You are on page 1of 16

pubs.acs.

org/acschemicalbiology Reviews

Configured for the Human Gut Microbiota: Molecular Mechanisms


of Dietary β‑Glucan Utilization
Benedikt Golisch,⊥ Zhenhuan Lei,⊥ Kazune Tamura, and Harry Brumer*

Cite This: ACS Chem. Biol. 2021, 16, 2087−2102 Read Online

ACCESS Metrics & More Article Recommendations


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via LNCC on October 9, 2023 at 11:04:41 (UTC).

ABSTRACT: The β-glucans are a disparate group of structurally diverse polysaccharides, whose members are widespread in human
diets as components of the cell walls of plants, algae, and fungi (including yeasts), and as bacterial exopolysaccharides. Individual β-
glucans from these sources have long been associated with positive effects on human health through metabolic and immunological
effects. Remarkably, the β-configured glucosidic linkages that define these polysaccharides render them inaccessible to the limited
repertoire of digestive enzymes encoded by the human genome. As a result, the various β-glucans become fodder for the human gut
microbiota (HGM) in the lower gastrointestinal tract, where they influence community composition and metabolic output, including
fermentation to short chain fatty acids (SCFAs). Only recently, however, have the specific molecular systems that enable the
utilization of β-glucans by select members of the HGM been fully elucidated by combined genetic, biochemical, and structural
biological approaches. In the context of β-glucan structures and their effects on human nutrition and health, we summarize here the
functional characterization of individual polysaccharide utilization loci (PULs) responsible for the saccharification of mixed-linkage
β(1→3)/β(1→4)-glucans, β(1→6)-glucans, β(1→3)-glucans, β(1→2)-glucans, and xyloglucans in symbiotic human gut bacteria.
These exemplar PULs serve as well-defined biomarkers for the prediction of β-glucan metabolic capability in individual bacterial taxa
and across the global human population.

■ INTRODUCTION
Glucose is the fundamental monosaccharide building block from
humans (and other mammals), because our genomes do not
encode the required β-glucanases. What a difference a bond
makes!
photosynthesis that is polymerized into plant energy storage and
In fact, the β-glucans comprise a disparate group of
cell wall polysaccharides via two possible anomeric config-
polysaccharides with varied backbone linkages and side chain
urations, alpha and beta.1,2 Of these, α-glucans (i.e., starch,
composition. The inherent recalcitrance of β-glucans to human
comprising amylose and amylopectin) in common fruits and
digestive enzymes makes them fodder for the human gut
vegetables, especially tubers and cereal grains, are readily
microbiota (HGM), particularly in the lower gastrointestinal
depolymerized by human digestive enzymes after cooking and
tract where they are deconstructed and fermented along with
thus constitute a major, rapidly accessible energy source in diets
other components of “dietary fiber”.7−9 We are thus holobionts,10
worldwide.3 On the other hand, our genomes do not equip us
with the enzymes necessary to digest the diverse β-glucans,
which are likewise common in our diets as structural Special Issue: Chemical Glycobiology
components of plant and fungal (including yeast) cell walls.4,5 Received: July 20, 2021
This dichotomy is best exemplified by the linear homopolymers Accepted: October 5, 2021
amylose (α(1→4)-glucan) and cellulose (β(1→4)-glucan, Published: October 28, 2021
Figure 1): Amylose is digestible by human amylases and α-
glucosidases,6 whereas cellulose is inherently undigestible by

© 2021 American Chemical Society https://doi.org/10.1021/acschembio.1c00563


2087 ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

Figure 1. Representative structures of common dietary α- and β-glucans. Monosaccharide residues are represented using the Consortium for
Functional Glycomics symbol nomenclature,34 as summarized in the key.

who are critically dependent on our resident microbes for the appreciated by microbiologists and nutritionists.8,11−20 On the
metabolism of β-glucans and other nonstarch polysaccharides other, recent years have witnessed the elucidation of highly
(NSPs). One on hand, this relationship has long been specialized molecular systems dedicated to the recognition,
2088 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

hydrolysis, and uptake of β-glucans by individual members of the humans,41 lichenan finds widespread use as a commercially
HGM, which underpins the role of these polysaccharides in available substrate for glycosidase enzymology.
nutrition and health.21 β(1→3)-Glucans. Polysaccharides comprising exclusively

■ A PRIMER ON DIETARY β-GLUCAN STRUCTURES


The β-glucans are, by definition, polysaccharides comprising a
β(1→3)-linked poly(glucosyl) backbones are widely distributed
in nature in both linear and branched forms. Bacteria, yeast,
microalgae, and terrestrial plants are known to produce linear
backbone of D-glucopyranosyl residues linked via β-glycosidic β(1→3)-glucans (Figure 1), which fulfill different biological
bonds.22,23 β-glucans are widespread in nature, from the roles and are historically referred to by source-specific names, for
extracellular polysaccharides of bacteria24 to the cell walls of example, curdlan (Agrobacterium sp.), zymosan (Saccharomyces
fungi,25,26 algae and seaweed,27 and terrestrial plants,2 all of cerevisiae), paramylon (Euglena gracilis), and callose (Viridi-
which may constitute dietary sources for humans. The inherent plantae).42−45 Of these, curdlan is an approved food gelling
diversity of potential glycosidic linkages, branching patterns, and agent (E424),46 while callose is produced in limited amounts by
degrees of polymerization28 enables a wide variation in β-glucan plants, including food crops, during particular developmental
structures among sources. In addition, environmental factors stages and in response to wounding or pathogen attack.2,45
during biosynthesis,25 as well as extraction and purification Branched β(1→3)-glucans are widespread in fungi and algae,
methods, can influence β-glucan molar mass and substitution in which they are key components of the cell wall. Both between
patterns.29 In turn, these molecular properties, as well as the and within these taxa, β(1→3)-glucans are distinguished by the
organization of β-glucans into higher-order composite struc- frequency and structures of the pendant side chains. In general,
tures in cell walls, will influence their accessibility to the HGM.30 β-glucans from fungi have comb-like structures comprising a
The structural intricacies and diverse sources of β-glucans β(1→3)-linked backbone that is branched by β-(1→6)-linked
have been revealed through a rich history of study.29,31−33 glucosyl side chains.47,48
Despite this, the exact nature of these polysaccharides in specific Pachyman, from the fungus Poria cocos, represents perhaps the
contexts is often obfuscated by the use of the generic term “β- most minimalist example, with a branching ratio of single β(1→
glucan” in scientific articles, popular media, and product labels. 6)-glucosyl residues of ca. 1:60.49,50 Other well-known examples
In the following sections, we provide a primer on representative of branched β(1→3)-glucans include lentinan from Lentinula
β-glucan structures of particular relevance to the HGM. edodes (shiitake mushroom), grifolan from Grifola f rondosa
β(1→4)-Glucans. Cellulose (β(1→4)-glucan, Figure 1) is (maitake), pleuran from Pleurotus ostreatus (oyster mushroom),
the primary load-bearing component of all terrestrial plant cell schizophyllan from Schizophyllum commune (split gill), and
walls, as well as many algae, and is therefore abundant in a well- scleroglucan from Sclerotium species.25,42,51,52 The first four of
balanced diet. Additionally, some bacteria, especially Komaga- these originate in fungi that are consumed by humans in
taeibacter xylinus (previously Acetobacter xylinum, Gluconaceto- different regions of the world and are likely to be generally
bacter xylinus), produce cellulose as an extracellular poly- representative of β-glucans in other edible fungi.53 All five
saccharide,35 which forms the basis of the gelatinous dessert nata contain single β(1→6)-Glc branches and are distinguished by
de coco. The linear, unbranched structure of cellulose enables small differences in the degree of branching,51 typically one
ordered packing of cellulose chains into semicrystalline branch for every three to four backbone residues.54−58 Yeasts
structures,2 which are essentially undigestible by the HGM. contain more elaborate branched β(1→3)-glucans in their cell
On the other hand, the incorporation of branching along the walls. In Saccharomyces cerevisiae, the most widely recognized
β(1→4)-glucan chain by some organisms disfavors ordered β(1→3)-glucan has long β(1→3)-Glc side chains extended
packing, yielding soluble polysaccharides. For example, xanthan from β(1→6) branch points59 (Figure 1). An additional β(1→
(Figure 1), a complex extracellular polysaccharide produced by 3)-glucan component has side chains of at least four β(1→6)-
the bacterium Xanthomonas campestris, is a widely used food Glc residues.60,61
rheology modifier (E415).36 Likewise, the xyloglucans (XyGs) β(1→3)-glucans from brown algae are known as laminarins,
comprise a family of ubiquitous fruit, vegetable, and grain cell the β(1→6)-branched structures of which are species dependent
wall matrix polysaccharides, the amorphous structures of which (Figure 1). Laminarins from the genus Laminaria invariably
result from frequent and extensive branching of the β(1→4)- have short branches,33 as exemplified by L. digitata laminarin,
glucan backbone via α(1→6)-xylosyl and other residues (Figure which contains single β(1→6)-glucosyl branches in a ratio of 1:7
1).2,8,37 versus backbone residues.33,62 In contrast, the laminarin from
β(1→3)/β(1→4)-Glucans (Mixed-Linkage β-Glucans). A Eisenia bicyclis exhibits a higher molecular weight and a 2:3 ratio
further way in which nature avoids the self-assembly of β(1→4)- of β(1→6) to β(1→3) linkages.62 Notably, the side chains of
glucans into insoluble, self-assembled structures is through the Eisenia bicyclis laminarin comprise up to three β(1→6)-glucosyl
introduction of alternate linkages in the backbone. Thus, residues per branch.42 Other unique β(1→3)-glucans include
“mixed-linkage β-glucans (MLGs)”, viz. β(1→3)/β(1→4)- the capsular polysaccharide of the bacterium Streptococcus
glucans, have a kinked structure composed of randomly pneumoniae Type 37, which contains Glcβ(1→2)−Glcβ(1→6)
distributed β(1→4)-linked cellotriosyl (Glc3) and cellotetraosyl side chains,63 and the cyclic, branched mixed-linkage β(1→3)/
(Glc4) units that are linked via β(1→3)-glycosidic bonds. MLGs β(1→6)-glucan from Bradyrhizobium japonicum.64
are found particularly in the starchy endosperm and aleurone β(1→6)-Glucans. Linear β(1→6)-glucans are exemplified
layer of common dietary cereals, in which the composition of by pustulan from the lichen Lasallia pustulata, which serves as a
β(1→3)- and β(1→4)-glycosidic bonds and random block commercial source of this polysaccharide (Figure 1). Linear
structure differs between species. The following Glc3/Glc4 ratios β(1→6)-glucan has also been isolated from the lichen
are observed: wheat, 3.0−4.5; barley, 1.8−3.5; rye, 1.9−3.0; and Umbilicaria caroliniana40 and the fungus Guignardia citricarpa.65
oat, 1.5−2.3.38,39 Lichens produce the eponymous MLG Dietary yeasts such as Saccharomyces cerevisiae also produce a
lichenan, which is predominantly composed of Glc3 units.40 linear β(1→6)-glucan, which is cross-linked with other cell wall
Although this MLG is more relevant to the diet of reindeer than components.59,66 Other fungi also contain branched β(1→6)-
2089 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

glucans. For example, Agaricus blazei (almond mushroom) and Actinobacteria (e.g. Bif idobacterium), Proteobacteria (e.g.,
an edible hybrid of Pleurotus f lorida and Lentinula edodes contain Escherichia), Fusobacteria (Fusobacterium), and Verrucomicro-
β(1→6)-glucans with single β(1→3)-linked glucosyl branches bia (Akkermansia).88 Due to their indigestibility by human
every three to four backbone residues.67,68 digestive enzymes, the β-glucans comprise “prebiotics” that can
β(1→2)-Glucans. Although β(1→4)-, β(1→3)-, and β(1→ modulate the growth and activity of beneficial gut bacteria.11
6)-glucans are widely distributed in nature and found in the Thus, there is a long history of fundamental microbiological
human diet, β(1→2)-glucans are rare. Thus far, only cyclic studies to identify key β-glucan-utilizing taxa, which predates
β(1→2)-glucans comprising ca. 20 residues are known from more recent innovations in human gut (meta)genomics.89−93
bacteria, where they function as osmolytes and virulence β(1→4)-Glucans. Cellulose breakdown in the human gut
factors69−71 (Figure 1). has been studied since the 19th century.13 In the early 20th

