier, could represent additional ways through which antimicrobial properties are exerted.
However, while the antimicrobial properties of bacterial and fungal polysaccharides are well known, those derived from marine species are yet to be widely studied, as is the case for their terrestrial counterparts. As will be discussed in subsequent sections, this could likely be due to the different challenges encountered in the study of marine microbial polysaccharides. Nevertheless, this undoubtedly represents a research gap that, if addressed, could potentially yield new classes of antibiotics to assist the fight against resistant pathogens [75].

3.1.3. Anti-Oxidant Activity

Bacterial or fungal polysaccharides, especially those that display biological activities, have been considered not only for their cytotoxicity or antibacterial effects but also for their anti-oxidant potential [76,77]. Nowadays, it is not uncommon to come across studies that highlight the anti-oxidant potential of compounds, probably because it is intrinsically linked to human health.
Tissues in the human body require oxygen for energy production, but as the oxygen is consumed, those tissues generate free radicals as by-products [78]. Free radicals are, basically, reactive and unstable molecules containing unpaired electrons, and they can be grouped into either reactive nitrogen species (RNSs) or reactive oxygen species (ROSs) [78,79]. These compounds are normally maintained at a suitable concentration by balancing the body’s production of free radicals through a defence system involving anti-oxidant enzymes [80]. However, in addition to cellular production, external sources such as radiation, chemicals, pollutants, cigarettes, alcohol, some drugs, or heavy metals, just to name a few, can also contribute substantially to the levels of ROS/RNS [80]. These can undoubtedly cause an imbalance, especially if the body is no longer able to counteract the additional production of free radicals, hence resulting in a condition of oxidative/nitrosative stress. Under this condition, these excess radicals, especially ROS, interact with various biological macromolecules such as proteins, DNA, RNA, and lipids, causing their structural and functional alterations [80,81]. Given that these macromolecules have important physiological functions, it is, therefore, not surprising that oxidative/nitrosative stress has been established as a major cause of human diseases, which include cardiovascular, organ disorders (pancreas, lungs, eyes, kidneys, and joints), neurodegenerative diseases, and even cancer [78,80].
In light of the above, it can be understood why anti-oxidant compounds are highly regarded as being beneficial for health. Indeed, anti-oxidants are molecules that, at low concentrations, inhibit or cause a significant delay in the oxidation of compounds [82], and as such, they help to hinder the negative effects of oxidative stress. Interestingly, reports on the anti-oxidant activities of polysaccharides from marine bacteria and fungi are not lacking, with some even involving novel species (e.g., Enterobacter cloacae MBB8) or novel polymers (e.g., EPSR4 from Bacillus subtilis) [25,37,83,84]. However, while this commonly studied activity of marine microbial EPS is now fairly established, the next step would undoubtedly be developing practical applications for such polymers. In this context, it is worth noting that providing anti-oxidants as supplements has been suggested as a means of mitigating the negative effects associated with oxidative stress [85]. Hence, the successful isolation of anti-oxidant EPS from two probiotic marine bacteria (Rhodotorula sp. and Pediococcus pentosaceus), as reported by Wang et al. [86] and Ayyash et al. [87], further highlights the potential of studying marine microorganisms for this purpose.

3.1.4. Drug Delivery

Besides their biological activities, microbial polysaccharides are also suitable to enhance the activities of other compounds, and for this purpose, they are often applied in the development of drug-delivery systems. Such nano-based or targeted delivery of therapeutic agents ensures that the latter are delivered at the required site and in a controlled manner, thereby overcoming the limitations (e.g., drug bioavailability, unwanted side effects, and non-specificity of drugs) that are encountered with current methods of drug delivery [88,89]. For this purpose, the selection of a suitable carrier molecule is a key parameter that needs to be considered, as its properties eventually influence the mechanism of drug release [90]. However, more importantly, the selected carrier would need to be biodegradable, biocompatible, and safe in order to be considered for such applications, and interestingly, microbial polysaccharides display such characteristics [90,91].
Dextran, composed of a linear chain of D-glucose linked by α-(1→6) bonds and commonly synthesized by lactic acid bacteria, is a widely used EPS for developing targeted delivery systems [92]. As a result of its non-toxicity, non-immunogenicity, and biocompatibility, dextran represents a suitable polymer for encapsulating or adsorbing therapeutic agents and ensuring their delivery while effectively providing protection against the immune system, as well as digestive enzymes [93]. The efficacy of such systems was demonstrated by Wang et al. and Fang et al., who developed dextran-based nanocarriers for the delivery of doxorubicin, with the results confirming the improved anticancer effects alongside reduced toxicity to the drug [94,95]. Similarly, chitosan (a derivative of chitin) and levan (a fructose homopolymer) also display properties such as biocompatibility, biodegradability, and low toxicity [96,97]. These features were further confirmed by studies whereby cisplatin, 5-fluorouracil, and resveratrol were successfully loaded onto those polysaccharides for delivery to cancer cells while being safe for healthy ones [98,99,100].
While there are reports on the production of the above polysaccharides from marine microorganisms (e.g., Penicillum spinulosum and Halomonas sp.) [101,102], it is surprising to note that such applications are yet to be established for those obtained from marine bacteria or fungi. This was reflected in an overview of two recent reviews on the application of marine microbial EPS as drug carriers, whereby the focus was largely, if not completely, on polysaccharides from non-marine sources [90,91]. However, the absence of a significant number of studies on the subject does not suggest that the potential of marine microbial polysaccharides has been overlooked. For instance, in an attempt to develop microgels as protein carriers, Zykwinska et al. successfully assembled EPSs from Vibrio diabolicus, a deep-sea hydrothermal bacterium, for the encapsulation of bovine serum albumin [103]. As a possible extension to that study, the authors subsequently isolated EPS from Alteromonas infernus to yield microcarriers that could encapsulate Transforming Growth Factor-β1 (TGF-β1) for applications in cartilage engineering [104]. Similarly, K1T-9, a 207 kDa heteropolysaccharide isolated from the novel marine bacterium Neorhizobium urealyticum, was successfully applied as an emulsifier for the encapsulation of astaxanthin [39]. Therefore, given the diversity and possibly unique properties of marine microbial polysaccharides, addressing the current research gap could potentially lay the foundations for the development of new or improved drug-delivery systems based on these polymers.

