Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Chem Biol. Author manuscript; available in PMC 2015 Jan 16.
Published in final edited form as:
PMCID: PMC3955176
NIHMSID: NIHMS555066
PMID: 24439204

The Challenge and Promise of Glycomics

Abstract

Glycomics is a broad and emerging scientific discipline focused on defining the structures and functional roles of glycans in biological systems. The staggering complexity of the glycome, minimally defined as the repertoire of glycans expressed in a cell or organism, has resulted in many challenges that must be overcome; these are being addressed by new advances in mass spectrometry, as well as expansion of genetic and cell biology studies. Conversely, identifying the specific glycan recognition determinants of glycan-binding proteins by employing the new technology of glycan microarrays is providing insights into how glycans function in recognition and signaling within an organism and with microbes and pathogens. The promises of a more complete knowledge of glycomes are immense in that glycan modifications of intracellular and extracellular proteins have critical functions in almost all biological pathways.

Introduction

Glycoconjugates exert their biological functions through complex molecular mechanisms involving both direct glycan recognition, and indirect glycan contributions to conformation and expression of the glycoconjugate. Glycans are directly recognized by glycan-binding proteins (GBPs) (Figure 1). Such interactions of glycans with GBPs can promote cell adhesion, cell-matrix interactions, cellular signaling, glycoprotein folding, and intracellular/extracellular targeting to organelles. In addition, glycans attached to macromolecules exert control through indirect mechanisms on glycoprotein conformation, stability, oligomerization, cell surface resident time, and turnover. Most secreted and membrane proteins are enzymatically glycosylated on one or more amino acids (Apweiler et al., 1999; Steentoft et al., 2013; Van den Steen et al., 1998; Zielinska et al., 2012; Zielinska et al., 2010), and virtually all nuclear and DNA binding proteins, cytoplasmic enzymes involved in metabolic regulation, and some mitochondrial proteins have the O-β-N-acetylglucosamine (O-GlcNAc) modification on Ser/Thr (Bond and Hanover, 2013; Copeland et al., 2013; Hart et al., 2011; Palaniappan et al., 2013; Yi et al., 2012). An emerging paradigm of modern glycomics recognizes that higher animals express a vast repertoire of glycan structures, which comprise their overall glycomes, and recent studies are unexpectedly revealing that glycans contribute in both general and specific ways to almost all biological regulatory pathways. Thus, as might be predicted, abnormalities in glycoconjugate synthesis or turnover are associated with hundreds of different human diseases and disorders, including the Congenital Disorders of Glycosylation (CDGs) and Dystroglycanopathies such as Congenital Muscular Dystrophy (Dennis et al., 2009; Filocamo and Morrone, 2011; Freeze, 2013; Hennet, 2012; Ju et al., 2013; Ohtsubo and Marth, 2006). Developing insights into glycan functions and the complexities of glycan structures and conformations represent both the challenge and promise of glycomics, the field of science now recognized as focused on glycans, just as genomics and proteomics are focused on nucleic acids and proteins, respectively.

An external file that holds a picture, illustration, etc.
Object name is nihms555066f1.jpg

The general roles of glycans in glycoproteins involve both their direct recognition by glycan-binding proteins (GBPs), and indirect effects of glycans on glycoprotein interactions, which may be dependent on protein-protein, protein-lipid, or even carbohydrate-to-carbohydrate interactions. This includes both glycoprotein on the cell surface, cellular organelles, and in secretions, as well as intracellular glycoproteins, e.g. O-GlcNAc glycoprotein. The bottom depicts many of the major classes of glycan linkages to proteins and lipids, along with the symbol key for representing glycan structures with abbreviations of monosaccharides and other substituents.

This article will focus on some of the major challenges and promises of the emerging field of glycomics, both structural and functional. While the term glycomics usually denotes the chemical aspects of glycobiology, we use this term here as a shorthand to denote a broad set of research and knowledge in chemistry and biology of glycans in terms of structure, function, biosynthesis, role in biology and disease, etc. Our major emphasis will be on human and animal systems, but we make a point to acknowledge that glycomics and glycoscience, represent broad areas of knowledge and research encompassing human and animal biology, as well as plant, fungal, and microbial systems.

Glycan and Glycan Binding Protein (GBP) Diversity

Glycans occur as both simple and complex structures in thousands of glycoconjugates, which include glycoproteins, proteoglycans, glycolipids, and as free or un-conjugated glycans. The factors regulating glycan expression and their molecular and functional roles have long been a supremely challenging puzzle. Moreover, our knowledge of the types of glycans and the number of glycan-amino acid linkages is growing, and the nature of the “core structures” of glycans in both glycoproteins, glycolipids, and glycosaminoglycans (GAGs), and glycosylphosphatidylinositol (GPI)-anchored glycoproteins is expanding at an astonishing rate, fueled by development of genomics, proteomics, and mass spectrometry-based tools (Figure 1). Mammalian glycomes are built from 9 common sugars (Glc, GlcNAc, Gal, GalNAc, Man, Fuc, GlcA, Xyl, and NeuAc (sialic acid) with a 10th sugar IdoA being created within pre-synthesized glycosaminoglycans. At least 9 amino acids are known to be glycosylated in nature (Asn, Arg, Ser, Thr, Tyr, Trp, Cys, Hydroxylysine and Hydroxyproline) (Spiro, 2002; Stepper et al., 2011). The surface membrane of cells may contain over 10 million glycans linked to Asn and Ser/Thr residues alone, with the concentration of terminal sugars such as sialic acid approaching 100 mM (Wang et al., 2013). The spatial and temporal organization and functions of all these mammalian glycans is largely unclear. While proteomics is defining protein-protein and protein-nucleic interactions that are estimated to be in the many hundreds of thousands (Garma et al., 2012; Venkatesan et al., 2009), with >70% of proteins encoded in the human genome containing at least one identifiable protein interaction domain (Liu et al., 2012a; Pawson and Warner, 2007), glycomics has not yet advanced enough to estimate the number of protein-glycan interactions, i.e., what can be termed the protein-glycan interactome. While there is a growing list of mammalian GBPs with defined carbohydrate-binding domains and carbohydrate-binding modules (see (Gupta et al., 2012; Taylor and Drickamer, 2011; Varki et al., 2009; Vasta and Ahmed, 2009) and www.cazy.org), many new interactions (Tateno, 2010), especially those involving GAGs, are being discovered and will require further mechanistic insights to generally define protein motifs governing glycan recognition. Because the historical term lectin refers to mainly soluble multivalent proteins capable of agglutinating cells, and lacking enzyme activity, the commonly understood definition of lectin would preclude many membrane bound monovalent proteins, antibodies that bind glycans, toxins, microbial adhesins, enzymes that bind glycans, engineered glycosidases, and GAG-binding proteins. Thus, the more general term GBPs encompasses all of these types of proteins, although the word lectin appears in many databases. We denote in Figure 1 “Direct Glycan Recognition,” where a glycan determinant is directly bound by a GBP, and “Indirect Glycan Effects” that range from glycan-glycan interactions (Handa and Hakomori, 2012), protein solubility, protection from proteolysis and immune surveillance to affecting protein conformation and associations of glycoconjugates in the plasma membrane, for example. Given the growing evidence that the human and individual animal glycomes contain thousands of glycan species (Cummings, 2009), as well as evidence for hundreds of different GBPs encoded in animal genomes (Hileman et al., 1998; Schulenburg et al., 2008; Zhang, 2010) Figure 2, it is possible that the number of protein-glycan interactions may approach that for protein-protein and protein-nucleic interactions, especially considering that many of the GBPs in animals interact with glycans derived from the microbiome, as well as those expressed by the myriad of pathogens that infect these animals (van den Berg et al., 2012).

An external file that holds a picture, illustration, etc.
Object name is nihms555066f2.jpg

Glycan-Related Genes. Compilation of data from several databases identifies glycosylation-related human and mouse genes (glycogenes). Modified and updated from Nairn et al., (Nairn et al., 2008).

Studies on glycan structures and functions are complicated by the facts that most glycans in human and animal tissues remain poorly defined structurally, expression is typically cell type-specific and developmentally- and differentiation-dependent, little is known about the factors governing glycosylation on specific proteins and at specific sites, and virtually nothing is known about the overall architecture and topology of glycan expression on and within cells. However, recent developments in this field have been spurred, not only by the important biological functions of glycans being revealed through genetic studies, but also through breakthroughs in technologies involving mass spectrometry, as well as glycan microarray analyses that demonstrate the wide distribution and binding specificities of GBPs. Together, the genetic information and GBP information are being combined to promote a functional understanding of glycans, which has been termed functional glycomics. While the current picture presented here underscores the tremendous challenges in the field, the landscape of Glycoscience is rapidly changing and the field is at the threshold of significant breakthroughs that will reverberate throughout biology. Such discoveries are highlighting the position of glycans as one of the four major classes of life’s macromolecules, on center stage of modern biomedical and chemical research (Marth, 2008).

Glycan Microarrays and Glycan Determinants

Protein-glycan interactions were historically studied using the hapten-inhibition approach in laborious precipitation or hemagglutination-type assays. Such approaches were far from high-throughput, lacked sensitivity, and required large quantities of valuable glycans, thus limiting binding studies to monosaccharides and short oligosaccharides or poorly defined, complex glycoconjugates isolated from natural sources. This approach led to the prior paradigm that GBPs have relatively low affinity and lack strong specificity. However, within the last decade and beginning with the introduction of ELISA-type assays in the late 1990’s, studies have used microarray technologies in which large libraries of complex glycans are immobilized on surfaces, akin to nucleic acid arrays. These glycan microarrays have helped to revolutionize the field of glycobiology and provided the ability to interrogate putative binding specificities of GBPs by researchers in a wide spectrum of biomedical and basic sciences, in a fashion unimaginable a decade ago. Glycan microarrays can be interrogated by indirect or direct fluorescence-based techniques to identify specific binding of glycans by lectins, antibodies, toxins, viruses, etc. (Figure 3).

An external file that holds a picture, illustration, etc.
Object name is nihms555066f3.jpg

To explore protein-glycan interactions one major technology in the field is glycan microarrays. In such an approach, individual glycans from natural sources or chemo/enzymatic syntheses, are modified to allow their automated printing and attachment, either covalently or non-covalently, to a slide-type surface in defined positions, akin to a gene array. The glycan microarray can be interrogated with a GBPs or other reagents or even cells and viruses, to identify those glycan “spots” that are recognized, which can be visualized by either direct or indirect fluorescent tagging. The spot pattern on the image, which should also incorporate replicates of each glycans, can be averaged to generate a histogram. In the example shown, the results indicate that that the GBPs in question bound strongly to one glycan, less strongly to another glycan, and did not bind appreciably to any other glycan. Such microarrays can also be prepared from glycolipids, glycopeptides, whole glycoproteins, or polysaccharides of animal, plant, or microbial origins.

