Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Glycobiology. 2017 Jan 1; 27(1): 3–49.
Published online 2016 Dec 28. doi: 10.1093/glycob/cww086
PMCID: PMC5884436
PMID: 27558841

Biological roles of glycans

Abstract

Simple and complex carbohydrates (glycans) have long been known to play major metabolic, structural and physical roles in biological systems. Targeted microbial binding to host glycans has also been studied for decades. But such biological roles can only explain some of the remarkable complexity and organismal diversity of glycans in nature. Reviewing the subject about two decades ago, one could find very few clear-cut instances of glycan-recognition-specific biological roles of glycans that were of intrinsic value to the organism expressing them. In striking contrast there is now a profusion of examples, such that this updated review cannot be comprehensive. Instead, a historical overview is presented, broad principles outlined and a few examples cited, representing diverse types of roles, mediated by various glycan classes, in different evolutionary lineages. What remains unchanged is the fact that while all theories regarding biological roles of glycans are supported by compelling evidence, exceptions to each can be found. In retrospect, this is not surprising. Complex and diverse glycans appear to be ubiquitous to all cells in nature, and essential to all life forms. Thus, >3 billion years of evolution consistently generated organisms that use these molecules for many key biological roles, even while sometimes coopting them for minor functions. In this respect, glycans are no different from other major macromolecular building blocks of life (nucleic acids, proteins and lipids), simply more rapidly evolving and complex. It is time for the diverse functional roles of glycans to be fully incorporated into the mainstream of biological sciences.

Keywords: biological roles, evolution, glycans

Introduction

In 1993, this journal published a review concluding that while limited evidence for all of the theories on the biological roles of glycans was available, exceptions to each could also be found (1). Some general principles were suggested. First, the reported biological consequences of experimental modification of glycosylation seemed highly variable, making it difficult to predict a priori the functions that a given glycan structure might be mediating, or its relative importance to the organism. Second, limited data suggested that the same glycan might mediate different functions at different locations within an organism, or at different times in its ontogeny. Third, the more specific intrinsic biological roles of glycans known at the time appeared to be mediated by unusual glycan sequences, unusual presentations of common sequences or further modifications of glycans. But it was also noted that such sequences were more likely to be targets for specific recognition by toxins and pathogenic microorganisms. It was therefore posited that ongoing host–pathogen interactions might contribute to the evolution of some aspects of intra- and interspecies glycan variation. Finally, some suggestions were made as to how one might elucidate more intrinsic biological functions for glycans. In particular, it was suggested that more studies of natural and induced mutations resulting in altered glycosylation within intact organisms would be required.

In the decade that followed, the author tracked several other more focused discussions of glycan functions, some examples of which are cited here (2104). For a while it was indeed possible for an individual to track and read such reviews on biological roles of glycans, but this became increasingly difficult over time. By a decade later it became impossible to do so, and one had to be content with tracking a sampling of reviews in areas of ongoing personal interest (105159). Meanwhile, some of the concepts in the original review were updated in a book chapter (160). Evolutionary and phylogenetic perspectives on the matter have also since been extensively addressed (161176). More recently, it has been emphasized that glycans are as universal in nature as nucleic acids, proteins, lipids and metabolites (177), and as essential to the existence of all known living organisms (178). But as depicted in Figure Figure1,1, glycans are also the most structurally diverse and rapidly evolving major class of molecules. Taken together with much greater technical difficulties in their analysis, one can understand why the knowledge base regarding these major building blocks of life has lagged so far behind.

An external file that holds a picture, illustration, etc.
Object name is cww086f01.jpg

Universal characteristics of all living cells. As indicated in the figure and discussed in the text, glycosylation is among the key features of all living cells. However, in contrast to the genetic code, the degree of chemical complexity and evolutionary diversification of glycans amongst various taxa is the greatest. The likely reasons for this difference are discussed in the text, and can help explain the still rather limited knowledge base regarding this class of molecules. But we now know that dense and complex glycosylation is universal to all living cells and even most viruses. Evidently, more than 3 billion years of evolution has failed to generate a free-living cell devoid of glycosylation. Thus, one can conclude that glycosylation is as essential to life as a genetic code. Figure modified from ref. 178 and used with permission from Varki A. 2011a. Cold Spring Harb Perspect Biol. 3, doi:pii: 10.1101/cshperspect. a005462. Copyright: Cold Spring Harbor Laboratory Press.

Despite these challenges, it is evident that information regarding the biological roles of glycans has vastly expanded in the last two decades. The present review first surveys the history of how our understanding of biological roles of glycans originally evolved, and then attempts to update the overview as of mid-2016. As a measure of how much progress has been made, any attempt at being comprehensive is now impractical, and the knowledge base of a single individual cannot do justice to this vast and complex field. Thus, one is only able to illustrate general principles with a few selected examples, and with a strong emphasis on the expertise of the author. For the same reason, the bibliography of citations cannot be comprehensive. Also, most of the broad implications of glycosylation for biotherapeutics are not addressed (179).

It is assumed that the reader is generally familiar with the major types and classes of glycans found in nature, and the conventional terminologies for describing them (180). Of course, it is important to also recognize that the full range of types and distributions of glycans in nature are still largely unexplored, and surprises continue to emerge. To cite just a few examples, the following glycans were mostly unknown when the previous version of this review was being written: functional sialylation in the fly nervous system (181); O-fucose and O-glucose glycans on Notch (182184), and O-fucose on thrombospondin repeats (185); mucin type O-glycosylation in protists (initiated with α-GlcNAc instead of α-GalNAc) (186); O-linked N-acetylglucosamine on cell surface/extracellular proteins (187188); the C-mannose linkage to proteins (189190); the complexities of O-mannose-linked glycans in tissues such as muscle (191194), including the novel glycosaminoglycan attached on α-dystroglycan (195200) generated by a dual function xylosyl/glucuronosyltransferase (201204) and attached via novel ribitol-phosphate bridge (205208); a plant cell wall proteoglycan wherein a core arabinogalactan protein is glycosylated with cell wall matrix xylan and pectin glycans (209); identification of β-galacturonic acid in a xyloglucan involved in plant root hair tip growth (210); the expanding diversity of milk oligosaccharides (211), recognition of immunomodulatory glycans in the gut microbiome (212213); N-linked glycans in prokaryotes (214216) and, discovery of the large family of prokaryotic nonulosonic acids (217219), the likely ancestors of sialic acids (220). The last two examples highlight the realization that many glycosylation types once thought to be unique to eukaryotes in fact have their origins in earlier evolved pathways in bacteria and archea (176, 221222). In this regard, it is notable that many major taxa of life forms such as archea, fungi, protists and algae still remain poorly explored with regard to glycan structure and functions (many such taxa are not much addressed in this review).

Historical background

The first half of the 20th century saw great strides in elucidation of the structure and biochemistry of simple and complex glycans found in nature, garnering many Nobel Prizes (180). Beyond their well-known roles in energy generation and metabolism, glycans obviously had many structural and biophysical roles in many systems, including nutritional storage. Given the dense coating of complex and diverse glycans on essentially all cell surfaces (sometimes called the “glycocalyx” in animal cells) as well as on most extracellular molecules, it was also not surprising to find many examples of infectious agents or symbiotic organisms that recognized such glycans with a high degree of specificity, mediating interactions with their hosts (223). Additionally many pathogens were found to express highly specific glycans on their own surfaces, which seemed to modulate their antigenicity and/or their susceptibility to bacteriophages. Meanwhile, pathogens were also found to elaborate highly specific exo- and endoglycosidases that could degrade host glycans. In fact, many of the structural details of eukaryotic glycans were initially deduced by using such microbial glycosidases as tools (224226).

The discovery of corresponding lysosomal glycosidases intrinsic to eukaryotic systems (227) then led to a better understanding of so-called “storage disorders”, wherein the deficiency of a single lysosomal glycosidase resulted in accumulation of the corresponding nondegraded product in lysosomes (228). Meanwhile, great strides were made in elucidating the structures of glycans in some taxa (particularly vertebrates), as well in understanding their biosynthetic pathways. The development of vertebrate cell lines with defined defects in most glycosylation pathways provided powerful and conclusive evidence for many complex glycan biosynthetic pathways, and tools for their in-depth study (229240), especially N-linked glycans on glycoproteins. Ironically, in vitro viability of these remarkable cell lines despite their gross defects in glycosylation raised questions in the minds of some scientists, as to whether complex glycans have specific and critically important functions intrinsic to intact vertebrate organisms.

Despite all these great strides in understanding the structure, biosynthesis and metabolism of glycans in several taxa, remarkably little was still known about their specific functions, beyond their metabolic, structural, biophysical and pathogen-facilitating roles. But early clues did exist for more specific intrinsic biological roles. Some of the “blood groups” that limited blood transfusion between individual humans could be explained by intraspecies variations in glycosylation (241). The effects of glycosidase pretreatment on the subsequent intravascular trafficking of blood cells in vivo raised the possibility that glycans might serve as targeting signals (242). The role of mammalian ɑ-lactalbumin in the generation of lactose in milk had also been elucidated (243), but the functional relevance of the resulting profusion of species-specific milk oligosaccharides elaborated from a lactose core oligosaccharide (244) remained elusive. A consistent finding of altered glycosylation in malignant cells suggested specific roles in cancer progression (245247). The selective reaggregation of dispersed sponge cells was shown to be due to carbohydrate–carbohydrate interactions between large acidic glycans (248).

Meanwhile, a major clue to a specific role of glycans in vertebrate systems emerged with the discovery of the asialoglycoprotein receptor, which recognized and bound to exposed β-linked galactose residues on desialylated glycoproteins, to rapidly clear them away in the liver (249250). But the intrinsic biological function of this highly specific hepatocyte endocytic receptor remained obscure at the time. Regardless, the concept that a terminal sugar on a glycan could act as an intraorganismal targeting signal was established, and evidence then emerged for a mannose receptor on macrophages (251252) and possibly one for mannose 6-phosphate on other cell types (253254). The isolation and characterization of many plant and animal glycan-binding proteins by methods such as affinity chromatography occurred in parallel (255263). During this period, the well-known pharmacological anticoagulant effect of the natural glycosaminoglycan heparin was shown to be due to a highly specific interaction of antithrombin with a particular 3-O-sulfated sequence (264266) within the heparin chain. These and other such findings provided indirect evidence that complex glycans might carry out specific functions of intrinsic value to the complex multicellular organisms that synthesized them. However, as of the end of the 1970s there remained no direct proof that glycans played such key biological roles. Even as late as 1988, the introduction to a major symposium on the topic stated that “...while the functions of DNA and proteins are generally known...it is much less clear what carbohydrates do...” (267).

In reality, a few specific examples had been defined earlier in the 1980s. The discovery and characterization of the rare human genetic disorder called I-cell disease (268) had led to the prediction that lysosomal enzymes shared a common recognition marker that mediated organelle-specific uptake into cells (269). The blockade of this uptake by mannose 6-phosphate (but not glucose 6-phosphate) (253) then led to the correct prediction that the glycans on these enzymes must selectively express a novel phosphomannosyl marker (254, 270) that was recognized by specific receptors, which might mediate both intra- and intercellular trafficking of these enzymes to their correct destination in lysosomes.

Elucidation of the biological significance of this presumed lysosomal enzyme trafficking pathway required the determination of the structures of the novel glycans involved (271273), and discovery of the enzymatic basis of the generation of this “phosphomannosyl recognition marker” (274281). All of this work culminated in the discovery of the biochemical defect in I-cell disease and related human genetic disorders, which turned out to be a failure of the initial phosphorylation mechanism (276278). Thus, for the first time one could state that specific recognition of a unique glycan mediated an equally specific and critical biological role, which was of intrinsic value to the organism that had synthesized the glycan.

A few years later, studies showed that small fungal cell wall glycan fragments could send highly specific signals to plants. Signal transmission depended on the precise stereochemistry of the glycans (282284). This concept of “oligosaccharins” was extended to other glycan fragments that could manipulate morphogenetic pathways of tobacco explants (285) providing preliminary evidence that glycans by themselves might act as signaling molecules internal to a species. Meanwhile, studies in animals indicated that sialidase treatment could abrogate the interaction of lymphocytes with high endothelial venules in lymph nodes (286) leading to the correct prediction that sialylated glycan signals were involved in the trafficking of lymphocytes out of the circulation. Along with other convergent lines of evidence, this eventually resulted in the definition of a family of cell adhesion molecules (287288) that were critical for leukocyte rolling on endothelium, prior to their exit from the circulation. These molecules were called “selectins” (289), and they recognized a common motif, consisting of sialylated fucosylated glycans (7, 13, 24, 91, 103, 128, 140, 290309); a topic that has continued to blossom, with implications for many fields.

While all this progress was occurring, it was generally assumed that glycosylation was only found on cell surface and secreted molecules, and that the nucleus and cytoplasm were devoid of this class of post-translational modification. The discovery of O-linked GlcNAc (310314) thus went unrecognized even by most other glycoscientists for years, until it was finally realized that this nucleocytoplasmic modification is the most common form of glycosylation in eukaryotic cells (314316), and that it mediates numerous modulatory functions on many proteins, including a complex interplay with protein phosphorylation (317320).

The increasing number of animal lectins that being discovered and characterized was then classified based on sequence homologies into C-type and S-type lectins (321), and the latter were eventually redesignated as galectins (322323). Discovery of the sialic acid-binding properties of sialoadhesin (324) and of CD22 (325), followed by the cloning of sialoadhesin (326), led to the definition of a new family of cell-type-specific vertebrate lectins initially called “sialoadhesins” (15, 327), but eventually designated as a subfamily of I-type lectins (328) and renamed as the Siglecs (329). The previously discovered phosphomannosyl receptors were now redesignated as “P-type” Lectins (85) and some previously known plant lectins became founding members of the “R-type” (Ricin-like) and “L-type” (Legume lectin-like) families (330). The power of phylogenomic sequence comparisons has since revealed many additional families of lectins such as the X-type lectins (intelectins) (331333), Ficolins (334335), etc. The earlier-mentioned somatic cell mutants in pathways of N- and O-glycosylation played key roles in this progress. Along with the discovery of new human genetic disorders in glycosylation (see next section), these and many other clues to intrinsic biological roles of glycans in the 1990s finally opened up this vast and uncharted territory of biology. Combined with the accelerating power of genomics and glycomics, we have now reached a point where numerous biological roles of glycans have been elucidated, to varying degrees of precision. For reasons of brevity, only a few examples are considered in this review.

Learning from natural or induced genetic alterations of glycosylation in multicellular organisms

Many different approaches have been used to elucidate the biological roles of glycans. Among these, one of the most instructive has been the study of genetic alterations of glycosylation in model organisms, and in human diseases. Indeed as mentioned earlier, it was the discovery of the genetic defect in I-cell disease that conclusively proved the biological significance and importance of the mannose 6-phosphate targeting pathway in vivo. At about the same time a defect in 3′-phosphoadenosine 5′-phosphosulfate (PAPS) formation was found in brachymorphic mice with multiple sulfation defects. While PAPS has many roles, the disproportionately short stature of the mice was apparently due to undersulfation of chondroitin sulfate in epiphyseal growth-plate cartilages (336). Another decade went by (see Figure Figure2)2) before the second human biosynthetic defect specific to glycosylation was discovered, a deficiency of a glycosaminoglycan core galactosyltransferase in a progeria-like syndrome (337). Meanwhile, the concept of “Carbohydrate Deficient Glycoprotein syndromes” (CDGs) had been suggested, based on the finding that children with previously unexplained multisystem disorders showed under-glycosylation of serum transferrin (338341)––a test originally devised to detect alcoholism via the general hypo-sialylation it causes in liver-derived serum glycoproteins (342)! The work of many investigators then led to the elucidation of the underlying enzymatic and genetic defects in these children (343350), eventually resulting in the repurposing of the acronym CDG to denote “Congenital Disorders of Glycosylation” (66, 351356). After a slow start in the early 1990s an international effort of many investigators has now resulted in a veritable explosion in discoveries of human genetic disorders of glycosylation (Figure (Figure2)2) (61, 67, 191, 352, 357382). These disorders continue to provide a goldmine of clues to biological roles of glycans, and this understanding has begun to benefit some patients via simple monosaccharide replacement therapies (374, 383388). Notably, in a recent study of consecutively enrolled patients with unexplained intellectual developmental disorder and metabolic phenotypes, whole exome sequencing showed that >10% were attributable to genetic defects in glycosylation pathways (389), mostly hypomorphic states of genes in which complete loss would have been lethal.

An external file that holds a picture, illustration, etc.
Object name is cww086f02.jpg

Accelerating progress in the discovery of human glycosylation disorders. The graph shows the cumulative number of human disorders with a major genetic defect in various glycosylation pathways and the year of their identification (2016 data for first 6 months). In early years, initial discovery was based on compelling biochemical evidence, and in later years by conclusive genetic proof. In most instances, the year indicates the occurrence of definitive proof of gene-specific mutations and correlations to biochemical results. Figure kindly provided by H. Freeze and Bobby Ng, updated from ref. 375 and reproduced with permission from Freeze HH, Chong JX, Bamshad MJ, Ng BG. 2014. Am J Hum Genet. 94:161–175. Copyright Elsevier. Reproduced with permission.

Beyond the clinically obvious CDGs, there is also increasing genetic evidence for the role of glycosylation-related genes in more subtle and common diseases in the population, as observed in genome wide association studies a few examples of which are mentioned here: NDST3 (390) and ST8SIA2 (391393) in schizophrenia and bipolar disorder; FUT2 nonsecretor status and blood group B associated with elevated serum lipase activity and risk for chronic pancreatitis (394); and type 2 diabetes susceptibility associated with ST6GAL1 (395).

Meanwhile, targeted genetic alterations of glycan biosynthetic pathways in mice (396401) also revealed a spectrum of abnormalities, again pointing to complex and varied functions of glycans in multicellular organisms. Since then, the list of mice with genetically altered glycosylation has expanded greatly, and resulting phenotypes have been highly instructive (98, 159). It is ironic that most of the glycosylation pathways that had earlier been dismissed because genetic defects caused “limited phenotypes” in the reductionist environment of the tissue culture dish later turned out to have clear and serious consequences in the intact organism, even in the form of hypomorphic alleles in humans. On the other hand, the phenotypic outcome of gene knockouts has been rather unpredictable. For example, while the MGAT1/GnT-I null state (which prevents the processing of N-glycans) caused embryonic lethality in mice (397398), it generated no grossly obvious phenotype in the Arabidopsis plant (402403), and limited phenotypes in Drosophila (404). Conversely while mice lacking ST3GAL5 seem to have only moderate phenotypes (405406), humans with similar defects suffer from severe multisystem disease (407408). Of course, any report of a “viable and fertile mouse with no major phenotype” must be taken with a large grain of salt. For example, the consequences of altering complex ganglioside biosynthesis (409411) or of knocking out one of the key ganglioside receptors called MAG (412413) was mostly evident later in the life, or when the mouse was subjected to specific challenges (414415). In contrast, complete elimination of ganglioside biosynthesis gave an early embryonic lethal phenotype (416). Further complexity has arisen from the realization that there are multiple isozymes of some glycosyltransferases (417418), and that post-transcriptional regulation by micro-RNAs is occurring (419420).

Summaries of all human and model organism phenotypes resulting from genetic alterations in glycosylation will not be attempted here, and are reviewed elsewhere (98, 159, 375, 378). In general, complete elimination of major classes or subclasses of glycans tends to result in embryonic lethality, while defects in outer terminal structures often give viable organisms with defects in specific functions and/or specific cell types, although these impacts are often species specific. As an example, null alleles preventing the synthesis of the core glycosaminoglycan backbone of heparan sulfate causes embryonic lethality (421422), but the prevention of proper sulfation of this backbone can give living mice with specific defects (422427). When embryonic lethality makes it difficult to define specific biological roles, tissue-specific targeted genetic alterations became important (428430). Experiments of nature such as somatic mutations in X-linked genes (431) and hypomorphic alleles of essential genes (388389, 432) have also helped our understanding of functions.

Note that the discussion above largely focused on examples from animals. While space does not allow a detailed discussion, loss of glycosylation in plants, fungi or prokaryotes can also lead to cell death. For example, a meristem-localized inducible expression of an UDP-glycosyltransferase gene is essential for growth and development in pea and alfalfa (433). Ethambutol (a traditional drug treatment for tuberculosis) is now known to target the arabinofuranosyltransferases EmbA and EmbB (434). Knockouts of these genes in mycobacteria are lethal, as is the case with some other glycosyltransferases (435). And in the fungus Aspergillus fumigatus, inhibition of cell wall β glucan synthesis is toxic (436437).

A broad classification of the biological roles of glycans

There are several different ways to classify the biological roles of glycans, based on the glycan types in question, on the glycan-binding protein involved, etc. A simple and broad classification (160) (see Figure Figure33 for a conceptual organization and Table TableII for a complete listing) divides glycan functions into four somewhat distinct categories. The first is structural and modulatory roles (including nutrient sequestration). The second category involves extrinsic (interspecies) recognition. The third is intrinsic (intraspecies) recognition. Finally, there is molecular mimicry of host glycans. All of these categories can involve glycan-binding proteins (see Figure Figure3).3). The next part of this review considers these classes of biological roles and discusses one or more examples of each. Given the vastness of relevant literature, the examples and citations are rather limited and biased towards the knowledge of the author. Examples of multifunctional roles of glycans and glycan-binding proteins that cross over between these somewhat arbitrary categories will be mentioned later.

An external file that holds a picture, illustration, etc.
Object name is cww086f03.jpg

General classification of the biological roles of glycans. A simplified and broad classification is presented, especially emphasizing the roles of organism-intrinsic and organism-extrinsic glycan-binding proteins in recognizing glycans. There is some overlap between the categories, e.g., some structural properties involve specific recognition of glycans. Binding shown on the left of the central “self” cell represents intrinsic recognition, and extrinsic recognition is represented by binding shown to the right of that cell. Molecular mimicry of host glycans adds further complexity to potential roles. Original drawing by R. Cummings, updated from ref. 160 with permission from the Consortium of Glycobiology Editors.

Table I.

Biological roles of glycans

Structural and modulatory roles
Physical structure
Physical protection and tissue elasticity
Water solubility of macromolecules
Lubrication
Physical expulsion of pathogens
Diffusion barriers
Glycoprotein folding
Protection from proteases
Modulation of membrane receptor signaling
Membrane organization
Modulation of transmembrane receptor spatial organization and function
Antiadhesive action
Depot functions
Nutritional storage
Gradient generation
Extracellular matrix organization
Protection from immune recognition
Effects of glycan branching on glycoprotein function
Cell surface glycan:lectin-based lattices
Masking or modification of ligands for glycan-binding proteins
Tuning a range of function
Molecular functional switching
Epigenetic histone modifications
Extrinsic (interspecies) recognition of glycans
Bacterial, fungal and parasite adhesins
Viral agglutinins
Bacterial and plant toxins
Soluble host proteins that recognize pathogens
Pathogen glycosidases
Host decoys
Herd immunity
Pathogen-associated molecular patterns
Immune modulation of host by symbiont/parasite
Antigen recognition, uptake and processing
Bacteriophage recognition of glycan targets
Intrinsic (intraspecies) recognition of glycans
Intracellular glycoprotein folding and degradation
Intracellular glycoprotein trafficking
Triggering of endocytosis and phagocytosis
Intercellular signaling
Intercellular adhesion
Cell–matrix interactions
Fertilization and reproduction
Clearance of damaged glycoconjugates and cells
Glycans as clearance receptors
Danger-associated molecular patterns
Self-associated molecular patterns
Antigenic epitopes
Xeno-autoantigens
Molecular mimicry of host glycans
Convergent evolution of host-like glycans
Appropriation of host glycans

Structural and modulatory roles

Given their ubiquitous presence and abundance in all cellular compartments, extracellular spaces and body fluids, glycans have many major biological effects mediated by their own primary structural properties, and/or by modulating functions of proteins and lipids to which they are attached. The randomly selected examples provided in each section do not focus on any particular glycan class per se.

Physical structure

β-Linked homopolymers of glucose or N-acetylglucosamine (cellulose or chitin, respectively) are among the most abundant organic molecules on the planet, providing strength and rigidity to structures such as plant and fungal cell walls and arthropod exoskeletons (438439). These polymers are also difficult to breakdown by physical, chemical or enzymatic means. Many other glycan polymers play major roles in fungal and plant cell wall structure and function (438447). For example, the hemicellulose xyloglucan not only plays a key role in the loosening and tightening of cellulose microfibrils, but also enables the plant cell to change its shape during growth and differentiation, and to retain its final shape after maturation (441). Needless to say, in the absence of these and many other major glycan polymers, the diversity of macroscopic structural variations in life forms on the planet would be far more limited.

Physical protection and tissue elasticity

There are many instances where thick layers of glycans provide an important physical protective role. In addition to the polymers mentioned above, the dense layer of mucins that coats many epithelial surfaces such as the inner lining of airways and intestines provides critical barrier functions, including protection against the invasion by microorganisms that live within the lumen (139, 418, 448452). Disruption of this layer by genetically altering mucin backbones, O-linked glycosyltransferases (or key chaperones like Cosmc) can have very serious consequences, including inflammation and carcinogenesis associated with microbial invasion (453456). Likewise the thick and biochemically robust cell walls of plants make it difficult for invading fungi and bacteria to reach the membrane of the plant cells (438, 441, 444, 446447). In other instances, the thick layer of glycans also provides tissue strength. Nature is rife with many more such examples including fungal and bacterial cell walls and polysaccharide coats, and the glycosaminoglycans of vertebrate cartilage, which are partly responsible for its elasticity, resiliency and compressibility (457).

Water solubility of macromolecules

It is interesting that many vertebrate internal body fluids such as blood plasma are rich in heavily glycosylated proteins. Apart from specific functional reasons for glycosylation, hydrophilic and acidic glycans also contribute significantly to the water solubility of these macromolecules. Indeed the remarkably high concentration of proteins in the blood plasma (~50–70 mg/mL in humans, carrying ~2 mM of bound sialic acids) would probably be impossible without this glycosylation. The antifreeze glycoproteins of certain fish alter the structure of bulk water itself, preventing nucleation of ice crystals in body fluids (458461). Antifreeze functions can also be mediated by certain polysaccharides with a lipid component (460).

Lubrication

The remarkably efficient lubrication provided by soluble and membrane-bound mucins on the lining of hollow organs may seem like a trivial “function”––until one realizes that deficiencies in oral salivary mucins caused by radiation damage to salivary glands (a side effect of head-and-neck cancer treatment) (462463) or by autoimmune disease (Sjogren's) (464) can be life-threatening, especially by limiting the ability to swallow food. Another example is the critical lubricating role of hyaluronan in body fluids (465467), such as the synovial fluid in joint cavities and tear fluid in the eyes, wherein deficiencies can be supplemented therapeutically (467471).

Physical expulsion of pathogens

Heavily glycosylated secretions produced in large amounts can serve as a response to physically expel intruders. For example, expulsion of N. brasiliensis worms from the rat intestine is associated not only with quantitative, but also with qualitative changes in the composition of mucins in goblet cells (472). On the microscopic level, a recent study shows that a “sentinel” goblet cell localized to the mouse colonic crypt entrance recognizes bacterial products, activating the Nlrp6 inflammasome, eventually inducing mucin secretion from adjacent goblet cells in the upper crypt, which expels bacterial intruders that have penetrated the protective inner mucus layer (473).

Diffusion barriers

Extracellular matrix glycosaminoglycans and/or heavily sialylated glycoproteins can comprise critical diffusion barriers. For example, the heavily sialylated protein podocalyxin on glomerular podocyte foot processes (474478) and heparan sulfate glycosaminoglycans within the glomerular basement membrane (479484) seem to play important roles in maintaining the integrity of blood plasma filtration by the kidney. Pathological or experimental damage to such glycans causes large molecules like albumin to escape into the urine, and is associated with glomerular diseases (478, 485491).

Glycoprotein folding

Protein molecules that are synthesized and secreted via the ER-Golgi pathway can be subjected to ER modifications such as O-fucosylation (492) and O-mannosylation (493), with important effects on facilitating proper folding in the ER lumen. A major fraction of such ER-synthesized proteins are also modified by N-linked glycans at Asn-X-Ser/Thr sequons (494), and it is reasonable to think that the large, generally hydrophilic sugar chains contribute to proper folding of nascent polypeptides emerging into the lumen of the ER. Indeed, it has long been known that preventing N-linked glycosylation using the inhibitor tunicamycin can have negative effects on the initial folding of such proteins (495496). In keeping with the highly conserved structure of the initial glycan added to asparagine residues, we now know that such N-glycans play a much more precise role in actually directing the folding, via specific recognition of certain features of N-glycans (see further discussion on quality control below). Even at the level of initial protein folding the exact context of the sequon can dictate the outcome. For example, experimentally placing a phenylalanine residue two or three positions before a glycosylated asparagine in distinct reverse turns facilitates stabilizing interactions between the aromatic side chain and the first GlcNAc residue of the glycan, while increasing glycosylation efficiency (497).

In this context, it is worth noting that the vast majority of published crystal structures of naturally occurring glycoproteins are derived from proteins that either had their glycosylation sites mutated, or had their glycans partially or completely degraded, prior to crystallization. The practical reason for making such a drastic change is that glycans are often heterogeneous and have a high range of motion, making it difficult to obtain an ordered crystal. Even if crystallization is possible, the glycans are typically disordered within the resulting image. In instances where glycans are left intact, the glycoproteins are often expressed in heterologous cells, resulting in nonspecies-specific glycosylation. This major technical artifact is rarely addressed in prominent protein crystallography papers. The bottom line is that when glycosylation sites are mutated or glycosylation is modified, there is a significant possibility that the folded form defined by crystallography may not be the native state. Solving this technical problem remains a major challenge for the future (498500), one in which new techniques such as cryo-electron microscopy (501502) may help. Meanwhile, it is definitely a worthwhile exercise to model the glycans back into the crystal structure to a best approximation (501, 503). Exceptions to the general lack of glycans in crystal structures can occur when the glycan is buried and partially immobilized within the folds of the protein, such as in the case of the Immunoglobulin-G (IgG) Fc-region (504506), or tightly packed on the surface, such as in the HIV virion (507509).

Protection from proteases

Heavily glycosylated proteins are protected from protease cleavage by glycans, likely by steric hindrance or negative charge. This effect is particularly prominent for mucins carrying densely packed O-glycans (451, 510). Indeed, extended segments of some mucins are even resistant to broad-spectrum proteases like proteinase K (511512). This property can actually be taken advantage of, to isolate mucin segments away from other proteins that could be more easily proteolyzed into smaller fragments, or even from whole tissues (511512). In a prokaryotic example, N-glycosylation in Campylobacter improves fitness, by providing protection against proteases in the gut (513). Conversely, there is evidence that glycosylation at single sites can regulate specific cleavage events with large impacts on protein activity (514), e.g., the protection of Tango1 by O-glycosylation is critical to apical secretion in Drosophila (515).

Modulation of membrane receptor signaling

Classic studies have shown that glycolipids can alter the signaling properties of protein receptors present within the same cell membrane (516). For example, subtly different forms of the sialylated ganglioside GM3 can have differential effects on tyrosine kinase signaling of the EGF receptor (517522) and elimination of GM3 affects insulin receptor action (405, 523524). Another classic example is the co-receptor activity of heparan sulfate in FGF signaling (525526). Glycosylation can also affect the signaling properties of the proteins to which it is attached. For example, ɑ1-6 core fucosylation of N-glycans affects transforming growth factor (TGF) signaling (527). Dysregulation of TGF-β1 receptor activation leads to abnormal lung development. While most core fucose-deficient mice die 3 d after birth, the survivors develop emphysematous changes of the lung. The underlying mechanism appears to be dysregulation of downstream TGF signaling, causing MMP gene activation, which eventually degrades alveolar membranes to give emphysema. In a similar vein, both sialylation and fucosylation modulate epidermal growth factor receptor-mediated intracellular signaling (528530).

An entirely new field opened up with the discovery that the Fringe molecule is a glycosyltransferase that modifies the important signaling protein Notch and thus modulates Notch–Delta interactions (182, 184). Before Fringe can act, Notch must first be glycosylated with an O-fucose, and the protein O-fucosyltransferase 1 is thus an essential component of Notch signaling pathways (531532). It was later discovered that the O-glucose modification on Notch added by a glucosyltransferase encoded by Rumi is also essential for Notch signaling and embryonic development (533534). Thus there are many roles of glycosylation in Notch signaling (535536). The structural basis of glycosylation-mediated Notch interactions with some of its ligands has been recently explored (537).

Membrane organization

Glycans can have profound effects on the organization of cell membranes. For example, GPI-anchored proteins are mainly associated with glycolipid-enriched membrane microdomains (538) and are organized in submicron domains at the cell surface (539). Cell surface lectins may also participate. Galectin-4 appears to be the major organizing factor of such “lipid rafts” on gastric epithelial cells (540). GPI-anchored proteins are selectively targeted to the apical surface in fully polarized epithelial cells (541). It also stands to reason that the glycans on cell surface glycoproteins can modulate membrane domain organization by their bulk and charge (542543). It now appears that glycans on one class of glycoproteins can even modulate the organization of other classes of glycans on other proteins present on the same cell surface, perhaps forming “clustered saccharide patches” (146, 302, 544). The formation of galectin-mediated lattices in the glycocalyx is discussed below.

Modulation of transmembrane receptor spatial organization and function

In addition to the role of heparan sulfate proteoglycans in modulating transmembrane receptor spatial organization discussed above, bulky glycoproteins in the cell surface glycocalyx can indirectly promote cell adhesion and signaling, facilitating integrin clustering by funneling active integrins into adhesions and altering their state, by applying tension to these matrix-bound molecules (545). This in turn promotes focal adhesion assembly and facilitates integrin-dependent growth factor signaling to support cell growth and survival. Since a bulky glycocalyx is a feature of malignant cells, it is suggested that these features could foster the spread of cancer by mechanically enhancing cell surface receptor function (545). However such mechanisms are also likely to operate in normal cells, which presumably exist in a continuum of biophysical states of the glycocalyx.

Antiadhesive action

Large acidic polymers such as hyaluronan and polysialic acid can inhibit cell–cell and cell–matrix interactions by virtue of both bulk and negative charge. These antiadhesive functions are particularly prominent during phases of development when cell migration is very active. The “plasticity” resulting from polysialic acid expression appears to be important for neuronal migration as well as reorganization following injury (546554).

Depot functions

Hydrophilic glycans on cell surfaces and extracellular matrices are capable of attracting and ordering water molecules (555). Beyond retaining water and cations (for unknown reasons, positively charged glycans are uncommon in nature), extracellular matrix glycosaminoglycans and polysialic acid can act as depots for growth factors and other bioactive molecules, which can be stored locally and released when needed, e.g., during injury and wound healing (86, 556561).

Nutritional storage

Polymeric glycans like glycogen in animal cells and starch in plants serve obvious roles in the long-term storage of glucose as an energy source, and marathon runners must build up liver glycogen stores before the big race. The earlier comment about O-linked GlcNAc being the first known form of cytosolic glycosylation is not strictly true, as the glycogenin protein was also known to self-glucosylate itself on a tyrosine residue with a short 8–12 glucose residue polymer in order to serve as the primer for glycogen synthesis (562567). In contrast, the mechanism of potato starch biosynthesis appeared to involve de novo synthesis, not an amylogenin primer (568).

Gradient generation

Gradients of growth factors can be generated by binding to extracellular matrix glycosaminoglycans such as heparan sulfate, especially in embryonic development (86). This organization of growth factors by glycosaminoglycans may contribute the morphogen gradients that are critical during development (569572).

Extracellular matrix organization

Many components of the extracellular matrix in vertebrates are large glycan polymers such as sulfated glycosaminoglycans and hyaluronan, that self-organize along with specific proteins into larger aggregates to generate structures such as basement membranes (573574) and cartilage (575577). Cartilage also acts as a template for primary and secondary ossification centers, development of the growth plates and the end of long bones, and the laying down of bone (578). Organizational roles are also obvious for glycans in the extracellular matrices of plants (see discussion above), and new roles are emerging for glycans in the biofilms surrounding bacteria, enabling them to form discrete multicellular communities (579583).

Protection from immune recognition

The adaptive immune system of vertebrate organisms functions largely by recognition of foreign peptide sequences, which are directly recognized by the B cell surface Ig receptor (584585), and are also loaded into the grooves of the major histocompatibility receptors to be presented to specific T-cell receptors (586). If the peptide carries a very small glycan, this moiety can contribute novel specificity to recognition (587591). However larger glycans typically disrupt peptide loading and/or T-cell receptor recognition, and often eliminate it altogether. This explains a common immune escape strategy of enveloped viruses, whose surface glycoproteins tend to be very heavily glycosylated (592593). Sometimes, such protective glycosylation can become so dense that it generates unique clustered epitopes recognized by specific antibodies, such as that seen on the surface of the HIV virion (507509, 594). In other instances, one type of glycan can block immune recognition of another, such as the Cryptococcus neoformans yeast cell wall, which is required for virulence (595596), apparently by protecting the deeper structures of the organism from recognition and attack by the host immune system.

Effects of glycan branching on glycoprotein function

The N-linked glycans on cell surface glycoproteins can have varying degrees of branching (597), and glycan branching is specifically upregulated in T-cell activation (598), and in malignant transformation (599604). Beyond their effects on protein structure per se, certain branched glycans can affect a variety of biological functions. Thus, there are reports of regulation of cytokine receptors by modulation of endocytosis rates by the type of glycan structure and branching (605), and N-glycan number and degree of branching can cooperate to regulate cell proliferation and differentiation (606) as well as thymocyte positive selection (607). The degree of branching of N-glycans is primarily dependent on the addition of β-linked GlcNAc residues donated by UDP-GlcNAc. Given that glucosamine and GlcNAc are major metabolic intermediates in most cells, the level of UDP-GlcNAc provides a likely connection between cellular metabolism, cell surface organization and disease (608). In keeping with this concept, the cell surface residency time of glucose transporter 2 is regulated by branching of its N-glycans, and alters insulin secretion as well (609). This provides a connection between diabetes, pancreatic β cell glycosylation and glucose transport (610). A different kind of N-glycan branching (so-called bissecting GlcNAc) inhibits growth factor signaling and retards mammary tumor progression (611) and E-cadherin may be a target molecule for this glycan modulating effect (612613).

