Running title: Synthesis of (1 β D-glycans* · 04/11/2010  · terminus of the acceptor glucan...

42
1 Running title: Synthesis of (14)-β-D-glycans* Title: Update on Mechanisms of Plant Cell Wall Biosynthesis: How plants make cellulose and other (14)-β-D-glycans Author: Nicholas C. Carpita Department of Botany & Plant Pathology, and Bindley Bioscience Center Purdue University 915 West State Street West Lafayette, IN 47907-2054 For correspondence: Tel. +1-765-494-4653 FAX: +1-765-494-0363 E-mail: [email protected] 1 This material is based upon work supported as part of the Center for Direct Catalytic Conversion of Biomass to Biofuels (C3Bio), an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Award Number DE-SC0000997 *Corresponding author; e-mail [email protected]. The author responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (www.plantphysiol.org) is: Nicholas C. Carpita ([email protected]). Plant Physiology Preview. Published on November 4, 2010, as DOI:10.1104/pp.110.163360 Copyright 2010 by the American Society of Plant Biologists https://plantphysiol.org Downloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

Transcript of Running title: Synthesis of (1 β D-glycans* · 04/11/2010  · terminus of the acceptor glucan...

  • 1

    Running title: Synthesis of (1→4)-β-D-glycans*

    Title: Update on Mechanisms of Plant Cell Wall Biosynthesis: How plants make

    cellulose and other (1→4)-β-D-glycans

    Author: Nicholas C. Carpita

    Department of Botany & Plant Pathology, and

    Bindley Bioscience Center

    Purdue University

    915 West State Street

    West Lafayette, IN 47907-2054

    For correspondence:

    Tel. +1-765-494-4653

    FAX: +1-765-494-0363

    E-mail: [email protected]

    1This material is based upon work supported as part of the Center for Direct Catalytic

    Conversion of Biomass to Biofuels (C3Bio), an Energy Frontier Research Center funded by the

    U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Award Number

    DE-SC0000997

    *Corresponding author; e-mail [email protected].

    The author responsible for distribution of materials integral to the findings presented in this

    article in accordance with the policy described in the Instructions for Authors

    (www.plantphysiol.org) is: Nicholas C. Carpita ([email protected]).

    Plant Physiology Preview. Published on November 4, 2010, as DOI:10.1104/pp.110.163360

    Copyright 2010 by the American Society of Plant Biologists

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 2

    The discovery of a gene that encodes a cotton (Gossypium hirsutum) cellulose synthase (Pear et

    al., 1996) revolutionized and invigorated the plant cell wall community to find the genes that

    encode the machinery of cell wall polysaccharide synthesis. The landscape was framed by the

    completion of the genome sequence of Arabidopsis thaliana (Arabidopsis Genome Initiative,

    2000), which gave a complete gene inventory for a model plant species, but one with many genes

    yet to be annotated for function. An estimated 10% of the plant genome, about 2,500 genes, is

    devoted to construction, dynamic architecture, sensing functions and metabolism of the plant cell

    wall. Based largely on prior discoveries of function in prokaryotic organisms, most of the

    tentatively annotated genes are organized into gene families for substrate generation, glycosyl

    transfer, targeting and trafficking, cell wall rearrangement and modification by hydrolases,

    esterases, and lyases (Yong et al., 2005; Penning et al., 2009). However, the biochemical

    activities of most enzymes involved in glycosyl transfer within these families remain to be

    verified, and an additional 40% of the genome encodes genes for whose function is not known.

    As many of these proteins contain secretory signal peptides (Arabidopsis Genome Initiative,

    2000), it is reasonable to infer that some have roles in cell wall construction.

    The Arabidopsis cellulose synthase/cellulose-synthase-like (CesA/Csl) gene superfamily,

    which includes 10 CesA genes and 29 Csl genes in six distinct groups, was one of the first large

    families to be described (Richmond and Somerville, 2000), and comparative analyses of a

    reference dicot, Arabidopsis, to a reference grass, rice (Oryza sativa), revealed substantive

    differences in family structures, adding two groups not seen in the dicot genome (Hazen et al.,

    2002). Extension of these annotations to compare all cell wall-related gene families of the

    grasses with those of the dicots reveals some correlation of family structure with the differences

    between plants with Type I walls, with those of the grasses with Type II walls (Fig. 1A; Penning

    et al., 2009). For CesAs and certain Csl genes, establishment of specific function for the

    synthases they encode comes from analysis of mutants lacking a particular function and in some

    specific examples, by heterologous expression. However, genetic approaches alone do not

    inform us about the biological mechanism of synthesis. The knowledge gained from molecular

    genetic approaches now needs to be augmented by biochemical and cell biological approaches to

    achieve a greater understanding of proteins and their interactions within a synthase complex,

    their organization at membranes, and their dynamics. This update focuses on the biochemical

    mechanisms of the synthesis of a single type of linkage, the (1→4)-β-D- linkage, in which one

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 3

    sugar is inverted nearly 180° with respect to each neighboring sugar in the chain. This linkage

    presents a unique steric problem for processive catalysis that all living organisms have solved,

    but we are still struggling to understand.

    This update reviews our present state of knowledge of the biochemical mechanisms of

    polysaccharide synthesis, including some classic discoveries, and presents an alternative

    hypothesis on the biochemical mechanisms and organization of complexes involved in synthase

    reactions that yield (1→4)-β-D- linkages.

    Cellulose synthesis

    In flowering plants, cellulose is a para-crystalline array of about two to three dozen (1→4)-β-D-

    glucan chains. Microfibrils of 36 glucan chains have a theoretical diameter of 3.8 nm, but X-ray

    scattering and NMR spectroscopy indicate some microfibril diameters could be as small as 2.4

    nm, or about two dozen chains (Kennedy et al., 2007). The microfibrils are synthesized at the

    plasma membrane by terminal complexes of six-membered ‘particle rosettes’ that produce a

    single microfibril (Giddings et al., 1980; Mueller and Brown, 1980). Thus, each of the six

    components of the particle rosette is expected to synthesize four to six of the glucan chains, and

    24 to 36 chains are then assembled into a functional microfibril (Doblin et al., 2002). In freeze-

    fracture, the particle rosettes, found only on the P-face of the membrane, are about 25 nm in

    diameter, but this size represents only the membrane-spanning and short exterior domains (Fig.

    2A). Hidden in surface views of rosette structures in the plasma membrane, the much larger

    catalytic domains of the cellulose synthases are estimated to be 50 nm wide and extend 35 nm

    into the cytoplasm (Bowling and Brown, 2008), a feature that has escaped consideration in many

    published models of the rosette structure (Fig. 2B and C).

    Cellulose synthase is an ancient enzyme (Nobles et al., 2001), and cellulose synthase genes

    in green algae are homologous to those of flowering plants (Roberts et al., 2002). The deduced

    amino acid sequences of CesAs share regions of similarity with the bacterial CesA proteins,

    namely the four catalytic motifs containing the D, DxD, D, Q/RxxRW that are highly conserved

    among those that synthesize several kinds of (1→4)-β-D-glycans (Saxena et al., 1995). The

    higher-plant CesA genes are predicted to encode polypeptides of about 110 kDa, each with a

    large, cytoplasmic N-terminal region containing zinc-finger (ZnF) domains, and eight membrane

    spans sandwiching the four U-motifs of the catalytic domain (Fig. 1B; Delmer, 1999).

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 4

    Further evidence for the function of plant CesA genes in cellulose synthesis came from

    Arabidopsis mutants of three of the CesA genes involved in primary wall synthesis: a

    temperature-sensitive radial swelling mutant rsw1 (AtCesA1; Arioli et al., 1998), a dwarf-

    hypocotyl mutant procuste (prc1) (AtCesA6; Fagard et al., 2000), and a stunted root phenotype

    with altered jasmonate and ethylene signaling (cev1) and ectopic lignification (eli1) mutant

    alleles (AtCesA3; Ellis and Turner, 2002; Caño-Delgado et al., 2003). Despite co-expression in

    the same cells and an expectation of redundancy, cellulose synthesis is impaired in each mutant.

    The same was observed with the irregular xylem mutants, irx1, irx3, and irx5, which display a

    phenotype of collapsed mature xylem cells as a result of lowered cellulose content during

    secondary cell wall deposition (Taylor et al., 2001, 2003). A widely accepted hypothesis is that

    the AtCesA1, AtCesA3, and AtCesA6 proteins assemble to function in primary wall cellulose

    synthesis, while the AtCesA4, AtCesA7 and AtCesA8 proteins assemble to make secondary wall

    cellulose (Fig. 1A), with each member of the trio performing a non-redundant function in the

    complex (Taylor, 2008). Lack of one CesA prevents incorporation of the other two into the

    plasma membrane (Gardiner et al., 2003). However, at least some of the subunits are potentially

    interchangeable, as inferred by the dominant negative inhibition of growth and primary wall

    thickness caused by constitutive expression of a mutated fra5 (irx3 allele) transgene (Zhong et

    al., 2003), and by the semi dominant-negative phenotype observed in the heterozygous AtCesA3

    mutant (Daras et al., 2009). AtCesA1 is essential for cellulose synthesis (Beeckman et al., 2002),

    whereas knock-outs of AtCesA3 (Ellis and Turner, 2002; Caño-Delgado et al., 2003) and AtCes6

    (Fagard et al., 2000) result in partially impaired synthesis but not total inhibition. Desprez et al.

    (2007) indicate that the AtCesA2 and AtCesA5 proteins have partially redundant functions with

    AtCesA6.

    A direct association of three distinct CesA polypeptides was demonstrated in vitro and by co-

    localization in vivo by Taylor et al. (2003). Domain swap experiments with wild-type and mutant

    AtCesA1 and AtCesA3 proteins in their respective mutants resulted in dominant-positive and

    dominant-negative effects, consistent with both catalytic and C-terminus domains being

    important for function (Wang et al., 2006). Direct interactions of three distinct CesA

    polypeptides in vivo were shown by bimolecular fluorescence complementation (Desprez et al.,

    2007). Although some complementary pairs gave stronger fluorescence than others, both homo-

    and heterodimers of AtCesA1, AtCesA3 and AtCesA6 are inferred. Wang et al. (2008) used pull-

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 5

    down experiments similar to those of Taylor et al. (2003) to show that these three primary wall

    CesA proteins interact. Further, they showed that Triton-soluble microsomal preparations

    subjected to native PAGE gave an 840 kDa complex, and that null mutants, but not missense

    mutations, gave smaller 420 kDa complexes (Wang et al., 2008). Atanassov et al. (2009)

    affinity-trapped a ladder of complexes of CesA oligomers to about 700-730 kDa. Consistent with

    the observations of Wang et al. (2008), only smaller oligomeric complexes of two of the CesAs

    are detected when the third is missing (Atanassov et al., 2009). Such an association of CesAs

    was indicated independently in yeast two-hybrid studies (Timmers et al., 2009).