The Consortium for Functional Glycomics (CFG), developed through a Large Scale Collaborative Project from NIH/NIGMS, made available a printed glycan microarray (Blixt et al., 2004) that continues to be available to the research community (Rillahan and Paulson, 2011; Smith and Cummings, 2013), and data generated is posted and publicly available (functionalglycomics.org/). Over 15,000 glycan microarrays or slides from the CFG have been used over the past decade by researchers in ~3,000 array experiments with >1,000 types of samples containing GBPs, including human and animal sera, as well as intact eukaryotic and prokaryotic cells. The database of the CFG is the world’s largest catalog of protein-glycan interactions. The work of the CFG is complementary to those of many other laboratories world-wide involved in developing glycan microarrays and probing them with a variety of GBPs and other reagents (Chevolot, 2012; de Paz and Seeberger, 2012; Liu et al., 2009; Park et al., 2013). Smaller libraries of glycans have also been usefully explored for binding to GBPs using solution-based approaches (Arata et al., 2001; Chang et al., 2011; Duverger et al., 2010; Gutierrez Gallego et al., 2004; Takeda et al., 2013). However, a need in the future is for all users of glycan microarrays to deposit their data in publicly-available databases for curation and cataloging in relation to the larger effort to define the protein-glycan interactome.

The results of glycan microarray studies using different platforms and different types of glycans are revealing that many GBPs including antibodies display relatively high affinity interactions with complex glycans, often involving 3–7 monosaccharide residues, that includes branched structures or saccharide modifications, such as sulfation or phosphorylation. Researchers are moving away from a focus on the minimal determinants that might inhibit a lectin binding to a single glycan, such as millimolar concentrations of monosaccharides or simple glycans, to the concept of the biological and physiological interactions between relevant naturally occurring GBPs and relatively complex glycan structures (Wang et al., 2013). The concept of glycan determinants (Cummings, 2009), which are the minimal glycan structures that confer the maximum glycan binding affinity, may be thought of as akin to antibody epitopes or glycotopes (Cao et al., 1996). This concept helps us to understand the contributions of individual glycan features and biosynthetic pathways to glycan recognition. However, such array studies naturally raise questions as to their physiological relevance, and there are concerns of glycan “presentation” on microarrays and the degree to which this presentation is biologically relevant to their presentation on cell surfaces or native glycoconjugates (Park et al., 2013).

It should be self-evident that binding or lack of binding of GBPs to glycans on glycan microarrays or other surfaces does not directly provide evidence of functionally significant interactions, but such results can provide hypotheses to be tested regarding particular glycan functions. In many cases, ligands predicted by glycan microarray analyses have been shown to have physiological relevance, whereas in some cases the connection is less clear. Examples of the former are results of virus binding, such as influenza viruses and noroviruses, which bind in microarrays and in vivo to sialylated (Blixt et al., 2004; Stevens et al., 2006; Walther et al., 2013) and blood group H-type glycans (Lindesmith et al., 2003; Shang et al., 2013), respectively. Another example of concordance of glycan microarray data and physiological functions are studies on CD22 (Siglec-2), which was shown to bind glycans with NeuAcα2-6-linked sialic acid on cells in vitro (Macauley et al., 2013; Powell et al., 1993; Sgroi et al., 1993), in solution studies (Powell and Varki, 1994), in glycan microarrays (Campanero-Rhodes et al., 2006; Tateno et al., 2008), and in mice (Hennet et al., 1998). A possible example of discordance between glycan microarray data and physiological ligands involves recent studies on Siglec-F in mice. Both human Siglec-8 and its murine homolog Siglec-F preferentially recognize the same sialylated glycan 6′-sulfo-sialyl Lewis X (6′su-SLex - NeuAcα2-3[6-SO3]Galβ1-4[Fucα1-3]GlcNAc-R), in part based on binding to immobilized arrayed glycans (Bochner et al., 2005; Kiwamoto et al., 2013b; Tateno et al., 2005). This Siglec is expressed in eosinophils and its binding to 6′su-SLex has been proposed to be important in mitigating allergic eosinophilic airway inflammation. Studies in vivo indicate, however, that while St3gal3 mutants lacking the sialyltransferase had diminished Siglec-F binding and more intense allergic eosinophilic airway inflammation (Kiwamoto et al., 2013a), deletion of the sulfotransferases capable of generating 6-sulfated galactose, as determined by glycomic analysis, does not appreciably affect Siglec-F binding in a mouse model (Patnode et al., 2013). Thus, in all cases where it is experimentally feasible, it is critical to link the results of glycan microarray analyses for GBPs glycan determinants with physiological evidence for such glycan determinants being functionally important. It is also important to note that current glycan microarrays lack the full presentation of structures found in the human and animal glycomes, thus many potentially important glycans and glycan determinants are lacking, and thus lack of binding of a protein to a glycan microarray may simply indicate that the relevant glycan ligand is missing.

While the number of structurally different glycans in the human glycome is presently unknown, it is likely to be many tens of thousands. The repertoire of glycan determinants in the human glycome, however, is estimable. An approach to this problem is to conceptually assemble partial determinants or segments of glycan structures with each other in biochemically allowed and defined ways based on known biosynthetic pathways and glycan structures (Cummings, 2009; Rademacher and Paulson, 2012; Werz et al., 2007). Such a combinatorial approach permits calculations as to how many such determinants might exist in N- and O-glycans and glycolipids, as well as glycosaminoglycans (GAGs). An illustration using partial determinants shows that that there are possibly many thousands of glycan determinants (Figure 4). Few of these glycan determinants are currently available on available microarrays (the current CFG glycan microarray has 611 different glycans); thus, there is a tremendous need for chemical and enzymatic syntheses of thousands of compounds. Fortunately, the NIH has funded the development of small libraries of synthetic glycans as a first step in expanding the availability of glycans to the research community (sbir.cancer.gov/funding/contracts/fy2013_09.asp). Coupled with advances in asymmetrical synthesis of glycans (Wang et al., 2013), these efforts should be encouraged since a large number of glycans are obviously needed for structural and functional studies and as standards for nuclear magnetic resonance and mass spectrometry (MS) analyses. In any case, the need for increased number of synthetic glycans in sufficient quantities to impact glycan analysis and functional studies far exceeds their availability.

An external file that holds a picture, illustration, etc.
Object name is nihms555066f4.jpg

GBPs often recognize 2–6 linear or branched monosaccharides with additional modifications, such as phosphorylation, sulfation, O-acetylation, etc. Such recognition can be termed a glycan determinant, which is the minimal glycan structure that confers maximum binding. These glycans determinants can be thought of as being assembled from partial determinants, that in the example shown are modified galactose, modified Gal-GlcNAc, and modified GlcNAc-Man. With the partial determinants shown, which are respectively, 6, 9, and 3 in number, it is possible to assemble these into 162 different glycan determinants. Shown also are two of these as Determinant 1 and 2, which are differently recognized by a GBPs or GRM. If one considers all such known partial determinants in human glycans defined to date, it is possible to predict that there are over 5,000 glycan determinants; if the GAG sequences are also included up to pentasaccharides, then there are an additional 10,000 or so. It is likely that this is an underestimate for the total theoretical glycan determinants, since it is likely that other partial determinants will be identified in the future. In addition, it is possible that in a single branched or linear glycan, the one set of glycan determinants may attenuate the recognition of the same or a different glycan determinant on the same molecule.

Milk and Glycolipid “Metaglycomes” as a Paradigm for Glycomics

While the human glycome certainly represents a challenge in terms of defining all its structures, the meta-glycomes of cells and tissues may be amenable to total analysis. We use the term, “metaglycome,” as denoting a constituent glycome of a specific cell or tissue, as well as a type or family of glyconjugates; summing all the human metaglycomes would then define the human glycome. In fact, targeting specific metaglycomes and providing their repertoire of glycans would likely constitute a milestone toward the direction of eventually defining the human glycome. Two such metaglycomes that would be amenable to such analyses are the soluble human milk oligosaccharides or glycans and human glycosphingolipids. The sizes of both of these metaglycomes have already been estimated, with the HMO comprising several hundred glycans (Bode, 2012; Urashima et al., 2011), and the human glycosphingolipid metaglycome comprising at least 500 different glycosylated neutral and acid species, based on the Lipid Maps database (www.lipidmaps.org/), without consideration of the aglycone lipid constituents. Importantly, technologies appear in place to allow quantitative and qualitative descriptions of both the free glycans in human milk (Bao et al., 2013; Ninonuevo et al., 2006) and releasable glycans from glycosphingolipid glycomes (Fujitani et al., 2011). Interestingly, both of these metaglycomes have terminal Glc as the non-reducing sugar (lactose-type glycans in milk and lactosylceramide-type glycans in glycosphingolipids), so technologies for defining the glycans may be overlapping and synergistic. Recent technological advances indicate that glycosphingolipid metaglycomes may be approached using intact lipids, maintaining information regarding aglycone dynamics as well ((Boccuto et al., 2013)

A common conundrum for the glycomics field is: how will researchers know that they have fully defined a metaglycome? The evidence that a metaglycome is largely, if not fully, defined would be based on the pace of discovery of new glycan species within such metaglycomes in the future; thus, at some point in time the pace of discovery would lessen to the point that one could estimate that >95% of the glycans below a certain reasonable size limit, e.g. 5,000 daltons, have been described for that metaglycome. The definitive description of a metaglycome to that level, with calculated and statistical reasoning, would go far in assisting the development of technologies and bioinformatic approaches that will provide researchers with their first breakthrough paradigm in defining at least two specific component metaglycomes of the human glycome. A clear challenge to the field of glycomics is that if we cannot fully define the HMO and glycosphingolipid metaglycomes, how can we possibly imagine defining the glycoprotein-derived metaglycomes, which are clearly much more complex?

The complexity of glycomes parallels the technical difficulties in analyses, as well as the number of glycan determinants that might exist (Figure 5). Thus, one could imagine that GAGs, with their multitude of potential disaccharide repeating units, would represent the most complex set of glycans in the human glycome, while GPI-anchors and human milk oligosaccharides, may represent the relatively least complex set. Thus, future studies in glycomics should focus to some degree on the metaglycomes that are most definable, and functional glycomics can then define the recognition of those glycans by GBPs, or the specific roles of those glycans in human physiology and disease. Such a strategy could lead to the identification of the protein-glycan interactome for that particular metaglycome.

An external file that holds a picture, illustration, etc.
Object name is nihms555066f5.jpg

The different classes of glycans, from the GPI glycans within GPI-anchored glycoproteins, to the GAGs, represent different levels of complexity and numbers and diversity of glycan determinants. Of course with complexity and increased numbers of determinants, the analytical difficulties expand tremendously. Thus, a great challenge of glycomic technologies is to difficult better methods of preparing and analyzing glycans to overcome the challenges of their complexities.

