Next Article in Journal
PAP8/pTAC6 Is Part of a Nuclear Protein Complex and Displays RNA Recognition Motifs of Viral Origin
Next Article in Special Issue
From Cancer Therapy to Winemaking: The Molecular Structure and Applications of β-Glucans and β-1, 3-Glucanases
Previous Article in Journal
Mesenchymal Stem-Cell Remodeling of Adsorbed Type-I Collagen—The Effect of Collagen Oxidation
Previous Article in Special Issue
The Emergence of New Catalytic Abilities in an Endoxylanase from Family GH10 by Removing an Intrinsically Disordered Region
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Discovery and Biotechnological Exploitation of Glycoside-Phosphorylases

Toulouse Biotechnology Institute (TBI), Centre National de la Recherche Scientifique (CNRS) Toulouse Biotechnology Institute (TBI), French National Research Institute for Agriculture, Food and Environment (INRAE), Institut National des Sciences Appliquées (INSA), Université de Toulouse, 31000 Toulouse, France
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(6), 3043; https://doi.org/10.3390/ijms23063043
Submission received: 11 February 2022 / Revised: 1 March 2022 / Accepted: 3 March 2022 / Published: 11 March 2022
(This article belongs to the Special Issue Biotechnological Application of Carbohydrate Active Enzymes)

Abstract

:
Among carbohydrate active enzymes, glycoside phosphorylases (GPs) are valuable catalysts for white biotechnologies, due to their exquisite capacity to efficiently re-modulate oligo- and poly-saccharides, without the need for costly activated sugars as substrates. The reversibility of the phosphorolysis reaction, indeed, makes them attractive tools for glycodiversification. However, discovery of new GP functions is hindered by the difficulty in identifying them in sequence databases, and, rather, relies on extensive and tedious biochemical characterization studies. Nevertheless, recent advances in automated tools have led to major improvements in GP mining, activity predictions, and functional screening. Implementation of GPs into innovative in vitro and in cellulo bioproduction strategies has also made substantial advances. Herein, we propose to discuss the latest developments in the strategies employed to efficiently discover GPs and make the best use of their exceptional catalytic properties for glycoside bioproduction.

1. Introduction

As numerous glycosides are used in the food, feed, cosmetics, health, and chemical industries, glycoside-synthesizing enzymes are very interesting tools to produce these compounds in vitro or in cellulo [1,2]. Indeed, while today most glycosides are chemically synthesized, biocatalysts can bypass or be used in combination with glycochemistry [3,4]. There are four types of glycoside-synthesizing enzymes.
In cellulo, glycosidic bond formation is mainly performed by glycosyl transferases (GTs), which catalyze the transfer of sugar moieties from an activated donor onto an acceptor molecule (often a carbohydrate) [5]. The advantages of GTs are the high synthesis yield they often offer [6]; while, major drawbacks are (i) the requirement for expensive donors [7], which can, nevertheless, be produced in cellulo, together with the target GTs to directly produce and excrete oligosaccharides [8,9], and (ii) the low number of stable GTs, produced as soluble forms for in vitro synthesis purposes [10].
Transglycosidases (TGs) catalyze the transfer of a glycosyl unit from a donor glycoside (often readily available and cheap, unlike nucleotide sugars) onto an acceptor molecule, to produce another type of carbohydrate or glycoconjugate, thanks to their wide acceptor promiscuity. More and more attention is being given to the discovery of new TGs [11] and their engineering [12,13,14,15], for oligosaccharide synthesis in vitro. However, only a few native TGs have been described, and their specificities are restricted to some substrates.
Glycosynthases (GSs) are engineered glycoside hydrolases (GHs), in which the substitution of the catalytic nucleophile by a non-nucleophilic residue (alanine, glycine, or serine) abolishes the innate hydrolytic activity. The role of the nucleophile is played by an activated donor, such as fluorine-containing substrates, providing the energy necessary to cross the barrier for the reaction to occur in the synthesis direction. They can, therefore, synthesize glycosidic bonds instead of hydrolyzing them. However, fluorine donor substrates are expensive or have to be prepared for each study, and suffer from instability [16,17]. Replacing them with sugar oxazoline [18] or (more toxic) glycosyl azide [19] donors could result in a substantial increase of synthesis yields [20].
Glycoside phosphorylases (GPs) catalyze, both glycoside degradation by using phosphate to breakdown osidic linkages (phosphorolysis), and synthesis by using sugar-phosphates as glycosyl donors (reverse phosphorolysis). GPs were discovered by the 1947 physiology and medicine Nobel Prize recipients Carl F. and Gerty T. Cori [21], who greatly contributed to the understanding of glycogen metabolism. Initially the term ‘phosphorylases’ was used to describe the enzyme found in the liver and skeletal muscle fibers that can produce α-d-glucose-1-phosphate (αGlc1P) from glycogen. GPs share structural and mechanistic characteristics with both glycoside hydrolases and glycosyl transferases. The two reactions of phosphorolysis and reverse phosphorolysis catalyzed by GPs oppose one another, resulting in an equilibrium. This is due to the fact that the free energy required for the cleavage of the glycosidic bond is close to the one required for cleavage of the ester linkage in glycosyl phosphates, as demonstrated by a study on cellobiose phosphorylase [22]. Herein, we propose to discuss these enzymes, with a particular focus on the latest advances in the approaches employed for their discovery and characterization. We will also emphasize the most up to date developments regarding the industrial applications of these particularly attractive biochemical tools, including in synthetic biology.

2. Glycoside-Phosphorylases

2.1. GP Catalytic Mechanism

Owing to the chimeric nature of GPs relative to GHs and GTs, they can be found in both families. A list of biochemically characterized GPs, as of November 2021, can be found in Table 1. GPs can be classified into retaining GPs or inverting GPs, based on the change in the anomeric carbon configuration after the reaction [23]. Km values span across two orders of magnitude, from 0.3 [24] to 9.6 mM [25], 1,46 [24] to 130 mM [26], and 0.42 [24] to 7.1 mM [27], for donors, acceptors and phosphate, respectively. Turnover numbers also widely vary. Kcat values can be as low as 0.8 s−1 for GH65 glucosylglycerate phosphorylase [28], up to 2866 s−1 for GH94 cellodextrin phosphorylase on cellopentaose [29].
Both inverting GHs and GPs follow a direct displacement, SN2-like reaction mechanism, (Figure 1) but differ in the catalytic machinery employed to perform the reactions [175]. In the case of inverting GHs, two catalytic amino acid residues are required: a proton donor that performs the nucleophilic attack on the anomeric carbon, and a catalytic base which activates a water molecule. Inverting GPs, on the contrary, require a single catalytic residue, since the catalytic base is the inorganic phosphate itself (Pi). The anomeric configuration inversion is achieved in a single step; the sugar at the non-reductive end is attacked by the catalytic base at the C1 position, a proton provided by the acid/base residue is then captured by the inter-osidic oxygen, which leads to the formation of an oxocarbenium ion-like transition state. Bond cleavage completes the reaction, yielding a glycosyl-phosphate of inverted configuration and an oligosaccharide of reduced chain length. Most retaining GPs (Figure 2) act on α-glycosides. Only a few GH3 family members are reported to be retaining GPs that convert β-glycosides to β-glycosidic phosphate [31]. The mechanism adopted by these enzymes is close to that of retaining GTs. They can either form a glycosyl-enzyme intermediate via double displacement mechanism, or operate through front-side displacement mechanism [175] (Figure 1). The double displacement mechanism occurs in two steps: First, the nucleophile attacks the C1 of the sugar at the non-reductive end, concomitantly to a proton capture by the inter-osidic oxygen on the proton donor. The inter-osidic oxygen-C1 bond cleavage results in the formation of a glycosyl-enzyme intermediate, plus a glycoside of reduced chain length. Following the withdrawal of this product from the catalytic site, the proton donor retrieves a proton from the inorganic phosphate. This activated inorganic phosphate participates in a concomitant nucleophilic attack onto the C1 of the glycosyl-enzyme intermediate. The reaction concludes with the cleavage of the glycosyl–enzyme bond, yielding a glycoside–phosphate of retained configuration (Figure 1).
The front-side displacement mechanism (also known as internal return-like or SNi-like), likely used by GPs of the GT4 and GT35 families, requires a single catalytic amino acid [47]. Only the nucleophile is required, since the proton donor role is played by the inorganic phosphate itself. A hydroxyl from the orthophosphate provides the proton caught by the glycosidic oxygen when a nucleophilic attack occurs. It is worth noting that both the protonation and the nucleophilic attack occur on the same side of the sugar ring, to ensure retention of anomery. The oxocarbenium-ion transition state is stabilized by a nucleophile, namely, in this case, a glutamine or an asparagine.

2.2. Classification and Substrate Specificity

In the Carbohydrate-Active Enzymes (CAZy) Database, GPs are classified into a total of 11 CAZymes families (Table 1). It is worth noting that a N-acetylglucosaminidase from the GH84 family has recently been converted into an efficient GP by a single point mutation [176]. However, as GH84 enzymes do not include any native GP, this family will not be further mentioned.

2.2.1. Retaining GPs

Retaining GPs are found in two GHs families (GH3 and GH13_18) and two GTs families (GT4 and GT35). Only two GPs have so far been characterized in the large GH3 family, which contains 42,678 members (November 2021). Nag3 from Cellulomonas fimi, which was initially characterized as a bifunctional N-acetyl-β-glucosaminidase/β-glucosidase [177], was actually found to also be a β-N-acetylglucosaminide phosphorylase releasing β-d-N-acetyl-glucosamine-1-phosphate (βGlcNAc1P). This new enzymatic activity is due to the presence of a His as a catalytic acid/base residue, instead of the canonical Glu or Asp. The preferred reaction of Nag3 is the cleavage of the disaccharide GlcNAc-anhydro-MurNAc for the recycling of peptidoglycan [31,178]. Another GH3 β-glucoside phosphorylase, BglP, was further discovered by activity-based metagenomics. BglP phosphorolyses the β-1,4-glucosidic linkages of cellulose and cello-oligosaccharides [30]. All the other retaining GPs display an α-linkage specificity, such as the retaining GPs from the GH13_18 subfamily, which act on α-glucosides to produce α-glucose 1-phosphate (αGlc1P).
The GH13_18 subfamily was thought for a long time to contain only sucrose phosphorylases. Nevertheless, sucrose 6′-phosphate phosphorylases (releasing αGlc1P and α-d-fructose-6-phosphate) were recently discovered from Thermoanaerobacterium thermosaccharolyticum [48] and Ruminococcus gnavus [44]. Later on, a novel glucosylglycerate phosphorylase from Meiothermus silvanus was discovered [28]. This enzyme is involved in the metabolism of α-1,2-glucosylglycerate, a molecule known to protect cells from water loss [179]. An α-1,2-glucosylglycerol phosphorylase from Marinobacter adhaerens was further identified. This enzyme is thought to help the bacterium tolerate high salt concentrations, thanks to the protective effect of glucosylglycerol [42]. The sucrose phosphorylases and related enzymes from GH13, their applications, and engineering were recently reviewed in [180].
GPs from the GT4 family are all trehalose phosphorylases [118,119,121,181], while all the characterized members of the GT35 family are α-1,4-glucan phosphorylases, acting on glycogen, starch, and maltodextrins [182,183].

2.2.2. Inverting GPs

Inverting GPs are currently grouped into seven GH families (GH65, GH94, GH112, GH130, GH149, GH161) and one GT family harboring dual glycosyltransferase/phosphorylase activities (GT108) [174]. All those of the GH65 family are involved in the cleavage of α-glucosides, to produce β-d-glucose-1-phosphate (βGlc1P). The first GP from this family was shown to effectively convert maltose and phosphate into βGlc1P and d-glucose [54]. Other enzymes with new specificities, playing key roles in various metabolic pathways, were further identified, such as the trehalose-6-phosphate phosphorylase from Lactococcus lactis subsp. Lactis [62], the trehalose phosphorylase from Thermoanaerobacter brockii [184], the kojibiose phosphorylase from Thermoanaerobacter brockii [185], the nigerose phosphorylase and the 3-O-α-d-Glucopyranosyl-l-rhamnose phosphorylase from Clostridium phytofermentans [25,65], and the 2-O-α-d-glucosylglycerol phosphorylase from Bacillus selenitireducens. The latter utilizes polyols such as glycerol as acceptors rather than longer oligosaccharides or polysaccharides, and could be involved in both the biosynthesis and cleavage of 2-O-α-d-glucosylglycerol, providing the microorganism with adaptation to environments with high salts concentration [67].
Inverting GPs from the GH94 family act on various β-linked glycosides to produce αGlc1P or α-d-N-acetyl-glucosamine-1-phosphate (αGlcNAc1P). The cellobiose phosphorylase and cellodextrin phosphorylases from Clostridium stercorarium [69] were the first to be reported in this family. Later, a N,N′-diacetylchitobiose phosphorylase able to generate α-d-N-acetyl-glucosamine-1-phosphate was identified from Vibrio furnissii [84]. All the GH94 GPs discovered later on were specific for αGlc1P. They include enzymes acting on various β-glucosidic linkages. Among them, the laminaribiose phosphorylase from Paenibacillus sp. [82]) plays a key role in the metabolism of paramylon, an intracellular storage form of β-1,3-glucan usually found in Euglena species [186]. The laminaribiose phosphorylase from Acholeplasma laidlawii was found to effectively produce shorter 1,3-β-d-glucosyl disaccharides [92]. The fungal (Neurospora crassa) and bacterial (Xanthomonas campestris) cellobionic acid phosphorylases catalyze the reversible phosphorolysis of 4-O-β-d-glucopyranosyl-d-gluconic acid, a motif found in recalcitrant cellulosic biomass [90]. Finally, the β-1,2-oligoglucan phosphorylase from Listeria innocua is hypothesized to participate in the metabolism of sophorose-containing glucans and/or exogenous 1,2-β-glucans in specific environmental conditions [87].
The GH112 family contains inverting phosphorylases acting on β-galactosides, which release α-d-galactose-1-phosphate (αGal1P). β-d-galactopyranosyl-1,3-N-acetylhexosamine phosphorylases are subdivided, based on their substrate preference, into galacto-N-biose phosphorylase (GNBP) [99,187], lacto-N-biose I phosphorylase (LNBP) [95], and galacto-N-biose/lacto-N-biose I phosphorylase (GLNBP, with no clear preference for either galacto-N-biose or lacto-N-biose I) [99,188]. The d-galactosyl-β-1,4-l-rhamnose phosphorylase from Clostridium phytofermentans probably plays a role in the degradation of rhamnogalacturonan I (RG-I), an important component of pectin [99].
The GH130 family contains inverting phosphorylases that act on β-mannosides to generate α-d-mannose-1-phosphate (αMan1P). The identification of a 4-O-β-d-mannosyl-d-glucose phosphorylase activity led to the creation of the GH130 family [105]. Then, many different substrate specificities were found, highlighting their role in the catabolism of plant, mammal, bacterial, and fungal mannosides, in particular in the human gut, where these bacterial enzymes are abundant [189]. The β-1,4 mannooligosaccharide phosphorylases, first discovered by Kawahara and colleagues [104], and the 4-O-β-d-mannosyl-d-glucose phosphorylases participate in the breakdown of plant cell wall mannans and glucomannans. The β-1,4-mannosyl-N-acetyl-glucosamine phosphorylase and β-1,4-mannopyranosyl-chitobiose phosphorylase are both involved in the catabolism of eukaryotic N-glycans [108,109]. The β-1,2-mannobiose phosphorylases [100,101] and β-1,2-oligomannan phosphorylases [101] act on exogenous yeast mannan. A β-1,3-mannobiose phosphorylase was later discovered [102], but its physiological function remains unelucidated. In 2013, Ladevèze and colleagues indicated that, based on sequence similarities, the GH130 family could be subdivided into at least two subgroups. The GH130_1 subgroup would include enzymes strictly acting on β-d-Manp-1,4-d-Glc; and the GH130_2 one, those acting on β-1,4-mannosides. The authors also mentioned the presence of a large heterogeneous sequence cluster, named GH130_NC, which was further shown to include β-1,2-mannosidases [190] and β-1,3-mannobiose phosphorylases [102]. Li and colleagues [112] revisited the functional diversity of GH130 enzymes, using a sequence similarity network approach to identify 15 iso-functional sequences meta-nodes. Three new enzymes originating from uncharacterized clusters were studied, which led to the identification of a new function, β-d-Manp-1,4-d-GlcA phosphorylase. This motif is found in the O-antigen of Shigella species lipopolysaccharides [191].
GH149 and GH161 are newly created families, with characterized enzymes originating from the photosynthetic excavate Euglena gracilis, the heterokont Ochromonas spp., and other bacteria that act on β-1,3-glucans, producing αGlc1P. The members of the GH149 and GH161 families likely evolved from a common ancestor, since they harbor the consensus amino acid characteristic of GP activity. However, as their sequences share only about 20% identity, they were segregated into the distinct GH149 [115] and GH161 families [116].
Four GPs were recently characterized in the newly created GT108 CAZy family. They act on mannogen, a β-1,2-mannan found in Leishmania parasites used for energy storage. Interestingly, these four enzymes are dual glycosyltransferase-phosphorylases, which take part in the recycling of mannogen. These Leishmania enzymes bear the same fold and adopt a similar reaction mechanism as bacterial GH130 mannan phosphorylases or hydrolases. However, their significant difference at the sequence level justified classifying them in another family [174].
Taken together, these studies show that GPs can, thus, degrade and synthesize a large array of glycosidic linkages, with the notable exceptions of α-/β-1,6 and β-1,1 linkages. As of November 2021, only 1893 phosphorylases had been characterized (Table 1), which represents about 1% of all the characterized CAZymes [192,193]. Taking into consideration the huge number of GTs and GHs that are listed in the CAZy database, many GPs with novel specificities might remain to be discovered.

2.3. Tertiary and Quaternary Structures

With the exception of GT4 retaining trehalose phosphorylases and GH161 inverting β-1,3 glucan phosphorylases, for which no structure has been elucidated yet, at least one GP crystallographic structure is available for each family in the protein data bank (PDB) (Table 2, Figure 3).
Thanks to these structures, in particular the ones in complex with substrates and products, the catalytic residues have been identified, such as those involved in phosphate binding. These structures complement the site-directed mutagenesis studies that have been performed to identify the critical residues for GP activity (e.g., [109] for the GH130 family). Additionally, when several GP members from the same family have been both functionally and structurally characterized, the structural comparison not only allows identifying the molecular determinants of GP activity but also those establishing donor and acceptor specificities ([189,194] for the GH130 family, for example), or even allows prediction of the phosphorolytic or hydrolytic nature of sequence clusters, together with the linkage specificity of the reactions [112].
From the structural point of view, GPs adopt a range of common 3D-folds, including (α/α)6 and (β/α)8 barrels, five-bladed beta-propeller fold, and GT-B folds. GPs containing CAZY families are univocally associated with a common fold, such as GH130 enzymes, which are all five-bladed beta-propellers. For enzymes from different CAZy families, there is no correlation between 3D-folds and GPs mechanism or substrate specificity, except for GPs belonging to the GH130 and GT108 families, which share a common five-bladed beta-propeller fold and act, for some of them, on the same substrates with the same inverting mechanism.
The vast majority of GPs are present in solution as dimers and, as such, are catalytically active, although no clear demonstration of allostery between the two protomers has been evidenced so far. A few of them are monomers or they assemble as larger homooligomeric complexes, such as the hexameric Uhgb_MP enzyme, for which the catalytic pocket results from the assembly of the different monomers (Table 2). However, the multimerization interface and the precise topology of GP active sites remains difficult to predict [195].

2.4. Biotechnological Applications

The different approaches for glycoside production include polysaccharide extraction from natural sources (e.g., microorganisms or plants) [227,228], hydrolysis and size-fractionation [229], chemical synthesis [7], (chemo-)enzymatic synthesis [7,230], and whole cell or cell-free synthesis [230,231]. Due to the high diversity of glycosides found in nature, direct extraction from natural sources remains the method of choice, especially for high-scale applications. However, several purification steps are required, often making use of harsh chemicals or mechanical extraction methods. Resource competition between food and feed industries can also appear when they are produced from raw plant material. This method is, thus, restrained to the extraction of abundant glycosides, such as sucrose (for the food industry) or cellulose (paper industry, biofuels) [227,232].
Chemical synthesis is well adapted for the synthesis of both structurally simple and more complex carbohydrates. Owing to its versatility, it is indeed possible to either introduce rare glycosyl units in glycosides or to assemble them in a more modular or complex fashion. Several automated systems were recently investigated [233,234], and this led to the chemical synthesis of complex glycoside libraries [235,236]. While being currently addressed [237], major drawbacks intrinsic to the use of these chemical syntheses are inevitable. One can think about the tedious manipulation of protecting groups leading to multi-step pathways, the use of (often toxic) solvents and harsh temperature conditions, or the lack of reaction selectivity [7]. Even though some carbohydrates cannot be accessed through chemical synthesis or at the cost of low yields, chemical synthesis still remains one of the most used methods to produce specialty and high added-value glycosides at industrial scale.
Finally, enzyme-based approaches, sometimes in combination with chemical steps to take advantage of the best of both worlds, are increasingly being envisioned for the synthesis of glycosides [7,110,238]. Enzymes, indeed, offer exquisite regio- and stereo-specificities, while being able to work under mild pH and temperature conditions in aqueous solvents. Enzymatic glycoside synthesis now benefits from an expanded catalogue of characterized biocatalysts found in many CAZy families, which can even be further engineered, when lacking the requisite properties [2,10]. Whole-cell biocatalysts also received growing interest, especially for the in cellulo coupling of carbohydrates to lipids or proteins, allowing the production of complex glycoconjugates for the pharmaceutical industries [230]. Examples of enzymatic syntheses strategies involving GPs are presented in Figure 4.
As explained above, GPs can be employed for the synthesis of high-value oligosaccharides, polysaccharides, and glycoconjugates, by reverse-phosphorolysis from glycosyl phosphates. Likewise, the phosphorolysis reaction can be used to regenerate glycosyl phosphates for multi-step enzymatic syntheses, for which a review was recently published [238]. Several examples of glycoside structures accessed through enzymatic synthesis using GPs were previously reported and reviewed [175,239,240]. Let us mention the case of the GH130 Uhgb_MP enzyme, which can use the relatively cheap N-acetyl-d-glucosamine and αMan1P substrates to produce the highly expensive β-1,4-mannopyranosyl-chitobiose; of major interest as the core part of human N-glycans [109]. Crystalline linear β-1,4-mannan, a major component of plant cell-wall hemicelluloses, has been successfully produced using the GH130 β-1,4-mannooligosaccharide phosphorylase TM1225 [110]. Similarly, sucrose-phosphorylases can use a cheap and renewable donor substrate, to synthesize α-1,2-glucosylglycerol, a compound used in cosmetics [241]. As previously mentioned, the ability of GPs to generate sugar phosphates from low-cost substrates by phosphorolysis can advantageously be used in enzymatic cascades, to regenerate nucleotide sugar donors essential to GTs (Figure 4). This was exemplified in vitro, with the kg-scale production of LNB from sucrose and N-acetyl-glucosamine, using an elegant one-pot four-enzyme system [239,240], and also in cellulo for the production of phenolic glucosides from sucrose in Escherichia coli [242], using GPs in tandem with glycosyltransferase. In these examples, the GPs were, indeed, used in the phosphorolysis direction, in order to increase the pool of sugar-phosphates from the phosphorolysis of low-cost oligosaccharides (e.g., sucrose in [74]). In these processes, sugar-phosphates are first converted to nucleotide-activated sugars, before being used as substrates for high added-value oligosaccharide/glycoconjugate synthesis by recombinant glycosyltransferases.
When using GPs in the reverse phosphorolysis direction for glycoside synthesis, one issue is the reversibility of the reaction, whcih could lead to low production yields when reaching equilibrium. A typical solution is to shift the equilibrium to promote the reaction in the desired direction, by withdrawing the reverse phosphorolysis product (Figure 4). This can easily be achieved in biphasic systems, when the product presents low solubility in the aqueous medium. Cellodextrin or cellobiose, for instance, produced with the GH94 cellodextrin phosphorylase, show a low solubility in the reaction medium and precipitate, so it can easily be recovered [243,244]. Biomineralization of phosphate under the form of struvite crystals (MgNH4PO4) has also been reported [245]. Another method for shifting the reaction equilibrium towards substrate consumption is to swap glycosyl phosphates for glycosyl fluorides as donor substrates. The release of fluoride does not cause an attack of the glycoside; thus, preventing phosphorolysis and significantly increasing the production yields [74]. However, glycoside fluorides are more expensive than their phosphate counterpart, and similarly to the fluoride donors for glycosynthases mentioned above, they are often unstable. However, when the desired product remains in the same phase as the substrates, the simplest way to shift the equilibrium towards completion remains to increase one of the substrates’ concentration. Obviously, the cheapest solution is often to increase the acceptor’s concentration, and relative excesses of 1:1.2 to 1:1.5 molar equivalents were shown to give appreciable results [246]. Multi-enzymatic cascades can lead to a similar result, without actually changing the initial concentration of substrates. Due to the interplay of multiple enzymes, the global equilibrium can be substantially changed. For example, in vitro continuous flow phosphorylation of the glycosyl donor using an immobilized phosphorylase [247] allowed successfully synthesizing oligosaccharides by immobilized multiple enzymes cascades [248] or using one-pot multi-enzymatic (OPME) systems [238,249]. Shifting the reaction’s equilibrium towards synthesis can also be successively achieved using an in vitro/in cellulo metabolic coupled system if part of the enzymatic cascade involves cellular catabolic enzymes. This is well exemplified by the synthesis of mannobiose, mannotriose, and mannosyl-glycerate from starch [250]. In this example, six consecutive in vitro steps allow the efficient generation of mannose from starch. This molecule provides the starting material for the intracellular production of β-1,2 mannobiose and mannotriose catalyzed by mannobiose phosphorylase in a Corynebacterium glutamicum strain. The synthetic pathway for mannosylglycerate synthesis has also been constructed by channeling the mannose taken up by the strain towards the production of GDP-mannose, the substrate of mannosylglycerate synthase. This can also be fully achieved in cellulo, but this generally requires strain engineering to increase the overall yield or to add the missing steps required to produce certain specialty compounds. Good examples include the production of 2-fucosyllactose, kojibiose or cellobiose using sucrose phosphorylase acting in the phosphorolysis direction in tandem with glycosyltransferases; or even by using two different GPs, one acting in the phosphorolysis direction, to accumulate the desired glycosyl-phosphate, and the other in the synthesis direction [251].
Lastly, when native enzymes do not meet the criteria to be considered for industrial scale production of carbohydrates, they can be modified using enzyme engineering. It is, therefore, possible to broaden the range of application of GPs, by altering their specificity towards defined donors or acceptors, or even improve the enzyme (thermo)stability or resistance to solvents, making them more suited to, or compatible with, certain industrial steps. Reviews illustrating various approaches and examples of GP engineering were recently published [2,32,180]. For example, a lactose phosphorylase was created by mutating the acceptor binding site of a GH94 cellobiose phosphorylase, to extend the range of recognized acceptors [252]. Similarly, the acceptor specificity of a cellobiose phosphorylase was modified in such a way that the engineered enzyme preferred methyl β-glucoside rather than the natural acceptor, glucose [253]. Other examples include Bifidobacterium adolescentis sucrose phosphorylase engineering, where a point mutation triggered a domain shift essential to accommodate a variety of large phenolic compounds, permitting resveratrol and quercetin glucosylation [254]. As industrial processes are usually performed near 60 °C, to avoid microbial contamination [255], several studies focused on the discovery and engineering of GPs from thermophilic microorganisms [110]. For instance, the sucrose phosphorylase from Bifidobacterium adolescentis showed high temperature tolerance properties compared to the other characterized sucrose GPs, making it a suitable template for thermostability engineering and understanding [32]. The diversity of known GPs is unfortunately restricted to one or a few enzymes for each family, limiting the understanding of their structure–specificity–stability relationships, and, thus, impeding rational engineering studies. Efforts are needed to discover new GP candidates for industrial applications. Even if the very nature of their catalytic mechanism makes them barely distinguishable from canonical GHs and GTs, a handful of approaches allow their detection using activity- and sequence-based methods.