3.2. Bioremediation

Over the past few decades, industrialization and other human activities such as the improper disposal of wastes and the excessive use of pesticides and fertilizers have been a major cause of environmental pollution [105]. In particular, water contamination via heavy metals such as chromium (Cr), cadmium (Cd), lead (Pb), arsenic (As), mercury (Hg), and silver (Ag), just to name a few, has been of concern not only due to their non-biodegradability and toxicity at certain concentrations but also due to their potential accumulation along the food chain [106]. Therefore, the removal of heavy metal contaminants has been devised as an effective strategy for treating polluted areas, and in this context, electroplating, ion exchange, precipitation, and membrane processes represent some of the most commonly used approaches for this purpose [106,107]. However, these methods are not without drawbacks, the most important of which include the high cost involved, its low efficiency, and the production of toxic by-products [106]. As a result, attention has shifted to better alternatives, with microbial-based treatments proving to be a suitable candidate for such applications.
Indeed, microorganisms, especially those from heavy-metal-polluted areas, have evolved to develop tolerance to such pollutants, and therefore, they can be ideal candidates for bioremediation processes [108,109]. While the microbial-based removal of toxic heavy metals from the environment can be mediated through different pathways [109], the current review will focus on the potential of their polysaccharides for this purpose. Microbial polysaccharides contain a number of functional groups such as carboxyl, hydroxyl, phosphate, amine, and uronic acids, with marine-derived ones being particularly rich in the latter [5,110]. These groups confer an overall negative charge to the polymers, thereby allowing them to bind to the positively charged heavy metals through the process of adsorption and subsequently leading to their removal [5]. For instance, a Bacillus cereus strain, isolated from a contaminated estuarine sediment, showed potential for water detoxification at concentrations of 25 to 150 mg/L due to its EPSs’ affinity for Pb, Cu, and Cd [106]. In this case, higher adsorption capacity, largely attributed to the functional groups present, occurred at the lower doses. Similarly, Concórdio-Reis et al. [111] reported the isolation of the EPS FucoPol from an Enterobacter species. This polymer, which showed specificity towards Pb, had an overall metal removal efficiency of 91.6–93.9% and this was achieved at a concentration of 5 g/L through its carboxyl and hydroxyl groups, under acidic conditions and within a temperature range of 5–40 °C. A different study further highlighted the potential of marine environments as a source of novel polysaccharides for bioremediation processes. Indeed, Zhang et al. [38] reported the ability of a novel polymer from an Alteromonas species to adsorb Cu, Ni, and Cr at a concentration of 1 g/L, thus indicating its suitability for the removal of heavy metals. However, despite its novel structure, the observed effects were still attributed to the functional groups present, especially O-H, C=O, and C-O-C, as is often the case for other microbial polysaccharides.
Closely related to the above are potential applications of marine microbial polysaccharides as bioflocculants. Flocculation refers to the aggregation of small suspended particles into larger flocs to aid their removal, and while it is not limited to heavy metals, it is nevertheless also applied to water treatment [112]. In this context, Chen et al. characterized a novel bioflocculant from the marine Alteromonas species [113]. Largely composed of polysaccharides, this polymer (20–220 mg/L) could effectively remove dyes such as Methylene Blue, Direct Black, and Congo Red at efficiencies between 72.3% and 98.5%, thus proving to be effective for the treatment of dyed wastewater. In a different study, a marine Bacillus species achieved 85% bioflocculant activity, with the optimum conditions for such activities also determined [112]. Overall, based on the above studies, it would not be unlikely that the huge diversity of marine microbial polysaccharides, as well as their specificity to pollutants, could drive the search for new polymers. At the same time, the fact that the presence of a polysaccharide backbone can enhance the thermal stability of bioflocculants could spark additional interest in the isolation of such polymers.
Although the above examples are not exhaustive, a key factor that explains the wide interest in polysaccharides is the remarkable structural diversity displayed by these polymers, as this translates into a wide range of properties, as well as potential applications. Thus, any prospective uses of polysaccharides are often dependent on structural characteristics such as the monosaccharide composition, the type of functional groups present, or even the conformation, including the degree of branching and the type of linkage [114]. For instance, the presence of neutral monosaccharides (e.g., glucose, mannose, fucose, arabinose, D-galactose, and glucuronic acid) has been reported as being more likely to induce anti-oxidant activities [76], while in their study on polysaccharides from Lactobacillus reuteri, Chen et al. [115] noted that the amount of galactose was related to the anti-inflammatory activity of the polymer, with higher amounts of the monosaccharide enhancing the biological activity. Similarly, in terms of conformations, the types of glycosidic linkages may affect the solubility and flexibility of polysaccharide chains, thus making them more or less suited to certain specific applications [67,76]. In other cases, polymers with mostly β-1,3-linkages were also reported as displaying greater antitumor activities in contrast to those containing mostly β-1,6- linkages [67]. Finally, as far as functional groups are concerned, biological activities are often observed when specific groups are present. For example, phosphate groups can contribute to the immunomodulatory effects of polysaccharides by improving their affinity to immune cells. In addition, compared with neutral polysaccharides, phosphorylated ones are also more likely to inhibit the growth of certain cancer cells [116,117]. Similarly, sulfated or acetylated polysaccharides can display better biological activities than non-sulfated ones, especially antibacterial, anti-oxidant, and antitumor effects [116,118,119]. As discussed in the subsequent section, the influence of the functional groups on the biological properties of polysaccharides is actually of great significance in generating specific polymers of practical value.

4. Challenges for the Commercialization of Marine Microbial Polysaccharides

Despite the wide interest in microbial polysaccharides, as well as their great potential for a number of applications, only a few of these polymers have actually been commercialized so far, with even fewer being from marine microorganisms (e.g., HE800 EPS, bearing the trademark Hyalurift® [120], and HYD657, commercially available as Abyssine® [121]). However, a survey of deposited patents related to marine microbial polysaccharides does suggest that the potential of these polymers for a wide range of applications is recognized (Table 3). Hence, while learning the barriers to commercialization may not necessarily result in more products, addressing them may potentially increase the likelihood that the results of research are eventually translated into practical applications.
One of the main factors that limit the production of polysaccharides is the high cost of production, even though microorganisms are relatively easier to grow and manipulate, are better producers, and are, overall, cheaper sources of biopolymers in comparison with non-microbial sources [5,122]. In this case, improving the yield of polysaccharides is a commonly applied strategy to make the process more cost-effective, and this can often be achieved through process optimization, whereby the growth conditions that maximize yields are identified and applied. For instance, Hereher et al. managed to increase EPS production from a Micrococcus roseus strain by over four times by modifying the amount of sucrose and ammonium sulfate, as well as the incubation temperature and pH [123]. Similarly, a three-fold increase in EPS yield was reported for a Halomonas xianhensis strain by optimizing culture conditions [124]. In addition, favoring the use of cheaper substrates has also been proposed as a means of bringing down the overall cost of production [125,126].
However, it should be noted that it is not uncommon for some microbial species to still have relatively low yields despite process optimization, and, in this case, the genetic modification of the microorganisms can provide a powerful alternative to improve yields, especially since they are more easily amenable to genetic changes compared to higher organisms [127]. Such modifications can range from increasing the availability of EPS precursors, overexpressing genes involved in EPS assembly, or even knocking out those that compete against EPS production. For instance, the marine yeast Aureobasidium melanogenum was successfully modified with the INU1 gene to improve pullulan production by more than five times compared with the wild strain [128]. At the same time, in addition to better yields, this approach may also help in generating tailor-made polysaccharides with better or novel physicochemical and biological properties [129]. Interestingly, as pointed out by Wang et al. [130], these benefits can be particularly suited to extremophilic microorganisms, which, despite their potential to synthesize EPS with distinct features and properties, are relatively poor producers of such polymers.
Another commonly applied strategy that has been proposed as a means of improving the value of isolated microbial polysaccharides is structural modification. As noted before, the properties of microbial polysaccharides are tightly linked to their structures, and therefore, this approach is particularly suited to cases where the isolation of closely related microbial species yields polysaccharides of nearly similar structures or those with no biological activities, both of which may not bring significant additional value to existing research [9,131]. In other cases, although polysaccharides may display the potential for specific applications, the absence of suitable physico-chemical properties (e.g., poor solubility) may hinder subsequent interest in their exploitation [50]. Thus, by altering the structures of such polymers through physical, biological, and chemical means, their existing properties can be enhanced and tailored for individual applications [50,131]. For instance, through a sulfation modification, Chopin et al. [132] improved the ability of GY785, an EPS from the bacterial strain Alteromonas infernus isolated from deep-sea hydrothermal vents, to drive the chondrogenic differentiation of mesenchymal stem cells, hence enhancing its potential for cartilage repair. Furthermore, there are also reports on phosphorylation modifications, which not only enhance existing biological activities of polysaccharides but also help to induce such activities in polymers in which they are naturally absent [131,133]. Overall, while reports of such modifications of polysaccharides from marine bacteria and fungi are not extensive, it is likely that this approach can be a powerful tool that could increase the likelihood of developing commercially viable marine microbial polymers. However, for most reported microbial polysaccharides, the relationship between structure and biological functions is not precisely established, hence making it more difficult for commercialization [57]. Therefore, the above examples clearly highlight the need to characterize the structures of microbial polysaccharides as part of studies, especially since such information would provide an understanding of potential modifications that could be undertaken in view of obtaining desirable properties.
According to Li et al. [58], the low application of polysaccharides can also be linked to the lack of in vivo research. Although studies involving animal models are not completely absent, integrating more in vivo research is more likely to provide a better insight into the actual potential of polysaccharides, especially as far as their biomedical applications are concerned. Finally, as noted before, most marine microorganisms remain unculturable, and this clearly limits studies on marine polysaccharides, as these require culturable samples for extracting and characterizing the polymers [134]. Therefore, devising new culture methods to improve the isolation of previously uncultured microorganisms can pave the way to the production of novel polysaccharides.

5. Conclusions and Future Perspectives

This paper provides an overview of marine microbial polysaccharides, their peculiarities, and current trends in their study. Research gaps, including challenges to their practical applications, were also briefly addressed. With much of the marine environment still relatively unexplored, there is no doubt that research on marine microbial polysaccharides remains promising. In particular, focusing on extreme environments may yield polymers which are not only unique in terms of their properties but also of potential high commercial value. In fact, as talks on climate change and the need to reduce petroleum-derived compounds step up, switching to sustainable alternatives is likely to take on even greater importance, thereby further fuelling interest in marine microbial polysaccharides.