In the absence of either knowing all the structures in the human glycome, or having all glycan determinants available as synthetic compounds, it is possible to use a shotgun glycomics approach, prior to definitive structural characterization, to obtain all the cellular glycans and use fractionated glycan species for functional studies on glycan microarrays and other surfaces. Such a shotgun approach has been successful in many cases for preparing total glycosphingolipid-derived glycan libraries and other types of natural glycans (Liu et al., 2012b; Song et al., 2011; van Diepen et al., 2012). The combined approach of general glycomics, the release of all the glycans from endogenous glycoconjugates and subsequent structural analysis, with functional glycomics, where individual purified glycans species are probed for their functional recognition by GBPs, is highly likely to make significant strides to unravel both the mysteries about the glycan structures that make up the glycome and their recognition by endogenous and pathogen-derived GBPs.

Glycan Recognition Molecules (GRM)

Over a hundred different plant lectins and a few from invertebrates have been identified, and many, such as Concanavalin A (Jack bean) and Helix pomatia agglutinin (snail), have been utilized successfully to study glycan expression and function. The use of glycan microarrays has allowed the binding determinants of a large number of lectins to be explored, thereby greatly facilitating their utility in testing hypotheses regarding glycan function. The repertoire of human and mouse glycan structures recognized by the plant lectins commonly available is quite limited, however, and often a single lectin binds to multiple glycan structures carrying minimal glycan determinants. In addition, analysis of the binding specificities of commercial plant lectins by the glycan microarray screening of the CFG has demonstrated remarkable inconsistencies among commercial preparations of the same lectin, including variable binding affinities and even inactive preparations. The field of Glycomics is at a point where more reliable and better-defined reagents that recognize a breadth of glycan structures are required. Moreover, if investigators encompassing the breadth of biomedical research had available these types of reliable reagents, a significant increase in our knowledge of glycan function would, no doubt, occur. The time is right to consider developing an extensive library of glycan recognition molecules (GRM) with specificities and determinants defined by binding to expansive glycan microarrays. These reagents, if made available to researchers with diverse health-related interests, would greatly accelerate the pace of glycan function discovery in physiology and disease. These reagents could also provide a direct means to target specific glycans for potential therapeutic and diagnostic applications. There are novel platform technologies in the early stages of development that can generate and select for molecules that recognize and bind specific glycan structures. These include antibodies or antibody-like proteins, as well as non-protein molecules such as nucleic acid aptamers (Li et al., 2008) and synthetic lectins (Ke et al., 2012). Since some of these glycan recognition molecules are not proteins, here we refer to this class as GRM, instead of GBP. Potential technologies for generation and selection of antibodies with specificity for specific glycan epitopes include engineering of glycans, glycopeptides, and novel adjuvants to elicit high-affinity IgG (Lakshminarayanan et al., 2012), yeast display of scFv antibodies (Zhao et al., 2011), and lamprey antibodies (Yu et al., 2012). In addition, the elucidation of glycan binding specificities of viruses, bacteria, both plant and animal lectins using glycan microarrays, as well as structural knowledge of these GBP plus many glycan-recognizing enzymes, provide a structural basis of glycan binding motifs. Knowledge of these motifs, coupled with detailed structural studies of glycan binding domains by crystallography and NMR spectroscopy, as well as molecular simulations of glycan-protein interactions, have led to an expanding understanding of how to engineer glycan recognition determinants, e.g., (Feinberg et al., 2013; Ford et al., 2003; Mercer et al., 2013). This knowledge could be exploited to assist the design and selection of GRM with glycan binding determinants and specificities of interest.

The prospect for the availability of libraries of GRM is somewhat analogous to the availability of antibodies directed against specific phospho-peptides for studies of cell signaling pathways, and the development of restriction enzymes for molecular biology. Availability of these sequence-specific reagents has allowed the investigation of signaling pathways and nucleic acids by non-expert investigators with diverse interests. Likewise, GRM would assist in testing for changes of glycan epitope expression and would also provide new tools for assessing functions by potentially blocking or crosslinking target glycans. Production of GRM will involve much more complex challenges than for peptide or phospho-peptide specific antibodies or restriction enzyme identification because of the diversity of glycan structures, and their precise binding determinants must be determined using glycan microarrays and other approaches, akin to what was done for the Cluster of Differentiation (CD) in immunology (www.sciencegateway.org/resources/prow/). Despite the obvious challenges, the availability of defined GRMs to pursue specific biological questions will likely be transformative.

Glycomic Analyses of Various Types of Glycoconjugates and Their Recognition by GBP

A clear challenge of glycomics is the complexity of analytical methods for individual classes of glycans, as discussed here. Each class requires different methods of extraction or glycan release, often different methods of analysis, or combinations of methods, such as HPLC, mass spectrometry, NMR, etc (Figure 6). Nevertheless, while complex, each of these methods has been adapted to provide deep insights into glycan structure for that particular glycan class. An all encompassing method of glycoprotein analysis at the top-down level would be ideal, and coupled with the bottom-up approaches below (Hanisch, 2012; Nicolardi et al., 2013), would help to define the relationships of glycans to their protein and lipid carriers and their relatively abundance to each other. Moreover, analytical approaches capable of probing the glycomes of living cells, as through biorthogonal methods (Belardi et al., 2012) or non-invasive magnetic resonance approaches, would provide breakthroughs into the dynamical aspects of glycomes in living tissues.

An external file that holds a picture, illustration, etc.
Object name is nihms555066f6.jpg

The general strategies for glycomic analyses typically involves isolating or generating free glycans from glycoproteins/proteoglycans and glycolipids. The obtained mixture of glycans can then be derivatized and directly analyzed by MS, or derivatized, separated by HPLC and other approaches, and further analyzed by MS or NMR. In shotgun glycomics for functional studies, the released glycans after separation can be printed to generate glycan microarrays. For site-specific glycosylation and identification of protein carriers, glycopeptides can be generated by proteolysis and then analyzed directly before or after glycan removal.

Glycosphingolipids

The majority of the current methodologies for preparing glycans beginning with cell/tissue extraction are depicted in Figure 6. Obviously, the challenges for structural determination for the glycans on these diverse classes of glycoconjugates (GSL) are formidable and call for more facile technologies with automated annotation. Beginning the discussion with GSL, sophisticated analytical platforms have been developed for their separation and resolution, ranging from thin-layer chromatography to HPLC (Levery, 2005; Suzuki et al., 2011). In many cases, the glycan moieties are released and analyzed separately from the ceramide aglycone, although the accuracy and speed of current mass spectrometers are facilitating the analysis of intact glycosphingolipids. In either case, MS is then used to characterize the glycans; MSn analyses obviously yield more-definitive structural assignments. Two new methodologies rely on either prior permethylation of the GSL after initial extraction, allowing LC-MSn or direct infusion MSn to resolve each permethylated species to quantify individual GSL with detailed structural information of both glycan and lipid moieties of each species (Boccuto et al., 2013) or the use of MALDI-quadrupole ion trap-TOF in negative ion mode (Ito et al., 2009). Neural tissue contains the most complex GSL, but they are expressed and have various functions in all cells. Some of these functions serve to regulate outside-in signaling pathways, particularly by modulating cell surface receptor association with other membrane proteins that can, in turn, affect efficacy of signaling (Figure 1). For example, the signaling of the insulin receptor can be strongly attenuated when it is in association with GM3; the levels of this ganglioside increase in response to inflammatory cytokines such as TNF-α, resulting in loss of insulin receptor function (Inokuchi, 2010; Kabayama et al., 2007). There are also many examples of direct glycan recognition (Figure 1) between GBP and GSL, as well, such as the binding of myelin-associated glycoprotein (Siglec 4) to GD1a and GT1b. Moreover, pathogens, such as Vibrio cholera, Shigella dysenteriae, and Clostridium tetani, all secrete toxins that bind specific GSL on cell surfaces. A discussion of microbial proteins that recognize cell surface GSL and other types of glycans can be found in Essentials of Glycobiology, Chapter 28 and other sources (Day et al., 2012; Karlsson, 1989; Topin et al., 2013).

N-linked glycoprotein glycan analysis and recognition by GBP

Asparagine-linked glycans (N-glycans) have been extensively studied in terms of their structures and precise functions (Moremen et al., 2012). Most of the endogenous GBPs for humans and mice whose ligand specificities are known bind to N-glycans (see http://www.imperial.ac.uk/research/animallectins/ctld/lectins.html and http://www.functionalglycomics.org/glycomics/publicdata/primaryscreen.jsp. Due to their larger size and extensive branching, the N-glycans are the most complicated non-glycosaminoglycan structures to analyze, since for example, a tetraantennary glycan can have heterogeneous branches that are even isobaric. These individual branches are difficult to resolve using any number of separation and MS-based platforms of analysis, although progress is being made (Yu et al., 2013a). The analysis of N-glycans, whereby they are released from glycoproteins or glycopeptides using PNGase F or chemical means such as hydrazinolysis, is depicted in Figure 6. Recombinant PNGase F releases all N-glycans found in humans and mice, leaving an Asp acid rather than Asn at the site of attachment to the protein and producing free glycan with an intact reducing terminus. Several methods have been developed to use this enzyme to incorporate 18O into the Asp where a glycan was attached (Kuster and Mann, 1999). Subsequent proteolysis can then produce peptides, some of which have incorporated 18O, that allow the identification of the specific peptides that had an N-glycan attached. Large scale analysis of N-glycans in tissues and serum has been performed by MSn using different techniques to enrich for this class of structures. One study utilized four mouse tissues and serum to identify 6,367 N-glycosylation sites on 2,352 proteins; over 99% of these contained the canonical Asn-X-Ser/Thr motif (where X is not Pro), however, a small fraction of these sites actually utilized the Asn-X-Cys motif (Zielinska et al., 2010).

Identification of released N-glycan structures has been pursued by many investigators and methods, reviewed in (Han and Costello, 2013; North et al., 2009). It is useful to distinguish between methods that are facile for discovery of novel glycans or discriminate between isobaric structures, in contrast to those that are optimal for detecting and quantifying structures in a more routine manner with potential for high-throughput analysis. One of the classic techniques used mainly for quantifying mixtures of N-glycans, after release from its protein or peptide, is tagging of the reducing termini of glycans with a fluorescent compound and separating the various species via HPLC or capillary electrophoresis (Doherty et al., 2012). This methodology has been very successfully scaled to high-throughput and automated analysis, allowing large populations of samples to be analyzed, although, since it is based mainly on retention time differences determined for standards whose structures have been demonstrated by a variety of means, glycans with unusual or unexpected structures may not be resolved. Glycosidase digestions can often aid in structural assignment. A recent study profiling the N-linked structures detected on the major serum glycoproteins of a large population, coupled with a Genome Wide Association Study (GWAS) of the genomes of those whose sera were analyzed, suggested that the transcription factor HNF1α is involved in regulating fucosylation of N-glycans on serum glycoproteins (Thanabalasingham et al., 2013). Fluorescent tagging of glycans still yields one of the lowest thresholds of glycan detection compared to other methods, including various types of mass spectrometry.