2.5. Methods for Glycoside Phosphorylase Activity Detection

Enzyme engineering and discovery from natural sources require the use of suitable screening methodologies. As the number of candidates to be tested can reach hundreds of millions, especially when mining metagenomic or combinatorial mutagenesis libraries, efficient screening methods that allow the detection of enzymes presenting the desired properties must be employed. A variety of approaches, including chromogenic assays, and chromatographic or spectroscopic analyses, are typically employed for detecting GPs and can be adapted to high throughput (Figure 5).

2.5.1. Chromogenic Assays

As the inorganic phosphate released during the reverse-phosphorolysis reaction or consumed during phosphorolysis directly reflects GP activity, measuring its concentration via chromogenic assays constitutes a very convenient proxy to identify GPs. The molybdenum blue activity assay was initially developed by Fiske and Subbarow, but not specifically for detection of GP activity [256]. With this method, the measurement of phosphate consumption or release, depending on the reaction screened for, is based on its conversion to a phosphomolybdate complex in the presence of the molybdate reactant. Under acidic conditions, phosphomolybdate is reduced to molybdenum, of which the blue color intensity can be quantified by absorbance measurement at 655 nm. This method was further modified in several ways, to adapt it to the micro-plate format and automate it [257,258,259]. It was later adapted to the detection of GP-catalyzed reverse-phosphorolysis activity in crude cellular extracts of E. coli, which cause high background and consumes glycosyl phosphates [79,260].
Phosphate concentration can also be measured with another chromogenic assay in liquid medium, using coupled enzyme systems. For instance, sucrose phosphorylase activity can be detected using a coupled assay with a phosphoglucomutase (which isomerizes G1P to G6P) and a glucose-6-phosphate dehydrogenase to release 6-phosphogluconate and NADPH. A final step involves a 6-phosphogluconate dehydrogenase yielding an additional NADPH molecule, together with ribulose-5-phosphate [261]. Similarly, the glyceraldehyde-3-phosphate dehydrogenase uses glyceraldehyde-3-P, NAD+ and inorganic phosphate to form 1,3-bisphosphoglycerate, releasing H+ and NADH. Both NADH and NADPH can then be detected at 340 nm [262]. It is, therefore, possible to couple this conversion to a preceding reaction that releases inorganic phosphate, resulting in an absorbance readout of NADH directly equivalent to the activity of the phosphorylase. Finally, a coupled enzymatic assay usable at pH 6 to 9 was recently developed and established in the context of DNA polymerase. It is based on a three-enzymes system: the concerted action of an inorganic pyrophosphatase, a purine nucleoside phosphorylase, and a xanthine oxidase generate uric acid. Absorbance readout at 293 nm allows monitoring the reaction and directly linking it to the release of phosphate [263]. Both assays are yet to be used to detect GP activity.
More recently, another assay in liquid format was employed for the screening of GP-catalyzed phosphorolysis reactions. It relies on the cleavage of a 2,4-dinitrophenyl glycoside donor in the presence of inorganic phosphate, yielding phenols that could be revealed by absorbance measurement at 400 nm. Even though compatibility with some GP activity detection was proved [30], others can be inactive or weakly active [109] on those chemically modified substrates that differ from the natural glycoside target.

2.5.2. Chromatographic and Spectroscopic Methods for Carbohydrate Detection and Characterization

In order to detect and characterize carbohydrate structures, various methods, with variable sensitivity, are available (Figure 5). These methods drastically vary, regarding e.g., the time for analysis, the sensitivity, the amounts of analyte required, or their destructive or non-destructive nature.
Thin layer chromatography (TLC) is a qualitative separation method that is widely applied for the separation of monosaccharides, oligosaccharides, and sugar phosphates. It is cheap and fast, easy to implement and perform, but suffers from a relative low separation efficiency and lack of information when dealing with unknown spots [264].
High performance liquid chromatography (HPLC) can be used as a quantitative analysis methodology. It generally relies on column-based separation, for example, with copolymer resins in protonated or metal ion forms that are suited for carbohydrate separation. While it offers good separation of carbohydrates (often after optimization of the elution gradient), one significant drawback is carbohydrate detection. Carbohydrates can commonly be detected using low-wavelength UV, but due to the similar absorbance between negatively charged sugars and the organic mobile phase (e.g., acetonitrile), UV detection is unsuited for detecting carbohydrates in mobile phases. Derivatization of the product by adding suitable chemical groups that can be detected at a different wavelengths [265] constitutes a way of dealing with this problem, but this requires tedious additional steps and is, therefore, restricted to qualitative analyses. Another common system for carbohydrate detection relies on the refractive index difference using refractive index detection (RI), although this often suffers from low sensitivity.
High-performance anion-exchange chromatography (HPAEC) offers high sensitivity for the separation of mono-, oligo-, and polysaccharides, based on their structural features. For example, several degrees of polymerization (DP) 2 or DP3 oligosaccharides, only differing by their osidic linkage types, can be separated by HPAEC. This is often coupled with pulsed amperometric detection (PAD), a highly specific and sensitive detection method (in the nanomole range) for carbohydrates [266]. HPAEC-PAD remains one of the most attractive and powerful methods for carbohydrate analysis. Nevertheless, as the response factor depends on the carbohydrate structures, it is impossible to determine the concentration of a product for which no pure standard is available. Although HPAEC allows separating glycosyl phosphates from low DP glycosides and monosaccharides, the very low PAD response factor for glycosyl phosphates makes the simultaneous determination of monosaccharide, oligosaccharide, and glycosyl phosphate concentrations difficult.
The determination of oligosaccharide structures requires information about the anomers, linkages, and their component glycosyl units. Permethylation analysis was developed in the late 1960s and is used to determine the position of glycosidic linkages [267]. Typically, the carbohydrates are derivatized to form acid-stable methyl ethers, then hydrolyzed and converted to alditol acetates, further separated by gas chromatography, and finally analyzed by mass spectrometry (MS) [268].
Nuclear magnetic resonance (NMR) spectroscopy can provide the configuration of glycosidic linkages, by comparing the chemical shifts with those of pure standard molecules. It is non-destructive, highly sensitive, and allows the analysis of product mixtures, even in crude cell extracts [269]. NMR is extensively used to determine the structure of homo-oligosaccharides and homo-polysaccharides synthesized by GPs [101,109,270].
Mass spectrometry (and more specifically, tandem MS) is a powerful analytic method allowing the identification of ionized species according to their mass-to-charge (m/z) ratios [271], and which is widely used to determine glycoside structures. The methylated alditol acetates are, therefore, identified by a combination of the relative retention times of the analytes and their fragmentation spectra.
Formal identification of donors and acceptors or osidic linkage type still relies on the use of the traditional workhorses of glycobiology, such as HPAEC-PAD or NMR. Reverse-phosphorolysis product structural determination also relies on a standard’s commercial availability or chemical synthesis. When no standards are available, highly resolutive MS-MS can be used to determine the linkage position. Recent methods, such as ion-mobility based separation, offer exquisite carbohydrate separation, even for isomers co-eluted by other separation techniques [272]. Associated with very sensitive and specific tandem MS methods, (complex) carbohydrate structure can now be completely elucidated [272,273]. This analysis only requires 1 µg of carbohydrate, directly from reactional mixtures. This new analytical technique paves the way for carbohydrate ‘sequencing’ using cyclic ion mobility mass spectrometry. Overall, great improvements in these methodologies allow a faster separation and the handling of more samples (with data acquisition per sample within minutes), with reduced compound consumption. These methods, previously used for precise product characterization in final screening steps, now offer real possibilities for screening large libraries, while generating good quality data [113].

2.6. Enzyme Discovery by Functional (Meta)Genomics

The vast majority (74%) of GPs discovered so far originate from archaebacteria and bacteria, which produce these enzymes for glycoside catabolism. More than 1030 bacterial and archaeal cells are present on Earth [274]. Depending on the habitats, uncultured species can make up to 99.9% of the ecosystem [270]. For instance, the human gut microbiota, one of the most studied ecosystems, due to its importance for health, counts about 70% uncultured species [275]. With the rapid development of next generation sequencing (NGS) technologies, sequencing costs have significantly dropped in the past two decades, opening the way to massive sequencing of metagenomes, the collective genomes of microbiota [276]. Most of these works are monogenic metagenomic studies, focusing on the sequencing of 16S (bacteria, archea) and 18S (eukaryotes) rRNA genes, to describe the microorganisms diversity in an environmental sample, without culturing them [277]. Nevertheless, the last decade has seen more and more functional metagenomics projects being launched. Functional metagenomics is the sequencing of functional genes to predict and/or prove their function [278]. These projects led to the establishment of several gene catalogs, issued from various microbiota, particularly those of human [279,280], mouse [281,282], pig [283], and cow [284] guts; totaling nearly 34 million genes and thousands metagenome-assembled genomes (MAGs). The drastic increase of available metagenomic sequences from various catalogs yielded hundreds of novel enzymes and activities, identified through activity- or sequence-based analyses (Figure 5).

2.6.1. Activity-Based Approaches

The activity-based approaches’ typical workflow begins with metagenomic DNA cloning (or cDNA for eukaryotes) from metatranscriptomic samples, in a vector (plasmids, fosmids, or bacterial artificial chromosomes). The genes issued from the microbiome are then expressed in a recombinant host [285]. In activity-based metagenomic studies, the host is often Escherichia coli, due to its high transformation efficiency and its ability to efficiently express genes issued from distant taxa [286]. The metagenomic libraries then have to be screened with a suitable screening assay, preferably at a high throughput. Activity-based approaches raise issues, such as the typical GC bias found in E. coli clone libraries, which may affect the DNA insert stability in the vector [287], or the potential toxic effect of the expressed exogenous gene on the cells.
To circumvent these problems, other host strains and vectors have been developed [288,289,290]. In activity-based metagenomics, the positive hit rate greatly depends on the metagenomic DNA source, its abundance in the metagenomic sample, the chosen cloning vector and expression host, the variability of heterologous gene expression, and the sensitivity and inherent bias of the screening assay. Hit rates typically range from 0.01‰ to 10‰ [291]. This approach, thus, requires high-throughput screening facilities to screen a sufficiently large sequence space (typically 104 − 105 clones, which corresponds to 0.3 × 106 to 3 × 106 genes for fosmid libraries). Depending on the assay format, the screening throughput varies between 104 (automated liquid assays in micro-plates) to 105 assays per day (automated assays on solid plates) [292], up to 106 with droplet-microfluidics [293].
To date, only three activity-based functional metagenomic studies have been performed to discover GPs. In the first one, a new sucrose-phosphorylase (Genbank accession FJ472846, not listed in the CAZy database, but likely a GH13_18 member) was isolated from a 5000-clone metagenomic library constructed from a sucrose refinery sample. In this case the screen consisted in a positive selection assay on a solid mineral medium containing sucrose as sole-carbon source [49]. More recently, the S. Withers group has started to use metagenomics to mine various ecosystems for GPs. Their first study was based on the screening of 17,168 clones library obtained from a passive mine tailings biochemical reactor system, on the activated substrate 2,4-dinitrophenyl β-d-glucoside (DNPGlc) and inorganic phosphate, leading to the discovery of the retaining β-glucoside-phosphorylase BglP, a rare GP from the GH3 family [30]. The last study focused on cello-oligosaccharide-degrading GPs, searched in ∼23,000 clones sourced from the same passive mine tailings biochemical reactor system, as in the previous study, and from beaver feces. In this work, MacDonald and colleagues used a liquid assay based on molybdenum blue formation to identify seven new GPs from the GH94 and GH149 families (cellobiose phosphorylases, cellodextrin phosphorylases, laminaribiose phosphorylases, and a β-1,3-glucan phosphorylase) [79]. By varying substrate combinations, the authors demonstrated the versatility of the screening method, with which we can expect GPs from other families to be discovered in the near future.

2.6.2. Sequence-Based Approaches

Contrary to activity-based approaches, sequence-based screening is performed using either in silico sequence homology-based methods, or in vitro, using DNA/DNA or DNA/RNA hybridization-based methods (such as PCR) to isolate sequences with known consensus motifs. The latter have only been used to discover CAZymes [294], and not yet to discover GPs. Since they rely on identifying sequences that possess the same motifs involved in substrate recognition and catalysis as those of already known enzymes, these methods perform poorly in discovering new enzyme activities and families.
In silico sequence homology-based methods rely on isolating sequences similar to a reference sequence for which the function has been experimentally confirmed, or which differs from it at particular locations. Two highly similar sequences are indeed expected to have evolved from a common ancestor, and, thus, most likely share a similar function. Alignment-search tools such as BLAST [295] are used to classify proteins into evolutionary families. It is, however, possible that two distant protein sequences may have the same function, owing to structurally conserved active site motifs. This is actually the case for members of the different CAZy families that share the same activity. For instance, β-mannosidases can be found in five different families (GH1, GH2, GH5, GH130, and GH164), even though some are structurally unrelated; with the GH1, GH2, and GH5 families harboring (β/α)8 folds and GH130 the five-bladed beta-propeller fold (unknown for GH164). In some cases, the opposite is found: the CAZy families may be related distantly, sharing less than 20–30% sequence identity, but still have high structural similarities, as for GH130 and GT108 families, which both have a five-bladed beta-propeller fold, and which both contain β-mannoside-phosphorylases [174].
Several databases, including COGs, Pfam, and KEGG, allow for automated functional annotation of metagenomic sequences. COGs identifies orthologous genes that typically share the same general function, such as carbohydrate transport and metabolism for the G cluster, in which CAZymes are found [296]. Pfam contains nearly 18,000 protein families (Pfam 32.0) [297]. Each family is created based on multiple sequence alignments and hidden Markov models (HMMs) [298]. Since Pfam families (like CAZy families) contain members with several activities, the activity of an enzyme cannot be deduced from the family it belongs to. Finally, KEGG can provide information on the involvement of a protein in the metabolism of an organism or an ecosystem [298]. Nevertheless, sequence similarity-based methods leave orphan genes (genes which do not share significant sequence similarity with any known sequence) without annotation. Furthermore, due (and participating in) to the amplification of incorrect functional annotations in databases, these tools can generate annotation mistakes in the preliminary functional annotation of sequences. For instance, when searching for the GH130 β-1,4-mannoside-phosphorylase Uhgb_MP (Genbank accession number ADD61463.1) in these databases, the results are the following:
  • COG: the protein belongs to COG2152: Predicted glycosyl hydrolase, GH43/DUF377 family (Carbohydrate transport and metabolism). The biochemical characterization of this enzyme [109], indicated that these results are false, this enzyme being a mannoside-phosphorylase of the GH130 family, which does not contain a GH43 module.
  • Pfam: the protein contains the domain Glyco_hydro_130 (PF04041), described as ‘beta-1,4-mannooligosaccharide phosphorylase’. This is true, but the same result is found for the GH130 β-1,2-Mannosidase AAO78885.1 [194].
  • Kegg: no result.
The IMG-M and MG-RAST servers [299,300] and the commercially available QIAGEN CLC Microbial Genomics Module are convenient tools which integrate data resulting from the interrogation of these different databases. They make it possible to quantify and compare the abundance of the main functional families in the target ecosystems.
For more accurate function prediction, especially for substrate specificity or mechanism prediction of an enzyme, sequence analysis can be complemented with local genomic neighborhood analysis. Indeed, the in silico analysis of polysaccharide utilization loci (PUL) composition, and thus, of the combinations of CAZymes involved in glycoside catabolism, can be a good way to predict the substrate targeted by a Bacteroidota PUL [301], and by each of its CAZyme components. This helped to discover the biological function of Uhgb_MP (for ‘unknown human gut bacterium_mannoside-phosphorylase’), the first GP issued from an uncultured bacterium [109], thanks to the cloning of entire metagenomic PULs in fosmids [286]. More recently, this strategy permitted elucidating the function of a novel sucrose-6F-phosphate-phosphorylase from the cultured human gut bacterium Ruminococcus gnavus E1 [44]. Nevertheless, such a strategy is restricted to long DNA sequences, and cannot be used with highly fragmented loci sequences, such as those found in datasets obtained by massive and random sequencing of metagenomes. Moreover, when multiple uncharacterized GPs from the same family are found together on the same PUL, predicting the substrate nature is very difficult, making the results uncomplete. Besides, given that the Bacteroidota PULs are the sole CAZy-containing bacterial operons referenced to date, the prediction of non-Bacteroidota PUL-like systems remains highly speculative, resulting in a large variability of the loci composition.
Mining unexplored clades of GP-containing CAZy families or subfamilies phylogenetic trees is also an efficient strategy for discovering new GP specificities, as exemplified with the GH13_18 subfamily [180]. This strategy led to the discovery of glucosylglycerate phosphorylases [28,42] and of a sucrose 6′-phosphate phosphorylases [48]. In the GH130 family, β-1,3- [102] and β-1,2-mannoside-phosphorylases [101] were found in the genomes of cultured strains.
Structure prediction and docking experiments can also be very useful [302]. Examination of the active site and catalytic residues conservation is indeed probably the best way to predict if a CAZyme is a GH or a GP. Inverting GH130 enzymes can either be GHs when they possess the two canonical carboxylic amino acids required to catalyze hydrolysis reactions, or GPs when only the proton donor and phosphate-binding residues are present [109,194]. For retaining enzymes, some GH3 members were shown to act as phosphorylases rather than hydrolases, because they use a His/Asp dyad as an acid/base catalyst rather than the standard Glu or Asp residues. This allows an anionic phosphate to enter the active site; while this is greatly disfavored when the active site bears a deprotonated carboxylate base directly adjacent to the anomeric center of the glycosyl enzyme [31]. This difference was confirmed recently, when a single point mutation of the acid/base, from Asp to Asn was shown to convert GH84 retaining hydrolases to efficient phosphorylases [176]. However, in-depth sequence analysis and structure prediction require specific expertise, to handle large amounts of data and align thousands of sequences.
Analyzing such high numbers of sequences more easily than by using multiple sequence alignments and phyologenetic tree analyses, to visualize sequence relationships in protein families, is essential. The Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST) was developed in 2015 [303] for this purpose. This web tool, dedicated to all members of the biological/biomedical community, including non-bioinformaticians, automatically generates sequence similarity networks (SSNs) [304], based on pairwise sequence comparison. In an SSN, each sequence is represented by a node (symbol), connected by an edge (line) to all the other nodes corresponding to sequences that share a sequence similarity greater than a user-specified value. This approach allows highlighting unexplored sequence spaces, guiding biochemists to discover novel functions. SSNs have been used to split a glycoyl radical enzyme superfamily into different iso-functional clusters, guiding the biochemical characterization efforts towards unexplored clusters of metagenomic sequences that are highly abundant in the human gut microbiome. This led to the discovery of a novel enzyme involved in the biosynthesis of L-proline, a key mediator of healthy microbiota–host symbioses [305]. Even though extremely powerful, the SSNs-based approach had not been tested for CAZyme diversity analysis until very recently, with the multispecific GH16 family subdivision into 23 robust subfamilies [306]. The GH130 family was also recently analyzed with this tool, which allowed revisiting the historical GH130_1 and GH130_2 subfamilies classification, to generate 15 robust sequence clusters (20+ sequences containing meta-nodes) [112]. This structure–sequence–function data integration permitted focusing biochemical characterization efforts onto five cluster representatives, to demonstrate the segregation of hydrolases and phosphorylases at the cluster level, and identify new functions, such as the GH130 Man-GlcA phosphorylase [112]. Finally, we can report the discovery of a new GP activity among GH94 enzymes using an SSN based strategy, 4-O-β-d-glucosyl-d-galactose phosphorylase [86]. However, the CAZy database and the publicly available metagenomic sequence datasets remain largely untapped, representing gold mines to be explored with such approaches, with the prospect of novel GP discoveries.

3. Conclusions and Perspectives

The necessary conversion towards eco-friendly solutions and cleaner processes in industry is largely based on the use of biocatalysts. Renewable sourcing of carbon mostly relies on the use of plant polysaccharides, meaning that an increasing number of processes will be dependent on CAZymes. Glycoside phosphorylases are fascinating CAZymes, with unique properties, combining the best of both worlds of GHs and GTs. The integration of such enzymes into the expanding catalogue of industrial applications, envisioned through the march towards global decarbonization, is a challenge facilitated by the exquisite plasticity and reactional diversity provided by these energy-optimized catalysts. Their unique biochemical properties, making them particularly adapted for biochemical processes improvements, have long been recognized, as testified by the interest in these enzymes manifested by both the scientific community and private companies. They, indeed, provide an elegant way to address the shortcomings of GHs and GTs, either by a direct replacement at the reaction level for improved efficiency or by complementation in fully integrated biological pathways.
Mostly used for their synthetic properties, polysaccharide degradation by GPs is energetically superior to hydrolysis, since transferring a glycosyl moiety onto a phosphate residue, rather than onto a water molecule, retains the inter-osidic bond energy in the newly created glycosyl-phosphate bond. The generated glycosyl-phosphate can subsequently be reused as an activated donor, by the same or another GP. Initially produced from a generally cheap and renewable material, this is a sustainable and elegant way to generate activated sugars. These are, most of the time, directly used as substrates for the actual reaction of interest, with the transfer of the glycosyl moiety onto an exogenous acceptor. Molecules of interest are extremely diverse, whether for the production of bio-compatible compounds designated for food/health applications or as an alternative for hazardous substances. GPs can also be part of a larger in vitro synthesis chain of reactions, helping to reduce the use of solvents, or improving the overall yield by providing exquisite stereoselectivity at a critical step. Nevertheless, even when only used for the degradation of polysaccharides, the energetic superiority of GPs over GHs can make a difference if the resulting monomers are to be fermented for direct energy supply, or as input carbohydrates for in cellulo synthesis in microbial cell factories, since sugar-phosphates have a higher energetic content than their isolated monosaccharide equivalents.
Nevertheless, the full potential of GPs remains largely underexploited, due to the very nature of their similarity to both GHs and GTs. As sequence-specific characteristics that would make them easy to identify are limited, case-by-case manual testing largely remains the method employed to demonstrate GP activity and characterize their acceptor tolerance. Thus, the known GP-catalyzed reactions are still largely less diverse than those catalyzed by GHs, and studies aiming at specifically isolating them from natural environments are also less common. However, in recent years technological advances have been made to address these limitations. GP-specific biochemical assays have been developed and used in high-throughput screening campaigns. Droplet microfluidics, in this respect, holds great promise for GP discoveries by functional metagenomics. Sequence-specific markers, such as the absence of signal peptides or unique phosphate binding residues can so far only be analyzed manually, family by family, when functional data are available. Multi-omics technologies can now be augmented by automated genomic data analysis, using methods such as SSNs, allowing glimpsing a bright future in the field, by specifically identifying potential GP-containing sequence clusters. These tools provide guidance to the experimenter, for focusing their efforts onto the most promising targets identified in the mountains of genomic data accumulated over the years. Reanalyzing GP-containing CAZy families in this new light would also help identifying novel activities from isolated sequence groups. SSNs are not only advantageous in highlighting groups of uncharacterized sequences, but also by splitting the diversity of each family into several smaller sequence groups, facilitating multiple sequence alignments, and searching for consensus motifs indicative of, e.g., GPs versus GHs or a particular activity. Splitting large datasets into sub-groups of similar sequences with SSNs is also an interesting strategy to decrease computational time and to improve structural alignments, based either on crystallographic protein structures or on tridimensional models. Indeed, SSNs alone cannot be used to predict enzyme mechanism, nor substrate or linkage specificities, which can vary due to differences in a single amino acid. When a sufficient number of enzymes with different mechanisms and linkage specificities have been biochemically characterized, primary sequence analysis can be sufficient to predict the GP or GH mechanism.
However, most of the time, SSNs and multiple sequence alignment analysis are insufficient to predict substrate specificities, since discrete differences of the active site topologies between enzymes belonging to the same meta-node can result in different activities. SSNs also have other limitations. First, they cannot be directly used to analyze CAZy families containing many multi-modular sequences. In this case, the multi-modularity must be analyzed and the sequences truncated, to analyze with SSNs, only the CAZy module of interest. This is what was done with the GH16 family [306]. Nevertheless, as a large share of GPs families are mono-modular, SSNs can often be used directly. The other limit of SSNs is their instability when the number and diversity of the analyzed sequences vary. This is why the creation of stable CAZy sub-families cannot be based solely on SSNs-based clustering, but must also use phylogenetic analyses, which are much more stable and have the advantage of providing evolutionary distances, contrary to SSNs. Nevertheless, again, SSNs can be used to facilitate the creation of the multiple sequence alignments required for the generation of such phylogenetic trees, by selecting a smaller number of sequences from each meta-node.
Artificial intelligence (AI)-based tools have attracted a lot of attention in recent years, holding promise for faster and more accurate structural predictions. Equipped with improved structural modeling tools, structural biology is thought to ease GP identification directly from protein sequences. Highlighting structural hotspots for GP activity or substrate specificity is thought to convey meaningful information for achieving fast and reliable production of desirable catalysts. AI is also envisioned to be an extremely valuable tool for data mining and integration, making sense of dispersed data by facilitating the combination of multiple data types and the summarization into structural projections easily understandable by the experimenter.
Ultra-high throughput (UHT) functional screening technologies, such as droplet microfluidics, which have only started to be used for CAZyme screening, have vast potential for GP identification and engineering [307]. Using, e.g., the coupled reactions detailed previously in this manuscript, it might be possible to specifically target phosphorolytic activities from libraries > 106 members large, making full use of the increased throughput offered by droplet microfluidics. With in-line coupling with analytical methods, such as MS or NMR, droplet microfluidics can become a valuable tool for glycomics, allowing the sorting of library members directly upon structural identification of the desired glycan patterns produced.