    Does synthesis of each (1→4)-β-D-glucan chain require one or two catalytic polypeptides?

    After over four decades of study, the biochemical mechanism by which cellulose is made

    remains a mystery (Delmer, 1999; Saxena and Brown, 2005; Somerville, 2006; Guerreiro et al.,

    2010), with only a few reports of cellulose synthesis in vitro with isolated membranes (e.g.

    Kudlicka and Brown, 1997; Lai-Kee-Him et al., 2002). For both cellulose and the related (1→4)-

    β-D-glycan, chitin, synthesis proceeds by attachment of glucosyl residues to the non-reducing

    terminus of the acceptor glucan chain (Koyama et al., 1997; Imai et al., 2003). The simplest

    hypothesis is that each CesA polypeptide synthesizes a single glucan chain. In the Delmer

    (1999) model, the eight membrane spans form a channel through which a single glucan chain is

    extruded (Fig. 3A). This mode of synthesis comes with a very big steric problem for synthesis.

    To make a (1→4)-β-D- linkage means that each glucosyl residue is turned 180° with respect to

    each neighbor. Thus, the O-4 position of non-reducing terminal sugar of the acceptor chain is

    displaced several Ångstroms upon addition of each successive unit (Fig. 3B). For the next

    glycosyl transfer to occur, then the site of catalysis must move several Ångstroms within the

    protein, the acceptor chain must swivel 180°, or the catalytic or acid-base amino acids must

    toggle between two forms to account for the displacement. To overcome this steric problem,

    several models have proposed that two sites or modes of glycosyl transfer reside within the

    catalytic complex, so that disaccharide units are added iteratively (Carpita et al., 1996; Koyama

    et al., 1997; Carpita and Vergara, 1998; Buckeridge et al., 1999, 2001; Saxena et al., 2001), or

    that two polypeptides associate to form two opposing catalytic sites (Vergara and Carpita, 2001;

    Buckeridge et al., 2001). In either model, glycobiosyl units, or any even-numbered oligomeric

    units, are added to the non-reducing end to ensure that the (1→4)-β-D- linkages are strictly

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 6

    preserved without inversion of substrate, active site, or terminus of the growing chain (Fig. 3C).

    Despite the rationale for a two-site model of catalysis, biochemical evidence from other

    types polysaccharide synthases indicate that a single polypeptide is sufficient. Hyaluronan (HA)

    is an unbranched polysaccharide composed of repeating units of (1→3)-β-D-N-

    acetylglucosamine and (1→4)-β-D-glucuronic acid (DeAngelis and Weigel, 1994). HA synthases

    exist in three distinct classes, with Class I containing integral membrane proteins to transport the

    HA across the membrane (Fig. 4). Bacterial and mammalian HA synthases have been shown to

    contain both transferase activities in a single polypeptide (DeAngelis and Weigel, 1994; Yoshida

    et al., 2000; Williams et al., 2006). Such a finding argues that synthesis of a single HA polymer

    requires only a single polypeptide.

    The crystal structure of a non-processive family 2 glycosyltransferase (GTs) sharing

    sequence similarity with a portion of the catalytic domain of a CesA, the Bacillus subtilis SpsA

    synthase, provided the first conformation of the active site and the role of the aspartyl residues in

    the positioning of the uridinyl group of a UDP-sugar (Fig. 5A; Charnock and Davies, 1999;

    http://www.pdb.org/pdb/explore/explore.do?structureId=1QGS). Charnock and colleagues

    (2001) argued that only a single site for a nucleotide-sugar substrate is accommodated within a

    single polypeptide of SpsA.

    A catalytic dimer hypothesis

    From the studies of the Class I HA synthases and the SpsA crystal structure, it is inferred that a

    single polypeptide alone has all the features needed for synthesis. However, these features still

    do not address mechanistically the fundamental steric problem of synthesis of a repeating

    (1→4)-β-D-glucosyl linkage of the cellulose glucan chains. In fact, two recent studies

    demonstrate unequivocally that at least some HA synthases and homologs of SpsA synthase

    form dimers. An identical structure as the SpsA synthase is the 3BCV polypeptide from

    Bacteroides fragilis, which is also predicted to contain a single substrate-binding site. In contrast

    to SpsA synthase, the 3BCV protein crystallizes as a dimer, and each monomer possesses a

    bound UDP (Fig. 5B; (http://www.pdb.org/pdb/explore/explore.do?structureId=3BCV). The

    dimerization occurs through the C-terminal regions, which appear to be flexible and, for this

    reason, were deleted from the crystal structure of the SpsA synthase (Charnock and Davies,

    1999). This is the first direct evidence by crystal structure of a homodimer formed by GT2

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 7

    proteins, but even more complicated structures are also found. For example, an E. coli strain K4

    chondroitin polymerase contains two ‘Rossmann fold-like’ domains within a single polypeptide,

    each binding a UDP-GlcA or UDP, and it also crystallizes as a dimer giving a total four

    nucleotide binding domains (Fig. 5C;

    (http://www.pdb.org/pdb/explore/explore.do?structureId=2Z86). Although the fit is not

    exceptionally good, the CesA catalytic domain threads through the E. coli chondroitin

    polymerase best of all for known crystal structures of glycosyl transferases involving nucleotide

    sugars (D. Kihara, Purdue University, personal communication), and it is consistent with the

    suggestion by Brown and Saxena (2000) and Saxena et al. (2001) of a conformation within

    which the catalytic domain of a single CesA would allow synthesis of cellobiose units of the

    chain within a single polypeptide.

    The Class II synthases contain two different types of GT2 modules but not the membrane-

    spanning domains (Fig. 4). One type of HA synthase possesses two repeats of the UDP-Glc and

    acceptor binding domains, so the synthesis of the characteristic disaccharide of HA by a single

    synthase is rationalized (Jing and DeAngelis, 2000). However, a direct interaction of two

    synthases is inferred for HA synthesis to explain the finding that host cells harboring constructs

    in which each site is independently disrupted are still able to make HA (Jing and DeAngelis,

    2000; Weigel and DeAngelis, 2007). Because the Class I HA synthases have both activities on a

    single peptide does not preclude the possibility that formation of homodimers of single isoforms

    of HA synthase is necessary for function, which, like the chrondroitin polymerase, would give

    four nucleotide sugar binding sites per dimer.

    Solving the steric problem aside, one must also ask if a channel of 8 membrane spans

    proposed for the cellulose synthase of sufficient size to extrude a (1→4)-β-D-glucan chain. The

    question about sufficient channel size was raised also with respect to the GT family 2 HA

    synthases by Weigel and DeAngelis (2007), who suggested that certain phospholipids required

    for activity possibly integrate with the membrane spans to widen the channel for extrusion.

    However, whether lipid interactions with a small number of domains would be significant is still

    in question. Callose synthases are about twice the size of CesAs and contain 16 membrane spans,

    i.e. double those of a CesA (Hong et al., 2001). Plasma membrane hexose and maltose

    transporters of prokaryotes and eukaryotes are homologous (Maiden et al., 1987), and virtually

    all of them contain a minimum of 11 to as many as 18 transmembrane spans per functional unit

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 8

    (Pao et al., 1998; Reifenberger et al., 1995; Sherson et al., 2000; Klepek et al., 2009).

    The sensitivity of detergent-solubilized CesAs complexes to DTT suggested to Atanassov et

    al. (2009) that disulfide bonds are involved in the coupling into the larger complexes. Other

    features of the protein outside the region of catalysis, such as the ZnFs, which show high

    similarity to RING-finger domains that bind Zn in a ‘cross-brace’ manner (Freemont, 2000),

    might function in the organization of these into the larger rosette structure. Kurek et al. (2002)

    proposed that CesAs are coupled through the ZnF domain in a redox-dependent manner,

    constituting the first step in the clustering of CesAs into rosettes. Moreover, the discovery that a

    thioredoxin-like protein associates with the CesA ZnF domain in a yeast two-hybrid screen led to

    the suggestion that reduction of the domains by oxidoreductases returns the CesAs to monomeric

    forms, which are directed to the ubiquitin-dependent turnover pathway (Kurek et al., 2002). The

    experimental herbicide CGA 325’615 blocks crystallization of the β-D-glucan chains into

    cellulose microfibrils, phenocopying the rsw1 swollen root tip (Peng et al., 2001). This

    phenotype can be abrogated completely in the presence of H2O2, suggesting the inhibitor blocks

    rosette assembly by enzymatic oxidation (Kurek et al., 2002).

    If ZnF domains of two CesAs couple as part of the recruitment into rosette particles, the

    question remaining is how all the others interact to form a complete complex. Although not

    discussed specifically, the study by Kurek et al. (2002) presented data that full-length CesA

    proteins formed tetramers and even higher ordered pairings, while the ZnF domains were limited

    to coupling of a single pair. Timmers et al. (2009) showed that heterodimer interactions indicated

    by yeast two-hybrid analysis do not require the ZnF domains. Taken together, these data provide

    evidence that domains other than the ZnFs of the CesA participate in coupling reactions if two to

    three dozen CesAs or more are aggregated to form a rosette complex.

    Comparison of CesA sequences suggests potential heterodimeric interaction domains

    within the catalytic domain. The initial scarcity of CesA protein sequences and the apparent

    variability within the so-called HVR led to the assumption that this region was probably not

    essential in catalysis (Pear et al., 1996). However, it is now understood that these regions are

    well conserved across grass and dicot species with distinct sub-clade structure. Potential protein-

    protein interactions through sub-domains of this region containing conserved Cys residues,

    clusters of consecutive basic Lys and Arg residues, and clusters of acidic Asp and Glu residues

    forming the basis of a Class Specific Region (Fig. 1B; Vergara and Carpita, 2001).