■ THE β-GLUCAN/HGM NEXUS


The United Nations Food and Agriculture Organization (FAO)
century, the disappearance of over 50% of cellulose extracted
from carrots and cabbage in the stool of subjects implicated
colonic microorganisms in cellulose degradation, although
estimates that over 871 million tonnes of fruits, 1.1 billion notably, purified cellulose remained indigestible.94,95 Subse-
tonnes of vegetables, and 2.9 billion tonnes of cereals were quently in the late 20th century, individual cellulolytic bacterial
produced worldwide in 2018.72 These agricultural products are strains were successfully isolated from human feces, including
indispensable parts of a well-balanced diet and are thus regular Ruminococcus, Clostridium, Eubacterium, and Enterococcus
sources of β-glucans. Dicotyledonous plants, which include most species.14,96−99 More recently, cellulolytic Bacteroides species
fruits and vegetables, are abundant in cellulose and xyloglucans, have been identified from the human gut.100 The effects of
the latter of which can comprise up to one-fourth of the dry amorphous plant cell wall matrix glycans, for example,
weight of the cell wall.2,8 Monocotyledonous plants of the xyloglucan, in facilitating cellulose degradation by human fecal
taxonomic Family Poaceae, which includes cereal crops, are samples has also been recently demonstrated.101
excellent sources of MLGs,2 which comprise nearly three- Considering the branched dietary β(1→4)-glucans, the ability
quarters of the soluble, nonstarch glycan content of the of Bacteroides ovatus, Bif idobacterium infantis, and several
endosperm.8,73 Although they are not a staple per se, the fungi Clostridium species, as well as crude human fecal slurries, to
(e.g., edible mushrooms and baker’s yeast) are regular dietary degrade xyloglucan polysaccharide has been long known.102
sources of β(1→3)- and β(1→6)-glucans, which constitute up More recently, the ability of several strains of Prevotella copri
to 85% of cell walls.33 Not least, bacterial β-glucans, such as (phylum Bacteroidetes) and Ruminococcus bicirculans (phylum
curdlan (E424) and xanthan (E415), are widely used in food Firmicutes) to grow on xyloglucan has been demonstra-
applications due to their physical−chemical effects on ted.103,104 A wider range of Bif idobacterium species are able to
thickening, whippability, emulsification, and gelation. 29 ferment xyloglucan oligosaccharides,102,105 which might be
Although it is difficult to estimate the total mass of β-glucans liberated by primary degraders, such as the Bacteroidetes. It has
consumed worldwide due to widely varying diets across cultures been recently shown that xyloglucans exert a significant impact
and regions, national health agencies generally recommend a on human fecal community structures in vitro, favoring the
daily intake by adults of 20−40 g of total dietary fiber. Moreover, expansion of Bacteroides,106,107 which are known to possess
daily consumption of 1−4 g of cereal (e.g., oat or barley) MLGs specific xyloglucan utilization loci (vide infra).108 Apart from
is associated with reduced circulating cholesterol levels and their use as a food gelling agent, xyloglucans have been explored
mitigation of rapid increases in postprandial blood glucose to resolve intestinal maladies.109,110 Although little is presently
levels.74,75 known about the interaction of xanthan with the HGM, exciting
The diverse β-glucans in our diet resist digestion in the upper new work has revealed a keystone Ruminococcaceae and a cross-
gastrointestinal tract and hence transit to the large intestine, feeding Bacteroides that can utilize xanthan and its hydrolysis
where they can be partially or completely metabolized by the products, respectively.111
HGM.76 The fermentation of complex carbohydrates such as β- β(1→3)/β(1→4)-Glucans. β-glucans from cereals, that is,
glucans by the HGM directly affects our nutrition and health by MLGs, have been widely advertised as health-promoting by the
generating short-chain fatty acids (SCFAs) that enter the food industry, due to cholesterol lowering properties112,113 and
circulatory system and contribute up to 10% of our daily caloric beneficial effects on postprandial glucose levels and insulin
intake.18−20 Butyrate, in particular, is the primary energy source response.114 The cholesterol-lowering effect is caused primarily
of the colonic endothelium and is critical to maintaining a by the viscosity of β-glucan and the entrapment of bile acids in
healthy gut barrier, including the prevention of cancer.12,17 β- the upper gastrointestinal tract. The resulting reduction triggers
glucans and their partial hydrolysis products may themselves an increase in bile acid biosynthesis, which consumes
mediate physiological effects,77 including immune system cholesterol.115,116 The HGM also contributes to the cholester-
stimulation, through direct interaction with cell-surface ol-lowering effect of MLGs through the production of
receptors.78−81 Furthermore, dietary β-glucans can modulate propionate; an increased propionate-to-acetate ratio in the
the composition and metabolism of the HGM to quench peripheral serum reduces cholesterol biosynthesis.117
reactive oxygen species (ROS) and promote carcinogen The viscosity of MLGs also modulates postprandial glucose
egestion.47 levels by enclosing the food bolus in a thick layer, with
The HGM is a complex and dynamic ecosystem comprising corresponding implications for hyperglycemia and type 2
trillions of cells and diverse taxa, the composition of which diabetes.15,118,119 On one hand, this bulk physical phenomenon
fluctuates depending upon daily dietary composition and other results in less mixing between food and enzymes in the
factors.82−87 The HGM is predominantly comprised of bacteria stomach.120 On the other, it slows down diffusion and
from the phyla Bacteroidetes (e.g., Bacteroides, Prevotella) and adsorption of glucose in the intestine, leading to a more
Firmicutes (e.g., Clostridium, Roseburia, Lactobacillus), which constant uptake. These phenomena extend the feeling of
constitute ca. 90% of taxa, together with lesser amounts of satiety.121 In addition, SCFA production by the HGM
2090 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

Figure 2. Glucan-specific polysaccharide utilization loci. (a) Genetic arrangement of exemplar Bacteroides β-glucan utilization loci (β-GULs) with
reference to the archetypal starch (α-glucan) utilization system (Sus). Boundary gene locus tags are indicated under each PUL. Proteins with
homologous functions are colored identically: blue, glycoside hydrolase (GH); green, cell-surface glycan-binding protein (SGBP); orange, TonB-
dependent transporter (TBDT); purple, sensor-regulator (SusR-like or hybrid-two component system (HTCS)). (b) Cellular organization of β-GUL-
encoded proteins that hydrolyze β-glucan homopolysaccharides, colored as in panel a.

2091 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

Figure 3. Cellular organization and structural biology of the XyGUL-encoded proteins that saccharify the heteropolysaccharide xyloglucan. The
XyGUL is shown in Figure 2a. GH and SGBP structures are represented by the following PDB IDs: BoGH5 (PDB ID 3ZMR108), BoGH9 (PDB ID
6DHT171), BoGH31 (PDB IDs 5JOU and 5JOV172), BoGH3B (PDB ID 5JP0172), BoGH43A (PDB IDs 5JOW, 5JOX, and 5JOY172), and BoGH43B
(PDB ID 5JOZ172).

modulates postprandial glucose levels by increasing the determined by the presence of specific polysaccharide utilization
production of peptide YY in intestinal cells, which is involved loci in genomes (vide infra).131−133
in the regulation of appetite and insulin mediated glucose uptake β(1→3)-Glucans. Whereas cereal MLGs are generally
in muscle and adipose tissue.122 However, dietary MLG only associated with metabolic health,119,134 fungal β(1→3)-glucans
shows a small impact on long-term effects such as body are more commonly implicated in effects on the immune
weight.123 system.51,135 Research in this area can be traced back to the
Motivated by the health-promoting effects of β(1→3)/β(1→ beginning of the 20th century, when the immunomodulating
4)-glucans, the fermentation of cereal MLGs by endogenous function of yeast cells was discovered and later linked to the
members of the HGM has received significant attention since crude cell wall preparation zymosan, which contains β(1→3)-
the late 1900s. For example, in vitro analyses revealed that glucans.26,136 Over time, it became apparent that branched
selected Bacteroidetes (Bacteroides and Prevotella) and Firmi- β(1→3)[β(1→6)]-glucans also play important roles in immune
cutes (Ruminococcus and Coprococcus) isolates grew on barley responses directly. In particular, M cells of Peyer’s patches in the
MLG (polysaccharide), whereas Lactobacillus, Bif idobacterium, small intestine can mediate the transport of fungal β-glucans
or Enterococcus species or E. coli strains did not.103,104,124,125 On through the epithelium,47 followed by binding to different
the other hand, there is evidence to suggest that Lactobacillus receptors on phagocytic and cytotoxic innate immune cells,
and Bif idobacterium species can utilize MLG hydrolysis which in turn triggers signaling pathways leading to immune
products,126 which likely accounts for growth observed in responses.135 Prominent receptors such as dectin-1 and
mixed cultures127 and changes in the ratio of Bacteroidetes to complement receptor 3 (CR3) are able to bind β(1→
Firmicutes.128,129 Such effects may arise from “sharing” of 3)[β(1→6)]-glucans and thereby enhance the release of
“public goods” in the complex ecosystem of the HGM.130 proinflammatory cytokines.79,137 Furthermore, β(1→3)[β(1→
Surveys of individual Bacteroides species have indicated that 6)]-glucans are able to induce the expression of receptors such as
growth on MLG is not ubiquitous across the genus but is Toll-like receptor 2 and can also shift the balance between T
2092 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

helper 1 (Th1) cells and T helper 2 (Th2) cells toward Th1, number of which scales with the linkage complexity of the
which is related to allergic responses.47,138 substrate154,155 (Figure 2, Figure 3). The general steps of PUL-
Fungal β(1→3)[β(1→6)]-glucans, including lentinan, grifo- mediated glycan saccharification are as follows:
lan, scleroglucan, and schizophyllan, have also received attention (1) substrate binding by one or more SGBPs at the bacterial
as potential antitumor compounds.17 Here, fungal β-glucans are cell surface,
first internalized by macrophages and transported to the bone
marrow where they are degraded into oligosaccharides, which (2) polysaccharide backbone cleavage by cell-surface anch-
bind to the CR3 receptor of neutrophils. The resulting activated ored endo-acting enzymes,
granulocytes are able to kill iC3b-coated tumor cells.135,139,140 (3) oligosaccharide import into the periplasm through the
Thus, fungal β(1→3)[β(1→6)]-glucans play an important role TBDT complex,
in the regulation of lymphocyte, neutrophil, and natural killer (4) ultimate conversion to monosaccharides (sometimes
cells in the innate immune system.139 disaccharides) for transport into the cytosol, through
The ability of bacterial species in the HGM to utilize β(1→3)- the concerted action of one or more exo-acting enzymes.
glucans was defined through pioneering work by Abigail Salyers The following sections will describe notable molecular
and co-workers in the late 1970s.141 Nine of 11 Bacteroides features of PULs responsible for the utilization of β(1→3)/
species, represented by ca. 200 strains in total, fermented β(1→4)-glucans, β(1→6)-glucans, β(1→3)-glucans, β(1→2)-
laminarin (and many terrestrial polysaccharides).142 In contrast, glucans, and xyloglucan in the context of this paradigm. We will
a survey of 154 strains from 22 species of human colonic specifically highlight the crucial role that the combined use of
bacteria, including Bif idobacterium, Peptostreptococcus, Lactoba- specific substrates and inhibitors, analytical biochemistry, and
cillus, Ruminococcus, Coprococcus, Eubacterium, and Fusobacte- structural biology has played in detailed SGBP and GH
rium species, revealed no laminarin fermenters, except for a structure−function studies.
single Peptostreptococcus productus strain.142 More recently, the Mixed-Linkage Glucan Utilization Loci. The mixed-
ability of Lactobacillus and Bif idobacterium species to utilize linkage β-glucan utilization locus (MLGUL) of B. ovatus (locus
laminarin for growth has been demonstrated.143 Notably, tags BACOVA_02740−02745) encodes, in order, a hybrid two-
growth of Bacteroides species on laminarin was found to induce component sensor (HTCS), a glycoside hydrolase family 16
specifically β(1→3)-glucanase activity,144 portending the (GH16) mixed-linkage endo-β-glucanase (MLGase), a TBDT/
discovery of β(1→3)-glucan-specific polysaccharide utilization SGBP (SusC/SusD-homologue) pair, an additional sequence-
loci in select Bacteroides and other Bacteroidetes (vide infra).133 divergent SGBP, and a glycoside hydrolase family 3 (GH3) β-
These results collectively underpin the interest in using glucosidase (Figure 2a).132,156 Initial reverse genetics (gene
laminarin as a prebiotic.145