Author Contributions

G.P.-V. conceived and designed the review. A.L., M.B., J.D. and G.P.-V. wrote the first draft. S.L. prepared the figures. A.L., G.C., E.L. and S.L. drafted the tables. S.L. and G.P.-V. finalized and submitted the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Union’s Horizon 2020 framework programmes (grant numbers LEIT-BIO-2015-685474, Metafluidics and FNR-2020-SEP-210639118, RadicalZ) and by the Agence Nationale de la Recherche (ANR-16-CE20-0006, Oligomet). A.L. was supported by INSA Toulouse and China Scholarship Council under the ‘CSC-UT/INSA Program’. The APC was funded by the European Union’s Horizon 2020 framework programme (grant number FNR-2020-SEP-210639118, RadicalZ).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. André, I.; Potocki-Véronèse, G.; Barbe, S.; Moulis, C.; Remaud-Siméon, M. CAZyme discovery and design for sweet dreams. Curr. Opin. Chem. Biol. 2014, 19, 17–24. [Google Scholar] [CrossRef] [PubMed]
  2. Benkoulouche, M.; Fauré, R.; Remaud-Siméon, M.; Moulis, C.; André, I. Harnessing glycoenzyme engineering for synthesis of bioactive oligosaccharides. Interface Focus 2019, 9, 20180069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Li, W.; McArthur, J.B.; Chen, X. Strategies for chemoenzymatic synthesis of carbohydrates. Carbohydr. Res. 2019, 472, 86–97. [Google Scholar] [CrossRef] [PubMed]
  4. Rudroff, F.; Mihovilovic, M.D.; Gröger, H.; Snajdrova, R.; Iding, H.; Bornscheuer, U.T. Opportunities and challenges for combining chemo- and biocatalysis. Nat. Catal. 2018, 1, 12–22. [Google Scholar] [CrossRef]
  5. Lairson, L.L.; Henrissat, B.; Davies, G.J.; Withers, S.G. Glycosyltransferases: Structures, Functions, and Mechanisms. Annu. Rev. Biochem. 2008, 77, 521–555. [Google Scholar] [CrossRef] [Green Version]
  6. Wong, C.-H. Practical synthesis of oligosaccharides based on glycosyltransferases and glycosylphosphites. In Modern Methods in Carbohydrate Synthesis; Khan, S.H., Ed.; CRC Press: Boca Raton, FL, USA, 1996; pp. 467–491. ISBN 978-3-7186-5785-8. [Google Scholar]
  7. Overkleeft, H.S.; Seeberger, P.H. Chemoenzymatic synthesis of glycans and glycoconjugates. In Essentials of Glycobiology; Varki, A., Cummings, R.D., Esko, J.D., Stanley, P., Hart, G.W., Aebi, M., Darvill, A.G., Kinoshita, T., et al., Eds.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2015. [Google Scholar]
  8. Fierfort, N.; Samain, E. Genetic engineering of Escherichia coli for the economical production of sialylated oligosaccharides. J. Biotechnol. 2008, 134, 261–265. [Google Scholar] [CrossRef]
  9. Samain, E.; Priem, B. Procede de Production D’oligosaccharides. European Patent EP1194584B1, 22 February 2006. [Google Scholar]
  10. McArthur, J.B.; Chen, X. Glycosyltransferase engineering for carbohydrate synthesis. Biochem. Soc. Trans. 2016, 44, 129–142. [Google Scholar] [CrossRef] [Green Version]
  11. Matsuzawa, T. The metagenome approach: A new resource for glycosidases. Trends Glycosci. Glycotechnol. 2019, 31, E15–E20. [Google Scholar] [CrossRef]
  12. Teze, D.; Hendrickx, J.; Czjzek, M.; Ropartz, D.; Sanejouand, Y.-H.; Tran, V.; Tellier, C.; Dion, M. Semi-rational approach for converting a GH1 β-glycosidase into a β-transglycosidase. Protein Eng. Des. Sel. 2014, 27, 13–19. [Google Scholar] [CrossRef] [Green Version]
  13. Daudé, D.; Vergès, A.; Cambon, E.; Emond, S.; Tranier, S.; André, I.; Remaud-Siméon, M. Neutral genetic drift-based engineering of a sucrose-utilizing enzyme toward glycodiversification. ACS Catal. 2019, 9, 1241–1252. [Google Scholar] [CrossRef] [Green Version]
  14. Zeuner, B.; Teze, D.; Muschiol, J.; Meyer, A.S. Synthesis of human milk oligosaccharides: Protein Engineering strategies for improved enzymatic transglycosylation. Molecules 2019, 24, 2033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Tran, L.T.; Blay, V.; Luang, S.; Eurtivong, C.; Choknud, S.; González-Díaz, H.; Cairns, J.R.K. Engineering faster transglycosidases and their acceptor specificity. Green Chem. 2019, 21, 2823–2836. [Google Scholar] [CrossRef]
  16. Albert, M.; Repetschnigg, W.; Ortner, J.; Gomes, J.; Paul, B.J.; Illaszewicz, C.; Weber, H.; Steiner, W.; Dax, K. Simultaneous detection of different glycosidase activities by 19F NMR spectroscopy. Carbohydr. Res. 2000, 327, 395–400. [Google Scholar] [CrossRef]
  17. Okuyama, M.; Mori, H.; Watanabe, K.; Kimura, A.; Chiba, S. α-Glucosidase mutant catalyzes “α-Glycosynthase”-type reaction. Biosci. Biotechnol. Biochem. 2002, 66, 928–933. [Google Scholar] [CrossRef] [Green Version]
  18. Higuchi, Y.; Eshima, Y.; Huang, Y.; Kinoshita, T.; Sumiyoshi, W.; Nakakita, S.-I.; Takegawa, K. Highly efficient transglycosylation of sialo-complex-type oligosaccharide using Coprinopsis cinerea endoglycosidase and sugar oxazoline. Biotechnol. Lett. 2017, 39, 157–162. [Google Scholar] [CrossRef]
  19. Cobucci-Ponzano, B.; Conte, F.; Bedini, E.; Corsaro, M.M.; Parrilli, M.; Sulzenbacher, G.; Lipski, A.; Dal Piaz, F.; Lepore, L.; Rossi, M.; et al. beta-Glycosyl azides as substrates for alpha-glycosynthases: Preparation of efficient Alpha-L-Fucosynthases. Chem. Biol. 2009, 16, 1097–1108. [Google Scholar] [CrossRef] [Green Version]
  20. Hayes, M.R.; Pietruszka, J. Synthesis of glycosides by glycosynthases. Molecules 2017, 22, 1434. [Google Scholar] [CrossRef]
  21. Cori, C.F.; Cori, G.T. Mechanism of formation of hexosemonophosphate in muscle and isolation of a new phosphate ester. Proc. Soc. Exp. Biol. Med. 1936, 34, 702–705. [Google Scholar] [CrossRef]
  22. Alexander, J.K. Characteristics of cellobiose phosphorylase. J. Bacteriol. 1961, 81, 903–910. [Google Scholar] [CrossRef] [Green Version]
  23. Koshland, D.E. Stereochemistry and the mechanism of enzymatic reactions. Biol. Rev. 1953, 28, 416–436. [Google Scholar] [CrossRef]
  24. Hamura, K.; Saburi, W.; Abe, S.; Morimoto, N.; Taguchi, H.; Mori, H.; Matsui, H. Enzymatic characteristics of cellobiose phosphorylase from Ruminococcus albus NE1 and kinetic mechanism of unusual substrate inhibition in reverse phosphorolysis. Biosci. Biotechnol. Biochem. 2012, 76, 812–818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nihira, T.; Nakai, H.; Kitaoka, M. 3-O-α-D-glucopyranosyl-L-rhamnose phosphorylase from Clostridium phytofermentans. Carbohydr. Res. 2012, 350, 94–97. [Google Scholar] [CrossRef] [PubMed]
  26. Chao, C.; Wim, S.; Tom, D. Characterization of β-Galactoside phosphorylases with diverging acceptor specificities. Enzym. Microb. Technol. 2011, 49, 59–65. [Google Scholar] [CrossRef] [PubMed]
  27. Hidaka, Y.; Hatada, Y.; Akita, M.; Yoshida, M.; Nakamura, N.; Takada, M.; Nakakuki, T.; Ito, S.; Horikoshi, K. Maltose phosphorylase from a deep-sea Paenibacillus Sp.: Enzymatic properties and nucleotide and amino-acid sequences. Enzym. Microb. Technol. 2005, 37, 185–194. [Google Scholar] [CrossRef]
  28. Franceus, J.; Pinel, D.; Desmet, T. Glucosylglycerate phosphorylase, an enzyme with novel specificity involved in compatible solute metabolism. Appl. Environ. Microbiol. 2017, 83, e01434-17. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, N.; Fosses, A.; Kampik, C.; Parsiegla, G.; Denis, Y.; Vita, N.; Fierobe, H.-P.; Perret, S. In Vitro and in vivo exploration of the cellobiose and cellodextrin phosphorylases panel in Ruminiclostridium Cellulolyticum: Implication for cellulose catabolism. Biotechnol. Biofuels 2019, 12, 208. [Google Scholar] [CrossRef] [Green Version]
  30. Macdonald, S.S.; Patel, A.; Larmour, V.L.C.; Morgan-Lang, C.; Hallam, S.J.; Mark, B.L.; Withers, S.G. Structural and mechanistic analysis of a β-glycoside phosphorylase identified by screening a metagenomic library. J. Biol. Chem. 2018, 293, 3451–3467. [Google Scholar] [CrossRef] [Green Version]
  31. Macdonald, S.S.; Blaukopf, M.; Withers, S.G. N-Acetylglucosaminidases from CAZy family GH3 are really glycoside phosphorylases, thereby explaining their use of histidine as an acid/base catalyst in place of glutamic Acid. J. Biol. Chem. 2015, 290, 4887–4895. [Google Scholar] [CrossRef] [Green Version]
  32. Van den Broek, L.A.M.; van Boxtel, E.L.; Kievit, R.P.; Verhoef, R.; Beldman, G.; Voragen, A.G.J. Physico-chemical and transglucosylation properties of recombinant sucrose phosphorylase from Bifidobacterium adolescentis DSM20083. Appl. Microbiol. Biotechnol. 2004, 65, 219–227. [Google Scholar] [CrossRef]
  33. Kullin, B.; Abratt, V.R.; Reid, S.J. A functional analysis of the Bifidobacterium longum cscA and scrP genes in sucrose utilization. Appl. Microbiol. Biotechnol. 2006, 72, 975–981. [Google Scholar] [CrossRef]
  34. Kim, M.; Kwon, T.; Joo Lee, H.; Heon Kim, K.; Kyun Chung, D.; Eog Ji, G.; Byeon, E.-S.; Lee, J.-H. Cloning and expression of sucrose phosphorylase gene from Bifidobacterium longum in E. coli and characterization of the recombinant enzyme. Biotechnol. Lett. 2003, 25, 1211–1217. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, H.; Sun, X.; Li, W.; Li, T.; Li, S.; Kitaoka, M. Expression and Characterization of Recombinant Sucrose Phosphorylase. Protein J. 2018, 37, 93–100. [Google Scholar] [CrossRef] [PubMed]
  36. Barrangou, R.; Altermann, E.; Hutkins, R.; Cano, R.; Klaenhammer, T.R. Functional and comparative genomic analyses of an operon involved in fructooligosaccharide utilization by Lactobacillus acidophilus. Proc. Natl. Acad. Sci. USA 2003, 100, 8957–8962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Goedl, C.; Schwarz, A.; Minani, A.; Nidetzky, B. Recombinant sucrose phosphorylase from Leuconostoc mesenteroides: Characterization, kinetic studies of transglucosylation, and application of immobilised enzyme for production of α-d-glucose 1-phosphate. J. Biotechnol. 2007, 129, 77–86. [Google Scholar] [CrossRef] [PubMed]
  38. Lee, J.-H.; Moon, Y.-H.; Kim, N.; Kim, Y.-M.; Kang, H.-K.; Jung, J.-Y.; Abada, E.; Kang, S.-S.; Kim, D. Cloning and expression of the sucrose phosphorylase gene from Leuconostoc mesenteroides in Escherichia coli. Biotechnol. Lett. 2008, 30, 749–754. [Google Scholar] [CrossRef]
  39. Kawasaki, H.; Nakamura, N.; Ohmori, M.; Sakai, T. Cloning and expression in Escherichia coli of sucrose phosphorylase gene from Leuconostoc mesenteroides No. 165. Biosci. Biotechnol. Biochem. 1996, 60, 322–324. [Google Scholar] [CrossRef]
  40. Lee, J.-H.; Yoon, S.-H.; Nam, S.-H.; Moon, Y.-H.; Moon, Y.-Y.; Kim, D. Molecular cloning of a gene encoding the sucrose phosphorylase from Leuconostoc mesenteroides B-1149 and the expression in Escherichia coli. Enzym. Microb. Technol. 2006, 39, 612–620. [Google Scholar] [CrossRef]
  41. Teixeira, J.S.; Abdi, R.; Su, M.S.-W.; Schwab, C.; Gänzle, M.G. Functional characterization of sucrose phosphorylase and scrR, a regulator of sucrose metabolism in Lactobacillus reuteri. Food Microbiol. 2013, 36, 432–439. [Google Scholar] [CrossRef]
  42. Franceus, J.; Decuyper, L.; D’hooghe, M.; Desmet, T. Exploring the sequence diversity in glycoside hydrolase family 13_18 reveals a novel glucosylglycerol phosphorylase. Appl. Microbiol. Biotechnol. 2018, 102, 3183–3191. [Google Scholar] [CrossRef]
  43. Trethewey, R.N.; Fernie, A.R.; Bachmann, A.; Fleischer-Notter, H.; Geigenberger, P.; Willmitzer, L. Expression of a bacterial sucrose phosphorylase in potato tubers results in a glucose-independent induction of glycolysis: Sugar signalling in heterotrophic plant tissue. Plant Cell Environ. 2001, 24, 357–365. [Google Scholar] [CrossRef]
  44. Tauzin, A.S.; Bruel, L.; Laville, E.; Nicoletti, C.; Navarro, D.; Henrissat, B.; Perrier, J.; Potocki-Veronese, G.; Giardina, T.; Lafond, M. Sucrose 6F-phosphate phosphorylase: A novel insight in the human gut microbiome. Microb. Genom. 2019, 5, e000253. [Google Scholar] [CrossRef] [PubMed]
  45. Russell, R.R.; Mukasa, H.; Shimamura, A.; Ferretti, J.J. Streptococcus mutans gtfA gene specifies sucrose phosphorylase. Infect. Immun. 1988, 56, 2763–2765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Shemesh, M.; Tam, A.; Steinberg, D. Differential gene expression profiling of Streptococcus mutans cultured under biofilm and planktonic conditions. Microbiology 2007, 153, 1307–1317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Goedl, C.; Nidetzky, B. Sucrose Phosphorylase Harbouring a Redesigned, Glycosyltransferase-Like Active Site Exhibits Retaining Glucosyl Transfer in the Absence of a Covalent Intermediate. ChemBioChem 2009, 10, 2333–2337. [Google Scholar] [CrossRef]
  48. Verhaeghe, T.; Aerts, D.; Diricks, M.; Soetaert, W.; Desmet, T. The quest for a thermostable sucrose phosphorylase reveals sucrose 6′-Phosphate phosphorylase as a novel specificity. Appl. Microbiol. Biotechnol. 2014, 98, 7027–7037. [Google Scholar] [CrossRef]
  49. Du, L.; Yang, H.; Huo, Y.; Wei, H.; Xu, Y.; Wei, Y.; Huang, R. A novel sucrose phosphorylase from the metagenomes of sucrose-rich environment: Isolation and characterization. World J. Microbiol. Biotechnol. 2012, 28, 2871–2878. [Google Scholar] [CrossRef]
  50. Franceus, J.; Capra, N.; Desmet, T.; Thunnissen, A.-M.W.H. Structural comparison of a promiscuous and a highly specific sucrose 6F-phosphate phosphorylase. Int. J. Mol. Sci. 2019, 20, 3906. [Google Scholar] [CrossRef] [Green Version]
  51. Nakai, H.; Baumann, M.J.; Petersen, B.O.; Westphal, Y.; Schols, H.; Dilokpimol, A.; Hachem, M.A.; Lahtinen, S.J.; Duus, J.Ø.; Svensson, B. The maltodextrin transport system and metabolism in Lactobacillus acidophilus NCFM and production of novel α-glucosides through reverse phosphorolysis by maltose phosphorylase: Maltose phosphorylase and reverse phosphorolysis. FEBS J. 2009, 276, 7353–7365. [Google Scholar] [CrossRef]
  52. Nihira, T.; Saito, Y.; Kitaoka, M.; Otsubo, K.; Nakai, H. Identification of Bacillus selenitireducens MLS10 maltose phosphorylase possessing synthetic ability for branched α-D-glucosyl trisaccharides. Carbohydr. Res. 2012, 360, 25–30. [Google Scholar] [CrossRef]
  53. Inoue, Y.; Yasutake, N.; Oshima, Y.; Yamamoto, Y.; Tomita, T.; Miyoshi, S.; Yatake, T. Cloning of the maltose phosphorylase gene from Bacillus sp. strain RK-1 and efficient production of the cloned gene and the trehalose phosphorylase gene from Bacillus stearothermophilus SK-1 in Bacillus subtilis. Biosci. Biotechnol. Biochem. 2002, 66, 2594–2599. [Google Scholar] [CrossRef]
  54. Ehrmann, M.A.; Vogel, R.F. Maltose metabolism of Lactobacillus sanfranciscensis: Cloning and heterologous expression of the key enzymes, maltose phosphorylase and phosphoglucomutase. FEMS Microbiol. Lett. 1998, 169, 81–86. [Google Scholar] [CrossRef] [PubMed]
  55. Mokhtari, A.; Blancato, V.S.; Repizo, G.D.; Henry, C.; Pikis, A.; Bourand, A.; de Fátima Álvarez, M.; Immel, S.; Mechakra-Maza, A.; Hartke, A.; et al. Enterococcus faecalis utilizes maltose by connecting two incompatible metabolic routes via a novel maltose 6′-phosphate phosphatase (MapP): Novel mode of maltose utilization by Enterococcus faecalis. Mol. Microbiol. 2013, 88, 234–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Hüwel, S.; Haalck, L.; Conrath, N.; Spener, F. Maltose phosphorylase from Lactobacillus brevis: Purification, characterization, and application in a biosensor for ortho-phosphate. Enzym. Microb. Technol. 1997, 21, 413–420. [Google Scholar] [CrossRef]
  57. Gao, Y.; Saburi, W.; Taguchi, Y.; Mori, H. Biochemical characteristics of maltose phosphorylase MalE from Bacillus sp. AHU2001 and chemoenzymatic synthesis of oligosaccharides by the enzyme. Biosci. Biotechnol. Biochem. 2019, 83, 2097–2109. [Google Scholar] [CrossRef]
  58. Chaen, H.; Nakada, T.; Nishimoto, T.; Kuroda, N.; Fukuda, S.; Sugimoto, T.; Kurimoto, M.; Tsujisaka, Y. Purification and characterization of thermostable Trehalose Phosphorylase from Thermoanaerobium brockii. J. Appl. Glycosci. 1999, 46, 399–405. [Google Scholar] [CrossRef]
  59. Inoue, Y.; Ishii, K.; Tomita, T.; Yatake, T.; Fukui, F. Characterization of Trehalose Phosphorylase from Bacillus stearothermophilus SK-1 and nucleotide sequence of the corresponding gene. Biosci. Biotechnol. Biochem. 2002, 66, 1835–1843. [Google Scholar] [CrossRef] [Green Version]
  60. Van der Borght, J.; Chen, C.; Hoflack, L.; Van Renterghem, L.; Desmet, T.; Soetaert, W. Enzymatic properties and substrate specificity of the trehalose phosphorylase from Caldanaerobacter subterraneus. Appl. Environ. Microbiol. 2011, 77, 6939–6944. [Google Scholar] [CrossRef] [Green Version]
  61. De Beul, E.; Jongbloet, A.; Franceus, J.; Desmet, T. Discovery of a kojibiose hydrolase by analysis of specificity-determining correlated positions in glycoside hydrolase family 65. Molecules 2021, 26, 6321. [Google Scholar] [CrossRef]
  62. Andersson, U.; Levander, F.; Rådström, P. Trehalose-6-phosphate phosphorylase is part of a novel metabolic pathway for trehalose utilization in Lactococcus lactis. J. Biol. Chem. 2001, 276, 42707–42713. [Google Scholar] [CrossRef] [Green Version]
  63. Mukherjee, K.; Narindoshvili, T.; Raushel, F.M. Discovery of a Kojibiose Phosphorylase in Escherichia coli K-12. Biochemistry 2018, 57, 2857–2867. [Google Scholar] [CrossRef]
  64. Yamamoto, T.; Nishio-Kosaka, M.; Izawa, S.; Aga, H.; Nishimoto, T.; Chaen, H.; Fukuda, S. Enzymatic Properties of recombinant kojibiose phosphorylase from Caldicellulosiruptor saccharolyticus ATCC43494. Biosci. Biotechnol. Biochem. 2011, 75, 1208–1210. [Google Scholar] [CrossRef] [PubMed]
  65. Nihira, T.; Nakai, H.; Chiku, K.; Kitaoka, M. Discovery of nigerose phosphorylase from Clostridium phytofermentans. Appl. Microbiol. Biotechnol. 2012, 93, 1513–1522. [Google Scholar] [CrossRef] [PubMed]
  66. Nihira, T.; Nishimoto, M.; Nakai, H.; Ohtsubo, K.; Kitaoka, M. Characterization of Two α-1,3-Glucoside Phosphorylases from Clostridium phytofermentans. J. Appl. Glycosci. 2014, 61, 59–66. [Google Scholar] [CrossRef] [Green Version]
  67. Nihira, T.; Saito, Y.; Ohtsubo, K.; Nakai, H.; Kitaoka, M. 2-O-α-D-Glucosylglycerol phosphorylase from Bacillus selenitireducens MLS10 possessing hydrolytic activity on β-D-Glucose 1-Phosphate. PLoS ONE 2014, 9, e86548. [Google Scholar] [CrossRef] [Green Version]
  68. Yernool, D.A.; McCarthy, J.K.; Eveleigh, D.E.; Bok, J.-D. Cloning and characterization of the glucooligosaccharide catabolic pathway β-Glucan glucohydrolase and cellobiose phosphorylase in the marine hyperthermophile Thermotoga neapolitana. J. Bacteriol. 2000, 182, 5172–5179. [Google Scholar] [CrossRef] [Green Version]
  69. Reichenbecher, M.; Lottspeich, F.; Bronnenmeier, K. Purification and properties of a cellobiose phosphorylase (CepA) and a cellodextrin phosphorylase (CepB) from the cellulolytic thermophile Clostridium Stercorarium. Eur. J. Biochem. 1997, 247, 262–267. [Google Scholar] [CrossRef]
  70. Rajashekhara, E.; Kitaoka, M.; Kim, Y.-K.; Hayashi, K. Characterization of a cellobiose phosphorylase from a Hyperthermophilic Eubacterium, Thermotoga maritima MSB8. Biosci. Biotechnol. Biochem. 2002, 66, 2578–2586. [Google Scholar] [CrossRef]
  71. Kim, Y.-K.; Kitaoka, M.; Krishnareddy, M.; Mori, Y.; Hayashi, K. Kinetic studies of a recombinant cellobiose phosphorylase (CBP) of the Clostridium thermocellum YM4 strain expressed in Escherichia coli. J. Biochem. 2002, 132, 197–203. [Google Scholar] [CrossRef]
  72. Nidetzky, B.; Griessler, R.; Schwarz, A.; Splechtna, B. Cellobiose phosphorylase from Cellulomonas uda: Gene cloning and expression in Escherichia coli, and application of the recombinant enzyme in a ‘glycosynthase-type’ reaction. J. Mol. Catal. B Enzym. 2004, 29, 241–248. [Google Scholar] [CrossRef]
  73. Ha, S.-J.; Galazka, J.M.; Joong Oh, E.; Kordić, V.; Kim, H.; Jin, Y.-S.; Cate, J.H.D. Energetic benefits and rapid cellobiose fermentation by Saccharomyces cerevisiae expressing cellobiose phosphorylase and mutant cellodextrin transporters. Metab. Eng. 2013, 15, 134–143. [Google Scholar] [CrossRef]
  74. Nakai, H.; Hachem, M.A.; Petersen, B.O.; Westphal, Y.; Mannerstedt, K.; Baumann, M.J.; Dilokpimol, A.; Schols, H.A.; Duus, J.Ø.; Svensson, B. Efficient chemoenzymatic oligosaccharide synthesis by reverse phosphorolysis using cellobiose phosphorylase and cellodextrin phosphorylase from Clostridium thermocellum. Biochimie 2010, 92, 1818–1826. [Google Scholar] [CrossRef] [PubMed]
  75. Fosses, A.; Maté, M.; Franche, N.; Liu, N.; Denis, Y.; Borne, R.; de Philip, P.; Fierobe, H.-P.; Perret, S. A seven-gene cluster in Ruminiclostridium cellulolyticum is essential for signalization, uptake and catabolism of the degradation products of cellulose hydrolysis. Biotechnol. Biofuels 2017, 10, 250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Alexander, J.K. Purification and specificity of cellobiose phosphorylase from Clostridium thermocellum. J. Biol. Chem. 1968, 243, 2899–2904. [Google Scholar] [CrossRef]