Author Contributions

Conceptualization, R.J. and A.A.A.; methodology, R.J. and A.A.A.; writing—original draft preparation, R.J. and A.A.A.; writing—review and editing, R.J., A.A.A., N.N., A.F.A., Y.Z. and D.P.; supervision, R.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the Distinguished Scientist Fellowship Program (DSFP) at King Saud University, Riyadh, Saudi Arabia. The University of Mauritius and Beijing Forestry University are also thanked for their research support. Nabeelah B. Sadeer is also thanked for her assistance with chemical structures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Baranwal, J.; Barse, B.; Fais, A.; Delogu, G.L.; Kumar, A. Biopolymer: A Sustainable Material for Food and Medical Applications. Polymers 2022, 14, 983. [Google Scholar] [CrossRef] [PubMed]
  2. Osemwegie, O.O.; Adetunji, C.O.; Ayeni, E.A.; Adejobi, O.I.; Arise, R.O.; Nwonuma, C.O.; Oghenekaro, A.O. Exopolysaccharides from bacteria and fungi: Current status and perspectives in Africa. Heliyon 2020, 6, e04205. [Google Scholar] [CrossRef] [PubMed]
  3. Zaghloul, E.H.; Ibrahim, M.I.A. Production and Characterization of Exopolysaccharide from Newly Isolated Marine Probiotic Lactiplantibacillus plantarum EI6 with in vitro Wound Healing Activity. Front. Microbiol. 2022, 13, 903363. [Google Scholar] [CrossRef] [PubMed]
  4. Khan, R.; Shah, M.D.; Shah, L.; Lee, P.-C.; Khan, I. Bacterial polysaccharides—A big source for prebiotics and therapeutics. Front. Nutr. 2022, 9, 1031935. [Google Scholar] [CrossRef]
  5. Ibrahim, H.A.H.; Elhassayeb, H.E.A.; El-Sayed, W.M.M. Potential functions and applications of diverse microbial exopolysaccharides in marine environments. J. Genet. Eng. Biotechnol. 2022, 20, 151. [Google Scholar] [CrossRef]
  6. Liu, Z.-X.; Huang, S.-L.; Hou, J.; Guo, X.-P.; Wang, F.-S.; Sheng, J.-Z. Cell-based high-throughput screening of polysaccharide biosynthesis hosts. Microb. Cell Factories 2021, 20, 62. [Google Scholar] [CrossRef]
  7. Mahmoud, Y.A.-G.; El-Naggar, M.E.; Abdel-Megeed, A.; El-Newehy, M. Recent Advancements in Microbial Polysaccharides: Synthesis and Applications. Polymers 2021, 13, 4136. [Google Scholar] [CrossRef]
  8. Zeidan, A.A.; Poulsen, V.K.; Janzen, T.; Buldo, P.; Derkx, P.M.F.; Øregaard, G.; Neves, A.R. Polysaccharide production by lactic acid bacteria: From genes to industrial applications. FEMS Microbiol. Rev. 2017, 41 (Suppl. 1), S168–S200. [Google Scholar] [CrossRef] [Green Version]
  9. Ahmad, N.H.; Mustafa, S.; Man, Y.B.C. Microbial Polysaccharides and Their Modification Approaches: A Review. Int. J. Food Prop. 2014, 18, 332–347. [Google Scholar] [CrossRef] [Green Version]
  10. Mohammed, A.S.A.; Naveed, M.; Jost, N. Polysaccharides; Classification, Chemical Properties, and Future Perspective Applications in Fields of Pharmacology and Biological Medicine (A Review of Current Applications and Upcoming Potentialities). J. Polym. Environ. 2021, 29, 2359–2371. [Google Scholar] [CrossRef]
  11. Zeitz, M.A.; Tanveer, Z.; Openshaw, A.T.; Schmidt, M. Genetic Regulators and Physiological Significance of Glycogen Storage in Candida albicans. J. Fungi 2019, 5, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Garcia-Rubio, R.; De Oliveira, H.C.; Rivera, J.; Trevijano-Contador, N. The Fungal Cell Wall: Candida, Cryptococcus, and Aspergillus Species. Front. Microbiol. 2020, 10, 2993. [Google Scholar] [CrossRef] [PubMed]
  13. Gow, N.A.R.; Latge, J.-P.; Munro, C.A. The Fungal Cell Wall: Structure, Biosynthesis, and Function. Microbiol. Spectr. 2017, 5, 28513415. [Google Scholar] [CrossRef] [Green Version]
  14. Agustinho, D.P.; Miller, L.C.; Li, L.X.; Doering, T.L. Peeling the onion: The outer layers of Cryptococcus neoformans. Memórias Inst. Oswaldo Cruz 2018, 113, e180040. [Google Scholar] [CrossRef] [Green Version]
  15. Mahapatra, S.; Banerjee, D. Fungal Exopolysaccharide: Production, Composition and Applications. Microbiol. Insights 2013, 6, MBI-S10957. [Google Scholar] [CrossRef] [Green Version]
  16. Pereira, S.B.; Sousa, A.; Santos, M.; Araújo, M.; Serôdio, F.; Granja, P.; Tamagnini, P. Strategies to Obtain Designer Polymers Based on Cyanobacterial Extracellular Polymeric Substances (EPS). Int. J. Mol. Sci. 2019, 20, 5693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Ruocco, N.; Costantini, S.; Guariniello, S.; Costantini, M. Polysaccharides from the Marine Environment with Pharmacological, Cosmeceutical and Nutraceutical Potential. Molecules 2016, 21, 551. [Google Scholar] [CrossRef] [PubMed]
  18. Qi, M.; Zheng, C.; Wu, W.; Yu, G.; Wang, P. Exopolysaccharides from Marine Microbes: Source, Structure and Application. Mar. Drugs 2022, 20, 512. [Google Scholar] [CrossRef]
  19. Casillo, A.; Lanzetta, R.; Parrilli, M.; Corsaro, M.M. Exopolysaccharides from Marine and Marine Extremophilic Bacteria: Structures, Properties, Ecological Roles and Applications. Mar. Drugs 2018, 16, 69. [Google Scholar] [CrossRef] [Green Version]
  20. Mac Rygaard, A.; Thøgersen, M.S.; Nielsen, K.F.; Gram, L.; Bentzon-Tilia, M. Effects of Gelling Agent and Extracellular Signaling Molecules on the Culturability of Marine Bacteria. Appl. Environ. Microbiol. 2017, 83, e00243-17. [Google Scholar] [CrossRef] [Green Version]
  21. Sen, K.; Sen, B.; Wang, G. Diversity, Abundance, and Ecological Roles of Planktonic Fungi in Marine Environments. J. Fungi 2022, 8, 491. [Google Scholar] [CrossRef] [PubMed]
  22. Gonçalves, M.F.M.; Esteves, A.C.; Alves, A. Marine Fungi: Opportunities and Challenges. Encyclopedia 2022, 2, 559–577. [Google Scholar] [CrossRef]
  23. Srivastava, N.; Kumari, S.; Kurmi, S.; Pinnaka, A.K.; Choudhury, A.R. Isolation, purification, and characterization of a novel exopolysaccharide isolated from marine bacteria Brevibacillus borstelensis M42. Arch. Microbiol. 2022, 204, 399. [Google Scholar] [CrossRef] [PubMed]
  24. Li, H.; Huang, L.; Zhang, Y.; Yan, Y. Production, Characterization and Immunomodulatory Activity of an Extracellular Polysaccharide from Rhodotorula mucilaginosa YL-1 Isolated from Sea Salt Field. Mar. Drugs 2020, 18, 595. [Google Scholar] [CrossRef] [PubMed]
  25. Shyam, K.P.; Rajkumar, P.; Ramya, V.; Sivabalan, S.; Kings, A.J.; Miriam, L.M. Exopolysaccharide production by optimized medium using novel marine Enterobacter cloacae MBB8 isolate and its antioxidant potential. Carbohydr. Polym. Technol. Appl. 2021, 2, 100070. [Google Scholar] [CrossRef]
  26. Wang, C.; Mao, W.; Chen, Z.; Zhu, W.; Chen, Y.; Zhao, C.; Li, N.; Yan, M.; Liu, X.; Guo, T. Purification, structural characterization and antioxidant property of an extracellular polysaccharide from Aspergillus terreus. Process. Biochem. 2013, 48, 1395–1401. [Google Scholar] [CrossRef]
  27. Chen, Y.; Mao, W.-J.; Yan, M.-X.; Liu, X.; Wang, S.-Y.; Xia, Z.; Xiao, B.; Cao, S.-J.; Yang, B.-Q.; Li, J. Purification, Chemical Characterization, and Bioactivity of an Extracellular Polysaccharide Produced by the Marine Sponge Endogenous Fungus Alternaria sp. SP-32. Mar. Biotechnol. 2016, 18, 301–313. [Google Scholar] [CrossRef]