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 9

    To test the catalytic dimer hypothesis, we expressed fusion proteins containing only the

    catalytic domain of Arabidopsis and maize (Zea mays) CesAs with affinity tags and observed

    that dimers and higher-order aggregates collapse reversibly to monomeric forms by thiol

    reducing agents (C. Rayon, A. Olek, L. Paul, S. Ghosh, unpublished results). Because the ZnF

    was absent in these constructs, dimerization must occur through thiol-sensitive sequences in the

    catalytic domain. Such an interaction of CesAs to form homo- or hetero-dimers solves the three

    basic problems of the single polypeptide-single polymer conundrum: 1) the steric problem is

    solved by coordinate synthesis and attachment of cellobiose units instead of monomers,

    preserving the integrity of the O-4 site of attachment at the non-reducing terminus of the chain,

    2) a channel of 16 membrane spanning domains is consistent with sugar transport and callose

    extrusion, and 3) the interaction produces two ZnF domains for recruitment of the catalytic dimer

    into rosette particles (Fig. 6). An exciting prospect is that conservation of space would be

    maintained if CesAs turn out to have a structure like the chondroitin polymerase dimers and

    function like Class II HA synthases, because a CesA catalytic dimer with four nucleotide binding

    domains would be capable of generating two (1→4)-β-D-glucan chains instead of just one.

    Further experiments are needed to establish preferred heterodimer interactions, the stoichiometry

    of UDP-Glc binding, and the role of the ZnF in recruitment of the catalytic dimers into larger

    complexes.

    The biological synthesis of cellulose

    Cellulose synthase has a half-life of less than thirty minutes, remarkably short for a

    membrane protein (Jacob-Wilk et al., 2006). Assembly of rosettes occurs in the Golgi stacks, and

    they must be continually secreted to the plasma membrane to maintain cellulose synthesis

    (Haigler and Brown, 1986). Additional proteins are suspected to be necessary for the formation

    of primers of polymer synthesis, metabolic channeling of substrates, chain crystallization of the

    chains, and termination of chains (Doblin et al., 2002; Somerville, 2006; Guerreiro et al., 2010).

    Further, a proteomics survey of plasma membrane proteins shows that certain CesAs are

    phosphorylated at several locations within the catalytic and N-terminal domains (Nühse et al.,

    2004). Modification of some potential phosphorylation sites with amino acids that either prevent

    (Ala) or mimick (Glu) phoshorylation have multiple effects that either reduce synthesis rates or

    interactions with the microtubule cytoskeleton independently (Chen et al., 2010).

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 10

    In affinity labeling experiments with [32P]UDP-Glc, an 84 kD polypeptide was found to be

    associated with a plasma membrane fraction containing the highest activity of callose synthase

    (Delmer et al., 1991), and subsequently was identified as sucrose synthase (SuSy). Confirmation

    of plasma membrane association was made immunocytochemically (Amor et al., 1995), and

    Delmer and Amor (1995) proposed that the association of SuSy represented a UDP-Glc delivery

    mechanism to cellulose synthase. β-Glucan microfibrils are synthesized from sucrose and UDP

    on immobilized tobacco plasma membrane sheets (Hirai et al., 1998). More recently, SuSy was

    immunologically associated with CesA proteins in the rosette structures (Fujii et al., 2010),

    strengthening the idea of an association of SuSy directly with cellulose synthases for metabolic

    channeling of glucose through a localized pool of UDP-Glc. However, a quadruple mutant that

    eliminates all detectable SuSy in vegetative tissue does not impair cellulose synthesis (Barratt et

    al., 2009). Overexpression of SuSy in developing vascular tissue of transgenic poplar yields

    small but significant increases in cellulose content (Coleman et al., 2009). Taken together, SuSy

    does not appear to be required for cellulose synthesis but may enhance rates by concentrating

    substrate at the site of synthesis.

    Peng et al. (2002) provided evidence that sitosterol-cellodextrins synthesized from

    sitosterol-β-glycoside serve as primers of glucan chain initiation, with the KORRIGAN

    glucanohydrolase trimming the sitosterol from the growing chain. Debolt and colleagues (2009)

    question this role, as they find double mutants of two major sterol-β-glucoside synthases result in

    severe defects in cuticle formation but not in cellulose synthesis. However, the sterol-glucosides

    are substantially reduced in the double mutant but not entirely eliminated, leaving open the

    question.

    Outside the Golgi stacks, a membrane compartment also containing the KORRIGAN

    (Robert et al., 2005) might represent a dynamic factory associated with both the microtubule

    network and plasma membrane that aligns and directs the cellulose synthase complex, and

    coordinate that function with the deposition of the many polysaccharides directed to it from

    packaged Golgi vesicles. Bimolecular fluorescence complementation techniques, as shown for

    CesA interactions (Desprez et al. 2007) can give important clues to selected participants in

    protein-protein interactions within a complex in vivo. Fluorescence tagging has also allowed

    visualization of the movement of cellulose synthase complexes at the plasma membrane (Paredez

    et al., 2006; Wightman and Turner, 2006; Gutierrez et al., 2009). These studies also established

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 11

    the dynamics of the relationships with the cortical microtubule network in real time. As reviewed

    by Baskin (2001) and Szymanski and Cosgrove (2009), these studies bring resolution to ideas on

    alignment of cortical microtubules and cellulose microfibrils generated long ago through

    observations with inhibitors (Green, 1962) and in the electron microscope (Ledbetter and Porter,

    1963). Improvements in imaging tools are still needed that permit visualization of the delivery

    of Golgi-derived vesicles to the sites of cellulose synthesis much progress has already been made

    along these lines (Held et al., 2008; Konopka and Bednarek, 2008; Crowell et al., 2009).

    The synthesis of non-cellulosic polysaccharides with (1→4)-β-D-glycan backbones

    Because of the same conserved domains of nucleotide-sugar binding and catalysis as those

    encoding CesAs, sub-families of Csl genes were predicted to encode the synthases of non-

    cellulosic polymers with (1→4)-β-D-glycan backbones, primarily (gluco)mannans and

    galacto(gluco)mannans, xyloglucans, glucuronoarabinoxylans (GAX), and the grass-specific

    (1→3),(1→4)-β-D-glucans (Delmer, 1999; Richmond and Somerville, 2000; Hazen et al., 2002).

    For the most part this has turned out to be true, but GAXs are a clear exception. There is still an

    incomplete knowledge of most of the gene products and interactions among them to make

    specific β-D-glycans, even among families where at least one member has a confirmed glycosyl

    transferase activity. There is also a growing disconnection between the classic studies on the

    synthesis of these polysaccharides in vitro and the discovery of genes encoding the machinery

    that warrants a revisit.

    CslF and CslH: Mixed-linkage (1→3),(1→4)-β-D-glucan synthase is the topological

    equivalent of cellulose synthase

    The mixed-linkage (1→3),(1→4)-β-D-glucan is made in the grasses (Poales)(Carpita, 1996;

    Buckeridge et al., 2004), certain lichens (Wood et al., 1994), and Equisetum (Fry et al., 2008;

    Sørensen et al., 2008), but differences in the distribution of their cellodextrin oligomers indicate

    they probably arose by convergent evolution of synthases. For the grasses, this glucan is not a

    random mixture of (1→3)-β-D- and (1→4)-β-D-glucosyl linkages, but is composed primarily of

    cellotriosyl and cellotetraosyl units in a ratio of about 2.5:1 connected by single (1→3)-β-D-

    linkages (Wood et al., 1994). Upon cleavage with a Trichoderma cellulase, smaller amounts of

    higher cellodextrin series are observed with the odd-numbered cellodextrin 2-fold higher in

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 12

    abundance than the next even-numbered unit in the series. The synthesis of the (1→3),(1→4)-β-

    D-glucan has been demonstrated in vitro with isolated, intact Golgi membranes and UDP-Glc

    (Gibeaut and Carpita, 1993; Buckeridge et al., 1999, 2001; Urbanowicz et al., 2004). Whereas

    μM concentrations of UDP-[14C]-glucose result in much shorter oligomers and polymers

    enriched in cellotetraosyl units rather than cellotriosyl units (Buckeridge et al. 1999), much

    larger polysaccharides enriched in cellotriosyl units are observed with higher concentrations of

    substrate (Buckeridge et al., 1999, 2001). The ratios of cellotriosyl : cellotetraosyl and

    cellopentosyl : cellohexaosyl are increased proportionally with substrate concentrations higher

    than 250 μM (Buckeridge et al., 1999), indicating that the mechanism of synthesis of the odd-

    numbered cellodextrin unit is fundamentally different than synthesis of the even-numbered units.

    Proteolysis protection assays show further that the active site of catalysis is on the outward

    facing Golgi membrane (Urbanowicz et al., 2004). Golgi membranes treated with proteinase K

    specifically lost their ability to make the odd-numbered cellodextrin units, whereas the synthesis

    of the cellotetraosyl and higher order even-numbered units was unaffected. Again, loss of the

    ability to make cellotriosyl units is correlated with significant loss in size of the (1→3),(1→4)-β-

    D-glucan product (Urbanowicz et al., 2004). We proposed a similar catalytic dimer model

    wherein even-numbered units are synthesized by core cellulose synthase-like proteins and the

    odd-numbered units arise by an additional glycosyl transferase (GT) that has yet to be identified

    (Buckeridge et al., 2001, 2004).

    Limited proteolysis and detergent reconstitution experiments with the mixed-linkage

    (1→3),(1→4)-β-D-glucan of grass species provides kinetic evidence for three sites of glycosyl

    transfer within the catalytic domain—two from the cellulose synthase-like core domain and a

    third, separable activity (Urbanowicz et al., 2004). (1→3),(1→4)-β-D-Glucan is the topological

    equivalent of cellulose synthase at the Golgi membrane. Limited proteolysis or detergent

    treatment causes loss of the ability to make the diagnostic odd-numbered cellotriose units for

    synthesis, without affecting the ability to generate the even-numbered cellotetraosyl unit

    (Urbanowicz et al., 2004).