Cell surface glycan:lectin-based lattices

The glycocalyx on the surface of vertebrate cells is often likened to a semi-randomly organized tropical rain forest (146), or to a sea-floor kelp bed (the latter analogy by P. Gagneux adds water and motion to the image, making it even more realistic). But the glycocalyx is also suggested to include self-organizing ordered lattices of glycans and lectins (83, 614615). An intriguing connection has been established between the glycan branching phenomena mentioned above and the formation of such cell surface lattices involving galectin recognition of polylactosaminoglycans, which tend to be enriched on highly branched glycans (117, 606, 608, 616). The concept is that an ordered lattice forms within the glycocalyx that alters interactions between cell surface molecules, also affecting their rates of clearance from the cell surface by endocytosis. Thus, evolutionary selection is suggested to have modulated the number of glycans of inhibitory versus activating growth factor receptors, such that branching (controlled by UDP-GlcNAc and GlcNAc transferases) can differentially affect their relative ratio on the cell surface, by altering cell surface residence times (606, 608, 616617). Complexity arises because capping of polylactosaminoglycans by sialic acids can modulate galectin recognition by its presence and or linkage type (618622).

Masking or modification of ligands for glycan-binding proteins

In some cases, modifications of monosaccharides and/or specific monosaccharides themselves can act as biological masks that prevent the recognition of the underlying glycan by specific glycan-binding proteins (623). Classic examples can be found in the case of terminal sialic acid wherein O-acetyl modifications can block the binding of some influenza viruses (22, 623624), and the removal of sialic acid itself can unmask binding sites for receptors or antibodies that recognize subterminal β-galactose residues (22, 623). In another example, certain enzymes called Sulfs mediate extracellular removal of binding sites for heparin sulfate ligands (625), which can then signal through other receptors, e.g., wnt/frizzled or IFN-β/IFNAR (626628).

Tuning a range of function

The size, number, branching and degree of sialylation of N-glycans can generate numerous glycoforms of a single polypeptide such as erythropoietin (629640) or GM-CSF (641644). It turns out that the nature of the glycosylation, extent of branching and level of sialylation modulate the activity of such cytokines over a range of function, by affecting its interaction with its cognate receptor, and also by altering the rate of clearance from the circulation. In passing, it is worth mentioning that differences in the sulfation and sialylation of the N-glycans expressed on endogenous versus exogenous erythropoietin are used by the Anti-Doping Agency to detect illicit administration, and has led to rescinding of many major sporting trophies (645646).

Molecular functional switching

The once obscure (310), but now well-known and widespread, O-GlcNAc modification of nuclear and cytoplasmic proteins, has been shown to be a multifunctional molecular switch, which can work with, or in competition against, Ser/Thr phosphorylation, altering the functions of a wide variety of modified proteins and affecting numerous physiological and pathological processes. This remarkable system and its numerous ramifications have been well reviewed elsewhere (21, 70, 647654), and will not be addressed in detail here. Other forms of nucleocytoplasmic glycosylation have since been discovered and characterized functionally in many organisms. For example, oxygen-sensing in diverse protozoa depends on prolyl-4-hydroxylation of the E3(SCF)ubiquitin ligase family subunit Skp1, and modification of the resulting hydroxyproline with a series of sugars. In the social amoeba Dictyostelium, O2 availability is rate limiting for hydroxylation of newly synthesized Skp1. Knockout mutants of the Skp1 prolyl hydroxylase and each of the Skp1 glycosyltransferases confirmed that O2-dependent post-translational glycosylation of Skp1 promotes association with F-box proteins and their engagement in functional E3(SCF)Ub ligases, which in turn regulate O2-dependent developmental progression (655659).

Returning to the extracellular compartment, another classic example is the modulation of IgG effector functions by the structural features of the N-glycans in the IgG-Fc region (660663). Incomplete galactosylation of these glycans has been associated with chronic inflammatory diseases (664666), and there are clear-cut effects of IgG-Fc glycan core fucosylation on antibody-dependent cellular cytotoxicity (667671) that are relevant to biotechnology (672673). Sialylation of a minor fraction of the IgG-Fc N-glycans also appears to convert the IgG molecule into an inhibitor of inflammation, and is thought to underlie the anti-inflammatory properties of therapeutically delivered intravenous immunoglobulin in humans. While many papers have been written about this effect, there is still some controversy about the extent of the effects, and the details of mechanisms in different models and species (674701). One possible explanation for the confusing results is that the immune responses are subject to “hormesis”. This is poorly appreciated but common biological phenomenon wherein low and high doses of the same stimulus can result in opposite biological responses and outcomes (702705). Regardless of controversies, it is clear that subtle changes in the glycosylation state of Ig-Fc regions of N-glycans can profoundly influence not only circulating half-life, but also the effector function of antibodies (706). Meanwhile, contrary to 40 years of X-ray crystallography suggesting immobility of the Fc-region N-glycan, recent NMR studies suggest that this glycan is actually mobile and dynamic in solution (707). Thus, it is possibly the range of motion of the glycan that is being altered by the various modifications, secondarily affecting interactions with Fc-receptors of various types via an allosteric mechanism (708). Finally, while much of what is stated above refers to IgGs prepared for therapeutic use, it appears that modulation of IgG-Fc glycosylation occurs naturally in vivo, in various inflammatory and infectious conditions (687) although the mechanisms of modulation are largely unknown. More recent evidence suggests that the Fc-region of other Ig classes may also alter effector function (709710).

Epigenetic histone modifications

It is now clear that the addition of O-GlcNAc residues to histone proteins surrounding chromosomal DNA is a key component of the histone code that regulates gene expression. O-GlcNAcylation targets key transcriptional and epigenetic regulators including RNA polymerase II, histones, histone deacetylase complexes and members of the Polycomb and Trithorax groups. Given its dependence on cytosolic UDP-GlcNAc levels, O-GlcNAc cycling is thought to serve as a homeostatic mechanism linking nutrient availability to higher-order chromatin organization. Evidence also suggests that this “simple” glycosylation mechanism can also influence X chromosome inactivation and genetic imprinting (650, 652653, 711), which may be related to the fact that the O-GlcNAc transferase is encoded on the X chromosome.

Extrinsic (interspecies) recognition of glycans

As mentioned earlier, it is not at all surprising that numerous pathogens and symbionts have evolved highly specific ways to recognize aspects of the dense and complex forest of cell surface glycans they encounter in host species. These interactions often involve glycan-binding proteins (see Figure Figure3),3), and can result in symbiosis, commensalism or disease, depending on the interaction in question and on the biological circumstances. A few examples from this vast field of knowledge are mentioned.

Bacterial, fungal and parasite adhesins

Among the numerous examples that can be cited for bacterial adhesins (223, 712), the example of Helicobacter recognition of gastric sialoglycans is particularly interesting, given its role in pathogenesis of gastric ulcers and cancers (713719). The F-pilus mediated glycan-dependent binding of uropathogenic Escherichia coli accounts for millions of urinary tract infections a year (720721) and small molecule inhibitors are being explored as therapies or prophylactics (722). With regard to parasites, a well-known example is the merozoite stage of Plasmodium falciparum, which initiates malaria via recognition of densely sialylated glycophorins on target erythrocytes (723730), with the types of sialic acids presented affecting species specificity (731733). At a later stage in malaria, heparan sulfate on endothelial cells mediates the binding of P. falciparum-infected erythrocytes via the DBL1ɑ domain of PfEMP1 (734), likely accounting for some of the most serious complications of the disease. Specificity for host glycans also plays a role in the binding of a pathogenic fungus (Candida glabrata) to various target tissues (735).

It is notable that in many instances expert researchers eventually find “glycan-independent” mechanisms of pathogen interaction with target cells and sometimes assume that the glycan-dependent process is therefore unimportant. However, such studies are often done in static conditions with long contact times, making the initial glycan “handshake” less critical. The situation is quite different in real life, where opportunities for contact and infection may be transient and difficult. Of course, with increasing evolutionary time a highly successful pathogen may come to rely more on glycan-independent mechanisms, as appears to be the case with endemic P. falciparum infections (727, 736738).

Viral agglutinins

By tradition, viral glycan-binding proteins are called hemagglutinins, because many were originally discovered by virtue of their ability to agglutinate erythrocytes (which ironically are noninfectable, because they do not have the machinery for viral replication). Of these the best known is probably influenza hemagglutinin (the “H” in “H1N1”), which plays a key role in the infection process of this highly successful group of viruses. Much has been written about the specificity of the binding of these pathogens in relation to the sialic acid ligand, particularly the specific linkage to the underlying sugar chain, which determines the preference of the virus for avian versus human hosts (739749). The evolution of the avian influenza viruses towards infecting humans involves selection for a change in binding specificity, which can be replicated experimentally (744). Interestingly, even our closest evolutionary cousins (chimpanzees) do not have a high density of human sialic acid composition or linkage on their airway epithelium (750), explaining the lack of nonhuman primate models and the unlikely choice of the ferret as a model for human influenza––because it happens to express the human-like linkage on its airway epithelium (751752), and also because it turns out to, like humans, be missing the nonhuman sialic acid Neu5Gc (753). There are other examples of even more exquisite sialic acid specificity of viruses, based on the presence of O-acetyl esters at specific positions: while a 4-O-acetyl ester on sialic acid targets is required for mouse hepatitis virus infection (754756), a 9-O-acetyl ester on the sialic acid side chain is required for the binding of certain other coronaviruses and influenza C and D viruses (624, 757761). The difference between these two specificities is dictated by only a few key amino acid changes in the viral receptors (762763).

Bacterial and plant toxins

Many soluble plant and bacterial toxins mediate their effects by binding to target glycans on cells of another species. Typically, a bacterial glycan-binding B subunit is multimeric and serves to bring the toxic A subunit close to the membrane, whereby the latter then crosses over to mediate its toxic actions in the cytosol (with or without prior endocytosis). Classic examples include cholera toxin, which binds GM1 ganglioside (256, 764), the infamous ricin toxin that binds to terminal β-linked galactose residues (765766), and the entero-hemorrhagic E. coli/Shiga verotoxin that recognizes globotriaosylceramide (Gb3Cer) and globotetraosylceramide (Gb4Cer) glycosphingolipids (767769). The precise spacing of target ligands can be very important to the optimal binding of the pentameric lectins to the target (770771). The single oxygen atom difference between the Neu5Ac and Neu5Gc forms of sialic acids can also determine the specificity of toxin binding, such as in the cases of the typhoid (772) and SubAB (773) toxins. In some cases, there is also evidence of dual specificity, e.g., fucosylated blood group structures on glycoproteins may contribute to cholera toxin binding (774), via an independent binding site (775).

Soluble host proteins that recognize pathogens

Vertebrates also express toxic glycan-recognizing peptides that can attack bacteria. For example, the small intestinal mucus layer is rich in RegIIIgamma, a secreted host antibacterial lectin, which is essential for maintaining partial sterility of a ~50-μm zone that physically separates the luminal microbes from the intestinal epithelial surface (776). Also, host galectins have been found to have unexpected toxicity towards bacteria via recognition of their surface glycans (777778). Killing occurs rapidly and independently of complement and is accompanied by disruption of membrane integrity. Galectin-3 may also play an important role in innate immunity against infection and colonization of Helicobacter pylori. (779). Galectin-1 can have dual and opposing effects on virus infection of human endothelial cells (780). In other instances, circulating soluble multimeric (typically pentameric) glycan-binding proteins recognize surface glycans of foreign pathogens but do not directly kill the pathogen. Instead they provide a signpost to attract other active components of the immune system such as complement and macrophages. Examples including collectins like the mannan-binding lectin, and the ficolins (334, 781). Indeed this kind of triggering of innate immune reactions via multivalent recognition of foreign glycans represents some of the most ancient and effective forms of immunity. For example, the hemolymph of horseshoe crabs recognizes invaders through a combinatorial approach, using lectins with different specificities against glycans exposed on pathogens (782), allowing these organisms to survive almost unchanged for >100 million years, without the benefit of adaptive immunity.

Pathogen glycosidases

Numerous pathogens generate a diverse array of cell surface and secreted glycosidases that serve to remodel or destroy the host glycocalyx, sometimes then utilizing the released monosaccharides as food sources and/or providing a nutritional resource for other microorganisms in the same milieu (783785). Some mammals also rely on symbiotic microorganisms within their digestive tract to gain energy from plant biomass that is resistant to mammalian digestive enzymes (786). In other instances, the glycosidase acts in a balance with the binding activity of the same pathogen (787). For example, the sialic acid-binding (“hemagglutinating”, H) activity of the influenza viruses is balanced by the activity of its sialic acid releasing enzyme (neuraminidase, N), the latter working both to allow the virus to gain access to cell surfaces by cutting through interfering molecules (788789), and also to allow release from cells after replication (790). The elegant structure-based design of a modified version of the previously known sialidase inhibitor Neu5Ac2en (791) gave rise to the potent and specific inhibitor zanamivir (Relenza) (792793), and later to the structurally related agent oseltamivir (Tamiflu) (794795). It is worth noting that oseltamivir is not a glycan, showing how chemical shapes can be designed to mimic glycans. In yet other cases, microbial glycosidases remodel host glycans to generate the optimal receptor for subsequent binding. For example the secreted neuraminidase of Vibrio cholerae removes all but one specific sialic acid residue from host gangliosides, leaving behind the GM1 monosialoganglioside, which is the specific receptor for the B subunit of the AB5 exotoxin secreted by the same organism (796).

Host decoys

It has been suggested that circulating erythrocytes might act as noninfectable decoy receptors for glycan-recognizing viruses that gain access to the bloodstream (164). The thick layer of mucin glycans on the surface of epithelial cells lining hollow organs (139, 418, 448452) also plays a critical role by providing decoy binding sites for pathogens, diverting them from their intended targets on the cells. Of course commensals may take advantage of such mucin binding to remain within their preferred ecological niche (797800) and to favor dental biofilm development (801). But on the rare occasions when such bacteria accidentally find their way into the bloodstream, these same commensalism-favoring adhesins become virulence factors, mediating interactions with platelets, which act as carriers of the organisms to eventual infection of damaged heart valves (798, 802807). This is a recurring theme at sites of interspecies interactions, wherein factors favoring routine commensalism turn into potent “virulence factors”, on the occasions when physical barriers are breached and/or host immunity is compromised.

Herd immunity

As discussed previously in the context of glycan evolution (164), herd immunity refers to a form of indirect protection from infectious disease, which occurs when a large percentage of a population is resistant to an infectious agent, effectively providing protection to individuals who are not immune. Since glycans are often the targets for many infectious agents, intra- and interspecies polymorphisms in the expression of such targets can provide herd immunity, and restrict the spread of disease. As an example the ABO(H) blood group system can affect the spread of a highly infectious noroviruses that selectively binds one blood group structure and not another (808811). This is likely why not everyone is sick by the time a cruise ship suffering a norovirus epidemic makes it back to port. ABO blood group polymorphisms also appear to affect susceptibility to cholera, as the cholera toxin has a secondary binding site for such glycans (775). The high levels of competitive oligosaccharides in human milk likely provide protection to breast-fed infants against some viruses and toxins (812814).

Pathogen-associated molecular patterns

It is now well recognized that innate immune cells also detect pathogen-associated molecular patterns (PAMPs) using Pattern Recognition Receptors (815816), particularly Toll-like receptors (TLRs) (817818), NOD-like receptors (NLRs) (819822) and C-type lectins (823826). Many PAMPs are glycoconjugates, e.g., bacterial lipo-oligosaccharides or glycan-based polymers, e.g., lipopolysaccharides and bacterial peptidoglycans, including bacterial DNA or viral RNA (which are (deoxy)ribose-based polymers) (827828). Glucan and oligochitin oligosaccharides released from fungal cell walls can also function as elicitors of plant defense (829).

Immune modulation of host by symbiont/parasite

In some instances, glycan molecules mediate symbiont or parasite modulation of host immune responses. For example, glycans such as polysaccharide A (an unusual pentasaccharide repeat), derived from important mammalian gut microbiome members, helps to modulate the host immune system into a more tolerant state (via T-reg engagement) (213). Similarly, glycans derived from parasitic worms alter the immune status of their long-term hosts (830), a process dubbed as “glycan gimmickry” (831).

Antigen recognition, uptake and processing

Antigenic proteins must first be taken up by antigen presenting cells (macrophages and especially dendritic cells), which process them into peptides, to be presented by MHC Class II molecules, for recognition by T lymphocytes. This process can be facilitated by glycans on the target protein. For example, the presence of high densities of terminal Man or GlcNAc residues on foreign proteins or microbes can trigger phagocytosis via C-type lectins on antigen presenting cells, with resulting delivery of the antigenic proteins to processing compartments (832834). As another example, nonhuman ɑ-Gal (835) or Neu5Gc (836) residues carried on injected glycoproteins can result in immune reactions, and in the formation of immune complexes, which in turn can enhance immune reactivity against the peptide backbone. An alternative form of self/nonself recognition is exemplified by foreign glycolipid presentation by CD1 molecules, which are detected by restricted or invariant TCRs of NKT cells (837842).

Bacteriophage recognition of glycan targets

The complexity and diversity of surface polysaccharides found on strains of a single bacterial species (>100 in the case of the pneumococcus) (843) might be explained not only by selection for evasion of the vertebrate antibody response (844), but also by the need to evade attack by environmental bacteriophages (843, 845), which often use bacteria surface glycans as targets for recognition, and sometimes subsequent cleavage. While studies are continuing (846849), this remains a poorly explored area. Given the very high density of bacteriophages in nature (10 million viruses per milliliter of surface seawater!) (850), there are likely a huge number of such as yet undiscovered viruses with exquisite specificities for diverse glycan structures. Indeed, it is possible that a cognate bacteriophage exists for every variant of every bacterial surface polysaccharide that occurs in nature. Thus bacteriophages are effectively a massive reservoir of glycan-binding and glycan-hydrolyzing proteins still waiting to be exploited for glycan analysis and bacterial diagnostics as well as therapeutics (851), e.g., a potential new source of therapeutic “enzybiotics” (852) or disrupters of biofilms (853). Early steps in this kind of systematic search are promising (854).

Intrinsic (intraspecies) recognition of glycans

As mentioned earlier, numerous pathogens and symbionts have evolved highly specific glycan-binding proteins that can recognize aspects of the cell surface glycans they encounter in host species. For a long time, examples of glycan-binding proteins with clear-cut functions intrinsic to the same species (see Figure Figure3)3) proved elusive. Even when candidates such as the asialoglycoprotein receptor were found, their intrinsic functions were not obvious. Beginning with the discovery of the specific functions of P-type lectins in mediating lysosomal enzyme trafficking (discussed earlier), many examples of glycan-binding proteins with intrinsic functions are now well known, and participate in a wide variety of functions. Only a few examples are mentioned below.

Intracellular glycoprotein folding and degradation

In addition to the biophysical effects of attached glycans on nascent glycoprotein folding discussed above, specific recognition of certain glycan residues plays a key role regulating the process of ER-associated degradation (ERAD). After the unusual Glc3Man9GlcNAc2-P-P-dolichol structure of the lipid-linked oligosaccharide donor for N-glycosylation was first fully defined (855), it turned out to be identical in almost all eukaryotes studied. Conservation of this structure for more than a billion years of evolution strongly suggested that it serves a very important purpose. But while many features of the structure were clearly needed to ensure optimal N-glycan transfer (856859), variations were possible in some parasites (860861) and mutant cell lines (235), and the exquisite conservation of the native structure remained largely unexplained. A clue finally emerged when it was discovered that the third glucose residue on N-glycans is repeatedly removed and then put back on again during glycoprotein folding in the ER (862863). This in turn led to the discovery that this terminal glucose residue is recognized by certain ER chaperones, calnexin and calreticulin (864868). The key role of this glucosylation/deglucosylation cycle in protein folding is now well established (63, 71, 869878). However, even after the last glucose residue has been permanently removed, there are further steps of recognition of the oligomannose type N-glycans that have been partially processed by ER mannosidases (879884). These recognition events are mediated in part by mannose 6-phosphate receptor homology domains in several chaperone proteins, as well as additional mannosidase-like proteins and recognition complexes (130, 879, 881888). Effectively, a byzantine array of glycan-modifying and glycan-recognizing proteins determines the final fate of a glycoprotein molecule in the ER––whether it will be allowed to go forward into the Golgi pathway towards its final destination, or be consigned for ERAD. And since most proteins that enter the ER are glycosylated, this system has a huge impact in normal and diseased states, as well as on unfolded protein stress responses (883, 889). As mentioned earlier, O-mannosylation and O-fucosylation can also play a role in monitoring the folding of newly synthesized proteins. Proteins that fail to fold are eventually removed from harmful futile protein folding cycles and prepared for disposal, via reverse translocation into the cytosol. There is even a sophisticated cytosolic pathway for removing and recycling the N-glycans from misfolded proteins prior to the action of proteasomes, beginning with the action of a previously mysterious cytosolic Peptide: N-glycanase (890). Notably O-GlcNAcylation of nucleocytoplasmic proteins can also occur cotranslationally, protecting nascent polypeptide chains from premature degradation by decreasing cotranslational ubiquitinylation (891).

Intracellular glycoprotein trafficking

As discussed earlier, the classic example of glycan roles in intracellular trafficking is that of the mannose 6-phosphate recognition system for the targeting of lysosomal enzymes to lysosomes. There is now evidence for other lectin-like molecules within the ER-Golgi pathway, which likely modulate the trafficking of specific classes of glycoproteins. For example, the LMAN1 gene product ERGIC-53 in the ER-Golgi intermediate compartment is a mannose-selective and calcium-dependent human homolog of leguminous L-type lectins (892) and acts as a critical chaperone for the coagulation factors V and VIII during their biosynthesis in hepatocytes (893) and endothelial cells (894) respectively, also potentially affecting the biosynthesis of some other glycoproteins (895). Other examples are VIPL and VIP36 (896901). Overall, it is reasonable to predict the existence of more such glycan-recognizing proteins in the ER-Golgi pathway, potentially involved in trafficking and/or chaperone functions.

Triggering of endocytosis and phagocytosis

A variety of cell surface receptors that recognize terminal glycans can trigger uptake of molecules (endocytosis), particles (phagocytosis) or even intact cells. The classic examples are the asialoglycoprotein receptor of hepatocytes and the mannose receptor of macrophages, mentioned in the introduction. A large variety of lectins are known to carry out endocytosis in macrophages and dendritic cells. Such recognition processes may be critical not only for providing antigens to process and present to T cells, but also for clearing away damaged cells or glycoproteins, such as occurs when microbial sialidases enter the circulation during sepsis and cause desialylation of platelets (902903), or when cancers secrete incompletely glycosylated mucins (512).

Intercellular signaling

In addition to the plant oligosaccharins already mentioned in the introduction, oligogalacturonides released from pectins can also act as regulators in plants (141, 904906). Another well-established example of intercellular signaling is represented by bacterial Nod factors (820) (not to be confused with the NLRs of vertebrate inflammasomes), which communicate signals between rhizobacteria and the roots of their leguminous plant hosts (907911), initiating the symbiosis that is eventually responsible for the bulk of the natural nitrogen fixation on the planet––a process key to the survival of many organisms that benefit from the resulting food chain. The chito-oligosaccharides that transmit the signal show structural specificity for organism and host (907908) and appear to be detected by a specific lectin (911). Transferring this capability into nonlegume crop species is obviously an exciting prospect. However, this may not be easy, since nonlegumes recognize the Nod factor via a mechanism that results in strong suppression of responses (912). In vertebrate systems hyaluronan fragments released during injury can be detected by TLRs, thus triggering host immune responses (467, 913915).

Intercellular adhesion

The mechanism of species-specific recognition of disaggregated sponge cells has already been mentioned in the introductory sections. The selectin-based system for cell–cell interactions of leukocytes, platelets and endothelial cells has also been discussed. The fact that oral fucose feeding results in correction of leukocyte adhesion deficiency-II by restoring selectin ligands provides the genetic proof of concept of this system in humans (385). The role of selectin interactions in a variety of normal and pathological conditions like inflammation and cancer is now understood (309, 916921), and therapeutic approaches are in evaluation (921923). One particularly promising therapeutic outcome is based on the finding that selectins interact with sickled red cells and leukocytes in the circulation to facilitate endothelial adhesion and other interactions (924927) ultimately contributing to vascular occlusion and “sickle cell crisis” (928). The effectiveness of the pan-selectin inhibitor GMI-1070 in reducing selectin-mediated cell adhesion and abrogating crisis shows much promise in early clinical trials (922, 929). Another classic example is the role of Myelin-associated glycoprotein (MAG, Siglec-4) in mediating key interactions between neurons and glia (930932), a process critical for maintaining the stability of the myelin sheath that insulates axons (933934).

Cell–matrix interactions

Evidence for critical matrix interactions with cell surface glycans can be found in the variety of muscular dystrophies resulting from altered glycosylation of the α-dystroglycan ligand for major matrix proteins such as laminin, described in the introduction (191204). In another example, hyaluronan matrices synthesized by stressed cells that recruit inflammatory cells are early events in many pathological processes (935936). Interestingly, this is a process that most if not all cells undergo, when dividing in a hyperglycemic environment. This phenomenon likely impacts experiments of many investigators who use “standard” commercial tissue culture media, which actually have unphysiologically high amounts of glucose (937), at levels that might even cause diabetic coma in a patient.

Fertilization and reproduction

Many early studies suggested that glycan-recognition processes were a critical part of many sperm–egg interactions (75, 938939). This field lagged behind for a while, partly because many researchers were looking for a single overarching glycan-recognition mechanism––until the realization that species-specific variations were in fact, exactly what one would expect! In a few instances such as in humans, specific glycans have now been identified as binding targets (940941). Glycans also appear to be involved at many steps in the reproductive process, and in the processes of sperm migration to the site of fertilization (942). During the latter process, there is even evidence that circulating antibodies can enter the uterine fluid and destroy sperm carrying nonspecies-specific glycan antigens (943). After fertilization, there is evidence that glycans and glycan-binding proteins are involved in the processes of implantation (944) and placental functions (945946) in mammals.

Clearance of damaged glycoconjugates and cells

Terminal sialic acids on circulating glycoproteins can be removed by endogenous sialidases during natural aging of the proteins (947), or suffer an attack by a pathogen expressing a sialidase. In either case, there would be exposure of underlying glycans recognized by specific receptors, such as the hepatocyte asialoglycoprotein receptor mentioned earlier. Data indicate that this kind of “eat me” signal may even mitigate the lethal coagulopathy of sepsis by clearing away damaged platelets (902903). There also appears to be a very high-capacity system for clearance of incompletely glycosylated mucins by the liver (512). Such molecules are released in large amounts by cancer cells, but could also potentially appear during damage to otherwise healthy organs. The subset of cancer-derived molecules (e.g., CA125, Sialyl-Tn and CA19-9) that survive such clearance then become useful markers of disease progression (948951). The value of such markers for early detection remains unclear (952), a problem known to plague many predictive serum markers (953).

Glycans as clearance receptors

Glycans themselves can act as clearance receptors for other molecules. For example, heparan sulfate proteoglycans in the liver space of Disse mediate clearance of triglyceride-rich lipoproteins independently of the well-known LDL receptor family members (954955).

Danger-associated molecular patterns 

Innate immune cells also recognize glycans released from tissue damage in vertebrates such as hyaluronan fragments (467, 914915) and some matrix proteoglycans (956957) as danger-associated molecular patterns (DAMPs) or “alarmins”, triggering responses similar to those generated by exogenous PAMPs (see earlier discussion). While fungal glycans can act as PAMPs to activate the host immune response, they can also instead mask other glycoconjugates to prevent such activation. Examples include Candida albicans (443, 958959) and Histoplasma capsulatum (960).

Self-associated molecular patterns

As mentioned above, signals initiated by DAMPs and PAMPs are transduced via similar pathways, activating innate immune inflammatory responses. It was recently proposed that glycans could also act as self-associated molecular patterns (SAMPs) (961), being recognized by intrinsic inhibitory receptors to maintain the baseline nonactivated state of innate immune cells, and to dampen their reactivity following an immune response. A clear example of glycan-based SAMPs has been reported in the form of inhibitory Siglec recognition of cell surface sialoglycans (962964), which may also provide a mechanism for the host to discriminate between infectious nonself and noninfectious self (965). Recent work (157, 966968) has also affirmed prior evidence that sialoglycan recognition by factor H can blunt immune responses by inhibiting the alternate pathway of complement activation (966, 969). Siglec-9 recognition of hyaluronan may be another example of a SAMP system (970). Not surprisingly, these very same self-glycans are also common candidates for molecular mimicry by commensals or pathogens that engage these inhibitory receptors (see below).

Antigenic epitopes

In addition to the blood groups already mentioned, intra- and interspecies variations in glycosylation can result in strongly antigenic epitopes. Indeed a significant fraction of circulating Ig found in normal humans may be directed against foreign glycan antigens (971973). Certain types of modifications of N-glycans found on plant and invertebrate glycoproteins can trigger immune reactions in humans (974977), including therapeutic glycoproteins (835836). In a more complex scenario, individuals exposed to Lone Star tick bites seem to develop IgE antibodies against ɑ-Gal epitopes (humans do not have these epitopes). Upon subsequent exposure to mammalian foods that are rich in ɑ-Gal motifs (such as red meats), individuals react (sometimes severely) in an apparent “red meat allergy” (978979). On another practical note, glycans like ɑ-Gal and the nonhuman sialic acid Neu5Gc represent the major xenoantigens that must be bypassed, to pursue the goal of xeno- (pig organ) transplantation into humans. In pursuit of this difficult goal, ɑ-Gal- and Neu5Gc-double null pigs have recently been generated (980983).

Xeno-autoantigens

It has recently been found that the nonhuman sialic acid Neu5Gc can become metabolically incorporated from dietary sources (particularly red meat) into certain cell types in the body, appearing on the surfaces of human cells as if it was synthesized by the individual (984985). These “xeno-autoantigens” are recognized by pre-existing circulating “xeno-autoantibodies”, and the resulting “xenosialitis” is suggested as one mechanism for the epidemiological association between red meat consumption and the exacerbation of some common disease states, such as carcinomas and complications of atherosclerosis (984985). It would not be surprising if other examples exist. One can imagine for example that bacterial nonulosonic acids or plant monosaccharides that are structurally related to host monosaccharides might get activated to their corresponding nucleotide sugars and then get transferred onto endogenous glycans at a low rate. While the efficiency of such a process would likely be lower than that of Neu5Gc, the resulting immune responses might be even stronger.

Molecular mimicry of host glycans

Given that the host immune system recognizes typical glycans found on many pathogens are PAMPs and that endogenous glycans function as SAMPs, it is not surprising that microorganisms have evolved ways to achieve molecular mimicry of host glycans. What is remarkable is the striking extent to which such mimicry has been achieved, via all imaginable mechanisms. Just a few examples are cited here, with an emphasis on sialoglycan mimicry by vertebrate pathogens.

Convergent evolution of host-like glycans

It was originally thought that pathogen molecular mimicry was being achieved via vertebrate to bacterial gene transfer. While there is continued controversy about the extent of horizontal gene transfer between prokaryotes and eukaryotes (986), most instances of glycan molecular mimicry by pathogens seem to involve convergent evolution of pre-existing pathogen biosynthetic pathways, or de novo generation of functional genes. Demonstrating the power of natural selection at the host-pathogen interface, Group B Streptococcus polysaccharides (987) display identity to specifics of host glycan structure such as the Neu5Acα2-3Galβ1-4GlcNAcβ1- (which perfectly matches the structure of N-glycan antennae on many human glycoproteins), and Campylobacter species carry out near-perfect mimicking of complex brain ganglioside glycans (988989). In the former case, it is evident that this mimicry allows the organism to imitate endogenous SAMPs and down-regulate innate immune responses by engaging the inhibitory Siglecs (962). In the latter instance rare human immune responses against the ganglioside-like structures can even result in serious illness, with the complement-fixing antibodies damaging peripheral nerves (Guillain-Barré syndrome) (990991). While Campylobacter sialylation may also modulate immune responses via Siglecs during sporadic contacts with humans (992995), it is unclear why the organism (which normally lives in the chicken intestine) (996) has evolved this remarkable degree of molecular mimicry of vertebrate ganglioside. Perhaps there are Siglec-like inhibitory pathways in the chicken that have yet to be discovered.?

Appropriation of host glycans

Continuing with the example of sialic acids as host molecular mimics, microorganisms seem to have evolved every other conceivable mechanism to achieve this goal. These mechanisms range from the simple acquisition of host sialoglycans (997) to the direct transfer of host sialic acids by trans-sialidases (998), to the highly efficient uptake of the small amounts of environmental free sialic acids (999) or even the direct utilization of trace amounts of CMP sialic acid present in host body fluids (968, 1000). In addition to acting as SAMPs recognized by Siglecs or limiting complement activation via factor H recruitment, such terminal sialic acids also serve to mask antibody recognition of underlying structures. The fact that numerous organisms have independently evolved so many different ways to decorate themselves with host-like sialoglycans (1001) speaks to the strong selection pressure for this mimicry.

As with examples mentioned earlier, these “virulence factors” may actually represent attempts at commensalism and symbiosis, which become pathological in some circumstances. Many other examples of host glycan mimicry can be cited (1002), such as the bacterial re-invention of hyaluronan (163, 1003), heparosan (the backbone of heparan sulfate) and chondroitin (the backbone of chondroitin sulfate) (10041005). Interestingly, there appear to be limits to the “inventiveness” (constraints to convergent evolution) of microorganisms. Despite hundreds of millions of years of selection, no prokaryotes seem to have reinvented sulfation of glycosaminoglycan backbones, nor recreated the difficult biosynthesis of the nonhuman sialic acid Neu5Gc. While not exactly full-blown “mimicry”, most viruses simply take over and use the host glycosylation machinery to install glycans that mask and protect them from immune destruction. Examples of molecular mimicry by pathogens of plants or invertebrates need to be further investigated.

Multifunctional roles of the single type of glycan

The above classification of biological roles falls apart when one considers certain glycan molecules that can mediate many different types of roles, depending on the circumstance. An example is the lipophosphoglycan of Leishmania species, which is needed for establishment of initial infections in vertebrate hosts, and not for persistence or pathology (1006)––but is later needed for binding to galectins located on the surface of midguts of their invertebrate vectors (1007). A more striking example is the myriad functions of heparan sulfate proteoglycans, with different roles being mediated by slight modifications of the molecule (155, 425, 427, 572, 628, 10081011). This is also exemplified by the widely disparate phenotypes arising from genetic modifications in various steps involved in biosynthesis of the molecule. Thus for example, mice deficient in heparan sulfate 6-O-sulfotransferase-1 exhibit defective heparan sulfate biosynthesis, abnormal placentation and late embryonic lethality (1012) and autism-like socio-communicative deficits and stereotypes appear in mice lacking heparan sulfate only in the brain (1013).

Even a structurally simple molecule like polysialic acid (an ɑ2-8-linked homopolymer of N-acetylneuraminic acid) can have a remarkable range of endogenous functions. For example, polySia has been implicated in numerous normal and pathological processes and phenotypes, including cell migration (10141016); cell differentiation (1014, 1016); neurite outgrowth (10171019); blockade of myelination (10201022); binding and modulation of neurotrophin function (559, 10231025); alteration of synaptic plasticity (10261028); effects on learning and memory (10291032); facilitation of repair following injury (546547, 10331036); schizophrenia pathogenesis (391, 10371042); major depression (553, 10431045); bipolar disorder (391392, 1044); alcoholism (1046); epilepsy (10471048) and social interaction (1049).

Some questions and issues arising

Some readers will likely feel that the examples of biological functions discussed are not the most striking ones, and others will doubtless complain that numerous additional functions have not been mentioned. Such deficiencies and limitations are simply an indication of how far the field has come in the last 20+ years since the last review in 1993. Let us conclude this incomplete attempt by discussing some questions and issues arising, and some future prospects.

Why did glycans become the preferred cell surface covering during evolution?

With the possible exception of transient bloodstream phase of certain parasites, there appears to be no exception thus far to the “rule” that the surfaces of all cells in nature are covered with a dense and complex coating of glycans, which is taxon-, species- and cell-type specific (178). If it had been biologically possible to evolve a living cell devoid of such a coating, such a cell would have no doubt emerged from >3 billion years of evolutionary selection. There is no single best explanation for this ubiquity of cell surface glycans, and several mutually nonexclusive ones can be considered. In addition to providing a physical barrier to protect the plasma membrane, glycans tend to be hydrophilic, often do not have rigid structures, and instead have significant freedom of motion in aqueous solution. These are the optimal properties for a class of molecules that interact at the interface with an aqueous environment. Also, it is difficult for cells coated only with proteins to evolve and escape mechanism from a pathogen that binds to a specific cell surface protein. Most amino acid changes are not usually well tolerated by proteins, impacting folding and/or stability or even rendering the molecule dysfunctional. In contrast, most intrinsic glycan functions are mediated not by a single absolutely required sequence, but by an ensemble of structures, spanning a continuum. Even the apparently “lock-and-key” example of Man-6-P recognition of lysosomal enzyme N-glycans discussed earlier actually involves a spectrum of Man-6-P bearing structures with a range of binding properties to two different M6PRs. In other words, many glycan functions are “analog” not “digital”. Thus, it is easier for a host to escape pathogens by subtly changing glycosylation (i.e., glycans may convey more robustness to the organism) without drastically altering intrinsic functions. Last but not least, a vastly greater number of structural variations can be generated via monosaccharide oligomerization and branching in comparison with nucleic acids or amino acids (1050). This increases the odds of evolutionary selection to escape from a cell surface interacting pathogen or toxin.

Red Queen effects in glycan evolution?

Given the above considerations, it is reasonable to suggest that glycans are particularly prone to Red Queen effects (running to stay in one place) (164). As illustrated in Figure Figure4,4, one can envisage several such effects involving glycan interactions, leading to a delicate balance between preserving endogenous function and evading pathogen attack (170). A more nuanced and sophisticated view that takes into account additional evolutionary considerations can be found in Figure Figure55 (173).

An external file that holds a picture, illustration, etc.
Object name is cww086f04.jpg

Red Queen effects in the evolutionary diversification of glycans. Each arrowed circle represents a potential evolutionary vicious cycle, driven by a Red Queen effect, in which hosts are constantly trying to evade the more rapidly evolving pathogens that infect them. Hosts require glycans for critical cellular functions but must constantly change them to evade glycan-binding pathogens, and yet do so without impairing their own fitness. Hosts also produce soluble glycans such as mucins, which act as decoys to divert pathogens from cell surfaces; but pathogens are constantly adjusting to these defenses. Hosts recognize pathogen-specific glycans as markers of “non-self,” but pathogens can modify their glycans to more closely mimic host glycans. There are also possible secondary Red Queen effects involving host glycan-binding proteins that recognize “self”. In each of these cycles, hosts with altered glycans that can still carry out adequate cellular functions are most likely to survive. Reproduced with permission from Varki A. 2006. Cell. 126:841–845. Copyright Elsevier.