  28. Alshawwa, S.Z.; Alshallash, K.S.; Ghareeb, A.; Elazzazy, A.M.; Sharaf, M.; Alharthi, A.; Abdelgawad, F.E.; El-Hossary, D.; Jaremko, M.; Emwas, A.-H.; et al. Assessment of Pharmacological Potential of Novel Exopolysaccharide Isolated from Marine Kocuria sp. Strain AG5: Broad-Spectrum Biological Investigations. Life 2022, 12, 1387. [Google Scholar] [CrossRef]
  29. Asker, M.S.; El Sayed, O.H.; Mahmoud, M.G.; Yahya, S.M.; Mohamed, S.S.; Selim, M.S.; El Awady, M.S.; Abdelnasser, S.M.; Elsoud, M.M.A. Production of exopolysaccharides from novel marine bacteria and anticancer activity against hepatocellular carcinoma cells (HepG2). Bull. Natl. Res. Cent. 2018, 42, 30. [Google Scholar] [CrossRef]
  30. Wang, Y.; Liu, G.; Liu, R.; Wei, M.; Zhang, J.; Sun, C. EPS364, a Novel Deep-Sea Bacterial Exopolysaccharide, Inhibits Liver Cancer Cell Growth and Adhesion. Mar. Drugs 2021, 19, 171. [Google Scholar] [CrossRef]
  31. Liu, G.; Liu, R.; Shan, Y.; Sun, C. Marine bacterial exopolysaccharide EPS11 inhibits migration and invasion of liver cancer cells by directly targeting collagen I. J. Biol. Chem. 2021, 297, 101133. [Google Scholar] [CrossRef]
  32. Cao, R.; Jin, W.; Shan, Y.; Wang, J.; Liu, G.; Kuang, S.; Sun, C. Marine Bacterial Polysaccharide EPS11 Inhibits Cancer Cell Growth via Blocking Cell Adhesion and Stimulating Anoikis. Mar. Drugs 2018, 16, 85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wang, J.; Liu, G.; Ma, W.; Lu, Z.; Sun, C. Marine Bacterial Polysaccharide EPS11 Inhibits Cancer Cell Growth and Metastasis via Blocking Cell Adhesion and Attenuating Filiform Structure Formation. Mar. Drugs 2019, 17, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Sran, K.S.; Bisht, B.; Mayilraj, S.; Choudhury, A.R. Structural characterization and antioxidant potential of a novel anionic exopolysaccharide produced by marine Microbacterium aurantiacum FSW-25. Int. J. Biol. Macromol. 2019, 131, 343–352. [Google Scholar] [CrossRef] [PubMed]
  35. Sran, K.S.; Sundharam, S.S.; Krishnamurthi, S.; Choudhury, A.R. Production, characterization and bio-emulsifying activity of a novel thermostable exopolysaccharide produced by a marine strain of Rhodobacter johrii CDR-SL 7Cii. Int. J. Biol. Macromol. 2019, 127, 240–249. [Google Scholar] [CrossRef]
  36. Liu, S.-B.; Chen, X.-L.; He, H.-L.; Zhang, X.-Y.; Xie, B.-B.; Yu, Y.; Chen, B.; Zhou, B.-C.; Zhang, Y.-Z. Structure and Ecological Roles of a Novel Exopolysaccharide from the Arctic Sea Ice Bacterium Pseudoalteromonas sp. Strain SM20310. Appl. Environ. Microbiol. 2013, 79, 224–230. [Google Scholar] [CrossRef] [Green Version]
  37. Abdel-Wahab, B.A.; El-Kareem, H.F.A.; Alzamami, A.; Fahmy, C.A.; Elesawy, B.H.; Mahmoud, M.M.; Ghareeb, A.; El Askary, A.; Nahas, H.H.A.; Attallah, N.G.M.; et al. Novel Exopolysaccharide from Marine Bacillus subtilis with Broad Potential Biological Activities: Insights into Antioxidant, Anti-Inflammatory, Cytotoxicity, and Anti-Alzheimer Activity. Metabolites 2022, 12, 715. [Google Scholar] [CrossRef]
  38. Zhang, Z.; Cai, R.; Zhang, W.; Fu, Y.; Jiao, N. A Novel Exopolysaccharide with Metal Adsorption Capacity Produced by a Marine Bacterium Alteromonas sp. JL2810. Mar. Drugs 2017, 15, 175. [Google Scholar] [CrossRef] [Green Version]
  39. Roychowdhury, R.; Srivastava, N.; Kumari, S.; Pinnaka, A.K.; Choudhury, A.R. Isolation of an exopolysaccharide from a novel marine bacterium Neorhizobium urealyticum sp. nov. and its utilization in nanoemulsion formation for encapsulation and stabilization of astaxanthin. LWT 2021, 151, 112105. [Google Scholar] [CrossRef]
  40. Spanò, A.; Gugliandolo, C.; Lentini, V.; Maugeri, T.L.; Anzelmo, G.; Poli, A.; Nicolaus, B. A Novel EPS-Producing Strain of Bacillus licheniformis Isolated from a Shallow Vent Off Panarea Island (Italy). Curr. Microbiol. 2013, 67, 21–29. [Google Scholar] [CrossRef]
  41. Gugliandolo, C.; Spanò, A.; Lentini, V.; Arena, A.; Maugeri, T.L. Antiviral and immunomodulatory effects of a novel bacterial exopolysaccharide of shallow marine vent origin. J. Appl. Microbiol. 2014, 116, 1028–1034. [Google Scholar] [CrossRef] [PubMed]
  42. Sahana, T.; Rekha, P. A novel exopolysaccharide from marine bacterium Pantoea sp. YU16-S3 accelerates cutaneous wound healing through Wnt/β-catenin pathway. Carbohydr. Polym. 2020, 238, 116191. [Google Scholar] [CrossRef] [PubMed]
  43. Wei, M.; Geng, L.; Wang, Q.; Yue, Y.; Wang, J.; Wu, N.; Wang, X.; Sun, C.; Zhang, Q. Purification, characterization and immunostimulatory activity of a novel exopolysaccharide from Bacillus sp. H5. Int. J. Biol. Macromol. 2021, 189, 649–656. [Google Scholar] [CrossRef] [PubMed]
  44. Singh, S.; Sran, K.S.; Pinnaka, A.K.; Choudhury, A.R. Purification, characterization and functional properties of exopolysaccharide from a novel halophilic Natronotalea sambharensis sp. nov. Int. J. Biol. Macromol. 2019, 136, 547–558. [Google Scholar] [CrossRef]
  45. Guo, S.; Mao, W.; Yan, M.; Zhao, C.; Li, N.; Shan, J.; Lin, C.; Liu, X.; Guo, T.; Guo, T.; et al. Galactomannan with novel structure produced by the coral endophytic fungus Aspergillus ochraceus. Carbohydr. Polym. 2014, 105, 325–333. [Google Scholar] [CrossRef]
  46. Wu, K.; Li, Y.; Lin, Y.; Xu, B.; Yang, J.; Mo, L.; Huang, R.; Zhang, X. Structural characterization and immunomodulatory activity of an exopolysaccharide from marine-derived Aspergillus versicolor SCAU141. Int. J. Biol. Macromol. 2023, 227, 329–339. [Google Scholar] [CrossRef]
  47. Shao, Z.; Tian, Y.; Liu, S.; Chu, X.; Mao, W. Anti-Diabetic Activity of a Novel Exopolysaccharide Produced by the Mangrove Endophytic Fungus Penicillium janthinellum N29. Mar. Drugs 2023, 21, 270. [Google Scholar] [CrossRef]
  48. Roca, C.; Lehmann, M.; Torres, C.A.; Baptista, S.; Gaudêncio, S.P.; Freitas, F.; Reis, M.A. Exopolysaccharide production by a marine Pseudoalteromonas sp. strain isolated from Madeira Archipelago ocean sediments. New Biotechnol. 2016, 33, 460–466. [Google Scholar] [CrossRef]
  49. Poli, A.; Finore, I.; Romano, I.; Gioiello, A.; Lama, L.; Nicolaus, B. Microbial Diversity in Extreme Marine Habitats and Their Biomolecules. Microorganisms 2017, 5, 25. [Google Scholar] [CrossRef] [Green Version]
  50. Laurienzo, P. Marine Polysaccharides in Pharmaceutical Applications: An Overview. Mar. Drugs 2010, 8, 2435–2465. [Google Scholar] [CrossRef] [Green Version]
  51. López-Ortega, M.A.; Chavarría-Hernández, N.; del Rocio Lopez-Cuellar, M.; Rodríguez-Hernández, A.I. A review of extracellular polysaccharides from extreme niches: An emerging natural source for the biotechnology. From the adverse to diverse! Int. J. Biol. Macromol. 2021, 177, 559–577. [Google Scholar] [CrossRef] [PubMed]
  52. Sun, M.-L.; Zhao, F.; Shi, M.; Zhang, X.-Y.; Zhou, B.-C.; Zhang, Y.-Z.; Chen, X.-L. Characterization and Biotechnological Potential Analysis of a New Exopolysaccharide from the Arctic Marine Bacterium Polaribacter sp. SM1127. Sci. Rep. 2015, 5, 18435. [Google Scholar] [CrossRef] [Green Version]
  53. Courtois, A.; Berthou, C.; Guézennec, J.; Boisset, C.; Bordron, A. Exopolysaccharides Isolated from Hydrothermal Vent Bacteria Can Modulate the Complement System. PLoS ONE 2014, 9, e94965. [Google Scholar] [CrossRef]