deletion) experiments established that this locus was exclusively
responsible for the growth of B. ovatus on MLGs. Subsequent
MOLECULAR SYSTEMS FOR β-GLUCAN biochemical analysis revealed that the vanguard BoGH16MLG
UTILIZATION (PDB ID 5NBO) hydrolyzes MLGs with much higher specificity
Despite the long-standing appreciation that the HGM is able to than the β(1→3)-glucans curdlan, laminarin, or yeast β-glucan.
saccharify β-glucans to glucose (and in the case of xyloglucan, Product analysis demonstrated that this endo-acting enzyme
other monosaccharides; Figure 1) to feed primary metabolism, hydrolyzes MLGs into only two distinct oligosaccharides,
detailed molecular insight into this process has only recently G4G3G and G4G4G3G (where “G” represents a β-glucosyl
been achieved through combined microbiological, biochemical, residue and the number indicates the linkage regiochemistry),
and structural biological studies on β-glucan-specific poly- which reflects the specific hydrolysis of β(1→4) linkages
saccharide utilization loci (PULs)146 in exemplar Bacteroides immediately preceding β(1→3)-linked glucosyl residues.
species. In this context, it is worth noting that Bacteroidetes are Further substrate analysis with chromogenic oligosaccharides
particularly prodigious utilizers of complex glycans, with showed that a β(1→3) linkage spanning enzyme subsites −2
numbers of glycoside hydrolases (GHs) and polysaccharide and −1 is necessary for catalysis, whereas β(1→4)-Glc binding
lyases (PLs) collectively numbering in the hundreds in some in the −3 and −4 subsites contributes to substrate recognition
species, which far exceeds other bacteria in the HGM.4 Despite (see ref 157 for an explanation of sugar-binding subsites). These
the organization of many of these enzymes into families in the kinetic data were rationalized by the tertiary structures of
carbohydrate-active enzymes (CAZy) classification, sequence BoGH16MLG in complex with the product G4G4G3G and
analysis alone is often insufficient to confidently assign structurally related covalent inhibitors (PDB IDs 5NBP, 6VHO,
molecular function at the level of substrate and product 7KR6), which illustrated the existence of three well-defined
specificity.147,148 Hence, detailed structure−function character- negative subsites and a fourth weakly interacting negative subsite
ization is necessitated. in the active-site cleft.132,158 BoGH16MLG possessed a β-jellyroll
Canonical PULs are multigene clusters found in Bacteroidetes fold characteristic of the family147 and was outer-membrane-
that encode molecular systems of carbohydrate-binding, anchored by N-terminal lipidation.
-cleaving, and -transporting proteins, the expression of which Biochemical analysis showed that the exo-β-glucosidase
is under the control of a glycan-specific sensor or regulator that GH3MLG can completely degrade the BoGH16MLG-derived
recognizes intermediate breakdown products.131,146,149−151 The oligosaccharides, G4G4G3G and G4G3G, into glucose by
archetypal PUL encodes the starch (α-glucan) utilization system sequential hydrolysis from the nonreducing end, after import
elucidated by Abigail Salyers and colleagues in the late 1980s and into the periplasm (Figure 2b). Initial-rate kinetics using series
early 1990s152 (comprising susABCDEFG/susR, locus tags of gluco-oligosaccharides with different linkages and degrees of
BT3698−3705, Figure 2a). Hallmarks of complete PULs are a polymerization demonstrated the preference of BoGH3MLG for
core pair of SusC/SusD homologues constituting a TonB- β(1→3)- and β(1→4)-linkages over β(1→6). Furthermore,
dependent transporter (TBDT) and cell-surface glycan-binding BoGH3MLG hydrolyzes oligosaccharides longer than trisacchar-
protein (SGBP) complex153 and a cohort of GHs and PLs, the ides with similar catalytic efficiencies due to the presence of one
2093 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

negative subsite and two positive subsites that contribute to immediately downstream of the SusD-like SGBP, as is common
catalysis. In the absence of an experimental tertiary structure, among most PULs. Second, the SusR-like regulator, which is
homology modeling indicated that the (β/α)8-barrel (triose strictly conserved in syntenic β1,6-GULs,66 is comparatively rare
phosphate isomerase (TIM)-barrel) and the α/β sandwich versus HTCS and ECF regulators in Bacteroides PULs; less than
domain define a narrow active-site cleft typical of GH3 10% of Bacteroidetes regulators are SusR-like.151,161
members.159 Commensurate with a role in initiating substrate capture, both
Detailed analysis of substrate recognition by the SGBP-A the SusD-like SGBP-A and the additional SGBP-B selectively
(SusD homologue) and SGBP-B, using pull-down assays, bind to β(1→6)-glucan and show no affinity toward β(1→3)-
affinity gel electrophoresis, and isothermal titration calorimetry, nor β(1→4)-glucans by affinity gel electrophoresis analysis.
revealed particular affinities of both proteins for mixed-linkage Isothermal titration calorimetry further showed that the SGBP-B
glucans over all β(1→4)-glucans (e.g., cellulose, xyloglucan). targets the polysaccharide with higher affinity than the SusD-like
Commensurate with roles in recognition of MLG polysacchar- SGBP-A and that both proteins are likely to present at least six
ide chains at the cell surface, both SGBPs bound longer glycosyl-binding subsites.66 Thus, in the canonical model of
oligosaccharides, such as the hexasaccharide G4G3G4G4G3G PUL-encoded systems (Figure 2b), SGBP-B likely provides
and the heptasaccharide G4G4G3G4G4G3G but not the limit- general affinity of the bacterium to the substrate, whereas SGBP-
digest products of BoGH16MLG, G4G3G and G4G4G3G.156 A assists in feeding β(1→6)-glucan fragments generated by the
The SGBP-A (PDB IDs 6E60, 6DMF, 6E61) has a tertiary GH30_3 endo-glucanase into the TBDT through a “pedal-bin”
structure based upon four tetratricopeptide repeat (TPR) that is action.153
typical of SusD homologues, whereas the SGBP-B (PDB IDs Indeed, BtGH30_3 (PDB IDs 5NGK, 5NGL66) was revealed
6E57, 6E9B) has an extended, multimodular domain structure by kinetic analyses to be a highly specific endo-β(1→6)-
that presents the binding site distal to the membrane-anchored glucanase with activity on pustulan, but not on β(1→3)-glucan,
N-terminus. Reverse genetics demonstrated that SGBP-A has an β(1→4)-glucan, nor β(1→6)-galactan. The BtGH30_3-cata-
irreplaceable function for the growth of B. ovatus on MLGs and lyzed hydrolysis of pustulan generated a distribution of
therefore plays a key role in carbohydrate uptake by the TBDT. oligosaccharides of different chain lengths, typical of endo-acting
In contrast, the SGBP-B was dispensable for growth but GHs and consistent with the observed release of these
increased the lag phase, suggesting an important complementary oligosaccharides from the cell surface under nonlimiting growth
role in MLG recognition.156 conditions. Further kinetic and product analyses using a range of
Systems-based approaches such as those described above, β(1→6)-gluco-oligosaccharides (Glc2 to Glc8) suggested that
which incorporate microbiological, biochemical, and structural BtGH30_3 requires three subsites, from −2 to +1, to be
data, underpin definition of the MLGUL as a reference to occupied for catalytic activity; gentibiose (Glcβ(1→6)-Glc) was
identify homologous PULs in other Bacteroidetes, for example, not hydrolyzed by the enzyme and thus constituted the limit-
Bacteroides, Prevotella, Dysgonomonas, and Xylanibacter132 digest product in vitro. Solution of the tertiary structure of
Furthermore, these PULs serve as valuable molecular markers BtGH30_3 (PDB ID 5NGK) revealed a classic (α/β)8 TIM
for the predictive analysis of specific polysaccharide utilization in barrel fold typical of the family.162 A complex with the bespoke
bacterial genomes and metagenomes132 (see ref 103 for a recent inhibitor β-glucosyl-1,6-deoxynojirimycin (GlcDNJ; PDB ID
survey of a large number of Prevotella copri isolates). For 5NGL), suggested that the active-site cleft presented four
example, screening of 121 Bacteroides strains revealed that only 7 subsites, −2 to +2, versus the three subsites indicated by the
were able to utilize MLGs, and all possessed a syntenic MLGUL. biochemical data. The authors explain this distinction by
Subsequent analysis of publicly available human gut meta- suggesting that the oligosaccharides used for kinetic analysis
genome data revealed, in turn, that MLG utilization is do not fully recapitulate the bent conformation of the native
ubiquitous in human populations, commensurate with diverse polysaccharide, resulting in suboptimal occupation of the four
cereals forming a stable part of the diet.132 Although Firmicutes subsites in the deep U-shaped cleft of the enzyme.66
systems are much less studied, individual CAZyme families (e.g., Final saccharification of β(1→6)-gluco-oligosaccharides in
GH16, GH9, and GH5) have been implicated in MLG the periplasm appears to be mediated by the exo-β-glucosidase
utilization by ‘omics and biochemical approaches,104,125 which BtGH3, which was 30-fold more specific for β(1→6)-glycosidic
may have similar predictive value. linkages than the corresponding β(1→3) and β(1→4) linkages.
β(1→6)-Glucan Utilization Loci. Like the MLGUL, the The authors note that the gene encoding this β-glucosidase,
β(1→6)-glucan utilization locus of B. thetaiotaomicron (locus BT3314, was transcribed at a 10-fold lower level than the other
tags BT3309−3314 (“PUL1,6‑β‑glucan” in ref 66, here “β1,6-GUL” genes of the β1,6-GUL and further speculate that this may allow
for consistency160) represents another minimalist PUL compris- the cognate oligosaccharide ligand of the SusR-like regulator to
ing only two GHs, viz. a GH30 and a GH3 member (Figure 2). A maintain a significant steady-state level in the periplasm, thereby
syntenic PUL from B. ovatus was originally predicted to target sustaining upregulation.66 Because the commercially available
MLGs.131 However, exemplary work by Lowe, Gilbert, and pustulan was used for all biochemical work, it is presently
colleagues revealed instead that the B. thetaiotaomicron β1,6- unknown how BtGH30_3 and BtGH3 (and the other β1,6-GUL
GUL is specifically upregulated during growth on pustulan and components) accommodate the β(1→3) branches that would
yeast cell wall extract, which contains β(1→6)-glucans.66 be typically encountered on fungal cell wall β(1→6)-glucans. At
Moreover, these authors demonstrated that the encoded the same time, the non-negligible activity of BtGH3 toward the
SGBPs and GHs were highly specific for β(1→6)-glucans and β(1→6) linkage suggests that this enzyme may be sufficient to
the corresponding oligosaccharides. In total, the β1,6-GUL fully degrade all incoming oligosaccharides from more complex
encodes a SusR homologue, a SusC/SusD-like TBDT/SGBP substrates.
pair, a GH30 subfamily 3 (GH30_3) member, an additional In this context, it is interesting to note that yeast β(1→6)-
SGBP, and a GH3 member. The overall organization of this PUL glucan is believed to be cross-linked to α-mannans and
is notable in two ways. First, the auxiliary SGBP is not mannoproteins in the cell wall, suggesting that the β1,6-GUL
2094 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