An external file that holds a picture, illustration, etc.
Object name is cww086f05.jpg

Evolutionary conflicts between alleles and individuals. For single allele-single individual, single alleles conflict with themselves when their positive effects in one context cause negative effects in another. Some examples are here. Selectins on epithelial cells bind glycans on leukocytes and guide them to sites of inflammation, but this can also be exploited by cancer cells. Regulatory or functional changes that separate conflicting tasks are expected to evolve in response. For single allele-multiple individuals, conflicts can extend across individuals that share an allele. Females that lack Neu5Gc raise antibodies against it. Males that lack Neu5Gc have higher rates of fertilization, and females have lower rates. Individual-specific regulation could resolve these conflicts. For multiple genes-single individual, selfish alleles can bias reproduction in their favor at the cost of individual reproduction, causing conflict with other genes in the genome. Mutant alleles that favor heterozygotes are passed more often than expected but increase the risk of congenital disorders of glycosylation. Other genes are selected to suppress the selfish allele, often by modification of chromosomal recombination and linkage. For multiple genes-multiple individuals, molecular markers of self cause cells to direct benefits toward identical genetic relatives, but they can be exploited by pathogen mimics. Co-evolution is a common outcome, as hosts develop more reliable markers of self, and pathogens develop more effective molecular mimics. Reproduced with permission from Springer and Gagneux, 2013, J Biol Chem, 288:904–6911.

What is the significance of lineage-specific deletions or additions of specific glycans?

In contrast to the genetic code, there are many more species-specific variations in glycans, ranging from entire classes of glycoconjugates, e.g., sulfated glycosaminoglycans not found in prokaryotes, to specific glycans, e.g., the absence of ɑ-Gal epitopes in old world monkeys (1051), and the independent losses of the sialic acid Neu5Gc in humans (1052), new world monkeys (1053), mustelids and related taxa (753), and apparently in sauropsids (the ancestors of birds and reptiles) (1054). In every instance of apparent lineage-specific loss, a careful phylogenetic analysis is needed to ascertain if the differences are due to gain or loss of a specific gene or pathway and/or due to convergent evolution (or particularly in the case of prokaryotes, horizontal gene transfer). Regardless of the underlying mechanisms, more studies are needed to understand not only the functions of taxon-specific glycans but also the biological significance of their loss in some lineages. Some data suggest that taxon-specific glycan losses may have played a role in protection from parasites like malaria (732, 1055), and even in speciation events, such as the origin of the genus Homo (943).

Why did evolution select O-GlcNAc as the dominant form of intracellular eukaryotic glycosylation?

In striking contrast to the bewildering diversity of extracellular glycosylation, there seems to be a limited number of forms of intracellular glycosylation, with a single modification (O-GlcNAc) numerically dominating the scene (647648). One possible explanation is that this intracellular environment is not subject to selection pressures by myriad pathogens that express diverse and specific glycan-binding proteins. More specifically with regard to O-GlcNAc it has been suggested that the donor molecule UDP-GlcNAc acts as an optimal metabolic sensor for multiple pathways, i.e., uridine, phosphate, glucose, nitrogen and acetate (649, 651653).

Do free oligosaccharides in the cytosol and serum have specific functions?

During the N-glycosylation of glycoproteins in the ER, considerable amounts of unconjugated polymannose-type glycans are generated from breakdown of the lipid-linked precursor (1056). Later, misfolded glycoproteins that are returned to the cytosol for proteasomal degradation are first subject to a cytosolic PNGase enzyme that releases free oligosaccharides (890). Such free oligosaccharides are then subject to either further cytosolic catabolism or pumped back into lysosomes for degradation (1056). Even free complex N-glycans bearing sialic acids can also be found in the cytosol (1057). The question arises as to whether such glycans mediate any specific functions in the nucleocytosolic compartment, before they are degraded, such affecting transcription. Meanwhile, free sialyloligosaccharides related to N-glycans have recently been found in serum (1058), and may also have novel functions yet to be discovered.

Can a glycan-binding protein recognize more than one class of glycan?

Glycan-binding proteins are often discovered based on their recognition properties, and given names related to their initially discovered binding targets, e.g., Galectins bind β-galactosides (323), and Siglecs recognize sialic acids (329). However, many examples have emerged wherein well-known glycan-binding proteins are discovered to also bind to unrelated glycan class, sometimes not even obviously similar in structure. In some instances, this is simply because the protein in question has two distinct binding modules, e.g., the L-type lectin mannose receptor can also have a separate R-type lectin module that recognizes sulfated GalNAc residues on pituitary glycoprotein hormones (1059). However, in many other cases, the binding region is shared or very close by. Thus for example, selectins that were originally defined by their binding to sialylated fucosylated glycans can bind quite well to certain subsets of heparan sulfate glycosaminoglycans (10601061). Likewise, Fibroblast Growth Factor-2 can bind both to polysialic acid and heparan sulfate (1025), and Siglec-9 binds both sialic acids and hyaluronan (970). Recently, it has even been shown that some galectins bind efficiently to as yet undefined motifs on bacterial surfaces (777) as well as to some sulfated glycosaminoglycans (1062), in a manner still inhibitable by its canonical ligand lactose. In most such instances, it is unclear what the shared glycan motif is. Given great dissimilarities in primary structure, cross-recognition perhaps arises from “clustered patch” combinations (146, 302) of components of more than one monosaccharide, such as hydroxyl, carboxyl, sulfate, acetyl groups, etc. Given existing difficulties in efficiently incorporating even small, defined cognate ligands into the binding pockets of crystal structures of glycan-binding proteins, it will be challenging to compare such disparate ligands and define the shared recognition components. Regardless, it is clear biological functions of a glycan-binding protein should not be assumed to be mediated by the canonical glycan ligand class that originally defined the name of the protein.

How many more glycan-binding proteins are there yet to be discovered?

Regarding extrinsic (interspecies) binding proteins, it is reasonable to predict that for every specific glycan found on the cell surfaces of a host there is somewhere, a pathogen or symbiont that has developed an exquisitely specific binding protein for the glycan. Indeed, if this vast array of binding proteins could be isolated and converted into useful probes, one could have a new approach to “glycomics”, which actually studies the intact (naturalistic) glycome, in a manner that is exactly as it is “seen” by binding proteins in nature. The first steps in this direction have already been taken, and the results are very promising (10631066).

With regard to intrinsic (intraspecies) binding proteins, the situation is less clear. However, the serendipitous mechanism by which many of them have been discovered suggests that a systematic approach to future discovery may be useful. Consider the case of sialic acid-binding proteins. As late as the 1970s, it was thought that sialic acids were just biological masks, and that there were no binding proteins intrinsic to the organism synthesizing them. Notably of the few sialic acid-binding proteins reported since then, i.e., Factor H (10671068), Selectins (286), Siglecs (324325), PILRs (1069) and PECAM-1 (10701071), almost all were discovered serendipitously to recognize sialic acid, based on an unexpected loss of a functional readout upon sialidase treatment. Given that sialic acids have been present on the glycocalyx of the Deuterostome lineage of animals for more than 500 million years, it would not be surprising if there are many more as yet undiscovered sialic acid-binding properties of other already known proteins. The same is likely to be true for other classes of glycans, especially terminal and exposed structures. On the other hand, given relatively low single-site-binding affinities, a systematic approach to discovering such proteins may not be trivial. Sialoglycan array studies recently revealed the sialic acid-binding properties of M-ficolin (1072). On a cautionary note, the very high density of targets in glycan arrays might also detect binding specificities that may not exist in nature.

Is glycan recognition by proteins really of “low affinity”?

Compared to protein–protein interactions that typically have measured binding affinities in the nanomolar range, studies of important glycan–protein interactions usually give values in the micromolar, or even millimolar range. While there can be a high degree of recognition specificity, the single-site affinity is typically poor. Various reasons are discussed, and this general observation underscores the frequent need for multivalent avidity, in order to generate effective biological functions or effective experimental probes (1073). Of course, multivalency is the general state of most biology at the cell surface as nothing is present in only one copy. Thus effective affinity in nature is actually quite high. Regardless, in reality most glycans in aqueous solution are in constant motion and constitute an ensemble of many different shapes generated by many mobile bond angles, which are constantly interchanging (10741075). In order for the more rigid and ordered binding pockets of glycan-binding proteins to bind such “shape-shifting” glycans, they must actually “trap” one of the numerous possible solution conformations of the cognate glycan into the pocket, where the immobilized glycan can be seen in a crystal structure. Strictly speaking then the effective concentration of the true cognate glycan is far lower than that of the total glycan concentration, likely in the nanomolar range. Exceptions may arise when the glycan targets are restricted in their motion, forming “clustered saccharide patches”, such as on the surface of densely glycosylated cells, mucins or viruses (146, 544).

Can we better define and name specific glycoform ligands for glycan-binding proteins?

In some cases, the natural ligands for glycan-binding proteins can be defined by the primary sequence of the cognate glycan, e.g., the Sambucus nigra agglutinin binds the motif Siaα2-6Gal(NAc). However, in other instances the ligand is a specific glycoform of a particular glycoprotein, which is difficult to define in terms of a cognate glycan sequence. This results in the inadvertently erroneous statements, implying that the polypeptide is the ligand, e.g., “PSGL-1 is the ligand for P-selectin” (1076); or “CD24 is the ligand for Siglec-10” (965) (PSGL-1 and CD24 are actually the designated names of the core polypeptides). It also leads to incorrect assumptions, e.g., that a glycoform of CD24 must automatically be the ligand for mouse Siglec-G, the mouse ortholog of human Siglec-10 (965). In these and many other such instances, the ligand is actually a specific glycoform of the named polypeptide, and is only synthesized by certain cell types with the right kind of glycosylation machinery, i.e., the same carrier polypeptide does not serve as a ligand in other situations. In the case of a CD44 glycoform from hematopoietic stem cells that is a specific ligand for E- and L-selectin (10771078), the authors reasonably chose to rename the molecule altogether as HCELL (hematopoietic cell E- and L-selectin ligand) (921). However, this leaves out the useful information that the underlying polypeptide is CD44. A compromise may be to list the name of the polypeptide and use the superscript to indicate that it is a specific glycoform that generated the ligand in question, e.g., CD44HCELL or HCELLCD44. As with most nomenclature issues, it may be hard to find consensus on this matter. Suffice it to say that there is need for a resolution, to make it easier to understand the literature on ligands for glycan-binding proteins.

Why has it taken so long to elucidate biological roles of glycans?

It is clear that glycans got “left out” of the initial phase of the molecular biology revolution of the 1980s, not only because they were more complex and difficult to study, but also because they were not part of the original “central dogma” (1079). This resulted in a peculiar distortion of the bioscience community, in which an entire generation of biologists (beginning in the 1980s) has been trained without much knowledge or appreciation about the structure, biosynthesis and roles of glycans in nature (1080). The relative lack of interest in glycans can also be partly traced to the early lack of understanding of their biological roles. So, why was it so difficult to elucidate biological roles of glycans? Some of the reasons are obvious, such as the technical difficulties in detection, analysis and manipulation in biological systems. Some additional considerations are outlined in Figure Figure6.6. Because of the information embedded in the template-driven biosynthesis of nucleic acids and proteins, it has been relatively easy to go from one to the other, using sophisticated yet facile experimental methods, and via bioinformatic predictions. Also as shown in the upper panel of Figure Figure6,6, the path to defining a specific function as being mediated by a specific protein has been relatively straightforward. In striking contrast the field of glycosciences originated in “descriptive” carbohydrate chemistry and biochemistry and remained in these domains for a long time. New glycans were discovered by a variety of means (such as those shown in lower panel of Figure Figure6)6) and their structure and biosynthesis were elucidated. Studies of changes in development and disease were almost guaranteed to show interesting findings, justifying further funding and research. It was also necessary to decipher the biosynthetic enzymes and mechanisms involved in generating each glycan. Thus there was plenty of interesting work to do, other than take on the most difficult task of elucidating function. Also, many of the functions of glycans tend to be “analog” and not “digital”, and many glycans have more than one disparate function. Finally, the rapid evolution of glycans has generated a lot of species-specific differences, making it difficult to find common themes applicable to all major model systems studies in biology. With the power of modern glycomics and the move to integrate glycosylation data into multiomic studies, it now possible to get past these difficulties and study the functions of glycans like never before. But we still then need to define the glycome.

An external file that holds a picture, illustration, etc.
Object name is cww086f06.jpg

Contrasts in early approaches to the discovery and characterization of proteins and glycans. Compared to the robust and relatively easy interdirectional progress in the early study of proteins, often originating from initial knowledge of their functions (upper panel), early approaches to the discovery and characterization of glycans (lower panel) did not often originate from functional clues. See text for discussion.

What is the glycome?

It is clear that the glycome of an organism is far, far more complex than that of its genome, transcriptome or proteome, and it is only recently that “glycomics” has become practically feasible (1065, 10811089). Daunting and sophisticated as it is, most of what is called glycomics in 2016 still amounts to generating a “parts list” of all the glycans one can find in a given cell type or tissue at a particular point in time and space, i.e., similar to a peptide map of a mixture of proteins. In addition, current methods partially or completely destroy or miss labile modifications like acetylation, sulfation, phosphorylation, lactylation, pyruvylation, etc. More efforts are needed to discover all the glycan attachment sites on proteins and lipids, in a cell type in question (1089). Eventually, we need not only to define all of the above, but also to understand and visualize the conformation and organization of glycans on individual cell types and surfaces, in the form of “clustered saccharide patches” (146, 302), or glycosynapses (87). In the final analysis, full understanding of the biology of glycans will require this comprehensive type of view for which analytical techniques are yet to be defined. But great advances are being made by many investigators in all of the above levels of glycomics, and the future looks bright. Moreover, we can take advantage of the fact that pathogens and commensals have already spent million of years adapting to interact with the glycans of their hosts, and have already evolved highly specific binding proteins for recognition. Thus as mentioned earlier, an entire array of probes for defining the glycome is already available in nature, waiting to be isolated, characterized and eventually converted into practical tools, if necessary with further mutations. Initial steps in this direction are also very promising (1090).

What biological roles do glycans not mediate?

This rhetorical question seeks to emphasize that the biological roles of glycans are highly varied, and span the spectrum of possibilities. So the exceptions are few. So far there does not seem to be an example of multigenerational information transfer directly mediated by glycans, such as that mediated by DNA or RNA. But it has recently become evident that O-GlcNAc can modify RNA polymerase II, histones, histone deacetylase complexes and members of the Polycomb and Trithorax groups (10911092). Thus, it is suggested that O-GlcNAc cycling serves as a mechanism linking nutrient availability to chromatin organization, histone modification and epigenetics (711). It remains to be seen if such epigenetic effects can mediate intergenerational transfer, in a manner similar to other epigenetic marks. There are also no clear-cut examples of glycans acting as enzymes [if one excludes RNA from being considered as a polysaccharide, and intramolecular self-cleavage of PolySia (1093) as a chemical anomaly].

Why the persisting lack of attention to this fundamental component of biology?

Glycans are a major and integral part of all biological systems, and >3 billion years of biological evolution has failed to generate any life form on the planet that is not absolutely dependent on glycan chains for its existence. Yet the current situation is comparable to that in cosmology, with a standard model based on extant knowledge––that functioned well until it was realized that the bulk of the universe consists of dark energy and dark matter, which had been previously ignored. In effect, glycans have become the “dark matter” of the biological universe (1094), important yet poorly understood and therefore deserving special attention. However the levels of funding, the number of scientists involved and the scientific popularity of Glycosciences remain low. Many of the reasons are evident from the foregoing discussion. As mentioned earlier, a major issue is the fact that an entire generation of scientists has been trained with a limited knowledge of this class of molecules, and they are unlikely to now turn to studying them. Thus, a new generation of young minds needs to be educated, in this aspect of biology. The other major factor is the lack of easily available technologies for the synthesis and analysis of glycans. In this regard, the 2012 US National Academies/National Research Council report on the future of glycoscience concluded by recommending: “…transforming Glycoscience from a field dominated by specialists to a widely studied and integrated discipline, which could lead to a more complete understanding of glycans and help solve key challenges in diverse fields”, and emphasized the need to invest in education and technology development (1095). A more recent NIH working group report further emphasizes the need for training in Glycoscience (1080).

Future prospects

Many functions of glycans will continue to be discovered by the conventional processes of scientific investigation that will serendipitously come upon such functions. However, there are several potential systematic approaches to uncovering these functions that are depicted in Figure Figure7.7. Each approach has its pros and cons, which are discussed in detail elsewhere (160). Also not fully shown in this figure are newer methods taking advantage of the power of chemoenzymatic synthesis (62, 127, 158) and introduction of modified sugars with bioorthogonal reporter groups into biological systems (111). As with any biological questions, there are pros and cons of studying isolated cells versus intact organisms. And one must always ask how species- or taxon-specific a given function might be. All of approaches depicted in Figure Figure77 are rendered difficult in the case of glycan types that have numerous nonoverlapping functions in the same biological system. As the late Philip Majerus once put it, trying to decipher the roles of such molecules by preventing their synthesis or by destroying them after the fact is like “sifting through the ashes to find out how dynamite works”. Of course this problem is not unique to glycans. Complexity and pleiotropy are inherent in all of biology, and the same could be said of other post-translational modifications.

An external file that holds a picture, illustration, etc.
Object name is cww086f07.jpg

Approaches towards elucidating biological roles of glycans. The figure assumes that a specific biological role is being mediated by recognition of a certain glycan structure by a specific glycan-binding protein. Clues about biological roles could be obtained by a variety of different approaches. For detailed discussion of each approach, see the original reference. Not shown are newer methods taking advantage of the power of chemoenzymatic synthesis and the introduction of modified sugars with bioorthogonal reporter groups into biological systems. Drawing by R. Cummings, updated from ref. 160 with permission from the Consortium of Glycobiology Editors.

One would have to go back very many decades to find reviews about “biological roles of nucleic acids” or “biological roles of proteins”. The fact that such a review on roles of glycans was necessary in 1993 indicates how far behind we were in our understanding of their biology. As this update after 23 years shows, we have come a very long way, and one author now can barely scratch the surface of the topic in a single review. The time has come for the biology of glycans to be “mainstreamed” with that of the other major macromolecules that are universal to all life forms. But this requires a concerted effort on the part of all biologists and naturalists, to fully integrate the roles of glycans into their thinking about living systems. Once that happens, there will no longer be any need for writing another review like this one.

Acknowledgements

The author apologizes to the numerous excellent scientists whose work is not adequately discussed or cited in this review, and thanks Linda Baum, Tamara Doering, Hudson Freeze, Pascal Gagneux, Rita Gerardy-Schahn, Robert Haltiwanger, Herbert Hildebrandt, Laura Kiessling, Stuart Kornfeld, Lara Mahal, Mike Pierce, Nancy Schwartz, Christine Szymanski, Mukund Thattai, Naoyuki Taniguchi and Chris West for valuable comments and suggestions. Much of the inspiration for this review has come from ongoing discussions with these and other colleagues, especially current and former members of my lab and editors of Essentials of Glycobiology. Long-term support from the National Institutes of Health and the Mathers Foundation of New York is gratefully acknowledged.

Conflict of interest statement

None declared.