  54. Carrión, O.; Delgado, L.; Mercade, E. New emulsifying and cryoprotective exopolysaccharide from Antarctic Pseudomonas sp. ID1. Carbohydr. Polym. 2015, 117, 1028–1034. [Google Scholar] [CrossRef] [PubMed]
  55. Sun, M.-L.; Liu, S.-B.; Qiao, L.-P.; Chen, X.-L.; Pang, X.; Shi, M.; Zhang, X.-Y.; Qin, Q.-L.; Zhou, B.-C.; Zhang, Y.-Z.; et al. A novel exopolysaccharide from deep-sea bacterium Zunongwangia profunda SM-A87: Low-cost fermentation, moisture retention, and antioxidant activities. Appl. Microbiol. Biotechnol. 2014, 98, 7437–7445. [Google Scholar] [CrossRef] [PubMed]
  56. Chikkanna, A.; Ghosh, D.; Kishore, A. Expression and characterization of a potential exopolysaccharide from a newly isolated halophilic thermotolerant bacteria Halomonas nitroreducensstrain WB1. PeerJ 2018, 6, e4684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Nadzir, M.M.; Nurhayati, R.W.; Idris, F.N.; Nguyen, M.H. Biomedical Applications of Bacterial Exopolysaccharides: A Review. Polymers 2021, 13, 530. [Google Scholar] [CrossRef]
  58. Li, N.; Wang, C.; Georgiev, M.I.; Bajpai, V.K.; Tundis, R.; Simal-Gandara, J.; Lu, X.; Xiao, J.; Tang, X.; Qiao, X. Advances in dietary polysaccharides as anticancer agents: Structure-activity relationship. Trends Food Sci. Technol. 2021, 111, 360–377. [Google Scholar] [CrossRef]
  59. Selim, M.S.; Amer, S.K.; Mohamed, S.S.; Mounier, M.M.; Rifaat, H.M. Production and characterisation of exopolysaccharide from Streptomyces carpaticus isolated from marine sediments in Egypt and its effect on breast and colon cell lines. J. Genet. Eng. Biotechnol. 2017, 16, 23–28. [Google Scholar] [CrossRef]
  60. Martos-Benítez, F.D.; Gutiérrez-Noyola, A.; Echevarría-Víctores, A. Postoperative complications and clinical outcomes among patients undergoing thoracic and gastrointestinal cancer surgery: A prospective cohort study. Rev. Bras. Ter. Intensiv. 2016, 28, 40–48. [Google Scholar] [CrossRef]
  61. Boogaard, W.M.C.V.D.; Komninos, D.S.J.; Vermeij, W.P. Chemotherapy Side-Effects: Not All DNA Damage Is Equal. Cancers 2022, 14, 627. [Google Scholar] [CrossRef] [PubMed]
  62. Do, T.T.H.; Lai, T.N.B.; Stephenson, S.L.; Tran, H.T.M. Cytotoxicity activities and chemical characteristics of exopolysaccharides and intracellular polysaccharides of Physarum polycephalum microplasmodia. BMC Biotechnol. 2021, 21, 28. [Google Scholar] [CrossRef] [PubMed]
  63. Abdelnasser, S.M.; Yahya, S.M.M.; Mohamed, W.F.; Asker, M.M.S.; Abu Shady, H.M.; Mahmoud, M.G.; Gadallah, M.A. Antitumor Exopolysaccharides Derived from Novel Marine Bacillus: Isolation, Characterization Aspect and Biological Activity. Asian Pac. J. Cancer Prev. 2017, 18, 1847–1854. [Google Scholar] [CrossRef]
  64. Amer, M.S.; Zaghloul, E.H.; Ibrahim, M.I. Characterization of exopolysaccharide produced from marine-derived Aspergillus terreus SEI with prominent biological activities. Egypt. J. Aquat. Res. 2020, 46, 363–369. [Google Scholar] [CrossRef]
  65. Mahgoub, A.M.; Mahmoud, M.G.; Selim, M.S.; EL Awady, M.E. Exopolysaccharide from Marine Bacillus velezensis MHM3 Induces Apoptosis of Human Breast Cancer MCF-7 Cells through a Mitochondrial Pathway. Asian Pac. J. Cancer Prev. 2018, 19, 1957–1963. [Google Scholar] [CrossRef] [PubMed]
  66. Ibrahim, A.Y.; Youness, E.R.; Mahmoud, M.G.; Asker, M.S.; El-Newary, S.A. Acidic Exopolysaccharide Produced from Marine Bacillus amyloliquefaciens 3MS 2017 for the Protection and Treatment of Breast Cancer. Breast Cancer Basic Clin. Res. 2020, 14, 1178223420902075. [Google Scholar] [CrossRef] [Green Version]
  67. Tukenmez, U.; Aktas, B.; Aslim, B.; Yavuz, S. The relationship between the structural characteristics of lactobacilli-EPS and its ability to induce apoptosis in colon cancer cells in vitro. Sci. Rep. 2019, 9, 8268. [Google Scholar] [CrossRef] [Green Version]
  68. Nussinov, R.; Tsai, C.-J.; Jang, H. Anticancer drug resistance: An update and perspective. Drug Resist. Updat. 2021, 59, 100796. [Google Scholar] [CrossRef]
  69. Terreni, M.; Taccani, M.; Pregnolato, M. New Antibiotics for Multidrug-Resistant Bacterial Strains: Latest Research Developments and Future Perspectives. Molecules 2021, 26, 2671. [Google Scholar] [CrossRef]
  70. Miethke, M.; Pieroni, M.; Weber, T.; Brönstrup, M.; Hammann, P.; Halby, L.; Arimondo, P.B.; Glaser, P.; Aigle, B.; Bode, H.B.; et al. Towards the sustainable discovery and development of new antibiotics. Nat. Rev. Chem. 2021, 5, 726–749. [Google Scholar] [CrossRef]
  71. Aullybux, A.A.; Puchooa, D.; Bahorun, T.; Jeewon, R. Phylogenetics and antibacterial properties of exopolysaccharides from marine bacteria isolated from Mauritius seawater. Ann. Microbiol. 2019, 69, 957–972. [Google Scholar] [CrossRef]
  72. Arayes, M.A.; Mabrouk, M.E.M.; Sabry, S.A.; Abdella, B. Exopolysaccharide production from Alkalibacillus sp. w3: Statistical optimization and biological activity. Biologia 2022, 78, 229–240. [Google Scholar] [CrossRef]
  73. Zhou, Y.; Cui, Y.; Qu, X. Exopolysaccharides of lactic acid bacteria: Structure, bioactivity and associations: A review. Carbohydr. Polym. 2018, 207, 317–332. [Google Scholar] [CrossRef] [PubMed]
  74. Rajoka, M.S.R.; Mehwish, H.M.; Hayat, H.F.; Hussain, N.; Sarwar, S.; Aslam, H.; Nadeem, A.; Shi, J. Characterization, the Antioxidant and Antimicrobial Activity of Exopolysaccharide Isolated from Poultry Origin Lactobacilli. Probiotics Antimicrob. Proteins 2018, 11, 1132–1142. [Google Scholar] [CrossRef]
  75. Abdalla, A.K.; Ayyash, M.M.; Olaimat, A.N.; Osaili, T.M.; Al-Nabulsi, A.A.; Shah, N.P.; Holley, R. Exopolysaccharides as Antimicrobial Agents: Mechanism and Spectrum of Activity. Front. Microbiol. 2021, 12, 664395. [Google Scholar] [CrossRef]
  76. Andrew, M.; Jayaraman, G. Structural features of microbial exopolysaccharides in relation to their antioxidant activity. Carbohydr. Res. 2019, 487, 107881. [Google Scholar] [CrossRef] [PubMed]
  77. Selim, M.S.; Mohamed, S.S.; Mahmoud, M.G.; Ibrahim, A.Y.; A Ghazy, E. Production, characterization, and antioxidant activities of bacterial exopolysaccharides extracted from petroleum oil water. Egypt. Pharm. J. 2019, 18, 42. [Google Scholar] [CrossRef]
  78. Martemucci, G.; Costagliola, C.; Mariano, M.; D’andrea, L.; Napolitano, P.; D’alessandro, A.G. Free Radical Properties, Source and Targets, Antioxidant Consumption and Health. Oxygen 2022, 2, 48–78. [Google Scholar] [CrossRef]
  79. Adwas, A.A.; Elsayed, A.; Azab, A.E.; Quwaydir, F.A. Oxidative stress and antioxidant mechanisms in human body. J. Appl. Biotechnol. Bioeng. 2019, 6, 43–47. [Google Scholar] [CrossRef]
  80. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxid. Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef] [Green Version]
  81. Sharifi-Rad, M.; Anil Kumar, N.V.; Zucca, P.; Varoni, E.M.; Dini, L.; Panzarini, E.; Rajkovic, J.; Tsouh Fokou, P.V.; Azzini, E.; Peluso, I.; et al. Lifestyle, Oxidative Stress, and Antioxidants: Back and Forth in the Pathophysiology of Chronic Diseases. Front. Physiol. 2020, 11, 694. [Google Scholar] [CrossRef] [PubMed]
  82. Sarangarajan, R.; Meera, S.; Rukkumani, R.; Sankar, P.; Anuradha, G. Antioxidants: Friend or foe? Asian Pac. J. Trop. Med. 2017, 10, 1111–1116. [Google Scholar] [CrossRef]