  77. Sasaki, T.; Tanaka, T.; Nakagawa, S.; Kainuma, K. Purification and properties of Cellvibrio gilvus cellobiose phosphorylase. Biochem. J. 1983, 209, 803–807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Dakhova, O.N.; Kurepina, N.E.; Zverlov, V.V.; Svetlichnyi, V.A.; Velikodvorskaya, G.A. Cloning and expression in Escherichia coli of Thermotoga neapolitana genes coding for enzymes of carbohydrate substrate degradation. Biochem. Biophys. Res. Commun. 1993, 194, 1359–1364. [Google Scholar] [CrossRef] [PubMed]
  79. Macdonald, S.S.; Armstrong, Z.; Morgan-Lang, C.; Osowiecka, M.; Robinson, K.; Hallam, S.J.; Withers, S.G. Development and application of a high-throughput functional metagenomic screen for glycoside phosphorylases. Cell Chem. Biol. 2019, 26, 1001.e5–1012.e5. [Google Scholar] [CrossRef]
  80. Sawano, T.; Saburi, W.; Hamura, K.; Matsui, H.; Mori, H. Characterization of Ruminococcus albus cellodextrin phosphorylase and identification of a key phenylalanine residue for acceptor specificity and affinity to the phosphate group. FEBS J. 2013, 280, 4463–4473. [Google Scholar] [CrossRef]
  81. Iñón de Iannino, N.; Briones, G.; Tolmasky, M.; Ugalde, R.A. Molecular cloning and characterization of cgs, the Brucella abortus Cyclic β(1-2) glucan synthetase gene: Genetic complementation of Rhizobium meliloti ndvB and Agrobacterium tumefaciens chvB mutants. J. Bacteriol. 1998, 180, 4392–4400. [Google Scholar] [CrossRef] [Green Version]
  82. Kitaoka, M.; Matsuoka, Y.; Mori, K.; Nishimoto, M.; Hayashi, K. Characterization of a bacterial laminaribiose phosphorylase. Biosci. Biotechnol. Biochem. 2012, 76, 343–348. [Google Scholar] [CrossRef] [Green Version]
  83. Kuhaudomlarp, S.; Walpole, S.; Stevenson, C.E.M.; Nepogodiev, S.A.; Lawson, D.M.; Angulo, J.; Field, R.A. Unravelling the Specificity of Laminaribiose Phosphorylase from Paenibacillus sp. YM-1 towards Donor Substrates Glucose/Mannose 1-Phosphate by Using X-ray Crystallography and Saturation Transfer Difference NMR Spectroscopy. ChemBioChem 2019, 20, 181–192. [Google Scholar] [CrossRef]
  84. Park, J.K.; Keyhani, N.O.; Roseman, S. Chitin Catabolism In the marine bacterium Vibrio Furnissii identification, molecular cloning, and characterization of An,N′-Diacetylchitobiose phosphorylase. J. Biol. Chem. 2000, 275, 33077–33083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Honda, Y.; Kitaoka, M.; Hayashi, K. Reaction mechanism of chitobiose phosphorylase from Vibrio proteolyticus: Identification of family 36 glycosyltransferase in Vibrio. Biochem. J. 2004, 377, 225–232. [Google Scholar] [CrossRef] [PubMed]
  86. De Doncker, M.; De Graeve, C.; Franceus, J.; Beerens, K.; Křen, V.; Pelantová, H.; Vercauteren, R.; Desmet, T. Exploration of GH94 sequence space for enzyme discovery reveals a novel Glucosylgalactose phosphorylase specificity. Chembiochem 2021, 22, 3319–3325. [Google Scholar] [CrossRef] [PubMed]
  87. Nakajima, M.; Toyoizumi, H.; Abe, K.; Nakai, H.; Taguchi, H.; Kitaoka, M. 1,2-β-Oligoglucan Phosphorylase from Listeria innocua. PLoS ONE 2014, 9, e92353. [Google Scholar] [CrossRef] [Green Version]
  88. Nakajima, M.; Tanaka, N.; Furukawa, N.; Nihira, T.; Kodutsumi, Y.; Takahashi, Y.; Sugimoto, N.; Miyanaga, A.; Fushinobu, S.; Taguchi, H.; et al. Mechanistic insight into the substrate specificity of 1,2-β-oligoglucan phosphorylase from LachnoClostridium phytofermentans. Sci. Rep. 2017, 7, 42671. [Google Scholar] [CrossRef]
  89. Kawaguchi, T.; Ikeuchi, Y.; Tsutsumi, N.; Kan, A.; Sumitani, J.-I.; Arai, M. Cloning, nucleotide sequence, and expression of the Clostridium thermocellum cellodextrin phosphorylase gene and its application to synthesis of cellulase inhibitors. J. Ferment. Bioeng. 1998, 85, 144–149. [Google Scholar] [CrossRef]
  90. Nihira, T.; Saito, Y.; Nishimoto, M.; Kitaoka, M.; Igarashi, K.; Ohtsubo, K.; Nakai, H. Discovery of cellobionic acid phosphorylase in cellulolytic bacteria and fungi. FEBS Lett. 2013, 587, 3556–3561. [Google Scholar] [CrossRef] [Green Version]
  91. Nam, Y.-W.; Nihira, T.; Arakawa, T.; Saito, Y.; Kitaoka, M.; Nakai, H.; Fushinobu, S. Crystal structure and substrate recognition of cellobionic acid phosphorylase, which plays a key role in oxidative cellulose degradation by microbes. J. Biol. Chem. 2015, 290, 18281–18292. [Google Scholar] [CrossRef] [Green Version]
  92. Nihira, T.; Saito, Y.; Kitaoka, M.; Nishimoto, M.; Otsubo, K.; Nakai, H. Characterization of a laminaribiose phosphorylase from Acholeplasma laidlawii PG-8A and production of 1,3-β-d-glucosyl disaccharides. Carbohydr. Res. 2012, 361, 49–54. [Google Scholar] [CrossRef]
  93. Nakajima, M.; Nishimoto, M.; Kitaoka, M. Characterization of d-galactosyl-β1→4-l-rhamnose phosphorylase from Opitutus terrae. Enzym. Microb. Technol. 2010, 46, 315–319. [Google Scholar] [CrossRef]
  94. Nakajima, M.; Nishimoto, M.; Kitaoka, M. Characterization of three β-Galactoside Phosphorylases from Clostridium phytofermentans. J. Biol. Chem. 2009, 284, 19220–19227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Nakajima, M.; Kitaoka, M. Identification of Lacto- N -Biose I Phosphorylase from Vibrio vulnificus CMCP6. Appl. Environ. Microbiol. 2008, 74, 6333–6337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Pichler, M.J.; Yamada, C.; Shuoker, B.; Alvarez-Silva, C.; Gotoh, A.; Leth, M.L.; Schoof, E.; Katoh, T.; Sakanaka, M.; Katayama, T.; et al. Butyrate producing colonic Clostridiales metabolise human milk oligosaccharides and cross feed on mucin via conserved pathways. Nat. Commun. 2020, 11, 3285. [Google Scholar] [CrossRef] [PubMed]
  97. Kitaoka, M.; Tian, J.; Nishimoto, M. Novel putative galactose operon involving Lacto- N -Biose phosphorylase in Bifidobacterium longum. Appl. Environ. Microbiol. 2005, 71, 3158–3162. [Google Scholar] [CrossRef] [Green Version]
  98. Nishimoto, M.; Kitaoka, M. Identification of the putative proton donor residue of Lacto- N -biose phosphorylase (EC 2.4.1.211). Biosci. Biotechnol. Biochem. 2007, 71, 1587–1591. [Google Scholar] [CrossRef] [Green Version]
  99. Nakajima, M.; Nishimoto, M.; Kitaoka, M. Characterization of β-1,3-galactosyl-N-acetylhexosamine phosphorylase from Propionibacterium acnes. Appl. Microbiol. Biotechnol. 2009, 83, 109–115. [Google Scholar] [CrossRef]
  100. Tsuda, T.; Nihira, T.; Chiku, K.; Suzuki, E.; Arakawa, T.; Nishimoto, M.; Kitaoka, M.; Nakai, H.; Fushinobu, S. Characterization and crystal structure determination of β-1,2-mannobiose phosphorylase from Listeria innocua. FEBS Lett. 2015, 589, 3816–3821. [Google Scholar] [CrossRef]
  101. Chiku, K.; Nihira, T.; Suzuki, E.; Nishimoto, M.; Kitaoka, M.; Ohtsubo, K.; Nakai, H. Discovery of two β-1,2-Mannoside Phosphorylases showing different chain-length specificities from Thermoanaerobacter sp. X-514. PLoS ONE 2014, 9, e114882. [Google Scholar] [CrossRef]
  102. Awad, F.N.; Laborda, P.; Wang, M.; Lu, A.M.; Li, Q.; Cai, Z.P.; Liu, L.; Voglmeir, J. Discovery and biochemical characterization of a mannose phosphorylase catalyzing the synthesis of novel β-1,3-mannosides. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2017, 1861, 3231–3237. [Google Scholar] [CrossRef]
  103. Saburi, W.; Tanaka, Y.; Muto, H.; Inoue, S.; Odaka, R.; Nishimoto, M.; Kitaoka, M.; Mori, H. Functional reassignment of Cellvibrio vulgaris EpiA to cellobiose 2-epimerase and an evaluation of the biochemical functions of the 4-O-β-d-mannosyl-d-glucose phosphorylase-like protein, UnkA. Biosci. Biotechnol. Biochem. 2015, 79, 969–977. [Google Scholar] [CrossRef]
  104. Kawahara, R.; Saburi, W.; Odaka, R.; Taguchi, H.; Ito, S.; Mori, H.; Matsui, H. Metabolic Mechanism of Mannan in a Ruminal Bacterium, Ruminococcus albus, Involving Two Mannoside Phosphorylases and Cellobiose 2-Epimerase. J. Biol. Chem. 2012, 287, 42389–42399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Senoura, T.; Ito, S.; Taguchi, H.; Higa, M.; Hamada, S.; Matsui, H.; Ozawa, T.; Jin, S.; Watanabe, J.; Wasaki, J.; et al. New microbial mannan catabolic pathway that involves a novel mannosylglucose phosphorylase. Biochem. Biophys. Res. Commun. 2011, 408, 701–706. [Google Scholar] [CrossRef] [PubMed]
  106. La Rosa, S.L.; Leth, M.L.; Michalak, L.; Hansen, M.E.; Pudlo, N.A.; Glowacki, R.; Pereira, G.; Workman, C.T.; Arntzen, M.Ø.; Pope, P.B.; et al. The human gut Firmicute Roseburia intestinalis is a primary degrader of dietary β-mannans. Nat. Commun. 2019, 10, 905. [Google Scholar] [CrossRef] [PubMed]
  107. Jaito, N.; Saburi, W.; Odaka, R.; Kido, Y.; Hamura, K.; Nishimoto, M.; Kitaoka, M.; Matsui, H.; Mori, H. Characterization of a thermophilic 4-O-β-d-mannosyl-d-glucose phosphorylase from Rhodothermus marinus. Biosci. Biotechnol. Biochem. 2014, 78, 263–270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Nihira, T.; Suzuki, E.; Kitaoka, M.; Nishimoto, M.; Ohtsubo, K.; Nakai, H. Discovery of β-1,4-d-Mannosyl-N-acetyl-d-glucosamine phosphorylase involved in the metabolism of N-Glycans. J. Biol. Chem. 2013, 288, 27366–27374. [Google Scholar] [CrossRef] [Green Version]
  109. Ladevèze, S.; Tarquis, L.; Cecchini, D.A.; Bercovici, J.; André, I.; Topham, C.M.; Morel, S.; Laville, E.; Monsan, P.; Lombard, V.; et al. Role of glycoside phosphorylases in mannose foraging by human gut bacteria. J. Biol. Chem. 2013, 288, 32370–32383. [Google Scholar] [CrossRef] [Green Version]
  110. Grimaud, F.; Pizzut-Serin, S.; Tarquis, L.; Ladevèze, S.; Morel, S.; Putaux, J.-L.; Potocki-Veronese, G. In Vitro Synthesis and crystallization of β-1,4-Mannan. Biomacromolecules 2019, 20, 846–853. [Google Scholar] [CrossRef]
  111. Chekan, J.R.; Kwon, I.H.; Agarwal, V.; Dodd, D.; Revindran, V.; Mackie, R.I.; Cann, I.; Nair, S.K. Structural and biochemical basis for mannan utilization by Caldanaerobius polysaccharolyticus Strain ATCC BAA-17. J. Biol. Chem. 2014, 289, 34965–34977. [Google Scholar] [CrossRef] [Green Version]
  112. Li, A.; Laville, E.; Tarquis, L.; Lombard, V.; Ropartz, D.; Terrapon, N.; Henrissat, B.; Guieysse, D.; Esque, J.; Durand, J.; et al. Analysis of the diversity of the glycoside hydrolase family 130 in mammal gut microbiomes reveals a novel mannoside-phosphorylase function. Microb. Genom. 2020, 6, mgen000404. [Google Scholar] [CrossRef]
  113. Ollivier, S.; Tarquis, L.; Fanuel, M.; Li, A.; Durand, J.; Laville, E.; Potocki-Veronese, G.; Ropartz, D.; Rogniaux, H. Anomeric retention of carbohydrates in multistage cyclic ion mobility (IMSn): De novo structural elucidation of enzymatically produced Mannosides. Anal. Chem. 2021, 93, 6254–6261. [Google Scholar] [CrossRef]
  114. Kuhaudomlarp, S.; Stevenson, C.E.M.; Lawson, D.M.; Field, R.A. The structure of a GH149 β-(1→3) glucan phosphorylase reveals a new surface oligosaccharide binding site and additional domains that are absent in the disaccharide-specific GH94 glucose-β-(1→3)-glucose (Laminaribiose) phosphorylase. Proteins 2019, 87, 885–892. [Google Scholar] [CrossRef] [PubMed]
  115. Kuhaudomlarp, S.; Patron, N.J.; Henrissat, B.; Rejzek, M.; Saalbach, G.; Field, R.A. Identification of Euglena gracilis β-1,3-glucan phosphorylase and establishment of a new glycoside hydrolase (GH) family GH149. J. Biol. Chem. 2018, 293, 2865–2876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Kuhaudomlarp, S.; Pergolizzi, G.; Patron, N.J.; Henrissat, B.; Field, R.A. Unraveling the subtleties of β-(1→3)-glucan phosphorylase specificity in the GH94, GH149, and GH161 glycoside hydrolase families. J. Biol. Chem. 2019, 294, 6483–6493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Yamamoto, Y.; Kawashima, D.; Hashizume, A.; Hisamatsu, M.; Isono, N. Purification and characterization of 1,3-β-d-Glucan Phosphorylase from Ochromonas danica. Biosci. Biotechnol. Biochem. 2013, 77, 1949–1954. [Google Scholar] [CrossRef]
  118. Saito, K.; Kase, T.; Takahashi, E.; Takahashi, E.; Horinouchi, S. Purification and characterization of a trehalose synthase from the basidiomycete Grifola frondosa. Appl. Environ. Microbiol. 1998, 64, 4340–4345. [Google Scholar] [CrossRef] [Green Version]
  119. Han, S.-E.; Kwon, H.-B.; Lee, S.-B.; Yi, B.-Y.; Murayama, I.; Kitamoto, Y.; Byun, M.-O. Cloning and characterization of a gene encoding trehalose phosphorylase (TP) from Pleurotus sajor-caju. Protein Expr. Purif. 2003, 30, 194–202. [Google Scholar] [CrossRef]
  120. Eis, C.; Nidetzky, B. Characterization of trehalose phosphorylase from Schizophyllum commune. Biochem. J. 1999, 341 Pt 2, 385–393. [Google Scholar] [CrossRef]
  121. Goedl, C.; Griessler, R.; Schwarz, A.; Nidetzky, B. Structure–function relationships for Schizophyllum commune trehalose phosphorylase and their implications for the catalytic mechanism of family GT-4 glycosyltransferases. Biochem. J. 2006, 397, 491–500. [Google Scholar] [CrossRef] [Green Version]
  122. Strazzulli, A.; Cobucci-Ponzano, B.; Iacono, R.; Giglio, R.; Maurelli, L.; Curci, N.; Schiano-di-Cola, C.; Santangelo, A.; Contursi, P.; Lombard, V.; et al. Discovery of hyperstable carbohydrate-active enzymes through metagenomics of extreme environments. FEBS J. 2020, 287, 1116–1137. [Google Scholar] [CrossRef]
  123. Lee, H.-S.; Shockley, K.R.; Schut, G.J.; Conners, S.B.; Montero, C.I.; Johnson, M.R.; Chou, C.-J.; Bridger, S.L.; Wigner, N.; Brehm, S.D.; et al. Transcriptional and biochemical analysis of starch metabolism in the Hyperthermophilic archaeon Pyrococcus furiosus. J. Bacteriol. 2006, 188, 2115–2125. [Google Scholar] [CrossRef] [Green Version]
  124. Mizanur, R.M.; Griffin, A.K.K.; Pohl, N.L. Recombinant production and biochemical characterization of a hyperthermostable α-glucan/maltodextrin phosphorylase from Pyrococcus furiosus. Archaea 2008, 2, 169–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Xavier, K.B.; Peist, R.; Kossmann, M.; Boos, W.; Santos, H. Maltose Metabolism in the Hyperthermophilic archaeon Thermococcus litoralis: Purification and characterization of key enzymes. J. Bacteriol. 1999, 181, 3358–3367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Wilson, C.M.; Rodriguez, M.; Johnson, C.M.; Martin, S.L.; Chu, T.M.; Wolfinger, R.D.; Hauser, L.J.; Land, M.L.; Klingeman, D.M.; Syed, M.H.; et al. Global transcriptome analysis of Clostridium thermocellum ATCC 27405 during growth on dilute acid pretreated Populus and switchgrass. Biotechnol. Biofuels 2013, 6, 179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. O’Neill, E.C.; Rashid, A.M.; Stevenson, C.E.M.; Hetru, A.-C.; Gunning, A.P.; Rejzek, M.; Nepogodiev, S.A.; Bornemann, S.; Lawson, D.M.; Field, R.A. Sugar-coated sensor chip and nanoparticle surfaces for the in vitro enzymatic synthesis of starch-like materials. Chem. Sci. 2014, 5, 341–350. [Google Scholar] [CrossRef]
  128. Ugalde, J.E.; Lepek, V.; Uttaro, A.; Estrella, J.; Iglesias, A.; Ugalde, R.A. Gene organization and transcription analysis of the Agrobacterium tumefaciens glycogen (glg) operon: Two transcripts for the single phosphoglucomutase gene. J. Bacteriol. 1998, 180, 6557–6564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Bhuiyan, S.H.; Rus’d, A.A.; Kitaoka, M.; Hayashi, K. Characterization of a hyperthermostable glycogen phosphorylase from Aquifex aeolicus expressed in Escherichia coli. J. Mol. Catal. B Enzym. 2003, 22, 173–180. [Google Scholar] [CrossRef]
  130. Kiel, J.A.K.W.; Boels, J.M.; Beldman, G.; Venema, G. Glycogen in Bacillus subtilis: Molecular characterization of an operon encoding enzymes involved in glycogen biosynthesis and degradation. Mol. Microbiol. 1994, 11, 203–218. [Google Scholar] [CrossRef]
  131. Chen, S.; Liu, J.; Pei, H.; Li, J.; Zhou, J.; Xiang, H. Molecular investigation of a novel thermostable glucan phosphorylase from Thermoanaerobacter tengcongensis. Enzym. Microb. Technol. 2007, 41, 390–396. [Google Scholar] [CrossRef]
  132. Griessler, R.; Psik, B.; Schwarz, A.; Nidetzky, B. Relationships between structure, function and stability for pyridoxal 5′-phosphate-dependent starch phosphorylase from Corynebacterium callunae as revealed by reversible cofactor dissociation studies: Cofactor dissociation studies of starch phosphorylase. Eur. J. Biochem. 2004, 271, 3319–3329. [Google Scholar] [CrossRef]
  133. Seibold, G.M.; Wurst, M.; Eikmanns, B.J. Roles of maltodextrin and glycogen phosphorylases in maltose utilization and glycogen metabolism in Corynebacterium glutamicum. Microbiology 2009, 155, 347–358. [Google Scholar] [CrossRef] [Green Version]
  134. Rybak, K.V.; Slivinskaya, E.A.; Voroshilova, E.B.; Kozlov, Y.I. Method for Producing an l-Amino Acid Using a Bacterium of the Enterobacteriaceae Family Having a Pathway of Glycogen Biosynthesis Disrupted. US Patent US-0275569, 17 January 2006. [Google Scholar]
  135. Palm, D.; Goerl, R.; Burger, K.J. Evolution of catalytic and regulatory sites in phosphorylases. Nature 1985, 313, 500–502. [Google Scholar] [CrossRef] [PubMed]
  136. O’Reilly, M.; Watson, K.A.; Schinzel, R.; Palm, D.; Johnson, L.N. Oligosaccharide substrate binding in Escherichia coli maltodextrin phosphorylase. Nat. Struct. Mol. Biol. 1997, 4, 405–412. [Google Scholar] [CrossRef] [PubMed]
  137. Takata, H.; Takaha, T.; Okada, S.; Takagi, M.; Imanaka, T. Characterization of a gene cluster for glycogen biosynthesis and a heterotetrameric ADP-glucose pyrophosphorylase from Bacillus stearothermophilus. J. Bacteriol. 1997, 179, 4689–4698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Sato, Y.; Okamoto-Shibayama, K.; Azuma, T. The malQ gene is essential for starch metabolism in Streptococcus mutans. J. Oral Microbiol. 2013, 5, 21285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Koksharova, O.A.; Brandt, U.; Cerff, R. The gap1 operon of the cyanobacterium Synechococcus PCC 7942 carries a gene encoding glycogen phosphorylase and is induced under anaerobic conditions. Mikrobiologiia 2004, 73, 388–392. [Google Scholar] [CrossRef]
  140. Bibel, M.; Brettl, C.; Gosslar, U.; Kriegshäuser, G.; Liebl, W. Isolation and analysis of genes for amylolytic enzymes of the hyperthermophilic bacterium Thermotoga maritima. FEMS Microbiol. Lett. 1998, 158, 9–15. [Google Scholar] [CrossRef]
  141. Takaha, T.; Yanase, M.; Takata, H.; Okada, S. Structure and properties of Thermus aquaticus. ALPHA.-Glucan phosphorylase expressed in Escherichia coli. J. Appl. Glycosci. 2001, 48, 71–78. [Google Scholar] [CrossRef] [Green Version]
  142. Alonso-Casajús, N.; Dauvillée, D.; Viale, A.M.; Muñoz, F.J.; Baroja-Fernández, E.; Morán-Zorzano, M.T.; Eydallin, G.; Ball, S.; Pozueta-Romero, J. Glycogen Phosphorylase, the Product of the glgP Gene, catalyzes glycogen breakdown by removing glucose units from the nonreducing ends in Escherichia coli. J. Bacteriol. 2006, 188, 5266–5272. [Google Scholar] [CrossRef] [Green Version]
  143. Tsujino, S.; Shanske, S.; Valberg, S.J.; Cardinet, G.H.; Smith, B.P.; DiMauro, S. Cloning of bovine muscle glycogen phosphorylase cDNA and identification of a mutation in cattle with myophosphorylase deficiency, an animal model for McArdle’s disease. Neuromuscul. Disord. 1996, 6, 19–26. [Google Scholar] [CrossRef]
  144. Dauvillée, D.; Chochois, V.; Steup, M.; Haebel, S.; Eckermann, N.; Ritte, G.; Ral, J.-P.; Colleoni, C.; Hicks, G.; Wattebled, F.; et al. Plastidial phosphorylase is required for normal starch synthesis in Chlamydomonas reinhardtii. Plant J. 2006, 48, 274–285. [Google Scholar] [CrossRef]
  145. Rogers, P.V.; Luo, S.; Sucic, J.F.; Rutherford, C.L. Characterization and cloning of glycogen phosphorylase 1 from Dictyostelium discoideum. Biochim. Biophys. Acta (BBA)-Gene Struct. Expr. 1992, 1129, 262–272. [Google Scholar] [CrossRef]
  146. Jones, T.H.D.; Wright, B.E. Partial purification and characterization of glycogen phosphorylase from Dictyostelium discoideum. J. Bacteriol. 1970, 104, 754–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Tick, G.; Cserpán, I.; Dombrádi, V.; Mechler, B.M.; Török, I.; Kiss, I. Structural and functional characterization of the DrosophilaGlycogen phosphorylase gene. Biochem. Biophys. Res. Commun. 1999, 257, 34–43. [Google Scholar] [CrossRef] [PubMed]
  148. Wu, G.; Muller, M. Glycogen Phosphorylase Sequences from the Amitochondriate Protists, Trichomonas vaginalis, Mastigamoeba balamuthi, Entamoeba histolytica and Giardia intestinalis1. J. Eukaryot. Microbiol. 2003, 50, 366–372. [Google Scholar] [CrossRef]
  149. Takeyasu, K.; Kawase, T.; Yoshimura, S.H. Intermolecular Interaction between Na +/K +-ATPase α subunit and glycogen phosphorylase. Ann. N. Y. Acad. Sci. 2003, 986, 522–524. [Google Scholar] [CrossRef]
  150. Newgard, C.B.; Littman, D.R.; van Genderen, C.; Smith, M.; Fletterick, R.J. Human brain glycogen phosphorylase. Cloning, sequence analysis, chromosomal mapping, tissue expression, and comparison with the human liver and muscle isozymes. J. Biol. Chem. 1988, 263, 3850–3857. [Google Scholar] [CrossRef]
  151. Newgard, C.B.; Nakano, K.; Hwang, P.K.; Fletterick, R.J. Sequence analysis of the cDNA encoding human liver glycogen phosphorylase reveals tissue-specific codon usage. Proc. Natl. Acad. Sci. USA 1986, 83, 8132–8136. [Google Scholar] [CrossRef] [Green Version]