References

1. Varki A. 1993. Biological roles of oligosaccharides: All of the theories are correct. Glycobiology. 3:97–130. [PMC free article] [PubMed] [Google Scholar]
2. Fiedler K, Simons K. 1995. The role of N-glycans in the secretory pathway. Cell. 81:309–312. [PubMed] [Google Scholar]
3. Lasky LA. 1995. Selectin–carbohydrate interactions and the initiation of the inflammatory response. Annu Rev Biochem. 64:113–139. [PubMed] [Google Scholar]
4. Nelson RM, Venot A, Bevilacqua MP, Linhardt RJ, Stamenkovic I. 1995. Carbohydrate–protein interactions in vascular biology. Annu Rev Cell Biol. 11:601–631. [PubMed] [Google Scholar]
5. Rudd PM, Woods RJ, Wormald MR, Opdenakker G, Downing AK, Campbell ID, Dwek RA. 1995. The effects of variable glycosylation on the functional activities of ribonuclease, plasminogen and tissue plasminogen activator. Biochim Biophys Acta Protein Struct Mol Enzymol. 1248:1–10. [PubMed] [Google Scholar]
6. Butcher EC, Picker LJ. 1996. Lymphocyte homing and homeostasis. Science. 272:60–66. [PubMed] [Google Scholar]
7. Crocker PR, Feizi T. 1996. Carbohydrate recognition systems: Functional triads in cell-cell interactions. Curr Opin Struct Biol. 6:679–691. [PubMed] [Google Scholar]
8. Dénarié J, Debellé F, Promé JC. 1996. Rhizobium lipo-chitooligosaccharide nodulation factors: Signaling molecules mediating recognition and morphogenesis. Annu Rev Biochem. 65:503–535. [PubMed] [Google Scholar]
9. Fukuda M. 1996. Possible roles of tumor-associated carbohydrate antigens. Cancer Res. 56:2237–2244. [PubMed] [Google Scholar]
10. Gahmberg CG, Tolvanen M. 1996. Why mammalian cell surface proteins are glycoproteins. Trends Biochem Sci. 21:308–311. [PubMed] [Google Scholar]
11. Hakomori S. 1996. Tumor malignancy defined by aberrant glycosylation and sphingo(glyco)lipid metabolism. Cancer Res. 56:5309–5318. [PubMed] [Google Scholar]
12. Hooper LV, Manzella SM, Baenziger JU. 1996. From legumes to leukocytes: Biological roles for sulfated carbohydrates. FASEB J. 10:1137–1146. [PubMed] [Google Scholar]
13. Kansas GS. 1996. Selectins and their ligands: Current concepts and controversies. Blood. 88:3259–3287. [PubMed] [Google Scholar]
14. Kasai K, Hirabayashi J. 1996. Galectins: A family of animal lectins that decipher glycocodes. J Biochem (Tokyo). 119:1–8. [PubMed] [Google Scholar]
15. Kelm S, Schauer R, Crocker PR. 1996. The sialoadhesins—A family of sialic acid-dependent cellular recognition molecules within the immunoglobulin superfamily. Glycoconj J. 13:913–926. [PubMed] [Google Scholar]
16. Prome JC. 1996. Signalling events elicited in plants by defined oligosaccharide structures. Curr Opin Struct Biol. 6:671–678. [PubMed] [Google Scholar]
17. Reuter G, Gabius HJ. 1996. Sialic acids structure-analysis-metabolism-occurrence-recognition. Biol Chem Hoppe Seyler. 377:325–342. [PubMed] [Google Scholar]
18. Rutishauser U. 1996. Polysialic acid and the regulation of cell interactions. Curr Opin Cell Biol. 8:679–684. [PubMed] [Google Scholar]
19. Spillmann D, Burger MM. 1996. Carbohydrate–carbohydrate interactions in adhesion. J Cell Biochem. 61:562–568. [PubMed] [Google Scholar]
20. Carbone FR, Gleeson PA. 1997. Carbohydrates and antigen recognition by T cells. Glycobiology. 7:725–730. [PubMed] [Google Scholar]
21. Hart GW. 1997. Dynamic O-linked glycosylation of nuclear and cytoskeletal proteins. Annu Rev Biochem. 66:315–335. [PubMed] [Google Scholar]
22. Kelm S, Schauer R. 1997. Sialic acids in molecular and cellular interactions. Int Rev Cytol. 175:137–240. [PMC free article] [PubMed] [Google Scholar]
23. McDowell G, Gahl WA. 1997. Inherited disorders of glycoprotein synthesis: Cell biological insights. Proc Soc Exp Biol Med. 215:145–157. [PubMed] [Google Scholar]
24. McEver RP. 1997. Selectin–carbohydrate interactions during inflammation and metastasis. Glycoconj J. 14:585–591. [PubMed] [Google Scholar]
25. Traub LM, Kornfeld S. 1997. The trans-Golgi network: A late secretory sorting station. Curr Opin Cell Biol. 9:527–533. [PubMed] [Google Scholar]
26. Von IM, Thomson RJ. 1997. Sialic acids and sialic acid-recognising proteins: Drug discovery targets and potential glycopharmaceuticals. Curr Med Chem. 4:185–210. [Google Scholar]
27. Etzler ME. 1998. Oligosaccharide signaling of plant cells. J Cell Biochem. 30-31:123–128. [PubMed] [Google Scholar]
28. Hileman RE, Fromm JR, Weiler JM, Linhardt RJ. 1998. Glycosaminoglycan-protein interactions: Definition of consensus sites in glycosaminoglycan binding proteins. BioEssays. 20:156–167. [PubMed] [Google Scholar]
29. Hirschberg CB, Robbins PW, Abeijon C. 1998. Transporters of nucleotide sugars, ATP, and nucleotide sulfate in the endoplasmic reticulum and Golgi apparatus. Annu Rev Biochem. 67:49–69. [PubMed] [Google Scholar]
30. Iozzo RV. 1998. Matrix proteoglycans: From molecular design to cellular function. Annu Rev Biochem. 67:609–652. [PubMed] [Google Scholar]
31. Lander AD. 1998. Proteoglycans: Master regulators of molecular encounter. Matrix Biol. 17:465–472. [PubMed] [Google Scholar]
32. Lindahl U, Kusche-Gullberg M, Kjellén L. 1998. Regulated diversity of heparan sulfate. J Biol Chem. 273:24979–24982. [PubMed] [Google Scholar]
33. Lloyd KO, Furukawa K. 1998. Biosynthesis and functions of gangliosides: recent advances. Glycoconj J. 15:627–636. [PubMed] [Google Scholar]
34. Rahmann H, Jonas U, Kappel T, Hildebrandt H. 1998. Differential involvement of gangliosides versus phospholipids in the process of temperature adaptation in vertebrates—A comparative phenomenological and physicochemical study. Ann NY Acad Sci. 845:72–91. [PubMed] [Google Scholar]
35. Bernfield M, Götte M, Park PW, Reizes O, Fitzgerald ML, Lincecum J, Zako M. 1999. Functions of cell surface heparan sulfate proteoglycans. Annu Rev Biochem. 68:729–777. [PubMed] [Google Scholar]
36. Brossay L, Kronenberg M. 1999. Highly conserved antigen-presenting function of CD1d molecules. Immunogenetics. 50:146–151. [PubMed] [Google Scholar]
37. Carchon H, Van SE, Matthijs G, Jaeken J. 1999. Carbohydrate-deficient glycoprotein syndrome type IA (phosphomannomutase-deficiency). Biochim Biophys Acta Mol Basis Dis. 1455:155–165. [PubMed] [Google Scholar]
38. Cooper DN, Barondes SH. 1999. God must love galectins; he made so many of them. Glycobiology. 9:979–984. [PubMed] [Google Scholar]
39. Freeze HH, Aebi M. 1999. Molecular basis of carbohydrate-deficient glycoprotein syndromes type I with normal phosphomannomutase activity. Biochim Biophys Acta Mol Basis Dis. 1455:167–178. [PubMed] [Google Scholar]
40. Kobata A, Takeuchi M. 1999. Structure, pathology and function of the N-linked sugar chains of human chorionic gonadotropin. Biochim Biophys Acta Mol Basis Dis. 1455:315–326. [PubMed] [Google Scholar]
41. Schachter H, Jaeken J. 1999. Carbohydrate-deficient glycoprotein syndrome type II. Biochim Biophys Acta Mol Basis Dis. 1455:179–192. [PubMed] [Google Scholar]
42. Schuette CG, Doering T, Kolter T, Sandhoff K. 1999. The glycosphingolipidoses-from disease to basic principles of metabolism. Biol Chem Hoppe Seyler. 380:759–766. [PubMed] [Google Scholar]
43. Baeg GH, Perrimon N. 2000. Functional binding of secreted molecules to heparan sulfate proteoglycans in Drosophila. Curr Opin Cell Biol. 12:575–580. [PubMed] [Google Scholar]
44. Funderburgh JL. 2000. Keratan sulfate: Structure, biosynthesis, and function. Glycobiology. 10:951–958. [PubMed] [Google Scholar]
45. Hakomori S. 2000. Traveling for the glycosphingolipid path. Glycoconj J. 17:627–647. [PubMed] [Google Scholar]
46. Hascall VC. 2000. Hyaluronan, a common thread. Glycoconj J. 17:607–616. [PubMed] [Google Scholar]
47. Hughes RC. 2000. So what do your sugars do. Glycoconj J. 17:567–575. [PubMed] [Google Scholar]
48. Kobata A. 2000. A journey to the world of glycobiology. Glycoconj J. 17:443–464. [PubMed] [Google Scholar]
49. Lander AD, Selleck SB. 2000. The elusive functions of proteoglycans: In vivo veritas. J Cell Biol. 148:227–232. [PMC free article] [PubMed] [Google Scholar]
50. Liu FT. 2000. Galectins: A new family of regulators of inflammation. Clin Immunol. 97:79–88. [PubMed] [Google Scholar]
51. Morgan WTJ, Watkins WM. 2000. Unravelling the biochemical basis of blood group ABO and Lewis antigenic specificity. Glycoconj J. 17:501–530. [PubMed] [Google Scholar]
52. Parodi AJ. 2000. Protein glucosylation and its role in protein folding. Annu Rev Biochem. 69:69–93. [PubMed] [Google Scholar]
53. Perrimon N, Bernfield M. 2000. Specificities of heparan sulphate proteoglycans in developmental processes. Nature. 404:725–728. [PubMed] [Google Scholar]
54. Roos MD, Hanover JA. 2000. Structure of O-linked GlcNAc transferase: Mediator of glycan-dependent signaling. Biochem Biophys Res Commun. 271:275–280. [PubMed] [Google Scholar]
55. Scanlin TF, Glick MC. 2000. Terminal glycosylation and disease: Influence on cancer and cystic fibrosis. Glycoconj J. 17:617–626. [PubMed] [Google Scholar]
56. Schachter H. 2000. The joys of HexNAc. The synthesis and function of N-and O-glycan branches. Glycoconj J. 17:465–483. [PubMed] [Google Scholar]
57. Selleck SB. 2000. Proteoglycans and pattern formation—Sugar biochemistry meets developmental genetics. Trends Genet. 16:206–212. [PubMed] [Google Scholar]
58. Yamaguchi Y. 2000. Lecticans: Organizers of the brain extracellular matrix. Cell Mol Life Sci. 57:276–289. [PMC free article] [PubMed] [Google Scholar]
59. Muramatsu T. 2000. a. Essential roles of carbohydrate signals in development, immune response and tissue functions, as revealed by gene targeting. J Biochem (Tokyo). 127:171–176. [PubMed] [Google Scholar]
60. Muramatsu T. 2000. b. Protein-bound carbohydrates on cell-surface as targets of recognition: An Odyssey in understanding them. Glycoconj J. 17:577–595. [PubMed] [Google Scholar]
61. Aebi M, Hennet T. 2001. Congenital disorders of glycosylation: Genetic model systems lead the way. Trends Cell Biol. 11:136–141. [PubMed] [Google Scholar]
62. Bertozzi CR, Kiessling LL. 2001. Chemical glycobiology. Science. 291:2357–2364. [PubMed] [Google Scholar]
63. Cabral CM, Liu Y, Sifers RN. 2001. Dissecting glycoprotein quality control in the secretory pathway. Trends Biochem Sci. 26:619–624. [PubMed] [Google Scholar]
64. Dennis JW, Warren CE, Granovsky M, Demetriou M. 2001. Genetic defects in N-glycosylation and cellular diversity in mammals. Curr Opin Struct Biol. 11:601–607. [PubMed] [Google Scholar]
65. Ellgaard L, Helenius A. 2001. ER quality control: Towards an understanding at the molecular level. Curr Opin Cell Biol. 13:431–437. [PubMed] [Google Scholar]
66. Freeze HH, Westphal V. 2001. Balancing N-linked glycosylation to avoid disease. Biochimie. 83:791–799. [PubMed] [Google Scholar]
67. Freeze HH. 2001. Update and perspectives on congenital disorders of glycosylation. Glycobiology. 11:129R–143R. [PubMed] [Google Scholar]
68. Fukuda M, Hiraoka N, Akama TO, Fukuda MN. 2001. Carbohydrate-modifying sulfotransferases: Structure, function and pathophysiology. J Biol Chem. 276:47747–47750. [PubMed] [Google Scholar]
69. Guha-Niyogi A, Sullivan DR, Turco SJ. 2001. Glycoconjugate structures of parasitic protozoa. Glycobiology. 11:45R–59R. [PubMed] [Google Scholar]
70. Hanover JA. 2001. Glycan-dependent signaling: O-linked N-acetylglucosamine. FASEB J. 15:1865–1876. [PubMed] [Google Scholar]
71. Helenius A, Aebi M. 2001. Intracellular functions of N-linked glycans. Science. 291:2364–2369. [PubMed] [Google Scholar]
72. Hooper LV, Gordon JI. 2001. Glycans as legislators of host-microbial interactions: Spanning the spectrum from symbiosis to pathogenicity. Glycobiology. 11:1R–10R. [PubMed] [Google Scholar]
73. Iozzo RV. 2001. Heparan sulfate proteoglycans: Intricate molecules with intriguing functions. J Clin Invest. 108:165–167. [PMC free article] [PubMed] [Google Scholar]
74. Lowe JB. 2001. Glycosylation, immunity, and autoimmunity. Cell. 104:809–812. [PubMed] [Google Scholar]
75. Mengerink KJ, Vacquier VD. 2001. Glycobiology of sperm–egg interactions in deuterostomes. Glycobiology. 11:37R–43R. [PubMed] [Google Scholar]
76. Nicolaou KC, Mitchell HJ. 2001. Adventures in carbohydrate chemistry: New synthetic technologies, chemical synthesis, molecular design, and chemical biology. Angew Chem Int Edit. 40:1576–1624. [PubMed] [Google Scholar]
77. Rudd PM, Elliott T, Cresswell P, Wilson IA, Dwek RA. 2001. Glycosylation and the immune system. Science. 291:2370–2376. [PubMed] [Google Scholar]
78. Saxon E, Bertozzi CR. 2001. Chemical and biological strategies for engineering cell surface glycosylation. Annu Rev Cell Dev Biol. 17:1–23. [PubMed] [Google Scholar]
79. Selva EM, Perrimon N. 2001. Role of heparan sulfate proteoglycans in cell signaling and cancer. Adv Cancer Res. 83:67–80. [PubMed] [Google Scholar]
80. Vyas AA, Schnaar RL. 2001. Brain gangliosides: Functional ligands for myelin stability and the control of nerve regeneration. Biochimie. 83:677–682. [PubMed] [Google Scholar]
81. Yarema KJ, Goon S, Bertozzi CR. 2001. Metabolic selection of glycosylation defects in human cells. Nat Biotechnol. 19:553–558. [PubMed] [Google Scholar]
82. Angata T, Brinkman-Van der Linden E. 2002. I-type lectins. Biochim Biophys Acta. 1572:294–316. [PubMed] [Google Scholar]
83. Brewer CF. 2002. Binding and cross-linking properties of galectins. Biochim Biophys Acta Gen Subj. 1572:255–262. [PubMed] [Google Scholar]
84. D'Haeze W, Holsters M. 2002. Nod factor structures, responses, and perception during initiation of nodule development. Glycobiology. 12:79R–105R. [PubMed] [Google Scholar]
85. Dahms NM, Hancock MK. 2002. P-type lectins. Biochim Biophys Acta Gen Subj. 1572:317–340. [PubMed] [Google Scholar]
86. Esko JD, Selleck SB. 2002. Order out of chaos: Assembly of ligand binding sites in heparan sulfate. Annu Rev Biochem. 71:435–471. [PubMed] [Google Scholar]
87. Hakomori S. 2002. The glycosynapse. Proc Natl Acad Sci USA. 99:225–232. [PMC free article] [PubMed] [Google Scholar]
88. Lu JH, Teh C, Kishore U, Reid KBM. 2002. Collectins and ficolins: Sugar pattern recognition molecules of the mammalian innate immune system. Biochim Biophys Acta Gen Subj. 1572:387–400. [PubMed] [Google Scholar]
89. Martin PT. 2002. Glycobiology of the synapse. Glycobiology. 12:1R–7R. [PubMed] [Google Scholar]
90. McDonald J, Hascall VC. 2002. Hyaluronan minireview series. J Biol Chem. 277:4575–4579. [PubMed] [Google Scholar]
91. McEver RP. 2002. Selectins: Lectins that initiate cell adhesion under flow. Curr Opin Cell Biol. 14:581–586. [PubMed] [Google Scholar]
92. Rabinovich GA, Baum LG, Tinari N, Paganelli R, Natoli C, Liu FT, Iacobelli S. 2002. Galectins and their ligands: Amplifiers, silencers or tuners of the inflammatory response. Trends Immunol. 23:313–320. [PubMed] [Google Scholar]
93. Raetz CRH, Whitfield C. 2002. Lipopolysaccharide endotoxins. Annu Rev Biochem. 71:635–700. [PMC free article] [PubMed] [Google Scholar]
94. Spiro RG. 2002. Protein glycosylation: Nature, distribution, enzymatic formation, and disease implications of glycopeptide bonds. Glycobiology. 12:43R–56R. [PubMed] [Google Scholar]
95. Toole BP, Wight TN, Tammi MI. 2002. Hyaluronan–cell interactions in cancer and vascular disease. J Biol Chem. 277:4593–4596. [PubMed] [Google Scholar]
96. Trowbridge JM, Gallo RL. 2002. Dermatan sulfate: New functions from an old glycosaminoglycan. Glycobiology. 12:117R–125R. [PubMed] [Google Scholar]
97. Vimr E, Lichtensteiger C. 2002. To sialylate, or not to sialylate: That is the question. Trends Microbiol. 10:254–257. [PubMed] [Google Scholar]
98. Lowe JB, Marth JD. 2003. A genetic approach to Mammalian glycan function. Annu Rev Biochem. 72:643–691. [PubMed] [Google Scholar]
99. Martin PT, Freeze HH. 2003. Glycobiology of neuromuscular disorders. Glycobiology. 13:67–75. [PubMed] [Google Scholar]
100. Brigl M, Brenner MB. 2004. CD1: Antigen presentation and T cell function. Annu Rev Immunol. 22:817–890. [PubMed] [Google Scholar]
101. Furukawa K, Tokuda N, Okuda T, Tajima O, Furukawa K. 2004. Glycosphingolipids in engineered mice: Insights into function. Semin Cell Dev Biol. 15:389–396. [PubMed] [Google Scholar]
102. Gu J, Taniguchi N. 2004. Regulation of integrin functions by N-glycans. Glycoconj J. 21:9–15. [PubMed] [Google Scholar]
103. Rosen SD. 2004. Ligands for L-selectin: Homing, inflammation, and beyond. Annu Rev Immunol. 22:129–156. [PubMed] [Google Scholar]
104. Haltiwanger RS, Lowe JB. 2004. Role of glycosylation in development. Annu Rev Biochem. 73:491–537. [PubMed] [Google Scholar]
105. Dube DH, Bertozzi CR. 2005. Glycans in cancer and inflammation—potential for therapeutics and diagnostics. Nat Rev Drug Discov. 4:477–488. [PubMed] [Google Scholar]
106. Sackstein R. 2005. The lymphocyte homing receptors: Gatekeepers of the multistep paradigm. Curr Opin Hematol. 12:444–450. [PubMed] [Google Scholar]
107. Yuki N, Odaka M. 2005. Ganglioside mimicry as a cause of Guillain-Barre syndrome. Curr Opin Neurol. 18:557–561. [PubMed] [Google Scholar]
108. Brockhausen I. 2006. Mucin-type O-glycans in human colon and breast cancer: Glycodynamics and functions. EMBO Rep. 7:599–604. [PMC free article] [PubMed] [Google Scholar]
109. Lehmann F, Tiralongo E, Tiralongo J. 2006. Sialic acid-specific lectins: Occurrence, specificity and function. Cell Mol Life Sci. 63:1331–1354. [PMC free article] [PubMed] [Google Scholar]
110. Ohtsubo K, Marth JD. 2006. Glycosylation in cellular mechanisms of health and disease. Cell. 126:855–867. [PubMed] [Google Scholar]
111. Prescher JA, Bertozzi CR. 2006. Chemical technologies for probing glycans. Cell. 126:851–854. [PubMed] [Google Scholar]
112. Rose MC, Voynow JA. 2006. Respiratory tract mucin genes and mucin glycoproteins in health and disease. Physiol Rev. 86:245–278. [PubMed] [Google Scholar]
113. Crocker PR, Paulson JC, Varki A. 2007. Siglecs and their roles in the immune system. Nat Rev Immunol. 7:255–266. [PubMed] [Google Scholar]
114. Hildebrandt H, Muhlenhoff M, Weinhold B, Gerardy-Schahn R. 2007. Dissecting polysialic acid and NCAM functions in brain development. J Neurochem. 103(Suppl 1): 56–64. [PubMed] [Google Scholar]
115. Jiang D, Liang J, Noble PW. 2007. Hyaluronan in tissue injury and repair. Annu Rev Cell Dev Biol. 23:435–461. [PubMed] [Google Scholar]
116. Trinchieri G, Sher A. 2007. Cooperation of Toll-like receptor signals in innate immune defence. Nat Rev Immunol. 7:179–190. [PubMed] [Google Scholar]
117. Garner OB, Baum LG. 2008. Galectin-glycan lattices regulate cell-surface glycoprotein organization and signalling. Biochem Soc Trans. 36:1472–1477. [PMC free article] [PubMed] [Google Scholar]
118. Marth JD, Grewal PK. 2008. Mammalian glycosylation in immunity. Nat Rev Immunol. 8:874–887. [PMC free article] [PubMed] [Google Scholar]
119. Todeschini A, Hakomori SI. 2008. Functional role of glycosphingolipids and gangliosides in control of cell adhesion, motility, and growth, through glycosynaptic microdomains. Biochim Biophys Acta. 1780:421–433. [PMC free article] [PubMed] [Google Scholar]
120. Bode L. 2009. Human milk oligosaccharides: Prebiotics and beyond. Nutr Rev. 67(Suppl 2):S183–S191. [PubMed] [Google Scholar]
121. Hsu KL, Mahal LK. 2009. Sweet tasting chips: Microarray-based analysis of glycans. Curr Opin Chem Biol. 13:427–432. [PubMed] [Google Scholar]
122. Jefferis R. 2009. Glycosylation as a strategy to improve antibody-based therapeutics. Nat Rev Drug Discov. 8:226–234. [PubMed] [Google Scholar]
123. Lommel M, Strahl S. 2009. Protein O-mannosylation: Conserved from bacteria to humans. Glycobiology. 19:816–828. [PubMed] [Google Scholar]
124. Taylor ME, Drickamer K. 2009. Structural insights into what glycan arrays tell us about how glycan-binding proteins interact with their ligands. Glycobiology. 19:1155–1162. [PMC free article] [PubMed] [Google Scholar]
125. Dam TK, Brewer CF. 2010. Lectins as pattern recognition molecules: The effects of epitope density in innate immunity. Glycobiology. 20:270–279. [PubMed] [Google Scholar]
126. Fadda E, Woods RJ. 2010. Molecular simulations of carbohydrates and protein-carbohydrate interactions: Motivation, issues and prospects. Drug Discov Today. 15:596–609. [PMC free article] [PubMed] [Google Scholar]
127. Kiessling LL, Splain RA. 2010. Chemical approaches to glycobiology. Annu Rev Biochem. 79:619–653. [PubMed] [Google Scholar]
128. Laubli H, Borsig L. 2010. Selectins promote tumor metastasis. Semin Cancer Biol. 20:169–177. [PubMed] [Google Scholar]
129. Monti E, Bonten E, D'Azzo A, Bresciani R, Venerando B, Borsani G, Schauer R, Tettamanti G. 2010. Sialidases in vertebrates a family of enzymes tailored for several cell functions. Adv Carbohydr Chem Biochem. 64:403–479. [PubMed] [Google Scholar]
130. Castonguay AC, Olson LJ, Dahms NM. 2011. Mannose 6-phosphate receptor homology (MRH) domain-containing lectins in the secretory pathway. Biochim Biophys Acta. 1810:815–826. [PMC free article] [PubMed] [Google Scholar]
131. Neelamegham S, Liu G. 2011. Systems glycobiology: Biochemical reaction networks regulating glycan structure and function. Glycobiology. 21:1541–1553. [PMC free article] [PubMed] [Google Scholar]
132. Rillahan CD, Paulson JC. 2011. Glycan microarrays for decoding the glycome. Annu Rev Biochem. 80:797–823. [PMC free article] [PubMed] [Google Scholar]
133. Stalnaker SH, Stuart R, Wells L. 2011. Mammalian O-mannosylation: Unsolved questions of structure/function. Curr Opin Struct Biol. 21:603–609. [PMC free article] [PubMed] [Google Scholar]
134. Dall'olio F, Malagolini N, Trinchera M, Chiricolo M. 2012. Mechanisms of cancer-associated glycosylation changes. Front Biosci. 17:670–699. [PubMed] [Google Scholar]
135. Dobson CM, Hempel SJ, Stalnaker SH, Stuart R, Wells L. 2012. O-Mannosylation and human disease. Cell Mol Life Sci. 70:2849–2857. [PMC free article] [PubMed] [Google Scholar]
136. Gille S, Pauly M. 2012. O-acetylation of plant cell wall polysaccharides. Front Plant Sci. 3:12. [PMC free article] [PubMed] [Google Scholar]
137. Imai M, Kawaoka Y. 2012. The role of receptor binding specificity in interspecies transmission of influenza viruses. Curr Opin Virol. 2:160–167. [PMC free article] [PubMed] [Google Scholar]
138. Kreisman LS, Cobb BA. 2012. Infection, inflammation and host carbohydrates: A Glyco-Evasion Hypothesis. Glycobiology. 22:1019–1030. [PMC free article] [PubMed] [Google Scholar]
139. Bergstrom KS, Xia L. 2013. Mucin-type O-glycans and their roles in intestinal homeostasis. Glycobiology. 23:1026–1037. [PMC free article] [PubMed] [Google Scholar]
140. Clark GF. 2013. The role of carbohydrate recognition during human sperm-egg binding. Hum Reprod. 28:566–577. [PubMed] [Google Scholar]
141. Ferrari S, Savatin DV, Sicilia F, Gramegna G, Cervone F, Lorenzo GD. 2013. Oligogalacturonides: Plant damage-associated molecular patterns and regulators of growth and development. Front Plant Sci. 4:49. [PMC free article] [PubMed] [Google Scholar]
142. Wiederschain GY. 2013. Glycobiology: Progress, problems, and perspectives. Biochemistry (Mosc). 78:679–696. [PubMed] [Google Scholar]
143. Baum LG, Garner OB, Schaefer K, Lee B. 2014. Microbe-host interactions are positively and negatively regulated by galectin-glycan interactions. Front Immunol. 5:284. [PMC free article] [PubMed] [Google Scholar]
144. Bull C, den Brok MH, Adema GJ. 2014. Sweet escape: Sialic acids in tumor immune evasion. Biochim Biophys Acta. 1846:238–246. [PubMed] [Google Scholar]
145. Chang YC, Nizet V. 2014. The interplay between Siglecs and sialylated pathogens. Glycobiology. 24:818–825. [PMC free article] [PubMed] [Google Scholar]
146. Cohen M, Varki A. 2014. Modulation of glycan recognition by clustered saccharide patches. Int Rev Cell Mol Biol. 308:75–125. [PubMed] [Google Scholar]
147. Colley KJ, Kitajima K, Sato C. 2014. Polysialic acid: Biosynthesis, novel functions and applications. Crit Rev Biochem Mol Biol. 49:498–532. [PubMed] [Google Scholar]
148. Dall'Olio F, Malagolini N, Trinchera M, Chiricolo M. 2014. Sialosignaling: Sialyltransferases as engines of self-fueling loops in cancer progression. Biochim Biophys Acta. 1840:2752–2764. [PubMed] [Google Scholar]
149. de Graaf M, Fouchier RA. 2014. Role of receptor binding specificity in influenza A virus transmission and pathogenesis. EMBO J. 33:823–841. [PMC free article] [PubMed] [Google Scholar]
150. Jiang T, Yu JT, Hu N, Tan MS, Zhu XC, Tan L. 2014. CD33 in Alzheimer's disease. Mol Neurobiol. 49:529–535. [PubMed] [Google Scholar]
151. Macauley MS, Crocker PR, Paulson JC. 2014. Siglec-mediated regulation of immune cell function in disease. Nat Rev Immunol. 14:653–666. [PMC free article] [PubMed] [Google Scholar]
152. Raju TS, Lang SE. 2014. Diversity in structure and functions of antibody sialylation in the Fc. Curr Opin Biotechnol. 30C:147–152. [PubMed] [Google Scholar]
153. Salama A, Evanno G, Harb J, Soulillou JP. 2014. Potential deleterious role of anti-Neu5Gc antibodies in xenotransplantation. Xenotransplantation. 22:85–94. [PubMed] [Google Scholar]
154. Stencel-Baerenwald JE, Reiss K, Reiter DM, Stehle T, Dermody TS. 2014. The sweet spot: Defining virus-sialic acid interactions. Nat Rev Microbiol. 12:739–749. [PMC free article] [PubMed] [Google Scholar]
155. Xu D, Esko JD. 2014. Demystifying heparan sulfate–protein interactions. Annu Rev Biochem. 83:129–157. [PMC free article] [PubMed] [Google Scholar]
156. Bochner BS, Zimmermann N. 2015. Role of siglecs and related glycan-binding proteins in immune responses and immunoregulation. J Allergy Clin Immunol. 135:598–608. [PMC free article] [PubMed] [Google Scholar]
157. Langford-Smith A, Day AJ, Bishop PN, Clark SJ. 2015. Complementing the sugar code: Role of GAGs and sialic acid in complement regulation. Front Immunol. 6:25. [PMC free article] [PubMed] [Google Scholar]
158. Krasnova L, Wong CH. 2016. Understanding the chemistry and biology of glycosylation with glycan synthesis. Annu Rev Biochem. 85:599–630. [PubMed] [Google Scholar]
159. Stanley P. 2016. What have we learned from glycosyltransferase knockouts in mice. J Mol Biol. 428:3166–3182. [PMC free article] [PubMed] [Google Scholar]
160. Varki A, Lowe JB. 2009. Biological roles of glycans In: Essentials of Glycobiology. Cold Spring Harbor (NY), p. 75–88. [PubMed] [Google Scholar]
161. Drickamer K, Taylor ME. 1998. Evolving views of protein glycosylation. Trends Biochem Sci. 23:321–324. [PubMed] [Google Scholar]
162. Burda P, Aebi M. 1999. The dolichol pathway of N-linked glycosylation. Biochim Biophys Acta Gen Subj. 1426:239–257. [PubMed] [Google Scholar]
163. DeAngelis PL. 1999. Hyaluronan synthases: Fascinating glycosyltransferases from vertebrates, bacterial pathogens, and algal viruses. Cell Mol Life Sci. 56:670–682. [PMC free article] [PubMed] [Google Scholar]
164. Gagneux P, Varki A. 1999. Evolutionary considerations in relating oligosaccharide diversity to biological function. Glycobiology. 9:747–755. [PubMed] [Google Scholar]
165. Apoil PA, Roubinet F, Despiau S, Mollicone R, Oriol R, Blancher A. 2000. Evolution of alpha2-fucosyltransferase genes in primates: Relation between an intronic Alu-Y element and red cell expression of ABH antigens. Mol Biol Evol. 17:337–351. [PubMed] [Google Scholar]
166. Keusch JJ, Manzella SM, Nyame KA, Cummings RD, Baenziger JU. 2000. Cloning of Gb3 synthase, the key enzyme in globo-series glycosphingolipid synthesis, predicts a family of beta1,4-glycosyltransferases conserved in plants, insects, and mammals. J Biol Chem. 275:25315–25321. [PubMed] [Google Scholar]
167. Srinivas VR, Reddy GB, Ahmad N, Swaminathan CP, Mitra N, Surolia A. 2001. Legume lectin family, the “natural mutants of the quaternary state” provide insights into the relationship between protein stability and oligomerization. Biochim Biophys Acta Gen Subj. 1527:102–111. [PubMed] [Google Scholar]
168. Valla S, Li JP, Ertesvåg H, Barbeyron T, Lindahl U. 2001. Hexuronyl C5-epimerases in alginate and glycosaminoglycan biosynthesis. Biochimie. 83:819–830. [PubMed] [Google Scholar]
169. Hirabayashi J, Hashidate T, Arata Y, Nishi N, Nakamura T, Hirashima M, Urashima T, Oka T, Futai M, Muller WEG, et al.. 2002. Oligosaccharide specificity of galectins: A search by frontal affinity chromatography. Biochim Biophys Acta Gen Subj. 1572:232–254. [PubMed] [Google Scholar]
170. Varki A. 2006. Nothing in glycobiology makes sense, except in the light of evolution. Cell. 126:841–845. [PubMed] [Google Scholar]
171. Bishop JR, Gagneux P. 2007. Evolution of carbohydrate antigens--microbial forces shaping host glycomes. Glycobiology. 17:23R–34R. [PubMed] [Google Scholar]
172. Nothaft H, Szymanski CM. 2013. Bacterial protein N-glycosylation: New perspectives and applications. J Biol Chem. 288:6912–6920. [PMC free article] [PubMed] [Google Scholar]
173. Springer SA, Gagneux P. 2013. Glycan evolution in response to collaboration, conflict, and constraint. J Biol Chem. 288:6904–6911. [PMC free article] [PubMed] [Google Scholar]
174. Lombard V, Golaconda Ramulu H, Drula E, Coutinho PM, Henrissat B. 2014. The carbohydrate-active enzymes database (CAZy) in 2013. Nucleic Acids Res. 42:D490–D495. [PMC free article] [PubMed] [Google Scholar]
175. Corfield AP, Berry M. 2015. Glycan variation and evolution in the eukaryotes. Trends Biochem Sci. 40:351–359. [PubMed] [Google Scholar]
176. Lombard J. 2016. The multiple evolutionary origins of the eukaryotic N-glycosylation pathway. Biol Direct. 11:36. [PMC free article] [PubMed] [Google Scholar]
177. Marth JD. 2008. A unified vision of the building blocks of life. Nat Cell Biol. 10:1015–1016. [PMC free article] [PubMed] [Google Scholar]
178. Varki A. 2011. a. Evolutionary forces shaping the Golgi glycosylation machinery: Why cell surface glycans are universal to living cells. Cold Spring Harb Perspect Biol. 3, doi:pii: a005462. 10.1101/cshperspect.a005462. [PMC free article] [PubMed] [Google Scholar]
179. Dalziel M, Crispin M, Scanlan CN, Zitzmann N, Dwek RA. 2014. Emerging principles for the therapeutic exploitation of glycosylation. Science. 343:1235681. [PubMed] [Google Scholar]
180. Varki A, Sharon N. 2009. Historical Background and Overview. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME editors. Essentials of Glycobiology Cold Spring Harbor (NY), Cold Spring Harbor Laboratory Press; p. 1–22. [PubMed] [Google Scholar]
181. Koles K, Repnikova E, Pavlova G, Korochkin LI, Panin VM. 2009. Sialylation in protostomes: A perspective from Drosophila genetics and biochemistry. Glycoconj J. 26:313–324. [PubMed] [Google Scholar]
182. Bruckner K, Perez L, Clausen H, Cohen S. 2000. Glycosyltransferase activity of Fringe modulates Notch-Delta interactions. Nature. 406:411–415. [PubMed] [Google Scholar]
183. Moloney DJ, Shair LH, Lu FM, Xia J, Locke R, Matta KL, Haltiwanger RS. 2000. a. Mammalian Notch1 is modified with two unusual forms of O-linked glycosylation found on epidermal growth factor-like modules. J Biol Chem. 275:9604–9611. [PubMed] [Google Scholar]
184. Moloney DJ, Panin VM, Johnston SH, Chen J, Shao L, Wilson R, Wang Y, Stanley P, Irvine KD, Haltiwanger RS, et al.. 2000. b. Fringe is a glycosyltransferase that modifies Notch. Nature. 406:369–375. [PubMed] [Google Scholar]
185. Hofsteenge J, Huwiler KG, Macek B, Hess D, Lawler J, Mosher DF, Peter-Katalinic J. 2001. C-mannosylation and O-fucosylation of the thrombospondin type 1 module. J Biol Chem. 276:6485–6498. [PubMed] [Google Scholar]
186. Wang F, Metcalf T, van der Wel H, West CM. 2003. Initiation of mucin-type O-glycosylation in dictyostelium is homologous to the corresponding step in animals and is important for spore coat function. J Biol Chem. 278:51395–51407. [PubMed] [Google Scholar]
187. Matsuura A, Ito M, Sakaidani Y, Kondo T, Murakami K, Furukawa K, Nadano D, Matsuda T, Okajima T. 2008. O-linked N-acetylglucosamine is present on the extracellular domain of notch receptors. J Biol Chem. 283:35486–35495. [PubMed] [Google Scholar]
188. Alfaro JF, Gong CX, Monroe ME, Aldrich JT, Clauss TR, Purvine SO, Wang Z, Camp DG, Shabanowitz J, Stanley P, et al.. 2012. Tandem mass spectrometry identifies many mouse brain O-GlcNAcylated proteins including EGF domain-specific O-GlcNAc transferase targets. Proc Natl Acad Sci USA. 109:7280–7285. [PMC free article] [PubMed] [Google Scholar]
189. Vliegenthart JFG, Casset F. 1998. Novel forms of protein glycosylation. Curr Opin Struct Biol. 8:565–571. [PubMed] [Google Scholar]
190. Gonzalez de Peredo A, Klein D, Macek B, Hess D, Peter-Katalinic J, Hofsteenge J. 2002. C-mannosylation and o-fucosylation of thrombospondin type 1 repeats. Mol Cell Proteomics. 1:11–18. [PubMed] [Google Scholar]
191. Yoshida A, Kobayashi K, Manya H, Taniguchi K, Kano H, Mizuno M, Inazu T, Mitsuhashi H, Takahashi S, Takeuchi M, et al.. 2001. Muscular dystrophy and neuronal migration disorder caused by mutations in a glycosyltransferase, POMGnT1. Dev Cell. 1:717–724. [PubMed] [Google Scholar]
192. Michele DE, Barresi R, Kanagawa M, Saito F, Cohn RD, Satz JS, Dollar J, Nishino I, Kelley RI, Somer H, et al.. 2002. Post-translational disruption of dystroglycan–ligand interactions in congenital muscular dystrophies. Nature. 418:417–422. [PubMed] [Google Scholar]
193. Grewal PK, Hewitt JE. 2003. Glycosylation defects: A new mechanism for muscular dystrophy. Hum Mol Genet. 12(Suppl 2):R259–R264. [PubMed] [Google Scholar]
194. Haliloglu G, Topaloglu H. 2004. Glycosylation defects in muscular dystrophies. Curr Opin Neurol. 17:521–527. [PubMed] [Google Scholar]
195. Grewal PK, Holzfeind PJ, Bittner RE, Hewitt JE. 2001. Mutant glycosyltransferase and altered glycosylation of alpha-dystroglycan in the myodystrophy mouse. Nature Genet. 28:151–154. [PubMed] [Google Scholar]
196. Longman C, Brockington M, Torelli S, Jimenez-Mallebrera C, Kennedy C, Khalil N, Feng L, Saran RK, Voit T, Merlini L, et al.. 2003. Mutations in the human LARGE gene cause MDC1D, a novel form of congenital muscular dystrophy with severe mental retardation and abnormal glycosylation of (alpha)-dystroglycan. Hum Mol Genet. 12:2853–2861. [PubMed] [Google Scholar]
197. Kanagawa M, Saito F, Kunz S, Yoshida-Moriguchi T, Barresi R, Kobayashi YM, Muschler J, Dumanski JP, Michele DE, Oldstone MB, et al.. 2004. Molecular recognition by LARGE is essential for expression of functional dystroglycan. Cell. 117:953–964. [PubMed] [Google Scholar]
198. Brockington M, Torelli S, Prandini P, Boito C, Dolatshad NF, Longman C, Brown SC, Muntoni F. 2005. Localization and functional analysis of the LARGE family of glycosyltransferases: Significance for muscular dystrophy. Hum Mol Genet. 14:657–665. [PubMed] [Google Scholar]
199. Fujimura K, Sawaki H, Sakai T, Hiruma T, Nakanishi N, Sato T, Ohkura T, Narimatsu H. 2005. LARGE2 facilitates the maturation of alpha-dystroglycan more effectively than LARGE. Biochem Biophys Res Commun. 329:1162–1171. [PubMed] [Google Scholar]
200. Patnaik SK, Stanley P. 2005. Mouse large can modify complex N- and mucin O-glycans on alpha-dystroglycan to induce laminin binding. J Biol Chem. 280:20851–20859. [PubMed] [Google Scholar]
201. Yoshida-Moriguchi T, Yu L, Stalnaker SH, Davis S, Kunz S, Madson M, Oldstone MB, Schachter H, Wells L, Campbell KP. 2010. O-mannosyl phosphorylation of alpha-dystroglycan is required for laminin binding. Science. 327:88–92. [PMC free article] [PubMed] [Google Scholar]
202. Inamori K, Yoshida-Moriguchi T, Hara Y, Anderson ME, Yu L, Campbell KP. 2012. Dystroglycan function requires xylosyl- and glucuronyltransferase activities of LARGE. Science. 335:93–96. [PMC free article] [PubMed] [Google Scholar]