  83. Yan, M.-X.; Mao, W.-J.; Liu, X.; Wang, S.-Y.; Xia, Z.; Cao, S.-J.; Li, J.; Qin, L.; Xian, H.-L. Extracellular polysaccharide with novel structure and antioxidant property produced by the deep-sea fungus Aspergillus versicolor N2bc. Carbohydr. Polym. 2016, 147, 272–281. [Google Scholar] [CrossRef] [PubMed]
  84. Prathyusha, A.; Sheela, G.M.; Bramhachari, P. Chemical characterization and antioxidant properties of exopolysaccharides from mangrove filamentous fungi Fusarium equiseti ANP2. Biotechnol. Rep. 2018, 19, e00277. [Google Scholar] [CrossRef] [PubMed]
  85. Meulmeester, F.L.; Luo, J.; Martens, L.G.; Mills, K.; van Heemst, D.; Noordam, R. Antioxidant Supplementation in Oxidative Stress-Related Diseases: What Have We Learned from Studies on Alpha-Tocopherol? Antioxidants 2022, 11, 2322. [Google Scholar] [CrossRef]
  86. Wang, Z.; Zhao, Y.; Jiang, Y.; Chu, W. Prebiotic, Antioxidant, and Immunomodulatory Properties of Acidic Exopolysaccharide from Marine Rhodotorula RY1801. Front. Nutr. 2021, 8, 710668. [Google Scholar] [CrossRef] [PubMed]
  87. Ayyash, M.; Abu-Jdayil, B.; Olaimat, A.; Esposito, G.; Itsaranuwat, P.; Osaili, T.; Obaid, R.S.; Kizhakkayil, J.; Liu, S.-Q. Physicochemical, bioactive and rheological properties of an exopolysaccharide produced by a probiotic Pediococcus pentosaceus M41. Carbohydr. Polym. 2020, 229, 115462. [Google Scholar] [CrossRef]
  88. Li, C.; Wang, J.; Wang, Y.; Gao, H.; Wei, G.; Huang, Y.; Yu, H.; Gan, Y.; Wang, Y.; Mei, L. Recent progress in drug delivery. Acta Pharm. Sin. B 2019, 9, 1145–1162. [Google Scholar] [CrossRef]
  89. Adepu, S.; Ramakrishna, S. Controlled Drug Delivery Systems: Current Status and Future Directions. Molecules 2021, 26, 5905. [Google Scholar] [CrossRef]
  90. Tabernero, A.; Cardea, S. Microbial Exopolysaccharides as Drug Carriers. Polymers 2020, 12, 2142. [Google Scholar] [CrossRef]
  91. Laubach, J.; Joseph, M.; Brenza, T.; Gadhamshetty, V.; Sani, R.K. Exopolysaccharide and biopolymer-derived films as tools for transdermal drug delivery. J. Control. Release 2021, 329, 971–987. [Google Scholar] [CrossRef] [PubMed]
  92. Díaz-Montes, E. Dextran: Sources, Structures, and Properties. Polysaccharides 2021, 2, 554–565. [Google Scholar] [CrossRef]
  93. Huang, G.; Huang, H. Application of dextran as nanoscale drug carriers. Nanomedicine 2018, 13, 3149–3158. [Google Scholar] [CrossRef]
  94. Wang, H.; Dai, T.; Zhou, S.; Huang, X.; Li, S.; Sun, K.; Zhou, G.; Dou, H. Self-Assembly Assisted Fabrication of Dextran-Based Nanohydrogels with Reduction-Cleavable Junctions for Applications as Efficient Drug Delivery Systems. Sci. Rep. 2017, 7, 40011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Fang, Y.; Wang, H.; Dou, H.-J.; Fan, X.; Fei, X.-C.; Wang, L.; Cheng, S.; Janin, A.; Zhao, W.L. Doxorubicin-loaded dextran-based nano-carriers for highly efficient inhibition of lymphoma cell growth and synchronous reduction of cardiac toxicity. Int. J. Nanomed. 2018, 13, 5673–5683. [Google Scholar] [CrossRef] [Green Version]
  96. de Siqueira, E.C.; Rebouças, J.d.S.; Pinheiro, I.O.; Formiga, F.R. Levan-based nanostructured systems: An overview. Int. J. Pharm. 2020, 580, 119242. [Google Scholar] [CrossRef]
  97. Garg, U.; Chauhan, S.; Nagaich, U.; Jain, N. Current Advances in Chitosan Nanoparticles Based Drug Delivery and Targeting. Adv. Pharm. Bull. 2019, 9, 195–204. [Google Scholar] [CrossRef] [Green Version]
  98. Sultan, M.H.; Moni, S.S.; Madkhali, O.A.; Bakkari, M.A.; Alshahrani, S.; Alqahtani, S.S.; Alhakamy, N.A.; Mohan, S.; Ghazwani, M.; Bukhary, H.A.; et al. Characterization of cisplatin-loaded chitosan nanoparticles and rituximab-linked surfaces as target-specific injectable nano-formulations for combating cancer. Sci. Rep. 2022, 12, 468. [Google Scholar] [CrossRef]
  99. Cinan, E.; Cesur, S.; Haskoylu, M.E.; Gunduz, O.; Oner, E.T. Resveratrol-Loaded Levan Nanoparticles Produced by Electrohydrodynamic Atomization Technique. Nanomaterials 2021, 11, 2582. [Google Scholar] [CrossRef]
  100. Sun, L.; Chen, Y.; Zhou, Y.; Guo, D.; Fan, Y.; Guo, F.; Zheng, Y.; Chen, W. Preparation of 5-fluorouracil-loaded chitosan nanoparticles and study of the sustained release in vitro and in vivo. Asian J. Pharm. Sci. 2017, 12, 418–423. [Google Scholar] [CrossRef]
  101. Amer, M.S.; Ibrahim, H.A.H. Chitosan from marine-derived Penicillum spinulosum MH2 cell wall with special emphasis on its antimicrobial and antifouling properties. Egypt. J. Aquat. Res. 2019, 45, 359–365. [Google Scholar] [CrossRef]
  102. Poli, A.; Kazak, H.; Gürleyendağ, B.; Tommonaro, G.; Pieretti, G.; Öner, E.T.; Nicolaus, B. High level synthesis of levan by a novel Halomonas species growing on defined media. Carbohydr. Polym. 2009, 78, 651–657. [Google Scholar] [CrossRef]
  103. Zykwinska, A.; Marquis, M.; Sinquin, C.; Cuenot, S.; Colliec-Jouault, S. Assembly of HE800 exopolysaccharide produced by a deep-sea hydrothermal bacterium into microgels for protein delivery applications. Carbohydr. Polym. 2016, 142, 213–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Zykwinska, A.; Marquis, M.; Godin, M.; Marchand, L.; Sinquin, C.; Garnier, C.; Jonchère, C.; Chédeville, C.; Le Visage, C.; Guicheux, J.; et al. Microcarriers Based on Glycosaminoglycan-Like Marine Exopolysaccharide for TGF-β1 Long-Term Protection. Mar. Drugs 2019, 17, 65. [Google Scholar] [CrossRef] [Green Version]
  105. Briffa, J.; Sinagra, E.; Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 2020, 6, e04691. [Google Scholar] [CrossRef] [PubMed]
  106. Mathivanan, K.; Chandirika, J.U.; Mathimani, T.; Rajaram, R.; Annadurai, G.; Yin, H. Production and functionality of exopolysaccharides in bacteria exposed to a toxic metal environment. Ecotoxicol. Environ. Saf. 2020, 208, 111567. [Google Scholar] [CrossRef] [PubMed]
  107. Velkova, Z.; Kirova, G.; Stoytcheva, M.; Kostadinova, S.; Todorova, K.; Gochev, V. Immobilized microbial biosorbents for heavy metals removal. Eng. Life Sci. 2018, 18, 871–881. [Google Scholar] [CrossRef] [Green Version]
  108. Mohapatra, R.K.; Parhi, P.K.; Patra, J.K.; Panda, C.R.; Thatoi, H.N. Biodetoxification of Toxic Heavy Metals by Marine Metal Resistant Bacteria—A Novel Approach for Bioremediation of the Polluted Saline Environment. In Microbial Biotechnology; Patra, J., Vishnuprasad, C., Das, G., Eds.; Springer: Singapore, 2017; pp. 343–376. [Google Scholar] [CrossRef]
  109. Nanda, M.; Kumar, V.; Sharma, D. Multimetal tolerance mechanisms in bacteria: The resistance strategies acquired by bacteria that can be exploited to ‘clean-up’ heavy metal contaminants from water. Aquat. Toxicol. 2019, 212, 1–10. [Google Scholar] [CrossRef]
  110. Gupta, P.; Diwan, B. Bacterial Exopolysaccharide mediated heavy metal removal: A Review on biosynthesis, mechanism and remediation strategies. Biotechnol. Rep. 2016, 13, 58–71. [Google Scholar] [CrossRef]
  111. Concórdio-Reis, P.; Reis, M.A.M.; Freitas, F. Biosorption of Heavy Metals by the Bacterial Exopolysaccharide FucoPol. Appl. Sci. 2020, 10, 6708. [Google Scholar] [CrossRef]