    As the topologic equivalent of cellulose synthase at the Golgi membrane, the

    (1→3),(1→4)-β-D-glucan synthase shares another feature with cellulose synthase. When its

    resident membranes are damaged, cellulose synthase (Delmer, 1977), and the (1→3),(1→4)-β-D-

    glucan synthase (Buckeridge et al., 2001) ‘default’ to synthesis of the (1→3)-β-D-glucan,

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 13

    callose, possibly by disruption of the complete active site to a single glycosyl transferase activity

    (Buckeridge et al., 2001; Urbanowicz et al., 2004). Loss of the cellobiosyl generating system to a

    single site of glycosyl transfer would thereafter make only callose (Buckeridge et al., 2001),

    whose synthesis does not require turning the catalytic site or acceptor 180° (Fig. 3B). While it

    can be argued that membrane disruption activates callose synthase in vitro in plasma membrane

    preparations of all angiosperms (Nishamura et al., 2003), only Golgi membranes from grasses–

    the only angiosperms that make the (1→3),(1→4)-β-D-glucan, make callose when damaged

    (Gibeaut and Carpita, 1993; Buckeridge et al., 1999). The most direct evidence for the default

    synthesis of callose from a damaged cellulose synthase comes from the experiments of Blanton

    et al. (2000), who showed that isolated membranes of a cellulose synthase mutant of the cellular

    slime mold Dictyostelium discoideum also lost the ability to make callose in vitro. Whereas

    some cellulose is made in vitro with wild-type membrane preparations, callose linkages

    predominate. Because cellulose synthase is a single gene in Dictyostelium, loss of the ability to

    make (1→3)-β-D-glucan in vitro as well as (1→4)-β-D-glucan in membranes from the mutant

    strongly suggests that the single polypeptide is responsible for both activities.

    Two groups of Csl genes, CslF and CslH, which are found only in grasses (Hazen et al.,

    2002), have been shown to catalyze (1→3),(1→4)-β-D-glucan biosynthesis. Heterologous

    expression of a rice CslF in Arabidopsis, a species that does not make (1→3),(1→4)-β-D-glucan,

    results in small amounts of the β-D-glucan in the cell walls (Burton et al., 2006). However,

    considerably greater amounts of the (1→3),(1→4)-β-D-glucan result when a CslH is co-

    expressed with CslF, suggesting a synergistic role for both CslH and CslF in the synthesis of the

    polysaccharide, and that a catalytic heterodimer enhances the activity (Doblin et al., 2009). If an

    accessory glycosyl transferase is necessary to make the odd-numbered cellodextrin unit, then

    Arabidopsis must produce a related isoform. This finding of concerted action by two distinct

    group members highlights the possibility that synthases of other cross-linking glycans might be

    encoded by Csls of different groups.

    CslA: Mannan and Glucomannan Biosynthesis

    One of the first cell wall polysaccharides to be synthesized in vitro was glucomannan. An early

    conclusion that GDP-Glc is the substrate for (1→4)-β-D- linkages of cellulose, and UDP-Glc is

    the substrate for (1→3)-β-D-glucans (Chambers and Elbein, 1970), had already been disproven.

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 14

    Still today GDP-Glc is still listed erroneously as the substrate for cellulose synthesis on most

    wall charts of biochemical pathways, even though kinetic evidence obtained with cotton fiber

    cells cultured in vitro showed unequivocally that UDP-Glc is the substrate for cellulose synthesis

    (Carpita and Delmer, 1981). Addition of both GDP-Glc and GDP-Man to membrane

    preparations resulted in marked stimulation of incorporation into a glucomannan product (Elbein

    and Hassid, 1966; Piro et al., 1993).

    The knowledge of GDP-nucleotide sugars as substrates were instrumental in the discovery

    that a CslA gene encodes a mannan synthase by expression profiling of guar seed development, a

    species that accumulates large amounts of galactomannan as a cell wall storage carbohydrate

    (Dhugga et al., 2004). Liepman et al. (2005, 2007) confirmed that at least four members of the

    CslA group function in mannan and/or glucomannan synthesis. However, mixed substrates of

    GDP-Glc and GDP-Man in the heterologous expression system (Liepman et al., 2007) do not

    give the marked enhancement of glucomannan synthesis long ago observed in vitro (Elbein and

    Hassid, 1966). While the CesA genes appear orthologous across several species, the Csl genes

    are not (Penning et al., 2009). In fact, the CslA group is resolved into three sub-groups that

    either are Arabidopsis-dominated, grass-dominated, or mixed. The CslA members defined as

    mannan synthase genes (Liepman et al., 2007), fall into both the Arabidopsis-dominated and the

    mixed sub-group (Penning et al., 2009). The functions of these other sub-group members of the

    grasses need to be defined.

    CslC: Xyloglucan Biosynthesis

    Xyloglucans were among the first complex cell wall polysaccharides whose synthesis was

    demonstrated in vitro. Early studies showed that labeled sugars from UDP-Glc and UDP-Xyl are

    incorporated into several polysaccharides using microsomal membranes, and were later refined

    by isolation of Golgi membranes (Ray et al., 1969, 1980). Small amounts of xyloglucan-like

    oligomers with the characteristic α-D-Xyl-(1→6)-D-glucosyl unit, isoprimeverose, are made with

    small amounts of UDP-Glc and UDP-Xyl, but Gordon and Maclachlan (1989) found that when

    concentrations of each nucleotide-sugar are increased to millimolar, large polymers containing

    the characteristic heptasaccharide XXXG (for nomenclature, see Table I) unit structure are

    synthesized. The tetraglucosyl unit of the xyloglucan backbone and the precisely repeated three

    xylosyl units added to make the XXXG structure is consistent with an even-number cellobiose

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 15

    unit synthase reaction for the glucan backbone. Even in the structural variant of Solanaceous

    xyloglucan, where just two xylosyl units are added, a tetraglucosyl unit backbone is preserved by

    the replacement of the third xylosyl group with an acetate (Sims et al., 1996). An apparent

    exception is the ability of certain tree legumes, such as jatobá (Hymenaea courbaril), to make

    XXXXG units in addition to XXXG (Buckeridge et al., 2000). Curiously, partial digestion of this

    polymer with a Trichoderma cellulase, which cleaves only at unbranched positions, yields

    octomer, nonamer, and decamer backbone oligomers whose ratios predict that the polymer

    consists of 4-5-5-4 frameworks separated by variable amounts of XXXG, rather than a random

    distribution of XXXXG and XXXG units (Tiné et al., 2006). This is an intriguing result for two

    reasons: 1) the 4-5-5-4 framework of these types of xyloglucans preserves the even-numbered

    unit symmetry of the backbone, and 2) to make such a framework, as many as 18 glucosyl

    residues might be contained within the complex in order to be ‘read’ properly to preserve the unit

    structure, making the complex much larger than expected.

    The topology of xyloglucan synthesis at the Golgi membrane is still uncertain. Several

    lines of evidence suggest that synthesis relies on transporters for UDP-Glc for backbone

    synthesis of (1→4)-β-D-glucans in vitro (Orellana, 2005), and UDP-GlcA is transported to

    provide UDP-Xyl for transfer within the lumen (Hayashi et al., 1988). Lerouxel et al. (2006)

    propose that synthesis of the backbone is the topological equivalent of cellulose synthase, but the

    addition of all subtending sugars of Xyl, Gal, and Fuc are within the lumen of the Golgi.

    Based on heterologous expression in Pichia, Cocuron et al. (2006) provide evidence that

    CslC genes encode the synthases of the xyloglucan backbone. Although Pichia is unable to

    make UDP-Xyl, co-expression of the xylosyl transferase with CslC is sufficient to induce

    extension of (1→4)-β-D-glucan chains, indicating that a close interaction of these proteins might

    stabilize the synthase to allow extension of the backbone. A complication to unequivocal

    annotation of function of this sub-family is the finding of a CslC at the plasma membrane instead

    of the Golgi membrane (Dwivany et al., 2009). Three members of the GT34 family are

    established as the xylosyl transferases involved in xyloglucan synthesis, and these Golgi resident

    proteins are predicted to face the lumen. (Cavalier et al. 2008; Zabotina et al., 2009)

    Xyloglucan is decorated in various ways in a species-specific manner (Hoffman et al.,

    2005; Peña et al., 2008). Most are primarily galactosylated, with a characteristic α-L-Fuc-

    (1→2)-β-D-Gal-(1→2)-α-D-Xyl- trisaccharide extension (Bauer et al., 1973), but others, such as

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 16

    those of Solanaceous species, have a truncated unit structure substituted with α-L-Ara-(1→2)-α-

    D-Xyl extensions instead of Gal (Sims et al., 1996), and the Asteridae and Oleales species have

    mixtures of these two forms of xyloglucan substitution (Hoffman et al., 2005). Further, the

    xyloglucans synthesized in the Golgi are modified in ways that make them structurally different

    than those that are assembled onto the cellulose microfibrils in the wall (Obel et al., 2009).

    The purification of a xyloglucan-specific fucosyl transferase that led to the discovery of a

    glycosyl transferase family 37 (GT37) gene encoding it (Perrin et al., 1999). While the synthase

    complex must involve a close interaction between the glucan synthases and the xylosyl

    transferases that decorate it (Cavalier et al., 2008), the association of the galactosyl and fucosyl

    transferases might be more transient. Transferases extracted from the membrane are able to add

    Gal from UDP-Gal (Madson et al., 2003) and Fuc from GDP-Fuc (Perrin et al., 1999; Vanzin et

    al., 2002) to exogenous xyloglucan in vitro.

    Xylan Biosynthesis

    The synthesis of grass (1→4)-β-D-xylans from UDP-Xyl with microsomal membranes was

    first demonstrated by Bailey and Hassid (1966). Cooperative action of two nucleotide-sugar

    substrates, in this instance UDP-Xyl and UDP-GlcA, resulted in the synthesis of (1→4)-β-D-

    xylans with subtending GlcA units (Waldron and Brett, 1983; Baydoun et al., 1989). Similar

    studies have shown that membrane preparations from grasses and mixtures of nucleotide sugars

    made GAXs, in part employing a nascent C-4 epimerase that interconverts UDP-Xyl and UDP-

    Ara (Porchia and Scheller, 2000; Porchia et al., 2002; Kuroyama and Tsumuraya, 2001; Zeng et

    al., 2008).

    Apart from a hint that the AtCslD5 may be involved (Bernal et al., 2007) in the synthesis of

    (1→4)-β-D-xylans, it has yet to be directly demonstrated that any Csl gene plays a role. In fact,

    informatics approaches yield non-Csl genes as more likely candidates for encoding the

    machinery for xylan synthesis (Mitchell et al., 2007), and several mutants with deficiencies in

    normal xylan synthesis, such as parvus (Lao et al., 2003; Lee et al., 2007), irx8 (Brown et al.,

    2005), irx7/fra8 (Zhong et al., 2005; Brown et al., 2007), irx9 and irx14 (Brown et al., 2007), and

    irx10 and irx10-L (Brown et al., 2009), do not include a member of the Csl gene family.