  152. Burke, J.; Hwang, P.; Anderson, L.; Lebo, R.; Gorin, F.; Fletterick, R. Intron/exon structure of the human gene for the muscle isozyme of glycogen phosphorylase. Proteins 1987, 2, 177–187. [Google Scholar] [CrossRef]
  153. Cuesta-Seijo, J.A.; Ruzanski, C.; Krucewicz, K.; Meier, S.; Hägglund, P.; Svensson, B.; Palcic, M.M. Functional and structural characterization of plastidic starch phosphorylase during barley endosperm development. PLoS ONE 2017, 12, e0175488. [Google Scholar] [CrossRef] [Green Version]
  154. Lin, C.-T.; Yeh, K.-W.; Lee, P.-D.; Su, J.-C. Primary structure of sweet potato starch phosphorylase deduced from its cDNA sequence. Plant Physiol. 1991, 95, 1250–1253. [Google Scholar] [CrossRef] [Green Version]
  155. Lu, Y.; Steichen, J.M.; Yao, J.; Sharkey, T.D. The role of cytosolic α-Glucan Phosphorylase in maltose metabolism and the comparison of amylomaltase in Arabidopsis and Escherichia coli. Plant Physiol. 2006, 142, 878–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Schliselfeld, L.H.; Danon, M.J. Inverse relationship of skeletal muscle glycogen from wild-type and genetically modified mice to their phosphorylase a activity. Biochem. Biophys. Res. Commun. 2002, 290, 874–877. [Google Scholar] [CrossRef] [PubMed]
  157. McInerney, M.; Serrano Rodriguez, G.; Pawlina, W.; Hurt, C.B.; Fletcher, B.S.; Laipis, P.J.; Frost, S.C. Glycogen phosphorylase is activated in response to glucose deprivation but is not responsible for enhanced glucose transport activity in 3T3-L1 adipocytes. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2002, 1570, 53–62. [Google Scholar] [CrossRef]
  158. Titani, K.; Koide, A.; Hermann, J.; Ericsson, L.H.; Kumar, S.; Wade, R.D.; Walsh, K.A.; Neurath, H.; Fischer, E.H. Complete amino acid sequence of rabbit muscle glycogen phosphorylase. Proc. Natl. Acad. Sci. USA 1977, 74, 4762–4766. [Google Scholar] [CrossRef] [Green Version]
  159. Fukui, T.; Shimomura, S.; Nakano, K. Potato and rabbit muscle phosphorylases: Comparative studies on the structure, function and regulation of regulatory and nonregulatory enzymes. Mol. Cell. Biochem. 1982, 42, 129–144. [Google Scholar] [CrossRef]
  160. Hwang, S.-K.; Nishi, A.; Satoh, H.; Okita, T.W. Rice endosperm-specific plastidial α-glucan phosphorylase is important for synthesis of short-chain malto-oligosaccharides. Arch. Biochem. Biophys. 2010, 495, 82–92. [Google Scholar] [CrossRef]
  161. Jha, A.B.; Dubey, R.S. Carbohydrate metabolism in growing rice seedlings under arsenic toxicity. J. Plant Physiol. 2004, 161, 867–872. [Google Scholar] [CrossRef]
  162. Walker, K.R. Characterisation of the Ovine Model of McArdle’s Disease: Development of Therapeutic Strategies. Ph.D. Thesis, Murdoch University, Perth, Australia, 2006. [Google Scholar]
  163. Tan, P.; Allen, J.G.; Wilton, S.D.; Akkari, P.A.; Huxtable, C.R.; Laing, N.G. A splice-site mutation causing ovine McArdle’s disease. Neuromuscul. Disord. 1997, 7, 336–342. [Google Scholar] [CrossRef]
  164. Crerar, M.M.; Hudson, J.W.; Matthews, K.E.; David, E.S.; Golding, G.B. Studies on the expression and evolution of the glycogen phosphorylase gene family in the rat. Genome 1988, 30, 582–590. [Google Scholar] [CrossRef]
  165. Tan, A.W.H.; Nuttall, F.Q. Characteristics of the dephosphoryated form of phosphorylase purified from rat liver and measurement of its activity in crude liver preparations. Biochim. Biophys. Acta (BBA)-Enzymol. 1975, 410, 45–60. [Google Scholar] [CrossRef]
  166. Berndt, N.; Rösen, P. Isolation and partial characterization of two forms of rat heart glycogen phosphorylase. Arch. Biochem. Biophys. 1984, 228, 143–154. [Google Scholar] [CrossRef]
  167. Hwang, P.K.; Tugendreich, S.; Fletterick, R.J. Molecular analysis of GPH1, the gene encoding glycogen phosphorylase in Saccharomyces cerevisiae. Mol. Cell. Biol. 1989, 9, 1659–1666. [Google Scholar] [CrossRef] [PubMed]
  168. Rath, V.L.; Hwang, P.K.; Fletterick, R.J. Purification and crystallization of glycogen phosphorylase from Saccharomyces cerevisiae. J. Mol. Biol. 1992, 225, 1027–1034. [Google Scholar] [CrossRef]
  169. Mori, H.; Tanizawa, K.; Fukui, T. Potato tuber type H phosphorylase isozyme. Molecular cloning, nucleotide sequence, and expression of a full-length cDNA in Escherichia coli. J. Biol. Chem. 1991, 266, 18446–18453. [Google Scholar] [CrossRef]
  170. Nakano, K.; Mori, H.; Fukui, T. Molecular cloning of cDNA encoding potato amyloplast a-Glucan phosphorylase and the structure of its transit peptide1. J. Biochem. 1989, 106, 691–695. [Google Scholar] [CrossRef]
  171. Sonnewald, U.; Basner, A.; Greve, B.; Steup, M. A second L-type isozyme of potato glucan phosphorylase: Cloning, antisense inhibition and expression analysis. Plant Mol Biol 1995, 27, 567–576. [Google Scholar] [CrossRef]
  172. Schupp, N.; Ziegler, P. The relation of starch phosphorylases to starch metabolism in wheat. Plant Cell Physiol. 2004, 45, 1471–1484. [Google Scholar] [CrossRef]
  173. Buchner, P.; Borisjuk, L.; Wobus, U. Glucan phosphorylases in Vicia faba L.: Cloning, structural analysis and expression patterns of cytosolic and plastidic forms in relation to starch. Planta 1996, 199, 64–73. [Google Scholar] [CrossRef]
  174. Sernee, M.F.; Ralton, J.E.; Nero, T.L.; Sobala, L.F.; Kloehn, J.; Vieira-Lara, M.A.; Cobbold, S.A.; Stanton, L.; Pires, D.E.V.; Hanssen, E.; et al. A family of dual-activity glycosyltransferase-phosphorylases mediates mannogen turnover and virulence in leishmania parasites. Cell Host Microbe 2019, 26, 385.e9–399.e9. [Google Scholar] [CrossRef] [Green Version]
  175. Puchart, V. Glycoside phosphorylases: Structure, catalytic properties and biotechnological potential. Biotechnol. Adv. 2015, 33, 261–276. [Google Scholar] [CrossRef]
  176. Teze, D.; Coines, J.; Raich, L.; Kalichuk, V.; Solleux, C.; Tellier, C.; André-Miral, C.; Svensson, B.; Rovira, C. A single point mutation converts GH84 O -GlcNAc Hydrolases into phosphorylases: Experimental and theoretical evidence. J. Am. Chem. Soc. 2020, 142, 2120–2124. [Google Scholar] [CrossRef] [PubMed]
  177. Mayer, C.; Vocadlo, D.J.; Mah, M.; Rupitz, K.; Stoll, D.; Warren, R.A.J.; Withers, S.G. Characterization of a β-N-acetylhexosaminidase and a β-N-acetylglucosaminidase/β-glucosidase from Cellulomonas fimi. FEBS J. 2006, 273, 2929–2941. [Google Scholar] [CrossRef] [PubMed]
  178. Reith, J.; Mayer, C. Peptidoglycan turnover and recycling in Gram-positive bacteria. Appl. Microbiol. Biotechnol. 2011, 92, 1–11. [Google Scholar] [CrossRef] [PubMed]
  179. Ferjani, A.; Mustardy, L.; Sulpice, R.; Marin, K.; Suzuki, I.; Hagemann, M.; Murata, N. Glucosylglycerol, a compatible solute, sustains cell division under salt stress. Plant Physiol. 2003, 131, 1628–1637. [Google Scholar] [CrossRef] [Green Version]
  180. Franceus, J.; Desmet, T. Sucrose phosphorylase and related enzymes in glycoside hydrolase family 13: Discovery, application and engineering. Int. J. Mol. Sci. 2020, 21, 2526. [Google Scholar] [CrossRef] [Green Version]
  181. Kitamoto, Y.; Akashi, H.; Tanaka, H.; Mori, N. α-Glucose-1-phosphate formation by a novel trehalose phosphorylase from Flammulina velutipes. FEMS Microbiol. Lett. 1988, 55, 147–150. [Google Scholar] [CrossRef] [Green Version]
  182. Cori, G.T.; Cori, C.F. Crystalline Muscle Phosphorylase Iv. Formation of Glycogen. J. Biol. Chem. 1943, 151, 57–63. [Google Scholar] [CrossRef]
  183. Lee, Y.P. Potato phosphorylase. I. Purification, physicochemical properties and catalytic activity. Biochim. Biophys. Acta 1960, 43, 18–24. [Google Scholar] [CrossRef]
  184. Maruta, K.; Mukai, K.; Yamashita, H.; Kubota, M.; Chaen, H.; Fukuda, S.; Kurimoto, M. Gene encoding a trehalose phosphorylase from Thermoanaerobacter brockii ATCC 35047. Biosci. Biotechnol. Biochem. 2002, 66, 1976–1980. [Google Scholar] [CrossRef] [Green Version]
  185. Yamamoto, T.; Maruta, K.; Mukai, K.; Yamashita, H.; Nishimoto, T.; Kubota, M.; Fukuda, S.; Kurimoto, M.; Tsujisaka, Y. Cloning and sequencing of kojibiose phosphorylase gene from Thermoanaerobacter brockii ATCC35047. J. Biosci. Bioeng. 2004, 98, 99–106. [Google Scholar] [CrossRef]
  186. Kiss, J.Z.; Vasconcelos, A.C.; Triemer, R.E. Structure of the Euglenoid storage carbohydrate, paramylon. Am. J. Bot. 1987, 74, 877–882. [Google Scholar] [CrossRef]
  187. Nakajima, M.; Nihira, T.; Nishimoto, M.; Kitaoka, M. Identification of galacto-N-biose phosphorylase from Clostridium perfringens ATCC13124. Appl. Microbiol. Biotechnol. 2008, 78, 465–471. [Google Scholar] [CrossRef] [PubMed]
  188. Derensy-Dron, D.; Krzewinski, F.; Brassart, C.; Bouquelet, S. β-1,3-Galactosyl-N-acetylhexosamine phosphorylase from Bifidobacterium bifidum DSM 20082: Characterization, partial purification and relation to mucin degradation. Biotechnol. Appl. Biochem. 1999, 29, 3–10. [Google Scholar] [CrossRef] [PubMed]
  189. Ladevèze, S. Functional and Structural Insights into Glycoside Hydrolase Family 130 Enzymes: Implications in Carbohydrate Foraging by Human Gut Bacteria. Ph.D. Thesis, INSA de Toulouse, Toulouse, France, 2015. [Google Scholar]
  190. Cuskin, F.; Lowe, E.C.; Temple, M.J.; Zhu, Y.; Cameron, E.A.; Pudlo, N.A.; Porter, N.T.; Urs, K.; Thompson, A.J.; Cartmell, A.; et al. Human gut Bacteroidetes can utilize yeast mannan through a selfish mechanism. Nature 2015, 517, 165–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. L’vov, V.L.; Shashkov, A.S.; Knirel, Y.A.; Arifulina, A.E.; Senchenkova, S.N.; Yakovlev, A.V.; Dmitriev, B.A. Structure of the O-specific polysaccharide chain of Shigella boydii type 5 lipopolysaccharide: A repeated study. Carbohydr. Res. 1995, 279, 183–192. [Google Scholar] [CrossRef]
  192. Garron, M.-L.; Henrissat, B. The continuing expansion of CAZymes and their families. Curr Opin Chem Biol 2019, 53, 82–87. [Google Scholar] [CrossRef] [PubMed]
  193. Drula, E.; Garron, M.-L.; Dogan, S.; Lombard, V.; Henrissat, B.; Terrapon, N. The carbohydrate-active enzyme database: Functions and literature. Nucleic Acids Res. 2022, 50, D571–D577. [Google Scholar] [CrossRef]
  194. Cuskin, F.; Baslé, A.; Ladevèze, S.; Day, A.M.; Gilbert, H.J.; Davies, G.J.; Potocki-Véronèse, G.; Lowe, E.C. The GH130 family of mannoside phosphorylases contains glycoside hydrolases that target β-1,2-mannosidic linkages in Candida mannan. J. Biol. Chem. 2015, 290, 25023–25033. [Google Scholar] [CrossRef] [Green Version]
  195. Ladevèze, S.; Cioci, G.; Roblin, P.; Mourey, L.; Tranier, S.; Potocki-Véronèse, G. Structural bases for N-glycan processing by mannoside phosphorylase. Acta Cryst. D 2015, 71, 1335–1346. [Google Scholar] [CrossRef]
  196. Sprogøe, D.; van den Broek, L.A.M.; Mirza, O.; Kastrup, J.S.; Voragen, A.G.J.; Gajhede, M.; Skov, L.K. Crystal structure of sucrose phosphorylase from Bifidobacterium adolescentis. Biochemistry 2004, 43, 1156–1162. [Google Scholar] [CrossRef]
  197. Egloff, M.-P.; Uppenberg, J.; Haalck, L.; van Tilbeurgh, H. Crystal structure of maltose phosphorylase from Lactobacillus brevis. Structure 2001, 9, 689–697. [Google Scholar] [CrossRef] [Green Version]
  198. Touhara, K.K.; Nihira, T.; Kitaoka, M.; Nakai, H.; Fushinobu, S. Structural basis for reversible phosphorolysis and hydrolysis reactions of 2-O-α-Glucosylglycerol Phosphorylase. J. Biol. Chem. 2014, 289, 18067–18075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Okada, S.; Yamamoto, T.; Watanabe, H.; Nishimoto, T.; Chaen, H.; Fukuda, S.; Wakagi, T.; Fushinobu, S. Structural and mutational analysis of substrate recognition in kojibiose phosphorylase. FEBS J. 2014, 281, 778–786. [Google Scholar] [CrossRef] [Green Version]
  200. Van Hoorebeke, A.; Stout, J.; Van der Meeren, R.; Kyndt, J.; Van Beeumen, J.; Savvides, S.N. Crystallization and X-ray diffraction studies of inverting trehalose phosphorylase from Thermoanaerobacter sp. Acta Cryst. F Struct. Biol. Cryst. Commun. 2010, 66, 442–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. O’Neill, E.C.; Pergolizzi, G.; Stevenson, C.E.M.; Lawson, D.M.; Nepogodiev, S.A.; Field, R.A. Cellodextrin phosphorylase from RuminiClostridium thermocellum: X-ray crystal structure and substrate specificity analysis. Carbohydr. Res. 2017, 451, 118–132. [Google Scholar] [CrossRef] [PubMed]
  202. Bianchetti, C.M.; Elsen, N.L.; Fox, B.G.; Phillips, G.N. Structure of cellobiose phosphorylase from Clostridium thermocellum in complex with phosphate. Acta Cryst. F Struct. Biol. Cryst. Commun. 2011, 67, 1345–1349. [Google Scholar] [CrossRef] [Green Version]
  203. Hidaka, M.; Kitaoka, M.; Hayashi, K.; Wakagi, T.; Shoun, H.; Fushinobu, S. Structural dissection of the reaction mechanism of cellobiose phosphorylase. Biochem. J. 2006, 398, 37–43. [Google Scholar] [CrossRef] [Green Version]
  204. Fushinobu, S.; Hidaka, M.; Hayashi, A.M.; Wakagi, T.; Shoun, H.; Kitaoka, M. Interactions between glycoside hydrolase family 94 cellobiose phosphorylase and glucosidase inhibitors. J. Appl. Glycosci. 2011, 58, 91–97. [Google Scholar] [CrossRef] [Green Version]
  205. Hidaka, M.; Honda, Y.; Kitaoka, M.; Nirasawa, S.; Hayashi, K.; Wakagi, T.; Shoun, H.; Fushinobu, S. Chitobiose Phosphorylase from Vibrio proteolyticus, a member of glycosyl transferase family 36, has a Clan GH-L-like (α/α)6 barrel fold. Structure 2004, 12, 937–947. [Google Scholar] [CrossRef]
  206. Hidaka, M.; Nishimoto, M.; Kitaoka, M.; Wakagi, T.; Shoun, H.; Fushinobu, S. The crystal structure of Galacto-N-biose/Lacto-N-biose I Phosphorylase. J. Biol. Chem. 2009, 284, 7273–7283. [Google Scholar] [CrossRef] [Green Version]
  207. Koyama, Y.; Hidaka, M.; Nishimoto, M.; Kitaoka, M. Directed evolution to enhance thermostability of galacto-N-biose/lacto-N-biose I phosphorylase. Protein Eng. Des. Sel. 2013, 26, 755–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Nakae, S.; Ito, S.; Higa, M.; Senoura, T.; Wasaki, J.; Hijikata, A.; Shionyu, M.; Ito, S.; Shirai, T. Structure of novel enzyme in mannan biodegradation process 4-O-β-d-Mannosyl-d-Glucose phosphorylase MGP. J. Mol. Biol. 2013, 425, 4468–4478. [Google Scholar] [CrossRef] [PubMed]
  209. Ye, Y.; Saburi, W.; Odaka, R.; Kato, K.; Sakurai, N.; Komoda, K.; Nishimoto, M.; Kitaoka, M.; Mori, H.; Yao, M. Structural insights into the difference in substrate recognition of two mannoside phosphorylases from two GH130 subfamilies. FEBS Lett. 2016, 590, 828–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Dai, L.; Chang, Z.; Yang, J.; Liu, W.; Yang, Y.; Chen, C.-C.; Zhang, L.; Huang, J.-W.; Sun, Y.; Guo, R.-T. Structural investigation of a thermostable 1,2-β-mannobiose phosphorylase from Thermoanaerobacter sp. X-514. Biochem. Biophys. Res. Commun. 2021, 579, 54–61. [Google Scholar] [CrossRef] [PubMed]
  211. Watson, K.A. Phosphorylase recognition and phosphorolysis of its oligosaccharide substrate: Answers to a long outstanding question. EMBO J. 1999, 18, 4619–4632. [Google Scholar] [CrossRef]
  212. Geremia, S.; Campagnolo, M.; Schinzel, R.; Johnson, L.N. Enzymatic catalysis in crystals of Escherichia coli maltodextrin phosphorylase. J. Mol. Biol. 2002, 322, 413–423. [Google Scholar] [CrossRef]
  213. Campagnolo, M.; Campa, C.; Zorzi, R.D.; Wuerges, J.; Geremia, S. X-ray studies on ternary complexes of maltodextrin phosphorylase. Arch. Biochem. Biophys. 2008, 471, 11–19. [Google Scholar] [CrossRef]
  214. Mathieu, C.; de la Sierra-Gallay, I.L.; Duval, R.; Xu, X.; Cocaign, A.; Léger, T.; Woffendin, G.; Camadro, J.-M.; Etchebest, C.; Haouz, A.; et al. Insights into brain glycogen metabolism. J. Biol. Chem. 2016, 291, 18072–18083. [Google Scholar] [CrossRef] [Green Version]
  215. Rath, V.L.; Ammirati, M.; Danley, D.E.; Ekstrom, J.L.; Gibbs, E.M.; Hynes, T.R.; Mathiowetz, A.M.; McPherson, R.K.; Olson, T.V.; Treadway, J.L.; et al. Human liver glycogen phosphorylase inhibitors bind at a new allosteric site. Chem. Biol. 2000, 7, 677–682. [Google Scholar] [CrossRef] [Green Version]
  216. Rath, V.L.; Ammirati, M.; LeMotte, P.K.; Fennell, K.F.; Mansour, M.N.; Danley, D.E.; Hynes, T.R.; Schulte, G.K.; Wasilko, D.J.; Pandit, J. Activation of human liver glycogen phosphorylase by alteration of the secondary structure and packing of the catalytic core. Mol. Cell. 2000, 6, 139–148. [Google Scholar] [CrossRef]
  217. Ekstrom, J.L.; Pauly, T.A.; Carty, M.D.; Soeller, W.C.; Culp, J.; Danley, D.E.; Hoover, D.J.; Treadway, J.L.; Gibbs, E.M.; Fletterick, R.J.; et al. Structure-Activity analysis of the purine binding site of human liver glycogen phosphorylase. Chem. Biol. 2002, 9, 915–924. [Google Scholar] [CrossRef] [Green Version]
  218. Wright, S.W.; Rath, V.L.; Genereux, P.E.; Hageman, D.L.; Levy, C.B.; McClure, L.D.; McCoid, S.C.; McPherson, R.K.; Schelhorn, T.M.; Wilder, D.E.; et al. 5-Chloroindoloyl glycine amide inhibitors of glycogen phosphorylase: Synthesis, In Vitro, in vivo, and X-ray crystallographic characterization. Bioorganic Med. Chem. Lett. 2005, 15, 459–465. [Google Scholar] [CrossRef] [PubMed]
  219. Klabunde, T.; Wendt, K.U.; Kadereit, D.; Brachvogel, V.; Burger, H.-J.; Herling, A.W.; Oikonomakos, N.G.; Kosmopoulou, M.N.; Schmoll, D.; Sarubbi, E.; et al. Acyl Ureas as human liver glycogen phosphorylase inhibitors for the treatment of Type 2 Diabetes. J. Med. Chem. 2005, 48, 6178–6193. [Google Scholar] [CrossRef] [PubMed]
  220. Pautsch, A.; Stadler, N.; Wissdorf, O.; Langkopf, E.; Moreth, W.; Streicher, R. Molecular recognition of the protein phosphatase 1 glycogen targeting subunit by glycogen phosphorylase. J. Biol. Chem. 2008, 283, 8913–8918. [Google Scholar] [CrossRef] [Green Version]
  221. Onda, K.; Suzuki, T.; Shiraki, R.; Yonetoku, Y.; Negoro, K.; Momose, K.; Katayama, N.; Orita, M.; Yamaguchi, T.; Ohta, M.; et al. Synthesis of 5-chloro-N-aryl-1H-indole-2-carboxamide derivatives as inhibitors of human liver glycogen phosphorylase a. Bioorganic Med. Chem. 2008, 16, 5452–5464. [Google Scholar] [CrossRef]
  222. Anderka, O.; Loenze, P.; Klabunde, T.; Dreyer, M.K.; Defossa, E.; Wendt, K.U.; Schmoll, D. Thermodynamic characterization of allosteric glycogen phosphorylase inhibitors. Biochemistry 2008, 47, 4683–4691. [Google Scholar] [CrossRef]
  223. Thomson, S.A.; Banker, P.; Bickett, D.M.; Boucheron, J.A.; Carter, H.L.; Clancy, D.C.; Cooper, J.P.; Dickerson, S.H.; Garrido, D.M.; Nolte, R.T.; et al. Anthranilimide based glycogen phosphorylase inhibitors for the treatment of type 2 diabetes. Part 3: X-ray crystallographic characterization, core and urea optimization and in vivo efficacy. Bioorganic Med. Chem. Lett. 2009, 19, 1177–1182. [Google Scholar] [CrossRef]
  224. Lukacs, C.M.; Oikonomakos, N.G.; Crowther, R.L.; Hong, L.-N.; Kammlott, R.U.; Levin, W.; Li, S.; Liu, C.-M.; Lucas-McGady, D.; Pietranico, S.; et al. The crystal structure of human muscle glycogen phosphorylase a with bound glucose and AMP: An intermediate conformation with T-state and R-state features. Proteins 2006, 63, 1123–1126. [Google Scholar] [CrossRef]
  225. Whittamore, P.R.O.; Addie, M.S.; Bennett, S.N.L.; Birch, A.M.; Butters, M.; Godfrey, L.; Kenny, P.W.; Morley, A.D.; Murray, P.M.; Oikonomakos, N.G.; et al. Novel thienopyrrole glycogen phosphorylase inhibitors: Synthesis, In Vitro SAR and crystallographic studies. Bioorganic Med. Chem. Lett. 2006, 16, 5567–5571. [Google Scholar] [CrossRef]
  226. Lin, K.; Rath, V.L.; Dai, S.C.; Fletterick, R.J.; Hwang, P.K. A protein phosphorylation switch at the conserved allosteric site in GP. Science 1996, 273, 1539–1541. [Google Scholar] [CrossRef]
  227. Bhaumik, P.; Dhepe, P.L. Conversion of Biomass into Sugars. In Biomass Sugars for Non-Fuel Applications; RSC Publishing: London, UK, 2015; pp. 1–53. [Google Scholar]
  228. Barcelos, M.C.S.; Vespermann, K.A.C.; Pelissari, F.M.; Molina, G. Current status of biotechnological production and applications of microbial exopolysaccharides. Crit. Rev. Food Sci. Nutr. 2020, 60, 1475–1495. [Google Scholar] [CrossRef] [PubMed]
  229. Dashtban, M.; Schraft, H.; Qin, W. Fungal bioconversion of lignocellulosic residues; opportunities & perspectives. Int. J. Biol. Sci. 2009, 5, 578–595. [Google Scholar] [CrossRef] [PubMed]
  230. Chen, R. Enzym. Microb. Technol. for synthesis of bioactive oligosaccharides: An update. Appl. Microbiol. Biotechnol. 2018, 102, 3017–3026. [Google Scholar] [CrossRef] [PubMed]
  231. Anderson, L.A.; Islam, M.A.; Prather, K.L.J. Synthetic biology strategies for improving microbial synthesis of “green” biopolymers. J. Biol. Chem. 2018, 293, 5053–5061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Raimo, A. Carbohydrate Chemistry: Fundamentals and Applications; World Scientific Publishing Company: Singapore, 2018; ISBN 978-981-322-366-0. [Google Scholar]
  233. Blow, N. Glycobiology: A spoonful of sugar. Nature 2009, 457, 617–620. [Google Scholar] [CrossRef] [PubMed]