may synergize with α-mannan-specific PULs. Yet, a broad survey G4G4G3G. Concordant with this specificity, phylogenetic
of Bacteroidetes indicated that syntenic β1,6-GULs are more analysis places BuGH16 in the “β-bulge laminarinase/MLGase”
widely distributed in a strain-specific manner than the α-mannan clade, corresponding to GH16 subfamily 3.147 This subfamily
PULs. Like the MLGUL, β1,6-GULs are found in human gut contains both laminarinases with secondary MLGase activity
resident Bacteroides, Prevotella, and Dysgonomonas strains, thus and specific MLGases,132 which require a β(1→3)-linkage
indicating a widespread importance of dietary β(1→6)-glucan between subsites −1 and −2 for substrate binding and catalysis,
metabolism.66 yet are tolerant of diverse linkages and branching in other
β(1→3)-Glucan Utilization Loci. Déjean et al. recently subsites.133
presented the functional characterization of three homologous In contrast to the broader substrate repertoire of BuGH16,
β(1→3)-glucan utilization loci (β1,3-GULs) from B. uniformis BuGH158 is a highly specific endo-β(1→3)-glucanase, with a
ATCC 8492 (locus tags BACUNI_01484−01490), B. f luxus strong preference for limited backbone branching. Family
YIT 12057 (locus tags HMPREF9446_00611−00616), and B. GH158 was recently discovered in 2019, and BuGH158 (PDB
thetaiotaomicron NLAE-zl-H207 (locus tags ID 6PAL) serves as the first structurally and mechanistically
H207DRAFT_02223−02229), which revealed that the ability characterized example. The tertiary structure of BuGH158
of individual taxa to utilize different β(1→3)- and β(1→3)/ revealed a (β/α)8 TIM barrel domain that places GH158 in the
β(1→4)-glucans arises through a combination of the individual large Clan GH-A and makes this the fourth family in GH-A
specificities of the cognate SGBPs and cell-surface GHs.133 Each known to contain endo-β(1→3)-glucanases (along with GH17,
of these β1,3-GULs contain a syntenic TBDT, SGBP-A (SusD GH128, and GH148). Superposition and analysis of the active-
homologue), SGBP-B, and GH3 member as a core set of site cleft vis-à-vis other endo-β(1→3)-glucanases revealed a
conserved genes yet notably differ in the nature of the outer- pocket formed by aromatic residues adjacent to the position of
membrane endo-glucanase. Specifically, B. thetaiotaomicron 6-OH of the glucose unit in the −1 subsite, suggesting a binding
possesses a GH16 member, B. f luxus possesses a GH158 site for a single β(1→6)-linked glucosyl branching residue
member, and B. uniformis possesses one of each (Figure 2a). typical of LdLam.133
Detailed enzyme kinetic studies on BuGH16 and BtGH16 Finally, Michaelis−Menten kinetics for a range of gluco-
demonstrated that both possessed broad substrate specificity, oligosaccharides indicated that BuGH3 is a specific exo-β(1→
that is, the ability to hydrolyze L. digitata laminarin (LdLam, 3)-glucosidase with a two-orders-of-magnitude higher kcat/Km
which contains single β(1→6)-linked glucosyl branches on the value for G3G disaccharide over G4G and G6G disaccharides.
β(1→3)-glucan backbone, Figure 1), Saccharomyces cerevisiae Despite this specificity, BuGH3 has the ability to completely
(yeast) β-glucan (yBG, which contains longer β(1→3)-linked hydrolyze into glucose all oligosaccharides produced by
side chains, Figure 1), and barley MLG (bMLG, with a β(1→ BuGH16 and BuGH158 from MLG and laminarin. It is also
3)/β(1→4)-linked backbone, Figure 1). In contrast, BuGH158 possible that other periplasmic glucosidases encoded outside of
and Bf GH158 were highly specific for LdLam. Thus, despite the β1,3GUL assist with alternate linkage degradation.133
possessing an SGBP-A capable of binding both LdLam and yBG, Using the β1,3GULs as markers of HGM metabolic potential,
and an SGBP-B capable of binding LdLam, yBG, and bMLG, the Déjean et al. also surveyed publicly available gut metagenomes
limited substrate range of Bf GH158 only permitted B. f luxus to from nearly 2500 adults on five different continents, viz. North
grow on the laminarin.133 America, South America, Africa, Europe, and Asia. The
In comparison, B. thetaiotaomicron could utilize both LdLam distribution of β1,3GULs was generally not restricted to any
and the more extensively branched yBG for growth, because particular geographic region or population, which likely reflects
BtGH16 could hydrolyze both of these β(1→3)-glucans, which the ubiquity of edible fungi and yeast fermentation products in
were also bound by the cognate BtSGBP-A and BtSGBP-B. On diets worldwide. Strikingly, however, β1,3GULs were not
the other hand, B. thetaiotaomicron was incapable of growth on detected in gut metagenomes of the indigenous Hadza and
bMLG, despite this being a substrate for BtGH16, because Yanomami tribes, perhaps due to the prevalence of Prevotella
neither BtSGBP-A nor BtSGBP-B recognized the β(1→3)/ over Bacteroides in these populations.133
β(1→4)-linked backbone. Finally, B. uniformis grew on all three β(1→2)-Glucan Utilization Loci. The extent to which
polysaccharides, viz. LdLam, yBG, and bMLG, due to the members of the HGM may utilize β(1→2)-glucans has not been
combined specificities of BuGH16 (hydrolyzing LdLam, yBG, explicitly studied. Intriguingly, “PUL61” from B. thetaiotaomi-
and bMLG), BuGH158 (hydrolyzing LdLam), and BuSGBP-B cron VPI-5482 (locus tags BT3566−3569) was first identified by
(binding LdLam, yBG, and bMLG). Notably, no binding affinity Martens et al. in 2008 as part of a genomic census, but at that
for these polysaccharides was observed for BuSGBP-A, although time, its substrate was unknown and difficult to predict based on
site-directed mutagenesis studies on other PUL systems have a limited knowledge of the GH complement (see Document S2,
demonstrated that a lack of substrate binding by SusD Figure S5 in ref 164). PUL61 is now known to encode both a
homologues can be compensated by a functional SGBP-B.156,163 GH3 member that is specific for β(1→2)-gluco-oligosacchar-
While the reader is referred to the original publication for a ides (PDB IDs 5XXL, 5XXM, 5XXN, 5XXO),165 and a GH144
detailed description of the biochemical analyses of all of the B. member (Figure 2a).150 GH144 was recently established as a
uniformis, B. thetaiotaomicron, and B. fluxus GHs and SGBPs, family of endo-β(1→2)-glucanases following seminal biochem-
some points regarding the structure and function of BuGH16 ical and structural work in 2017 on a Chitinophaga pinensis
and BuGH158 deserve special mention. Michaelis−Menten homologue.166 Thus far, no Bacteroides homologues have been
kinetics analysis revealed that BuGH16 was most active toward biochemically characterized, although unliganded tertiary
yBG and exhibited lower, similar activity against laminarins with structures of B. f ragilis (PDB ID 3EU8), B. uniformis (PDB ID
different branching frequency, as well as MLG. BuGH16 4GL3), and B. caccae (PDB ID 4QT9) have been deposited as
hydrolyzes the β(1→3)-glucans LdLam and yBG to glucose, part of a structural genomics initiative.167 Biochemical and
laminaribiose, and a branched trisaccharide, while bMLG is structural characterization of a GH144 member from Para-
completely converted into the oligosaccharides G4G3G and bacteroides distasonis revealed this enzyme to be an exo-sophoro-
2095 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

biohydrolase (i.e., releasing Glcβ(1→2)-Glc), the specificity of xyloglucanase BoGH5, possibly with the assistance of BoGH9,
which is dictated by the presence of a loop blocking one end of hydrolyzes the XyG backbone specifically at unbranched Glc
the active site (PDB ID 5Z06).168 With the benefit of hindsight units to generate shorter fragments. Following import into the
and experimental data on orthologous GH144 members, it is periplasm by the TBDT/SGBP-A (SusC/SusD-homologue)
highly probable that PUL61 and related PULs in other taxa are complex, the six exo-acting GHs deconstruct the resulting
responsible for β(1→2)-glucan utilization in the HGM. oligosaccharides in a concerted manner (Figure 3): Two GH43
Xyloglucan Utilization Loci. The xyloglucans, represented α-L-arabinofuranosidases and one GH2 β-galactosidase remove
by (fucogalacto)xyloglucan, (arabinogalacto)xyloglucan, and side chain capping Arafα(1→2) and Galpβ(1→2) residues,
other variants, are demonstrably the structurally most complex respectively, to free underlying Xylpα(1→6) residues. Hydrol-
of the β-glucans (Figure 1). Correspondingly, xyloglucan ysis of the side chain Xyl residue at the nonreducing end by the
utilization loci (XyGULs) comprise a greater number of GHs GH31 α-xylosidase frees the corresponding backbone Glc
than other β-GUL to address the monosaccharide diversity in residue for removal by the GH3 β-glucosidases. This α-
these plant cell wall polysaccharides169 (Figure 2a). The xylosidase/β-glucosidase cycle repeats to fully hydrolyze the
characterization of an exemplar XyGUL from B. ovatus (locus xyloglucan fragment.108
tags Bacova_02644−02659) was reported in 2014.108 This In addition to the novel structures of the SGBPs described
study utilized a combination of bacterial genetic, biochemical, above, structural biology of the XyGUL revealed specific insight
and structural biological approaches, which would be recapitu- into substrate specificity of the GHs through the solution of free
lated in subsequent PUL functional studies,170 including those and complexed structures of BoGH5 (PDB ID 3ZMR), BoGH9
described above. (PDB ID 6DHT), BoGH31 (PDB IDs 5JOU, 5JOV), BoGH3B
Foremost, genetic deletion of the XyGUL established its (PDB ID 5JP0), BoGH43A (PDB IDs 5JOW, 5JOX, 5JOY), and
unique requirement for growth on xyloglucan.108 Individual BoGH43B (PDB ID 5JOZ). While a full description of these
deletion of both SGBPs the GH5 endo-xyloglucanase defined the structures is beyond the scope of this review, particularly notable
essential roles of these XyGUL gene products in capturing the features include the following:
polysaccharide and initiating chain cleavage at the cell surface, • The endo-xyloglucanase BoGH5 possesses an N-terminal
respectively.108,163 Curiously, an additional cell-surface endo- PFAM PF13004 domain, which has previously been
glucanase, BoGH9, which is unique to B. ovatus among referred to as a “BACON (Bacteroidetes-associated
homologous XyGULs, exhibited feeble endo-xyloglucanase carbohydrate-binding),” although no experimental evi-
activity and was not required for growth on xyloglucan. A dence for a carbohydrate-binding function has been
subsequent study revealed that this enzyme is in fact more observed in this or other PF13004-containing pro-
specific for MLGs (including a requirement for calcium for teins.108,133 The BoGH5 structure represents the first
optimal activity) and may coordinate with the SGBPs for tertiary structure of a PF13004 domain, and the conserved
substrate capture.108,171 N-terminal positioning and linker flexibility suggest that
Tauzin et al. specifically revealed the structural basis of the these domains act as spacers to distance the correspond-
SusD homologue (SGBP-A; PDB IDs 5E75, 5E76) and the ing GH module from the cell surface.108
accessory SGBP (SGBP-B; PDB IDs 5E7G, 5E7H) in the
• The endo-(xylo)glucanase BoGH9 has an N-terminal Ig-
noncatalytic recognition of xyloglucan by B. ovatus (Figure
like fold domain and an (α/α)6 catalytic domain with a
3).163 Both SGBPs exhibit specificity for XyG and XyG
comparatively narrow active-site cleft that is typical of
oligosaccharides, while SGBP-A also exhibits limited affinity to
GH9 endo-glucanases (cellulases). Superposition with
cellohexaose, glucomannan, and MLGs. Thus, both SGBPs bind
other GH−oligosaccharide complexes, including
the β-glucan backbone residues, with increased affinity resulting
BoGH5:XXXG, suggests that the binding of xyloglucan
from XyG side-chain recognition. Oligosaccharide-complexed
is likely to be disfavored due to active-side clashes,
structures of SGBP-A and SGBP-B rationalized these bio-
whereas unbranched MLG is better accommodated,
chemical observations. The SGBP-A possessed the canonical
commensurate with kinetic analyses.171
SusD-fold, yet displayed an extended linear binding platform of
aromatic residues that accommodated seven Glc residues. In • The tertiary structures of the α-L-arabinofuranosidases
contrast, SGBP-B had a distinct four-domain extended BoGH43A and BoGH43B, which share limited sequence
architecture, which presented a single binding platform that identity (41%), were typical of GH43, comprising two
specifically accommodated five Glc resides on the C-terminal domains: a 5-bladed β-propeller domain embracing the
domain. These features distinguish the SGBP-B from the catalytic active site at the N-terminus and a β-sandwich
noncatalytic, multimodular amylose-binding proteins SusE and domain at the C-terminus. A complex of BoGH43A with
SusF, which are highly sequence-divergent despite their similar the bespoke inhibitor 1,4-dideoxy-1,4-imino-L-arabinitol
genomic organization within the Sus vis-à-vis the XyGUL. (“arabino-deoxynojirimycin”, AraDNJ) revealed key
Detailed reverse genetics experiments revealed that the presence substrate-binding interactions in the active-site pocket
of the SGBP-A but not its XyG-binding ability was strictly and the positions of the catalytic residues. BoGH43A was
essential to enable growth on XyG, commensurate with a tight also observed to have a wider active site pocket than
association to the TBDT. Whereas the SGBP-B was BoGH43B, which possibly explains the higher activity of
independently dispensable, its presence improved foraging of BoGH43A toward XyG oligosaccharide substrates.
shorter oligosaccharides, which might be released by other gut Although the reason for the presence of two GH43 is
bacteria.163 still unclear, the structural differences between active sites
Detailed enzyme kinetics and product analysis of the GHs may reflect tuning to as-yet unspecified side chains.172
encoded by the XyGUL enabled a complete pathway for the • The α-xylosidase BoGH31A consists of a central (β/α)8
saccharification of (arabinogalacto)xyloglucan to be established (TIM) barrel fold that is typical of GH31 catalytic
in B. ovatus. Operating at the cell surface, the vanguard endo- modules. Notably, a β-sandwich domain and a PA14
2096 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

domain at the N-terminus extend the active-site and systems grows, so too does our ability to use PULs as molecular
support the recognition of larger XyG oligosaccharide markers to predict the individual and combined metabolic
substrates.172 Specifically, BoGH31 shows higher activity capacity of members of the HGM.87 We anticipate that this will
toward XyG hepta- (“XXXG”, that is, Xyl3Glc4) and greatly enable microbiome-based “personalized medicine” to
nonasaccharides (“XLLG”, that is, Gal2Xyl3Glc4) than p- improve human nutrition and prevent or ameliorate disease.
nitrophenyl-α-xyloside, suggesting multiple substrate-
binding subsites.108
• The β-glucosidases BoGH3A and BoGH3B remove
■ AUTHOR INFORMATION
Corresponding Author
glucose from nonreducing ends of XyG oligosaccharides Harry Brumer − Michael Smith Laboratories, Department of
and cello-oligosaccharides. Although the β-glucosidase Chemistry, Department of Biochemistry and Molecular
BoGH3A resisted crystallization, the solution of the Biology, and Department of Botany, University of British
BoGH3B tertiary structure in complex with product Columbia, Vancouver, British Columbia V6T 1Z4, Canada;
(glucose) revealed three domains, including an N- orcid.org/0000-0002-0101-862X; Email: brumer@
terminal (TIM) barrel-like domain, a central α/β msl.ubc.ca
sandwich domain, and a C-terminal fibronectin type-III
(FN-III)-like domain.172 As for the GH43 paralogs in the Authors
XyGUL, the reason for the prima facie catalytic Benedikt Golisch − Michael Smith Laboratories and
redundancy of this GH3 pair is currently unclear. Department of Chemistry, University of British Columbia,
Vancouver, British Columbia V6T 1Z4, Canada
As described above for other β-glucan utilization loci, Zhenhuan Lei − Michael Smith Laboratories and Department
functional analysis of the B. ovatus XyGUL enabled the of Chemistry, University of British Columbia, Vancouver,
confident identification of homologous XyGULs in other British Columbia V6T 1Z4, Canada
Bacteroidetes. A perfect concordance was observed between Kazune Tamura − Michael Smith Laboratories and
the presence of a XyGUL and growth on xyloglucan in 70 strains Department of Biochemistry and Molecular Biology, University
(representing 6 individual species) of over 290 Bacteroidetes of British Columbia, Vancouver, British Columbia V6T 1Z4,
(primarily Bacteroides) strains surveyed. Thus, homologous Canada
XyGULs serve as molecular markers to reveal XyG metabolic
capability in human metagenomes, in which it was found to be Complete contact information is available at:
essentially ubiquitous across global populations.108 Of note, the https://pubs.acs.org/10.1021/acschembio.1c00563
GH43 α-L-arabinofuranosidases of the B. ovatus XyGUL suggest
specific adaptation to (arabinogalacto)xyloglucans found in Author Contributions

solanaceous fruits such as tomato, peppers, and eggplant. In B.G. and Z.L. made equal contributions. H.B. and K.T.
contrast, the presence of GH95 α-L-fucosidases in homologous conceived the overall structure of this review, which was drafted
XyGULs extends the substrate range to include (fucogalacto)- by B.G. and Z.L. and revised with input from all of the authors.
xyloglucans found generally in dicots, including most vegeta- Funding
bles.169 Presently, Firmicutes systems for xyloglucan utilization We thank the Canadian Institutes of Health Research (CIHR,
have not been extensively studied,104 but we anticipate that Operating Grant MOP-142472) and the Natural Sciences and
future functional analysis work in this area will enable more Engineering Research Council of Canada (NSERC, Discovery
informed analyses of (meta)genomes. Grant RGPIN-2018-03892) for funding. B.G. and K.T. are