    (1→4)-β-D-Glucuronoxylan (GX) synthesis appears to involve a complex initiation

    sequence. Peña et al. (2007) discovered that collapsed xylem mutants deficient in xylan, irx8

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 17

    and fra8, were essentially devoid of a complex tetrasaccharide, β-D-Xyl-(1→3)-α-L-Rha-(1→2)-

    α-D-GalA-(1→4)-D-Xyl, located at the reducing end of the xylan polymer, whereas an irx9

    mutant also severely deficient in xylan contained an overabundance of the tetrasaccharide. The

    presence or absence of this tetrasaccharide greatly affected the size distribution of the xylans. In

    irx8 and fra8 a broader distribution is observed, with some polymers longer than observed in

    wild type. Xylans of irx9 have short chains, with nearly all of them containing the

    tetrasaccharide (Peña et al., 2007). These results suggested a model whereby short chains of

    (1→4)-β-D-xylan are primed by the tetrasaccharide, and these are spliced, cleaving the primer,

    to make the long polysaccharides (York and O’Neill, 2008).

    The IRX8 (GAUT12) and PARVUS (GATL1) encode members of the GT8 of Group C, and

    IRX7/FRA8 of the GT47 Group E, that synthesize the primer tetrasaccharide, whereas IRX9 and

    IRX14 encode members of GT43 that are likely to encode the synthases of the (1→4)-β-D-xylan

    oligomeric backbones that are stitched together by a yet unidentified glycosyl transferase (Brown

    et al., 2007; Peña et al., 2007). GT8 family members encode retaining-type transferases and the

    GT47 members encode inverting-type transferases with respect to the anomeric linkages formed

    compared to the anomeric linkage in the nucleotide sugar, so GT8 members make α-D-linkages,

    whereas the GT47 members make β-D- or α-L- linkages. While an α-D-GalA is found in the

    tetrasaccharide, (1→2)-α-D-GlcA (4-O-Me-GlcA) side groups are also attached at precise

    intervals along the xylan chain (Nishitani and Nevins, 1991), and a general blockwise synthesis

    of six consecutive branched xylosyl residues in grass xylans is observed (Carpita and Whittern,

    1986). IRX10 and IRX10-L appear to encode xylosyl transferases also from GT47, but double

    mutants of these genes have greatly reduced GlcA substitutions along the shorter chains (Brown

    et al., 2009). It will be interesting to determine if addition of these GlcA side-groups play a role

    as an attachment or recognition point where short xylan chains are grafted together to make a

    long chain–a suggestion foretold by the original work on the cooperative action of UDP-GlcA

    and UDP-Xyl in the in vitro synthesis of glucuronoxylan (Waldron and Brett, 1983; Baydoun et

    al., 1989).

    York and O’Neill (2008) take a broader perspective of xylan synthesis and have suggested

    that reducing end addition should not been ruled out. In their model a type of alternating

    ‘pendulum’ mechanism is proposed for the introduction of the (1→4)-β-D-xylan linkages, a

    mechanism analogous to the dimer synthesis described here for (1→4)-β-D- linkages in general.

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 18

    We are just learning the identity of the GlcA and Araf transferases of the GX and GAX polymers

    and the distinctions between the GX devoid of arabinose that is abundant in the secondary xylem

    and the GAX with its rich Ara substitution that is the major polymer of the primary cell walls of

    grasses (Scheller and Ulvskov, 2010). Recently, proteomic approaches of isolated GAX synthase

    complexes from wheat membranes demonstrated a close interaction of GT43 and GT47 family

    members with a GT75 UDP-Ara mutase, an enzyme that interconverts the UDP-arabinopyranose

    and UDP-arabinofuranose conformations required for incorporation of the latter into the

    polysaccharide (Zeng et al., 2010). These kinds of studies represent the path forward to identify

    all the components of synthase complexes whose activities are preserved in vitro.

    Concluding remarks

    While the plant cell wall community is making steady progress on defining biochemical

    functions of backbone synthases and the glycosyl transferases that decorate them, the

    biochemical details of these large, coordinated complexes are still unknown. Fading from

    memory are a great many foundation studies of polysaccharide synthesis in vitro that give

    insights to the differences between an isolated activity and the function of a complex as a whole.

    The major differences between the behaviors of protein complexes in vitro and in vivo are

    compounded by the topology of the synthases and associated glycosyl transferases at the plasma

    membrane and Golgi membranes, across which both pH gradients and electrical potentials are

    generated. Early studies with excised cotton fibers cultured in vitro showed that resealing of the

    plasma membrane was essential to reconstitute cellulose synthesis, and evidence provided that

    regeneration of the electrical potential rather than the pH gradient led to that was critical to

    preservation of synthesis (Carpita and Delmer, 1980). This work was extended to show that

    artificial potentials stimulate (1→3)-β-D-glucan synthesis in vitro (Bacic and Delmer, 1981). In

    Gluconacetobacter, rates of cellulose synthesis could be modulated directly through adjustment

    of the electrical potential in living cells under conditions that did not impair metabolism (Delmer

    et al., 1982). Such gradients exist across other compartments, and, in contrast to cellulose

    synthesis at the plasma membrane, it is maintenance of a pH gradient that prolongs the synthesis

    of the mixed-linkage (1→3),(1→4)-β-D-glucan in vitro in isolated maize Golgi membranes

    (Gibeaut and Carpita, 1993). The physiological and biochemical bases for the effect of these

    potentials and gradients on polysaccharide synthesis are still not understood. They do serve to

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 19

    illustrate that, beyond the biochemical mechanism of synthesis, technologies that preserve not

    only the protein complexes but also their cellular context need to be developed to truly

    understand synthesis of macromolecules across membrane surfaces.

    Acknowledgments

    I thank Maureen McCann (Purdue University) and Peter Ulvskov (University of Copenhagen)

    for their review of the manuscript and their many helpful suggestions. I also thank Daisuke

    Kihara (Purdue University) for his contributions to the discussion on protein structure and

    modeling. The artwork in Figure 6 is by Pamela Burroff-Murr (Purdue University).

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 20

    References

    Amor Y, Haigler CH, Johnson S, Wainscott M, Delmer DP (1995) A membrane-associated

    form of sucrose synthase and its potential role in synthesis of cellulose and callose in

    plants. Proc Natl Acad Sci USA 92: 9353-9357

    Arabidopsis Genome Initiative (2000) Analysis of the genome sequence of the flowering plant

    Arabidopsis thaliana. Nature 408: 796-815

    Arioli T, Peng LC, Betzner AS, Burn J, Wittke W, Herth W, Camilleri C, Höfte H,

    Plazinski J, Birch R, Cork A, Glover J, Redmond J, Williamson RE (1998) Molecular

    analysis of cellulose biosynthesis in Arabidopsis. Science 279: 717-720

    Atanassov II, Pittman JK, Turner SR (2009) Elucidating the mechanisms of assembly and

    subunit interaction of the cellulose synthase complex of Arabidopsis secondary cell walls. J

    Biol Chem 284: 3833-3841

    Bailey RW, Hassid WZ (1966) Xylan synthesis from uridine-diphosphate-D-xylose by

    particulate preparations from immature corncobs. Biochemistry 56: 1586–1593

    Bacic A, Delmer DP (1981) Stimulation of membrane-associated polysaccharide synthetases by

    a membrane potential in developing cotton fibers. Planta 152: 346-351

    Barratt DHP. Derbyshire P, Findlay K, Pike M, Wellner N, Lunn J, Feil R, Simpson C,

    Maule AJ, Smith AM (2009) Normal growth of Arabidopsis requires cytosolic invertase

    but not sucrose synthase. Proc Natl Acad Sci USA 106: 13124-13129

    Baskin TI (2001) On the alignment of cellulose microfibrils by cortical microtubules: a review

    and a model. Protoplasma 215: 150-171

    Bauer WD, Talmadge KW, Keegstra K, Albersheim P (1973) The structure of plant cell

    walls. II. The hemicellulose of the walls of suspension-cultured sycamore cells. Plant

    Physiol 51: 174-187

    Baydoun EA-H, Waldron KW, Brett CT (1989) The interaction of xylosyltransferase and

    glucuronyltransferase involved in glucuronoxylan synthesis in pea (Pisum sativum)

    epicotyls. Biochem J 257: 853-858

    Beeckman T, Przemeck G K, Stamatiou G, Lau R, Terryn N, De Rycke R, Inze D, Berleth

    T (2002) Genetic complexity of cellulose synthase A gene function in Arabidopsis

    embryogenesis. Plant Physiol 130: 1883-1893

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 21

    Bernal AJ, Jensen JK, Harholt J, Sørensen S, Moller I, Blaukopf C, Johansen B, de Lotto

    R, Pauly M, Scheller HV, Willats WG (2007) Disruption of ATCSLD5 results in reduced

    growth, reduced xylan and homogalacturonan synthase activity and altered xylan

    occurrence in Arabidopsis. Plant J 52: 791-802

    Blanton RL, Fuller D, Iranfar N, Grimson MJ, Loomis WF (2000) The cellulose synthase

    gene of Dictyostelium. Proc Natl Acad Sci USA 97: 2391-2396

    Bowling AJ, Brown RM (2008) Cytoplasmic domain of cellulose synthesizing complexes in

    plants. Protoplasma 233: 115-127

    Brown DM, Zeef LAH, Ellis J, Goodacre R, Turner SR (2005) Identification of novel genes

    in Arabidopsis involved in secondary cell wall formation using expression profiling and

    reverse genetics. Plant Cell 17: 2281-2295

    Brown DM, Goubet F, Vicky WWA, Goodacre R, Stephens E, Dupree P, Turner SR (2007)

    Comparison of five xylan synthesis mutants reveals new insight into the mechanisms of

    xylan synthesis. Plant J 52: 1154-1168

    Brown DM, Zhang Z, Stephens E, Dupree P, Turner SR (2009) Characterization of IRX10

    and IRX10-like reveals an essential role in glucuronoxylan biosynthesis in Arabidopsis.