  234. Panza, M.; Pistorio, S.G.; Stine, K.J.; Demchenko, A.V. Automated chemical oligosaccharide synthesis: Novel approach to traditional challenges. Chem. Rev. 2018, 118, 8105–8150. [Google Scholar] [CrossRef]
  235. Eller, S.; Collot, M.; Yin, J.; Hahm, H.S.; Seeberger, P.H. Automated solid-phase synthesis of chondroitin sulfate glycosaminoglycans. Angew. Chem. Int. Ed. Engl. 2013, 52, 5858–5861. [Google Scholar] [CrossRef]
  236. Walvoort, M.T.C.; van den Elst, H.; Plante, O.J.; Kröck, L.; Seeberger, P.H.; Overkleeft, H.S.; van der Marel, G.A.; Codée, J.D.C. Automated solid-phase synthesis of β-mannuronic acid alginates. Angew. Chem. Int. Ed. Engl. 2012, 51, 4393–4396. [Google Scholar] [CrossRef]
  237. Nielsen, M.M.; Pedersen, C.M. Catalytic Glycosylations in oligosaccharide synthesis. Chem. Rev. 2018, 118, 8285–8358. [Google Scholar] [CrossRef]
  238. Pergolizzi, G.; Kuhaudomlarp, S.; Kalita, E.; Field, R.A. Glycan Phosphorylases in multi-enzyme synthetic processes. Protein Pept. Lett. 2017, 24, 696–709. [Google Scholar] [CrossRef] [Green Version]
  239. Nishimoto, M.; Kitaoka, M. Practical preparation of lacto-N-biose I, a candidate for the bifidus factor in human milk. Biosci. Biotechnol. Biochem. 2007, 71, 2101–2104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Nishimoto, M. Large scale production of lacto-N-biose I, a building block of type I human milk oligosaccharides, using sugar phosphorylases. Biosci. Biotechnol. Biochem. 2020, 84, 17–24. [Google Scholar] [CrossRef] [PubMed]
  241. Goedl, C.; Sawangwan, T.; Wildberger, P.; Nidetzky, B. Sucrose phosphorylase: A powerful transglucosylation catalyst for synthesis of α-d-glucosides as industrial fine chemicals. Biocatal. Biotransformation 2010, 28, 10–21. [Google Scholar] [CrossRef]
  242. De Bruyn, F.; De Paepe, B.; Maertens, J.; Beauprez, J.; De Cocker, P.; Mincke, S.; Stevens, C.; De Mey, M. Development of an in vivo glucosylation platform by coupling production to growth: Production of phenolic glucosides by a glycosyltransferase of Vitis vinifera. Biotechnol. Bioeng. 2015, 112, 1594–1603. [Google Scholar] [CrossRef]
  243. Samain, E.; Lancelon-Pin, C.; Férigo, F.; Moreau, V.; Chanzy, H.; Heyraud, A.; Driguez, H. Phosphorolytic synthesis of cellodextrins. Carbohydr. Res. 1995, 271, 217–226. [Google Scholar] [CrossRef]
  244. Hiraishi, M.; Igarashi, K.; Kimura, S.; Wada, M.; Kitaoka, M.; Samejima, M. Synthesis of highly ordered cellulose II In Vitro using cellodextrin phosphorylase. Carbohydr. Res. 2009, 344, 2468–2473. [Google Scholar] [CrossRef]
  245. Rivadeneyra, A.; Gonzalez-Martinez, A.; Gonzalez-Lopez, J.; Martin-Ramos, D.; Martinez-Toledo, M.V.; Rivadeneyra, M.A. Precipitation of phosphate minerals by microorganisms isolated from a fixed-biofilm reactor used for the treatment of domestic wastewater. Int. J. Environ. Res. Public Health 2014, 11, 3689–3704. [Google Scholar] [CrossRef] [Green Version]
  246. Liu, J.; Yin, X.; Li, Z.; Wu, X.; Zheng, Z.; Fang, J.; Gu, G.; Wang, P.G.; Liu, X. Facile Enzymatic Synthesis of Diverse Naturally-Occurring β-d-Mannopyranosides catalyzed by glycoside phosphorylases. ACS Catal. 2021, 11, 2763–2768. [Google Scholar] [CrossRef]
  247. Britton, J.; Majumdar, S.; Weiss, G.A. Continuous flow biocatalysis. Chem. Soc. Rev. 2018, 47, 5891–5918. [Google Scholar] [CrossRef]
  248. Zhong, C.; Duić, B.; Bolivar, J.M.; Nidetzky, B. Three-Enzyme phosphorylase cascade immobilized on solid support for biocatalytic synthesis of cello–oligosaccharides. ChemCatChem 2020, 12, 1350–1358. [Google Scholar] [CrossRef]
  249. Yu, H.; Chen, X. One-pot multienzyme (OPME) systems for chemoenzymatic synthesis of carbohydrates. Org. Biomol. Chem. 2016, 14, 2809–2818. [Google Scholar] [CrossRef] [PubMed]
  250. Tian, C.; Yang, J.; Li, Y.; Zhang, T.; Li, J.; Ren, C.; Men, Y.; Chen, P.; You, C.; Sun, Y.; et al. Artificially designed routes for the conversion of starch to value-added mannosyl compounds through coupling In Vitro and in vivo Metab. Eng. strategies. Metab. Eng. 2020, 61, 215–224. [Google Scholar] [CrossRef] [PubMed]
  251. Maertens, J.; Beauprez, J.; Mey, M.D. Metabolically Engineered Organisms for the Production of Added Value Bio-Products. Patent WO2012007481A3, 19 January 2012. [Google Scholar]
  252. De Groeve, M.R.M.; De Baere, M.; Hoflack, L.; Desmet, T.; Vandamme, E.J.; Soetaert, W. Creating lactose phosphorylase enzymes by directed evolution of cellobiose phosphorylase. Protein Eng. Des. Sel. 2009, 22, 393–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Groeve, M.R.M.D.; Remmery, L.; Hoorebeke, A.V.; Stout, J.; Desmet, T.; Savvides, S.N.; Soetaert, W. Construction of cellobiose phosphorylase variants with broadened acceptor specificity towards anomerically substituted glucosides. Biotechnol. Bioeng. 2010, 107, 413–420. [Google Scholar] [CrossRef]
  254. Kraus, M.; Grimm, C.; Seibel, J. Switching enzyme specificity from phosphate to resveratrol glucosylation. Chem. Commun. 2017, 53, 12181–12184. [Google Scholar] [CrossRef]
  255. Waites, M.J.; Morgan, N.L.; Rockey, J.S.; Higton, G. Industrial Microbiology: An Introduction; John Wiley & Sons: Hoboken, NJ, USA, 2009; ISBN 978-1-4443-1158-7. [Google Scholar]
  256. Fiske, C.H.; Subbarow, Y. The colorimetric determination of phosphorus. J. Biol. Chem. 1925, 66, 375–400. [Google Scholar] [CrossRef]
  257. Drueckes, P.; Schinzel, R.; Palm, D. Photometric microtiter assay of inorganic phosphate in the presence of acid-labile organic phosphates. Anal. Biochem. 1995, 230, 173–177. [Google Scholar] [CrossRef]
  258. Cogan, E.B.; Birrell, G.B.; Griffith, O.H. A robotics-based automated assay for inorganic and organic phosphates. Anal. Biochem. 1999, 271, 29–35. [Google Scholar] [CrossRef] [Green Version]
  259. Gawronski, J.D.; Benson, D.R. Microtiter assay for glutamine synthetase biosynthetic activity using inorganic phosphate detection. Anal. Biochem. 2004, 327, 114–118. [Google Scholar] [CrossRef]
  260. De Groeve, M.R.M.; Tran, G.H.; Van Hoorebeke, A.; Stout, J.; Desmet, T.; Savvides, S.N.; Soetaert, W. Development and application of a screening assay for glycoside phosphorylases. Anal. Biochem. 2010, 401, 162–167. [Google Scholar] [CrossRef] [Green Version]
  261. Tedokon, M.; Suzuki, K.; Kayamori, Y.; Fujita, S.; Katayama, Y. Enzymatic assay of inorganic phosphate with use of sucrose phosphorylase and phosphoglucomutase. Clin. Chem. 1992, 38, 512–515. [Google Scholar] [CrossRef] [PubMed]
  262. Serrano, A.; Löffelhardt, W. Identification of two different glyceraldehyde-3-phosphate dehydrogenases (phosphorylating) in the photosynthetic protist Cyanophora paradoxa. Arch. Microbiol. 1994, 162, 14–19. [Google Scholar] [CrossRef]
  263. Suárez, A.S.G.; Stefan, A.; Lemma, S.; Conte, E.; Hochkoeppler, A. Continuous enzyme-coupled assay of phosphate- or pyrophosphate-releasing enzymes. BioTechniques 2012, 53, 99–103. [Google Scholar] [CrossRef] [PubMed]
  264. Qureshi, M.N.; Stecher, G.; Sultana, T.; Abel, G.; Popp, M.; Bonn, G.K. Determination of carbohydrates in medicinal plants-comparison between TLC, mf-MELDI-MS and GC-MS. Phytochem. Anal. 2011, 22, 296–302. [Google Scholar] [CrossRef]
  265. Harvey, D.J. Derivatization of carbohydrates for analysis by chromatography; electrophoresis and mass spectrometry. J. Chromatogr. B 2011, 879, 1196–1225. [Google Scholar] [CrossRef]
  266. Corradini, C.; Cavazza, A.; Bignardi, C. High-Performance Anion-Exchange Chromatography Coupled with Pulsed Electrochemical Detection as a Powerful Tool to Evaluate Carbohydrates of Food Interest: Principles and Applications. Available online: https://www.hindawi.com/journals/ijcc/2012/487564/ (accessed on 21 April 2020).
  267. Hellerqvist, C.G.; Lindberg, B.; Svensson, S.; Holme, T.; Lindberg, A.A. Structural studies on the O-specific side-chains of the cell-wall lipopolysaccharide from Salmonella typhimurium 395 ms. Carbohydr. Res. 1968, 8, 43–55. [Google Scholar] [CrossRef]
  268. Ciucanu, I. Per-O-methylation reaction for structural analysis of carbohydrates by mass spectrometry. Anal. Chim. Acta 2006, 576, 147–155. [Google Scholar] [CrossRef]
  269. Irague, R.; Massou, S.; Moulis, C.; Saurel, O.; Milon, A.; Monsan, P.; Remaud-Siméon, M.; Portais, J.-C.; Potocki-Véronèse, G. NMR-Based structural Glycomics for high-throughput screening of carbohydrate-active enzyme specificity. Anal. Chem. 2011, 83, 1202–1206. [Google Scholar] [CrossRef]
  270. Handelsman, J. Metagenomics: Application of genomics to uncultured microorganisms. Microbiol. Mol. Biol. Rev. 2004, 68, 669–685. [Google Scholar] [CrossRef] [Green Version]
  271. Han, L.; Costello, C.E. Mass spectrometry of glycans. Biochemistry 2013, 78, 710–720. [Google Scholar] [CrossRef] [Green Version]
  272. Ropartz, D.; Fanuel, M.; Ujma, J.; Palmer, M.; Giles, K.; Rogniaux, H. Structure determination of large isomeric oligosaccharides of natural origin through multipass and multistage cyclic traveling-wave ion mobility mass spectrometry. Anal. Chem. 2019, 91, 12030–12037. [Google Scholar] [CrossRef] [PubMed]
  273. Tsai, S.-T.; Liew, C.Y.; Hsu, C.; Huang, S.-P.; Weng, W.-C.; Kuo, Y.-H.; Ni, C.-K. Automatic full glycan structural determination through logically derived sequence tandem mass spectrometry. ChemBioChem 2019, 20, 2351–2359. [Google Scholar] [CrossRef] [PubMed]
  274. Flemming, H.-C.; Wuertz, S. Bacteria and archaea on Earth and their abundance in biofilms. Nat. Rev. Microbiol. 2019, 17, 247–260. [Google Scholar] [CrossRef] [PubMed]
  275. Van de Guchte, M.; Blottière, H.M.; Doré, J. Humans as holobionts: Implications for prevention and therapy. Microbiome 2018, 6, 81. [Google Scholar] [CrossRef]
  276. Handelsman, J.; Rondon, M.R.; Brady, S.F.; Clardy, J.; Goodman, R.M. Molecular biological access to the chemistry of unknown soil microbes: A new frontier for natural products. Chem. Biol. 1998, 5, R245–R249. [Google Scholar] [CrossRef] [Green Version]
  277. Pace, N.R.; Stahl, D.A.; Lane, D.J.; Olsen, G.J. The analysis of natural microbial populations by ribosomal RNA Sequences. In Advances in Microbial Ecology; Marshall, K.C., Ed.; Advances in Microbial Ecology; Springer US: Boston, MA, USA, 1986; pp. 1–55. ISBN 978-1-4757-0611-6. [Google Scholar]
  278. Healy, F.G.; Ray, R.M.; Aldrich, H.C.; Wilkie, A.C.; Ingram, L.O.; Shanmugam, K.T. Direct isolation of functional genes encoding cellulases from the microbial consortia in a thermophilic, anaerobic digester maintained on lignocellulose. Appl. Microbiol. Biotechnol. 1995, 43, 667–674. [Google Scholar] [CrossRef]
  279. Li, J.; Jia, H.; Cai, X.; Zhong, H.; Feng, Q.; Sunagawa, S.; Arumugam, M.; Kultima, J.R.; Prifti, E.; Nielsen, T.; et al. An integrated catalog of reference genes in the human gut microbiome. Nat. Biotechnol. 2014, 32, 834–841. [Google Scholar] [CrossRef]
  280. Tierney, B.T.; Yang, Z.; Luber, J.M.; Beaudin, M.; Wibowo, M.C.; Baek, C.; Mehlenbacher, E.; Patel, C.J.; Kostic, A.D. The Landscape of genetic content in the gut and oral human microbiome. Cell Host Microbe 2019, 26, 283.e8–295.e8. [Google Scholar] [CrossRef]
  281. Xiao, L.; Feng, Q.; Liang, S.; Sonne, S.B.; Xia, Z.; Qiu, X.; Li, X.; Long, H.; Zhang, J.; Zhang, D.; et al. A catalog of the mouse gut metagenome. Nat. Biotechnol. 2015, 33, 1103–1108. [Google Scholar] [CrossRef]
  282. Lesker, T.R.; Durairaj, A.C.; Gálvez, E.J.C.; Lagkouvardos, I.; Baines, J.F.; Clavel, T.; Sczyrba, A.; McHardy, A.C.; Strowig, T. An integrated metagenome catalog reveals new insights into the murine gut microbiome. Cell Rep. 2020, 30, 2909.e6–2922.e6. [Google Scholar] [CrossRef] [Green Version]
  283. Xiao, L.; Estellé, J.; Kiilerich, P.; Ramayo-Caldas, Y.; Xia, Z.; Feng, Q.; Liang, S.; Pedersen, A.Ø.; Kjeldsen, N.J.; Liu, C.; et al. A reference gene catalogue of the pig gut microbiome. Nat. Microbiol. 2016, 1, 16161. [Google Scholar] [CrossRef] [PubMed]
  284. Li, J.; Zhong, H.; Ramayo-Caldas, Y.; Terrapon, N.; Lombard, V.; Potocki-Veronese, G.; Estellé, J.; Popova, M.; Yang, Z.; Zhang, H.; et al. A catalog of microbial genes from the bovine rumen unveils a specialized and diverse biomass-degrading environment. GigaScience 2020, 9, giaa057. [Google Scholar] [CrossRef] [PubMed]
  285. Bailly, J.; Fraissinet-Tachet, L.; Verner, M.-C.; Debaud, J.-C.; Lemaire, M.; Wésolowski-Louvel, M.; Marmeisse, R. Soil eukaryotic functional diversity, a metatranscriptomic approach. ISME J. 2007, 1, 632–642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Tasse, L.; Bercovici, J.; Pizzut-Serin, S.; Robe, P.; Tap, J.; Klopp, C.; Cantarel, B.L.; Coutinho, P.M.; Henrissat, B.; Leclerc, M.; et al. Functional metagenomics to mine the human gut microbiome for dietary fiber catabolic enzymes. Genome Res. 2010, 20, 1605–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Lam, K.N.; Charles, T.C. Strong spurious transcription likely contributes to DNA insert bias in typical metagenomic clone libraries. Microbiome 2015, 3, 22. [Google Scholar] [CrossRef] [Green Version]
  288. Taupp, M.; Mewis, K.; Hallam, S.J. The art and design of functional metagenomic screens. Curr. Opin. Biotechnol. 2011, 22, 465–472. [Google Scholar] [CrossRef]
  289. Lewin, A.; Lale, R.; Wentzel, A. Expression Platforms for Functional Metagenomics: Emerging Technology Options Beyond Escherichia coli. In Functional Metagenomics: Tools and Applications; Charles, T.C., Liles, M.R., Sessitsch, A., Eds.; Springer International Publishing: Berlin/Heidelberg, Germany, 2017; ISBN 978-3-319-61508-0. [Google Scholar]
  290. Lam, K.N.; Martens, E.C.; Charles, T.C. Developing a Bacteroides system for function-based screening of DNA from the human gut microbiome. mSystems 2018, 3, e00195-17. [Google Scholar] [CrossRef] [Green Version]
  291. Tauzin, A.S.; Laville, E.; Cecchini, D.; Blottière, H.M.; Leclerc, M.; Doré, J.; Potocki-Veronese, G. Human gut metagenomics: Success and limits of the activity-based approaches. In Functional Metagenomics: Tools and Applications; Charles, T.C., Liles, M.R., Sessitsch, A., Eds.; Springer International Publishing: Cham, The Netherlands, 2017; pp. 161–178. ISBN 978-3-319-61510-3. [Google Scholar]
  292. Ufarté, L.; Potocki-Veronese, G.; Laville, É. Discovery of new protein families and functions: New challenges in functional metagenomics for biotechnologies and microbial ecology. Front. Microbiol. 2015, 6, 563. [Google Scholar] [CrossRef]
  293. Colin, P.-Y.; Kintses, B.; Gielen, F.; Miton, C.M.; Fischer, G.; Mohamed, M.F.; Hyvönen, M.; Morgavi, D.P.; Janssen, D.B.; Hollfelder, F. Ultrahigh-throughput discovery of promiscuous enzymes by picodroplet functional metagenomics. Nat. Commun. 2015, 6, 10008. [Google Scholar] [CrossRef]
  294. Cottrell, M.T.; Yu, L.; Kirchman, D.L. Sequence and expression analyses of Cytophaga-like hydrolases in a Western arctic metagenomic library and the Sargasso Sea. Appl. Environ. Microbiol. 2005, 71, 8506–8513. [Google Scholar] [CrossRef] [Green Version]
  295. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  296. Tatusov, R.L.; Koonin, E.V.; Lipman, D.J. A genomic perspective on protein families. Science 1997, 278, 631–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. El-Gebali, S.; Mistry, J.; Bateman, A.; Eddy, S.R.; Luciani, A.; Potter, S.C.; Qureshi, M.; Richardson, L.J.; Salazar, G.A.; Smart, A.; et al. The Pfam protein families database in 2019. Nucleic Acids Res. 2019, 47, D427–D432. [Google Scholar] [CrossRef]
  298. Coggill, P.; Finn, R.D.; Bateman, A. Identifying Protein Domains with the Pfam Database. Curr. Protoc. Bioinform. 2008, 23, 2.5.1–2.5.17. [Google Scholar] [CrossRef] [PubMed]
  299. Markowitz, V.M.; Ivanova, N.N.; Szeto, E.; Palaniappan, K.; Chu, K.; Dalevi, D.; Chen, I.-M.A.; Grechkin, Y.; Dubchak, I.; Anderson, I.; et al. IMG/M: A data management and analysis system for metagenomes. Nucleic Acids Res 2008, 36, D534–D538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Meyer, F.; Paarmann, D.; D’Souza, M.; Olson, R.; Glass, E.; Kubal, M.; Paczian, T.; Rodriguez, A.; Stevens, R.; Wilke, A.; et al. The metagenomics RAST server–a public resource for the automatic phylogenetic and functional analysis of metagenomes. BMC Bioinform. 2008, 9, 386. [Google Scholar] [CrossRef] [Green Version]
  301. Lapébie, P.; Lombard, V.; Drula, E.; Terrapon, N.; Henrissat, B. Bacteroidetes use thousands of enzyme combinations to break down glycans. Nat. Commun. 2019, 10, 1–7. [Google Scholar] [CrossRef]
  302. Lobb, B.; Doxey, A.C. Novel function discovery through sequence and structural data mining. Curr. Opin. Struct. Biol. 2016, 38, 53–61. [Google Scholar] [CrossRef]
  303. Gerlt, J.A.; Bouvier, J.T.; Davidson, D.B.; Imker, H.J.; Sadkhin, B.; Slater, D.R.; Whalen, K.L. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A Web Tool for Generating Protein Sequence Similarity Networks. Biochim. Biophys. Acta (BBA)—Proteins Proteom. 2015, 1854, 1019–1037. [Google Scholar] [CrossRef] [Green Version]
  304. Atkinson, H.J.; Morris, J.H.; Ferrin, T.E.; Babbitt, P.C. Using sequence similarity networks for visualization of relationships across diverse protein superfamilies. PLoS ONE 2009, 4, e4345. [Google Scholar] [CrossRef]
  305. Levin, B.J.; Huang, Y.Y.; Peck, S.C.; Wei, Y.; Campo, A.M.; Marks, J.A.; Franzosa, E.A.; Huttenhower, C.; Balskus, E.P. A prominent glycyl radical enzyme in human gut microbiomes metabolizes trans-4-hydroxy-l-proline. Science 2017, 355, eaai8386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Viborg, A.H.; Terrapon, N.; Lombard, V.; Michel, G.; Czjzek, M.; Henrissat, B.; Brumer, H. A Subfamily Roadmap for Functional Glycogenomics of the Evolutionarily Diverse Glycoside Hydrolase Family 16 (GH16). J. Biol. Chem. 2019, 294, 15973–15986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Romero, P.A.; Tran, T.M.; Abate, A.R. Dissecting enzyme function with microfluidic-based deep mutational scanning. Proc. Natl. Acad. Sci. USA 2015, 112, 7159–7164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Inverting GPs and GHs catalytic mechanisms and family classification. The carbohydrates indicated for each family correspond to the glycosyl donors used by GPs of this family.
Figure 1. Inverting GPs and GHs catalytic mechanisms and family classification. The carbohydrates indicated for each family correspond to the glycosyl donors used by GPs of this family.
Ijms 23 03043 g001
Figure 2. Retaining GPs catalytic mechanisms and family classification. The carbohydrates indicated for each family correspond to the glycosyl donors used by GPs of this family.
Figure 2. Retaining GPs catalytic mechanisms and family classification. The carbohydrates indicated for each family correspond to the glycosyl donors used by GPs of this family.
Ijms 23 03043 g002
Figure 3. Structural representatives of the phosphorylase families with their main substrate (glycoside donnor, or acceptor, in green, or inorganic phosphate in red, when available). (A) GH3 β-glucoside phosphorylase BglX bound to 2FGlc (5VQE); (B) GH13_18 sucrose phosphorylase E232Q mutant from Bifidobacterium adolescentis in complex with sucrose (2GDU); (C) GH65 kojibiose phosphorylase from Caldicellulosiruptor saccharolyticus DSM 8903 in complex with kojibiose (3WIQ); (D) GH94 chitobiose phosphorylase from Vibrio proteolyticus in complex with GlcNAc (1V7W); (E) GH112 galacto-N-biose phosphorylase from Bifidobacterium longum subsp. longum JCM 1217 in complex with GalNAc (2ZUT). (F) GH130 MannosylGlucose phosphorylase from Bacteroides fragilis NCTC 9343 in complex with mannosyl-glucose and phosphate (3WAS); (G) GH149 β-1,3-oligoglucan phosphorylase in complex with laminaribiose (6HQ8); (H) GT35 maltodextrin phosphorylase from Escherichia coli str. K-12 substr. MG1655 in complex with maltose (1AHP); (I) GT108 Dual-Activity Glycosyltransferase-Phosphorylase from Leishmania mexicana MHOM/GT/2001/U1103 in complex with inorganic phosphate (6Q50).
Figure 3. Structural representatives of the phosphorylase families with their main substrate (glycoside donnor, or acceptor, in green, or inorganic phosphate in red, when available). (A) GH3 β-glucoside phosphorylase BglX bound to 2FGlc (5VQE); (B) GH13_18 sucrose phosphorylase E232Q mutant from Bifidobacterium adolescentis in complex with sucrose (2GDU); (C) GH65 kojibiose phosphorylase from Caldicellulosiruptor saccharolyticus DSM 8903 in complex with kojibiose (3WIQ); (D) GH94 chitobiose phosphorylase from Vibrio proteolyticus in complex with GlcNAc (1V7W); (E) GH112 galacto-N-biose phosphorylase from Bifidobacterium longum subsp. longum JCM 1217 in complex with GalNAc (2ZUT). (F) GH130 MannosylGlucose phosphorylase from Bacteroides fragilis NCTC 9343 in complex with mannosyl-glucose and phosphate (3WAS); (G) GH149 β-1,3-oligoglucan phosphorylase in complex with laminaribiose (6HQ8); (H) GT35 maltodextrin phosphorylase from Escherichia coli str. K-12 substr. MG1655 in complex with maltose (1AHP); (I) GT108 Dual-Activity Glycosyltransferase-Phosphorylase from Leishmania mexicana MHOM/GT/2001/U1103 in complex with inorganic phosphate (6Q50).