203. Inamori K, Hara Y, Willer T, Anderson ME, Zhu Z, Yoshida-Moriguchi T, Campbell KP. 2013. Xylosyl- and glucuronyltransferase functions of LARGE in alpha-dystroglycan modification are conserved in LARGE2. Glycobiology. 23:295–302. [PMC free article] [PubMed] [Google Scholar]
204. Yoshida-Moriguchi T, Campbell KP. 2015. Matriglycan: A novel polysaccharide that links dystroglycan to the basement membrane. Glycobiology. 25:702–713. [PMC free article] [PubMed] [Google Scholar]
205. Riemersma M, Froese DS, van Tol W, Engelke UF, Kopec J, van Scherpenzeel M, Ashikov A, Krojer T, von Delft F, Tessari M, et al.. 2015. Human ISPD is a cytidyltransferase required for dystroglycan O-mannosylation. Chem Biol. 22:1643–1652. [PubMed] [Google Scholar]
206. Gerin I, Ury B, Breloy I, Bouchet-Seraphin C, Bolsée J, Halbout M, Graff J, Vertommen D, Muccioli GG, Seta N, et al.. 2016. ISPD produces CDP-ribitol used by FKTN and FKRP to transfer ribitol phosphate onto α-dystroglycan. Nat Commun. 7:11534. [PMC free article] [PubMed] [Google Scholar]
207. Kanagawa M, Kobayashi K, Tajiri M, Manya H, Kuga A, Yamaguchi Y, Akasaka-Manya K, Furukawa J, Mizuno M, Kawakami H, et al.. 2016. Identification of a post-translational modification with Ribitol-phosphate and its defect in muscular dystrophy. Cell Rep. 14:2209–2223. [PubMed] [Google Scholar]
208. Praissman JL, Willer T, Sheikh MO, Toi A, Chitayat D, Lin YY, Lee H, Stalnaker SH, Wang S, Prabhakar PK, et al.. 2016. The functional O-mannose glycan on α-dystroglycan contains a phospho-ribitol primed for matriglycan addition. Elife. 5:e14473. [PMC free article] [PubMed] [Google Scholar]
209. Tan L, Eberhard S, Pattathil S, Warder C, Glushka J, Yuan C, Hao Z, Zhu X, Avci U, Miller JS, et al.. 2013. An Arabidopsis cell wall proteoglycan consists of pectin and arabinoxylan covalently linked to an arabinogalactan protein. Plant Cell. 25:270–287. [PMC free article] [PubMed] [Google Scholar]
210. Peña MJ, Kong Y, York WS, O'Neill MA. 2012. A galacturonic acid-containing xyloglucan is involved in Arabidopsis root hair tip growth. Plant Cell. 24:4511–4524. [PMC free article] [PubMed] [Google Scholar]
211. Smilowitz JT, Lebrilla CB, Mills DA, German JB, Freeman SL. 2014. Breast milk oligosaccharides: Structure–function relationships in the neonate. Annu Rev Nutr. 34:143–169. [PMC free article] [PubMed] [Google Scholar]
212. Eynon EE, Zenewicz LA, Flavell RA. 2005. Sugar-coated regulation of T cells. Cell. 122:2–4. [PubMed] [Google Scholar]
213. Mazmanian SK, Liu CH, Tzianabos AO, Kasper DL. 2005. An immunomodulatory molecule of symbiotic bacteria directs maturation of the host immune system. Cell. 122:107–118. [PubMed] [Google Scholar]
214. Szymanski CM, Yao R, Ewing CP, Trust TJ, Guerry P. 1999. Evidence for a system of general protein glycosylation in Campylobacter jejuni. Mol Microbiol. 32:1022–1030. [PubMed] [Google Scholar]
215. Wacker M, Linton D, Hitchen PG, Nita-Lazar M, Haslam SM, North SJ, Panico M, Morris HR, Dell A, Wren BW, et al.. 2002. N-linked glycosylation in Campylobacter jejuni and its functional transfer into E. coli. Science. 298:1790–1793. [PubMed] [Google Scholar]
216. Young NM, Brisson JR, Kelly J, Watson DC, Tessier L, Lanthier PH, Jarrell HC, Cadotte N St, Michael F, Aberg E, et al.. 2002. Structure of the N-linked glycan present on multiple glycoproteins in the Gram-negative bacterium, Campylobacter jejuni. J Biol Chem. 277:42530–42539. [PubMed] [Google Scholar]
217. Knirel YA, Vinogradov EV, Shashkov AS, Dmitriev BA, Kochetkov NK, Stanislavsky ES, Mashilova GM. 1986. Somatic antigens of Pseudomonas aeruginosa. The structure of O-specific polysaccharide chains of P. aeruginosa O10 (Lanyi) lipopolysaccharides. Eur J Biochem. 157:129–138. [PubMed] [Google Scholar]
218. Thibault P, Logan SM, Kelly JF, Brisson JR, Ewing CP, Trust TJ, Guerry P. 2001. Identification of the carbohydrate moieties and glycosylation motifs in Campylobacter jejuni flagellin. J Biol Chem. 276:34862–34870. [PubMed] [Google Scholar]
219. Schoenhofen IC, Vinogradov E, Whitfield DM, Brisson JR, Logan SM. 2009. The CMP-legionaminic acid pathway in Campylobacter: Biosynthesis involving novel GDP-linked precursors. Glycobiology. 19:715–725. [PubMed] [Google Scholar]
220. Lewis AL, Desa N, Hansen EE, Knirel YA, Gordon JI, Gagneux P, Nizet V, Varki A. 2009. Innovations in host and microbial sialic acid biosynthesis revealed by phylogenomic prediction of nonulosonic acid structure. Proc Natl Acad Sci USA. 106:13552–13557. [PMC free article] [PubMed] [Google Scholar]
221. Brodsky FM, Thattai M, Mayor S. 2012. Evolutionary cell biology: Lessons from diversity. Nat Cell Biol. 14:651. [PubMed] [Google Scholar]
222. Dey G, Thattai M, Baum B. 2016. On the archaeal origins of eukaryotes and the challenges of inferring phenotype from genotype. Trends Cell Biol. 10.1016/j.tcb.2016.03.009. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
223. Esko JD, Sharon N. 2009. Microbial Lectins: Hemagglutinins, Adhesins, and Toxins. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; p. 489–500. [PubMed] [Google Scholar]
224. Neufeld EF, Ginsburg V. 1966. Complex Carbohydrates—Part A. New York: Academic Press. [Google Scholar]
225. Ginsburg V. 1972. Complex Carbohydrates—Part B. New York: Academic Press. [Google Scholar]
226. Ginsburg V. 1978. Complex Carbohydrates—Part C. New York: Academic Press. [Google Scholar]
227. De Duve C, Wattiaux R. 1966. Functions of lysosomes. Annu Rev Physiol. 28:435–492. [PubMed] [Google Scholar]
228. Neufeld EF, Fratantoni JC. 1970. Inborn errors of mucopolysaccharide metabolism. Science. 169:141–146. [PubMed] [Google Scholar]
229. Gottlieb C, Skinner AM, Kornfeld S. 1974. Isolation of a clone of Chinese hamster ovary cells deficient in plant lectin-binding sites. Proc Natl Acad Sci USA. 71:1078–1082. [PMC free article] [PubMed] [Google Scholar]
230. Gottlieb C, Baenziger J, Kornfeld S. 1975. Deficient uridine diphosphate-N-acetylglucosamine:glycoprotein N-acetylglucosaminyltransferase activity in a clone of Chinese hamster ovary cells with altered surface glycoproteins. J Biol Chem. 250:3303–3309. [PubMed] [Google Scholar]
231. Stanley P, Narasimhan S, Siminovitch L, Schachter H. 1975. Chinese hamster ovary cells selected for resistance to the cytotoxicity of phytohemagglutinin are deficient in a UDP-N-acetylglucosamine–glycoprotein N-acetylglucosaminyltransferase activity. Proc Natl Acad Sci USA. 72:3323–3327. [PMC free article] [PubMed] [Google Scholar]
232. Briles EB, Li E, Kornfeld S. 1977. Isolation of wheat germ agglutinin-resistant clones of Chinese hamster ovary cells deficient in membrane sialic acid and galactose. J Biol Chem. 252:1107–1116. [PubMed] [Google Scholar]
233. Narasimhan S, Stanley P, Schachter H. 1977. Control of glycoprotein synthesis. Lectin-resistant mutant containing only one of two distinct N-acetylglucosaminyltransferase activities present in wild type Chinese hamster ovary cells. J Biol Chem. 252:3926–3933. [PubMed] [Google Scholar]
234. Li E, Kornfeld S. 1978. Structure of the altered oligosaccharide present in glycoproteins from a clone of Chinese hamster ovary cells deficient in N-acetylglucosaminyltransferase activity. J Biol Chem. 253:6426–6431. [PubMed] [Google Scholar]
235. Chapman A, Trowbridge IS, Hyman R, Kornfeld S. 1979. Structure of the lipid-linked oligosaccharides that accumulate in class E Thy-1-negative mutant lymphomas. Cell. 17:509–515. [PubMed] [Google Scholar]
236. Reitman ML, Trowbridge IS, Kornfeld S. 1980. Mouse lymphoma cell lines resistant to pea lectin are defective in fucose metabolism. J Biol Chem. 255:9900–9906. [PubMed] [Google Scholar]
237. Cummings RD, Trowbridge IS, Kornfeld S. 1982. A mouse lymphoma cell line resistant to the leukoagglutinating lectin from Phaseolus vulgaris is deficient in UDP-GlcNAc: Alpha-D-mannoside beta 1,6 N-acetylglucosaminyltransferase. J Biol Chem. 257:13421–13427. [PubMed] [Google Scholar]
238. Reitman ML, Trowbridge IS, Kornfeld S. 1982. A lectin-resistant mouse lymphoma cell line is deficient in glucosidase II, a glycoprotein-processing enzyme. J Biol Chem. 257:10357–10363. [PubMed] [Google Scholar]
239. Esko JD, Stewart TE, Taylor WH. 1985. Animal cell mutants defective in glycosaminoglycan biosynthesis. Proc Natl Acad Sci USA. 82:3197–3201. [PMC free article] [PubMed] [Google Scholar]
240. Esko JD, Weinke JL, Taylor WH, Ekborg G, Roden L, Anantharamaiah G, Gawish A. 1987. Inhibition of chondroitin and heparan sulfate biosynthesis in Chinese hamster ovary cell mutants defective in galactosyltransferase I. J Biol Chem. 262:12189–12195. [PubMed] [Google Scholar]
241. Watkins WM, Morgan WT. 1955. Inhibition by simple sugars of enzymes which decompose the blood-group substances. Nature. 175:676–677. [PubMed] [Google Scholar]
242. Gesner BM, Ginsburg V. 1964. Effect of glycoosidases on the fate of transfused lymphocytes. Proc Natl Acad Sci USA. 52:750–755. [PMC free article] [PubMed] [Google Scholar]
243. Brew K, Vanaman TC, Hill RL. 1968. The role of alpha-lactalbumin and the A protein in lactose synthetase: A unique mechanism for the control of a biological reaction. Proc Natl Acad Sci USA. 59:491–497. [PMC free article] [PubMed] [Google Scholar]
244. Kobata A, Ginsburg V, Tsuda M. 1969. Oligosaccharides of human milk. I. Isolation and characterization. Arch Biochem Biophys. 130:509–513. [PubMed] [Google Scholar]
245. Buck CA, Glick MC, Warren L. 1971. Glycopeptides from the surface of control and virus-transformed cells. Science. 172:169–171. [PubMed] [Google Scholar]
246. Noonan KD, Renger HC, Basilico C, Burger MM. 1973. Surface changes in temperature-sensitive Simian virus 40-transformed cells. Proc Natl Acad Sci USA. 70:347–349. [PMC free article] [PubMed] [Google Scholar]
247. Wickus GG, Robbins PW. 1973. Plasma membrane proteins of normal and Rous sarcoma virus-transformed chick-embryo fibroblasts. Nature. 245:65–67. [PubMed] [Google Scholar]
248. Turner RS, Burger MM. 1973. Involvement of a carbohydrate group in the active site for surface guided reassociation of animal cells. Nature. 244:509–510. [PubMed] [Google Scholar]
249. Morell AG, Gregoriadis G, Scheinberg IH, Hickman J, Ashwell G. 1971. The role of sialic acid in determining the survival of glycoproteins in the circulation. J Biol Chem. 246:1461–1467. [PubMed] [Google Scholar]
250. Ashwell G, Harford J. 1982. Carbohydrate-specific receptors of the liver. Annu Rev Biochem. 51:531–554. [PubMed] [Google Scholar]
251. Schlesinger PH, Doebber TW, Mandell BF, White R, DeSchryver C, Rodman JS, Miller MJ, Stahl P. 1978. Plasma clearance of glycoproteins with terminal mannose and N-acetylglucosamine by liver non-parenchymal cells. Studies with beta-glucuronidase, N-acetyl-beta-D-glucosaminidase, ribonuclease B and agalacto-orosomucoid. Biochem J. 176:103–109. [PMC free article] [PubMed] [Google Scholar]
252. Stahl PD, Rodman JS, Miller MJ, Schlesinger PH. 1978. Evidence for receptor-mediated binding of glycoproteins, glycoconjugates, and lysosomal glycosidases by alveolar macrophages. Proc Natl Acad Sci USA. 75:1399–1403. [PMC free article] [PubMed] [Google Scholar]
253. Kaplan A, Achord DT, Sly WS. 1977. Phosphohexosyl components of a lysosomal enzyme are recognized by pinocytosis receptors on human fibroblasts. Proc Natl Acad Sci U S A. 74:2026–2030. [PMC free article] [PubMed] [Google Scholar]
254. Natowicz MR, Chi MM, Lowry OH, Sly WS. 1979. Enzymatic identification of mannose 6-phosphate on the recognition marker for receptor-mediated pinocytosis of beta-glucuronidase by human fibroblasts. Proc Natl Acad Sci USA. 76:4322–4326. [PMC free article] [PubMed] [Google Scholar]
255. Roche AC, Monsigny M. 1974. Purification and properties of limulin: A lectin (agglutinin) from hemolymph of Limulus polyphemus. Biochim Biophys Acta. 371:242–254. [PubMed] [Google Scholar]
256. Richards RL, Moss J, Alving CR, Fishman PH, Brady RO. 1979. Choleragen (cholera toxin): A bacterial lectin. Proc Natl Acad Sci USA. 76:1673–1676. [PMC free article] [PubMed] [Google Scholar]
257. Bishayee S, Dorai DT. 1980. Isolation and characterisation of a sialic acid-binding lectin (carcinoscorpin) from Indian horseshoe crab Carcinoscorpius rotunda cauda. Biochim Biophys Acta. 623:89–97. [PubMed] [Google Scholar]
258. Monsigny M, Roche AC, Sene C, Maget DR, Delmotte F. 1980. Sugar–lectin interactions: How does wheat-germ agglutinin bind sialoglycoconjugates. Eur J Biochem. 104:147–153. [PubMed] [Google Scholar]
259. Kornfeld K, Reitman ML, Kornfeld R. 1981. The carbohydrate-binding specificity of pea and lentil lectins. Fucose is an important determinant. J Biol Chem. 256:6633–6640. [PubMed] [Google Scholar]
260. Firon N, Ofek I, Sharon N. 1982. Interaction of mannose-containing oligosaccharides with the fimbrial lectin of Escherichia coli. Biochem Biophys Res Commun. 105:1426–1432. [PubMed] [Google Scholar]
261. Miller RL. 1982. A sialic acid-specific lectin from the slug Limax flavus. J Invertebr Pathol. 39:210–214. [PubMed] [Google Scholar]
262. Murray PA, Levine MJ, Tabak LA, Reddy MS. 1982. Specificity of salivary-bacterial interactions: II. Evidence for a lectin on Streptococcus sanguis with specificity for a NeuAc alpha 2, 3Ga1 beta 1, 3Ga1NAc sequence. Biochem Biophys Res Commun. 106:390–396. [PubMed] [Google Scholar]
263. Sharon N, Lis H. 1989. Lectins as cell recognition molecules. Science. 246:227–234. [PubMed] [Google Scholar]
264. Lindahl U, Backstrom G, Hook M, Thunberg L, Fransson LA, Linker A. 1979. Structure of the antithrombin-binding site in heparin. Proc Natl Acad Sci USA. 76:3198–3202. [PMC free article] [PubMed] [Google Scholar]
265. Rosenberg RD, Lam L. 1979. Correlation between structure and function of heparin. Proc Natl Acad Sci USA. 76:1218–1222. [PMC free article] [PubMed] [Google Scholar]
266. Lindahl U, Backstrom G, Thunberg L, Leder IG. 1980. Evidence for a 3-O-sulfated D-glucosamine residue in the antithrombin-binding sequence of heparin. Proc Natl Acad Sci USA. 77:6551–6555. [PMC free article] [PubMed] [Google Scholar]
267. Ruoslahti E. 1988. Introduction. In: Bock G, Harnett S editors. Ciba Foundation Symposium: Carbohydrate Recognition in Cellular Function. New York: Wiley. p. 1–5. [Google Scholar]
268. Leroy JG, Ho MW, MacBrinn MC, Zielke K, Jacob J, OBrien JS. 1972. I-cell disease: Biochemical studies. Pediatr Res. 6:752–757. [PubMed] [Google Scholar]
269. Hickman S, Neufeld EF. 1972. A hypothesis for I-cell disease: Defective hydrolases that do not enter lysosomes. Biochem Biophys Res Commun. 49:992–999. [PubMed] [Google Scholar]
270. Hasilik A, Neufeld EF. 1980. Biosynthesis of lysosomal enzymes in fibroblasts. Phosphorylation of mannose residues. J Biol Chem. 255:4946–4950. [PubMed] [Google Scholar]
271. Tabas I, Kornfeld S. 1980. Biosynthetic intermediates of beta-glucuronidase contain high mannose oligosaccharides with blocked phosphate residues. J Biol Chem. 255:6633–6639. [PubMed] [Google Scholar]
272. Varki A, Kornfeld S. 1980. b. Structural studies of phosphorylated high mannose-type oligosaccharides. J Biol Chem. 255:10847–10858. [PubMed] [Google Scholar]
273. Varki A, Kornfeld S. 1983. The spectrum of anionic oligosaccharides released by endo-beta-N-acetylglucosaminidase H from glycoproteins. Structural studies and interactions with the phosphomannosyl receptor. J Biol Chem. 258:2808–2818. [PubMed] [Google Scholar]
274. Varki A, Kornfeld S. 1980. a. Identification of a rat liver alpha-N-acetylglucosaminyl phosphodiesterase capable of removing “blocking” alpha-N-acetylglucosamine residues from phosphorylated high mannose oligosaccharides of lysosomal enzymes. J Biol Chem. 255:8398–8401. [PubMed] [Google Scholar]
275. Goldberg DE, Kornfeld S. 1981. The phosphorylation of beta-glucuronidase oligosaccharides in mouse P388D1 cells. J Biol Chem. 256:13060–13067. [PubMed] [Google Scholar]
276. Hasilik A, Waheed A, von Figura K. 1981. Enzymatic phosphorylation of lysosomal enzymes in the presence of UDP-N-acetylglucosamine. Absence of the activity in I-cell fibroblasts. Biochem Biophys Res Commun. 98:761–767. [PubMed] [Google Scholar]
277. Reitman ML, Varki A, Kornfeld S. 1981. Fibroblasts from patients with I-cell disease and pseudo-Hurler polydystrophy are deficient in uridine 5’-diphosphate-N-acetylglucosamine:glycoprotein N-acetylglucosaminylphosphotransferase activity. J Clin Invest. 67:1574–1579. [PMC free article] [PubMed] [Google Scholar]
278. Varki A, Reitman ML, Kornfeld S. 1981. Identification of a variant of mucolipidosis III (pseudo-Hurler polydystrophy): A catalytically active N-acetylglucosaminylphosphotransferase that fails to phosphorylate lysosomal enzymes. Proc Natl Acad Sci USA. 78:7773–7777. [PMC free article] [PubMed] [Google Scholar]
279. Waheed A, Hasilik A, von Figura K. 1981. Processing of the phosphorylated recognition marker in lysosomal enzymes. Characterization and partial purification of a microsomal alpha-N-acetylglucosaminyl phosphodiesterase. J Biol Chem. 256:5717–5721. [PubMed] [Google Scholar]
280. Reitman ML, Kornfeld S. 1981. a. Lysosomal enzyme targeting. N-Acetylglucosaminylphosphotransferase selectively phosphorylates native lysosomal enzymes. J Biol Chem. 256:11977–11980. [PubMed] [Google Scholar]
281. Reitman ML, Kornfeld S. 1981. b. UDP-N-acetylglucosamine:glycoprotein N-acetylglucosamine-1-phosphotransferase. Proposed enzyme for the phosphorylation of the high mannose oligosaccharide units of lysosomal enzymes. J Biol Chem. 256:4275–4281. [PubMed] [Google Scholar]
282. Sharp JK, Albersheim P, Ossowski P, Pilotti A, Garegg P, Lindberg B. 1984. a. Comparison of the structures and elicitor activities of a synthetic and a mycelial-wall-derived hexa(beta-D-glucopyranosyl)-D-glucitol. J Biol Chem. 259:11341–11345. [PubMed] [Google Scholar]
283. Sharp JK, Valent B, Albersheim P. 1984. c. Purification and partial characterization of a beta-glucan fragment that elicits phytoalexin accumulation in soybean. J Biol Chem. 259:11312–11320. [PubMed] [Google Scholar]
284. Sharp JK, McNeil M, Albersheim P. 1984. b. The primary structures of one elicitor-active and seven elicitor-inactive hexa(beta-D-glucopyranosyl)-D-glucitols isolated from the mycelial walls of Phytophthora megasperma f. sp. glycinea. J Biol Chem. 259:11321–11336. [PubMed] [Google Scholar]
285. Van TKT, Toubart P, Cousson A, Darvill AG, Gollin DJ, Chelf P, Albersheim P. 1985. Manipulation of the morphogenetic pathways of tobacco explants by oligosaccharins. Nature. 314:615–617. [Google Scholar]
286. Rosen SD, Singer MS, Yednock TA, Stoolman LM. 1985. Involvement of sialic acid on endothelial cells in organ-specific lymphocyte recirculation. Science. 228:1005–1007. [PubMed] [Google Scholar]
287. Watson ML, Kingsmore SF, Johnston GI, Siegelman MH, Le BMM, Lemons RS, Bora NS, Howard TA, Weissman IL, McEver RP, et al.. 1990. Genomic organization of the selectin family of leukocyte adhesion molecules on human and mouse chromosome 1. J Exp Med. 172:263–272. [PMC free article] [PubMed] [Google Scholar]
288. Spertini O, Luscinskas FW, Kansas GS, Munro JM, Griffin JD, Gimbrone MAJ, Tedder TF. 1991. Leukocyte adhesion molecule-1 (LAM-1, L-selectin) interacts with an inducible endothelial cell ligand to support leukocyte adhesion. J Immunol. 147:2565–2573. [PubMed] [Google Scholar]
289. Bevilacqua M, Butcher E, Furie B, Gallatin M, Gimbrone M, Harlan J, Kishimoto K, Lasky L, McEver R, Paulson J, et al.. 1991. Selectins: A family of adhesion receptors. Cell. 67:233–233. [PubMed] [Google Scholar]
290. McEver RP. 1991. Selectins: Novel receptors that mediate leukocyte adhesion during inflammation. Thromb Haemost. 65:223–228. [PubMed] [Google Scholar]
291. Tyrrell D, James P, Rao N, Foxall C, Abbas S, Dasgupta F, Nashed M, Hasegawa A, Kiso M, Asa D, et al.. 1991. Structural requirements for the carbohydrate ligand of E-selectin. Proc Natl Acad Sci USA. 88:10372–10376. [PMC free article] [PubMed] [Google Scholar]
292. Zhou Q, Moore KL, Smith DF, Varki A, McEver RP, Cummings RD. 1991. The selectin GMP-140 binds to sialylated, fucosylated lactosaminoglycans on both myeloid and nonmyeloid cells. J Cell Biol. 115:557–564. [PMC free article] [PubMed] [Google Scholar]
293. Berg EL, Magnani J, Warnock RA, Robinson MK, Butcher EC. 1992. Comparison of L-selectin and E-selectin ligand specificities: The L-selectin can bind the E-selectin ligands sialyl Lex and sialyl Lea. Biochem Biophys Res Commun. 184:1048–1055. [PubMed] [Google Scholar]
294. Cummings RD, Smith DF. 1992. The selectin family of carbohydrate-binding proteins: Structure and importance of carbohydrate ligands for cell adhesion. BioEssays. 14:849–856. [PubMed] [Google Scholar]
295. Green PJ, Tamatani T, Watanabe T, Miyasaka M, Hasegawa A, Kiso M, Yuen C-T, Stoll MS, Feizi T. 1992. High affinity binding of the leucocyte adhesion molecule L-selectin to 3’-sulphated-Lea and -Lex oligosaccharides and the predominance of sulphate in this interaction demonstrated by binding studies with a series of lipid-linked oligosaccharides. Biochem Biophys Res Commun. 188:244–251. [PubMed] [Google Scholar]
296. Imai Y, Lasky LA, Rosen SD. 1992. Further characterization of the interaction between L-selectin and its endothelial ligands. Glycobiology. 2:373–381. [PubMed] [Google Scholar]
297. Larkin M, Ahern TJ, Stoll MS, Shaffer M, Sako D, O'Brien J, Yuen C-T, Lawson AM, Childs RA, Barone KM, et al.. 1992. Spectrum of sialylated and nonsialylated fuco-oligosaccharides bound by the endothelial-leukocyte adhesion molecule E-selectin. Dependence of the carbohydrate binding activity on E-selectin density. J Biol Chem. 267:13661–13668. [PubMed] [Google Scholar]
298. Larsen GR, Sako D, Ahern TJ, Shaffer M, Erban J, Sajer SA, Gibson RM, Wagner DD, Furie BC, Furie B. 1992. P-selectin and E-selectin. Distinct but overlapping leukocyte ligand specificities. J Biol Chem. 267:11104–11110. [PubMed] [Google Scholar]
299. Moore KL, Stults NL, Diaz S, Smith DF, Cummings RD, Varki A, McEver RP. 1992. Identification of a specific glycoprotein ligand for P-selectin (CD62) on myeloid cells. J Cell Biol. 118:445–456. [PMC free article] [PubMed] [Google Scholar]
300. Paavonen T, Renkonen R. 1992. Selective expression of sialyl-Lewis X and Lewis A epitopes, putative ligands for L-selectin, on peripheral lymph-node high endothelial venules. Am J Pathol. 141:1259–1264. [PMC free article] [PubMed] [Google Scholar]
301. Rosen SD. 1993. L-selectin and its biological ligands. Histochemistry. 100:185–191. [PubMed] [Google Scholar]
302. Varki A. 1994. Selectin ligands. Proc Natl Acad Sci USA. 91:7390–7397. [PMC free article] [PubMed] [Google Scholar]
303. Tedder TF, Steeber DA, Chen A, Engel P. 1995. The selectins: Vascular adhesion molecules. FASEB J. 9:866–873. [PubMed] [Google Scholar]
304. Kannagi R. 1997. Carbohydrate-mediated cell adhesion involved in hematogenous metastasis of cancer. Glycoconj J. 14:577–584. [PubMed] [Google Scholar]
305. Varki A. 1997. Selectin ligands: Will the real ones please stand up. J Clin Invest. 99:158–162. [PMC free article] [PubMed] [Google Scholar]
306. Furie B, Furie BC, Flaumenhaft R. 2001. A journey with platelet P-selectin: The molecular basis of granule secretion, signalling and cell adhesion. Thromb Haemost. 86:214–221. [PubMed] [Google Scholar]
307. Varki NM, Varki A. 2002. Heparin inhibition of selectin-mediated interactions during the hematogenous phase of carcinoma metastasis: Rationale for clinical studies in humans. Semin Thromb Hemost. 28:53–66. [PubMed] [Google Scholar]
308. Sackstein R. 2012. Engineering cellular trafficking via glycosyltransferase-programmed stereosubstitution. Ann NY Acad Sci. 1253:193–200. [PMC free article] [PubMed] [Google Scholar]
309. Schnaar RL. 2016. Glycobiology simplified: Diverse roles of glycan recognition in inflammation. J Leukoc Biol. 99:825–838. [PMC free article] [PubMed] [Google Scholar]
310. Torres CR, Hart GW. 1984. Topography and polypeptide distribution of terminal N-acetylglucosamine residues on the surfaces of intact lymphocytes evidence for O-linked GlcNAc. J Biol Chem. 259:3308–3317. [PubMed] [Google Scholar]
311. Holt GD, Hart GW. 1986. The subcellular distribution of terminal N-acetylglucosamine moieties. Localization of a novel protein-saccharide linkage, O-linked GlcNAc. J Biol Chem. 261:8049–8057. [PubMed] [Google Scholar]
312. Holt GD, Haltiwanger RS, Torres CR, Hart GW. 1987. a. Erythrocytes contain cytoplasmic glycoproteins. O-linked GlcNAc on Band 4.1. J Biol Chem. 262:14847–14850. [PubMed] [Google Scholar]
313. Holt GD, Snow CM, Senior A, Haltiwanger RS, Gerace L, Hart GW. 1987. b. Nuclear pore complex glycoproteins contain cytoplasmically disposed O-linked N-acetylglucosamine. J Cell Biol. 104:1157–1164. [PMC free article] [PubMed] [Google Scholar]
314. D'Onofrio M, Starr CM, Park MK, Holt GD, Haltiwanger RS, Hart GW, Hanover JA. 1988. Partial cDNA sequence encoding a nuclear pore protein modified by O-linked N-acetylglucosamine. Proc Natl Acad Sci USA. 85:9595–9599. [PMC free article] [PubMed] [Google Scholar]
315. Hart GW, Haltiwanger RS, Holt GD, Kelly WG. 1989. Glycosylation in the nucleus and cytoplasm. Annu Rev Biochem. 58:841–874. [PubMed] [Google Scholar]
316. Starr CM, Hanover JA. 1990. Glycosylation of nuclear pore protein p62. Reticulocyte lysate catalyzes O-linked N-acetylglucosamine addition in vitro. J Biol Chem. 265:6868–6873. [PubMed] [Google Scholar]
317. Benko DM, Haltiwanger RS, Hart GW, Gibson W. 1988. Virion basic phosphoprotein from human cytomegalovirus contains O-linked N-acetylglucosamine. Proc Natl Acad Sci USA. 85:2573–2577. [PMC free article] [PubMed] [Google Scholar]
318. Hayes BK, Hart GW. 1994. Novel forms of protein glycosylation. Curr Opin Struct Biol. 4:692–696. [PubMed] [Google Scholar]
319. Murphy, J-E, Hanover JA, Froehlich M, DuBois G, Keen JH. 1994. Clathrin assembly protein AP-3 is phosphorylated and glycosylated on the 50-kDa structural domain. J Biol Chem. 269:21346–21352. [PubMed] [Google Scholar]
320. Chou T-Y, Hart GW, Dang CV. 1995. c-Myc is glycosylated at threonine 58, a known phosphorylation site and a mutational hot spot in lymphomas. J Biol Chem. 270:18961–18965. [PubMed] [Google Scholar]
321. Drickamer K. 1988. Two distinct classes of carbohydrate-recognition domains in animal lectins. J Biol Chem. 263:9557–9560. [PubMed] [Google Scholar]
322. Barondes SH, Cooper DNW, Gitt MA, Leffler H. 1994. a. Galectins. Structure and function of a large family of animal lectins. J Biol Chem. 269:20807–20810. [PubMed] [Google Scholar]
323. Barondes SH, Castronovo V, Cooper DNW, Cummings RD, Drickamer K, Feizi T, Gitt MA, Hirabayashi J, Hughes C, Kasai K, et al.. 1994. b. Galectins: A family of animal beta-galactoside-binding lectins. Cell. 76:597–598. [PubMed] [Google Scholar]
324. Crocker PR, Kelm S, Dubois C, Martin B, McWilliam AS, Shotton DM, Paulson JC, Gordon S. 1991. Purification and properties of sialoadhesin, a sialic acid-binding receptor of murine tissue macrophages. EMBO J. 10:1661–1669. [PMC free article] [PubMed] [Google Scholar]
325. Powell LD, Sgroi D, Sjoberg ER, Stamenkovic I, Varki A. 1993. Natural ligands of the B cell adhesion molecule CD22beta carry N-linked oligosaccharides with alpha-2,6-linked sialic acids that are required for recognition. J Biol Chem. 268:7019–7027. [PubMed] [Google Scholar]
326. Crocker PR, Mucklow S, Bouckson V, McWilliam A, Willis AC, Gordon S, Milon G, Kelm S, Bradfield P. 1994. Sialoadhesin, a macrophage sialic acid binding receptor for haemopoietic cells with 17 immunoglobulin-like domains. EMBO J. 13:4490–4503. [PMC free article] [PubMed] [Google Scholar]
327. Kelm S, Pelz A, Schauer R, Filbin MT, Tang S, De Bellard M-E, Schnaar RL, Mahoney JA, Hartnell A, Bradfield P, et al.. 1994. Sialoadhesin, myelin-associated glycoprotein and CD22 define a new family of sialic acid-dependent adhesion molecules of the immunoglobulin superfamily. Curr Biol. 4:965–972. [PubMed] [Google Scholar]
328. Powell LD, Varki A. 1995. I-type lectins. J Biol Chem. 270:14243–14246. [PubMed] [Google Scholar]
329. Crocker PR, Clark EA, Filbin M, Gordon S, Jones Y, Kehrl JH, Kelm S, Le Douarin N, Powell L, Roder J, et al.. 1998. Siglecs: A family of sialic-acid binding lectins [letter]. Glycobiology. 8:v. [PubMed] [Google Scholar]
330. Drickamer K, Taylor ME. 1993. Biology of animal lectins. Annu Rev Cell Biol. 9:237–264. [PubMed] [Google Scholar]
331. Lee JK, Buckhaults P, Wilkes C, Teilhet M, King ML, Moremen KW, Pierce M. 1997. Cloning and expression of a Xenopus laevis oocyte lectin and characterization of its mRNA levels during early development. Glycobiology. 7:367–372. [PubMed] [Google Scholar]
332. Lee JK, Baum LG, Moremen K, Pierce M. 2004. The X-lectins: A new family with homology to the Xenopus laevis oocyte lectin XL-35. Glycoconj J. 21:443–450. [PubMed] [Google Scholar]
333. Wesener DA, Wangkanont K, McBride R, Song X, Kraft MB, Hodges HL, Zarling LC, Splain RA, Smith DF, Cummings RD, et al.. 2015. Recognition of microbial glycans by human intelectin-1. Nat Struct Mol Biol. 22:603–610. [PMC free article] [PubMed] [Google Scholar]
334. Matsushita M, Fujita T. 2001. Ficolins and the lectin complement pathway. Immunol Rev. 180:78–85. [PubMed] [Google Scholar]
335. Holmskov U, Thiel S, Jensenius JC. 2003. Collections and ficolins: Humoral lectins of the innate immune defense. Annu Rev Immunol. 21:547–578. [PubMed] [Google Scholar]
336. Sugahara K, Schwartz NB. 1979. Defect in 3’-phosphoadenosine 5’-phosphosulfate formation in brachymorphic mice. Proc Natl Acad Sci USA. 76:6615–6618. [PMC free article] [PubMed] [Google Scholar]
337. Quentin E, Gladen A, Roden L, Kresse H. 1990. A genetic defect in the biosynthesis of dermatan sulfate proteoglycan: Galactosyltransferase I deficiency in fibroblasts from a patient with a progeroid syndrome. Proc Natl Acad Sci USA. 87:1342–1346. [PMC free article] [PubMed] [Google Scholar]
338. Jaeken J, van Eijk HG, van der Heul C, Corbeel L, Eeckels R, Eggermont E. 1984. Sialic acid-deficient serum and cerebrospinal fluid transferrin in a newly recognized genetic syndrome. Clin Chim Acta. 144:245–247. [PubMed] [Google Scholar]
339. Jaeken J, Eggermont E, Stibler H. 1987. An apparent homozygous X-linked disorder with carbohydrate-deficient serum glycoproteins [letter]. Lancet. 2:1398–1398. [PubMed] [Google Scholar]
340. Stibler H, Jaeken J. 1990. Carbohydrate deficient serum transferrin in a new systemic hereditary syndrome. Arch Dis Child. 65:107–111. [PMC free article] [PubMed] [Google Scholar]
341. Ramaekers VT, Stibler H, Kint J, Jaeken J. 1991. A new variant of the carbohydrate deficient glycoproteins syndrome. J Inherited Metab Dis. 14:385–388. [PubMed] [Google Scholar]
342. Stibler H, Allgulander C, Borg S, Kjellin KG. 1978. Abnormal microheterogeneity of transferrin in serum and cerebrospinal fluid in alcoholism. Acta Med Scand. 204:49–56. [PubMed] [Google Scholar]
343. Yamashita K, Ohkura T, Ideo H, Ohno K, Kanai M. 1993. a. Electrospray ionization-mass spectrometric analysis of serum transferrin isoforms in patients with carbohydrate-deficient glycoprotein syndrome. J Biochem (Tokyo). 114:766–769. [PubMed] [Google Scholar]
344. Yamashita K, Ideo H, Ohkura T, Fukushima K, Yuasa I, Ohno K, Takeshita K. 1993. b. Sugar chains of serum transferrin from patients with carbohydrate deficient glycoprotein syndrome. Evidence of asparagine-N-linked oligosaccharide transfer deficiency. J Biol Chem. 268:5783–5789. [PubMed] [Google Scholar]
345. Jaeken J, Schachter H, Carchon H, De CP, Coddeville B, Spik G. 1994. Carbohydrate deficient glycoprotein syndrome type II: A deficiency in Golgi localised N-acetyl-glucosaminyltransferase II. Arch Dis Child. 71:123–127. [PMC free article] [PubMed] [Google Scholar]
346. Powell LD, Paneerselvam K, Vij R, Diaz S, Manzi AE, Buist N, Freeze H, Varki A. 1994. Carbohydrate-deficient glycoprotein syndrome: Not an N-linked oligosaccharide processing defect, but an abnormality in lipid-linked oligosaccharide biosynthesis. J Clin Invest. 94:1901–1909. [PMC free article] [PubMed] [Google Scholar]
347. Charuk JHM, Tan J, Bernardini M, Haddad S, Reithmeier RAF, Jaeken J, Schachter H. 1995. Carbohydrate-deficient glycoprotein syndrome type II--An autosomal recessive N-acetylglucosaminyltransferase II deficiency different from typical hereditary erythroblastic multinuclearity, with a positive acidified-serum lysis test (HEMPAS). Eur J Biochem. 230:797–805. [PubMed] [Google Scholar]
348. Krasnewich DM, Holt GD, Brantly M, Skovby F, Redwine J, Gahl WA. 1995. Abnormal synthesis of dolichol-linked oligosaccharides in carbohydrate-deficient glycoprotein syndrome. Glycobiology. 5:503–510. [PubMed] [Google Scholar]
349. Marquardt T, Ullrich K, Zimmer P, Hasilik A, Deufel T, Harms E. 1995. Carbohydrate-deficient glycoprotein syndrome (CDGS)—Glycosylation, folding and intracellular transport of newly synthesized glycoproteins. Eur J Cell Biol. 66:268–273. [PubMed] [Google Scholar]
350. Tan J, Dunn J, Jaeken J, Schachter H. 1996. Mutations in the MGAT2 gene controlling complex N-glycan synthesis cause carbohydrate-deficient glycoprotein syndrome type II, an autosomal recessive disease with defective brain development. Am J Hum Genet. 59:810–817. [PMC free article] [PubMed] [Google Scholar]
351. Cormier-Daire V, Amiel J, Vuillaumier-Barrot S, Tan J, Durand G, Munnich A, Le MM, Seta N. 2000. Congenital disorders of glycosylation IIa cause growth retardation, mental retardation, and facial dysmorphism. J Med Genet. 37:875–877. [PMC free article] [PubMed] [Google Scholar]
352. Grünewald S, Imbach T, Huijben K, Rubio-Gozalbo ME, Verrips A, de Klerk JBC, Stroink H, Andel JFD, Van Hove JLK, Wendel U, et al.. 2000. Clinical and biochemical characteristics of congenital disorder of glycosylation type Ic, the first recognized endoplasmic reticulum defect in N-glycan synthesis. Ann Neurol. 47:776–781. [PubMed] [Google Scholar]
353. Imbach T, Schenk B, Schollen E, Burda P, Stutz A, Grünewald S, Bailie NM, King MD, Jaeken J, Matthijs G, et al.. 2000. Deficiency of dolichol-phosphate-mannose synthase-1 causes congenital disorder of glycosylation type Ie. J Clin Invest. 105:233–239. [PMC free article] [PubMed] [Google Scholar]
354. Kim S, Westphal V, Srikrishna G, Mehta DP, Peterson S, Filiano J, Karnes PS, Patterson MC, Freeze HH. 2000. Dolichol phosphate mannose synthase (DPM1) mutations define congenital disorder of glycosylation Ie (CDG-Ie). J Clin Invest. 105:191–198. [PMC free article] [PubMed] [Google Scholar]
355. Matthijs G. 2000. Congenital disorders of glycosylation. Trends Biochem Sci. 25:428–428. [PubMed] [Google Scholar]
356. Orlean P. 2000. Congenital disorders of glycosylation caused by defects in mannose addition during N-linked oligosaccharide assembly. J Clin Invest. 105:131–132. [PMC free article] [PubMed] [Google Scholar]
357. Grünewald S, Matthijs G, Jaeken J. 2002. Congenital disorders of glycosylation: A review. Pediatr Res. 52:618–624. [PubMed] [Google Scholar]
358. Schwartz NB, Domowicz M. 2002. Chondrodysplasias due to proteoglycan defects. Glycobiology. 12:57R–68R. [PubMed] [Google Scholar]
359. Manya H, Sakai K, Kobayashi K, Taniguchi K, Kawakita M, Toda T, Endo T. 2003. Loss-of-function of an N-acetylglucosaminyltransferase, POMGnT1, in muscle-eye-brain disease. Biochem Biophys Res Commun. 306:93–97. [PubMed] [Google Scholar]
360. Marquardt T, Denecke J. 2003. Congenital disorders of glycosylation: Review of their molecular bases, clinical presentations and specific therapies. Eur J Pediatr. 162:359–379. [PubMed] [Google Scholar]