    Plant J 57: 732-746

    Brown RM Jr, Saxena IM (2000) Cellulose biosynthesis: a model for understanding the

    assembly of biopolymers. Plant Physiol Biochem 38: 57-67

    Buckeridge MS, Vergara CE, Carpita NC (1999) Mechanism of synthesis of a cereal mixed-

    linkage (1→3)(1→4)-β-D-glucan: evidence for multiple sites of glucosyl transfer in the

    synthase complex. Plant Physiol 120: 1105-1116

    Buckeridge MS, dos Santos HP, Tiné MAS (2000) Mobilisation of storage cell wall

    polysaccharides in seeds. Plant Physiol Biochem 38: 141-152

    Buckeridge MS, Vergara CE, Carpita NC (2001) Insight into multi-site mechanisms of

    glycosyl transfer in (1→4)-β-D-glycan synthases provided by the cereal mixed-linkage

    (1→3)(1→4)-β-D-glucan synthase. Phytochemistry 57: 1045-1053

    Buckeridge MS, Rayon C, Vergara CE, Carpita NC (2004) Mixed linkage (1→3)(1→4)-β-D-

    glucans of grasses. Cereal Chem 81: 115-127

    Burton RA, Wilson SM, Hrmova M, Harvey AJ, Shirley NJ, Medhurst A, Stone BA,

    Newbigin EJ, Bacic A, Fincher GB (2006) Cellulose synthase-like CSLF genes mediate

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 22

    the synthesis of cell wall (1,3;1,4)-β-D-glucans. Science 311: 1940–1942

    Caño-Delgado A, Penfield S, Smith C, Catley M, Bevan M (2003) Reduced cellulose

    synthesis evokes lignification and defense responses in Arabidopsis thaliana. Plant J 34:

    351-362

    Carpita NC (1996) Structure and biogenesis of the cell walls of grasses. Annu Rev Plant

    Physiol Plant Molec Biol 47: 445-476

    Carpita NC, Delmer DP (1980) Protection of cellulose synthesis in detached cotton fibers by

    polyethylene glycol. Plant Physiol 66: 911-916

    Carpita NC, Delmer DP (1981) Concentration and metabolic turnover of UDP-glucose in

    developing cotton fibers. J Biol Chem 256: 308-315

    Carpita NC, Vergara CE (1998) A recipe for cellulose. Science 279: 672-673

    Carpita NC, Whittern D (1986) A highly substituted glucuronoarabinoxylan from developing

    maize coleoptiles. Carbohydr Res 146: 129-140

    Carpita N, McCann M, Griffing LR (1996) The plant extracellular matrix: News from the

    cell’s frontier. Plant Cell 8: 1451-1463

    Cavalier DM, Lerouxel O, Neumetzler L, Yamauchi K, Reinecke A, Freshour G, Zabotina

    OA, Hahn MG, Burgert I, Pauly M, Raikhel NV, Keegstra K (2008) Disrupting two

    Arabidopsis thaliana xylosyltransferase genes results in plants deficient in xyloglucan, a

    major primary cell wall component. Plant Cell 20: 1519-1537

    Chambers J, Elbein AD (1970) Biosynthesis of glucans in mung bean seedlings – Formation of

    β-(1→4)-glucans from GDP-glucose and β-(1→3)-glucans from UDP-glucose. Arch

    Biochem Biophys 138: 620-631

    Charnock S J Davies G J (1999) Structure of the nucleotidediphospho-sugar transferase SpsA

    from Bacillus subtilis in native and nucleotide-complexed forms. Biochemistry 38: 6380-

    6385

    Charnock S J Henrissat B Davies G J (2001) Three-dimensional structures of UDP-sugar

    glycosyltransferases illuminate the biosynthesis of plant polysaccharides. Plant Physiol

    125: 527-531

    Chen S, Ehrhardt DW, Somerville CR (2010) Mutations of cellulose synthase (CESA1)

    phosphorylation sites modulate anisotropic cell expansion and bidirectional mobility of

    cellulose synthases. Proc Natl Acad Sci USA 107: 17188-17193

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 23

    Cocuron JC, Lerouxel O, Drakakaki G, Alonso AP, Liepman AH, Keegstra K, Raikhel N,

    Wilkerson CG (2007) A gene from the cellulose synthase-like C family encodes a β-1,4

    glucan synthase. Proc Natl Acad Sci USA 104: 8550-8555

    Coleman HD, Yan J, Mansfield SD (2009) Sucrose synthase affects carbon partitioning to

    increase cellulose production and altered cell wall ultrastructure. Proc Natl Acad Sci USA

    106: 13118-13123

    Crowell EF, Bischoff V, Desprez T, Rolland A, Stierhof Y-D, Schumacher K, Gonneau M,

    Vernhettes S (2009) Pausing of Golgi bodies on microtubules regulates secretion of

    cellulose synthase complexes in Arabidopsis. Plant Cell 21: 1141-1154

    Daras G, Rigas S, Penning B, Milioni D, McCann MC, Carpita NC, Fasseas C, Hatzopoulos

    P (2009) The thanatos mutation in Arabidopsis thaliana cellulose synthase3 (AtCesA3) has

    a dominant-negative effect on cellulose synthesis and plant growth. New Phytol 184: 114-

    126

    DeAngelis PL, Weigel PH (1994) Immunochemical confirmation of the primary structure of

    streptococcal hyaluronan synthase and synthesis of high molecular weight product by

    recombinant enzyme. Biochemistry 33: 9033-9039

    Debolt S, Scheible W-R, Schrick K, Auer M, Beisson F, Bischoff V, Bouvier-Navé P,

    Carroll A, Hematy K, Li Y, Milne J, Nair M, Schaller H, Zemla M, Somerville C

    (2009) Mutations in UDP-glucose:sterol glucosyltransferase in Arabidopsis cause

    transparent testa phenotype and suberization defect in seeds. Plant Physiol 151: 78-87

    Delmer DP (1977) Biosynthesis of cellulose and other plant cell wall polysaccharides. Rec Adv

    Phytochem 11: 105-153

    Delmer DP (1999) Cellulose biosynthesis: exciting times for a difficult field of study. Annu Rev

    Plant Physiol Plant Mol Biol 50: 245-276

    Delmer DP, Amor Y (1995) Cellulose biosynthesis. Plant Cell 7: 987-1000

    Delmer DP, Benziman M, Padan E (1982) Requirement for a membrane potential for cellulose

    synthesis in intact cells of Acetobacter xylinum. Proc Natl Acad Sci USA 79: 5282-5286

    Delmer DP, Solomon M, Read SM (1991) Direct labeling with [32P]UDP-glucose for

    identification of a subunit of cotton fiber callose synthase. Plant Physiol 95: 556-563

    Desprez T, Juraniec M, Crowell EF, Jouy H, Pochylova Z, Parcy F, Höfte H, Gonneau M,

    Vernhettes S (2007) Organization of cellulose synthase complexes involved in primary

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 24

    cell wall synthesis in Arabidopsis thaliana. Proc Natl Acad Sci USA 104: 15572-15577

    Dhugga KS, Barreiro R, Whitten B, Stecca K, Hazebroek J, Randhawa GS, Dolan M,

    Kinney AJ, Tomes D, Nichols S, Anderson P (2004) Guar seed β-mannan synthase is a

    member of the cellulose synthase super gene family. Science 303: 363-366

    Doblin MS, Kurek I, Jacob-Wilk D, Delmer DP (2002) Cellulose biosynthesis in plants: from

    genes to rosettes. Plant Cell Physiol 43: 1407-1420

    Doblin MS, Pettolino FA, Wilson SM, Campbell R, Burton RA, Fincher GB, Newbigin E,

    Bacic A (2009) A barley cellulose synthase-like CSLH gene mediates (1,3;1,4)-β-D-glucan

    synthesis in transgenic Arabidopsis. Proc Natl Acad Sci USA 106: 5996-6001

    Dwivany FM, Yulia D, Burton RA, Shirley NJ, Wilson SM, Fincher GB, Bacic A, Newbigin

    E, Doblin MS (2009) The CELLULOSE SYNTHASE-LIKE C (CSLC) family of barley

    includes members that are integral membrane proteins targeted to the plasma membrane.

    Molecular Plant 2: 1025-1039

    Elbein AD, Hassid WZ (1966) The enzymatic synthesis of a glucomannan. Biochem Biophys

    Res Comm 23: 311-318

    Ellis C, Turner JG (2001) The Arabidopsis mutant cev1 has constitutively active jasmonate and

    ethylene signal pathways and enhanced resistance to pathogens. Plant Cell 13: 1025-1033

    Fagard M, Desnos T, Desprez T, Goubet F, Refregier G, Mouille G, McCann M, Rayon C,

    Vernhettes S, Höfte H (2000) PROCUSTE1 encodes a cellulose synthase required for

    normal cell elongation specifically in roots and dark-grown hypocotyls of Arabidopsis.

    Plant Cell 12: 2409-2423

    Freemont PS (2000) Ubiquitination: RING for destruction? Curr Biol 10: R84-R87

    Fry SC, York WS, Albersheim P, Darvill A, Hayashi T, Joseleau JP, Kato Y, Lorences EP,

    Maclachlan GA, McNeil M, Mort AJ, Reid JSG, Seitz HU, Selvendran RR, Voragen

    AGJ, White AR (1993) An unambiguous nomenclature for xyloglucan-derived

    oligosaccharides. Physiol Plant 89: 1-3

    Fry SC, Nesselrode BHWA, Miller JG, Mewburn BR (2008) Mixed-linkage (1 3),(1 4)-β-D-

    glucan is a major hemicellulose of Equisetum (horsetail) cell walls. New Phytol 179: 104-

    115

    Fujii S, Hayashi T, Mizuno K (2010) Sucrose synthase is an integral component of the

    cellulose synthase machinery. Plant Cell Physiol 51: 294-301

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 25

    Gardiner JC, Taylor NG, Turner SR (2003) Control of cellulose synthase complex

    localization in developing xylem. Plant Cell 15: 1740-1748

    Gibeaut DM, Carpita NC (1993) Synthesis of (1→3),(1→4)-β-D-glucan in the Golgi apparatus

    of maize coleoptiles. Proc Natl Acad Sci USA 90: 3850-3854

    Giddings TH, Brower DL, Staehelin LA (1980) Visualization of particle complexes in the

    plasma membrane of Micrasterias denticulata associated with the formation of cellulose

    fibrils in primary and secondary cell walls. J Cell Biol 84: 327-339

    Gordon R, Maclachlan G (1986) Incorporation of UDP-[14C]Glucose into xyloglucan by pea

    membranes. Plant Physiol 91: 373-378

    Green PB (1962) Mechanism for plant cellular morphogenesis. Science 138: 1404-1405