Ijms 23 03043 g003
Figure 4. Biotechnological applications of GPs: the different strategies employed for improving glycoside production yields are marked in red.
Figure 4. Biotechnological applications of GPs: the different strategies employed for improving glycoside production yields are marked in red.
Ijms 23 03043 g004
Figure 5. Sequence- and activity-based approaches for GP discovery from (meta)genomes and functional characterization.
Figure 5. Sequence- and activity-based approaches for GP discovery from (meta)genomes and functional characterization.
Ijms 23 03043 g005
Table 1. List of functionally characterized GPs. Some of the enzymes (labelled with a star) referenced in this table do not currently appear in the list of characterized enzymes of the CAZy Database.
Table 1. List of functionally characterized GPs. Some of the enzymes (labelled with a star) referenced in this table do not currently appear in the list of characterized enzymes of the CAZy Database.
FamilyMechanismGenBank
Uniprot
Protein NameOrganismPDBReverse-Phosphorolysis ActivityReference (DOI)
DonorAcceptorProduct
GH3RetainingAUG44408.1β-glucoside phosphorylase
(BglP)
uncultured bacterium5VQD, 5VQEβ-GlcNac1P
β-Glc1P
p-NP/DNPp-Nitrophenyl-N-acetyl-β-d-glucosaminide
p-nitrophenyl-β-d-glucopyranoside
2,4-Dinitrophenyl-β- d-glucopyranoside
[30]
GH3RetainingAAQ05801.1Bifunctional N-acetyl-β-glucosaminidase/β-glucosidase
(Nag3)
Cellulomonas fimi ATCC 484 β-Glc1PDNP2,4-Dinitrophenyl-β-d-glucopyranoside[31]
GH13_18RetainingAAO33821.1Sucrose phosphorylase
(SucP;SP;BaSP)
Bifidobacterium adolescentis DSM 200831R7A, 2GDU, 2GDV, 5C8B, 5M9X, 5MAN, 5MB2, 6FMEα-Glc1Pd-fructoseSucrose:
(α-d-glucopyranosyl-(1→2)-β-d-fructofuranoside)
[32]
GH13_18RetainingAAN24362.1Sucrose phosphorylase
(Spl;BL0536)
Bifidobacterium longum NCC2705 α-Glc1Pd-fructoseSucrose[33]
GH13_18RetainingAAO84039.1Sucrose phosphorylase
(SplP)
Bifidobacterium longum SJ32 α-Glc1Pd fructoseSucrose[34]
GH13_18RetainingBAF62433.1Sucrose phosphorylase
(Spl)
Bifidobacterium longum subsp. longum JCM 1217 α-Glc1Pd-fructoseSucrose[35]
GH13_18RetainingAAC74391.2Glucosylglycerate phosphorylase
(YcjM)
Escherichia coli str. K-12 substr. MG1655 α-Glc1Pd glycerateGlucosyl-glycerate[28]
GH13_18RetainingAAO21868.1Sucrose phosphorylase
(LaSP;GtfA2;LBA1437)
Lactobacillus acidophilus NCFM α-Glc1Pd fructoseSucrose[36]
GH13_18RetainingBAA14344.1Sucrose phosphorylase
(LmSPase)
Leuconostoc mesenteroides ATCC 12291 α-Glc1Pd-fructoseSucrose[37]
GH13_18RetainingABS59292.1Sucrose phosphorylase
(742sp)
Leuconostoc mesenteroides B-742 α-Glc1Pd-fructoseSucrose[38]
GH13_18Retaining2207198ASucrose phosphorylaseLeuconostoc mesenteroides NO. 165 α-Glc1Pd-fructoseSucrose[39]
GH13_18RetainingAAX33736.1Sucrose phosphorylase
(LmSP1)
Leuconostoc mesenteroides NRRL B-1149 α-Glc1Pd-fructoseSucrose[40]
GH13_18RetainingAGK37834.1Sucrose phosphorylase
(ScrP)
Limosilactobacillus reuteri LTH5448 α-Glc1Pd-fructoseSucrose[41]
GH13_18RetainingADP98617.1Glucosylglycerol phosphorylase
(HP15_2853)
Marinobacter adhaerens HP15 α-Glc1Pd-glycerolGlucosyl-glycerol:
(α-d-glucopyranosyl-(1→2)-glycerol)
[42]
GH13_18RetainingADH62582.1Glucosylglycerate phosphorylase
(MSGGaP;Mesil_0665)
Meiothermus silvanus DSM 9946 α-Glc1Pd-glycerateGlucosyl-glycerate:
(α-d-glucopyranosyl-(1→2)-glycerate)
[28]
GH13_18RetainingAAD40317.1Sucrose phosphorylasePelomonas saccharophila IAM 14368 α-Glc1Pd-fructoseSucrose[43]
GH13_18RetainingCCA61958.1Sucrose 6(F)-phosphate phosphorylase
(RUGNEv3_61221;SucP)
Ruminococcus gnavus E1 α-Glc1Pd-fructose-6-PSucrose-6-phosphate[44]
GH13_18RetainingAEJ61152.1Glucosylglycerate phosphorylase (Spith_0877)Spirochaeta thermophila DSM 6578 α-Glc1Pd-glycerateGlucosyl-glycerate[28]
GH13_18RetainingCAA30846.1Sucrose phosphorylase
(GftA;SmSP)
Streptococcus mutans INGBRITT/GS5 α-Glc1Pd-fructoseSucrose[45]
GH13_18RetainingAAN58596.1Sucrose phosphorylase
(GtfA;SMU.881)
Streptococcus mutans UA159 α-Glc1Pd-fructoseSucrose[46]
GH13_18RetainingCAA80424.1*Sucrose phosphorylaseAgrobacterium vitis α-Glc1Pd-fructoseSucrose[47]
GH13_18RetainingADL69407.1Sucrose 6(F)-phosphate phosphorylase
(SPP;TtSPP;Tthe_1921)
Thermoanaerobacterium thermosaccharolyticum DSM 5716S9Vα-Glc1Pd-fructose-6-PSucrose-6-phosphate[48]
GH13_18RetainingACJ83248.1Sucrose phosphorylase
(Unspase)
Uncultured bacterium α-Glc1Pd-fructoseSucrose[49]
GH13_18RetainingBAN03569.1Sucrose 6(F)-phosphate phosphorylase
(YM304_32550)
Ilumatobacter coccineus YM16-3046S9Uα-Glc1Pd-fructose-6-PSucrose-6-phosphate[50]
GH65InvertingAAV43670.1Maltose phosphorylase
(MalP;LBA1870)
Lactobacillus acidophilus NCFM β-Glc1Pd-glucoseMaltose:
(α-d-glucosyl-(1→4)-d-glucose)
[51]
GH65InvertingADH99560.1Maltose phosphorylase
(Bsel_2056)
Bacillus selenitireducens MLS10 β-Glc1Pd-glucoseMaltose[52]
GH65InvertingBAC54904.1Maltose phosphorylase
(MPase)
Bacillus sp. RK-1 β-Glc1Pd-glucoseMaltose[53]
GH65InvertingBAD97810.1Maltose phosphorylase
(MapA)
Paenibacillus sp. SH-55 β-Glc1Pd-glucoseMaltose[27]
GH65InvertingCAA11905.1Maltose phosphorylaseFructilactobacillus sanfranciscensis DSM 20451(T) β-Glc1Pd-glucoseMaltose[54]
GH65InvertingAAO80764.1Maltose phosphorylase
(MalP;EF0957
Enterococcus faecalis V583 β-Glc1Pd-glucoseMaltose[55]
GH65InvertingQ7SIE1Maltose phosphorylaseLevilactobacillus brevis ATCC 8287 1H54β-Glc1Pd-glucoseMaltose[56]
GH65InvertingBAQ19546.1Maltose phosphorylase
(MalE)
Bacillus sp. AHU2001 β-Glc1Pd-glucoseMaltose[57]
GH65InvertingBAB97299.1Trehalose phosphorylase
(TreP)
Thermoanaerobacter brockii ATCC 35047Preliminary data
PMID: 20383018
β-Glc1Pd-glucoseTrehalose:
(α-d-glucosyl-(1→1)-d-glucose)
[58]
GH65InvertingBAC20640.1Trehalose phosphorylase
(TPase)
Geobacillus stearothermophilus SK-1 β-Glc1Pd-glucoseTrehalose[59]
GH65InvertingKKC30106.1Trehalose phosphorylase
(TP)
Caldanaerobacter subterraneus subsp. Pacificus DSM 12653 β-Glc1Pd-glucoseTrehalose[60]
GH65InvertingADG89586.1Trehalose phosphorylase (TbGP; Tbis_2887)Thermobispora bispora DSM 43833 β-Glc1Pd-glucoseTrehalose[61]
GH65InvertingAAK04526.1Trehalose-6-phosphate phosphorylase (TrePP;YeeA;L39593;
LL0428)
Lactococcus lactis subsp. lactis Il1403 β-Glc1Pd-glucose-6-phosphateTrehalose-6-phosphate:
(α-d-glucosyl-(1→1)-d-glucopyranoside-6-phosphate)
[62]
GH65InvertingAPF18594.1Trehalose-6-phosphate phosphorylase
(CaGP;Cabys_1845)
Caldithrix abyssi DSM 13497 LF13 β-Glc1Pd-glucose-6-phosphateTrehalose-6-phosphate[61]
GH65InvertingAAE30762.1Kojibiose phosphorylase
(KojP;KPase)
Thermoanaerobacter brockii ATCC 35047 β-Glc1Pd-glucoseKojibiose:
(α-d-glucosyl-(1→2)-d-glucose)
[58]
GH65InvertingAAC74398.1Kojibiose phosphorylase
(YcjT)
Escherichia coli str. K-12 substr. MG1655 β-Glc1Pd-glucoseKojibiose[63]
GH65InvertingABP66077.1Kojibiose phosphorylase
(CsKP;Csac_0439)
Caldicellulosiruptor saccharolyticus DSM 89033WIQ, 3WIRβ-Glc1Pd-glucoseKojibiose[64]
GH65InvertingACL68803.1Kojibiose phosphorylase
(HoGP; Hore_00420)
Halothermothrix orenii H 168 β-Glc1Pd-glucoseKojibiose[61]
GH65InvertingABX42243.1Nigerose phosphorylase
(Cphy_1874)
Lachnoclostridium phytofermentans ISDg β-Glc1Pd-glucoseNigerose:
(α-d-glucosyl-(1→3)-d-glucose)
[65]
GH65InvertingABX43667.1Nigerose phosphorylase
(Cphy_3313)
Lachnoclostridium phytofermentans ISDg β-Glc1Pd-glucoseNigerose[66]
GH65InvertingADQ05832.11,3-α-oligoglucan phosphorylase (ChGP;Calhy_0070)Caldicellulosiruptor hydrothermalis 108 β-Glc1PMaltose
Kojibiose
Nigerose
α-1,3-oligoglucan[61]
GH65InvertingADC90669.1Trisaccharide α-glucan phosphorylase
(MiGP;HMPREF0868_0806)
Mageeibacillus indolicus UPII9-5 β-Glc1PIsomaltose
Isomaltulose
Maltose
Kojibiose
Nigerose
α-1,3-oligoglucan[61]
GH65InvertingABX43668.1α-1,3-oligoglucan phosphorylase
(Cphy_3314)
Lachnoclostridium phytofermentans ISDg β-Glc1Pd-glucoseα-1,3-oligoglucan[66]
GH65InvertingABX41399.13-O-α-glucopyranosyl-L-rhamnose phosphorylase
(Cphy_1019)
Lachnoclostridium phytofermentans ISDg β-Glc1Pl-rhamnoseα-d-glucopyranosyl-(1,3)-l-rhamnose[25]
GH65InvertingADI00307.1α-1,2-glucosylglycerol phosphorylase
(Bsel_2816)
Bacillus selenitireducens MLS104KTP, 4KTRβ-Glc1Pd-glycerol(α-d-glucosyl-(1→2)-glycerol)[67]
GH94InvertingAAB95491.2Cellobiose phosphorylase
(CbpA)
Thermotoga neapolitana α-Glc1Pd-glucoseCellobiose:
(β-d-glucopyranosyl-(1,4)-d-glucopyranose)
[68]
GH94InvertingAAC45510.1Cellobiose phosphorylase
(CepA)
Thermoclostridium stercorarium NCIB 11754 α-Glc1Pd-glucoseCellobiose[69]
GH94InvertingAAD36910.1Cellobiose phosphorylase
(CepA;TM1848;Tmari_1863
Thermotoga maritima MSB8 α-Glc1Pd-glucoseCellobiose[70]
GH94InvertingAAL67138.1Cellobiose phosphorylase
(Cbp)
Acetivibrio thermocellus YM43QDEα-Glc1Pd-glucoseCellobiose[71]
GH94InvertingAAQ20920.1Cellobiose phosphorylase
(CbP)
Cellulomonas uda DSM 20108 3RRS, 3RSY, 3S4A, 3S4B, 3S4C, 3S4Dα-Glc1Pd-glucoseCellobiose[72]
GH94InvertingABD80580.1Cellobiose phosphorylase 94A
(Cbp;SdCBP;Sde1318, Cep94A)
Saccharophagus degradans 2-40 α-Glc1Pd-glucoseCellobiose[73]
GH94InvertingABN51514.1Cellobiose phosphorylase
(Cbp;CtCBP;Cthe_0275)
Acetivibrio thermocellus ATCC 27405 α-Glc1Pd-glucoseCellobiose[74]
GH94InvertingACL76454.1Cellobiose phosphorylase
(CbpA;Ccel_2109)
Ruminiclostridium cellulolyticum H10 α-Glc1Pd-glucoseCellobiose[75]
GH94InvertingADU20744.1Cellobiose phosphorylase
(CBP;RaCBP;Rumal_0187)
Ruminococcus albus 7 (DSM 20455) α-Glc1Pd-glucoseCellobiose[24]
GH94InvertingBAA25846.1Cellobiose-phosphorylase
(Cbp)
Acetivibrio thermocellus YM4 α-Glc1Pd-glucoseCellobiose[76]
GH94InvertingBAA28631.1Cellobiose phosphorylase
(Cbp;CgCBP)
Cellulomonas gilvus ATCC131272CQS, 2CQT, 3ACS, 3ACT, 3AFJ, 3QFY, 3QFZ, 3QG0α-Glc1Pd-glucoseCellobiose[77]
GH94InvertingCAB16926.1*Cellobiose phosphorylase
(CepA)
Thermotoga neapolitana Z2706-MC24 α-Glc1Pd-glucoseCellobiose[78]
GH94InvertingAAC45511.1Cellodextrin phosphorylase
(CepB; CsCdP)
Thermoclostridium stercorarium DSM8532 α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[69]
GH94InvertingACL76454.1Cellobiose phosphorylase (CbpA;Ccel_2109)Ruminiclostridium cellulolyticum H10 α-Glc1Pd-glucoseCellobiose[29]
GH94InvertingQCO92836.1Cellobiose phosphorylase
(GH94_G)
Uncultured bacterium α-Glc1Pd-glucoseCellobiose[79]
GH94InvertingQCO92799.1Cellobiose phosphorylase
(GH94_E)
Uncultured bacterium α-Glc1Pd-glucoseCellobiose[79]
GH94InvertingADU22883.1Cellodextrin phosphorylase
(CDP;RaCDP;Rumal_2403)
Ruminococcus albus 7 (DSM 20455) α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[80]
GH94InvertingACD71661.1Cyclic β-1,2-glucan synthetase
(Cgs)
Brucella abortus S19 α-Glc1P[(1→2)-β-d-glucosyl]n[(1→2)-β-d-glucosyl]n+1[81]
GH94InvertingACL75793.1Cellodextrin phosphorylase
(CdpA; Ccel_1439)
Ruminiclostridium cellulolyticum H10 α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[29]
GH94InvertingACL77700.1Cellodextrin phosphorylase
(CdpC; Ccel_3412)
Ruminiclostridium cellulolyticum H10 α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[29]
GH94InvertingACL76688.1Cellodextrin phosphorylase
(CdpB; Ccel_2354)
Ruminiclostridium cellulolyticum H10 α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[29]
GH94InvertingQCO92991.1Cellodextrin phosphorylase
(GH94_A)
Uncultured bacterium α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[29]
GH94InvertingQCO92946.1Cellodextrin phosphorylase
(GH94_D)
Uncultured bacterium α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[29]
GH94InvertingBAJ10826.1Laminaribiose phosphorylase
(LbpA)
Paenibacillus sp. YM16GGY, 6GH2, 6GH3 α-Glc1Pd-glucoseLaminaribiose:
(β-d-glucopyranosyl-(1,3)-d-glucopyranose)
[82,83]
GH94InvertingAAG23740.1Diacetylchitobiose phosphorylase
(ChbP)
Vibrio furnissii α-GlcNAc-1PN-acetyl-d-glucosamineN,N’-diacetylchitobiose:
(N-acetyl-d-glucosaminyl-(1,4)-N-acetyl-d-glucosamine)
[84]
GH94InvertingBAC87867.1Chitobiose phosphorylase
(ChbP;VpChBP)
Vibrio proteolyticus1V7V, 1V7W, 1V7Xα-GlcNAc-1PN-acetyl-d-glucosamineN,N’-diacetylchitobiose[85]
GH94InvertingAFH48717.1Chitobiose phosphorylase (IaGP;IALB_1006)Ignavibacterium album JCM 16511 α-GlcNAc-1PN-acetyl-d-glucosamineN,N’-diacetylchitobiose[86]
GH94InvertingCAC97070.1β-1,2-oligoglucan phosphorylase
(LiSOGP;Lin1839)
Listeria innocua Clip11262 α-Glc1P[(1→2)-β-d-glucosyl]n[(1→2)-β-d-glucosyl]n+1[87]
GH94InvertingABX41081.1β-1,2-oligoglucan phosphorylase
(LpSOGP;Cphy_0694)
Lachnoclostridium phytofermentans ISDg5H3Z, 5H40, 5H41, 5H42 α-Glc1P[(1→2)-β-d-glucosyl]n[(1→2)-β-d-glucosyl]n+1[88]
GH94InvertingQCO92990.1Laminaribiose phosphorylase
(GH94_B)
Uncultured bacterium α-Glc1Pd-glucoseLaminaribiose[79]
GH94InvertingQCO92945.1Laminaribiose phosphorylase
(GH94_C)
Uncultured bacterium α-Glc1Pd-glucoseLaminaribiose[79]
GH94InvertingBAB71818.1Cellodextrin phosphorylase
(Cdp-ym4)
Acetivibrio thermocellus YM4 α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[71]
GH94InvertingABN54185.1Cellodextrin-phosphorylase
(Cdp;CtCDP;Cthe_2989)
Acetivibrio thermocellus ATCC 274055NZ7, 5NZ8α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[89]
GH94InvertingADZ85667.1*Cellodextrin phosphorylase
(Cdp;ClCDP;Clole_3989)
Clostridium lentocellum α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[73]
GH94InvertingAAM43298.1Cellobionic acid phosphorylase
(CelAP;NdvB;XCC4077)
Xanthomonas campestris pv. campestris str. ATCC 33913 α-Glc1Pd-gluconateβ-d-glucopyranosyl-(1,4)-d-gluconate[90]
GH94InvertingABD80168.1Cellobionic acid phosphorylase 94B
(Cep94B;CBAP;
SdCBAP;Sde_0906)
Saccharophagus degradans 2-40 4ZLE, 4ZLF, 4ZLG, 4ZLIα-Glc1Pd-gluconateβ-d-glucopyranosyl-(1,4)-d-gluconate[91]
GH94InvertingEAA28929.1Cellobionic acid phosphorylase
(CelAP;NdvB;
NCU09425)
Neurospora crassa OR74A α-Glc1Pd-gluconateβ-d-glucopyranosyl-(1,4)-d-gluconate[90]
GH94InvertingAAC45511.1Cellodextrin phosphorylase
(CepB;CsCdP)
Thermoclostridium stercorarium DSM8532 α-Glc1P[(1→4)-β-d-glucosyl]n[(1→4)-β-d-glucosyl]n+1[69]
GH94InvertingAHC20114.1Glucosylgalactose Phosphorylase
(GGalP)
Paenibacillus polymyxa CR1 α-Glc1Pd-galactoseβ-d-glucopyranosyl-(1,4)-d-galactose[86]
GH94InvertingABX81345.1Laminaribiose phosphorylase
(ACL_0729;ACL0729)
Acholeplasma laidlawii PG-8A α-Glc1Pd-glucoseLaminaribiose[92]
GH112InvertingACB74662.1D-galactosyl-β-1,4-L-rhamnose phosphorylase
(GalRhaP;Oter_1377)
Opitutus terrae PB90-1 α-Gal1Pl-rhamnose(β-d-galactosyl-(1,4)-l-rhamnose)[93]
GH112InvertingABX42289.1D-galactosyl-β-1,4-L-rhamnose phosphorylase
(Cphy_1920)
Lachnoclostridium phytofermentans ISDg α-Gal1Pl-rhamnose(β-d-galactosyl-(1,4)-l-rhamnose)[94]
GH112InvertingAAO07997.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase
(GalGlyNAcP;VV2_1091)
Vibrio vulnificus CMCP6 α-Gal1PN-acetyl-d-glucosamine(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)[95]
GH112InvertingABX40964.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase
(Cphy_0577)
Lachnoclostridium phytofermentans ISDg α-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[94]
GH112InvertingABX43387.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase
(Cphy_3030)
Lachnoclostridium phytofermentans ISDg α-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[94]
GH112InvertingAEN95946.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase (RhGLnbp112;RHOM_04120)Roseburia hominis A2-183 α-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[96]
GH112InvertingEEG94248.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase
(RiGLnbp112; ROSEINA2194_01886; GnpA)
Roseburia inulinivorans DSM 16841 α-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[96]
GH112InvertingBAD80751.1Galacto-N-biose/lacto-N-biose phosphorylase
(LnpA1;LnbP;GLNBP;
BLLJ_1623)
Bifidobacterium longum subsp. longum JCM 12172ZUS, 2ZUT, 2ZUU, 2ZUV, 2ZUW, 3WFZα-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[97]
GH112InvertingBAD80752.1Lacto-N-biose phosphorylase
(LnpA1;LnbP)
Bifidobacterium bifidum JCM 1254 α-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[98]
GH112InvertingACV29689.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase/galacto-N-biose/lacto-N-biose I phosphorylase
(GLNBP;Apre_1669)
Anaerococcus prevotii DSM 20548 α-Gal1PN-acetyl-d-glucosamine
N-acetyl-d-galactosamine
(β-d-galactosyl-(1,3)-N-acetyl-d-glucosamine)
(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)
[26]
GH112InvertingZP_05748149.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase/galacto-N-biose phosphorylase
(GNBP;HMPREF0357_1319)
Erysipelothrix rhusiopathiae ATCC 19414 α-Gal1PN-acetyl-d-galactosamine(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)[26]
GH112InvertingACZ00636.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase/galacto-N-biose phosphorylase
(GNBP;Smon_0146)
Streptobacillus moniliformis DSM 12112 α-Gal1PN-acetyl-d-galactosamine(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)[26]
GH112InvertingABG83511.1Galacto-N-biose phosphorylase
(CPF_0553)
Clostridium perfringens ATCC 13124 α-Gal1PN-acetyl-d-galactosamine(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)[95]
GH112InvertingBAH10636.1β-1,3-galactosyl-N-acetylhexosamine phosphorylase
(GnpA)
Cutibacterium acnes JCM 6425 α-Gal1PN-acetyl-d-galactosamine(β-d-galactosyl-(1,3)-N-acetyl-d-galactosamine)[99]
GH130InvertingCAC96089.1β-1,2-mannobiose phosphorylase
(Lin0857)
Listeria innocua Clip112625B0P, 5B0Q, 5B0R, 5B0Sα-Man1Pd-mannoseβ-1,2-mannobiose[100]
GH130InvertingABY93074.1β-1,2-mannobiose phosphorylase
(Teth514_1789)
Thermoanaerobacter sp. X5147FIP, 7FIQ, 7FIR, 7FISα-Man1Pd-mannoseβ-1,2-mannobiose[101]
GH130InvertingABY93073.1β-1,2-oligomannan phosphorylase
(Teth514_1788)
Thermoanaerobacter sp. X514 α-Man1P(β-1,2-d-mannose)n(β-1,2-d-mannose)n+1[101]
GH130InvertingCAZ94304.1β-1,3-mannooligosaccharide phosphorylase
(zobellia_231)
Zobellia galactanivorans DsijT α-Man1P(β-1,3-d-mannose)n(β-1,3-d-mannose)n+1[102]
GH130InvertingAAS19693.1β-1,4-mannosylglucose phosphorylase
(UnkA, CvMGP)
Cellvibrio vulgaris NCIMB8633 α-Man1Pd-glucoseβ-mannopyranosyl-(1→4)-d-glucopyranose[103]
GH130InvertingADU21379.1β-1,4-mannosylglucose phosphorylase
(RaMP1; RaMGP; Rumal_0852)
Ruminococcus albus 7 (DSM 20455)5AY9, 5AYCα-Man1Pd-glucoseβ-mannopyranosyl-(1→4)-d-glucopyranose[104]
GH130InvertingCAH06518.1β-1,4-mannosylglucose phosphorylase
(MGP; BF0772)
Bacteroides fragilis NCTC 93433WAS, 3WAT, 3WAU, 4KMIα-Man1Pd-glucoseβ-mannopyranosyl-(1→4)-d-glucopyranose[105]
GH130InvertingVCV21228.1β-1,4-mannosylglucose phosphorylase
(RIL182_01099; ROSINTL182_07685)
Roseburia intestinalis L1-82 α-Man1Pd-glucoseβ-mannopyranosyl-(1→4)-d-glucopyranose[106]
GH130InvertingACY49318.1β-1,4-mannosylglucose phosphorylase
(MGP; RmMGP; Rmar_2440)
Rhodothermus marinus R-10T (ATCC43812) α-Man1Pd-glucoseβ-mannopyranosyl-(1→4)-d-glucopyranose[107]
GH130InvertingAAO76140.1β-1,4-mannosyl-N-acetyl-glucosamine phosphorylase
(BT1033)
Bacteroides thetaiotaomicron VPI-5482 α-Man1PN-acetyl-d-glucosamineβ-mannopyranosyl-(1→4)-N-acetyl-d-glucosamine[108]
GH130InvertingADD61463.1β-1,4-mannopyranosyl-[N-glycan] phosphorylase
β-1,4-mannopyranosyl-chitobiose phosphorylase
(Uhgb_MP)
Uncultured bacterium4UDG, 4UDI, 4UDJ, 4UDKα-Man1Pβ-N-acetyl-d-glucosaminyl-(1,4)-N-acetyl-d-glucosamineβ-1,4-mannopyranosyl-chitobiose
(β-mannopyranosyl-(1→4)-N-acetyl-β-d-glucosaminyl-(1→4)-N-acetyl-d-glucosamine)
[109]
GH130InvertingADU20661.1β-1,4-mannooligosaccharide phosphorylase
(MOP; RaMP2; Rumal_0099)
Ruminococcus albus 7 (DSM 20455)5AYD, 5AYEα-Man1P(β-1,4-d-mannose)n(β-1,4-d-mannose)n+1[104]
GH130InvertingVCV21229.1β-1,4-mannooligosaccharide phosphorylase
(RIL182_01100; ROSINTL182_05474)
Roseburia intestinalis L1-82 α-Man1P(β-1,4-d-mannose)n(β-1,4-d-mannose)n+1[106]
GH130InvertingAAD36300.1β-1,4-mannooligosaccharide phosphorylase
(TM1225)
Thermotoga maritima MSB81VKDα-Man1P(β-1,4-d-mannose)n(β-1,4-d-mannose)n+1[110]
GH130InvertingWP_026485574.1β-1,4-mannooligosaccharide phosphorylase
(CalpoDRAFT_0075)
Caldanaerobius polysaccharolyticus ATCC BAA-17 α-Man1P(β-1,4-d-mannose)n(β-1,4-d-mannose)n+1[111]
GH130InvertingWP_026486530.1β-1,4-mannooligosaccharide phosphorylase
(CalpoDRAFT_1209)
Caldanaerobius polysaccharolyticus ATCC BAA-17 α-Man1P(β-1,4-d-mannose)n(β-1,4-d-mannose)n+1[111]
GH130Inverting
CAH1385115.1
β-1,4-mannosyl-glucuronic acid phosphorylaseUncultured bacterium α-Man1Pd-glucuronic acidβ-mannopyranosyl-(1→4)-d-glucuronic acid[112]
GH130Inverting
CAH1385116.1
β-1,4-mannooligosaccharide phosphorylaseUncultured bacterium α-Man1Pd-mannose
β-1,4-mannobiose
β-1,4-mannobiose
β-1,4-mannotriose
[112]
GH130Inverting
CAH1385117.1
β-1,3-mannosyl-glucose phosphorylase
β-1,3-mannooligosaccharide phosphorylase
Uncultured bacterium α-Man1Pd-glucose
d-mannose
β-mannopyranosyl-(1→3)-d-glucopyranose
β-1,3-mannobiose
[112]
GH130Inverting
CAH1385118.1
Mixed-linkage mannoside synthase
Uhgb_MS
Uncultured bacterium α-Man1Pd-mannose
α-1,2-mannobiose
β-1,2-mannobiose