Analyses of released N-glycans by mass spectrometric methods range from simple MALDI-TOF to those that employ permethylation, followed by purification and MSn techniques. Permethylation assists in extraction of glycans from peptides and other contaminants, neutralizes negative charges, amplifies differences in masses between somewhat similar species, and is able to decrease variability between various molecular ions to be detected by the mass spectrometer. In addition, fragmentation of permethylated glycans produces scars at previous sites of glycosidic linkage, providing additional structural information. Obviously, because of the many isobaric structures found in mixtures of N-glycans, definitive characterization requires fragmentation and extensive analysis (see MIRAGE, below). Quantitation can be done by the total ion mapping method (Aoki et al., 2007), but other methods, also based on those developed for proteomics, have been developed (North et al., 2009; Orlando, 2010). Attempts are being made to automate some of these types of analyses, which will significantly accelerate throughput and move glycan analysis from a specialty to a more common practice.

N-glycans are essential in the quality-control of glycoprotein folding and the intracellular trafficking of glycoproteins (Aebi, 2013; Braakman and Bulleid, 2011). Many of the endogenous GBP discovered thus far in humans and mice bind to N-glycans and function in these pathways. The first description of an animal GBP, (Fig. 1 Type I), the “Ashwell-Morell receptor”, also termed the hepatic asialoglycoprotein receptor, and founding member of the C-type lectins, is expressed on hepatocytes and binds to N-glycans on glycoproteins when they have terminal galactose, due to exposure to neuraminidase (Hudgin et al., 1974). Recent studies have demonstrated that this receptor shows selectivity to ligands that are exposed on desialylated glycoproteins involved in prothrombosis induced by bacterial infection (Grewal et al., 2008). N-glycans also have prominent functions in the Type II, indirect glycan effects in regulating glycoprotein function, regulating glucose transporter (Ohtsubo et al., 2005), receptor half-life on the cell surface (Partridge et al., 2004), and modulating the effector functions of the Fc moiety of IgG. Evidence has been presented indicating that IVIG (intravenous immunoglobulin) is therapeutically anti-inflammatory (Nimmerjahn and Ravetch, 2008) and that α2-6-linked sialic acid on N-glycans of IgG are immunosuppressive (Anthony et al., 2008a; Kaneko et al., 2006) and may bind to DC-SIGN and murine SIGN-R1 (Anthony et al., 2008b; Schwab et al., 2012) to signal the downstream expression of immunosuppressive cytokines and receptors, but more recent studies question some of these prior interpretations (Yu et al., 2013b).

O-glycan analysis and recognition by GBP

O-glycans in humans and mice range from the shockingly ubiquitous O-GlcNAc modification of Ser/Thr found in intracellular and nuclear proteins (thereby qualifying them to be considered as glycoproteins) (Copeland et al., 2013), to the large family of proteins that contain Ser/Thr-O-GalNAc that is normally extended by other glycans and glycan modifications. Almost all proteins with a predicted signal sequence are also predicted to have one or more Ser/Thr-O-GalNAc modifications (Steentoft et al., 2013). Other O-linked glycan modifications include O-Fuc, O-Glc, and O-Man; in fact, one-third of O-linked protein modifications in rat brain has been shown to be O-Man (Chai et al., 1999). Compared to N-glycans, only a few GBP have thus far been shown to bind O-glycans, suggesting that there are many more of these GBP to be discovered. The most prominent of these GBP is P-selectin, which is expressed on activated endothelia and functions in the first step of the inflammatory response involving cell adhesion (Wilkins et al., 1995). Of particular note is the recent study showing a unique O-Man-linked glycan with an unusual repeating polymer of xylose and glucuronate that serves as the ligand for laminin found in basement membranes (Inamori et al., 2013; Yoshida-Moriguchi et al., 2010). This laminin-glycan interaction is responsible for adhesion between many cell types, including muscle and nerve, epithelia and basal lamina, and is affected in cancer cells, as well as in several of the Congenital Disorders of Glycosylation (Freeze, 2013) that affect muscle and nerve function, e.g., (Yang et al., 2013). This result also demonstrates that novel glycan structures are still being discovered and that O-linked glycans can clearly function as ligands for endogenous GBP. The cadherin family of proteins has also recently been shown to express significant levels of O-Man-containing glycans (Vester-Christensen et al., 2013).

Because of lack of an equivalent of a pan-specific PNGase for O-linked glycans, structures attached to Ser/Thr residues are typically released by beta-elimination via mild base/borohydride or other basic reagents. After release, their separation and quantitation are accomplished by methods similar to N-glycans. There are examples of proteins that have over 1,000 O-GalNAc glycans, while others have only a single O-linked GalNAc modification. The diversity of O-GalNAc glycans in a single mucin is staggering when modifications such as sulfation are taken into account; for example, human Muc5ac from patients with cystic fibrosis contains over 260 distinct glycan structures, determined by release, fluorescent tagging, HPLC separation and MS analysis (Xia et al., 2005). Glycosylation of mucins is often cell- and tissue-specific and altered in disease states (Larsson et al., 2011). The factors that determine selectivity of particular O-GalNAc modifications are only beginning to be understood (Steentoft et al., 2013). The generation of technologies to generate cultured cells that lack particular glycosyltransferases represents a significant step toward our understanding the mechanisms that regulate this diversity of glycan structures. Application of MS methods that select specific molecular ions that when fragmented serve as signatures for particular O-glycan structures, usually employing tandem MS, can quantify O-glycans in a mixture released from an isolated glycoprotein, serum or cell-derived glycoproteins. This method of selected reaction monitoring and consecutive reaction monitoring or their variants offers the possibility of automation and relatively high-throughput of O-glycan analysis, and can be used to identify specific glycoprotein glycoforms, e.g., (Sanda et al., 2013; Zhang et al., 2012).

Glycosaminoglycan analysis and recognition by GBP

Analysis of glycosaminoglycans relies mainly on characterization of disaccharides after enzymatic cleavage and further analysis using tandem mass spectrometry sequencing, reviewed in (Zaia, 2013). Clearly, GBP bind glycosaminoglycans; the example of antithrombin III binding to heparin is perhaps the most prominent, although these GAG-binding GBP are often not classified as lectins because they lack the signature fold that have been identified for various lectin families. Nonetheless, these GBP are selective in their binding. Using both literature-based and affinity proteomics approaches, the number of GBP that interact with heparin/heparan sulfate was estimated to be in the hundreds (Ori et al., 2011). Many hyaluronic acid-binding proteins and receptors are also known, notably CD44 and TLR4 (Day and Prestwich, 2002).

Glycoproteomics

The highest resolution of the glycome would also allow an assignment of individual glycan structures that are expressed at each site on a particular glycoprotein. This assignment is obviously much easier for glycoproteins that express only a few glycans compared to those that express hundreds. The higher the number of glycosylation sites on a protein, the greater the amount of material required for analysis, as well as the greater the complexity and time of analysis. Recently, site-specific glycosylation of alpha-dystroglycan, a glycoprotein with >20 glycosylation sites, has been characterized (Harrison et al., 2012; Stalnaker et al., 2010). Newly developed ion-trap mass spectrometry techniques to fragment glycopeptides such that glycans can be fragmented by CID, followed by ECD peptide fragmentation and identification, have increased the possibility of assigning glycans to particular amino acid sites in a mixture of glycopeptides, e.g., (Halim et al., 2012). Nonetheless, the goal of assigning glycan structures to specific sites on proteins in a complex mixture is still formidable. A novel strategy to metabolically label glycans and glycoconjugates of cultured cells using 15N-glutamine has been developed. This methodology will allow using MS to study the turnover of individual glycans on particular glycoproteins, as well as a detailed comparison of glycans in two populations of cells that differ in some way, such as after differentiation, oncogenic transformation or exposure to a cytokine (Fang et al., 2010; Orlando et al., 2009).

Glycogenes and Glycotranscriptomes

Among the initial steps to understand how the glycomes of human and mouse cells are regulated, the first requirement was to produce a list of glycosylation-related enzymes and proteins. The advent of the CaZY database (http://www.cazy.org), an up-to-date, curated collection of enzymes/proteins from all sources that act on carbohydrates, has allowed visualization of the breadth of proteins that recognize glycans or are involved in their metabolism (Nairn et al., 2008). This database, along with several others, has been used to generate a “parts list” of transcripts encoding proteins known or hypothesized (because of sequence identity or similarity) to bind, metabolize, or be directly involved in complex glycan synthesis, breakdown, or transport. The current list of transcripts involved in glycan recognition suggests a total of around 200–210 transcripts in human and mouse. The number of transcripts for putative GBPs, including GAG-binding proteins and others that do not fit into the lectin rubric, is obviously higher. Identification of the amounts of glycotranscripts expressed at any point in time by a particular cell type and how those transcripts change during differentiation, disease progression or experimental perturbation reveal important clues to understanding which glycans are expressed and their regulation. The nature of the many agents that regulated glycosylation, coupled with competition between biosynthetic enzymes, however, makes it very difficult to extrapolate from transcriptome data to predict precise levels of the expression of particular glycans. Nevertheless, a recent study focused on the glycomics of mouse embryonic stem cell differentiation to embryoid bodies (EB) and to extra-embryonic endoderm (EEE), in terms of both glyco-transcript changes and glycan changes (Nairn et al., 2012). The results suggested that in about 60% of the cases studied between ESC and EB and between ESC and EEE differentiation, changes in glycosyltransferase transcript levels were consistent with the observed changes in glycan expression.

Glycobioinformatics

An overarching goal of bioinformatics focusing on glycomics, also called glycobioinformatics, is to develop and provide tools and algorithms that facilitate the study and identification of glycans, their regulation and function. This includes software capable of interpreting analytical data to assign glycan structures, databases storing of glycan structures and meta information and algorithms linking these structural databases to protein databases, such as UniProt. (http://www.uniprot.org). Although analytical tools have not reached the degree of automation as is common in proteomic analysis, they are becoming more essential in the interpretation and display of experimental glycomics structural data. Two widely used tools are GlycanBuilder and GlycoWorkbench (Damerell et al., 2012). The most widely used databases are listed in the Glycomics Portal (http://glycomics.ccrc.uga.edu/GlycomicsPortal/) and allow not only finding and retrieving information about glycan structures, but also information about the interaction of these structures with other macromolecules. For example, searching the CFG and other databases containing glycan array binding data can assist the identification of GBP that recognize particular glycan structures or substructures.