361. Jaeken J, Carchon H. 2004. Congenital disorders of glycosylation: A booming chapter of pediatrics. Curr Opin Pediatr. 16:434–439. [PubMed] [Google Scholar]
362. Kim DS, Hayashi YK, Matsumoto H, Ogawa M, Noguchi S, Murakami N, Sakuta R, Mochizuki M, Michele DE, Campbell KP, et al.. 2004. POMT1 mutation results in defective glycosylation and loss of laminin-binding activity in alpha-DG. Neurology. 62:1009–1011. [PubMed] [Google Scholar]
363. Freeze HH, Aebi M. 2005. Altered glycan structures: The molecular basis of congenital disorders of glycosylation. Curr Opin Struct Biol. 15:490–498. [PubMed] [Google Scholar]
364. Martin PT. 2005. The dystroglycanopathies: The new disorders of O-linked glycosylation. Semin Pediatr Neurol. 12:152–158. [PMC free article] [PubMed] [Google Scholar]
365. Matthijs G. 2005. Research network: EUROGLYCANET: A European network focused on congenital disorders of glycosylation. Eur J Hum Genet. 13:395–397. [PubMed] [Google Scholar]
366. Jaeken J, Matthijs G. 2007. Congenital disorders of glycosylation: A rapidly expanding disease family. Annu Rev Genomics Hum Genet. 8:261–278. [PubMed] [Google Scholar]
367. Haeuptle MA, Hennet T. 2009. Congenital disorders of glycosylation: An update on defects affecting the biosynthesis of dolichol-linked oligosaccharides. Hum Mutat. 30:1628–1641. [PubMed] [Google Scholar]
368. Hennet T. 2009. From glycosylation disorders back to glycosylation: What have we learned. Biochim Biophys Acta. 1792:921–924. [PubMed] [Google Scholar]
369. Schachter H, Freeze HH. 2009. Glycosylation diseases: Quo vadis. Biochim Biophys Acta. 1792:925–930. [PMC free article] [PubMed] [Google Scholar]
370. Jaeken J. 2010. Congenital disorders of glycosylation. Ann NY Acad Sci. 1214:190–198. [PubMed] [Google Scholar]
371. Reynders E, Foulquier F, Annaert W, Matthijs G. 2011. How Golgi glycosylation meets and needs trafficking: The case of the COG complex. Glycobiology. 21:853–863. [PubMed] [Google Scholar]
372. Hennet T. 2012. Diseases of glycosylation beyond classical congenital disorders of glycosylation. Biochim Biophys Acta. 1820:1306–1317. [PubMed] [Google Scholar]
373. Willer T, Lee H, Lommel M, Yoshida-Moriguchi T, de Bernabe DB, Venzke D, Cirak S, Schachter H, Vajsar J, Voit T, et al.. 2012. ISPD loss-of-function mutations disrupt dystroglycan O-mannosylation and cause Walker-Warburg syndrome. Nat Genet. 44:575–580. [PMC free article] [PubMed] [Google Scholar]
374. Freeze HH. 2013. Understanding human glycosylation disorders: Biochemistry leads the charge. J Biol Chem. 288:6936–6945. [PMC free article] [PubMed] [Google Scholar]
375. Freeze HH, Chong JX, Bamshad MJ, Ng BG. 2014. Solving glycosylation disorders: Fundamental approaches reveal complicated pathways. Am J Hum Genet. 94:161–175. [PMC free article] [PubMed] [Google Scholar]
376. Scott K, Gadomski T, Kozicz T, Morava E. 2014. Congenital disorders of glycosylation: New defects and still counting. J Inherit Metab Dis. 37:609–617. [PMC free article] [PubMed] [Google Scholar]
377. Endo T. 2015. Glycobiology of α-dystroglycan and muscular dystrophy. J Biochem. 157:1–12. [PubMed] [Google Scholar]
378. Freeze HH, Eklund EA, Ng BG, Patterson MC. 2015. Neurological aspects of human glycosylation disorders. Annu Rev Neurosci. 38:105–125. [PMC free article] [PubMed] [Google Scholar]
379. He P, Grotzke JE, Ng BG, Gunel M, Jafar-Nejad H, Cresswell P, Enns GM, Freeze HH. 2015. A congenital disorder of deglycosylation: Biochemical characterization of N-glycanase 1 deficiency in patient fibroblasts. Glycobiology. 25:836–844. [PMC free article] [PubMed] [Google Scholar]
380. Heeley J, Shinawi M. 2015. Multi-systemic involvement in NGLY1-related disorder caused by two novel mutations. Am J Med Genet A. 167A:816–820. [PubMed] [Google Scholar]
381. Hennet T, Cabalzar J. 2015. Congenital disorders of glycosylation: A concise chart of glycocalyx dysfunction. Trends Biochem Sci. 40:377–384. [PubMed] [Google Scholar]
382. Suzuki T, Huang C, Fujihira H. 2016. The cytoplasmic peptide:N-glycanase (NGLY1)—Structure, expression and cellular functions. Gene. 577:1–7. [PMC free article] [PubMed] [Google Scholar]
383. Panneerselvam K, Freeze HH. 1996. Mannose corrects altered N-glycosylation in carbohydrate-deficient glycoprotein syndrome fibroblasts. J Clin Invest. 97:1478–1487. [PMC free article] [PubMed] [Google Scholar]
384. Niehues R, Hasilik M, Alton G, Körner C, Schiebe-Sukumar M, Koch HG, Zimmer KP, Wu RR, Harms E, Reiter K, et al.. 1998. Carbohydrate-deficient glycoprotein syndrome type Ib—Phosphomannose isomerase deficiency and mannose therapy. J Clin Invest. 101:1414–1420. [PMC free article] [PubMed] [Google Scholar]
385. Marquardt T, Luhn K, Srikrishna G, Freeze HH, Harms E, Vestweber D. 1999. Correction of leukocyte adhesion deficiency type II with oral fucose. Blood. 94:3976–3985. [PubMed] [Google Scholar]
386. Rush JS, Panneerselvam K, Waechter CJ, Freeze HH. 2000. Mannose supplementation corrects GDP-mannose deficiency in cultured fibroblasts from some patients with Congenital Disorders of Glycosylation (CDG). Glycobiology. 10:829–835. [PubMed] [Google Scholar]
387. Lühn K, Marquardt T, Harms E, Vestweber D. 2001. Discontinuation of fucose therapy in LADII causes rapid loss of selectin ligands and rise of leukocyte counts. Blood. 97:330–332. [PubMed] [Google Scholar]
388. van Karnebeek CD, Bonafé L, Wen XY, Tarailo-Graovac M, Balzano S, Royer-Bertrand B, Ashikov A, Garavelli L, Mammi I, Turolla L, et al.. 2016. NANS-mediated synthesis of sialic acid is required for brain and skeletal development. Nat Genet. 48:777–784. [PubMed] [Google Scholar]
389. Tarailo-Graovac M, Shyr C, Ross CJ, Horvath GA, Salvarinova R, Ye XC, Zhang LH, Bhavsar AP, Lee JJ, Drögemöller BI, et al.. 2016. Exome sequencing and the management of neurometabolic disorders. N Engl J Med. 374:2246–2255. [PMC free article] [PubMed] [Google Scholar]
390. Lencz T, Guha S, Liu C, Rosenfeld J, Mukherjee S, DeRosse P, John M, Cheng L, Zhang C, Badner JA, et al.. 2013. Genome-wide association study implicates NDST3 in schizophrenia and bipolar disorder. Nat Commun. 4:2739. [PMC free article] [PubMed] [Google Scholar]
391. Vazza G, Bertolin C, Scudellaro E, Vettori A, Boaretto F, Rampinelli S, De Sanctis G, Perini G, Peruzzi P, Mostacciuolo ML. 2007. Genome-wide scan supports the existence of a susceptibility locus for schizophrenia and bipolar disorder on chromosome 15q26. Mol Psychiatry. 12:87–93. [PubMed] [Google Scholar]
392. Lee MT, Chen CH, Lee CS, Chen CC, Chong MY, Ouyang WC, Chiu NY, Chuo LJ, Chen CY, Tan HK, et al.. 2010. Genome-wide association study of bipolar I disorder in the Han Chinese population. Mol Psychiatry. 16:548–556. [PubMed] [Google Scholar]
393. McAuley EZ, Scimone A, Tiwari Y, Agahi G, Mowry BJ, Holliday EG, Donald JA, Weickert CS, Mitchell PB, Schofield PR, et al.. 2012. Identification of sialyltransferase 8B as a generalized susceptibility gene for psychotic and mood disorders on chromosome 15q25-26. PLoS One. 7:e38172. [PMC free article] [PubMed] [Google Scholar]
394. Weiss FU, Schurmann C, Guenther A, Ernst F, Teumer A, Mayerle J, Simon P, Völzke H, Radke D, Greinacher A, et al.. 2015. Fucosyltransferase 2 (FUT2) non-secretor status and blood group B are associated with elevated serum lipase activity in asymptomatic subjects, and an increased risk for chronic pancreatitis: A genetic association study. Gut. 64:646–656. [PubMed] [Google Scholar]
395. Saxena R, Saleheen D, Been LF, Garavito ML, Braun T, Bjonnes A, Young R, Ho WK, Rasheed A, Frossard P, et al.. 2013. Genome-wide association study identifies a novel locus contributing to type 2 diabetes susceptibility in Sikhs of Punjabi origin from India In: Diabetes. 62 p. 1746–1755. [PMC free article] [PubMed] [Google Scholar]
396. Varki A, Hooshmand F, Diaz S, Varki NM, Hedrick SM. 1991. Developmental abnormalities in transgenic mice expressing a sialic acid-specific 9-O-acetylesterase. Cell. 65:65–74. [PubMed] [Google Scholar]
397. Ioffe E, Stanley P. 1994. Mice lacking N-acetylglucosaminyltransferase I activity die at mid-gestation, revealing an essential role for complex or hybrid N-linked carbohydrates. Proc Natl Acad Sci USA. 91:728–732. [PMC free article] [PubMed] [Google Scholar]
398. Metzler M, Gertz A, Sarkar M, Schachter H, Schrader JW, Marth JD. 1994. Complex asparagine-linked oligosaccharides are required for morphogenic events during post-implantation development. EMBO J. 13:2056–2065. [PMC free article] [PubMed] [Google Scholar]
399. Thall AD, Maly P, Lowe JB. 1995. Oocyte Gal-alpha1,3Gal epitopes implicated in sperm adhesion to the zona pellucida glycoprotein ZP3 are not required for fertilization in the mouse. J Biol Chem. 270:21437–21440. [PubMed] [Google Scholar]
400. Maly P, Thall AD, Petryniak B, Rogers GE, Smith PL, Marks RM, Kelly RJ, Gersten KM, Cheng GY, Saunders TL, et al.. 1996. The alpha(1,3)Fucosyltransferase Fuc-TVII controls leukocyte trafficking through an essential role in L-, E-, and P-selectin ligand biosynthesis. Cell. 86:643–653. [PubMed] [Google Scholar]
401. Asano M, Furukawa K, Kido M, Matsumoto S, Umesaki Y, Kochibe N, Iwakura Y. 1997. Growth retardation and early death of beta-1,4-galactosyltransferase knockout mice with augmented proliferation and abnormal differentiation of epithelial cells. EMBO J. 16:1850–1857. [PMC free article] [PubMed] [Google Scholar]
402. von Schaewen A, Sturm A, O'Neill J, Chrispeels MJ. 1993. Isolation of a mutant Arabidopsis plant that lacks N-acetyl glucosaminyl transferase I and is unable to synthesize Golgi-modified complex N-linked glycans. Plant Physiol. 102:1109–1118. [PMC free article] [PubMed] [Google Scholar]
403. Strasser R, Stadlmann J, Svoboda B, Altmann F, Glossl J, Mach L. 2005. Molecular basis of N-acetylglucosaminyltransferase I deficiency in Arabidopsis thaliana plants lacking complex N-glycans. Biochem J. 387:385–391. [PMC free article] [PubMed] [Google Scholar]
404. Sarkar M, Leventis PA, Silvescu CI, Reinhold VN, Schachter H, Boulianne GL. 2006. Null mutations in Drosophila N-acetylglucosaminyltransferase I produce defects in locomotion and a reduced life span. J Biol Chem. 281:12776–12785. [PubMed] [Google Scholar]
405. Yamashita T, Hashiramoto A, Haluzik M, Mizukami H, Beck S, Norton A, Kono M, Tsuji S, Daniotti JL, Werth N, et al.. 2003. Enhanced insulin sensitivity in mice lacking ganglioside GM3. Proc Natl Acad Sci USA. 100:3445–3449. [PMC free article] [PubMed] [Google Scholar]
406. Yoshikawa M, Go S, Suzuki S, Suzuki A, Katori Y, Morlet T, Gottlieb SM, Fujiwara M, Iwasaki K, Strauss KA, et al.. 2015. Ganglioside GM3 is essential for the structural integrity and function of cochlear hair cells. Hum Mol Genet. 24:2796–2807. [PMC free article] [PubMed] [Google Scholar]
407. Simpson MA, Cross H, Proukakis C, Priestman DA, Neville DC, Reinkensmeier G, Wang H, Wiznitzer M, Gurtz K, Verganelaki A, et al.. 2004. Infantile-onset symptomatic epilepsy syndrome caused by a homozygous loss-of-function mutation of GM3 synthase. Nat Genet. 36:1225–1229. [PubMed] [Google Scholar]
408. Boccuto L, Aoki K, Flanagan-Steet H, Chen CF, Fan X, Bartel F, Petukh M, Pittman A, Saul R, Chaubey A, et al.. 2014. 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. 23:418–433. [PMC free article] [PubMed] [Google Scholar]
409. Takamiya K, Yamamoto A, Furukawa K, Yamashiro S, Shin M, Okada M, Fukumoto S, Haraguchi M, Takeda N, Fujimura K, et al.. 1996. Mice with disrupted GM2/GD2 synthase gene lack complex gangliosides but exhibit only subtle defects in their nervous system. Proc Natl Acad Sci USA. 93:10662–10667. [PMC free article] [PubMed] [Google Scholar]
410. Fukumoto S, Yamamoto A, Hasegawa T, Abe K, Takamiya K, Okada M, Min ZJ, Furukawa K, Miyazaki H, Tsuji Y, et al.. 1997. Genetic remodeling of gangliosides resulted in the enhanced reactions to the foreign substances in skin. Glycobiology. 7:1111–1120. [PubMed] [Google Scholar]
411. Kawai H, Sango K, Mullin KA, Proia RL. 1998. Embryonic stem cells with a disrupted GD3 synthase gene undergo neuronal differentiation in the absence of b-series gangliosides. J Biol Chem. 273:19634–19638. [PubMed] [Google Scholar]
412. Collins BE, Ito H, Sawada N, Ishida H, Kiso M, Schnaar RL. 1999. Enhanced binding of the neural siglecs, myelin-associated glycoprotein and Schwann cell myelin protein, to Chol-1 (alpha-series) gangliosides and novel sulfated Chol-1 analogs. J Biol Chem. 274:37637–37643. [PubMed] [Google Scholar]
413. Vyas AA, Patel HV, Fromholt SE, Heffer-Lauc M, Vyas KA, Dang J, Schachner M, Schnaar RL. 2002. Gangliosides are functional nerve cell ligands for myelin-associated glycoprotein (MAG), an inhibitor of nerve regeneration. Proc Natl Acad Sci USA. 99:8412–8417. [PMC free article] [PubMed] [Google Scholar]
414. Sheikh KA, Sun J, Liu Y, Kawai H, Crawford TO, Proia RL, Griffin JW, Schnaar RL. 1999. Mice lacking complex gangliosides develop Wallerian degeneration and myelination defects. Proc Natl Acad Sci USA. 96:7532–7537. [PMC free article] [PubMed] [Google Scholar]
415. Okada M, Itoh M, Haraguchi M, Okajima T, Inoue M, Oishi H, Matsuda Y, Iwamoto T, Kawano T, Fukumoto S, et al.. 2002. b-Series ganglioside deficiency exhibits no definite changes in the neurogenesis and the sensitivity to Fas-mediated apoptosis but impairs regeneration of the lesioned hypoglossal nerve. J Biol Chem. 277:1633–1636. [PubMed] [Google Scholar]
416. Yamashita T, Wada R, Sasaki T, Deng C, Bierfreund U, Sandhoff K, Proia RL. 1999. A vital role for glycosphingolipid synthesis during development and differentiation. Proc Natl Acad Sci USA. 96:9142–9147. [PMC free article] [PubMed] [Google Scholar]
417. Marth JD. 1996. a. Complexity in O-linked oligosaccharide biosynthesis engendered by multiple polypeptide N-acetylgalactosaminyltransferases. Glycobiology. 6:701–705. [PubMed] [Google Scholar]
418. Bennett EP, Mandel U, Clausen H, Gerken TA, Fritz TA, Tabak LA. 2012. Control of mucin-type O-glycosylation: A classification of the polypeptide GalNAc-transferase gene family. Glycobiology. 22:736–756. [PMC free article] [PubMed] [Google Scholar]
419. Pedersen ME, Snieckute G, Kagias K, Nehammer C, Multhaupt HA, Couchman JR, Pocock R. 2013. An epidermal microRNA regulates neuronal migration through control of the cellular glycosylation state. Science. 341:1404–1408. [PubMed] [Google Scholar]
420. Agrawal P, Kurcon T, Pilobello KT, Rakus JF, Koppolu S, Liu Z, Batista BS, Eng WS, Hsu KL, Liang Y, et al.. 2014. Mapping posttranscriptional regulation of the human glycome uncovers microRNA defining the glycocode. Proc Natl Acad Sci USA. 111:4338–4343. [PMC free article] [PubMed] [Google Scholar]
421. Lin X, Wei G, Shi Z, Dryer L, Esko JD, Wells DE, Matzuk MM. 2000. Disruption of gastrulation and heparan sulfate biosynthesis in EXT1-deficient mice. Dev Biol. 224:299–311. [PubMed] [Google Scholar]
422. Forsberg E, Kjellén L. 2001. Heparan sulfate: Lessons from knockout mice. J Clin Invest. 108:175–180. [PMC free article] [PubMed] [Google Scholar]
423. Shworak NW, HajMohammadi S, De Agostini AI, Rosenberg RD. 2002. Mice deficient in heparan sulfate 3-O-sulfotransferase-1: Normal hemostasis with unexpected perinatal phenotypes. Glycoconj J. 19:355–361. [PubMed] [Google Scholar]
424. Bishop JR, Schuksz M, Esko JD. 2007. Heparan sulphate proteoglycans fine-tune mammalian physiology. Nature. 446:1030–1037. [PubMed] [Google Scholar]
425. Lindahl U. 2014. A personal voyage through the proteoglycan field. Matrix Biol. 35:3–7. [PubMed] [Google Scholar]
426. Mizumoto S, Yamada S, Sugahara K. 2014. Human genetic disorders and knockout mice deficient in glycosaminoglycan. Biomed Res Int. 2014:495764. [PMC free article] [PubMed] [Google Scholar]
427. Poulain FE, Yost HJ. 2015. Heparan sulfate proteoglycans: A sugar code for vertebrate development. Development. 142:3456–3467. [PMC free article] [PubMed] [Google Scholar]
428. Marth JD. 1996. b. Recent advances in gene mutagenesis by site-directed recombination. J Clin Invest. 97:1999–2002. [PMC free article] [PubMed] [Google Scholar]
429. Tarutani M, Itami S, Okabe M, Ikawa M, Tezuka T, Yoshikawa K, Kinoshita T, Takeda J. 1997. Tissue-specific knockout of the mouse Pig-a gene reveals important roles for GPI-anchored proteins in skin development. Proc Natl Acad Sci USA. 94:7400–7405. [PMC free article] [PubMed] [Google Scholar]
430. Shafi R, Lyer SPN, Ellies LG, O'Donnell N, Marek KW, Chui D, Hart GW, Marth JD. 2000. The O-GlcNAc transferase gene resides on the X chromosome and is essential for embryonic stem cell viability and mouse ontogeny. Proc Natl Acad Sci USA. 97:5735–5739. [PMC free article] [PubMed] [Google Scholar]
431. Takeda J, Miyata T, Kawagoe K, Iida Y, Endo Y, Fujita T, Takahashi M, Kitani T, Kinoshita T. 1993. Deficiency of the GPI anchor caused by a somatic mutation of the PIG-A gene in paroxysmal nocturnal hemoglobinuria. Cell. 73:703–711. [PubMed] [Google Scholar]
432. Almeida AM, Murakami Y, Layton DM, Hillmen P, Sellick GS, Maeda Y, Richards S, Patterson S, Kotsianidis I, Mollica L, et al.. 2006. Hypomorphic promoter mutation in PIGM causes inherited glycosylphosphatidylinositol deficiency. Nat Med. 12:846–851. [PubMed] [Google Scholar]
433. Woo HH, Orbach MJ, Hirsch AM, Hawes MC. 1999. Meristem-localized inducible expression of a UDP-glycosyltransferase gene is essential for growth and development in pea and alfalfa. Plant Cell. 11:2303–2315. [PMC free article] [PubMed] [Google Scholar]
434. Belanger AE, Besra GS, Ford ME, Mikusová K, Belisle JT, Brennan PJ, Inamine JM. 1996. The embAB genes of Mycobacterium avium encode an arabinosyl transferase involved in cell wall arabinan biosynthesis that is the target for the antimycobacterial drug ethambutol. Proc Natl Acad Sci USA. 93:11919–11924. [PMC free article] [PubMed] [Google Scholar]
435. Sassetti CM, Boyd DH, Rubin EJ. 2003. Genes required for mycobacterial growth defined by high density mutagenesis. Mol Microbiol. 48:77–84. [PubMed] [Google Scholar]
436. Beauvais A, Bruneau JM, Mol PC, Buitrago MJ, Legrand R, Latgé JP. 2001. Glucan synthase complex of Aspergillus fumigatus. J Bacteriol. 183:2273–2279. [PMC free article] [PubMed] [Google Scholar]
437. Bowman JC, Hicks PS, Kurtz MB, Rosen H, Schmatz DM, Liberator PA, Douglas CM. 2002. The antifungal echinocandin caspofungin acetate kills growing cells of Aspergillus fumigatus in vitro. Antimicrob Agents Chemother. 46:3001–3012. [PMC free article] [PubMed] [Google Scholar]
438. McFarlane HE, Doring A, Persson S. 2014. The cell biology of cellulose synthesis. Annu Rev Plant Biol. 65:69–94. [PubMed] [Google Scholar]
439. Koch BE, Stougaard J, Spaink HP. 2015. Keeping track of the growing number of biological functions of chitin and its interaction partners in biomedical research. Glycobiology. 25:469–482. [PMC free article] [PubMed] [Google Scholar]
440. O'Neill MA, Ishii T, Albersheim P, Darvill AG. 2004. Rhamnogalacturonan II: Structure and function of a borate cross-linked cell wall pectic polysaccharide. Annu Rev Plant Biol. 55:109–139. [PubMed] [Google Scholar]
441. Hayashi T, Kaida R. 2011. Functions of xyloglucan in plant cells. Mol Plant. 4:17–24. [PubMed] [Google Scholar]
442. Kumar P, Yang M, Haynes BC, Skowyra ML, Doering TL. 2011. Emerging themes in cryptococcal capsule synthesis. Curr Opin Struct Biol. 21:597–602. [PMC free article] [PubMed] [Google Scholar]
443. Gow NA, Hube B. 2012. Importance of the Candida albicans cell wall during commensalism and infection. Curr Opin Microbiol. 15:406–412. [PubMed] [Google Scholar]
444. Atmodjo MA, Hao Z, Mohnen D. 2013. Evolving views of pectin biosynthesis. Annu Rev Plant Biol. 64:747–779. [PubMed] [Google Scholar]
445. Free SJ. 2013. Fungal cell wall organization and biosynthesis. Adv Genet. 81:33–82. [PubMed] [Google Scholar]
446. Pauly M, Gille S, Liu L, Mansoori N, de Souza A, Schultink A, Xiong G. 2013. Hemicellulose biosynthesis. Planta. 238:627–642. [PubMed] [Google Scholar]
447. Burton RA, Fincher GB. 2014. Evolution and development of cell walls in cereal grains. Front Plant Sci. 5:456. [PMC free article] [PubMed] [Google Scholar]
448. Schwientek T, Bennett EP, Flores C, Thacker J, Hollmann M, Reis CA, Behrens J, Mandel U, Keck B, Schafer MA, et al.. 2002. Functional conservation of subfamilies of putative UDP-N-acetylgalactosamine:polypeptide N-acetylgalactosaminyltransferases in Drosophila, Caenorhabditis elegans, and mammals. One subfamily composed of l(2)35Aa is essential in Drosophila. J Biol Chem. 277:22623–22638. [PubMed] [Google Scholar]
449. Ten Hagen KG, Tran DT. 2002. A UDP-GalNAc:polypeptide N-acetylgalactosaminyltransferase is essential for viability in Drosophila melanogaster. J Biol Chem. 277:22616–22622. [PubMed] [Google Scholar]
450. Johansson ME, Sjövall H, Hansson GC. 2013. The gastrointestinal mucus system in health and disease. Nat Rev Gastroenterol Hepatol. 10:352–361. [PMC free article] [PubMed] [Google Scholar]
451. Lillehoj EP, Kato K, Lu W, Kim KC. 2013. Cellular and molecular biology of airway mucins. Int Rev Cell Mol Biol. 303:139–202. [PMC free article] [PubMed] [Google Scholar]
452. Pelaseyed T, Bergstrom JH, Gustafsson JK, Ermund A, Birchenough GM, Schutte A, van der Post S, Svensson F, Rodriguez-Pineiro AM, Nystrom EE, et al.. 2014. The mucus and mucins of the goblet cells and enterocytes provide the first defense line of the gastrointestinal tract and interact with the immune system. Immunol Rev. 260:8–20. [PMC free article] [PubMed] [Google Scholar]
453. An G, Wei B, Xia B, McDaniel JM, Ju T, Cummings RD, Braun J, Xia L. 2007. Increased susceptibility to colitis and colorectal tumors in mice lacking core 3-derived O-glycans. J Exp Med. 204:1417–1429. [PMC free article] [PubMed] [Google Scholar]
454. Wang Y, Ju T, Ding X, Xia B, Wang W, Xia L, He M, Cummings RD. 2010. Cosmc is an essential chaperone for correct protein O-glycosylation. Proc Natl Acad Sci USA. 107:9228–9233. [PMC free article] [PubMed] [Google Scholar]
455. Ju T, Aryal RP, Kudelka MR, Wang Y, Cummings RD. 2014. The Cosmc connection to the Tn antigen in cancer. Cancer Biomark. 14:63–81. [PMC free article] [PubMed] [Google Scholar]
456. Roy MG, Livraghi-Butrico A, Fletcher AA, McElwee MM, Evans SE, Boerner RM, Alexander SN, Bellinghausen LK, Song AS, Petrova YM, et al.. 2014. Muc5b is required for airway defence. Nature. 505:412–416. [PMC free article] [PubMed] [Google Scholar]
457. Watanabe H, Yamada Y, Kimata K. 1998. Roles of aggrecan, a large chondroitin sulfate proteoglycan, in cartilage structure and function. J Biochem (Tokyo). 124:687–693. [PubMed] [Google Scholar]
458. Davies PL, Hew CL. 1990. Biochemistry of fish antifreeze proteins. FASEB J. 4:2460–2468. [PubMed] [Google Scholar]
459. Bouvet V, Ben RN. 2003. Antifreeze glycoproteins: Structure, conformation, and biological applications. Cell Biochem Biophys. 39:133–144. [PubMed] [Google Scholar]
460. Duman JG. 2015. Animal ice-binding (antifreeze) proteins and glycolipids: An overview with emphasis on physiological function. J Exp Biol. 218:1846–1855. [PubMed] [Google Scholar]
461. Bar Dolev M, Braslavsky I, Davies PL. 2016. Ice-binding proteins and their function. Annu Rev Biochem. 85:515–542. [PubMed] [Google Scholar]
462. Furness S, Bryan G, McMillan R, Birchenough S, Worthington HV. 2013. Interventions for the management of dry mouth: Non-pharmacological interventions. Cochrane Database Syst Rev. 9:CD009603. [PMC free article] [PubMed] [Google Scholar]
463. Saleh J, Figueiredo MA, Cherubini K, Salum FG. 2015. Salivary hypofunction: An update on aetiology, diagnosis and therapeutics. Arch Oral Biol. 60:242–255. [PubMed] [Google Scholar]
464. Rogus-Pulia NM, Logemann JA. 2011. Effects of reduced saliva production on swallowing in patients with Sjogren's syndrome. Dysphagia. 26:295–303. [PubMed] [Google Scholar]
465. McDonald JA, Camenisch TD. 2002. Hyaluronan: Genetic insights into the complex biology of a simple polysaccharide. Glycoconj J. 19:331–339. [PubMed] [Google Scholar]
466. Hascall V, Esko JD. 2009. Hyaluronan. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology. Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press; p. 219–228. [PubMed] [Google Scholar]
467. Maytin EV. 2016. Hyaluronan: More than just a wrinkle filler. Glycobiology. 26:553–559. [PMC free article] [PubMed] [Google Scholar]
468. Grimshaw J, Trocha-Grimshaw J, Fisher W, Rice A, Smith S, Spedding P, Duffy J, Mollan R. 1996. Quantitative analysis of hyaluronan in human synovial fluid using capillary electrophoresis. Electrophoresis. 17:396–400. [PubMed] [Google Scholar]
469. Cianflocco AJ. 2013. Viscosupplementation in patients with osteoarthritis of the knee. Postgrad Med. 125:97–105. [PubMed] [Google Scholar]
470. Hwang HS, Sung YM, Lee WS, Kim EC. 2014. Additive effect of preservative-free sodium hyaluronate 0.1% in treatment of dry eye syndrome with diquafosol 3% eye drops. Cornea. 33:935–941. [PubMed] [Google Scholar]
471. Tseng CL, Hung YJ, Chen ZY, Fang HW, Chen KH. 2016. Synergistic effect of artificial tears containing epigallocatechin gallate and hyaluronic acid for the treatment of rabbits with dry eye syndrome. PLoS One. 11:e0157982. [PMC free article] [PubMed] [Google Scholar]
472. Koninkx JF, Mirck MH, Hendriks HG, Mouwen JM, van Dijk JE. 1988. Nippostrongylus brasiliensis: Histochemical changes in the composition of mucins in goblet cells during infection in rats. Exp Parasitol. 65:84–90. [PubMed] [Google Scholar]
473. Birchenough GM, Nyström EE, Johansson ME, Hansson GC. 2016. A sentinel goblet cell guards the colonic crypt by triggering Nlrp6-dependent Muc2 secretion. Science. 352:1535–1542. [PMC free article] [PubMed] [Google Scholar]
474. Kerjaschki D, Vernillo AT, Farquhar MG. 1985. Reduced sialylation of podocalyxin—the major sialoprotein of the rat kidney glomerulus—in aminonucleoside nephrosis. Am J Pathol. 118:343–349. [PMC free article] [PubMed] [Google Scholar]
475. Sawada H, Stukenbrok H, Kerjaschki D, Farquhar MG. 1986. Epithelial polyanion (podocalyxin) is found on the sides but not the soles of the foot processes of the glomerular epithelium. Am J Pathol. 125:309–318. [PMC free article] [PubMed] [Google Scholar]
476. Dekan G, Gabel C, Farquhar MG. 1991. Sulfate contributes to the negative charge of podocalyxin, the major sialoglycoprotein of the glomerular filtration slits. Proc Natl Acad Sci USA. 88:5398–5402. [PMC free article] [PubMed] [Google Scholar]
477. Ito M, Sugihara K, Asaka T, Toyama T, Yoshihara T, Furuichi K, Wada T, Asano M. 2012. Glycoprotein hyposialylation gives rise to a nephrotic-like syndrome that is prevented by sialic acid administration in GNE V572L point-mutant mice. PLoS One. 7:e29873. [PMC free article] [PubMed] [Google Scholar]
478. Weinhold B, Sellmeier M, Schaper W, Blume L, Philippens B, Kats E, Bernard U, Galuska SP, Geyer H, Geyer R, et al.. 2012. Deficits in sialylation impair podocyte maturation. J Am Soc Nephrol. 23:1319–1328. [PMC free article] [PubMed] [Google Scholar]
479. Parthasarathy N, Spiro RG. 1984. Isolation and characterization of the heparan sulfate proteoglycan of the bovine glomerular basement membrane. J Biol Chem. 259:12749–12755. [PubMed] [Google Scholar]
480. Miettinen A, Stow JL, Mentone S, Farquhar MG. 1986. Antibodies to basement membrane heparan sulfate proteoglycans bind to the laminae rarae of the glomerular basement membrane (GBM) and induce subepithelial GBM thickening. J Exp Med. 163:1064–1084. [PMC free article] [PubMed] [Google Scholar]
481. Shimomura H, Spiro RG. 1987. Studies on macromolecular components of human glomerular basement membrane and alterations in diabetes. Decreased levels of heparan sulfate proteoglycan and laminin. Diabetes. 36:374–381. [PubMed] [Google Scholar]
482. Stow JL, Soroka CJ, MacKay K, Striker L, Striker G, Farquhar MG. 1989. Basement membrane heparan sulfate proteoglycan is the main proteoglycan synthesized by glomerular epithelial cells in culture. Am J Pathol. 135:637–646. [PMC free article] [PubMed] [Google Scholar]
483. van den Born J, van den Heuvel LP, Bakker MA, Veerkamp JH, Assmann KJ, Weening JJ, Berden JH. 1993. Distribution of GBM heparan sulfate proteoglycan core protein and side chains in human glomerular diseases. Kidney Int. 43:454–463. [PubMed] [Google Scholar]
484. Harvey SJ, Miner JH. 2008. Revisiting the glomerular charge barrier in the molecular era. Curr Opin Nephrol Hypertens. 17:393–398. [PubMed] [Google Scholar]
485. Gelberg H, Healy L, Whiteley H, Miller LA, Vimr E. 1996. In vivo enzymatic removal of alpha2->6-linked sialic acid from the glomerular filtration barrier results in podocyte charge alteration and glomerular injury. Lab Invest. 74:907–920. [PubMed] [Google Scholar]
486. Groffen AJA, Hop FWH, Tryggvason K, Dijkman H, Assmann KJM, Veerkamp JH, Monnens LAH, Van den Heuvel LPWJ. 1997. Evidence for the existence of multiple heparan sulfate proteoglycans in the human glomerular basement membrane and mesangial matrix. Eur J Biochem. 247:175–182. [PubMed] [Google Scholar]
487. Levidiotis V, Freeman C, Punler M, Martinello P, Creese B, Ferro V, van der Vlag J, Berden JH, Parish CR, Power DA. 2004. a. A synthetic heparanase inhibitor reduces proteinuria in passive Heymann nephritis. J Am Soc Nephrol. 15:2882–2892. [PubMed] [Google Scholar]
488. Levidiotis V, Freeman C, Tikellis C, Cooper ME, Power DA. 2004. b. Heparanase is involved in the pathogenesis of proteinuria as a result of glomerulonephritis. J Am Soc Nephrol. 15:68–78. [PubMed] [Google Scholar]
489. Nakayama K, Natori Y, Sato T, Kimura T, Sugiura A, Sato H, Saito T, Ito S, Natori Y. 2004. Altered expression of NDST-1 messenger RNA in puromycin aminonucleoside nephrosis. J Lab Clin Med. 143:106–114. [PubMed] [Google Scholar]
490. Quaggin SE. 2007. Sizing up sialic acid in glomerular disease. J Clin Invest. 117:1480–1483. [PMC free article] [PubMed] [Google Scholar]
491. Macauley MS, Arlian BM, Rillahan CD, Pang PC, Bortell N, Marcondes MC, Haslam SM, Dell A, Paulson JC. 2014. Systemic blockade of sialylation in mice with a global inhibitor of sialyltransferases. J Biol Chem. 289:35149–35158. [PMC free article] [PubMed] [Google Scholar]
492. Vasudevan D, Takeuchi H, Johar SS, Majerus E, Haltiwanger RS. 2015. Peters plus syndrome mutations disrupt a noncanonical ER quality-control mechanism. Curr Biol. 25:286–295. [PMC free article] [PubMed] [Google Scholar]
493. Xu C, Ng DT. 2015. O-mannosylation: The other glycan player of ER quality control. Semin Cell Dev Biol. 41:129–134. [PubMed] [Google Scholar]
494. Weerapana E, Imperiali B. 2006. Asparagine-linked protein glycosylation: From eukaryotic to prokaryotic systems. Glycobiology. 16:91R–101R. [PubMed] [Google Scholar]
495. Gibson R, Schlesinger S, Kornfeld S. 1979. The nonglycosylated glycoprotein of vesicular stomatitis virus is temperature-sensitive and undergoes intracellular aggregation at elevated temperatures. J Biol Chem. 254:3600–3607. [PubMed] [Google Scholar]
496. Roth MG, Fitzpatrick JP, Compans RW. 1979. Polarity of influenza and vesicular stomatitis virus maturation in MDCK cells: Lack of a requirement for glycosylation of viral glycoproteins. Proc Natl Acad Sci USA. 76:6430–6434. [PMC free article] [PubMed] [Google Scholar]
497. Culyba EK, Price JL, Hanson SR, Dhar A, Wong CH, Gruebele M, Powers ET, Kelly JW. 2011. Protein native-state stabilization by placing aromatic side chains in N-glycosylated reverse turns. Science. 331:571–575. [PMC free article] [PubMed] [Google Scholar]
498. Kang HJ, Lee C, Drew D. 2013. Breaking the barriers in membrane protein crystallography. Int J Biochem Cell Biol. 45:636–644. [PubMed] [Google Scholar]
499. Okamoto R, Izumi M, Kajihara Y. 2014. Decoration of proteins with sugar chains: Recent advances in glycoprotein synthesis. Curr Opin Chem Biol. 22:92–99. [PubMed] [Google Scholar]
500. Columbus L. 2015. Post-expression strategies for structural investigations of membrane proteins. Curr Opin Struct Biol. 32:131–138. [PMC free article] [PubMed] [Google Scholar]
501. Lyumkis D, Julien JP, de Val N, Cupo A, Potter CS, Klasse PJ, Burton DR, Sanders RW, Moore JP, Carragher B, et al.. 2013. Cryo-EM structure of a fully glycosylated soluble cleaved HIV-1 envelope trimer. Science. 342:1484–1490. [PMC free article] [PubMed] [Google Scholar]
502. Wibmer CK, Moore PL, Morris L. 2015. HIV broadly neutralizing antibody targets. Curr Opin HIV AIDS. 10:135–143. [PMC free article] [PubMed] [Google Scholar]
503. Julien JP, Cupo A, Sok D, Stanfield RL, Lyumkis D, Deller MC, Klasse PJ, Burton DR, Sanders RW, Moore JP, et al.. 2013. Crystal structure of a soluble cleaved HIV-1 envelope trimer. Science. 342:1477–1483. [PMC free article] [PubMed] [Google Scholar]
504. Gilhespy-Muskett AM, Partridge J, Jefferis R, Homans SW. 1994. A novel 13C isotopic labelling strategy for probing the structure and dynamics of glycan chains in situ on glycoproteins. Glycobiology. 4:485–489. [PubMed] [Google Scholar]
505. Wormald MR, Rudd PM, Harvey DJ, Chang SC, Scragg IG, Dwek RA. 1997. Variations in oligosaccharide-protein interactions in immunoglobulin G determine the site-specific glycosylation profiles and modulate the dynamic motion of the Fc oligosaccharides. Biochemistry. 36:1370–1380. [PubMed] [Google Scholar]
506. Crispin M, Yu X, Bowden TA. 2013. Crystal structure of sialylated IgG Fc: Implications for the mechanism of intravenous immunoglobulin therapy. Proc Natl Acad Sci USA. 110:E3544–E3546. [PMC free article] [PubMed] [Google Scholar]
507. McLellan JS, Pancera M, Carrico C, Gorman J, Julien JP, Khayat R, Louder R, Pejchal R, Sastry M, Dai K, et al.. 2011. Structure of HIV-1 gp120 V1/V2 domain with broadly neutralizing antibody PG9. Nature. 480:336–343. [PMC free article] [PubMed] [Google Scholar]
508. Pejchal R, Doores KJ, Walker LM, Khayat R, Huang PS, Wang SK, Stanfield RL, Julien JP, Ramos A, Crispin M, et al.. 2011. A potent and broad neutralizing antibody recognizes and penetrates the HIV glycan shield. Science. 334:1097–1103. [PMC free article] [PubMed] [Google Scholar]
509. Garces F, Sok D, Kong L, McBride R, Kim HJ, Saye-Francisco KF, Julien JP, Hua Y, Cupo A, Moore JP, et al.. 2014. Structural evolution of glycan recognition by a family of potent HIV antibodies. Cell. 159:69–79. [PMC free article] [PubMed] [Google Scholar]
510. Loomes KM, Senior HE, West PM, Roberton AM. 1999. Functional protective role for mucin glycosylated repetitive domains. Eur J Biochem. 266:105–111. [PubMed] [Google Scholar]
511. Wahrenbrock M, Borsig L, Le D, Varki N, Varki A. 2003. Selectin–mucin interactions as a probable molecular explanation for the association of Trousseau syndrome with mucinous adenocarcinomas. J Clin Invest. 112:853–862. [PMC free article] [PubMed] [Google Scholar]
512. Wahrenbrock MG, Varki A. 2006. Multiple hepatic receptors cooperate to eliminate secretory mucins aberrantly entering the bloodstream: Are circulating cancer mucins the “tip of the iceberg”. Cancer Res. 66:2433–2441. [PubMed] [Google Scholar]
513. Alemka A, Nothaft H, Zheng J, Szymanski CM. 2013. N-glycosylation of Campylobacter jejuni surface proteins promotes bacterial fitness. Infect Immun. 81:1674–1682. [PMC free article] [PubMed] [Google Scholar]
514. Schjoldager KT, Clausen H. 2012. Site-specific protein O-glycosylation modulates proprotein processing—deciphering specific functions of the large polypeptide GalNAc-transferase gene family. Biochim Biophys Acta. 1820:2079–2094. [PubMed] [Google Scholar]