    Guerreiro G, Fugelstad J, Bulone V (2010) What do we really know about cellulose

    biosynthesis in higher plants? J Integr Plant Biol 52: 161-175

    Gutierrez R, Lindeboom JJ, Paredez AR, Emons AMC, Ehrhardt DW (2009) Arabidopsis

    cortical microtubules position cellulose synthase delivery to the plasma membrane and

    interact with cellulose synthase trafficking compartments. Nature Cell Biol 11: 797-808

    Haigler CH, Brown RM Jr (1986) Transport of rosettes from the Golgi apparatus to the plasma

    membrane in isolated mesophyll cells of Zinnia elegans during differentiation to tracheary

    elements in suspension culture. Protoplasma 134: 111-120

    Hayashi T, Koyama T, Matsuda K (1988) Formation of UDP-xylose and xyloglucan in

    soybean Golgi membranes. Plant Physiol 87: 341-345

    Hazen SP, Scott-Craig JS, Walton JD (2002) Cellulose synthase-like genes of rice. Plant

    Physiol 128: 336-340

    Held MA, Boulaflous A, Brandizzi F (2008) Advances in fluorescent protein-based imaging for

    the analysis of plant endomembranes. Plant Physiol 147: 1469-1481

    Herth W (1985) Plasma-membrane rosettes involved in localized wall thickening during xylem

    vessel formation of Lepidium sativum L. Planta 164: 12-21

    Hirai N, Sonobe S, Hayashi T (1998) In situ synthesis of β-glucan microfibrils on tobacco

    plasma membrane sheets. Proc Natl Acad Sci USA 95: 15102-15106

    Hoffman M, Jia ZH, Peña MJ, Cash M, Harper A, Blackburn AR, Darvill A, York WS

    (2005) Structural analysis of xyloglucans in the primary cell walls of plants in the subclass

    Asteridae. Carbohydr Res 340: 1826-1840

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 26

    Hong Z, Delauney AJ, Verma DPS (2001) A cell plate-specific callose synthase and its

    interaction with phragmoplastin. Plant Cell 13: 755-768

    Imai T, Watanabe T, Yui T, Sugiyama J (2003) The directionality of chitin biosynthesis: a

    revisit. Biochem J 374: 755-760

    Jacob-Wilk D, Kurek I, Hogan P, Delmer DP (2006) The cotton fiber zinc-binding domain of

    cellulose synthase A1 from Gossypium hirsutum displays rapid turnover in vitro and in

    vivo. Proc Nat Acad Sci USA 103: 12191-12196

    Jing W, DeAngelis PL (2000) Dissection of the two transferase activities of the Pasteurella

    multocida hyaluronan synthase: two active sites exist in one polypeptide. Glycobiology 10:

    883-889

    Kennedy CJ, Cameron GJ, ˇSturcová A, Apperley DC, Altaner C, Wess TJ, Jarvis MC

    (2007) Microfibril diameter in celery collenchyma cellulose: X-ray scattering and NMR

    evidence. Cellulose 14: 235-246

    Klepek Y-S, Volke M, Konrad KR, Wippel K, Hoth S, Hedrich R, Sauer N (2009)

    Arabidopsis thaliana POLYOL/MONOSACCHARIDE TRANSPORTERS 1 and 2:

    fructose and xylitol/H+ symporters in pollen and young xylem cells. J Exp Bot 61: 537-550

    Konopka CA, Bednarek SY (2008) Variable-angle epifluorescence microscopy: a new way to

    look at protein dynamics in the plant cell cortex. Plant J 53: 186-196

    Koyama M, Helbert W, Imai T, Sugiyama J, Henrissat B (1997) Parallel-up structure

    evidences the molecular directionality during biosynthesis of bacterial cellulose. Proc Natl

    Acad Sci USA 94: 9091-9095

    Kudlicka K, Brown RM (1997) Cellulose and callose biosynthesis in higher plants. 1.

    Solubilization and separation of (1→3)- and (1→4)-β-glucan synthase activities from mung

    bean. Plant Physiol 115: 643-656

    Kurek I, Kawagoe Y, Jacob-Wilk D, Doblin M, Delmer D (2002) Dimerization of cotton fiber

    cellulose synthase catalytic subunits occurs via oxidation of the zinc-binding domains. Proc

    Natl Acad Sci USA 99: 11109-11114

    Kuroyama H, Tsumuraya Y (2001) A xylosyltransferase that synthesizes β-(1→4)-xylans in

    wheat (Triticum aestivum L.) seedlings. Planta 213: 231-240

    Lai-Kee-Him J, Chanzy H, Muller M, Putaux JT, Imai T, Bulone V (2002) In vitro versus in

    vivo cellulose microfibrils from plant primary wall synthases: Structural differences. J Biol

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 27

    Chem 277: 36931-36939

    Lao NT, Long D, Kiang S, Coupland G, Shoue DA, Carpita NC, Kavanagh TA (2003)

    Mutation of a family 8 glycosyl transferase gene alters cell wall carbohydrate composition

    and cause a humidity-sensitive semi-sterile dwarf phenotype in Arabidopsis. Plant Mol

    Biol 53: 647-661

    Ledbetter MC, Porter KR (1963) A “microtubule” in plant cell fine structure. J Cell Biol 19:

    239-250

    Lee CH, Zhong RQ, Richardson EA, Himmelsbach DS, McPhail BT, Ye ZH (2007) The

    PARVUS gene is expressed in cells undergoing secondary wall thickening and is essential

    for glucuronoxylan biosynthesis. Plant Cell Physiol 48: 1659-1672

    Lerouxel O, Cavalier DM, Liepman AH, Keegstra K (2006) Biosynthesis of plant cell wall

    polysaccharides. Curr Opin Plant Biol 9: 621-630

    Liepman AH, Wilkerson CG, Keegstra K (2005) Expression of cellulose synthase-like (Csl)

    genes in insect cells reveals that CslA family members encode mannan synthases. Proc

    Natl Acad Sci USA 102: 2221-2226

    Liepman AH, Nairn CJ, Willats WGT, Sørensen I, Roberts AW, Keegstra K (2007)

    Functional genomic analysis conservation of function among cellulose synthase-like A

    gene family members and suggests diverse roles of mannans in plants. Plant Physiol 143:

    1881-1893

    Madson M, Dunand C, Verma R, Vanzin GF, Caplan J, Li X, Shoue DA, Carpita NC,

    Reiter W-D (2003) The MUR3 gene of Arabidopsis encodes a xyloglucan

    galactosyltransferase that is evolutionarily related to animal exostosins. Plant Cell 15:

    1662-1670

    Maiden MCJ, Davis EO, Baldwin SA, Moore DCM, Henderson PJF (1987) Mammalian and

    bacterial sugar transport proteins are homologous. Nature 325: 641-643

    Mitchell RAC, Dupree P, Shewry PR (2007) A novel bioinformatics approach identifies

    candidate genes for the synthesis and feruloylation of arabinoxylan. Plant Physiol 144: 43-

    53

    Mueller SC, Brown Jr RM (1980) Evidence for an intramembrane component associated with a

    cellulose microfibril synthesizing complex in higher plants. J Cell Biol 84: 315-326

    Nishimura MT, Stein M, Hou B-H, Vogel JP, Edwards H, Somerville SC (2003) Loss of a

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 28

    callose synthase results in salicylic acid-dependent disease resistance. Science 301: 969-

    972

    Nishitani K, Nevins DJ (1991) Glucuronoxylan xylanohydrolase. A unique xylanase with a

    requirement for appendant glucuronosyl units. J Biol Chem 266: 6539-6543

    Nobles DR, Romanovicz DK, Brown Jr RM (2001) Cellulose in Cyanobacter: Origin of

    vascular plant cellulose synthase? Plant Physiol 127: 529-542

    Nühse TS, Stensballe A, Jensen ON, Peck SC (2004) Phosphoproteomics of the Arabidopsis

    plasma membrane and a new phosphorylation site database. Plant Cell 16: 2394-2405

    Obel N, Erben V, Schwartz T, Kuhnel S, Fodor A, Pauly M (2009) Microanalysis of plant

    cell wall polysaccharides. Molecular Plant 2: 922-932

    Orellana A (2005) Biosynthesis of non-cellulosic polysaccharides in the Golgi apparatus.

    Topological considerations. Plant Biosystems 139: 42-45

    Pao SS, Paulsen IT, Saier Jr MH (1998) Major facilitator superfamily. Microbiol Mol Biol

    Rev 62: 1-34

    Paredez AR, Somerville CR, Ehrhardt DW (2006) Visualization of cellulose synthase

    demonstrates functional association with microtubules. Science 312: 1491-1495

    Pear JR, Kawagoe Y, Schreckengost WE, Delmer DP, Stalker D M (1996) Higher plants

    contain homologs of the bacterial celA genes encoding the catalytic subunit of cellulose

    synthase. Proc Natl Acad Sci USA 93: 12637-12642

    Peña MJ, Zhong R, Zhou G-K, Richardson EA, O’Neill MA, Darvill AG, York WS, Ye Z-

    H (2007) Arabidopsis irregular xylem8 and irregular xylem9: Implications for the

    complexity of glucuronoxylan biosynthesis. Plant Cell 19: 549-563

    Peña MJ, Darvill AG, Eberhard S, York WS, O’Neill MA (2008) Moss and liverwort

    xyloglucans contain galacturonic acid and are structurally distinct from the xyloglucans

    synthesized by hornworts and vascular plants. Glycobiology 18: 891-904

    Peng LC, Xiang F, Roberts E, Kawagoe Y, Greve LC, Kreuz K, Delmer DP (2001) The

    experimental herbicide CGA 325'615 inhibits synthesis of crystalline cellulose and causes

    accumulation of non-crystalline β-14-glucan associated with CesA protein. Plant Physiol

    126: 981-992

    Peng L, Kawagoe Y, Hogan P, Delmer D (2002) Sitosterol-β-glucoside as primer for cellulose

    synthesis in plants. Science 295: 147-150

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 29

    Penning B, Hunter CT, Tayengwa R, Eveland AR, Dugard CK, Olek A, Vermerris W,

    Davis M, Thomas SR, Koch KE, McCarty DR, McCann MC, Carpita NC (2009)

    Genetic resources for maize cell wall biology. Plant Physiol 151: 1703-1728

    Perrin RM, DeRocher AE, Bar-Peled M, Zeng W, Norambuena L, Orellana A, Raikhel

    NV, Keegstra K (1999) Xyloglucan fucosyltransferase, an enzyme involved in plant cell

    wall biosynthesis. Science 284: 1976-1979

    Piro G, Zuppa A, Dalessandro G, Northcote DH (1993) Glucomannan synthesis in pea

    epicotyls: The mannose and glucose transferases. Planta 190: 206–220

    Porchia AC, Scheller HV (2000) Arabinoxylan biosynthesis: identification and partial

    characterization of β-1,4-xylosyltransferase from wheat. Physiol Plant 110: 350-356

    Porchia AC, Sørensen SO, Scheller HV (2002) Arabinoxylan biosynthesis in wheat.