Although progress has been made in developing and applying such glycobioinformatic tools, many challenges remain. It appears timely to present several proposals for development of bioinformatics tools in order to accelerate glycomics research: 1. An international, open access/open source registry for glycan structures must be developed. In order to facilitate communication between individuals, databases and in the scientific literature an uncurated database that simply associates specific structures with a unique identifier (i.e., accession number) is required. Such a “glycan namespace” will allow each structure, whether it is confirmed or not, to be unambiguously specified by a single identifier. These identifiers can then be used in the communication between tools and databases and will overcome the diverse, incompatible glycan sequence formats currently in use. In addition, a highly curated database containing vetted glycan structures along with meta-data (literature references, species information, attachment of the glycan to other macro molecules such as proteins, etc) is required to facilitate the annotation of laboratory data and provide conceptual links to related biological entities and concepts. UniCarb (http://unicarb-db.biomedicine.gu.seare) is making important, initial efforts toward such a curated database. Both of the databases must be freely accessible and readable by both scientists and computer programs. The curated database will be most useful if it conforms to MIRAGE standards (Minimum Information Required for A Glycomics Experiment), as published in Molecular and Cellular Proteomics (Kolarich et al., 2013). In addition storing attachment information of glycan structures to glycoconjugates will provide an entry point that can be used by databases of other research fields (e.g. proteomics). 2. Independent databases for experimental glycomics data must also be developed, based on various methodologies used in glycomics, e.g., MS, NMR, LC-MS, qRT-PCR/deep sequencing. These databases, which will be linked via the registries noted above with other data sets, must not just contain the annotated data but also the raw data and meta-data about experimental procedures used to generate and annotate the data. The information in these databases must be accessible for researchers but also presented in machine-readable formats and interfaces that allow interaction of the databases with annotation tools e.g., via Web services. To facilitate this communication, at some point there needs to be standardization and agreement on common formats for glycan analysis data (i.e., MS annotation, glycan array data). 3. Manuscripts published in journals and other publications must conform to MIRAGE standards, and clear data deposition requirements, analogous to the requirement to submit coordinate files to PDB for three-dimensional structure reports, must be established for publication. 4. (Semi-)automatic annotation tools, similar to those in use for proteomics, must be developed for interpreting and annotating tandem MS and MSn data for glycan structure assignment. These tools must allow sharing of data between users and assist deposition of data into common registries and databases.

To Glycoscience and Beyond

Glycoscience has made spectacular progress in the past few years with the advent of new technologies for exploring glycan structure and function, along with insights into the genetic and molecular aspects of glycan expression and regulation. The field is challenged, however, by the complexity and dynamic nature of the glycome and lack of understanding of how glycoconjugates are expressed topologically and temporally. Understanding the functions of glycans and their higher order contributions typically require physiological studies of organisms and identification of altered pathways of anabolism and catabolism in patients (Figure 7). Thus, combinations of studies in single cell systems may fail to identify or even predict these higher order functions. As for any type of biological system, the degree of understanding can be gauged by its predictive ability, and on this score glycoscience is truly challenged. One Holy Grail for glycomics researchers is to be able to predict the glycan structure(s) to be found on any particular amino acid site on any glycoprotein of interest; another is for functional glycomics researchers to predict the molecular interactions of a glycan with GBPs and GRMs. At present, our state of knowledge is far from predictive on any level, and we are limited to identifying potential sites of glycosylation on proteins given their amino acid sequences. Defining glycosylation sites and structures of glycans at specific sites is still technically difficult and limited to laboratories with advanced tools, especially considering all the types of amino acid modifications that can occur (Fig. 1). A further challenge for glycomics is to understand the links between expression patterns and levels of glycosyltransferases/glycosidases and their localization, spatially and temporally, relative to particular glycan structures, as well as the consequences of such expressions, and that of GBPs, on biological activities. In spite of these challenges, the future is growing brighter, as our knowledge grows about glycan structures and both their direct and indirect functions through studies performed across many biological disciplines. Technological advances in mass spectrometry and other sequencing methods, glycan synthesis, bioinformatics, and growing knowledge of biological roles of glycans in development, health, and disease, notably the CDGs, provide great hope for the future of glycomics. The possibility exists to finally tie the knot linking glycans to nucleic acids, proteins, and lipids in the grand theory of everything biological, assigning glycans their place as a pillar among those macromolecules that are essential for life.

An external file that holds a picture, illustration, etc.
Object name is nihms555066f7.jpg

The higher order functions of glycans are typically revealed in physiological studies of organisms, and to a lesser extent perhaps in studies based on isolated tissues and cells. Thus, mutations of genes involved in specific glycosylation pathways of anabolism or catabolism might have little effects on cultured cells, but are deleterious to organismal development; increasing biological complexity is denoted by the green arrow. Of course, the analytical ease of defining the glycome itself is simplified by considering free or released glycans, compared to individual glycoconjugates, such as a glycoprotein, and becomes increasingly difficult with larger biological complexity to the organism itself, denoted by the blue arrow.

Acknowledgments

Acknowledgements and Resources:

We gratefully acknowledge input and discussions from our colleagues, Kelley Moremen, Alison Nairn, René Ranzinger, David Smith, Michael Tiemeyer, Lance Wells, Rob Woods, and William York. We also acknowledge research support from NIH Grants to JMP (U01CA128454 and P41GM103490) and to RDC (P41GM103694; R24GM098791; U01CA168930; and R01AI101982).

Abbreviations

Glcglucose
GlcNAcN-acetylglucosamine
Galgalactose
GalNAcN-acetylgalactosamine
Manmannose
Fucfucose
GlcAglucuronate
IdoAiduronate
Xylxylose
NeuAcN-acetylneuraminate
GBPglycan-binding protein
GRMglycan recognition molecule
CFGConsortium for Functional Glycomics
MSmass spectrometry
CDGCongenital Disorders of Glycosylation
LLOlipid-linked oligosaccharide