515. Zhang L, Syed ZA, van Dijk Härd I, Lim JM, Wells L, Ten Hagen KG. 2014. O-glycosylation regulates polarized secretion by modulating Tango1 stability. Proc Natl Acad Sci USA. 111:7296–7301. [PMC free article] [PubMed] [Google Scholar]
516. Lingwood CA, Hakomori S. 1977. Selective inhibition of cell growth and associated changes in glycolipid metabolism induced by monovalent antibodies to glycolipids. Exp Cell Res. 108:385–391. [PubMed] [Google Scholar]
517. Bremer EG, Schlessinger J, Hakomori S. 1986. Ganglioside-mediated modulation of cell growth. Specific effects of GM3 on tyrosine phosphorylation of the epidermal growth factor receptor. J Biol Chem. 261:2434–2440. [PubMed] [Google Scholar]
518. Zhou Q, Hakomori S, Kitamura K, Igarashi Y. 1994. GM3 directly inhibits tyrosine phosphorylation and De-N-acetyl-GM3 directly enhances serine phosphorylation of epidermal growth factor receptor, independently of receptor–receptor interaction. J Biol Chem. 269:1959–1965. [PubMed] [Google Scholar]
519. Rebbaa A, Hurh J, Yamamoto H, Kersey DS, Bremer EG. 1996. Ganglioside GM3 inhibition of EGF receptor mediated signal transduction. Glycobiology. 6:399–406. [PubMed] [Google Scholar]
520. Wang XQ, Sun P, Paller AS. 2003. Ganglioside GM3 blocks the activation of epidermal growth factor receptor induced by integrin at specific tyrosine sites. J Biol Chem. 278:48770–48778. [PubMed] [Google Scholar]
521. Yoon SJ, Nakayama K, Hikita T, Handa K, Hakomori SI. 2006. Epidermal growth factor receptor tyrosine kinase is modulated by GM3 interaction with N-linked GlcNAc termini of the receptor. Proc Natl Acad Sci USA. 103:18987–18991. [PMC free article] [PubMed] [Google Scholar]
522. Coskun U, Grzybek M, Drechsel D, Simons K. 2011. Regulation of human EGF receptor by lipids. Proc Natl Acad Sci USA. 108:9044–9048. [PMC free article] [PubMed] [Google Scholar]
523. Tagami S, Inokuchi J, Kabayama K, Yoshimura H, Kitamura F, Uemura S, Ogawa C, Ishii A, Saito M, Ohtsuka Y, et al.. 2002. Ganglioside GM3 participates in the pathological conditions of insulin resistance. J Biol Chem. 277:3085–3092. [PubMed] [Google Scholar]
524. Kabayama K, Sato T, Kitamura F, Uemura S, Kang BW, Igarashi Y, Inokuchi J. 2005. TNFalpha-induced insulin resistance in adipocytes as a membrane microdomain disorder: Involvement of ganglioside GM3. Glycobiology. 15:21–29. [PubMed] [Google Scholar]
525. Rapraeger AC, Krufka A, Olwin BB. 1991. Requirement of heparan sulfate for bFGF-mediated fibroblast growth and myoblast differentiation. Science. 252:1705–1708. [PubMed] [Google Scholar]
526. Yayon A, Klagsbrun M, Esko JD, Leder P, Ornitz DM. 1991. Cell surface, heparin-like molecules are required for binding of basic fibroblast growth factor to its high affinity receptor. Cell. 64:841–848. [PubMed] [Google Scholar]
527. Wang X, Inoue S, Gu J, Miyoshi E, Noda K, Li W, Mizuno-Horikawa Y, Nakano M, Asahi M, Takahashi M, et al.. 2005. Dysregulation of TGF-beta1 receptor activation leads to abnormal lung development and emphysema-like phenotype in core fucose-deficient mice. Proc Natl Acad Sci USA. 102:15791–15796. [PMC free article] [PubMed] [Google Scholar]
528. Wang X, Gu J, Ihara H, Miyoshi E, Honke K, Taniguchi N. 2006. Core fucosylation regulates epidermal growth factor receptor-mediated intracellular signaling. J Biol Chem. 281:2572–2577. [PubMed] [Google Scholar]
529. Liu YC, Yen HY, Chen CY, Chen CH, Cheng PF, Juan YH, Chen CH, Khoo KH, Yu CJ, Yang PC, et al.. 2011. Sialylation and fucosylation of epidermal growth factor receptor suppress its dimerization and activation in lung cancer cells. Proc Natl Acad Sci USA. 108:11332–11337. [PMC free article] [PubMed] [Google Scholar]
530. Yen HY, Liu YC, Chen NY, Tsai CF, Wang YT, Chen YJ, Hsu TL, Yang PC, Wong CH. 2015. Effect of sialylation on EGFR phosphorylation and resistance to tyrosine kinase inhibition. Proc Natl Acad Sci USA. 112:6955–6960. [PMC free article] [PubMed] [Google Scholar]
531. Okajima T, Irvine KD. 2002. Regulation of notch signaling by O-linked fucose. Cell. 111:893–904. [PubMed] [Google Scholar]
532. Shi S, Stanley P. 2003. Protein O-fucosyltransferase 1 is an essential component of Notch signaling pathways. Proc Natl Acad Sci USA. 100:5234–5239. [PMC free article] [PubMed] [Google Scholar]
533. Acar M, Jafar-Nejad H, Takeuchi H, Rajan A, Ibrani D, Rana NA, Pan H, Haltiwanger RS, Bellen HJ. 2008. Rumi is a CAP10 domain glycosyltransferase that modifies Notch and is required for Notch signaling. Cell. 132:247–258. [PMC free article] [PubMed] [Google Scholar]
534. Fernandez-Valdivia R, Takeuchi H, Samarghandi A, Lopez M, Leonardi J, Haltiwanger RS, Jafar-Nejad H. 2011. Regulation of mammalian Notch signaling and embryonic development by the protein O-glucosyltransferase Rumi. Development. 138:1925–1934. [PMC free article] [PubMed] [Google Scholar]
535. Stanley P, Okajima T. 2010. Roles of glycosylation in Notch signaling. Curr Top Dev Biol. 92:131–164. [PubMed] [Google Scholar]
536. Takeuchi H, Haltiwanger RS. 2014. Significance of glycosylation in Notch signaling. Biochem Biophys Res Commun. 453:235–242. [PMC free article] [PubMed] [Google Scholar]
537. Luca VC, Jude KM, Pierce NW, Nachury MV, Fischer S, Garcia KC. 2015. Structural biology. Structural basis for Notch1 engagement of Delta-like 4. Science. 347:847–853. [PMC free article] [PubMed] [Google Scholar]
538. Brown DA, Rose JK. 1992. Sorting of GPI-anchored proteins to glycolipid-enriched membrane subdomains during transport to the apical cell surface. Cell. 68:533–544. [PubMed] [Google Scholar]
539. Varma R, Mayor S. 1998. GPI-anchored proteins are organized in submicron domains at the cell surface. Nature. 394:798–801. [PubMed] [Google Scholar]
540. Hansen GH, Immerdal L, Thorsen E, Niels-Christiansen LL, Nystrøm BT, Demant EJ, Danielsen EM. 2001. Lipid rafts exist as stable cholesterol-independent microdomains in the brush border membrane of enterocytes. J Biol Chem. 276:32338–32344. [PubMed] [Google Scholar]
541. Paladino S, Pocard T, Catino MA, Zurzolo C. 2006. GPI-anchored proteins are directly targeted to the apical surface in fully polarized MDCK cells. J Cell Biol. 172:1023–1034. [PMC free article] [PubMed] [Google Scholar]
542. Wier M, Edidin M. 1988. Constraint of the translational diffusion of a membrane glycoprotein by its external domains. Science. 242:412–414. [PubMed] [Google Scholar]
543. Barbour S, Edidin M. 1992. Cell-specific constraints to the lateral diffusion of a membrane glycoprotein. J Cell Physiol. 150:526–533. [PubMed] [Google Scholar]
544. Cohen M, Hurtado-Ziola N, Varki A. 2009. ABO blood group glycans modulate sialic acid recognition on erythrocytes. Blood. 114:3668–3676. [PMC free article] [PubMed] [Google Scholar]
545. Paszek MJ, DuFort CC, Rossier O, Bainer R, Mouw JK, Godula K, Hudak JE, Lakins JN, Wijekoon AC, Cassereau L, et al.. 2014. The cancer glycocalyx mechanically primes integrin-mediated growth and survival. Nature. 511:319–325. [PMC free article] [PubMed] [Google Scholar]
546. El Maarouf A, Petridis AK, Rutishauser U. 2006. Use of polysialic acid in repair of the central nervous system. Proc Natl Acad Sci USA. 103:16989–16994. [PMC free article] [PubMed] [Google Scholar]
547. Zhang Y, Ghadiri-Sani M, Zhang X, Richardson PM, Yeh J, Bo X. 2007. Induced expression of polysialic acid in the spinal cord promotes regeneration of sensory axons. Mol Cell Neurosci. 35:109–119. [PubMed] [Google Scholar]
548. Rockle I, Seidenfaden R, Weinhold B, Muhlenhoff M, Gerardy-Schahn R, Hildebrandt H. 2008. Polysialic acid controls NCAM-induced differentiation of neuronal precursors into calretinin-positive olfactory bulb interneurons. Dev Neurobiol. 68:1170–1184. [PubMed] [Google Scholar]
549. Rutishauser U. 2008. Polysialic acid in the plasticity of the developing and adult vertebrate nervous system. Nat Rev Neurosci. 9:26–35. [PubMed] [Google Scholar]
550. Battista D, Rutishauser U. 2010. Removal of polysialic acid triggers dispersion of subventricularly derived neuroblasts into surrounding CNS tissues. J Neurosci. 30:3995–4003. [PMC free article] [PubMed] [Google Scholar]
551. Vitureira N, Andres R, Perez-Martinez E, Martinez A, Bribian A, Blasi J, Chelliah S, Lopez-Domenech G, De Castro F, Burgaya F, et al.. 2010. Podocalyxin is a novel polysialylated neural adhesion protein with multiple roles in neural development and synapse formation. PLoS One. 5:e12003. [PMC free article] [PubMed] [Google Scholar]
552. Jungnickel J, Eckhardt M, Haastert-Talini K, Claus P, Bronzlik P, Lipokatic-Takacs E, Maier H, Gieselmann V, Grothe C. 2012. Polysialyltransferase overexpression in Schwann cells mediates different effects during peripheral nerve regeneration. Glycobiology. 22:107–115. [PubMed] [Google Scholar]
553. McCall T, Weil ZM, Nacher J, Bloss EB, El Marouf A, Rutishauser U, McEwen BS. 2012. Depletion of polysialic acid from neural cell adhesion molecule (PSA-NCAM) increases CA3 dendritic arborization and increases vulnerability to excitotoxicity. Exp Neurol. 241C:5–12. [PMC free article] [PubMed] [Google Scholar]
554. Hildebrandt H, Dityatev A. 2015. Polysialic acid in brain development and synaptic plasticity. Top Curr Chem. 366:55–96. [PubMed] [Google Scholar]
555. Espinosa-Marzal RM, Fontani G, Reusch FB, Roba M, Spencer ND, Crockett R. 2013. Sugars communicate through water: Oriented glycans induce water structuring. Biophys J. 104:2686–2694. [PMC free article] [PubMed] [Google Scholar]
556. Moscatelli D. 1992. Basic fibroblast growth factor (bFGF) dissociates rapidly from heparan sulfates but slowly from receptors. Implications for mechanisms of bFGF release from pericellular matrix. J Biol Chem. 267:25803–25809. [PubMed] [Google Scholar]
557. Rahmoune H, Chen HL, Gallagher JT, Rudland PS, Fernig DG. 1998. Interaction of heparan sulfate from mammary cells with acidic fibroblast growth factor (FGF) and basic FGF—Regulation of the activity of basic FGF by high and low affinity binding sites in heparan sulfate. J Biol Chem. 273:7303–7310. [PubMed] [Google Scholar]
558. Li J, Dai G, Cheng YB, Qi X, Geng MY. 2011. Polysialylation promotes neural cell adhesion molecule-mediated cell migration in a fibroblast growth factor receptor-dependent manner, but independent of adhesion capability. Glycobiology. 21:1010–1018. [PubMed] [Google Scholar]
559. Ono S, Hane M, Kitajima K, Sato C. 2012. Novel regulation of fibroblast growth factor 2 (FGF2)-mediated cell growth by polysialic acid. J Biol Chem. 287:3710–3722. [PMC free article] [PubMed] [Google Scholar]
560. Cruz-Chu ER, Malafeev A, Pajarskas T, Pivkin IV, Koumoutsakos P. 2014. Structure and response to flow of the glycocalyx layer. Biophys J. 106:232–243. [PMC free article] [PubMed] [Google Scholar]
561. Andersson-Sjöland A, Hallgren O, Rolandsson S, Weitoft M, Tykesson E, Larsson-Callerfelt AK, Rydell-Törmänen K, Bjermer L, Malmström A, Karlsson JC, et al.. 2015. Versican in inflammation and tissue remodeling: The impact on lung disorders. Glycobiology. 25:243–251. [PMC free article] [PubMed] [Google Scholar]
562. Alonso MD, Lomako J, Lomako WM, Whelan WJ. 1995. a. Catalytic activities of glycogenin additional to autocatalytic self-glucosylation. J Biol Chem. 270:15315–15319. [PubMed] [Google Scholar]
563. Alonso MD, Lomako J, Lomako WM, Whelan WJ. 1995. b. A new look at the biogenesis of glycogen. FASEB J. 9:1126–1137. [PubMed] [Google Scholar]
564. Carrizo ME, Miozzo MC, Goldraij A, Curtino JA. 1997. Purification of rabbit skeletal muscle proteoglycogen: Studies on the glucosyltransferase activity of polysaccharide-free and -bound glycogenin. Glycobiology. 7:571–578. [PubMed] [Google Scholar]
565. Zeqiraj E, Tang X, Hunter RW, García-Rocha M, Judd A, Deak M, von Wilamowitz-Moellendorff A, Kurinov I, Guinovart JJ, Tyers M, et al.. 2014. Structural basis for the recruitment of glycogen synthase by glycogenin. Proc Natl Acad Sci USA. 111:E2831–E2840. [PMC free article] [PubMed] [Google Scholar]
566. Zeqiraj E, Sicheri F. 2015. Getting a handle on glycogen synthase—Its interaction with glycogenin. Mol Aspects Med. 46:63–69. [PMC free article] [PubMed] [Google Scholar]
567. Adeva-Andany MM, González-Lucán M, Donapetry-García C, Fernández-Fernández C, Ameneiros-Rodríguez E. 2016. Glycogen metabolism in humans. BBA Clin. 5:85–100. [PMC free article] [PubMed] [Google Scholar]
568. Mukerjea R, Robyt JF. 2013. Tests for the mechanism of starch biosynthesis: De novo synthesis or an amylogenin primer synthesis. Carbohydr Res. 372:55–59. [PubMed] [Google Scholar]
569. Inatani M, Irie F, Plump AS, Tessier-Lavigne M, Yamaguchi Y. 2003. Mammalian brain morphogenesis and midline axon guidance require heparan sulfate. Science. 302:1044–1046. [PubMed] [Google Scholar]
570. Chen E, Stringer SE, Rusch MA, Selleck SB, Ekker SC. 2005. A unique role for 6-O sulfation modification in zebrafish vascular development. Dev Biol. 284:364–376. [PubMed] [Google Scholar]
571. Schwartz NB, Domowicz MS. 2014. Chemistry and function of glycosaminoglycans in the nervous system. Adv Neurobiol. 9:89–115. [PubMed] [Google Scholar]
572. Balasubramanian R, Zhang X. 2016. Mechanisms of FGF gradient formation during embryogenesis. Semin Cell Dev Biol. 53:94–100. [PMC free article] [PubMed] [Google Scholar]
573. Iozzo RV. 2005. Basement membrane proteoglycans: From cellar to ceiling. Nat Rev Mol Cell Biol. 6:646–656. [PubMed] [Google Scholar]
574. Farach-Carson MC, Warren CR, Harrington DA, Carson DD. 2014. Border patrol: Insights into the unique role of perlecan/heparan sulfate proteoglycan 2 at cell and tissue borders. Matrix Biol. 34:64–79. [PMC free article] [PubMed] [Google Scholar]
575. Aspberg A. 2012. The different roles of aggrecan interaction domains. J Histochem Cytochem. 60:987–996. [PMC free article] [PubMed] [Google Scholar]
576. Mohammadi H, Mequanint K, Herzog W. 2013. Computational aspects in mechanical modeling of the articular cartilage tissue. Proc Inst Mech Eng H. 227:402–420. [PubMed] [Google Scholar]
577. Pap T, Bertrand J. 2013. Syndecans in cartilage breakdown and synovial inflammation. Nat Rev Rheumatol. 9:43–55. [PubMed] [Google Scholar]
578. Melrose J, Shu C, Whitelock JM, Lord MS. 2016. The cartilage extracellular matrix as a transient developmental scaffold for growth plate maturation. Matrix Biol. 52–54:363–383. [PubMed] [Google Scholar]
579. Davies DG, Parsek MR, Pearson JP, Iglewski BH, Costerton JW, Greenberg EP. 1998. The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science. 280:295–298. [PubMed] [Google Scholar]
580. Neu TR, Swerhone GDW, Lawrence JR. 2001. Assessment of lectin-binding analysis for in situ detection of glycoconjugates in biofilm systems. Microbiology. 147:299–313. [PubMed] [Google Scholar]
581. Flemming HC, Wingender J. 2010. The biofilm matrix. Nat Rev Microbiol. 8:623–633. [PubMed] [Google Scholar]
582. Brussow H. 2013. Bacteriophage-host interaction: From splendid isolation into a messy reality. Curr Opin Microbiol. 16:500–506. [PubMed] [Google Scholar]
583. Gunn JS, Bakaletz LO, Wozniak DJ. 2016. What's on the outside matters: The role of the extracellular polymeric substance of Gram-negative biofilms in evading host immunity and as a target for therapeutic intervention. J Biol Chem. 291:12538–12546. [PMC free article] [PubMed] [Google Scholar]
584. Cobey S, Wilson P, Matsen FA. 2015. The evolution within us. Philos Trans R Soc Lond B Biol Sci. 370, 10.1098/rstb.2014.0235. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
585. Hoogeboom R, Tolar P. 2016. Molecular mechanisms of B cell antigen gathering and endocytosis. Curr Top Microbiol Immunol. 393:45–63. [PubMed] [Google Scholar]
586. Brazin KN, Mallis RJ, Das DK, Feng Y, Hwang W, Wang JH, Wagner G, Lang MJ, Reinherz EL. 2015. Structural features of the αβTCR mechanotransduction apparatus that promote pMHC siscrimination. Front Immunol. 6:441. [PMC free article] [PubMed] [Google Scholar]
587. Haurum JS, Arsequell G, Lellouch AC, Wong SYC, Dwek RA, McMichael AJ, Elliott T. 1994. Recognition of carbohydrate by major histocompatibility complex class I-restricted, glycopeptide-specific cytotoxic T lymphocytes. J Exp Med. 180:739–744. [PMC free article] [PubMed] [Google Scholar]
588. Abdel-Motal UM, Berg L, Rostén A, Bengtsson M, Thorpe CJ, Kihlberg J, Dahmén J, Magnusson G, Karlsson KA, Jondal M. 1996. Immunization with glycosylated Kb-binding peptides generates carbohydrate-specific, unrestricted cytotoxic T cells. Eur J Immunol. 26:544–551. [PubMed] [Google Scholar]
589. Galli-Stampino L, Meinjohanns E, Frische K, Meldal M, Jensen T, Werdelin O, Mouritsen S. 1997. T-cell recognition of tumor-associated carbohydrates: The nature of the glycan moiety plays a decisive role in determining glycopeptide immunogenicity. Cancer Res. 57:3214–3222. [PubMed] [Google Scholar]
590. Corthay A, Bäcklund J, Holmdahl R. 2001. Role of glycopeptide-specific T cells in collagen-induced arthritis: An example how post-translational modification of proteins may be involved in autoimmune disease. Ann Med. 33:456–465. [PubMed] [Google Scholar]
591. Gad M, Werdelin O, Meldal M, Komba S, Jensen T. 2002. Characterization of T cell hybridomas raised against a glycopeptide containing the tumor-associated T antigen, (betaGal (1-3) alphaGalNAc-O/Ser). Glycoconj J. 19:59–65. [PubMed] [Google Scholar]
592. Wang CC, Chen JR, Tseng YC, Hsu CH, Hung YF, Chen SW, Chen CM, Khoo KH, Cheng TJ, Cheng YS, et al.. 2009. Glycans on influenza hemagglutinin affect receptor binding and immune response. Proc Natl Acad Sci USA. 106:18137–18142. [PMC free article] [PubMed] [Google Scholar]
593. Das SR, Hensley SE, David A, Schmidt L, Gibbs JS, Puigbo P, Ince WL, Bennink JR, Yewdell JW. 2011. Fitness costs limit influenza A virus hemagglutinin glycosylation as an immune evasion strategy. Proc Natl Acad Sci USA. 108:E1417–E1422. [PMC free article] [PubMed] [Google Scholar]
594. Walker LM, Huber M, Doores KJ, Falkowska E, Pejchal R, Julien JP, Wang SK, Ramos A, Chan-Hui PY, Moyle M, et al.. 2011. Broad neutralization coverage of HIV by multiple highly potent antibodies. Nature. 477:466–470. [PMC free article] [PubMed] [Google Scholar]
595. Reese AJ, Doering TL. 2003. Cell wall alpha-1,3-glucan is required to anchor the Cryptococcus neoformans capsule. Mol Microbiol. 50:1401–1409. [PubMed] [Google Scholar]
596. Reese AJ, Yoneda A, Breger JA, Beauvais A, Liu H, Griffith CL, Bose I, Kim MJ, Skau C, Yang S, et al.. 2007. Loss of cell wall alpha(1-3) glucan affects Cryptococcus neoformans from ultrastructure to virulence. Mol Microbiol. 63:1385–1398. [PMC free article] [PubMed] [Google Scholar]
597. Stanley P, Schachter H, Taniguchi N. 2009. N-Glycans. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME editors. Essentials of Glycobiology Cold Spring Harbor (NY), Cold Spring Harbor Laboratory Press; p. 101–114. [PubMed] [Google Scholar]
598. Demetriou M, Granovsky M, Quaggin S, Dennis JW. 2001. Negative regulation of T-cell activation and autoimmunity by Mgat5 N-glycosylation. Nature. 409:733–739. [PubMed] [Google Scholar]
599. Dennis JW, Laferte S, Waghorne C, Breitman ML, Kerbel RS. 1987. Beta 1-6 branching of Asn-linked oligosaccharides is directly associated with metastasis. Science. 236:582–585. [PubMed] [Google Scholar]
600. Pierce M, Buckhaults P, Chen L, Fregien N. 1997. Regulation of N-acetylglucosaminyltransferase V and Asn-linked oligosaccharide beta(1,6) branching by a growth factor signaling pathway and effects on cell adhesion and metastatic potential. Glycoconj J. 14:623–630. [PubMed] [Google Scholar]
601. Chen L, Zhang WJ, Fregien N, Pierce M. 1998. The her-2/neu oncogene stimulates the transcription of N-acetylglucosaminyltransferase V and expression of its cell surface oligosaccharide products. Oncogene. 17:2087–2093. [PubMed] [Google Scholar]
602. Granovsky M, Fata J, Pawling J, Muller WJ, Khokha R, Dennis JW. 2000. Suppression of tumor growth and metastasis in Mgat5-deficient mice. Nat Med. 6:306–312. [PubMed] [Google Scholar]
603. Guo HB, Lee I, Kamar M, Akiyama SK, Pierce M. 2002. Aberrant N-glycosylation of beta1 integrin causes reduced alpha5beta 1 integrin clustering and stimulates cell migration. Cancer Res. 62:6837–6845. [PubMed] [Google Scholar]
604. Lau KS, Dennis JW. 2008. N-Glycans in cancer progression. Glycobiology. 18:750–760. [PubMed] [Google Scholar]
605. Partridge EA, Le RC, Di GGM, Pawling J, Cheung P, Granovsky M, Nabi IR, Wrana JL, Dennis JW. 2004. Regulation of cytokine receptors by Golgi N-glycan processing and endocytosis. Science. 306:120–124. [PubMed] [Google Scholar]
606. Lau KS, Partridge EA, Grigorian A, Silvescu CI, Reinhold VN, Demetriou M, Dennis JW. 2007. Complex N-glycan number and degree of branching cooperate to regulate cell proliferation and differentiation. Cell. 129:123–134. [PubMed] [Google Scholar]
607. Zhou RW, Mkhikian H, Grigorian A, Hong A, Chen D, Arakelyan A, Demetriou M. 2014. N-glycosylation bidirectionally extends the boundaries of thymocyte positive selection by decoupling Lck from Ca2+ signaling. Nat Immunol. 15:1038–1045. [PubMed] [Google Scholar]
608. Dennis JW, Nabi IR, Demetriou M. 2009. a. Metabolism, cell surface organization, and disease. Cell. 139:1229–1241. [PMC free article] [PubMed] [Google Scholar]
609. Ohtsubo K, Takamatsu S, Minowa MT, Yoshida A, Takeuchi M, Marth JD. 2005. Dietary and genetic control of glucose transporter 2 glycosylation promotes insulin secretion in suppressing diabetes. Cell. 123:1307–1321. [PubMed] [Google Scholar]
610. Ohtsubo K, Chen MZ, Olefsky JM, Marth JD. 2011. Pathway to diabetes through attenuation of pancreatic beta cell glycosylation and glucose transport. Nat Med. 17:1067–1075. [PMC free article] [PubMed] [Google Scholar]
611. Song Y, Aglipay JA, Bernstein JD, Goswami S, Stanley P. 2010. The bisecting GlcNAc on N-glycans inhibits growth factor signaling and retards mammary tumor progression. Cancer Res. 70:3361–3371. [PMC free article] [PubMed] [Google Scholar]
612. Yoshimura M, Nishikawa A, Ihara Y, Taniguchi S, Taniguchi N. 1995. Suppression of lung metastasis of B16 mouse melanoma by N-acetylglucosaminyltransferase III gene transfection. Proc Natl Acad Sci USA. 92:8754–8758. [PMC free article] [PubMed] [Google Scholar]
613. Yoshimura M, Ihara Y, Matsuzawa Y, Taniguchi N. 1996. Aberrant glycosylation of E-cadherin enhances cell–cell binding to suppress metastasis. J Biol Chem. 271:13811–13815. [PubMed] [Google Scholar]
614. Bhattacharyya L, Khan MI, Fant J, Brewer CF. 1989. Formation of highly ordered cross-linked lattices between asparagine-linked oligosaccharides and lectins observed by electron microscopy. J Biol Chem. 264:11543–11545. [PubMed] [Google Scholar]
615. Brewer CF, Miceli MC, Baum LG. 2002. Clusters, bundles, arrays and lattices: Novel mechanisms for lectin-saccharide-mediated cellular interactions. Curr Opin Struct Biol. 12:616–623. [PubMed] [Google Scholar]
616. Dennis JW, Lau KS, Demetriou M, Nabi IR. 2009. b. Adaptive regulation at the cell surface by N-glycosylation. Traffic. 10:1569–1578. [PubMed] [Google Scholar]
617. Boscher C, Dennis JW, Nabi IR. 2011. Glycosylation, galectins and cellular signaling. Curr Opin Cell Biol. 23:383–392. [PubMed] [Google Scholar]
618. Cha SK, Ortega B, Kurosu H, Rosenblatt KP, Kuro-O M, Huang CL. 2008. Removal of sialic acid involving Klotho causes cell-surface retention of TRPV5 channel via binding to galectin-1. Proc Natl Acad Sci USA. 105:9805–9810. [PMC free article] [PubMed] [Google Scholar]
619. Zhuo Y, Bellis SL. 2011. Emerging role of alpha2,6-sialic acid as a negative regulator of galectin binding and function. J Biol Chem. 286:5935–5941. [PMC free article] [PubMed] [Google Scholar]
620. Mo D, Costa SA, Ihrke G, Youker RT, Pastor-Soler N, Hughey RP, Weisz OA. 2012. Sialylation of N-linked glycans mediates apical delivery of endolyn in MDCK cells via a galectin-9-dependent mechanism. Mol Biol Cell. 23:3636–3646. [PMC free article] [PubMed] [Google Scholar]
621. Murugaesu N, Iravani M, van Weverwijk A, Ivetic A, Johnson DA, Antonopoulos A, Fearns A, Jamal-Hanjani M, Sims D, Fenwick K, et al.. 2014. An in vivo functional screen identifies ST6GalNAc2 sialyltransferase as a breast cancer metastasis suppressor. Cancer Discov. 4:304–317. [PubMed] [Google Scholar]
622. Suzuki O, Abe M, Hashimoto Y. 2015. Sialylation by β-galactoside α-2,6-sialyltransferase and N-glycans regulate cell adhesion and invasion in human anaplastic large cell lymphoma. Int J Oncol. 46:973–980. [PMC free article] [PubMed] [Google Scholar]
623. Schauer R. 1985. Sialic acids and their role as biological masks. Trends Biochem Sci. 10:357–360. [Google Scholar]
624. Muchmore E, Varki A. 1987. Inactivation of influenza C esterase decreases infectivity without loss of binding; a probe for 9-O-acetylated sialic acids. Science. 236:1293–1295. [PubMed] [Google Scholar]
625. Uchimura K, Lemjabbar-Alaoui H, van Kuppevelt TH, Rosen SD. 2010. Use of a phage display antibody to measure the enzymatic activity of the Sulfs. Methods Enzymol. 480:51–64. [PubMed] [Google Scholar]
626. Morimoto-Tomita M, Uchimura K, Werb Z, Hemmerich S, Rosen SD. 2002. Cloning and characterization of two extracellular heparin-degrading endosulfatases in mice and humans. J Biol Chem. 277:49175–49185. [PMC free article] [PubMed] [Google Scholar]
627. Gorsi B, Stringer SE. 2007. Tinkering with heparan sulfate sulfation to steer development. Trends Cell Biol. 17:173–177. [PubMed] [Google Scholar]
628. Vives RR, Seffouh A, Lortat-Jacob H. 2014. Post-synthetic regulation of HS structure: The Yin and Yang of the Sulfs in cancer. Front Oncol. 3:331. [PMC free article] [PubMed] [Google Scholar]
629. Lukowsky WA, Painter RH. 1972. Studies on the role of sialic acid in the physical and biological properties of erythropoietin. Can J Biochem. 50:909–917. [PubMed] [Google Scholar]
630. Takeuchi M, Takasaki S, Miyazaki H, Kato T, Hoshi S, Kochibe N, Kobata A. 1988. Comparative study of the asparagine-linked sugar chains of human erythropoietins purified from urine and the culture medium of recombinant Chinese hamster ovary cells. J Biol Chem. 263:3657–3663. [PubMed] [Google Scholar]
631. Takeuchi M, Inoue N, Strickland TW, Kubota M, Wada M, Shimizu R, Hoshi S, Kozutsumi H, Takasaki S, Kobata A. 1989. Relationship between sugar chain structure and biological activity of recombinant human erythropoietin produced in Chinese hamster ovary cells. Proc Natl Acad Sci USA. 86:7819–7822. [PMC free article] [PubMed] [Google Scholar]
632. Fukuda M, Sasaki H, Fukuda MN. 1990. Structure and role of carbohydrate in human erythropoietin. Adv Exp Med Biol. 271:53–68. [PubMed] [Google Scholar]
633. Takeuchi M, Kobata A. 1991. Structures and functional roles of the sugar chains of human erythropoietins. Glycobiology. 1:337–346. [PubMed] [Google Scholar]
634. Yamaguchi K, Akai K, Kawanishi G, Ueda M, Masuda S, Sasaki R. 1991. Effects of site-directed removal of N-glycosylation sites in human erythropoietin on its production and biological properties. J Biol Chem. 266:20434–20439. [PubMed] [Google Scholar]
635. Misaizu T, Matsuki S, Strickland TW, Takeuchi M, Kobata A, Takasaki S. 1995. Role of antennary structure of N-linked sugar chains in renal handling of recombinant human erythropoietin. Blood. 86:4097–4104. [PubMed] [Google Scholar]
636. Yuen CT, Storring PL, Tiplady RJ, Izquierdo M, Wait R, Gee CK, Gerson P, Lloyd P, Cremata JA. 2003. Relationships between the N-glycan structures and biological activities of recombinant human erythropoietins produced using different culture conditions and purification procedures. Br J Haematol. 121:511–526. [PubMed] [Google Scholar]
637. Sinclair AM, Elliott S. 2005. Glycoengineering: The effect of glycosylation on the properties of therapeutic proteins. J Pharm Sci. 94:1626–1635. [PubMed] [Google Scholar]
638. Llop E, Gutierrez-Gallego R, Segura J, Mallorqui J, Pascual JA. 2008. Structural analysis of the glycosylation of gene-activated erythropoietin (epoetin delta, Dynepo). Anal Biochem. 383:243–254. [PubMed] [Google Scholar]
639. Kiss Z, Elliott S, Jedynasty K, Tesar V, Szegedi J. 2010. Discovery and basic pharmacology of erythropoiesis-stimulating agents (ESAs), including the hyperglycosylated ESA, darbepoetin alfa: An update of the rationale and clinical impact. Eur J Clin Pharmacol. 66:331–340. [PubMed] [Google Scholar]
640. Su D, Zhao H, Xia H. 2010. Glycosylation-modified erythropoietin with improved half-life and biological activity. Int J Hematol. 91:238–244. [PubMed] [Google Scholar]
641. Cebon J, Nicola N, Ward M, Gardner I, Dempsey P, Layton J, Dührsen U, Burgess AW, Nice E, Morstyn G. 1990. Granulocyte-macrophage colony stimulating factor from human lymphocytes. The effect of glycosylation on receptor binding and biological activity. J Biol Chem. 265:4483–4491. [PubMed] [Google Scholar]
642. Altmann SW, Prystowsky MB. 1992. Evaluation of human N-linked glycosylation sites in murine granulocyte-macrophage colony-stimulating factor. Arch Biochem Biophys. 293:349–355. [PubMed] [Google Scholar]
643. Niu LH, Heaney ML, Vera JC, Golde DW. 2000. High-affinity binding to the GM-CSF receptor requires intact N-glycosylation sites in the extracellular domain of the β subunit. Blood. 95:3357–3362. [PubMed] [Google Scholar]
644. Forno G, Bollati FM, Oggero M, Kratje R, Etcheverrigaray M, Conradt HS, Nimtz M. 2004. N- and O-linked carbohydrates and glycosylation site occupancy in recombinant human granulocyte-macrophage colony-stimulating factor secreted by a Chinese hamster ovary cell line. Eur J Biochem. 271:907–919. [PubMed] [Google Scholar]
645. Belalcazar V, Gutiérrez Gallego R, Llop E, Segura J, Pascual JA. 2006. Assessing the instability of the isoelectric focusing patterns of erythropoietin in urine. Electrophoresis. 27:4387–4395. [PubMed] [Google Scholar]
646. Desharnais P, Naud JF, Ayotte C. 2013. Desialylation improves the detection of recombinant erythropoietins in urine samples analyzed by SDS-PAGE. Drug Test Anal. 5:870–876. [PubMed] [Google Scholar]
647. Hart GW, Housley MP, Slawson C. 2007. Cycling of O-linked beta-N-acetylglucosamine on nucleocytoplasmic proteins. Nature. 446:1017–1022. [PubMed] [Google Scholar]
648. Hart GW, Slawson C, Ramirez-Correa G, Lagerlof O. 2011. Cross talk between O-GlcNAcylation and phosphorylation: Roles in signaling, transcription, and chronic disease. Annu Rev Biochem. 80:825–858. [PMC free article] [PubMed] [Google Scholar]
649. Hart GW. 2014. Three decades of research on O-GlcNAcylation—A major nutrient sensor that regulates signaling, transcription and cellular metabolism. Front Endocrinol (Lausanne). 5:183. [PMC free article] [PubMed] [Google Scholar]
650. Lewis BA, Hanover JA. 2014. O-GlcNAc and the epigenetic regulation of gene expression. J Biol Chem. 289:34440–34448. [PMC free article] [PubMed] [Google Scholar]
651. Bond MR, Hanover JA. 2015. A little sugar goes a long way: The cell biology of O-GlcNAc. J Cell Biol. 208:869–880. [PMC free article] [PubMed] [Google Scholar]
652. Olivier-Van Stichelen S, Hanover JA. 2015. You are what you eat: O-linked N-acetylglucosamine in disease, development and epigenetics. Curr Opin Clin Nutr Metab Care. 18:339–345. [PMC free article] [PubMed] [Google Scholar]
653. Hardivillé S, Hart GW. 2016. Nutrient regulation of gene expression by O-GlcNAcylation of chromatin. Curr Opin Chem Biol. 33:88–94. [PMC free article] [PubMed] [Google Scholar]
654. Levine ZG, Walker S. 2016. The biochemistry of O-GlcNAc transferase: Which functions make it essential in mammalian cells. Annu Rev Biochem. 85:631–657. [PubMed] [Google Scholar]
655. Teng-umnuay P, van der Wel H, West CM. 1999. Identification of a UDP-GlcNAc:Skp1-hydroxyproline GlcNAc-transferase in the cytoplasm of Dictyostelium. J Biol Chem. 274:29144–29144. [PubMed] [Google Scholar]
656. West CM, Wang ZA, van der Wel H. 2010. A cytoplasmic prolyl hydroxylation and glycosylation pathway modifies Skp1 and regulates O2-dependent development in Dictyostelium. Biochim Biophys Acta. 1800:160–171. [PMC free article] [PubMed] [Google Scholar]
657. Xu Y, Brown KM, Wang ZA, van der Wel H, Teygong C, Zhang D, Blader IJ, West CM. 2012. The Skp1 protein from Toxoplasma is modified by a cytoplasmic prolyl 4-hydroxylase associated with oxygen sensing in the social amoeba Dictyostelium. J Biol Chem. 287:25098–25110. [PMC free article] [PubMed] [Google Scholar]
658. Zhang D, van der Wel H, Johnson JM, West CM. 2012. Skp1 prolyl 4-hydroxylase of dictyostelium mediates glycosylation-independent and -dependent responses to O2 without affecting Skp1 stability. J Biol Chem. 287:2006–2016. [PMC free article] [PubMed] [Google Scholar]
659. Sheikh MO, Xu Y, van der Wel H, Walden P, Hartson SD, West CM. 2015. Glycosylation of Skp1 promotes formation of Skp1-Cullin-1-F-box protein complexes in dictyostelium. Mol Cell Proteomics. 14:66–80. [PMC free article] [PubMed] [Google Scholar]
660. Winkelhake JL, Kunicki TJ, Elcombe BM, Aster RH. 1980. Effects of pH treatments and deglycosylation of rabbit immunoglobulin G on the binding of C1q. J Biol Chem. 255:2822–2828. [PubMed] [Google Scholar]
661. Heyman B, Nose M, Weigle WO. 1985. Carbohydrate chains on IgG2b: A requirement for efficient feedback immunosuppression. J Immunol. 134:4018–4023. [PubMed] [Google Scholar]
662. Fukushima K, Takasaki S. 1993. Processing inhibition of N-linked sugar chains associated with induction of Fc receptor-mediated phagocytosis in the mouse monocytoid cells. Glycobiology. 3:15–22. [PubMed] [Google Scholar]
663. Newkirk MM, Rauch J. 1993. Binding of human monoclonal IgG rheumatoid factors to Fc is influenced by carbohydrate. J Rheumatol. 20:776–780. [PubMed] [Google Scholar]
664. Parekh RB, Dwek RA, Sutton BJ, Fernandes DL, Leung A, Stanworth D, Rademacher TW, Mizuochi T, Taniguchi T, Matsuta K, et al.. 1985. Association of rheumatoid arthritis and primary osteoarthritis with changes in the glycosylation pattern of total serum IgG. Nature. 316:452–457. [PubMed] [Google Scholar]
665. Rademacher TW, Williams P, Dwek RA. 1994. Agalactosyl glycoforms of IgG autoantibodies are pathogenic. Proc Natl Acad Sci USA. 91:6123–6127. [PMC free article] [PubMed] [Google Scholar]
666. Malhotra R, Wormald MR, Rudd PM, Fischer PB, Dwek RA, Sim RB. 1995. Glycosylation changes of IgG associated with rheumatoid arthritis can activate complement via the mannose-binding protein. Nature Med. 1:237–243. [PubMed] [Google Scholar]
667. Shields RL, Lai J, Keck R, O'Connell LY, Hong K, Meng YG, Weikert SH, Presta LG. 2002. Lack of fucose on human IgG1 N-linked oligosaccharide improves binding to human Fcgamma RIII and antibody-dependent cellular toxicity. J Biol Chem. 277:26733–26740. [PubMed] [Google Scholar]
668. Okazaki A, Shoji-Hosaka E, Nakamura K, Wakitani M, Uchida K, Kakita S, Tsumoto K, Kumagai I, Shitara K. 2004. Fucose depletion from human IgG1 oligosaccharide enhances binding enthalpy and association rate between IgG1 and FcgammaRIIIa. J Mol Biol. 336:1239–1249. [PubMed] [Google Scholar]
669. Mori K, Iida S, Yamane-Ohnuki N, Kanda Y, Kuni-Kamochi R, Nakano R, Imai-Nishiya H, Okazaki A, Shinkawa T, Natsume A, et al.. 2007. Non-fucosylated therapeutic antibodies: The next generation of therapeutic antibodies. Cytotechnology. 55:109–114. [PMC free article] [PubMed] [Google Scholar]