    Characterization of arabinosyltransferase activity in Golgi membranes. Plant Physiol 130:

    432-441

    Ray PM, Shininger TL, Ray MM (1969) Isolation of β-glucan synthetase particles from plant

    cells and identification with Golgi membranes. Proc Natl Acad Sci USA 64: 605-612

    Ray PM (1980) Cooperative action of β-glucan synthetase and UDP-xylose xylosyl transferase

    of Golgi membranes in the synthesis of xyloglucan-like polysaccharide. Biochim Biophys

    Acta 629: 431-444

    Reifenberger E, Freidel K, Ciriacy M (1995) Identification of novel HXT genes in

    Saccharomyces cerevisiae reveals the impact of individual hexose transporters on

    glycolytic flux. Mol Microbiol 16: 157-167

    Richmond TA, Somerville CR (2000) The cellulose synthase superfamily. Plant Physiol 124:

    495-498

    Robert S, Bichet A, Grandjean O, Kierzkowski D, Satiat-Jeunemaître B, Pelletier S,

    Hauser M-T, Höfte H, Vernhettes S (2005) An Arabidopsis endo-1,4-β-D-glucanase

    involved in cellulose synthesis undergoes regulated intracellular cycling. Plant Cell 17:

    3378-3389

    Roberts AW, Roberts EM, Delmer DP (2002) Cellulose synthase (CesA) genes in the green

    alga Mesotaenium caldariorum. Eukaryotic Cell 1: 847-855

    Saxena IM, Brown Jr RM (2005) Cellulose biosynthesis: Current views and evolving concepts.

    Ann Bot 96: 9-21

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 30

    Saxena IM, Brown Jr RM, Fevre M, Geremia RA, Henrissat B (1995) Multidomain

    architecture of β-glucosyl transferases: implications for mechanism of action. J Bacteriol

    177: 1419-1424

    Saxena IM, Brown Jr RM, Dandekar T (2001) Structure-function characterization of cellulose

    synthase: relationship to other glycosyltransferases. Phytochemistry 57: 1135-1148

    Scheller HV, Ulvskov P (2010) Hemicelluloses. Annu Rev Plant Biol 61: 263-289

    Sherson SM, Hemmann G, Wallace G, Forbes S, Germain V, Stadler R, Bechtold N, Sauer

    N, Smith SM (2000) Monosaccharide/proton symporter AtSTP1 plays a major role in

    uptake and response of Arabidopsis seeds and seedlings to sugars. Plant J 24: 849-857

    Sims IM, Munro SLA, Currie G, Craik D, Bacic A (1996) Structural characterization of

    xyloglucan secreted by suspension cultured cells of Nicotiana plumbaginifolia. Carbohydr

    Res 293: 147-172

    Somerville C (2006) Cellulose synthesis in higher plants. Annu Rev Cell Devel Biol 22: 53-78

    Sørensen I, Pettolino FA, Wilson SM, Doblin MS, Johansen B, Bacic A, Willats WGT

    (2008) Mixed-linkage (1→3),(1→4)-β-D-glucan is not unique to the Poales and is an

    abundant component of Equisetum arvense cell walls. Plant J 54: 510-521

    Szymanski DB, Cosgrove DJ (2009) Dynamic coordination of cytoskeletal and cell wall

    systems during plant cell morphogenesis. Curr Biol 19: R800-R811

    Taylor NG (2008) Cellulose biosynthesis and deposition in higher plants. New Phytol 178: 239-

    252

    Taylor NG, Laurie S, Turner SR (2001) Multiple cellulose synthase catalytic subunits are

    required for cellulose synthesis in Arabidopsis. Plant Cell 12: 2529-2539

    Taylor NG, Howells RM, Huttly AK, Vickers K, Turner SR (2003) Interactions among three

    distinct CesA proteins essential for cellulose synthesis. Proc Natl Acad Sci USA 100:

    1450-1455

    Timmers J, Vernhettes S, Desprez T, Vincken J-P, Visser RGF, Trindade LM (2009)

    Interactions between membrane-bound cellulose synthases involved in the synthesis of the

    secondary wall. FEBS Lett 583: 978-982

    Tiné MAS, Silva CO, de Lima DU, Carpita NC, Buckeridge MS (2006) Fine structure of a

    mixed-oligomer storage xyloglucan from seeds of Hymenaea courbaril. Carbohydr Polym

    66: 444-454

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 31

    Urbanowicz BR, Rayon C, Carpita NC (2004) Topology of the maize mixed-linkage

    (1→3)(1→4)-β-D-glucan synthase at the Golgi membrane. Plant Physiol 134: 758-768

    Vanzin GF, Madson M, Carpita NC, Raikhel NV, Keegstra K, Reiter WD (2002) The mur2

    mutant of Arabidopsis thaliana lacks fucosylated xyloglucan because of a lesion in

    fucosyltransferase AtFUT1. Proc Natl Acad Sci USA 99: 3340-3345

    Vergara CE, Carpita NC (2001) β-D-Glucan synthases and the CesA gene family: lessons to

    learned from the mixed-linkage (1→3)(1→4)-β-D-glucan synthase. Plant Mol Biol 47:

    145-160

    Waldron KW, Brett CT (1983) A glucuronosyltransferase involved in glucuronoxylan

    synthesis in pea (Pisum sativum) epicotyls. Biochem J 213: 115-122

    Wang J, Howles PA, Cork AH, Birch R, Williamson RE (2006) Chimeric proteins suggest

    that the catalytic and/or C-terminal domains give CesA1 and CesA3 access to their specific

    sites in the cellulose synthase of primary walls. Plant Physiol 142: 685-695

    Wang J, Elliott JE, Williamson RE (2008) Features of the primary wall CESA complex in wild

    type and cellulose deficient mutants of Arabidopsis thaliana. J Exp Bot 59: 2627-2637

    Weigel PH, DeAngelis PL (2007) Hyaluronan synthases: a decade-plus of novel

    glycosyltransferases. J Biol Chem 282: 36777-36781

    Wightman R, Turner SR (2008) The roles of the cytoskeleton during cellulose deposition at the

    secondary wall. Plant J 54: 794-805

    Williams KJ, Halkes KM, Kamerling JP, DeAngelis PL (2006) Critical elements of

    oligosaccharide acceptor substrates for the Pasteurella multocida hyaluronan synthase. J

    Biol Chem 281: 5391-5397

    Wood PJ, Weisz J, Blackwell BA (1994) Structural studies of (1→3)(1→4)-β-D-glucan by 13C-

    nuclear magnetic resonance spectroscopy and by rapid analysis of cellulose-like regions

    using high-performance anion exchange chromatography of oligosaccharides released by

    lichenase. Cereal Chem 71: 301-307

    Yong W, O’Malley R, Link B, Binder B, Bleecker A, Koch KE, McCann MC, McCarty

    DR, Patterson S, Reiter W-D, Staiger CJ, Thomas S, Vermerris W, Carpita NC (2005)

    Plant cell wall genomics. Planta 221: 747-751

    York WS, O’Neill MA (2008) Biochemical control of xylan biosynthesis – which end is up?

    Curr Opin Plant Biol 11: 258-265

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 32

    Yoshida M, Itano N, Yamada Y, Kimata K (2000) In vitro synthesis of hyaluronan by a single

    protein derived from mouse HAS1 gene and characterization of amino acid residues

    essential for activity. J Biol Chem 275: 497-506

    Zabotina OA, van de Ven WTG, Freshour G, Drakakaki G., Cavalier D, Mouille G, Hahn

    MG, Keegstra K, Raikhel NV (2008) Arabidopsis XXT5 gene encodes a putative α-1,6-

    xylosyltransferase that is involved in xyloglucan biosynthesis. Plant J 56: 101-115

    Zeng W, Chatterjee M, Faik A (2008) UDP-xylose-stimulated glucuronyltransferase activity in

    wheat microsomal membranes: Characterization and role in glucurono(arabino)xylan

    biosynthesis. Plant Physiol 147: 78-91

    Zeng W, Jiang N, Nadella R, Killen TL, Nadella V, Faik A (2010) A glucurono(arabino)

    xylan synthase complex from wheat contains members of the GT43, GT47, and GT75

    families and functions cooperatively. Plant Physiol 154: 78-97

    Zhong R, Morrison III WH, Freshour G, Hahn MG, Ye Z-H (2003) Expression of a mutant

    form of cellulose synthase AtCesA7 causes dominant negative effect on cellulose

    biosynthesis. Plant Physiol 132: 786-795

    Zhong R, Peña MJ, Zhou G-K, Nairn CJ, Wood-Jones A, Richardson EA, Morrison III

    WH, Darvill AG, York WS, Ye Z-H (2005) Arabidopsis Fragile Fiber8, which encodes a

    putative glucuronyltransferase, is essential for normal secondary wall synthesis. Plant Cell

    17: 3390-3408

    https://plantphysiol.orgDownloaded on April 2, 2021. - Published by Copyright (c) 2020 American Society of Plant Biologists. All rights reserved.

    https://plantphysiol.org

  • 33

    Table I. Six possible xyloglucan oligomers are produced by digestion with a Trichoderma

    endoglucanase. Because of characteristic side-group construction of these

    complex glucans, a single-letter code was devised to denote the non-reducing

    terminal sugar of the side chain, with other constituents understood (Fry et al.,

    1993). Thus, XXXG signifies a tetraglucosyl backbone with three residues

    bearing a xylosyl side-groups. For many angiosperms, this heptasaccharide is

    decorated further with variable amounts of Gal at the two Xyl residues closer to

    the reducing end, and if Gal is added to the first Xyl residue, a Fuc residue is

    usually added. In addition to the glucan backbone synthase, at least three types of

    glycosyl transferases, and as many as six, are needed to construct all the side-

    gro