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • Aebi M. N-linked protein glycosylation in the ER. Biochim Biophys Acta. 2013;1833:2430–2437. [PubMed] [Google Scholar]
  • Anthony RM, Nimmerjahn F, Ashline DJ, Reinhold VN, Paulson JC, Ravetch JV. Recapitulation of IVIG anti-inflammatory activity with a recombinant IgG Fc. Science. 2008a;320:373–376. [PMC free article] [PubMed] [Google Scholar]
  • Anthony RM, Wermeling F, Karlsson MC, Ravetch JV. Identification of a receptor required for the anti-inflammatory activity of IVIG. Proc Natl Acad Sci U S A. 2008b;105:19571–19578. [PMC free article] [PubMed] [Google Scholar]
  • Aoki K, Perlman M, Lim JM, Cantu R, Wells L, Tiemeyer M. Dynamic developmental elaboration of N-linked glycan complexity in the Drosophila melanogaster embryo. J Biol Chem. 2007;282:9127–9142. [PubMed] [Google Scholar]
  • Apweiler R, Hermjakob H, Sharon N. On the frequency of protein glycosylation, as deduced from analysis of the SWISS-PROT database. Biochim Biophys Acta. 1999;1473:4–8. [PubMed] [Google Scholar]
  • Arata Y, Hirabayashi J, Kasai K. Sugar binding properties of the two lectin domains of the tandem repeat-type galectin LEC-1 (N32) of Caenorhabditis elegans. Detailed analysis by an improved frontal affinity chromatography method. J Biol Chem. 2001;276:3068–3077. [PubMed] [Google Scholar]
  • Bao Y, Chen C, Newburg DS. Quantification of neutral human milk oligosaccharides by graphitic carbon high-performance liquid chromatography with tandem mass spectrometry. Anal Biochem. 2013;433:28–35. [PMC free article] [PubMed] [Google Scholar]
  • Belardi B, O’Donoghue GP, Smith AW, Groves JT, Bertozzi CR. Investigating cell surface galectin-mediated cross-linking on glycoengineered cells. J Am Chem Soc. 2012;134:9549–9552. [PMC free article] [PubMed] [Google Scholar]
  • Blixt O, Head S, Mondala T, Scanlan C, Huflejt ME, Alvarez R, Bryan MC, Fazio F, Calarese D, Stevens J, et al. Printed covalent glycan array for ligand profiling of diverse glycan binding proteins. Proc Natl Acad Sci U S A. 2004;101:17033–17038. [PMC free article] [PubMed] [Google Scholar]
  • Boccuto L, Aoki K, Flanagan-Steet H, Chen C, Fan X, Bartel F, Petukh M, Pittman A, Saul R, Chaubey A, et al. A mutation in a ganglioside biosynthetic enzyme, ST3GAL5, results in salt & pepper syndrome, a neurocutaneous disorder with altered glycolipid and glycoprotein glycosylation. Hum Mol Genet 2013 [PMC free article] [PubMed] [Google Scholar]
  • Bochner BS, Alvarez RA, Mehta P, Bovin NV, Blixt O, White JR, Schnaar RL. Glycan array screening reveals a candidate ligand for Siglec-8. J Biol Chem. 2005;280:4307–4312. [PubMed] [Google Scholar]
  • Bode L. Human milk oligosaccharides: every baby needs a sugar mama. Glycobiology. 2012;22:1147–1162. [PMC free article] [PubMed] [Google Scholar]
  • Bond MR, Hanover JA. O-GlcNAc cycling: a link between metabolism and chronic disease. Annu Rev Nutr. 2013;33:205–229. [PMC free article] [PubMed] [Google Scholar]
  • Braakman I, Bulleid NJ. Protein folding and modification in the mammalian endoplasmic reticulum. Annu Rev Biochem. 2011;80:71–99. [PubMed] [Google Scholar]
  • Campanero-Rhodes MA, Childs RA, Kiso M, Komba S, Le Narvor C, Warren J, Otto D, Crocker PR, Feizi T. Carbohydrate microarrays reveal sulphation as a modulator of siglec binding. Biochem Biophys Res Commun. 2006;344:1141–1146. [PubMed] [Google Scholar]
  • Cao Y, Stosiek P, Springer GF, Karsten U. Thomsen-Friedenreich-related carbohydrate antigens in normal adult human tissues: a systematic and comparative study. Histochem Cell Biol. 1996;106:197–207. [PubMed] [Google Scholar]
  • Chai W, Yuen CT, Kogelberg H, Carruthers RA, Margolis RU, Feizi T, Lawson AM. High prevalence of 2-mono- and 2,6-di-substituted manol-terminating sequences among O-glycans released from brain glycopeptides by reductive alkaline hydrolysis. European journal of biochemistry/FEBS. 1999;263:879–888. [PubMed] [Google Scholar]
  • Chang CF, Pan JF, Lin CN, Wu IL, Wong CH, Lin CH. Rapid characterization of sugar-binding specificity by in-solution proximity binding with photosensitizers. Glycobiology. 2011;21:895–902. [PubMed] [Google Scholar]
  • Chevolot Y. Carbohydrate Microarrays: Methods and Protocols. New York: Humana Press; 2012. [Google Scholar]
  • Copeland RJ, Han G, Hart GW. O-GlcNAcomics-Revealing roles of O-GlcNAcylation in disease mechanisms and development of potential diagnostics. Proteomics Clin Appl 2013 [PMC free article] [PubMed] [Google Scholar]
  • Cummings RD. The repertoire of glycan determinants in the human glycome. Mol Biosyst. 2009;5:1087–1104. [PubMed] [Google Scholar]
  • Damerell D, Ceroni A, Maass K, Ranzinger R, Dell A, Haslam SM. The GlycanBuilder and GlycoWorkbench glycoinformatics tools: updates and new developments. Biol Chem. 2012;393:1357–1362. [PubMed] [Google Scholar]
  • Day AJ, Prestwich GD. Hyaluronan-binding proteins: tying up the giant. J Biol Chem. 2002;277:4585–4588. [PubMed] [Google Scholar]
  • Day CJ, Semchenko EA, Korolik V. Glycoconjugates play a key role in Campylobacter jejuni Infection: interactions between host and pathogen. Front Cell Infect Microbiol. 2012;2:9. [PMC free article] [PubMed] [Google Scholar]
  • de Paz JL, Seeberger PH. Recent advances and future challenges in glycan microarray technology. Methods Mol Biol. 2012;808:1–12. [PubMed] [Google Scholar]
  • Dennis JW, Nabi IR, Demetriou M. Metabolism, cell surface organization, and disease. Cell. 2009;139:1229–1241. [PMC free article] [PubMed] [Google Scholar]
  • Doherty M, McManus CA, Duke R, Rudd PM. High-throughput quantitative N-glycan analysis of glycoproteins. Methods Mol Biol. 2012;899:293–313. [PubMed] [Google Scholar]
  • Duverger E, Lamerant-Fayel N, Frison N, Monsigny M. Carbohydrate-lectin interactions assayed by SPR. Methods Mol Biol. 2010;627:157–178. [PubMed] [Google Scholar]
  • Fang M, Lim JM, Wells L. Quantitative Glycomics of Cultured Cells Using Isotopic Detection of Aminosugars with Glutamine (IDAWG) Curr Protoc Chem Biol. 2010;2:55–69. [PMC free article] [PubMed] [Google Scholar]
  • Feinberg H, Rowntree TJ, Tan SL, Drickamer K, Weis WI, Taylor ME. Common polymorphisms in human langerin change specificity for glycan ligands. J Biol Chem 2013 [PMC free article] [PubMed] [Google Scholar]
  • Filocamo M, Morrone A. Lysosomal storage disorders: molecular basis and laboratory testing. Hum Genomics. 2011;5:156–169. [PMC free article] [PubMed] [Google Scholar]
  • Ford MG, Weimar T, Kohli T, Woods RJ. Molecular dynamics simulations of galectin-1-oligosaccharide complexes reveal the molecular basis for ligand diversity. Proteins. 2003;53:229–240. [PMC free article] [PubMed] [Google Scholar]
  • Freeze HH. Understanding human glycosylation disorders: biochemistry leads the charge. J Biol Chem. 2013;288:6936–6945. [PMC free article] [PubMed] [Google Scholar]
  • Fujitani N, Takegawa Y, Ishibashi Y, Araki K, Furukawa J, Mitsutake S, Igarashi Y, Ito M, Shinohara Y. Qualitative and quantitative cellular glycomics of glycosphingolipids based on rhodococcal endoglycosylceramidase-assisted glycan cleavage, glycoblotting-assisted sample preparation, and matrix-assisted laser desorption ionization tandem time-of-flight mass spectrometry analysis. J Biol Chem. 2011;286:41669–41679. [PMC free article] [PubMed] [Google Scholar]
  • Garma L, Mukherjee S, Mitra P, Zhang Y. How many protein-protein interactions types exist in nature? PLoS One. 2012;7:e38913. [PMC free article] [PubMed] [Google Scholar]
  • Grewal PK, Uchiyama S, Ditto D, Varki N, Le DT, Nizet V, Marth JD. The Ashwell receptor mitigates the lethal coagulopathy of sepsis. Nat Med. 2008;14:648–655. [PMC free article] [PubMed] [Google Scholar]
  • Gupta GS, Gupta A, Gupta RK. Animal Lectins: Form, Function and Clinical Applications. Heidelberg: Springer; 2012. [Google Scholar]
  • Gutierrez Gallego R, Haseley SR, van Miegem VF, Vliegenthart JF, Kamerling JP. Identification of carbohydrates binding to lectins by using surface plasmon resonance in combination with HPLC profiling. Glycobiology. 2004;14:373–386. [PubMed] [Google Scholar]
  • Halim A, Nilsson J, Ruetschi U, Hesse C, Larson G. Human urinary glycoproteomics; attachment site specific analysis of N- and O-linked glycosylations by CID and ECD. Mol Cell Proteomics. 2012;11 M111 013649. [PMC free article] [PubMed] [Google Scholar]
  • Han L, Costello CE. Mass spectrometry of glycans. Biochemistry Biokhimiia. 2013;78:710–720. [PMC free article] [PubMed] [Google Scholar]
  • Handa K, Hakomori SI. Carbohydrate to carbohydrate interaction in development process and cancer progression. Glycoconj J. 2012;29:627–637. [PubMed] [Google Scholar]
  • Hanisch FG. O-glycoproteomics: site-specific O-glycoprotein analysis by CID/ETD electrospray ionization tandem mass spectrometry and top-down glycoprotein sequencing by in-source decay MALDI mass spectrometry. Methods in molecular biology. 2012;842:179–189. [PubMed] [Google Scholar]
  • Harrison R, Hitchen PG, Panico M, Morris HR, Mekhaiel D, Pleass RJ, Dell A, Hewitt JE, Haslam SM. Glycoproteomic characterization of recombinant mouse alpha-dystroglycan. Glycobiology. 2012;22:662–675. [PMC free article] [PubMed] [Google Scholar]
  • Hart GW, Slawson C, Ramirez-Correa G, Lagerlof O. Cross talk between O-GlcNAcylation and phosphorylation: roles in signaling, transcription, and chronic disease. Annu Rev Biochem. 2011;80:825–858. [PMC free article] [PubMed] [Google Scholar]
  • Hennet T. Diseases of glycosylation beyond classical congenital disorders of glycosylation. Biochim Biophys Acta. 2012;1820:1306–1317. [PubMed] [Google Scholar]
  • Hennet T, Chui D, Paulson JC, Marth JD. Immune regulation by the ST6Gal sialyltransferase. Proc Natl Acad Sci U S A. 1998;95:4504–4509. [PMC free article] [PubMed] [Google Scholar]
  • Hileman RE, Fromm JR, Weiler JM, Linhardt RJ. Glycosaminoglycan-protein interactions: definition of consensus sites in glycosaminoglycan binding proteins. Bioessays. 1998;20:156–167. [PubMed] [Google Scholar]
  • Hudgin RL, Pricer WE, Jr, Ashwell G, Stockert RJ, Morell AG. The isolation and properties of a rabbit liver binding protein specific for asialoglycoproteins. J Biol Chem. 1974;249:5536–5543. [PubMed] [Google Scholar]