■ CONCLUSION
The metabolism of β-glucans by the human gut microbiota is a
recipients of Four-Year Doctoral Fellowships from the
University of British Columbia.
Notes
key contributor to our nutrition and health. The molecular The authors declare no competing financial interest.


systems that address β-glucans in the HGM are as complex and
varied as the structures of the polysaccharides themselves. As ACKNOWLEDGMENTS
such, the fine structural analysis of dietary glycans and the
biochemical analysis of the proteins and enzymes responsible for We thank previous members of the Brumer lab and our
their metabolism are intrinsically linked.173 Although our focus collaborators for their invaluable contributions to our collective
in this review has been on the metabolism of β-glucans in the work on β-glucan metabolism in the microbiota.
human diet, it should be understood that the molecular systems
described here will, in many cases, have parallels in diverse
monogastric animals and other complex ecosystems. For
■ KEYWORDS
Dietary fiber: The components of plant-based foods that are
example, homologous β1,3-GULs have been identified and not broken down by human digestive enzymes.
biochemically characterized in marine Bacteroidetes.62,174 And Nonstarch polysaccharides (NSP): Complex carbohydrates
while the currently characterized β-glucan-specific PULs serve as with diverse structures that comprise the major component of
exemplar representatives, further functional characterization of dietary fiber.
homologues will undoubtedly be warranted, given the micro- Complex carbohydrates: Oligosaccharides and polysacchar-
heterogenity observed in PUL content across ides typically comprised of multiple monosaccharide residues
taxa.66,108,132,133,169 In the future, we also look forward to the and linkages.
genetic, biochemical, and structural characterization of potential β-Glucans: A family of polysaccharides that are related by a
β-glucan utilization systems in Firmicutes, Actinobacteria, and poly(glucosyl) backbone comprising β-anomeric linkages.
other members of the HGM, which are currently unexplored at Individual β-glucans are distinguished by the regiochemistry
this level of resolution.21 As our collective knowledge of such of the linkages joining two glucosyl residues in the chain.
2097 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

Human gut microbiota (HGM): The composite of all (12) Shoukat, M.; Sorrentino, A. Cereal beta-glucan: a promising
microorganisms that residue in the human gastrointestinal prebiotic polysaccharide and its impact on the gut health. Int. J. Food Sci.
tract. Technol. 2021, 56, 2088−2097.
Polysaccharide utilization locus (PUL): Physically linked and (13) Cummings, J. H.; Macfarlane, G. T. The control and
consequences of bacterial fermentation in the human colon. J. Appl.
coordinately regulated bacterial genes that encode a suite of
Bacteriol. 1991, 70, 443−459.
enzymes and other proteins responsible for the uptake and (14) Betian, H. G.; Linehan, B. A.; Bryant, M. P.; Holdeman, L. V.
hydrolysis to monosaccharides of complex glycans. Isolation of a cellulolytic Bacteroides sp. from human feces. Appl.
Bacteroidetes: A phylum of Gram-negative bacteria, members Environ. Microbiol. 1977, 33, 1009−1010.
of which predominate in the HGM. PULs were originally (15) Tosh, S. M.; Bordenave, N. Emerging science on benefits of
defined in Bacteroides species of this phylum. whole grain oat and barley and their soluble dietary fibers for heart
Firmicutes: A phylum of Gram-positive bacteria, members of health, glycemic response, and gut microbiota. Nutr. Rev. 2020, 78, 13−
which predominate in the HGM. The term “Gram-positive 20.
PUL (gpPUL)” has been coined to describe the distinct (16) Thomas, F.; Hehemann, J. H.; Rebuffet, E.; Czjzek, M.; Michel,
clusters of polysaccharide utilization genes in some species of G. Environmental and gut Bacteroidetes: the food connection. Front.
this phylum. Microbiol. 2011, 2, 93.
(17) Jayachandran, M.; Chen, J. L.; Chung, S. S. M.; Xu, B. J. A critical
Carbohydrate-active enzymes (CAZymes): Protein-based
review on the impacts of beta-glucans on gut microbiota and human
catalysts involved in the biosynthesis and biodegradation of health. J. Nutr. Biochem. 2018, 61, 101−110.
complex carbohydrates, comprising glycosyltransferases (18) Cummings, J. H.; Macfarlane, G. T. Role of intestinal bacteria in
(GTs), glycoside hydrolases (GHs), polysaccharide lyases nutrient metabolism. Clin. Nutr. 1997, 16, 3−11.
(PLs), carbohydrate esterases (CEs), and auxiliary activity (19) McNeil, N. I. The contribution of the large-intestine to energy
(AA) enzymes. supplies in man. Am. J. Clin. Nutr. 1984, 39, 338−342.
Surface glycan binding proteins (SGBP): Noncatalytic (20) Cummings, J. H. Short chain fatty-acids in the human-colon. Gut
carbohydrate-binding proteins located on the exterior of the 1981, 22, 763−779.
outer membrane of bacteria that facilitate the capture and (21) Fernandez-Julia, P. J.; Munoz-Munoz, J.; van Sinderen, D. A
uptake of complex carbohydrates. comprehensive review on the impact of beta-glucan metabolism by
Glycobiology: The study of the structure, biochemistry, and Bacteroides and Bifidobacterium species as members of the gut
microbiota. Int. J. Biol. Macromol. 2021, 181, 877−889.
biology of carbohydrates.


(22) IUB-IUPAC Joint Commission On Biochemical Nomenclature
(JCBN). Polysaccharide Nomenclature - Recommendations 1980. J.
REFERENCES Biol. Chem. 1982, 257, 3352−3354.
(1) Stein, O.; Granot, D. An Overview of Sucrose Synthases in Plants. (23) McNaught, A. D. International Union of Pure and Applied
Front. Plant Sci. 2019, 10, 95. Chemistry and International Union of Biochemistry and Molecular
(2) Carpita, N. C.; McCann, M. C. Cell walls. In Biochemistry and Biology - Joint Commission on Biochemical Nomenclature -
Molecular Biology of Plants; Buchanan, B. B., Gruissem, W., Jones, R. L., Nomenclature of carbohydrates (Recommendations 1996) (Reprinted
from Pure Appl Chem, vol 68, pg 1919−2008, 1996). Adv. Carbohydr.
Eds.; American Society of Plant Physiologists: Rockville, MD, 2000; pp
Chem. Biochem. 1997, 52 (52), 43−177.
52−108.
(24) McIntosh, M.; Stone, B. A.; Stanisich, V. A. Curdlan and other
(3) Magallanes-Cruz, P. A.; Flores-Silva, P. C.; Bello-Perez, L. A.
bacterial (1 -> 3)-beta-D-glucans. Appl. Microbiol. Biotechnol. 2005, 68,
Starch Structure Influences Its Digestibility: A Review. J. Food Sci. 2017,
163−173.
82, 2016−2023.
(25) Rop, O.; Mlcek, J.; Jurikova, T. Beta-glucans in higher fungi and
(4) El Kaoutari, A.; Armougom, F.; Gordon, J. I.; Raoult, D.;
their health effects. Nutr. Rev. 2009, 67, 624−631.
Henrissat, B. The abundance and variety of carbohydrate-active (26) Samuelsen, A. B. C.; Schrezenmeir, J.; Knutsen, S. H. Effects of
enzymes in the human gut microbiota. Nat. Rev. Microbiol. 2013, 11, orally administered yeast-derived beta-glucans: A review. Mol. Nutr.
497−504. Food Res. 2014, 58, 183−193.
(5) Cantarel, B. L.; Lombard, V.; Henrissat, B. Complex Carbohydrate (27) Tanna, B.; Mishra, A. Nutraceutical Potential of Seaweed
Utilization by the Healthy Human Microbiome. PLoS One 2012, 7, Polysaccharides: Structure, Bioactivity, Safety, and Toxicity. Compr.
e28742. Rev. Food Sci. Food Saf. 2019, 18, 817−831.
(6) Rose, D. R.; Chaudet, M. M.; Jones, K. Structural Studies of the (28) Laine, R. A. A calculation of all possible oligosaccharide isomers
Intestinal alpha-Glucosidases, Maltase-glucoamylase and Sucrase- both branched and linear yields 1.05 × 10(12) structures for a reducing
isomaltase. J. Pediatr. Gastroenterol. Nutr. 2018, 66, S11−S13. hexasaccharide - the isomer-barrier to development of single-method
(7) Williams, B. A.; Mikkelsen, D.; Flanagan, B. M.; Gidley, M. J. saccharide sequencing or synthesis systems. Glycobiology 1994, 4, 759−
″Dietary fibre″: moving beyond the ″soluble/insoluble″ classification 767.
for monogastric nutrition, with an emphasis on humans and pigs. J. (29) Zhu, F. M.; Du, B.; Xu, B. J. A critical review on production and
Anim. Sci. Biotechnol. 2019, 10, 45. industrial applications of beta-glucans. Food Hydrocolloids 2016, 52,
(8) McDougall, G. J.; Morrison, I. M.; Stewart, D.; Hillman, J. R. Plant 275−288.
cell walls as dietary fibre: Range, structure, processing and function. J. (30) Grundy, M. M. L.; Edwards, C. H.; Mackie, A. R.; Gidley, M. J.;
Sci. Food Agric. 1996, 70, 133−150. Butterworth, P. J.; Ellis, P. R. Re-evaluation of the mechanisms of
(9) Hutkins, R. W.; Krumbeck, J. A.; Bindels, L. B.; Cani, P. D.; Fahey, dietary fibre and implications for macronutrient bioaccessibility,
G.; Goh, Y. J.; Hamaker, B.; Martens, E. C.; Mills, D. A.; Rastal, R. A.; digestion and postprandial metabolism. Br. J. Nutr. 2016, 116, 816−
Vaughan, E.; Sanders, M. E. Prebiotics: why definitions matter. Curr. 833.
Opin. Biotechnol. 2016, 37, 1−7. (31) Burkus, Z.; Temelli, F. Effect of extraction conditions on yield,
(10) van de Guchte, M.; Blottiere, H. M.; Dore, J. Humans as composition, and viscosity stability of barley beta-glucan gum. Cereal
holobionts: implications for prevention and therapy. Microbiome 2018, Chem. 1998, 75, 805−809.
6, 81. (32) Garcia-Vaquero, M.; Rajauria, G.; O’Doherty, J. V.; Sweeney, T.
(11) Gibson, G. R.; Roberfroid, M. B. Dietary modulation of the Polysaccharides from macroalgae: Recent advances, innovative
human colonic microbiota - introducing the concept of prebiotics. J. technologies and challenges in extraction and purification. Food Res.
Nutr. 1995, 125, 1401−1412. Int. 2017, 99, 1011−1020.