  112. Ntozonke, N.; Okaiyeto, K.; Okoli, A.S.; Olaniran, A.O.; Nwodo, U.U.; Okoh, A.I. A Marine Bacterium, Bacillus sp. Isolated from the Sediment Samples of Algoa Bay in South Africa Produces a Polysaccharide-Bioflocculant. Int. J. Environ. Res. Public Health 2017, 14, 1149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Chen, Z.; Li, Z.; Liu, P.; Liu, Y.; Wang, Y.; Li, Q.; Zhen, C. Characterization of a novel bioflocculant from a marine bacterium and its application in dye wastewater treatment. BMC Biotechnol. 2017, 17, 84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Chaisuwan, W.; Jantanasakulwong, K.; Wangtueai, S.; Phimolsiripol, Y.; Chaiyaso, T.; Techapun, C.; Phongthai, S.; You, S.; Regenstein, J.M.; Seesuriyachan, P. Microbial exopolysaccharides for immune enhancement: Fermentation, modifications and bioactivities. Food Biosci. 2020, 35, 100564. [Google Scholar] [CrossRef]
  115. Chen, Y.-C.; Wu, Y.-J.; Hu, C.-Y. Monosaccharide composition influence and immunomodulatory effects of probiotic exopolysaccharides. Int. J. Biol. Macromol. 2019, 133, 575–582. [Google Scholar] [CrossRef]
  116. Rajoka, M.S.R.; Wu, Y.; Mehwish, H.M.; Bansal, M.; Zhao, L. Lactobacillus exopolysaccharides: New perspectives on engineering strategies, physiochemical functions, and immunomodulatory effects on host health. Trends Food Sci. Technol. 2020, 103, 36–48. [Google Scholar] [CrossRef]
  117. Verkhnyatskaya, S.; Ferrari, M.; de Vos, P.; Walvoort, M.T.C. Shaping the Infant Microbiome with Non-digestible Carbohydrates. Front. Microbiol. 2019, 10, 343. [Google Scholar] [CrossRef] [Green Version]
  118. Wang, K.; Li, W.; Rui, X.; Li, T.; Chen, X.; Jiang, M.; Dong, M. Chemical modification, characterization and bioactivity of a released exopolysaccharide (r-EPS1) from Lactobacillus plantarum. Glycoconj. J. 2014, 32, 17–27. [Google Scholar] [CrossRef]
  119. Zhang, Z.; Liu, Z.; Tao, X.; Wei, H. Characterization and sulfated modification of an exopolysaccharide from Lactobacillus plantarum ZDY2013 and its biological activities. Carbohydr. Polym. 2016, 153, 25–33. [Google Scholar] [CrossRef]
  120. Senni, K.; Gueniche, F.; Changotade, S.; Septier, D.; Sinquin, C.; Ratiskol, J.; Lutomski, D.; Godeau, G.; Guezennec, J.; Colliec-Jouault, S. Unusual Glycosaminoglycans from a Deep Sea Hydrothermal Bacterium Improve Fibrillar Collagen Structuring and Fibroblast Activities in Engineered Connective Tissues. Mar. Drugs 2013, 11, 1351–1369. [Google Scholar] [CrossRef] [Green Version]
  121. Di Donato, P.; Poli, A.; Taurisano, V.; Abbamondi, G.R.; Nicolaus, B.; Tommonaro, G. Recent Advances in the Study of Marine Microbial Biofilm: From the Involvement of Quorum Sensing in Its Production up to Biotechnological Application of the Polysaccharide Fractions. J. Mar. Sci. Eng. 2016, 4, 34. [Google Scholar] [CrossRef] [Green Version]
  122. García, A.; Fernández-Sandoval, M.T.; Morales-Guzmán, D.; Martínez-Morales, F.; Trejo-Hernández, M.R. Advances in exopolysaccharide production from marine bacteria. J. Chem. Technol. Biotechnol. 2022, 97, 2694–2705. [Google Scholar] [CrossRef]
  123. Hereher, F.; ElFallal, A.; Abou-Dobara, M.; Toson, E.; Abdelaziz, M.M. Cultural optimization of a new exopolysaccharide producer Micrococcus roseus. Beni-Suef Univ. J. Basic Appl. Sci. 2018, 7, 632–639. [Google Scholar] [CrossRef]
  124. Biswas, J.; Paul, A.K. Optimization of factors influencing exopolysaccharide production by Halomonas xianhensis SUR308 under batch culture. AIMS Microbiol. 2017, 3, 564–579. [Google Scholar] [CrossRef]
  125. Nguyen, P.-T.; Nguyen, T.-T.; Bui, D.-C.; Hong, P.-T.; Hoang, Q.-K.; Nguyen, H.-T. Exopolysaccharide production by lactic acid bacteria: The manipulation of environmental stresses for industrial applications. AIMS Microbiol. 2020, 6, 451–469. [Google Scholar] [CrossRef] [PubMed]
  126. Gudiña, E.J.; Couto, M.R.; Silva, S.P.; Coelho, E.; Coimbra, M.A.; Teixeira, J.A.; Rodrigues, L.R. Sustainable Exopolysaccharide Production by Rhizobium viscosum CECT908 Using Corn Steep Liquor and Sugarcane Molasses as Sole Substrates. Polymers 2022, 15, 20. [Google Scholar] [CrossRef]
  127. Delbarre-Ladrat, C.; Sinquin, C.; Lebellenger, L.; Zykwinska, A.; Colliec-Jouault, S. Exopolysaccharides produced by marine bacteria and their applications as glycosaminoglycan-like molecules. Front. Chem. 2014, 2, 85. [Google Scholar] [CrossRef] [Green Version]
  128. Ma, Z.-C.; Liu, N.-N.; Chi, Z.; Liu, G.-L.; Chi, Z.-M. Genetic Modification of the Marine-Isolated Yeast Aureobasidium melanogenum P16 for Efficient Pullulan Production from Inulin. Mar. Biotechnol. 2015, 17, 511–522. [Google Scholar] [CrossRef]
  129. Sun, X.; Zhang, J. Bacterial exopolysaccharides: Chemical structures, gene clusters and genetic engineering. Int. J. Biol. Macromol. 2021, 173, 481–490. [Google Scholar] [CrossRef]
  130. Li, S.; Xiong, Q.; Lai, X.; Li, X.; Wan, M.; Zhang, J.; Yan, Y.; Cao, M.; Lu, L.; Guan, J.; et al. Molecular Modification of Polysaccharides and Resulting Bioactivities. Compr. Rev. Food Sci. Food Saf. 2015, 15, 237–250. [Google Scholar] [CrossRef]
  131. Wang, J.; Salem, D.R.; Sani, R.K. Extremophilic exopolysaccharides: A review and new perspectives on engineering strategies and applications. Carbohydr. Polym. 2018, 205, 8–26. [Google Scholar] [CrossRef]
  132. Chopin, N.; Sinquin, C.; Ratiskol, J.; Zykwinska, A.; Weiss, P.; Cérantola, S.; Le Bideau, J.; Colliec-Jouault, S. A Direct Sulfation Process of a Marine Polysaccharide in Ionic Liquid. BioMed Res. Int. 2015, 2015, 508656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Xia, S.; Zhai, Y.; Wang, X.; Fan, Q.; Dong, X.; Chen, M.; Han, T. Phosphorylation of polysaccharides: A review on the synthesis and bioactivities. Int. J. Biol. Macromol. 2021, 184, 946–954. [Google Scholar] [CrossRef] [PubMed]
  134. Vitorino, L.C.; Bessa, L.A. Microbial Diversity: The Gap between the Estimated and the Known. Diversity 2018, 10, 46. [Google Scholar] [CrossRef] [Green Version]
Table 1. Examples of novel polysaccharides- and/or novel exopolysaccharides-producing marine bacteria and fungi, reported over the past decade, that highlight the large diversity of these polymers in terms of size, composition, and biological activities. (* tr—trace amounts; N/A—not available)—Note: when not available, figures of polysaccharide structures were inferred based on information provided in the relevant studies and should not be taken as representing the actual or most stable conformations of the polymers.
Table 1. Examples of novel polysaccharides- and/or novel exopolysaccharides-producing marine bacteria and fungi, reported over the past decade, that highlight the large diversity of these polymers in terms of size, composition, and biological activities. (* tr—trace amounts; N/A—not available)—Note: when not available, figures of polysaccharide structures were inferred based on information provided in the relevant studies and should not be taken as representing the actual or most stable conformations of the polymers.
Organism NameCompositionMolecular Weight (kDa)Possible Structure(s)Biological Activity (If Any)Level of CharacterizationReference