  • Inamori K, Hara Y, Willer T, Anderson ME, Zhu Z, Yoshida-Moriguchi T, Campbell KP. Xylosyl- and glucuronyltransferase functions of LARGE in alpha-dystroglycan modification are conserved in LARGE2. Glycobiology. 2013;23:295–302. [PMC free article] [PubMed] [Google Scholar]
  • Inokuchi J. Membrane microdomains and insulin resistance. FEBS letters. 2010;584:1864–1871. [PubMed] [Google Scholar]
  • Ito H, Kuno A, Sawaki H, Sogabe M, Ozaki H, Tanaka Y, Mizokami M, Shoda J, Angata T, Sato T, et al. Strategy for glycoproteomics: identification of glyco-alteration using multiple glycan profiling tools. J Proteome Res. 2009;8:1358–1367. [PubMed] [Google Scholar]
  • Ju T, Wang Y, Aryal RP, Lehoux SD, Ding X, Kudelka MR, Cutler C, Zeng J, Wang J, Sun X, et al. Tn and sialyl-Tn antigens, aberrant O-glycomics as human disease markers. Proteomics Clin Appl 2013 [PMC free article] [PubMed] [Google Scholar]
  • Kabayama K, Sato T, Saito K, Loberto N, Prinetti A, Sonnino S, Kinjo M, Igarashi Y, Inokuchi J. Dissociation of the insulin receptor and caveolin-1 complex by ganglioside GM3 in the state of insulin resistance. Proc Natl Acad Sci U S A. 2007;104:13678–13683. [PMC free article] [PubMed] [Google Scholar]
  • Kaneko Y, Nimmerjahn F, Ravetch JV. Anti-inflammatory activity of immunoglobulin G resulting from Fc sialylation. Science. 2006;313:670–673. [PubMed] [Google Scholar]
  • Karlsson KA. Animal glycosphingolipids as membrane attachment sites for bacteria. Annu Rev Biochem. 1989;58:309–350. [PubMed] [Google Scholar]
  • Ke C, Destecroix H, Crump MP, Davis AP. A simple and accessible synthetic lectin for glucose recognition and sensing. Nature chemistry. 2012;4:718–723. [PubMed] [Google Scholar]
  • Kiwamoto T, Brummet ME, Wu F, Motari MG, Smith DF, Schnaar RL, Zhu Z, Bochner BS. Mice deficient in the St3gal3 gene product alpha2,3 sialyltransferase (ST3Gal-III) exhibit enhanced allergic eosinophilic airway inflammation. J Allergy Clin Immunol 2013a [PMC free article] [PubMed] [Google Scholar]
  • Kiwamoto T, Katoh T, Tiemeyer M, Bochner BS. The role of lung epithelial ligands for Siglec-8 and Siglec-F in eosinophilic inflammation. Curr Opin Allergy Clin Immunol. 2013b;13:106–111. [PMC free article] [PubMed] [Google Scholar]
  • Kolarich D, Rapp E, Struwe WB, Haslam SM, Zaia J, McBride R, Agravat S, Campbell MP, Kato M, Ranzinger R, et al. The minimum information required for a glycomics experiment (MIRAGE) project: improving the standards for reporting mass-spectrometry-based glycoanalytic data. Mol Cell Proteomics. 2013;12:991–995. [PMC free article] [PubMed] [Google Scholar]
  • Kuster B, Mann M. 18O-labeling of N-glycosylation sites to improve the identification of gel-separated glycoproteins using peptide mass mapping and database searching. Anal Chem. 1999;71:1431–1440. [PubMed] [Google Scholar]
  • Lakshminarayanan V, Thompson P, Wolfert MA, Buskas T, Bradley JM, Pathangey LB, Madsen CS, Cohen PA, Gendler SJ, Boons GJ. Immune recognition of tumor-associated mucin MUC1 is achieved by a fully synthetic aberrantly glycosylated MUC1 tripartite vaccine. Proc Natl Acad Sci U S A. 2012;109:261–266. [PMC free article] [PubMed] [Google Scholar]
  • Larsson JM, Karlsson H, Crespo JG, Johansson ME, Eklund L, Sjovall H, Hansson GC. Altered O-glycosylation profile of MUC2 mucin occurs in active ulcerative colitis and is associated with increased inflammation. Inflamm Bowel Dis. 2011;17:2299–2307. [PubMed] [Google Scholar]
  • Levery SB. Glycosphingolipid structural analysis and glycosphingolipidomics. Methods Enzymol. 2005;405:300–369. [PubMed] [Google Scholar]
  • Li M, Lin N, Huang Z, Du L, Altier C, Fang H, Wang B. Selecting aptamers for a glycoprotein through the incorporation of the boronic acid moiety. J Am Chem Soc. 2008;130:12636–12638. [PMC free article] [PubMed] [Google Scholar]
  • Lindesmith L, Moe C, Marionneau S, Ruvoen N, Jiang X, Lindblad L, Stewart P, LePendu J, Baric R. Human susceptibility and resistance to Norwalk virus infection. Nat Med. 2003;9:548–553. [PubMed] [Google Scholar]
  • Liu BA, Engelmann BW, Nash PD. High-throughput analysis of peptide-binding modules. Proteomics. 2012a;12:1527–1546. [PMC free article] [PubMed] [Google Scholar]
  • Liu Y, Childs RA, Palma AS, Campanero-Rhodes MA, Stoll MS, Chai W, Feizi T. Neoglycolipid-based oligosaccharide microarray system: preparation of NGLs and their noncovalent immobilization on nitrocellulose-coated glass slides for microarray analyses. Methods Mol Biol. 2012b;808:117–136. [PubMed] [Google Scholar]
  • Liu Y, Palma AS, Feizi T. Carbohydrate microarrays: key developments in glycobiology. Biol Chem. 2009;390:647–656. [PubMed] [Google Scholar]
  • Macauley MS, Pfrengle F, Rademacher C, Nycholat CM, Gale AJ, von Drygalski A, Paulson JC. Antigenic liposomes displaying CD22 ligands induce antigen-specific B cell apoptosis. J Clin Invest. 2013;123:3074–3083. [PMC free article] [PubMed] [Google Scholar]
  • Marth JD. A unified vision of the building blocks of life. Nature cell biology. 2008;10:1015–1016. [PMC free article] [PubMed] [Google Scholar]
  • Mercer N, Ramakrishnan B, Boeggeman E, Verdi L, Qasba PK. Use of novel mutant galactosyltransferase for the bioconjugation of terminal N-acetylglucosamine (GlcNAc) residues on live cell surface. Bioconjugate chemistry. 2013;24:144–152. [PMC free article] [PubMed] [Google Scholar]
  • Moremen KW, Tiemeyer M, Nairn AV. Vertebrate protein glycosylation: diversity, synthesis and function. Nat Rev Mol Cell Biol. 2012;13:448–462. [PMC free article] [PubMed] [Google Scholar]
  • Nairn AV, Aoki K, dela Rosa M, Porterfield M, Lim JM, Kulik M, Pierce JM, Wells L, Dalton S, Tiemeyer M, et al. Regulation of glycan structures in murine embryonic stem cells: combined transcript profiling of glycan-related genes and glycan structural analysis. J Biol Chem. 2012;287:37835–37856. [PMC free article] [PubMed] [Google Scholar]
  • Nairn AV, York WS, Harris K, Hall EM, Pierce JM, Moremen KW. Regulation of glycan structures in animal tissues: transcript profiling of glycan-related genes. J Biol Chem. 2008;283:17298–17313. [PMC free article] [PubMed] [Google Scholar]
  • Nicolardi S, van der Burgt YE, Dragan I, Hensbergen PJ, Deelder AM. Identification of new apolipoprotein-CIII glycoforms with ultrahigh resolution MALDI-FTICR mass spectrometry of human sera. Journal of proteome research. 2013;12:2260–2268. [PubMed] [Google Scholar]
  • Nimmerjahn F, Ravetch JV. Anti-inflammatory actions of intravenous immunoglobulin. Annu Rev Immunol. 2008;26:513–533. [PubMed] [Google Scholar]
  • Ninonuevo MR, Park Y, Yin H, Zhang J, Ward RE, Clowers BH, German JB, Freeman SL, Killeen K, Grimm R, et al. A strategy for annotating the human milk glycome. J Agric Food Chem. 2006;54:7471–7480. [PubMed] [Google Scholar]
  • North SJ, Hitchen PG, Haslam SM, Dell A. Mass spectrometry in the analysis of N-linked and O-linked glycans. Curr Opin Struct Biol. 2009;19:498–506. [PMC free article] [PubMed] [Google Scholar]
  • Ohtsubo K, Marth JD. Glycosylation in cellular mechanisms of health and disease. Cell. 2006;126:855–867. [PubMed] [Google Scholar]
  • Ohtsubo K, Takamatsu S, Minowa MT, Yoshida A, Takeuchi M, Marth JD. Dietary and genetic control of glucose transporter 2 glycosylation promotes insulin secretion in suppressing diabetes. Cell. 2005;123:1307–1321. [PubMed] [Google Scholar]
  • Ori A, Wilkinson MC, Fernig DG. A systems biology approach for the investigation of the heparin/heparan sulfate interactome. J Biol Chem. 2011;286:19892–19904. [PMC free article] [PubMed] [Google Scholar]
  • Orlando R. Quantitative glycomics. Methods Mol Biol. 2010;600:31–49. [PubMed] [Google Scholar]
  • Orlando R, Lim JM, Atwood JA, 3rd, Angel PM, Fang M, Aoki K, Alvarez-Manilla G, Moremen KW, York WS, Tiemeyer M, et al. IDAWG: Metabolic incorporation of stable isotope labels for quantitative glycomics of cultured cells. J Proteome Res. 2009;8:3816–3823. [PMC free article] [PubMed] [Google Scholar]
  • Palaniappan KK, Hangauer MJ, Smith TJ, Smart BP, Pitcher AA, Cheng EH, Bertozzi CR, Boyce M. A Chemical Glycoproteomics Platform Reveals O-GlcNAcylation of Mitochondrial Voltage-Dependent Anion Channel 2. Cell Rep 2013 [PMC free article] [PubMed] [Google Scholar]
  • Park S, Gildersleeve JC, Blixt O, Shin I. Carbohydrate microarrays. Chem Soc Rev. 2013;42:4310–4326. [PMC free article] [PubMed] [Google Scholar]
  • Partridge EA, Le Roy C, Di Guglielmo GM, Pawling J, Cheung P, Granovsky M, Nabi IR, Wrana JL, Dennis JW. Regulation of cytokine receptors by Golgi N-glycan processing and endocytosis. Science. 2004;306:120–124. [PubMed] [Google Scholar]
  • Patnode ML, Cheng CW, Chou CC, Singer MS, Elin MS, Uchimura K, Crocker PR, Khoo KH, Rosen SD. Galactose 6-O-sulfotransferases are not required for the generation of Siglec-F ligands in leukocytes or lung tissue. J Biol Chem. 2013;288:26533–26545. [PMC free article] [PubMed] [Google Scholar]
  • Pawson T, Warner N. Oncogenic re-wiring of cellular signaling pathways. Oncogene. 2007;26:1268–1275. [PubMed] [Google Scholar]
  • Powell LD, Sgroi D, Sjoberg ER, Stamenkovic I, Varki A. Natural ligands of the B cell adhesion molecule CD22 beta carry N-linked oligosaccharides with alpha-2,6-linked sialic acids that are required for recognition. J Biol Chem. 1993;268:7019–7027. [PubMed] [Google Scholar]
  • Powell LD, Varki A. The oligosaccharide binding specificities of CD22 beta, a sialic acid-specific lectin of B cells. J Biol Chem. 1994;269:10628–10636. [PubMed] [Google Scholar]
  • Rademacher C, Paulson JC. Glycan fingerprints: calculating diversity in glycan libraries. ACS Chem Biol. 2012;7:829–834. [PMC free article] [PubMed] [Google Scholar]
  • Rillahan CD, Paulson JC. Glycan microarrays for decoding the glycome. Annu Rev Biochem. 2011;80:797–823. [PMC free article] [PubMed] [Google Scholar]
  • Sanda M, Pompach P, Brnakova Z, Wu J, Makambi K, Goldman R. Quantitative liquid chromatography-mass spectrometry-multiple reaction monitoring (LC-MS-MRM) analysis of site-specific glycoforms of haptoglobin in liver disease. Mol Cell Proteomics. 2013;12:1294–1305. [PMC free article] [PubMed] [Google Scholar]