2098 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

(33) Stone, B.; Clarke, A. E. Chemistry and biology of (1 -->3)-[beta]- cultured Grifola frondosa by solution NMR experiments. Carbohydr.
glucans; La Trobe University Press: Bundoora, Vic, 1992. Res. 2009, 344, 400−404.
(34) Varki, A.; Cummings, R. D.; Aebi, M.; Packer, N. H.; Seeberger, (56) Falch, B. H.; Espevik, T.; Ryan, L.; Stokke, B. T. The cytokine
P. H.; Esko, J. D.; Stanley, P.; Hart, G.; Darvill, A.; Kinoshita, T.; et al. stimulating activity of (1 -> 3)-beta-D-glucans is dependent on the
Symbol Nomenclature for Graphical Representations of Glycans. triple helix conformation. Carbohydr. Res. 2000, 329, 587−596.
Glycobiology 2015, 25, 1323−1324. (57) Karacsonyi, S.; Kuniak, L. Polysaccharides of Pleurotus ostreatus -
(35) Morgan, J. L. W.; Strumillo, J.; Zimmer, J. Crystallographic isolation and structure of pleuran, an alkali-insoluble beta-D-glucan.
snapshot of cellulose synthesis and membrane translocation. Nature Carbohydr. Polym. 1994, 24, 107−111.
2013, 493, 181−U192. (58) Komatsu, N.; Okubo, S.; Kikumoto, S.; Kimura, K.; Saito, G.;
(36) Mortensen, A.; Aguilar, F.; Crebelli, R.; Di Domenico, A.; Frutos, Sakai, S. Host-mediated antitumor action of schizophyllan: a glucan
M. J.; Galtier, P.; Gott, D.; Gundert-Remy, U.; Lambre, C.; et al. Re- produced by Schizophyllum commune. Gann 1969, 60, 137.
evaluation of xanthan gum (E415) as a food additive. EFSA J. 2017, 15, (59) Manners, D. J.; Masson, A. J.; Patterson, J. C.; Bjorndal, H.;
e04909. Lindberg, B. Structure of a beta-(1- 6)-D-glucan from yeast-cell walls.
(37) Chen, H.; Jiang, X. J.; Li, S. S.; Qin, W.; Huang, Z. Q.; Luo, Y. H.; Biochem. J. 1973, 135, 31−36.
Li, H.; Wu, D. T.; Zhang, Q.; Zhao, Y.; et al. Possible beneficial effects of (60) Magnani, M.; Calliari, C. M.; de Macedo, F. C.; Mori, M. P.;
xyloglucan from its degradation by gut microbiota. Trends Food Sci. Colus, I. M. D.; Castro-Gomez, R. J. H. Optimized methodology for
Technol. 2020, 97, 65−75. extraction of (1 -> 3)(1 -> 6)-beta-D-glucan from Saccharomyces
(38) Mikkelsen, M. S.; Jespersen, B. M.; Larsen, F. H.; Blennow, A.; cerevisiae and in vitro evaluation of the cytotoxicity and genotoxicity of
Engelsen, S. B. Molecular structure of large-scale extracted beta-glucan the corresponding carboxymethyl derivative. Carbohydr. Polym. 2009,
from barley and oat: Identification of a significantly changed block 78, 658−665.
structure in a high beta-glucan barley mutant. Food Chem. 2013, 136, (61) Kogan, G.; Alfoldi, J.; Masler, L. C-13-NMR spectroscopic
130−138. investigation of 2 yeast-cell wall beta-D-glucans. Biopolymers 1988, 27,
(39) Lazaridou, A.; Biliaderis, C. G. Molecular aspects of cereal beta- 1055−1063.
glucan functionality: Physical properties, technological applications and (62) Mystkowska, A. A.; Robb, C.; Vidal-Melgosa, S.; Vanni, C.;
physiological effects. J. Cereal Sci. 2007, 46, 101−118. Fernandez-Guerra, A.; Hohne, M.; Hehemann, J. H. Molecular
(40) Olafsdottir, E. S.; Ingolfsdottir, K. Polysaccharides from lichens: recognition of the beta-glucans laminarin and pustulan by a SusD-like
Structural characteristics and biological activity. Planta Med. 2001, 67, glycan-binding protein of a marine Bacteroidetes. FEBS J. 2018, 285,
199−208. 4465−4481.
(41) Salgado-Flores, A.; Hagen, L. H.; Ishaq, S. L.; Zamanzadeh, M.; (63) Knecht, J. C.; Schiffman, G.; Austrian, R. Some biological
Wright, A. D. G.; Pope, P. B.; Sundset, M. A. Rumen and Cecum properties of pneumococcus type-37 and chemistry of its capsular
Microbiomes in Reindeer (Rangifer tarandus tarandus) Are Changed in polysaccharide. J. Exp. Med. 1970, 132, 475.
Response to a Lichen Diet and May Affect Enteric Methane Emissions. (64) Rolin, D. B.; Pfeffer, P. E.; Osman, S. F.; Szwergold, B. S.;
PLoS One 2016, 11, e0155213. Kappler, F.; Benesi, A. J. Structural studies of a phosphocholine
(42) Synytsya, A.; Novak, M. Structural analysis of glucans. Annals of substituted beta-(1,3),(1,6) macrocyclic glucan from Bradyrhizobium
translational medicine 2014, 2, 17. japonicum USDA-110. Biochim. Biophys. Acta, Gen. Subj. 1992, 1116,
(43) Yotsuzuka, F. Curdlan. In Handbook of Dietary Fiber, 1st ed.; 215−225.
Cho, S. S., Dreher, M. L., Eds.; CRC Press, 2001; pp 737−757. (65) Sassaki, G. L.; Ferreira, J. C.; Glienke-Blanco, C.; Torri, G.; De
(44) Ohno, N.; Miura, T.; Miura, N. N.; Adachi, Y.; Yadomae, T. Toni, F.; Gorin, P. A. J.; Iacomini, M. Pustulan and branched beta-
Structure and biological activities of hypochlorite oxidized zymosan. galactofuranan from the phytopathogenic fungus Guignardia citricarpa,
Carbohydr. Polym. 2001, 44, 339−349. excreted from media containing glucose and sucrose. Carbohydr. Polym.
(45) Chen, X. Y.; Kim, J. Y. Callose synthesis in higher plants. Plant 2002, 48, 385−389.
Signaling Behav. 2009, 4, 489−492. (66) Temple, M. J.; Cuskin, F.; Basle, A.; Hickey, N.; Speciale, G.;
(46) Zhan, X. B.; Lin, C. C.; Zhang, H. T. Recent advances in curdlan Williams, S. J.; Gilbert, H. J.; Lowe, E. C. A Bacteroidetes locus
biosynthesis, biotechnological production, and applications. Appl. dedicated to fungal 1,6–glucan degradation: Unique substrate
Microbiol. Biotechnol. 2012, 93, 525−531. conformation drives specificity of the key endo-1,6–glucanase. J. Biol.
(47) Nakashima, A.; Yamada, K.; Iwata, O.; Sugimoto, R.; Atsuji, K.; Chem. 2017, 292, 10639−10650.
Ogawa, T.; Ishibashi-Ohgo, N.; Suzuki, K. beta-Glucan in Foods and Its (67) Maji, P. K.; Sen, I. K.; Behera, B.; Maiti, T. K.; Mallick, P.; Sikdar,
Physiological Functions. J. Nutr. Sci. Vitaminol. 2018, 64, 8−17. S. R.; Islam, S. S. Structural characterization and study of
(48) Kaur, R.; Sharma, M.; Ji, D. W.; Xu, M.; Agyei, D. Structural immunoenhancing properties of a glucan isolated from a hybrid
Features, Modification, and Functionalities of Beta-Glucan. Fibers mushroom of Pleurotus florida and Lentinula edodes. Carbohydr. Res.
2020, 8, 1. 2012, 358, 110−115.
(49) Saito, H.; Misaki, A.; Harada, T. A Comparison of the Structure (68) Dong, Q.; Yao, J.; Yang, X. T.; Fang, J. N. Structural
of Curdlan and Pachyman. Agric. Biol. Chem. 1968, 32, 1261−1269. characterization of a water-soluble beta-D-glucan from fruiting bodies
(50) Hoffmann, G. C.; Simson, B. W.; Timell, T. E. Structure and of Agaricus blazei Murr. Carbohydr. Res. 2002, 337, 1417−1421.
molecular size of pachyman. Carbohydr. Res. 1971, 20, 185. (69) Ciocchini, A. E.; Guidolin, L. S.; Casabuono, A. C.; Couto, A. S.;
(51) Novak, M.; Vetvicka, V. beta-glucans, history, and the present: de Iannino, N. I.; Ugalde, R. A. A glycosyltransferase with a length-
Immunomodulatory aspects and mechanisms of action. J. Immunotox- controlling activity as a mechanism to regulate the size of
icol. 2008, 5, 47−57. polysaccharides. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 16492−16497.
(52) Synytsya, A.; Novak, M. Structural diversity of fungal glucans. (70) Nakajima, M.; Tanaka, N.; Furukawa, N.; Nihira, T.; Kodutsumi,
Carbohydr. Polym. 2013, 92, 792−809. Y.; Takahashi, Y.; Sugimoto, N.; Miyanaga, A.; Fushinobu, S.; Taguchi,
(53) Sari, M.; Prange, A.; Lelley, J. I.; Hambitzer, R. Screening of beta- H.; Nakai, H. Mechanistic insight into the substrate specificity of 1,2-
glucan contents in commercially cultivated and wild growing beta-oligoglucan phosphorylase from Lachnoclostridium phytofermen-
mushrooms. Food Chem. 2017, 216, 45−51. tans. Sci. Rep. 2017, 7, 42671.
(54) Zhang, Y. Y.; Li, S.; Wang, X. H.; Zhang, L. N.; Cheung, P. C. K. (71) Abe, K.; Sunagawa, N.; Terada, T.; Takahashi, Y.; Arakawa, T.;
Advances in lentinan: Isolation, structure, chain conformation and Igarashi, K.; Samejima, M.; Nakai, H.; Taguchi, H.; Nakajima, M.; et al.
bioactivities. Food Hydrocolloids 2011, 25, 196−206. Structural and thermodynamic insights into beta-1,2-glucooligosac-
(55) Tada, R.; Adachi, Y.; Ishibashi, K. I.; Ohno, N. An unambiguous charide capture by a solute-binding protein in Listeria innocua. J. Biol.
structural elucidation of a 1,3-beta-D-glucan obtained from liquid- Chem. 2018, 293, 8812−8828.