α-1,2-mannotriose
β-d-mannopyranosyl-(1→2)-α-1,2-mannotriose
[113]
GH149Inverting β-1,3-oligoglucan phosphorylase
(Pro_7066)
Uncultured bacterium6HQ6, 6HQ8, 6HQ6α-Glc1P[(1→3)-β-d-glucosyl]n[(1→3)-β-d-glucosyl]n+1[114]
GH149InvertingAUO30192.1β-1,3-glucan phosphorylase
(EgP1)
Euglena gracilis α-Glc1P[(1→3)-β-d-glucosyl]n[(1→3)-β-d-glucosyl]n+1[115]
GH149InvertingQCO92814.1β-1,3-glucan phosphorylase
(GH149_H)
Uncultured bacterium α-Glc1P[(1→3)-β-d-glucosyl]n[(1→3)-β-d-glucosyl]n+1[79]
GH161InvertingWP_019688419.1β-1,3-glucan phosphorylase
(PPT_RS0121460; PapP)
Paenibacillus polymyxa ATCC 842 α-Glc1P[(1→3)-β-d-glucosyl]n[(1→3)-β-d-glucosyl]n+1[116]
GH161InvertingACJ76363.1β-1,3-glucan phosphorylase
(THA_1941; TaCDP)
Thermosipho africanus TCF52B α-Glc1P[(1→3)-β-d-glucosyl]n[(1→3)-β-d-glucosyl]n+1[116]
GH161InvertingBAU78234.1β-1,3-glucan phosphorylase
(OdBGP)
Ochromonas danica NIES-2142 α-Glc1P[(1→3)-β-d-glucosyl]n[(1→3)-β-d-glucosyl]n+1[117]
GT4RetainingBAA31350.1Trehalose synthase
(Tsase)
Grifola frondosa α-Glc1Pd-glucoseTrehalose:
(α-d-glucosyl-(1,1)-d-glucose)
[118]
GT4RetainingAAF22230.1Threalose phosphorylase
(PsTP)
Lentinus sajor-caju ASI2070
(Pleurotus sajor-caju)
α-Glc1Pd-glucosetrehalose[119]
GT4RetainingABC84380.1Threalose phosphorylaseSchizophyllum commune BT2115 α-Glc1Pd-glucosetrehalose[120,121]
GT35RetainingABP51432.1Glucan/maltodextrin phosphorylase
(GlgP; PyglgP; Pars_1881)
Pyrobaculum arsenaticum DSM 13514 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[122]
GT35RetainingAAL81659.1α-glucan/maltodextrin phosphorylase
(PF1535)
Pyrococcus furiosus DSM 3638 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[123,124]
GT35RetainingAAD28735.1Maltodextrin phosphorylase
(MalP)
Thermococcus litoralis α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[125]
GT35RetainingABN51595.1α-glucan phosphorylase
(α-GP; Cthe_0357)
Acetivibrio thermocellus ATCC 27405 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[126]
GT35RetainingCAB61943.1α-glucan phosphorylase
(PHS2; AtPHS2; At3g46970)
Arbidopsis thaliana α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[127]
GT35RetainingAAD03471.1Glycogen phosphorylase
(GlgP)
Agrobacterium tumefaciens A348 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[128]
GT35RetainingAAC06896.1Glycogen phosphorylase
(GlgP; Aq_717)
Aquifex aeolicus VF5 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[129]
GT35RetainingAAC00218.1Glycogen phosphorylase
(GlgP; BSU30940)
Bacillus subtilis subsp. subtilis str. 168 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[130]
GT35RetainingAAM24997.1α-glucan phosphorylase
(GlgP; TTE1805)
Caldanaerobacter subterraneus subsp. tengcongensis MB4 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[131]
GT35RetainingAAM52219.1Glycogen phosphorylaseCorynebacterium callunae 2C4Mα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[132]
GT35RetainingBAB98701.1Maltodextrin/glycogen phosphorylase
(MalP; GlgP1; cg1479)
Corynebacterium glutamicum ATCC 13032 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[133]
GT35RetainingBAB99480.1Glycogen phosphorylase
(GlgP; GlgP2; cg2289)
Corynebacterium glutamicum ATCC 13032 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[133]
GT35RetainingAAC76453.1Glycogen phosphorylase
(GlgP; b3428)
Escherichia coli str. K-12 substr. MG1655 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[134]
GT35RetainingAAC76442.1Maltodextrin phosphorylase
(MalP; b3417)
Escherichia coli str. K-12 substr. MG16551AHP, 1E4O, 1L5V, 1L5W, 1L6I, 1QM5, 2ASV, 2AV6, 2AW3, 2AZD, 2ECPα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[135]
[136]
GT35RetainingBAA19592.1Glycogen phosphorylase
(GlgP)
Geobacillus stearothermophilus TRBE14 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[137]
GT35RetainingAAN59210.1Maltodextrin phosphorylase
(GlgP; SMU.1564)
Streptococcus mutans UA1594L22α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[138]
GT35RetainingAAL26558.1Glycogen phosphorylase
(Glg)
Synechococcus elongatus PCC 7942 = FACHB-805 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[139]
GT35RetainingAGL50099.1α-glucan phosphorylase
(AgpA; Tmari_1175)
Thermotoga maritima MSB8 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[140]
GT35RetainingBAB11741.1α-glucan phosphorylase
(GlgP)
Thermus aquaticus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[141]
GT35RetainingCAC93400.1Glycogen phosphorylase
(glgP, YPO3938)
Yersinia pestis CO92 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[142]
GT35RetainingAAB46846.1Glycogen phosphorylase
(myophosphorylase)
Bos taurus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[143]
GT35RetainingABB88567.1Plastidial starch phosphorylase
(PhoB; Sta4)
Chlamydomonas reinhardtii 137C α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[144]
GT35RetainingCAA44069.1Glycogen phosphorylase 1
(GlpV; GP1)
Dictyostelium discoideum α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[145]
GT35RetainingAAA33211.1Glycogen phosphorylase 2
(GlpD; GP2)
Dictyostelium discoideum α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[146]
GT35RetainingAAD46887.1glycogen phosphorylase
(GlyP; GLYP; CG7254; Dmel_CG7254)
Drosophila melanogaster α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[147]
GT35RetainingAAN17338.1Glycogen phosphorylase-2Entamoeba histolytica α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[148]
GT35RetainingAAL23578.1Glycogen phosphorylaseEntamoeba histolytica α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[148]
GT35RetainingAAP33020.1Glycogen phosphorylaseGallus gallus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[149]
GT35RetainingAAK69600.1Glycogen phosphorylaseGiardia intestinalis ATCC30957 α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[148]
GT35RetainingAAB60395.1Glycogen phosphorylase (brain)
(bGP)
Homo sapiens5IKO, 5IKPα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[150]
GT35RetainingCAA75517.1Glycogen phosphorylase (liver)Homo sapiens1EM6, 1EXV, 1FA9, 1FC0, 1L5Q, 1L5R, 1L5S, 1L7X, 1XOI, 2ATI, 2QLL, 2ZB2, 3CEH, 3CEJ, 3CEM, 3DD1, 3DDS, 3DDWα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[151]
GT35RetainingAAC17451.1Glycogen phosphorylase (muscle)
(PigM)
Homo sapiens1Z8Dα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[152]
GT35RetainingBAK00834.1Plastidial α-1,4-glucan phosphorylase
(Pho1; HvPho1)
Hordeum vulgare subsp. Vulgare5LR8, 5LRA, 5LRBα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[153]
GT35RetainingAAA63271.1α-glucan phosphorylase LIpomoea batatas α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[154]
GT35RetainingAAK01137.1Starch phosphorylase
(fragment)
Ipomoea batatas α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[155]
GT35RetainingAAL23577.1Glycogen phosphorylaseMastigamoeba balamuthi α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[148]
GT35RetainingAAD30476.1Glycogen phosphorylase (muscle)
(PygM)
Mus musculus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[156]
GT35RetainingAAG00588.1Glycogen phosphorylaseMus musculus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[157]
GT35RetainingACJ76617.1Glycogen phosphorylase (muscle)
(PygM)
Oryctolagus cuniculusDozens available—2GJ4[A] best solutionα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[158,159]
GT35RetainingAAK15695.1α-1,4-glucan phosphorylase LOryza sativa α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[160]
GT35RetainingBAB92854.1α-1,4-glucan phosphorylase (Os01g0851700)Oryza sativa Japonica Group α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[161]
GT35RetainingAAV87308.1brain glycogen phosphorylase (PYGB)Ovis aries α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[162]
GT35RetainingAAB68800.1glycogen phosphorylase (muscle)Ovis aries α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[163]
GT35RetainingAAA41252.1glycogen phosphorylase (Brain)Rattus norvegicus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[164]
GT35RetainingAAH70901.1glycogen phosphorylase (liver)Rattus norvegicus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[165]
GT35RetainingAAA41253.1glycogen phosphorylase (muscle)Rattus norvegicus α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[166]
GT35RetainingAAB68057.1Glycogen phosphorylase
(Gph1; YPR160w)
Saccharomyces cerevisiae S288c1YGPα-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[167,168]
GT35RetainingAAA33809.1α-glucan phosphorylase HSolanum tuberosum α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[169]
GT35RetainingBAA00407.1α-glucan phosphorylase L1Solanum tuberosum α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[170]
GT35RetainingCAA52036.1α-glucan phosphorylase L2Solanum tuberosum α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[171]
GT35RetainingCAA59464.1α-glucan phosphorylaseSpinacia oleracea α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[159]
GT35RetainingAAL23579.1glycogen phosphorylaseTrichomonas vaginalis α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[148]
GT35RetainingAAF82787.1α-glucan phosphorylaseTriticum aestivum α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[172]
GT35RetainingCAA84494.1α-glucan phosphorylase (Pho2;VfPho2)Vicia faba var. minor α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[173]
GT35RetainingCAA85354.1α-glucan phosphorylase LVicia faba var. minor α-Glc1P[(1→4)-α-d-glucosyl]n[(1→4)-α-d-glucosyl]n+1[173]
GT108InvertingCBZ24448.1MTP3
(LMXM_10_1250)
Leishmania mexicana MHOM/GT/2001/U1103 α-Man1P(β-1,2-d-mannose)n(β-1,2-d-mannose)n+1[174]
GT108InvertingCBZ24451.1MTP6
(LMXM_10_1280)
Leishmania mexicana MHOM/GT/2001/U1103 α-Man1P(β-1,2-d-mannose)n(β-1,2-d-mannose)n+1[174]
GT108InvertingCBZ24449.1MTP4
(LMXM_10_1260)
Leishmania mexicana MHOM/GT/2001/U11036Q50α-Man1P(β-1,2-d-mannose)n(β-1,2-d-mannose)n+1[174]
GT108InvertingCBZ24452.1MTP7
(LMXM_10_1290)
Leishmania mexicana MHOM/GT/2001/U1103 α-Man1P(β-1,2-d-mannose)n(β-1,2-d-mannose)n+1[174]
Table 2. Structurally characterized GPs. Only the enzymes proven to have GP activity are listed. ACJ76617.1 has dozens of PDBs, only the PDB with the best solution is listed.
Table 2. Structurally characterized GPs. Only the enzymes proven to have GP activity are listed. ACJ76617.1 has dozens of PDBs, only the PDB with the best solution is listed.
CAZy FamilyGenBank/
UniProtKB Accession
Organism NamePDB Accession NumbersProtein NameFoldQuaternary StructureReference
GH3AUG44408.1Uncultured bacterium5VQD[A], 5VQE[A]β-glucoside phosphorylase
(BglP)
(β/α)8Monomer[30]
GH13_18AAO33821.1Bifidobacterium adolescentis DSM 200831R7A[A,B], 2GDU[A,B], 2GDV[A,B], 5C8B[B], 5M9X[B], 5MAN[B], 5MB2[B], 6FME[A,B]Sucrose phosphorylase
(SucP; SP; BaSP)
(β/α)8Dimer[196]
GH13_18BAN03569.1Ilumatobacter coccineus YM16-3046S9U[A]Sucrose 6(F)-phosphate phosphorylase
(YM304_32550)
(β/α)8Monomer[50]
GH13_18ADL69407.1Thermoanaerobacterium thermosaccharolyticum DSM 5716S9V[A,B]Sucrose 6(F)-phosphate phosphorylase
(SPP; TtSPP; Tthe_1921)
(β/α)8Monomer[50]
GH65Q7SIE1Levilactobacillus brevis ATCC 82871H54[A]Maltose phosphorylase(α/α)6Dimer[197]
GH65ADI00307.1Bacillus selenitireducens MLS104KTP[A,B], 4KTR[A,B,C,D,E,F,G,H]2-O-α-glucopyranosylglycerol: phosphate β-glucosyltransferase/
2-O-α-glucosylglycerol phosphorylase
(GGP; Bsel_2816)
(α/α)6Dimer[198]
GH65ABP66077.1Caldicellulosiruptor saccharolyticus DSM 89033WIQ[A], 3WIR[A,B,C,D]Kojibiose phosphorylase
(CsKP; Csac_0439
(α/α)6Dimer[199]
GH65/Uncultured bacterium6W0P[B,C,D], 6W0P[A]Kojibiose phosphorylase(α/α)6DimerUnpublished
GH65BAB97299.1Thermoanaerobacter brockii ATCC 35047 Trehalose phosphorylase
(TreP)
(α/α)6Dimer[200]
GH94ABN54185.1Acetivibrio thermocellus ATCC 274055NZ7[A,B], 5NZ8[A,B]Cellodextrin-phosphorylase
(Cdp; CtCDP; Cthe_2989)
(α/α)6Dimer[201]
GH94AAL67138.1Acetivibrio thermocellus YM43QDE[A,B]Cellobiose phosphorylase
(Cbp)
(α/α)6Dimer[202]
GH94BAA28631.1Cellulomonas gilvus ATCC131272CQS[A,B], 2CQT[A,B], 3ACS[A,B], 3ACT[A,B], 3AFJ[A,B], 3QFY[A,B], 3QFZ[A,B], 3QG0[A,B]Cellobiose phosphorylase
(Cbp; CgCBP)
(α/α)6Dimer[203,204]
GH94AAQ20920.1Cellulomonas uda DSM 201083RRS[A,B], 3RSY[A,B], 3S4A[A,B], 3S4B[A,B], 3S4C[A], 3S4D[A]Cellobiose phosphorylase
(CbP)
(α/α)6DimerUnpublished
GH94ABX41081.1 Lachnoclostridium phytofermentans ISDg5H3Z[A,B], 5H40[A,B], 5H41[A,B], 5H42[A,B]β-1,2-oligoglucan phosphorylase
(LpSOGP; Cphy_0694)
(α/α)6Monomer[88]
GH94BAJ10826.1Paenibacillus sp. YM16GGY[A,B], 6GH2[A,B], 6GH3[A,B]Laminaribiose phosphorylase
(LbpA)
(α/α)6Dimer[116]
GH94ABD80168.1Saccharophagus degradans 2-404ZLE[A], 4ZLF[A], 4ZLG[A], 4ZLI[A]Cellobionic acid phosphorylase 94B
(Cep94B; CBAP; SdCBAP; Sde_0906)
(α/α)6Dimer[91]
GH94BAC87867.1Vibrio proteolyticus1V7V[A], 1V7W[A], 1V7X[A]Chitobiose phosphorylase
(ChbP; VpChBP)
(α/α)6Dimer[205]
GH112BAD80751.1Bifidobacterium longum subsp. longum JCM 12172ZUS[A,B,C,D], 2ZUT[A,B,C,D], 2ZUU[A,B,C,D], 2ZUV[A,B], 2ZUW[A,B,C,D], 3WFZ[A,B,C,D]Galacto-N-biose phosphorylase/Lacto-N-biose phosphorylase
(LnpA1; LnbP; GLNBP; BLLJ_1623)
(β/α)8Dimer[206,207]
GH130CAH06518.1 Bacteroides fragilis NCTC 93433WAS[A,B], 3WAT[A,B], 3WAU[A,B], 4KMI[A,Bβ-1,4-mannosylglucose phosphorylase
(MGP; BF0772)
5-fold β-propellerHexamer[208]
GH130CAC96089.1Listeria innocua Clip112625B0P[A,B], 5B0Q[A,B], 5B0R[A,B], 5B0S[A,B]β-1,2-mannobiose phosphorylase
(Lin0857)
5-fold β-propellerDimer[100]
GH130ADU21379.1Ruminococcus albus 7 (DSM 20455)5AY9[A], 5AYC[A]β-1,4-mannosylglucose phosphorylase
(RaMP1; RaMGP; Rumal_0852)
5-fold β-propellerTrimer[209]
GH130ADU20661.1Ruminococcus albus 7 (DSM 20455)5AYD[A,B,C,D,E,F], 5AYE[A,B,C,D,E,F]β-1,4-mannooligosaccharide phosphorylase
(MOP; RaMP2; Rumal_0099)
5-fold β-propellerHexamer[209]
GH130ABY93074.1Thermoanaerobacter sp. X5147FIP, 7FIQ, 7FIR, 7FISβ-1,2-mannobiose phosphorylase
(Teth514_1789)
5-fold β-propellerMonomer[210]
GH130AAD36300.1Thermotoga maritima MSB81VKD[A,B,C,D,E,F]β-1,4-mannooligosaccharide phosphorylase (TM1225)5-fold β-propellerDimerUnpublished
GH130ADD61463.1Uncultured bacterium4UDG[A,B,C,D,E,F], 4UDI[A,B,C,D,E,F], 4UDJ[A,B,C,D,E,F], 4UDK[A,B,C,D,E,F]β-1,4-mannopyranosyl-[N-glycan] phosphorylase/β-1,4-mannopyranosyl-chitobiose phosphorylase
(UhgbMP; Uhgb_MP)
5-fold β-propellerHexamer[195]
GH149/Uncultured bacterium6HQ6[B], 6HQ8[A,B], 6HQ6[A]β-1,3-oligoglucan phosphorylase
(Pro_7066)
(α/α)6Dimer[114]
GT35AAM52219.1Corynebacterium callunae2C4M[A,B,C,D]Glycogen phosphorylaseGT-BDimerUnpublished
GT35AAC76442.1Escherichia coli str. K-12 substr. MG16551AHP[A], 1E4O[A], 1L5V[A], 1L5W[A], 1L6I[A], 1QM5[A], 2ASV[A,B], 2AV6[A,B], 2AW3[A,B], 2AZD[A,B], 2ECP[AMaltodextrin phosphorylase
(MalP; b3417)
GT-BDimer[136,211,212,213]
GT35AAN59210.1Streptococcus mutans UA1594L22[A]Maltodextrin phosphorylase
(GlgP; SMU.1564)
GT-BMonomerUnpublished
GT35CAB61943.1Arabidopsis thaliana4BQE[A,B], 4BQF[A,B], 4BQI[A,B]α-glucan phosphorylase
(PHS2; AtPHS2; At3g46970)
GT-BDimer[127]
GT35AAB60395.1Homo sapiens5IKO[A], 5IKP[A]Glycogen phosphorylase (brain)
(bGP)
GT-BDimer[214]
GT35CAA75517.1Homo sapiens1EM6[A,B], 1EXV[A,B], 1FA9[A], 1FC0[A,B], 1L5Q[A,B], 1L5R[A,B], 1L5S[A,B], 1L7X[A,B], 1XOI[A,B], 2ATI[A,B], 2QLL[A], 2ZB2[A,B], 3CEH[A,B], 3CEJ[A,B], 3CEM[A,B], 3DD1[A,B], 3DDS[A,B], 3DDW[A,B]Glycogen phosphorylase (liver)GT-BDimer[215,216,217,218,219,220,221,222,223]
GT35AAC17451.1Homo sapiens1Z8D[A]Glycogen phosphorylase (muscle)
(PigM)
GT-BDimer[224]
GT35BAK00834.1Hordeum vulgare subsp. Vulgare5LR8[A,B], 5LRA[A,B]Plastidial α-1,4-glucan phosphorylase
(Pho1; HvPho1)
GT-BDimer[153]
GT35ACJ76617.1Oryctolagus cuniculusDozens available-2GJ4[A] best solutionGlycogen phosphorylase (muscle)
(PygM)
GT-BDimer[225]
GT35AAB68057.1Saccharomyces cerevisiae S288c1YGP[A,B]Glycogen phosphorylase
(Gph1; YPR160w)
GT-BDimer[226]
GT108CBZ24449.1Leishmania mexicana MHOM/GT/2001/U11036Q50[A]MTP4
(LMXM_10_1260)
5-fold β-propellerMonomer[174]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, A.; Benkoulouche, M.; Ladeveze, S.; Durand, J.; Cioci, G.; Laville, E.; Potocki-Veronese, G. Discovery and Biotechnological Exploitation of Glycoside-Phosphorylases. Int. J. Mol. Sci. 2022, 23, 3043. https://doi.org/10.3390/ijms23063043

AMA Style

Li A, Benkoulouche M, Ladeveze S, Durand J, Cioci G, Laville E, Potocki-Veronese G. Discovery and Biotechnological Exploitation of Glycoside-Phosphorylases. International Journal of Molecular Sciences. 2022; 23(6):3043. https://doi.org/10.3390/ijms23063043

Chicago/Turabian Style

Li, Ao, Mounir Benkoulouche, Simon Ladeveze, Julien Durand, Gianluca Cioci, Elisabeth Laville, and Gabrielle Potocki-Veronese. 2022. "Discovery and Biotechnological Exploitation of Glycoside-Phosphorylases" International Journal of Molecular Sciences 23, no. 6: 3043. https://doi.org/10.3390/ijms23063043

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