Vibrio alginolyticusMannose, glucosamine, gluconic acid, galactosamine and arabinose (5:9:3.4:0.5:0.8)14.8N/AAntitumor activityMolecular weight, monosaccharide composition, functional groups, surface morphology, and element composition[30]
Bacillus sp.Mannose, glucosamine, galacturonic acid, glucose and xylose (1:2.58:0.68:0.13:3.09:1.41)22.3N/AAnticancer through different mechanismsMolecular weight and monosaccharide composition[31,32,33]
Microbacterium aurantiacumGlucose, mannose, fucose and glucuronic acid7000Marinedrugs 21 00420 i001Anti-oxidant activity and viscosifying propertyMolecular weight, total sugar, total protein and uronic acid content, functional groups, element analysis, monosaccharide composition, and partial structure[34]
Rhodobacter johriiGlucose, galactose, rhamnose and glucuronic acid (3:1.5:0.25:0.25)2000Marinedrugs 21 00420 i002Bioemulsification propertyMolecular weight, total sugar, total protein and uronic acid content, functional groups, element analysis, monosaccharide composition, and partial structure[35]
Pseudoalteromonas sp.Mannose, glucose, galactose, rhamnose, xylose, N-acetylgalactosamine and N-acetylglucosamine>2000Marinedrugs 21 00420 i003-Molecular weight, total sugar content, monosaccharide composition, linkage analysis, and structure[36]
Bacillus subtilisGlucose, rhamnose and arabinose14.8Marinedrugs 21 00420 i004Anti-oxidant, anti-inflammatory, cytotoxicity, and anti-Alzheimer activitiesMolecular weight, functional groups, uronic acid and sulfate content, monosaccharide composition, surface morphology, and crystal structure[37]
Alteromonas sp.Rhamnose, mannose and galacturonic acid167Marinedrugs 21 00420 i005Biosorption of heavy metalsMolecular weight, total sugar content, functional groups, linkage analysis and structure[38]
Neorhizobium urealyticum K1T sp. nov.Glucose and galacturonic acid207N/AEmulsificationMolecular weight, total sugar and total protein content, functional groups, element analysis, monosaccharide composition, and partial structure[39]
B. licheniformisFructose, fucose, glucose, galactosamine and mannose (1.0:0.75:0.28:tr:tr) *1000Marinedrugs 21 00420 i006Cytotoxic, antiviral and immunomodulatory propertiesMolecular weight, monosaccharide composition, polar lipids and fatty acids content, and structure[40,41]
Pantoea sp.Glucose, galactose, N-acetyl galactosamine and glucosamine (1.9:1:0.4:0.02)175N/ACutaneous wound healingMolecular weight, monosaccharide composition, total sugar and total protein contents, and functional groups[42]
Bacillus sp.Mannose, glucosamine, glucose, and galactose (1.00:0.02:0.07:0.02)89Marinedrugs 21 00420 i007ImmunomodulationMolecular weight, monosaccharide composition, linkage analysis, functional groups, and structure[43]
Natronotalea sambharensis sp. nov.Mannose, glucose and glucuronic acid4.6 × 106Marinedrugs 21 00420 i008Anti-oxidantMolecular weight, total sugar, total protein, nucleic acid and uronic acid content, element analysis, functional groups, surface morphology, monosaccharide composition, and partial structure[44]
Aspergillus ochraceusMannose and galactose (2.16:1.00)29Marinedrugs 21 00420 i009-Molecular weight, total sugar, protein and uronic acid content, monosaccharide composition, linkage analysis, sugar configuration, and structure[45]
Aspergillus versicolorGlucose5.1Marinedrugs 21 00420 i010ImmunomodulationMolecular weight, functional groups, monosaccharide composition, and linkage[46]
Penicillium janthinellumMannose and galactose10.24Marinedrugs 21 00420 i011Anti-diabetic activityMolecular weight, total sugar and total protein content, functional groups, monosaccharide composition, linkage analysis, sugar configurations, and structure[47]
Table 2. Examples of polysaccharides isolated from extremophilic microorganisms.
Table 2. Examples of polysaccharides isolated from extremophilic microorganisms.
Species NameType of EnvironmentEPS NameEPS CompositionDistinguishing FeaturesReference
Polaribacter sp.Polar region (Arctic)EPSN-acetyl glucosamine, mannose, glucuronic acid, moderate amounts of galactose and fucose, and minor amounts of glucose and rhamnoseTolerance to high salinity and a wide pH range[52]
Alteromonas infernusDeep-sea hydrothermal ventGY785Glucose, galactose, galacturonic acid and glucuronic acid-[53]
Vibrio diabolicusDeep-sea hydrothermal ventHE800N-acetyl glucosamine, N-acetyl galactosamine and glucuronic acid-[53]
Pseudomonas sp.Polar region (Antarctica)EPSGlucose, galactose, fucose, and uronic acidCryoprotection and emulsification[54]
Zunongwangia profundaDeep-sea (1245 m)EPS-High moisture retention and anti-oxidant potential[55]
Halomonas nitroreducensHydrothermal ventEPSThree different EPSs made up of glucose, mannose, galactose, and small quantities of rhamnose, arabinose, and galacturonic acid in variable amountsPseudoplastic nature with high emulsifying, anti-oxidant, and heavy metal-binding activities[56]
Table 3. Examples of patents that have been deposited for polysaccharides derived from marine bacteria or fungi during the last decade.
Table 3. Examples of patents that have been deposited for polysaccharides derived from marine bacteria or fungi during the last decade.
Patent NumberSpeciesSourceCharacteristic of PolysaccharidePatented Application
AU2016330332B2Alteromonas sp.Deep-sea hydrothermal environment15 kDa over-sulfated exopolysaccharide (GYS15)Anti-metastatic and/or related uses for various cancers
CN116120477A-Antarctic sea4350–4360 kDa extracellular polysaccharide with low temperature resistance and moisturizing functionsPreparation method and application
ES2585398B1Pseudomonas sp.Marine sediment2000 kDa exopolysaccharide with cryoprotective, emulsifying, thickening, stabilizing, or texturing propertiesCosmetic application
CN105087450AAlteromonas marina-167 kDa exopolysaccharideCulture of organism and preparation of polysaccharide
US10993434B2Pseudoalteromonas sp.Polar region100–430 kDa exopolysaccharideCryoprotection of cells
CN107523515APseudoalteromonas sp.--Absorption of heavy metals from drinking water
LU501700B1Aerococcus urinaeequi--Growth of microorganism and polysaccharide production
CN109457001A--Polysaccharide with mannose, glucosamine, ribose, rhamnose, glucuronic acid, galacturonic acid, glucose, galactolipin, xylose, and arabinosePreparation and application as decoloring agent
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jeewon, R.; Aullybux, A.A.; Puchooa, D.; Nazurally, N.; Alrefaei, A.F.; Zhang, Y. Marine Microbial Polysaccharides: An Untapped Resource for Biotechnological Applications. Mar. Drugs 2023, 21, 420. https://doi.org/10.3390/md21070420

AMA Style

Jeewon R, Aullybux AA, Puchooa D, Nazurally N, Alrefaei AF, Zhang Y. Marine Microbial Polysaccharides: An Untapped Resource for Biotechnological Applications. Marine Drugs. 2023; 21(7):420. https://doi.org/10.3390/md21070420

Chicago/Turabian Style

Jeewon, Rajesh, Aadil Ahmad Aullybux, Daneshwar Puchooa, Nadeem Nazurally, Abdulwahed Fahad Alrefaei, and Ying Zhang. 2023. "Marine Microbial Polysaccharides: An Untapped Resource for Biotechnological Applications" Marine Drugs 21, no. 7: 420. https://doi.org/10.3390/md21070420

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