2099 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

(72) Tayyib, S. Crop statistics; Food and Agriculture Organization of (92) Gill, S. R.; Pop, M.; DeBoy, R. T.; Eckburg, P. B.; Turnbaugh, P.
the United Nations (FAOStat), 2018. J.; Samuel, B. S.; Gordon, J. I.; Relman, D. A.; Fraser-Liggett, C. M.;
(73) Burton, R. A.; Fincher, G. B. (1,3;1,4)-beta-D-Glucans in Cell Nelson, K. E. Metagenomic analysis of the human distal gut
Walls of the Poaceae, Lower Plants, and Fungi: A Tale of Two Linkages. microbiome. Science 2006, 312, 1355−1359.
Mol. Plant 2009, 2, 873−882. (93) Mancabelli, L.; Milani, C.; Lugli, G. A.; Turroni, F.; Ferrario, C.;
(74) Stephen, A. M.; Champ, M. M. J.; Cloran, S. J.; Fleith, M.; van van Sinderen, D.; Ventura, M. Meta-analysis of the human gut
Lieshout, L.; Mejborn, H.; Burley, V. J. Dietary fibre in Europe: current microbiome from urbanized and pre-agricultural populations. Environ.
state of knowledge on definitions, sources, recommendations, intakes Microbiol. 2017, 19, 1379−1390.
and relationships to health. Nutr. Res. Rev. 2017, 30, 149−190. (94) Williams, R. D.; Olmsted, W. H. The manner in which food
(75) Henrion, M.; Francey, C.; Le, K. A.; Lamothe, L. Cereal B- controls the bulk of the feces. Ann. Intern. Med. 1936, 10, 717−727.
Glucans: The Impact of Processing and How It Affects Physiological (95) Williams, R. D.; Olmsted, W. H.; Hamann, C. H.; Fiorito, J. A.;
Responses. Nutrients 2019, 11, 1729. Duckles, D. The effect of cellulose, hemicellulose and lignin on the
(76) Fardet, A. New hypotheses for the health-protective mechanisms weight of the stool: A contribution to the study of laxation in man. J.
of whole-grain cereals: what is beyond fibre? Nutr. Res. Rev. 2010, 23, Nutr. 1936, 11, 433−449.
65−134. (96) Montgomery, L. Isolation of human colonic fibrolytic bacteria.
(77) Murphy, E. J.; Rezoagli, E.; Major, I.; Rowan, N. J.; Laffey, J. G. Lett. Appl. Microbiol. 1988, 6, 55−57.
beta-Glucan Metabolic and Immunomodulatory Properties and (97) Wedekind, K. J.; Mansfield, H. R.; Montgomery, L. Enumeration
Potential for Clinical Application. Journal of Fungi 2020, 6, 356. and isolation of cellulolytic and hemicellulolytic bacteria from human
(78) Goodridge, H. S.; Wolf, A. J.; Underhill, D. M. beta-glucan feces. Appl. Environ. Microbiol. 1988, 54, 1530−1535.
recognition by the innate immune system. Immunol. Rev. 2009, 230, (98) Robert, C.; Bernalier-Donadille, A. The cellulolytic microflora of
38−50. the human colon: evidence of microcrystalline cellulose-degrading
(79) Chan, G. C. F.; Chan, W. K.; Sze, D. M. Y. The effects of beta- bacteria in methane-excreting subjects. FEMS Microbiol. Ecol. 2003, 46,
glucan on human immune and cancer cells. J. Hematol. Oncol. 2009, 2, 81−89.
25. (99) Chassard, C.; Delmas, E.; Robert, C.; Lawson, P. A.; Bernalier-
(80) Wu, L. J.; Zhao, J.; Zhang, X. N.; Liu, S.; Zhao, C. Y. Antitumor Donadille, A. Ruminococcus champanellensis sp nov., a cellulose-
effect of soluble l3-glucan as an immune stimulant. Int. J. Biol. Macromol. degrading bacterium from human gut microbiota. Int. J. Syst. Evol.
2021, 179, 116−124. Microbiol. 2012, 62, 138−143.
(81) De Marco Castro, E.; Calder, P. C.; Roche, H. M. beta-1,3/1,6- (100) Robert, C.; Chassard, C.; Lawson, P. A.; Bernalier-Donadille, A.
Glucans and Immunity: State of the Art and Future Directions. Mol. Bacteroides cellulosilyticus sp nov., a cellulolytic bacterium from the
Nutr. Food Res. 2021, 65, 1901071. human gut microbial community. Int. J. Syst. Evol. Microbiol. 2007, 57,
(82) David, L. A.; Maurice, C. F.; Carmody, R. N.; Gootenberg, D. B.; 1516−1520.
Button, J. E.; Wolfe, B. E.; Ling, A. V.; Devlin, A. S.; Varma, Y.; (101) Lu, S. Y.; Mikkelsen, D.; Flanagan, B. M.; Williams, B. A.;
Fischbach, M. A.; Biddinger, S. B.; Dutton, R. J.; Turnbaugh, P. J. Diet Gidley, M. J. Interaction of cellulose and xyloglucan influences in vitro
rapidly and reproducibly alters the human gut microbiome. Nature fermentation outcomes. Carbohydr. Polym. 2021, 258, 117698.
2014, 505, 559. (102) Hartemink, R.; VanLaere, K. M. J.; Mertens, A. K. C.;
(83) Sonnenburg, E. D.; Smits, S. A.; Tikhonov, M.; Higginbottom, S. Rombouts, F. M. Fermentation of xyloglucan by intestinal bacteria.
K.; Wingreen, N. S.; Sonnenburg, J. L. Diet-induced extinctions in the Anaerobe 1996, 2, 223−230.
gut microbiota compound over generations. Nature 2016, 529, 212− (103) Fehlner-Peach, H.; Magnabosco, C.; Raghavan, V.; Scher, J. U.;
U208. Tett, A.; Cox, L. M.; Gottsegen, C.; Watters, A.; Wiltshire-Gordon, J.
(84) Delannoy-Bruno, O.; Desai, C.; Raman, A. S.; Chen, R. Y.; D.; Segata, N.; et al. Distinct Polysaccharide Utilization Profiles of
Hibberd, M. C.; Cheng, J.; Han, N.; Castillo, J. J.; Couture, G.; Lebrilla, Human Intestinal Prevotella copri Isolates. Cell Host Microbe 2019, 26,
C. B.; et al. Evaluating microbiome-directed fibre snacks in gnotobiotic 680.
mice and humans. Nature 2021, 595, 91−95. (104) Wegmann, U.; Louis, P.; Goesmann, A.; Henrissat, B.; Duncan,
(85) Gupta, V. K.; Paul, S.; Dutta, C. Geography, Ethnicity or S. H.; Flint, H. J. Complete genome of a new Firmicutes species
Subsistence-SpecificVariations in Human Microbiome Composition belonging to the dominant human colonic microbiota (’Ruminococcus
and Diversity. Front. Microbiol. 2017, 8, 1162. bicirculans’) reveals two chromosomes and a selective capacity to utilize
(86) Wilson, A. S.; Koller, K. R.; Ramaboli, M. C.; Nesengani, L. T.; plant glucans. Environ. Microbiol. 2014, 16, 2879−2890.
Ocvirk, S.; Chen, C. X.; Flanagan, C. A.; Sapp, F. R.; Merritt, Z. T.; (105) Ozcan, E.; Sun, J. D.; Rowley, D. C.; Sela, D. A. A Human Gut
Bhatti, F.; et al. Diet and the Human Gut Microbiome: An International Commensal Ferments Cranberry Carbohydrates To Produce Formate.
Review. Dig. Dis. Sci. 2020, 65, 723−740. Appl. Environ. Microbiol. 2017, 83, 16.
(87) Flint, H. J.; Duncan, S. H.; Louis, P. The impact of nutrition on (106) Feng, G. L.; Mikkelsen, D.; Hoedt, E. C.; Williams, B. A.;
intestinal bacterial communities. Curr. Opin. Microbiol. 2017, 38, 59− Flanagan, B. M.; Morrison, M.; Gidley, M. J. In vitro fermentation
65. outcomes of arabinoxylan and galactoxyloglucan depend on fecal
(88) Rinninella, E.; Raoul, P.; Cintoni, M.; Franceschi, F.; Miggiano, inoculum more than substrate chemistry. Food Funct. 2020, 11, 7892−
G. A. D.; Gasbarrini, A.; Mele, M. C. What is the Healthy Gut 7904.
Microbiota Composition? A Changing Ecosystem across Age, Environ- (107) Moro Cantu-Jungles, T.; do Nascimento, G. E.; Zhang, X. W.;
ment, Diet, and Diseases. Microorganisms 2019, 7, 14. Iacomini, M.; Cordeiro, L. M. C.; Hamaker, B. R. Soluble xyloglucan
(89) Almeida, A.; Mitchell, A. L.; Boland, M.; Forster, S. C.; Gloor, G. generates bigger bacterial community shifts than pectic polymers
B.; Tarkowska, A.; Lawley, T. D.; Finn, R. D. A new genomic blueprint during in vitro fecal fermentation. Carbohydr. Polym. 2019, 206, 389−
of the human gut microbiota. Nature 2019, 568, 499. 395.
(90) Qin, J. J.; Li, R. Q.; Raes, J.; Arumugam, M.; Burgdorf, K. S.; (108) Larsbrink, J.; Rogers, T. E.; Hemsworth, G. R.; McKee, L. S.;
Manichanh, C.; Nielsen, T.; Pons, N.; Levenez, F.; Yamada, T.; et al. ) A Tauzin, A. S.; Spadiut, O.; Klinter, S.; Pudlo, N. A.; Urs, K.; Koropatkin,
human gut microbial gene catalogue established by metagenomic N. M.; et al. A discrete genetic locus confers xyloglucan metabolism in
sequencing. Nature 2010, 464, 59−U70. select human gut Bacteroidetes. Nature 2014, 506, 498.
(91) Methe, B. A.; Nelson, K. E.; Pop, M.; Creasy, H. H.; Giglio, M. (109) Pleşea Condratovici, C.; Bacarea, V.; Pique, N. Xyloglucan for
G.; Huttenhower, C.; Gevers, D.; Petrosino, J. F.; Abubucker, S.; the Treatment of Acute Gastroenteritis in Children: Results of a
Badger, J. H.; et al. A framework for human microbiome research. Randomized, Controlled, Clinical Trial. Gastroenterology Research and
Nature 2012, 486, 215−221. Practice 2016, 2016, 6874207.

2100 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

(110) Gnessi, L.; Bacarea, V.; Marusteri, M.; Pique, N. Xyloglucan for Microbiota Toward Reduced Cardiovascular Disease Risk. Front.
the treatment of acute diarrhea: results of a randomized, controlled, Microbiol. 2016, 7, 129.
open-label, parallel group, multicentre, national clinical trial. BMC (129) Bai, J. Y.; Li, T. T.; Zhang, W. H.; Fan, M. C.; Qian, H. F.; Li, Y.;
Gastroenterol. 2015, 15, 153. Wang, L. Systematic assessment of oat beta-glucan catabolism during in
(111) Ostrowski, M. P.; La Rosa, S. L.; Kunath, B. J.; Robertson, A.; vitro digestion and fermentation. Food Chem. 2021, 348, 129116.
Pereira, G.; Hagen, L. H.; Varghese, N. J.; Qiu, L.; Yao, T.; Flint, G. The (130) Briggs, J. A.; Grondin, J. M.; Brumer, H. Communal living:
Food Additive Xanthan Gum Drives Adaptation of the Human Gut glycan utilization by the human gut microbiota. Environ. Microbiol.
Microbiota. bioRxiv 2021, DOI: 10.1101/2021.06.02.446819. 2021, 23, 15.
(112) Wang, Y. N.; Harding, S. V.; Thandapilly, S. J.; Tosh, S. M.; (131) Martens, E. C.; Lowe, E. C.; Chiang, H.; Pudlo, N. A.; Wu, M.;
Jones, P. J. H.; Ames, N. P. Barley beta-glucan reduces blood cholesterol McNulty, N. P.; Abbott, D. W.; Henrissat, B.; Gilbert, H. J.; Bolam, D.
levels via interrupting bile acid metabolism. Br. J. Nutr. 2017, 118, 822− N.; Gordon, J. I. Recognition and Degradation of Plant Cell Wall
829. Polysaccharides by Two Human Gut Symbionts. PLoS Biol. 2011, 9,
(113) Cicero, A. F. G.; Fogacci, F.; Veronesi, M.; Strocchi, E.; Grandi, e1001221.
E.; Rizzoli, E.; Poli, A.; Marangoni, F.; Borghi, C. A Randomized (132) Tamura, K.; Hemsworth, G. R.; DeJean, G.; Rogers, T. E.;
Placebo-Controlled Clinical Trial to Evaluate the Medium-Term Pudlo, N. A.; Urs, K.; Jain, N.; Davies, G. J.; Martens, E. C.; Brumer, H.
Effects of Oat Fibers on Human Health: The Beta-Glucan Effects on Molecular Mechanism by which Prominent Human Gut Bacteroidetes
Lipid Profile, Glycemia and inTestinal Health (BELT) Study. Nutrients Utilize Mixed-Linkage Beta-Glucans, Major Health-Promoting Cereal
2020, 12, 686. Polysaccharides. Cell Rep. 2017, 21, 417−430.
(114) Bjorck, I.; Granfeldt, Y.; Liljeberg, H.; Tovar, J.; Asp, N. G. Food (133) Déjean, G.; Tamura, K.; Cabrera, A.; Jain, N.; Pudlo, N. A.;
properties affecting the digestion and absorption of carbohydrates. Am. Pereira, G.; Viborg, A. H.; Van Petegem, F.; Martens, E. C.; Brumer, H.
J. Clin. Nutr. 1994, 59, 699S−705S. Synergy between Cell Surface Glycosidases and Glycan-Binding
(115) Lia, A.; Andersson, H.; Mekki, N.; Juhel, C.; Senft, M.; Lairon, Proteins Dictates the Utilization of Specific Beta(1,3)- Glucans by
D. Postprandial lipemia in relation to sterol and fat excretion in Human Gut Bacteroides. mBio 2020, 11, e00095-20.
ileostomy subjects given oat-bran and wheat test meals. Am. J. Clin. (134) El Khoury, D.; Cuda, C.; Luhovyy, B. L.; Anderson, G. H. Beta
Nutr. 1997, 66, 357−365. Glucan: Health Benefits in Obesity and Metabolic Syndrome. J. Nutr.
(116) Ellegard, L.; Andersson, H. Oat bran rapidly increases bile acid Metab. 2012, 2012, 851362.
excretion and bile acid synthesis: an ileostomy study. Eur. J. Clin. Nutr. (135) Camilli, G.; Tabouret, G.; Quintin, J. The Complexity of Fungal
2007, 61, 938−945. beta-Glucan in Health and Disease: effects on the Mononuclear
(117) Theuwissen, E.; Mensink, R. P. Water-soluble dietary fibers and Phagocyte System. Front. Immunol. 2018, 9, 673.
cardiovascular disease. Physiol. Behav. 2008, 94, 285−292. (136) Patidar, A.; Mahanty, T.; Raybarman, C.; Sarode, A. Y.; Basak,
(118) Rebello, C. J.; Greenway, F. L.; Finley, J. W. Whole Grains and
S.; Saha, B.; Bhattacharjee, S. Barley beta-Glucan and Zymosan induce
Pulses: A Comparison of the Nutritional and Health Benefits. J. Agric.
Dectin-1 and Toll-like receptor 2 co-localization and anti-leishmanial
Food Chem. 2014, 62, 7029−7049.
immune response in Leishmania donovani-infected BALB/c mice.
(119) Daou, C.; Zhang, H. Oat Beta-Glucan: Its Role in Health
Scand. J. Immunol. 2020, 92, e12952.
Promotion and Prevention of Diseases. Compr. Rev. Food Sci. Food Saf.
(137) Brown, G. D.; Gordon, S. Immune recognition of fungal beta-
2012, 11, 355−365.
glucans. Cell. Microbiol. 2005, 7, 471−479.
(120) Tosh, S. M. Review of human studies investigating the post-
(138) Jesenak, M.; Banovcin, P.; Rennerova, Z.; Majtan, J. beta-
prandial blood-glucose lowering ability of oat and barley food products.
Glucans in the treatment and prevention of allergic diseases. Allergol.
Eur. J. Clin. Nutr. 2013, 67, 310−317.
(121) Wanders, A. J.; van den Borne, J.; de Graaf, C.; Hulshof, T.; Immunopathol. 2014, 42, 149−156.
Jonathan, M. C.; Kristensen, M.; Mars, M.; Schols, H. A.; Feskens, E. J. (139) Hong, F.; Yan, J.; Baran, J. T.; Allendorf, D. J.; Hansen, R. D.;
M. Effects of dietary fibre on subjective appetite, energy intake and Ostroff, G. R.; Xing, P. X.; Cheung, N. K. V.; Ross, G. D. Mechanism by
body weight: a systematic review of randomized controlled trials. Obes. which orally administered beta-1,3-glucans enhance the tumoricidal
Rev. 2011, 12, 724−739. activity of antitumor monoclonal antibodies in murine tumor models. J.
(122) den Besten, G.; van Eunen, K.; Groen, A. K.; Venema, K.; Immunol. 2004, 173, 797−806.
Reijngoud, D. J.; Bakker, B. M. The role of short-chain fatty acids in the (140) Li, B.; Allendorf, D. J.; Hansen, R.; Marroquin, J.; Ding, C. L.;
interplay between diet, gut microbiota, and host energy metabolism. J. Cramer, D. E.; Yan, J. Yeast beta-glucan amplifies phagocyte killing of
Lipid Res. 2013, 54, 2325−2340. iC3b-opsonized tumor cells via complement receptor 3-Syk-phospha-
(123) Beck, E. J.; Tapsell, L. C.; Batterham, M. J.; Tosh, S. M.; Huang, tidylinositol 3-kinase pathway. J. Immunol. 2006, 177, 1661−1669.
X. F. Oat beta-glucan supplementation does not enhance the (141) Whitaker, R. J. Abigail Salyers: An Almost Unbeatable Force.
effectiveness of an energy-restricted diet in overweight women. Br. J. Women in Microbiology 2018, 243−251.
Nutr. 2010, 103, 1212−1222. (142) Salyers, A. A.; Vercellotti, J. R.; West, S. E. H.; Wilkins, T. D.
(124) Crittenden, R.; Karppinen, S.; Ojanen, S.; Tenkanen, M.; Fermentation of mucin and plant polysaccharides by strains of
Fagerstrom, R.; Matto, J.; Saarela, M.; Mattila-Sandholm, T.; Poutanen, Bacteroides from human colon. Appl. Environ. Microbiol. 1977, 33,
K. In vitro fermentation of cereal dietary fibre carbohydrates by 319−322.
probiotic and intestinal bacteria. J. Sci. Food Agric. 2002, 82, 781−789. (143) Seong, H.; Bae, J. H.; Seo, J. S.; Kim, S. A.; Kim, T. J.; Han, N. S.
(125) Alessi, A. M.; Gray, V.; Farquharson, F. M.; Flores-Lopez, A.; Comparative analysis of prebiotic effects of seaweed polysaccharides
Shaw, S.; Stead, D.; Wegmann, U.; Shearman, C.; Gasson, M.; Collie- laminaran, porphyran, and ulvan using in vitro human fecal
Duguid, E. S. R.; et al. beta-Glucan is a major growth substrate for fermentation. J. Funct. Foods 2019, 57, 408−416.
human gut bacteria related to Coprococcus eutactus. Environ. Microbiol. (144) Salyers, A. A.; Palmer, J. K.; Wilkins, T. D. Laminarinase (beta-
2020, 22, 2150−2164. glucanase) activity in bacteroides from human colon. Appl. Environ.
(126) Jaskari, J.; Kontula, P.; Siitonen, A.; Jousimies-Somer, H.; Microbiol. 1977, 33, 1118−1124.
Mattila-Sandholm, T.; Poutanen, K. Oat beta-glucan and xylan (145) Shang, Q. S.; Jiang, H.; Cai, C.; Hao, J. J.; Li, G. Y.; Yu, G. L. Gut
hydrolysates as selective substrates for Bifidobacterium and Lactoba- microbiota fermentation of marine polysaccharides and its effects on
cillus strains. Appl. Microbiol. Biotechnol. 1998, 49, 175−181. intestinal ecology: An overview. Carbohydr. Polym. 2018, 179, 173−
(127) Zhang, G. Y.; Hamaker, B. R. Cereal Carbohydrates and Colon 185.
Health. Cereal Chem. 2010, 87, 331−341. (146) Grondin, J. M.; Tamura, K.; Dejean, G.; Abbott, D. W.; Brumer,
(128) Wang, Y. A.; Ames, N. P.; Tun, H. M.; Tosh, S. M.; Jones, P. J.; H. Polysaccharide Utilization Loci: Fueling Microbial Communities. J.
Khafipour, E. High Molecular Weight Barley beta-Glucan Alters Gut Bacteriol. 2017, 199, e00860-16.