  • Schulenburg H, Hoeppner MP, Weiner J, 3rd, Bornberg-Bauer E. Specificity of the innate immune system and diversity of C-type lectin domain (CTLD) proteins in the nematode Caenorhabditis elegans. Immunobiology. 2008;213:237–250. [PubMed] [Google Scholar]
  • Schwab I, Biburger M, Kronke G, Schett G, Nimmerjahn F. IVIg-mediated amelioration of ITP in mice is dependent on sialic acid and SIGNR1. Eur J Immunol. 2012;42:826–830. [PubMed] [Google Scholar]
  • Sgroi D, Varki A, Braesch-Andersen S, Stamenkovic I. CD22, a B cell-specific immunoglobulin superfamily member, is a sialic acid-binding lectin. J Biol Chem. 1993;268:7011–7018. [PubMed] [Google Scholar]
  • Shang J, Piskarev VE, Xia M, Huang P, Jiang X, Likhosherstov LM, Novikova OS, Newburg DS, Ratner DM. Identifying human milk glycans that inhibit norovirus binding using surface plasmon resonance. Glycobiology 2013 [PMC free article] [PubMed] [Google Scholar]
  • Smith DF, Cummings RD. Application of microarrays for deciphering the structure and function of the human glycome. Mol Cell Proteomics. 2013;12:902–912. [PMC free article] [PubMed] [Google Scholar]
  • Song X, Lasanajak Y, Xia B, Heimburg-Molinaro J, Rhea JM, Ju H, Zhao C, Molinaro RJ, Cummings RD, Smith DF. Shotgun glycomics: a microarray strategy for functional glycomics. Nat Methods. 2011;8:85–90. [PMC free article] [PubMed] [Google Scholar]
  • Spiro RG. Protein glycosylation: nature, distribution, enzymatic formation, and disease implications of glycopeptide bonds. Glycobiology. 2002;12:43R–56R. [PubMed] [Google Scholar]
  • Stalnaker SH, Hashmi S, Lim JM, Aoki K, Porterfield M, Gutierrez-Sanchez G, Wheeler J, Ervasti JM, Bergmann C, Tiemeyer M, et al. Site mapping and characterization of O-glycan structures on alpha-dystroglycan isolated from rabbit skeletal muscle. J Biol Chem. 2010;285:24882–24891. [PMC free article] [PubMed] [Google Scholar]
  • Steentoft C, Vakhrushev SY, Joshi HJ, Kong Y, Vester-Christensen MB, Schjoldager KT, Lavrsen K, Dabelsteen S, Pedersen NB, Marcos-Silva L, et al. Precision mapping of the human O-GalNAc glycoproteome through SimpleCell technology. EMBO J. 2013;32:1478–1488. [PMC free article] [PubMed] [Google Scholar]
  • Stepper J, Shastri S, Loo TS, Preston JC, Novak P, Man P, Moore CH, Havlicek V, Patchett ML, Norris GE. Cysteine S-glycosylation, a new post-translational modification found in glycopeptide bacteriocins. FEBS Lett. 2011;585:645–650. [PubMed] [Google Scholar]
  • Stevens J, Blixt O, Paulson JC, Wilson IA. Glycan microarray technologies: tools to survey host specificity of influenza viruses. Nat Rev Microbiol. 2006;4:857–864. [PMC free article] [PubMed] [Google Scholar]
  • Suzuki A, Miyazaki M, Matsuda J, Yoneshige A. High-performance thin-layer chromatography/mass spectrometry for the analysis of neutral glycosphingolipids. Biochim Biophys Acta. 2011;1811:861–874. [PubMed] [Google Scholar]
  • Takeda Y, Seko A, Sakono M, Hachisu M, Koizumi A, Fujikawa K, Ito Y. Parallel quantification of lectin-glycan interaction using ultrafiltration. Carbohydr Res. 2013;375:112–117. [PubMed] [Google Scholar]
  • Tateno H. SUEL-related lectins, a lectin family widely distributed throughout organisms. Biosci Biotechnol Biochem. 2010;74:1141–1144. [PubMed] [Google Scholar]
  • Tateno H, Crocker PR, Paulson JC. Mouse Siglec-F and human Siglec-8 are functionally convergent paralogs that are selectively expressed on eosinophils and recognize 6′-sulfo-sialyl Lewis X as a preferred glycan ligand. Glycobiology. 2005;15:1125–1135. [PubMed] [Google Scholar]
  • Tateno H, Mori A, Uchiyama N, Yabe R, Iwaki J, Shikanai T, Angata T, Narimatsu H, Hirabayashi J. Glycoconjugate microarray based on an evanescent-field fluorescence-assisted detection principle for investigation of glycan-binding proteins. Glycobiology. 2008;18:789–798. [PubMed] [Google Scholar]
  • Taylor ME, Drickamer K. Introduction to Glycobiology. New York: Oxford University Press; 2011. [Google Scholar]
  • Thanabalasingham G, Huffman JE, Kattla JJ, Novokmet M, Rudan I, Gloyn AL, Hayward C, Adamczyk B, Reynolds RM, Muzinic A, et al. Mutations in HNF1A result in marked alterations of plasma glycan profile. Diabetes. 2013;62:1329–1337. [PMC free article] [PubMed] [Google Scholar]
  • Topin J, Arnaud J, Sarkar A, Audfray A, Gillon E, Perez S, Jamet H, Varrot A, Imberty A, Thomas A. Deciphering the glycan preference of bacterial lectins by glycan array and molecular docking with validation by microcalorimetry and crystallography. PLoS One. 2013;8:e71149. [PMC free article] [PubMed] [Google Scholar]
  • Urashima T, Kobata A, Fukuda K, Ohnishi M, Kitaoka M. Milk Oligosaccharides. Hauppauge, NY: Nova Science Publishers, Inc; 2011. [Google Scholar]
  • van den Berg LM, Gringhuis SI, Geijtenbeek TB. An evolutionary perspective on C-type lectins in infection and immunity. Ann N Y Acad Sci. 2012;1253:149–158. [PubMed] [Google Scholar]
  • Van den Steen P, Rudd PM, Dwek RA, Opdenakker G. Concepts and principles of O-linked glycosylation. Crit Rev Biochem Mol Biol. 1998;33:151–208. [PubMed] [Google Scholar]
  • van Diepen A, Smit CH, van Egmond L, Kabatereine NB, Pinot de Moira A, Dunne DW, Hokke CH. Differential anti-glycan antibody responses in Schistosoma mansoni-infected children and adults studied by shotgun glycan microarray. PLoS Negl Trop Dis. 2012;6:e1922. [PMC free article] [PubMed] [Google Scholar]
  • Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME. Essentials of Glycobiology. 2. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 2009. [PubMed] [Google Scholar]
  • Vasta GR, Ahmed H. Animal Lectins: A Functional View. New York: CRC Press; 2009. [Google Scholar]
  • Venkatesan K, Rual JF, Vazquez A, Stelzl U, Lemmens I, Hirozane-Kishikawa T, Hao T, Zenkner M, Xin X, Goh KI, et al. An empirical framework for binary interactome mapping. Nat Methods. 2009;6:83–90. [PMC free article] [PubMed] [Google Scholar]
  • Vester-Christensen MB, Halim A, Joshi HJ, Steentoft C, Bennett EP, Levery SB, Vakhrushev SY, Clausen H. Mining the O-mannose glycoproteome reveals cadherins as major O-mannosylated glycoproteins. Proc Natl Acad Sci U S A 2013 [PMC free article] [PubMed] [Google Scholar]
  • Walther T, Karamanska R, Chan RW, Chan MC, Jia N, Air G, Hopton C, Wong MP, Dell A, Malik Peiris JS, et al. Glycomic analysis of human respiratory tract tissues and correlation with influenza virus infection. PLoS Pathog. 2013;9:e1003223. [PMC free article] [PubMed] [Google Scholar]
  • Wang Z, Chinoy ZS, Ambre SG, Peng W, McBride R, de Vries RP, Glushka J, Paulson JC, Boons GJ. A general strategy for the chemoenzymatic synthesis of asymmetrically branched N-glycans. Science. 2013;341:379–383. [PMC free article] [PubMed] [Google Scholar]
  • Werz DB, Ranzinger R, Herget S, Adibekian A, von der Lieth CW, Seeberger PH. Exploring the structural diversity of mammalian carbohydrates (“glycospace”) by statistical databank analysis. ACS Chem Biol. 2007;2:685–691. [Google Scholar]
  • Wilkins PP, Moore KL, McEver RP, Cummings RD. Tyrosine sulfation of P-selectin glycoprotein ligand-1 is required for high affinity binding to P-selectin. J Biol Chem. 1995;270:22677–22680. [PubMed] [Google Scholar]
  • Xia B, Royall JA, Damera G, Sachdev GP, Cummings RD. Altered O-glycosylation and sulfation of airway mucins associated with cystic fibrosis. Glycobiology. 2005;15:747–775. [PubMed] [Google Scholar]
  • Yang AC, Ng BG, Moore SA, Rush J, Waechter CJ, Raymond KM, Willer T, Campbell KP, Freeze HH, Mehta L. Congenital disorder of glycosylation due to DPM1 mutations presenting with dystroglycanopathy-type congenital muscular dystrophy. Molecular genetics and metabolism. 2013;110:345–351. [PMC free article] [PubMed] [Google Scholar]
  • Yi W, Clark PM, Mason DE, Keenan MC, Hill C, Goddard WA, 3rd, Peters EC, Driggers EM, Hsieh-Wilson LC. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science. 2012;337:975–980. [PMC free article] [PubMed] [Google Scholar]
  • Yoshida-Moriguchi T, Yu L, Stalnaker SH, Davis S, Kunz S, Madson M, Oldstone MB, Schachter H, Wells L, Campbell KP. O-mannosyl phosphorylation of alpha-dystroglycan is required for laminin binding. Science. 2010;327:88–92. [PMC free article] [PubMed] [Google Scholar]
  • Yu C, Ali S, St-Germain J, Liu Y, Yu X, Jaye DL, Moran MF, Cooper MD, Ehrhardt GR. Purification and identification of cell surface antigens using lamprey monoclonal antibodies. Journal of immunological methods. 2012;386:43–49. [PMC free article] [PubMed] [Google Scholar]
  • Yu X, Jiang Y, Chen Y, Huang Y, Costello CE, Lin C. Detailed glycan structural characterization by electronic excitation dissociation. Anal Chem. 2013a;85:10017–10021. [PMC free article] [PubMed] [Google Scholar]
  • Yu X, Vasiljevic S, Mitchell DA, Crispin M, Scanlan CN. Dissecting the molecular mechanism of IVIg therapy: the interaction between serum IgG and DC-SIGN is independent of antibody glycoform or Fc domain. J Mol Biol. 2013b;425:1253–1258. [PubMed] [Google Scholar]
  • Zaia J. Glycosaminoglycan glycomics using mass spectrometry. Mol Cell Proteomics. 2013;12:885–892. [PMC free article] [PubMed] [Google Scholar]
  • Zhang H, Wang Z, Stupak J, Ghribi O, Geiger JD, Liu QY, Li J. Targeted glycomics by selected reaction monitoring for highly sensitive glycan compositional analysis. Proteomics. 2012;12:2510–2522. [PMC free article] [PubMed] [Google Scholar]
  • Zhang L. Glycosaminoglycan (GAG) biosynthesis and GAG-binding proteins. Prog Mol Biol Transl Sci. 2010;93:1–17. [PubMed] [Google Scholar]
  • Zhao A, Nunez-Cruz S, Li C, Coukos G, Siegel DL, Scholler N. Rapid isolation of high-affinity human antibodies against the tumor vascular marker Endosialin/TEM1, using a paired yeast-display/secretory scFv library platform. Journal of immunological methods. 2011;363:221–232. [PMC free article] [PubMed] [Google Scholar]
  • Zielinska DF, Gnad F, Schropp K, Wisniewski JR, Mann M. Mapping N-glycosylation sites across seven evolutionarily distant species reveals a divergent substrate proteome despite a common core machinery. Mol Cell. 2012;46:542–548. [PubMed] [Google Scholar]
  • Zielinska DF, Gnad F, Wisniewski JR, Mann M. Precision mapping of an in vivo N-glycoproteome reveals rigid topological and sequence constraints. Cell. 2010;141:897–907. [PubMed] [Google Scholar]
-