2101 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102
ACS Chemical Biology pubs.acs.org/acschemicalbiology Reviews

(147) Viborg, A. H.; Terrapon, N.; Lombard, V.; Michel, G.; Czjzek, beta-1,2-glucanase reveal a new glycoside hydrolase family. J. Biol.
M.; Henrissat, B.; Brumer, H. A subfamily roadmap of the Chem. 2017, 292, 7487−7506.
evolutionarily diverse glycoside hydrolase family 16 (GH16). J. Biol. (167) CAZypedia Consortium. Ten years of CAZypedia: a living
Chem. 2019, 294, 15973−15986. encyclopedia of carbohydrate-active enzymes. Glycobiology 2018, 28,
(148) Lombard, V.; Ramulu, H. G.; Drula, E.; Coutinho, P. M.; 3−8.
Henrissat, B. The carbohydrate-active enzymes database (CAZy) in (168) Shimizu, H.; Nakajima, M.; Miyanaga, A.; Takahashi, Y.;
2013. Nucleic Acids Res. 2014, 42, D490−D495. Tanaka, N.; Kobayashi, K.; Sugimoto, N.; Nakai, H.; Taguchi, H.
(149) Schwalm, N. D.; Groisman, E. A. Navigating the Gut Buffet: Characterization and Structural Analysis of a Novel exo-Type Enzyme
Control of Polysaccharide Utilization in Bacteroides spp. Trends Acting on beta-1,2-Glucooligosaccharides from Parabacteroides dis-
Microbiol. 2017, 25, 1005−1015. tasonis. Biochemistry 2018, 57, 3849−3860.
(150) Terrapon, N.; Lombard, V.; Drula, E.; Lapebie, P.; Al-Masaudi, (169) Dejean, G.; Tauzin, A. S.; Bennett, S. W.; Creagh, A. L.; Brumer,
S.; Gilbert, H. J.; Henrissat, B. PULDB: the expanded database of H. Adaptation of Syntenic Xyloglucan Utilization Loci of Human Gut
Polysaccharide Utilization Loci. Nucleic Acids Res. 2018, 46, D677− Bacteroidetes to Polysaccharide Side Chain Diversity. Appl. Environ.
D683. Microbiol. 2019, 85, e01491-19.
(151) Terrapon, N.; Lombard, V.; Gilbert, H. J.; Henrissat, B. (170) Ye, M.; Yu, J.; Shi, X.; Zhu, J.; Gao, X.; Liu, W. Polysaccharides
Automatic prediction of polysaccharide utilization loci in Bacteroidetes catabolism by the human gut bacterium -Bacteroides thetaiotaomicron:
species. Bioinformatics 2015, 31, 647−655. advances and perspectives. Crit. Rev. Food Sci. Nutr. 2020, 1−20.
(152) Foley, M. H.; Cockburn, D. W.; Koropatkin, N. M. The Sus (171) Foley, M. H.; Dejean, G.; Hemsworth, G. R.; Davies, G. J.;
operon: a model system for starch uptake by the human gut Brumer, H.; Koropatkin, N. M. A Cell-Surface GH9 Endo-Glucanase
Bacteroidetes. Cell. Mol. Life Sci. 2016, 73, 2603−2617. Coordinates with Surface Glycan-Binding Proteins to Mediate
(153) Pollet, R. M.; Martin, L. M.; Koropatkin, N. M. TonB- Xyloglucan Uptake in the Gut Symbiont Bacteroides ovatus. J. Mol.
dependent transporters in the Bacteroidetes: Unique domain structures Biol. 2019, 431, 981−995.
and potential functions. Mol. Microbiol. 2021, 115, 490−501. (172) Hemsworth, G. R.; Thompson, A. J.; Stepper, J.; Sobala, F.;
(154) Martens, E. C.; Kelly, A. G.; Tauzin, A. S.; Brumer, H. The Devil Coyle, T.; Larsbrink, J.; Spadiut, O.; Goddard-Borger, E. D.; Stubbs, K.
Lies in the Details: How Variations in Polysaccharide Fine-Structure A.; Brumer, H.; et al. Structural dissection of a complex Bacteroides
Impact the Physiology and Evolution of Gut Microbes. J. Mol. Biol. ovatus gene locus conferring xyloglucan metabolism in the human gut.
2014, 426, 3851−3865. Open Biol. 2016, 6, 160142.
(155) Hemsworth, G. R.; Dejean, G.; Davies, G. J.; Brumer, H. (173) Lapebie, P.; Lombard, V.; Drula, E.; Terrapon, N.; Henrissat, B.
Learning from microbial strategies for polysaccharide degradation. Bacteroidetes use thousands of enzyme combinations to break down
Biochem. Soc. Trans. 2016, 44, 94−108. glycans. Nat. Commun. 2019, 10, 2043.
(156) Tamura, K.; Foley, M. H.; Gardill, B. R.; Dejean, G.; Schnizlein, (174) Becker, S.; Scheffel, A.; Polz, M. F.; Hehemann, J. H. Accurate
M.; Bahr, C. M. E.; Creagh, A. L.; van Petegem, F.; Koropatkin, N. M.; Quantification of Laminarin in Marine Organic Matter with Enzymes
Brumer, H. Surface glycan-binding proteins are essential for cereal beta- from Marine Microbes. Appl. Environ. Microbiol. 2017, 83, e03389-16.
glucan utilization by the human gut symbiont Bacteroides ovatus. Cell.
Mol. Life Sci. 2019, 76, 4319−4340.
(157) Davies, G. J.; Wilson, K. S.; Henrissat, B. Nomenclature for
sugar-binding subsites in glycosyl hydrolases. Biochem. J. 1997, 321,
557−559.
(158) Jain, N.; Tamura, K.; Déjean, G.; Van Petegem, F.; Brumer, H.
Orthogonal Active-Site Labels for Mixed-Linkage endo-β-Glucanases.
ACS Chem. Biol. 2021, 16, 1968.
(159) Ketudat Cairns, J. R.; Esen, A. beta-Glucosidases. Cell. Mol. Life
Sci. 2010, 67, 3389−3405.
(160) McKee, L. S.; La Rosa, S. L.; Westereng, B.; Eijsink, V. G.; Pope,
P. B.; Larsbrink, J. Polysaccharide degradation by the Bacteroidetes:
mechanisms and nomenclature. Environ. Microbiol. Rep. 2021, 13, 559−
581.
(161) Ravcheev, D. A.; Godzik, A.; Osterman, A. L.; Rodionov, D. A.
Polysaccharides utilization in human gut bacterium Bacteroides
thetaiotaomicron: comparative genomics reconstruction of metabolic
and regulatory networks. BMC Genomics 2013, 14, 873.
(162) St John, F. J.; Gonzalez, J. M.; Pozharski, E. Consolidation of
glycosyl hydrolase family 30: A dual domain 4/7 hydrolase family
consisting of two structurally distinct groups. FEBS Lett. 2010, 584,
4435−4441.
(163) Tauzin, A. S.; Kwiatkowski, K. J.; Orlovsky, N. I.; Smith, C. J.;
Creagh, A. L.; Haynes, C. A.; Wawrzak, Z.; Brumer, H.; Koropatkin, N.
M. Molecular Dissection of Xyloglucan Recognition in a Prominent
Human Gut Symbiont. mBio 2016, 7, e02134-15.
(164) Martens, E. C.; Chiang, H. C.; Gordon, J. I. Mucosal Glycan
Foraging Enhances Fitness and Transmission of a Saccharolytic Human
Gut Bacterial Symbiont. Cell Host Microbe 2008, 4, 447−457.
(165) Ishiguro, R.; Tanaka, N.; Abe, K.; Nakajima, M.; Maeda, T.;
Miyanaga, A.; Takahashi, Y.; Sugimoto, N.; Nakai, H.; Taguchi, H.
Function and structure relationships of beta-1,2-glucooligosaccharide-
degrading beta-glucosidase. FEBS Lett. 2017, 591, 3926−3936.
(166) Abe, K.; Nakajima, M.; Yamashita, T.; Matsunaga, H.;
Kamisuki, S.; Nihira, T.; Takahashi, Y.; Sugimoto, N.; Miyanaga, A.;
Nakai, H.; et al. Biochemical and structural analyses of a bacterial endo-

2102 https://doi.org/10.1021/acschembio.1c00563
ACS Chem. Biol. 2021, 16, 2087−2102

You might also like