670. Peipp M, Lammerts van Bueren JJ, Schneider-Merck T, Bleeker WW, Dechant M, Beyer T, Repp R, van Berkel PH, Vink T, van de Winkel JG, et al.. 2008. Antibody fucosylation differentially impacts cytotoxicity mediated by NK and PMN effector cells. Blood. 112:2390–2399. [PubMed] [Google Scholar]
671. Kapur R, Kustiawan I, Vestrheim A, Koeleman CA, Visser R, Einarsdottir HK, Porcelijn L, Jackson D, Kumpel B, Deelder AM, et al.. 2014. A prominent lack of IgG1-Fc fucosylation of platelet alloantibodies in pregnancy. Blood. 123:471–480. [PMC free article] [PubMed] [Google Scholar]
672. Raju TS. 2008. Terminal sugars of Fc glycans influence antibody effector functions of IgGs. Curr Opin Immunol. 20:471–478. [PubMed] [Google Scholar]
673. Reusch D, Tejada ML. 2015. Fc glycans of therapeutic antibodies as critical quality attributes. Glycobiology. 25:1325–1334. [PMC free article] [PubMed] [Google Scholar]
674. Nimmerjahn F, Ravetch JV. 2007. The antiinflammatory activity of IgG: The intravenous IgG paradox. J Exp Med. 204:11–15. [PMC free article] [PubMed] [Google Scholar]
675. Nimmerjahn F, Anthony RM, Ravetch JV. 2007. Agalactosylated IgG antibodies depend on cellular Fc receptors for in vivo activity. Proc Natl Acad Sci USA. 104:8433–8437. [PMC free article] [PubMed] [Google Scholar]
676. Scallon BJ, Tam SH, McCarthy SG, Cai AN, Raju TS. 2007. Higher levels of sialylated Fc glycans in immunoglobulin G molecules can adversely impact functionality. Mol Immunol. 44:1524–1534. [PubMed] [Google Scholar]
677. Anthony RM, Wermeling F, Karlsson MC, Ravetch JV. 2008. a. Identification of a receptor required for the anti-inflammatory activity of IVIG. Proc Natl Acad Sci USA. 105:19571–19578. [PMC free article] [PubMed] [Google Scholar]
678. Anthony RM, Nimmerjahn F, Ashline DJ, Reinhold VN, Paulson JC, Ravetch JV. 2008. b. Recapitulation of IVIG anti-inflammatory activity with a recombinant IgG Fc. Science. 320:373–376. [PMC free article] [PubMed] [Google Scholar]
679. Bayry J, Bansal K, Kazatchkine MD, Kaveri SV. 2009. DC-SIGN and alpha2,6-sialylated IgG Fc interaction is dispensable for the anti-inflammatory activity of IVIg on human dendritic cells. Proc Natl Acad Sci USA. 106:E24; author reply E25. [PMC free article] [PubMed] [Google Scholar]
680. Anthony RM, Kobayashi T, Wermeling F, Ravetch JV. 2011. Intravenous gammaglobulin suppresses inflammation through a novel T(H)2 pathway. Nature. 475:110–113. [PMC free article] [PubMed] [Google Scholar]
681. Guhr T, Bloem J, Derksen NI, Wuhrer M, Koenderman AH, Aalberse RC, Rispens T. 2011. Enrichment of sialylated IgG by lectin fractionation does not enhance the efficacy of immunoglobulin G in a murine model of immune thrombocytopenia. PLoS One. 6:e21246. [PMC free article] [PubMed] [Google Scholar]
682. Barb AW, Meng L, Gao Z, Johnson RW, Moremen KW, Prestegard JH. 2012. NMR characterization of immunoglobulin g fc glycan motion on enzymatic sialylation. Biochemistry. 51:4618–4626. [PMC free article] [PubMed] [Google Scholar]
683. Gelfand EW. 2012. Intravenous immune globulin in autoimmune and inflammatory diseases. N Engl J Med. 367:2015–2025. [PubMed] [Google Scholar]
684. Jones MB, Nasirikenari M, Lugade AA, Thanavala Y, Lau JT. 2012. Anti-inflammatory IgG production requires functional P1 promoter in beta-galactoside alpha2,6-sialyltransferase 1 (ST6Gal-1) gene. J Biol Chem. 287:15365–15370. [PMC free article] [PubMed] [Google Scholar]
685. Kasermann F, Boerema DJ, Ruegsegger M, Hofmann A, Wymann S, Zuercher AW, Miescher S. 2012. Analysis and functional consequences of increased Fab-sialylation of intravenous immunoglobulin (IVIG) after lectin fractionation. PLoS One. 7:e37243. [PMC free article] [PubMed] [Google Scholar]
686. Schwab I, Seeling M, Biburger M, Aschermann S, Nitschke L, Nimmerjahn F. 2012. B cells and CD22 are dispensable for the immediate antiinflammatory activity of intravenous immunoglobulins in vivo. Eur J Immunol. 42:3302–3309. [PubMed] [Google Scholar]
687. Ogata S, Shimizu C, Franco A, Touma R, Kanegaye JT, Choudhury BP, Naidu NN, Kanda Y, Hoang LT, Hibberd ML, et al.. 2013. Treatment response in kawasaki disease is associated with sialylation levels of endogenous but not therapeutic intravenous immunoglobulin g. PLoS One. 8:e81448. [PMC free article] [PubMed] [Google Scholar]
688. Yu X, Vasiljevic S, Mitchell DA, Crispin M, Scanlan CN. 2013. 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. 425:1253–1258. [PubMed] [Google Scholar]
689. Ahmed AA, Giddens J, Pincetic A, Lomino JV, Ravetch JV, Wang LX, Bjorkman PJ. 2014. Structural characterization of anti-inflammatory immunoglobulin G Fc proteins. J Mol Biol. 426:3166–3179. [PMC free article] [PubMed] [Google Scholar]
690. Ballow M. 2014. Mechanisms of immune regulation by IVIG. Curr Opin Allergy Clin Immunol. 14:509–515. [PubMed] [Google Scholar]
691. Böhm S, Kao D, Nimmerjahn F. 2014. Sweet and sour: The role of glycosylation for the anti-inflammatory activity of immunoglobulin G. Curr Top Microbiol Immunol. 382:393–417. [PubMed] [Google Scholar]
692. Campbell IK, Miescher S, Branch DR, Mott PJ, Lazarus AH, Han D, Maraskovsky E, Zuercher AW, Neschadim A, Leontyev D, et al.. 2014. Therapeutic effect of IVIG on inflammatory arthritis in mice is dependent on the Fc portion and independent of sialylation or basophils. J Immunol. 192:5031–5038. [PMC free article] [PubMed] [Google Scholar]
693. Fokkink WJ, Selman MH, Wuhrer M, Jacobs BC. 2014. Immunoglobulin G Fc N-glycosylation in Guillain-Barré syndrome treated with intravenous immunoglobulin. Clin Exp Immunol. 178(Suppl 1):105–107. [PMC free article] [PubMed] [Google Scholar]
694. Nagelkerke SQ, Dekkers G, Kustiawan I, van de Bovenkamp FS, Geissler J, Plomp R, Wuhrer M, Vidarsson G, Rispens T, van den Berg TK, et al.. 2014. Inhibition of FcgammaR-mediated phagocytosis by IVIg is independent of IgG-Fc sialylation and FcgammaRIIb in human macrophages. Blood. 124:3709–3718. [PubMed] [Google Scholar]
695. Quast I, Lunemann JD. 2014. Fc glycan-modulated immunoglobulin G effector functions. J Clin Immunol. 34(Suppl 1): S51–S55. [PubMed] [Google Scholar]
696. Schwab I, Mihai S, Seeling M, Kasperkiewicz M, Ludwig RJ, Nimmerjahn F. 2014. a. Broad requirement for terminal sialic acid residues and FcgammaRIIB for the preventive and therapeutic activity of intravenous immunoglobulins in vivo. Eur J Immunol. 44:1444–1453. [PubMed] [Google Scholar]
697. Schwab I, Lux A, Nimmerjahn F. 2014. b. Reply to—IVIG pluripotency and the concept of Fc-sialylation: Challenges to the scientist. Nat Rev Immunol. 14:349. [PubMed] [Google Scholar]
698. Sudo M, Yamaguchi Y, Spath PJ, Matsumoto-Morita K, Ong BK, Shahrizaila N, Yuki N. 2014. Different IVIG glycoforms affect in vitro inhibition of anti-ganglioside antibody-mediated complement deposition. PLoS One. 9:e107772. [PMC free article] [PubMed] [Google Scholar]
699. von Gunten S, Shoenfeld Y, Blank M, Branch DR, Vassilev T, Kasermann F, Bayry J, Kaveri S, Simon HU. 2014. IVIG pluripotency and the concept of Fc-sialylation: Challenges to the scientist. Nat Rev Immunol. 14:349. [PubMed] [Google Scholar]
700. Lünemann JD, Nimmerjahn F, Dalakas MC. 2015. Intravenous immunoglobulin in neurology—mode of action and clinical efficacy. Nat Rev Neurol. 11:80–89. [PubMed] [Google Scholar]
701. Quast I, Keller CW, Maurer MA, Giddens JP, Tackenberg B, Wang LX, Münz C, Nimmerjahn F, Dalakas MC, Lünemann JD. 2015. Sialylation of IgG Fc domain impairs complement-dependent cytotoxicity. J Clin Invest. 125:4160–4170. [PMC free article] [PubMed] [Google Scholar]
702. Calabrese EJ. 2008. Hormesis and medicine. Br J Clin Pharmacol. 66:594–617. [PMC free article] [PubMed] [Google Scholar]
703. Calabrese EJ, Blain RB. 2011. The hormesis database: The occurrence of hormetic dose responses in the toxicological literature. Regul Toxicol Pharmacol. 61:73–81. [PubMed] [Google Scholar]
704. Pearce OM, Laubli H, Bui J, Varki A. 2014. a. Hormesis in cancer immunology: Does the quantity of an immune reactant matter. Oncoimmunology. 3:e29312. [PMC free article] [PubMed] [Google Scholar]
705. Pearce OM, Laubli H, Verhagen A, Secrest P, Zhang J, Varki NM, Crocker PR, Bui JD, Varki A. 2014. b. Inverse hormesis of cancer growth mediated by narrow ranges of tumor-directed antibodies. Proc Natl Acad Sci USA. 111:5998–6003. [PMC free article] [PubMed] [Google Scholar]
706. Chung AW, Crispin M, Pritchard L, Robinson H, Gorny MK, Yu X, Bailey-Kellogg C, Ackerman ME, Scanlan C, Zolla-Pazner S, et al.. 2014. Identification of antibody glycosylation structures that predict monoclonal antibody Fc-effector function. AIDS. 28:2523–2530. [PMC free article] [PubMed] [Google Scholar]
707. Barb AW, Prestegard JH. 2011. NMR analysis demonstrates immunoglobulin G N-glycans are accessible and dynamic. Nat Chem Biol. 7:147–153. [PMC free article] [PubMed] [Google Scholar]
708. Subedi GP, Hanson QM, Barb AW. 2014. Restricted motion of the conserved immunoglobulin G1 N-glycan is essential for efficient FcgammaRIIIa binding. Structure. 22:1478–1488. [PMC free article] [PubMed] [Google Scholar]
709. Sondermann P, Pincetic A, Maamary J, Lammens K, Ravetch JV. 2013. General mechanism for modulating immunoglobulin effector function. Proc Natl Acad Sci USA. 110:9868–9872. [PMC free article] [PubMed] [Google Scholar]
710. Shade KT, Platzer B, Washburn N, Mani V, Bartsch YC, Conroy M, Pagan JD, Bosques C, Mempel TR, Fiebiger E, et al.. 2015. A single glycan on IgE is indispensable for initiation of anaphylaxis. J Exp Med. 212:457–467. [PMC free article] [PubMed] [Google Scholar]
711. Hanover JA, Krause MW, Love DC. 2012. Bittersweet memories: Linking metabolism to epigenetics through O-GlcNAcylation. Nat Rev Mol Cell Biol. 13:312–321. [PubMed] [Google Scholar]
712. Ofek I, Sharon N. 1990. Adhesins as lectins: Specificity and role in infection. Curr Top Microbiol Immunol. 151:91–114. [PubMed] [Google Scholar]
713. Marshall BJ, Warren JR. 1984. Unidentified curved bacilli in the stomach of patients with gastritis and peptic ulceration. Lancet. 1:1311–1315. [PubMed] [Google Scholar]
714. Boren T, Falk P, Roth KA, Larson G, Normark S. 1993. Attachment of Helicobacter pylori to human gastric epithelium mediated by blood group antigens. Science. 262:1892–1895. [PubMed] [Google Scholar]
715. Ilver D, Arnqvist A, Ogren J, Frick IM, Kersulyte D, Incecik ET, Berg DE, Covacci A, Engstrand L, Boren T. 1998. Helicobacter pylori adhesin binding fucosylated histo-blood group antigens revealed by retagging. Science. 279:373–377. [PubMed] [Google Scholar]
716. Teneberg S, Leonardsson I, Karlsson H, Jovall PA, Angstrom J, Danielsson D, Naslund I, Ljungh A, Wadstrom T, Karlsson KA. 2002. Lactotetraosylceramide, a novel glycosphingolipid receptor for Helicobacter pylori, present in human gastric epithelium. J Biol Chem. 277:19709–19719. [PubMed] [Google Scholar]
717. Walz A, Odenbreit S, Mahdavi J, Boren T, Ruhl S. 2005. Identification and characterization of binding properties of Helicobacter pylori by glycoconjugate arrays. Glycobiology. 15:700–708. [PubMed] [Google Scholar]
718. Linden S, Mahdavi J, Semino-Mora C, Olsen C, Carlstedt I, Boren T, Dubois A. 2008. Role of ABO secretor status in mucosal innate immunity and H. pylori infection. PLoS Pathog. 4:e2. [PMC free article] [PubMed] [Google Scholar]
719. Kobayashi M, Lee H, Nakayama J, Fukuda M. 2009. Roles of gastric mucin-type O-glycans in the pathogenesis of Helicobacter pylori infection. Glycobiology. 19:453–461. [PMC free article] [PubMed] [Google Scholar]
720. Kline KA, Falker S, Dahlberg S, Normark S, Henriques-Normark B. 2009. Bacterial adhesins in host–microbe interactions. Cell Host Microbe. 5:580–592. [PubMed] [Google Scholar]
721. Lillington J, Geibel S, Waksman G. 2014. Biogenesis and adhesion of type 1 and P pili. Biochim Biophys Acta. 1840:2783–2793. [PubMed] [Google Scholar]
722. Jiang X, Abgottspon D, Kleeb S, Rabbani S, Scharenberg M, Wittwer M, Haug M, Schwardt O, Ernst B. 2012. Antiadhesion therapy for urinary tract infections—A balanced PK/PD profile proved to be key for success. J Med Chem. 55:4700–4713. [PubMed] [Google Scholar]
723. Orlandi PA, Klotz FW, Haynes JD. 1992. A malaria invasion receptor, the 175-kilodalton erythrocyte binding antigen of Plasmodium falciparum recognizes the terminal Neu5Ac(alpha 2-3)Gal-sequences of glycophorin A. J Cell Biol. 116:901–909. [PMC free article] [PubMed] [Google Scholar]
724. Sim BK, Chitnis CE, Wasniowska K, Hadley TJ, Miller LH. 1994. Receptor and ligand domains for invasion of erythrocytes by Plasmodium falciparum. Science. 264:1941–1944. [PubMed] [Google Scholar]
725. Duraisingh MT, Maier AG, Triglia T, Cowman AF. 2003. Erythrocyte-binding antigen 175 mediates invasion in Plasmodium falciparum utilizing sialic acid-dependent and -independent pathways. Proc Natl Acad Sci USA. 100:4796–4801. [PMC free article] [PubMed] [Google Scholar]
726. Chattopadhyay D, Rayner J, McHenry AM, Adams JH. 2006. The structure of the Plasmodium falciparum EBA175 ligand domain and the molecular basis of host specificity. Trends Parasitol. 22:143–145. [PMC free article] [PubMed] [Google Scholar]
727. Persson KE, McCallum FJ, Reiling L, Lister NA, Stubbs J, Cowman AF, Marsh K, Beeson JG. 2008. Variation in use of erythrocyte invasion pathways by Plasmodium falciparum mediates evasion of human inhibitory antibodies. J Clin Invest. 118:342–351. [PMC free article] [PubMed] [Google Scholar]
728. Jiang L, Gaur D, Mu J, Zhou H, Long CA, Miller LH. 2011. Evidence for erythrocyte-binding antigen 175 as a component of a ligand-blocking blood-stage malaria vaccine. Proc Natl Acad Sci USA. 108:7553–7558. [PMC free article] [PubMed] [Google Scholar]
729. Ambroggio X, Jiang L, Aebig J, Obiakor H, Lukszo J, Narum DL. 2013. The epitope of monoclonal antibodies blocking erythrocyte invasion by Plasmodium falciparum map to the dimerization and receptor glycan binding sites of EBA-175. PLoS One. 8:e56326. [PMC free article] [PubMed] [Google Scholar]
730. Malpede BM, Lin DH, Tolia NH. 2013. Molecular basis for sialic acid-dependent receptor recognition by Plasmodium falciparum erythrocyte binding antigen 140/BAEBL. J Biol Chem. 288:12406–12415 [PMC free article] [PubMed] [Google Scholar]
731. Klotz FW, Orlandi PA, Reuter G, Cohen SJ, Haynes JD, Schauer R, Howard RJ, Palese P, Miller LH. 1992. Binding of Plasmodium falciparum 175-kilodalton erythrocyte binding antigen and invasion of murine erythrocytes requires N-acetylneuraminic acid but not its O-acetylated form. Mol Biochem Parasitol. 51:49–54. [PMC free article] [PubMed] [Google Scholar]
732. Martin MJ, Rayner JC, Gagneux P, Barnwell JW, Varki A. 2005. Evolution of human-chimpanzee differences in malaria susceptibility: Relationship to human genetic loss of N-glycolylneuraminic acid. Proc Natl Acad Sci USA. 102:12819–12824. [PMC free article] [PubMed] [Google Scholar]
733. Dankwa S, Lim C, Bei AK, Jiang RH, Abshire JR, Patel SD, Goldberg JM, Moreno Y, Kono M, Niles JC, et al.. 2016. Ancient human sialic acid variant restricts an emerging zoonotic malaria parasite. Nat Commun. 7:11187. [PMC free article] [PubMed] [Google Scholar]
734. Vogt AM, Barragan A, Chen Q, Kironde F, Spillmann D, Wahlgren M. 2003. Heparan sulfate on endothelial cells mediates the binding of Plasmodium falciparum-infected erythrocytes via the DBL1alpha domain of PfEMP1. Blood. 101:2405–2411. [PubMed] [Google Scholar]
735. Zupancic ML, Frieman M, Smith D, Alvarez RA, Cummings RD, Cormack BP. 2008. Glycan microarray analysis of Candida glabrata adhesin ligand specificity. Mol Microbiol. 68:547–559. [PubMed] [Google Scholar]
736. Spadafora C, Awandare GA, Kopydlowski KM, Czege J, Moch JK, Finberg RW, Tsokos GC, Stoute JA. 2010. Complement receptor 1 is a sialic acid-independent erythrocyte receptor of Plasmodium falciparum. PLoS Pathog. 6:e1000968. [PMC free article] [PubMed] [Google Scholar]
737. Tham WH, Wilson DW, Lopaticki S, Schmidt CQ, Tetteh-Quarcoo PB, Barlow PN, Richard D, Corbin JE, Beeson JG, Cowman AF. 2010. Complement receptor 1 is the host erythrocyte receptor for Plasmodium falciparum PfRh4 invasion ligand. Proc Natl Acad Sci USA. 107:17327–17332. [PMC free article] [PubMed] [Google Scholar]
738. Mensah-Brown HE, Amoako N, Abugri J, Stewart LB, Agongo G, Dickson EK, Ofori MF, Stoute JA, Conway DJ, Awandare GA. 2015. Analysis of erythrocyte invasion mechanisms of Plasmodium falciparum clinical isolates across 3 malaria-endemic areas in Ghana. J Infect Dis. 212:1288–1297. [PMC free article] [PubMed] [Google Scholar]
739. Paulson JC, Sadler JE, Hill RL. 1979. Restoration of specific myxovirus receptors to asialoerythrocytes by incorporation of sialic acid with pure sialyltransferases. J Biol Chem. 254:2120–2124. [PubMed] [Google Scholar]
740. Rogers GN, Paulson JC. 1983. Receptor determinants of human and animal influenza virus isolates: Differences in receptor specificity of the H3 hemagglutinin based on species of origin. Virology. 127:361–373. [PubMed] [Google Scholar]
741. Ito T, Suzuki Y, Mitnaul L, Vines A, Kida H, Kawaoka Y. 1997. Receptor specificity of influenza A viruses correlates with the agglutination of erythrocytes from different animal species. Virology. 227:493–499. [PubMed] [Google Scholar]
742. Ha Y, Stevens DJ, Skehel JJ, Wiley DC. 2001. X-ray structures of H5 avian and H9 swine influenza virus hemagglutinins bound to avian and human receptor analogs. Proc Natl Acad Sci USA. 98:11181–11186. [PMC free article] [PubMed] [Google Scholar]
743. Gambaryan A, Yamnikova S, Lvov D, Tuzikov A, Chinarev A, Pazynina G, Webster R, Matrosovich M, Bovin N. 2005. Receptor specificity of influenza viruses from birds and mammals: New data on involvement of the inner fragments of the carbohydrate chain. Virology. 334:276–283. [PubMed] [Google Scholar]
744. Glaser L, Stevens J, Zamarin D, Wilson IA, Garcia-Sastre A, Tumpey TM, Basler CF, Taubenberger JK, Palese P. 2005. A single amino acid substitution in 1918 influenza virus hemagglutinin changes receptor binding specificity. J Virol. 79:11533–11536. [PMC free article] [PubMed] [Google Scholar]
745. Stevens J, Blixt O, Tumpey TM, Taubenberger JK, Paulson JC, Wilson IA. 2006. Structure and receptor specificity of the hemagglutinin from an H5N1 influenza virus. Science. 312:404–410. [PubMed] [Google Scholar]
746. Belser JA, Blixt O, Chen LM, Pappas C, Maines TR, Van Hoeven N, Donis R, Busch J, McBride R, Paulson JC, et al.. 2008. Contemporary North American influenza H7 viruses possess human receptor specificity: Implications for virus transmissibility. Proc Natl Acad Sci USA. 105:7558–7563. [PMC free article] [PubMed] [Google Scholar]
747. Chandrasekaran A, Srinivasan A, Raman R, Viswanathan K, Raguram S, Tumpey TM, Sasisekharan V, Sasisekharan R. 2008. Glycan topology determines human adaptation of avian H5N1 virus hemagglutinin. Nat Biotechnol. 26:107–113. [PubMed] [Google Scholar]
748. Walther T, Karamanska R, Chan RW, Chan MC, Jia N, Air G, Hopton C, Wong MP, Dell A, Malik Peiris JS, et al.. 2013. Glycomic analysis of human respiratory tract tissues and correlation with influenza virus infection. PLoS Pathog. 9:e1003223. [PMC free article] [PubMed] [Google Scholar]
749. Xu R, de Vries RP, Zhu X, Nycholat CM, McBride R, Yu W, Paulson JC, Wilson IA. 2013. Preferential recognition of avian-like receptors in human influenza A H7N9 viruses. Science. 342:1230–1235. [PMC free article] [PubMed] [Google Scholar]
750. Gagneux P, Cheriyan M, Hurtado-Ziola N, Van Der Linden EC, Anderson D, McClure H, Varki A, Varki NM. 2003. Human-specific regulation of Alpha2-6 linked sialic acids. J Biol Chem. 278:48245–48250. [PubMed] [Google Scholar]
751. Baum LG, Paulson JC. 1990. Sialyloligosaccharides of the respiratory epithelium in the selection of human influenza virus receptor specificity. Acta Histochem (Jena). 89(Suppl. 40):35–38. [PubMed] [Google Scholar]
752. Leigh MW, Connor RJ, Kelm S, Baum LG, Paulson JC. 1995. Receptor specificity of influenza virus influences severity of illness in ferrets. Vaccine. 13:1468–1473. [PubMed] [Google Scholar]
753. Ng PS, Bohm R, Hartley-Tassell LE, Steen JA, Wang H, Lukowski SW, Hawthorne PL, Trezise AE, Coloe PJ, Grimmond SM, et al.. 2014. Ferrets exclusively synthesize Neu5Ac and express naturally humanized influenza A virus receptors. Nat Commun. 5:5750. [PMC free article] [PubMed] [Google Scholar]
754. Regl G, Kaser A, Iwersen M, Schmid H, Kohla G, Strobl B, Vilas U, Schauer R, Vlasak R. 1999. The hemagglutinin-esterase of mouse hepatitis virus strain S is a sialate-4-O-acetylesterase. J Virol. 73:4721–4727. [PMC free article] [PubMed] [Google Scholar]
755. Wurzer WJ, Obojes K, Vlasak R. 2002. The sialate-4-O-acetylesterases of coronaviruses related to mouse hepatitis virus: a proposal to reorganize group 2 Coronaviridae. J Gen Virol. 83:395–402. [PubMed] [Google Scholar]
756. Langereis MA, van Vliet AL, Boot W, de Groot RJ. 2010. Attachment of mouse hepatitis virus to O-acetylated sialic acid is mediated by hemagglutinin-esterase and not by the spike protein. J Virol. 84:8970–8974. [PMC free article] [PubMed] [Google Scholar]
757. Schultze B, Gross H-J, Klenk H-D, Brossmer R, Herrler G. 1990. Differential reactivity of bovine coronavirus (BCV) and influenza C virus with N-acetyl-9-O-acetylneuraminic acid (NEU5,9AC2)-containing receptors. Adv Exp Med Biol. 276:115–119. [PubMed] [Google Scholar]
758. Schultze B, Herrler G. 1992. Bovine coronavirus uses N-acetyl-9-O-acetylneuraminic acid as a receptor determinant to initiate the infection of cultured cells. J Gen Virol. 73:901–906. [PubMed] [Google Scholar]
759. Schwegmann-Wessels C, Herrler G. 2006. Sialic acids as receptor determinants for coronaviruses. Glycoconj J. 23:51–58. [PMC free article] [PubMed] [Google Scholar]
760. Langereis MA, Bakkers MJ, Deng L, Padler-Karavani V, Vervoort SJ, Hulswit RJ, van Vliet AL, Gerwig GJ, de Poot SA, Boot W, et al.. 2015. Complexity and diversity of the mammalian sialome revealed by nidovirus virolectins. Cell Rep. 11:1966–1978. [PMC free article] [PubMed] [Google Scholar]
761. Song H, Qi J, Khedri Z, Diaz S, Yu H, Chen X, Varki A, Shi Y, Gao GF. 2016. An open receptor-binding cavity of hemagglutinin-esterase-fusion glycoprotein from newly-identified influenza D virus: Basis for its broad cell tropism. PLoS Pathog. 12:e1005411. [PMC free article] [PubMed] [Google Scholar]
762. Langereis MA, Zeng Q, Heesters BA, Huizinga EG, de Groot RJ. 2012. The murine coronavirus hemagglutinin-esterase receptor-binding site: A major shift in ligand specificity through modest changes in architecture. PLoS Pathog. 8:e1002492. [PMC free article] [PubMed] [Google Scholar]
763. Bakkers MJ, Zeng Q, Feitsma LJ, Hulswit RJ, Li Z, Westerbeke A, van Kuppeveld FJ, Boons GJ, Langereis MA, Huizinga EG, et al.. 2016. Coronavirus receptor switch explained from the stereochemistry of protein-carbohydrate interactions and a single mutation. Proc Natl Acad Sci USA. 113:E3111–E3119. [PMC free article] [PubMed] [Google Scholar]
764. Holmgren J, Lonnroth I, Mansson J, Svennerholm L. 1975. Interaction of cholera toxin and membrane GM1 ganglioside of small intestine. Proc Natl Acad Sci USA. 72:2520–2524. [PMC free article] [PubMed] [Google Scholar]
765. Baenziger JU, Fiete D. 1979. Structural determinants of Ricinus communis agglutinin and toxin specificity for oligosaccharides. J Biol Chem. 254:9795–9799. [PubMed] [Google Scholar]
766. Lambert JM, McIntyre G, Gauthier MN, Zullo D, Rao V, Steeves RM, Goldmacher VS, Blättler WA. 1991. The galactose-binding sites of the cytotoxic lectin ricin can be chemically blocked in high yield with reactive ligands prepared by chemical modification of glycopeptides containing triantennary N-linked oligosaccharides. Biochemistry. 30:3234–3247. [PubMed] [Google Scholar]
767. Karmali MA, Petric M, Lim C, Fleming PC, Arbus GS, Lior H. 1985. The association between idiopathic hemolytic uremic syndrome and infection by verotoxin-producing Escherichia coli. J Infect Dis. 151:775–782. [PubMed] [Google Scholar]
768. Kiarash A, Boyd B, Lingwood CA. 1994. Glycosphingolipid receptor function is modified by fatty acid content. Verotoxin 1 and verotoxin 2c preferentially recognize different globotriaosyl ceramide fatty acid homologues. J Biol Chem. 269:11138–11146. [PubMed] [Google Scholar]
769. Steil D, Schepers CL, Pohlentz G, Legros N, Runde J, Humpf HU, Karch H, Müthing J. 2015. Shiga toxin glycosphingolipid receptors of Vero-B4 kidney epithelial cells and their membrane microdomain lipid environment. J Lipid Res. 56:2322–2336. [PMC free article] [PubMed] [Google Scholar]
770. Kitov PI, Sadowska JM, Mulvey G, Armstrong GD, Ling H, Pannu NS, Read RJ, Bundle DR. 2000. Shiga-like toxins are neutralized by tailored multivalent carbohydrate ligands. Nature. 403:669–672. [PubMed] [Google Scholar]
771. Kitova EN, Kitov PI, Bundle DR, Klassen JS. 2001. The observation of multivalent complexes of Shiga-like toxin with globotriaoside and the determination of their stoichiometry by nanoelectrospray Fourier-transform ion cyclotron resonance mass spectrometry. Glycobiology. 11:605–611. [PubMed] [Google Scholar]
772. Deng L, Song J, Gao X, Wang J, Yu H, Chen X, Varki N, Naito-Matsui Y, Galan JE, Varki A. 2014. Host adaptation of a bacterial toxin from the human pathogen salmonella typhi. Cell. 159:1290–1299. [PMC free article] [PubMed] [Google Scholar]
773. Byres E, Paton AW, Paton JC, Lofling JC, Smith DF, Wilce MC, Talbot UM, Chong DC, Yu H, Huang S, et al.. 2008. Incorporation of a non-human glycan mediates human susceptibility to a bacterial toxin. Nature. 456:648–652. [PMC free article] [PubMed] [Google Scholar]
774. Wands AM, Fujita A, McCombs JE, Cervin J, Dedic B, Rodriguez AC, Nischan N, Bond MR, Mettlen M, Trudgian DC, et al.. 2015. Fucosylation and protein glycosylation create functional receptors for cholera toxin. Elife. 4:e09545. [PMC free article] [PubMed] [Google Scholar]
775. Vasile F, Reina JJ, Potenza D, Heggelund JE, Mackenzie A, Krengel U, Bernardi A. 2014. Comprehensive analysis of blood group antigen binding to classical and El Tor cholera toxin B-pentamers by NMR. Glycobiology. 24:766–778. [PubMed] [Google Scholar]
776. Vaishnava S, Yamamoto M, Severson KM, Ruhn KA, Yu X, Koren O, Ley R, Wakeland EK, Hooper LV. 2011. The antibacterial lectin RegIIIgamma promotes the spatial segregation of microbiota and host in the intestine. Science. 334:255–258. [PMC free article] [PubMed] [Google Scholar]
777. Stowell SR, Arthur CM, Dias-Baruffi M, Rodrigues LC, Gourdine JP, Heimburg-Molinaro J, Ju T, Molinaro RJ, Rivera-Marrero C, Xia B, et al.. 2010. Innate immune lectins kill bacteria expressing blood group antigen. Nat Med. 16:295–301. [PMC free article] [PubMed] [Google Scholar]
778. Arthur CM, Cummings RD, Stowell SR. 2015. Evaluation of the bactericidal activity of galectins. Methods Mol Biol. 1207:421–430. [PMC free article] [PubMed] [Google Scholar]
779. Park AM, Hagiwara S, Hsu DK, Liu FT, Yoshie O. 2016. Galectin-3 plays an important role in innate immunity to gastric infection by Helicobacter pylori. Infect Immun. 84:1184–1193. [PMC free article] [PubMed] [Google Scholar]
780. Garner OB, Yun T, Pernet O, Aguilar HC, Park A, Bowden TA, Freiberg AN, Lee B, Baum LG. 2015. Timing of galectin-1 exposure differentially modulates Nipah virus entry and syncytium formation in endothelial cells. J Virol. 89:2520–2529. [PMC free article] [PubMed] [Google Scholar]
781. Degn SE, Thiel S. 2013. Humoral pattern recognition and the complement system. Scand J Immunol. 78:181–193. [PubMed] [Google Scholar]
782. Kawabata S, Iwanaga S. 1999. Role of lectins in the innate immunity of horseshoe crab. Dev Comp Immunol. 23:391–400. [PubMed] [Google Scholar]
783. Sonnenburg JL, Xu J, Leip DD, Chen CH, Westover BP, Weatherford J, Buhler JD, Gordon JI. 2005. Glycan foraging in vivo by an intestine-adapted bacterial symbiont. Science. 307:1955–1959. [PubMed] [Google Scholar]
784. Koropatkin NM, Cameron EA, Martens EC. 2012. How glycan metabolism shapes the human gut microbiota. Nat Rev Microbiol. 10:323–335. [PMC free article] [PubMed] [Google Scholar]
785. Marcobal A, Southwick AM, Earle KA, Sonnenburg JL. 2013. A refined palate: Bacterial consumption of host glycans in the gut. Glycobiology. 23:1038–1046. [PMC free article] [PubMed] [Google Scholar]
786. White BA, Lamed R, Bayer EA, Flint HJ. 2014. Biomass utilization by gut microbiomes. Annu Rev Microbiol. 68:279–296. [PubMed] [Google Scholar]
787. Matrosovich M, Herrler G, Klenk HD. 2015. Sialic acid receptors of viruses. Top Curr Chem. 367:1–28. [PMC free article] [PubMed] [Google Scholar]
788. Matrosovich MN, Matrosovich TY, Gray T, Roberts NA, Klenk HD. 2004. Neuraminidase is important for the initiation of influenza virus infection in human airway epithelium. J Virol. 78:12665–12667. [PMC free article] [PubMed] [Google Scholar]
789. Cohen M, Zhang XQ, Senaati HP, Chen HW, Varki NM, Schooley RT, Gagneux P. 2013. Influenza A penetrates host mucus by cleaving sialic acids with neuraminidase. Virol J. 10:321. [PMC free article] [PubMed] [Google Scholar]
790. Zhang L, Bukreyev A, Thompson CI, Watson B, Peeples ME, Collins PL, Pickles RJ. 2005. Infection of ciliated cells by human parainfluenza virus type 3 in an in vitro model of human airway epithelium. J Virol. 79:1113–1124. [PMC free article] [PubMed] [Google Scholar]
791. Meindl P, Bodo G, Palese P, Schulman J, Tuppy H. 1974. Inhibition of neuraminidase activity by derivatives of 2-deoxy-2,3-dehydro-N-acetylneuraminic acid. Virology. 58:457–463. [PubMed] [Google Scholar]
792. von Itzstein M, Wu WY, Kok GB, Pegg MS, Dyason JC, Jin B, Van Phan T, Smythe ML, White HF, Oliver SW, et al.. 1993. Rational design of potent sialidase-based inhibitors of influenza virus replication. Nature. 363:418–423. [PubMed] [Google Scholar]
793. Hayden FG, Osterhaus ADME, Treanor JJ, Fleming DM, Aoki FY, Nicholson KG, Bohnen AM, Hirst HM, Keene O, Wightman K. 1997. Efficacy and safety of the neuraminidase inhibitor zanamivir in the treatment of influenza virus infections. N Engl J Med. 337:874–880. [PubMed] [Google Scholar]
794. Hayden FG, Treanor JJ, Fritz RS, Lobo M, Betts RF, Miller M, Kinnersley N, Mills RG, Ward P, Straus SE. 1999. Use of the oral neuraminidase inhibitor oseltamivir in experimental human influenza—Randomized controlled trials for prevention and treatment. JAMA. 282:1240–1246. [PubMed] [Google Scholar]
795. Hayden FG, Atmar RL, Schilling M, Johnson C, Poretz D, Paar D, Huson L, Ward P, Mills RG, Oseltamivir SG. 1999. Use of the selective oral neuraminidase inhibitor oseltamivir to prevent influenza. N Engl J Med. 341:1336–1343. [PubMed] [Google Scholar]
796. Moustafa I, Connaris H, Taylor M, Zaitsev V, Wilson JC, Kiefel MJ, von Itzstein M, Taylor G. 2004. Sialic acid recognition by Vibrio cholerae neuraminidase. J Biol Chem. 279:40819–40826. [PubMed] [Google Scholar]
797. Takahashi Y, Sandberg AL, Ruhl S, Muller J, Cisar JO. 1997. A specific cell surface antigen of Streptococcus gordonii is associated with bacterial hemagglutination and adhesion to alpha2-3-linked sialic acid-containing receptors. Infect Immun. 65:5042–5051. [PMC free article] [PubMed] [Google Scholar]
798. Plummer C, Wu H, Kerrigan SW, Meade G, Cox D, Ian Douglas CW. 2005. A serine-rich glycoprotein of Streptococcus sanguis mediates adhesion to platelets via GPIb. Br J Haematol. 129:101–109. [PubMed] [Google Scholar]
799. Takamatsu D, Bensing BA, Prakobphol A, Fisher SJ, Sullam PM. 2006. Binding of the streptococcal surface glycoproteins GspB and Hsa to human salivary proteins. Infect Immun. 74:1933–1940. [PMC free article] [PubMed] [Google Scholar]
800. Bensing BA, Khedri Z, Deng L, Yu H, Prakobphol A, Fisher SJ, Chen X, Iverson TM, Varki A, Sullam PM. 2016. Novel aspects of sialoglycan recognition by the Siglec-like domains of streptococcal SRR glycoproteins. Glycobiology. 10.1093/glycob/cww042. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
801. Zhou P, Liu J, Li X, Takahashi Y, Qi F. 2015. The sialic acid binding protein, Hsa, in Streptococcus gordonii DL1 also mediates intergeneric coaggregation with Veillonella species. PLoS One. 10:e0143898. [PMC free article] [PubMed] [Google Scholar]
802. Sullam PM, Valone FH, Mills J. 1987. Mechanisms of platelet aggregation by viridans group streptococci. Infect Immun. 55:1743–1750. [PMC free article] [PubMed] [Google Scholar]
803. Yajima A, Takahashi Y, Konishi K. 2005. Identification of platelet receptors for the Streptococcus gordonii DL1 sialic acid-binding adhesin. Microbiol Immunol. 49:795–800. [PubMed] [Google Scholar]
804. Plummer C, Douglas CW. 2006. Relationship between the ability of oral streptococci to interact with platelet glycoprotein Ibalpha and with the salivary low-molecular-weight mucin, MG2. FEMS Immunol Med Microbiol. 48:390–399. [PubMed] [Google Scholar]
805. Xiong YQ, Bensing BA, Bayer AS, Chambers HF, Sullam PM. 2008. Role of the serine-rich surface glycoprotein GspB of Streptococcus gordonii in the pathogenesis of infective endocarditis. Microb Pathog. 45:297–301. [PMC free article] [PubMed] [Google Scholar]
806. Pyburn TM, Bensing BA, Xiong YQ, Melancon BJ, Tomasiak TM, Ward NJ, Yankovskaya V, Oliver KM, Cecchini G, Sulikowski GA, et al.. 2011. A structural model for binding of the serine-rich repeat adhesin GspB to host carbohydrate receptors. PLoS Pathog. 7:e1002112. [PMC free article] [PubMed] [Google Scholar]
807. Deng L, Bensing BA, Thamadilok S, Yu H, Lau K, Chen X, Ruhl S, Sullam PM, Varki A. 2014. Oral streptococci utilize a siglec-like domain of serine-rich repeat adhesins to preferentially target platelet sialoglycans in human blood. PLoS Pathog. 10:e1004540. [PMC free article] [PubMed] [Google Scholar]
808. Hutson AM, Atmar RL, Graham DY, Estes MK. 2002. Norwalk virus infection and disease is associated with ABO histo-blood group type. J Infect Dis. 185:1335–1337. [PubMed] [Google Scholar]
809. Huang P, Farkas T, Zhong W, Tan M, Thornton S, Morrow AL, Jiang X. 2005. Norovirus and histo-blood group antigens: Demonstration of a wide spectrum of strain specificities and classification of two major binding groups among multiple binding patterns. J Virol. 79:6714–6722. [PMC free article] [PubMed] [Google Scholar]
810. Choi JM, Hutson AM, Estes MK, Prasad BV. 2008. Atomic resolution structural characterization of recognition of histo-blood group antigens by Norwalk virus. Proc Natl Acad Sci USA. 105:9175–9180. [PMC free article] [PubMed] [Google Scholar]
811. Le Pendu J, Nyström K, Ruvoën-Clouet N. 2014. Host-pathogen co-evolution and glycan interactions. Curr Opin Virol. 7:88–94. [PubMed] [Google Scholar]
812. Bode L. 2012. Human milk oligosaccharides: Every baby needs a sugar mama. Glycobiology. 22:1147–1162. [PMC free article] [PubMed] [Google Scholar]
813. Etzold S, Bode L. 2014. Glycan-dependent viral infection in infants and the role of human milk oligosaccharides. Curr Opin Virol. 7C:101–107. [PubMed] [Google Scholar]
814. El-Hawiet A, Kitova EN, Klassen JS. 2015. Recognition of human milk oligosaccharides by bacterial exotoxins. Glycobiology. 25:845–854. [PubMed] [Google Scholar]
815. Medzhitov R, Janeway CAJ. 1997. Innate immunity: The virtues of a nonclonal system of recognition. Cell. 91:295–298. [PubMed] [Google Scholar]
816. Janeway CAJ, Medzhitov R. 2002. Innate immune recognition. Annu Rev Immunol. 20:197–216. [PubMed] [Google Scholar]
817. Beutler BA. 2009. TLRs and innate immunity. Blood. 113:1399–1407. [PMC free article] [PubMed] [Google Scholar]
818. Moresco EM, LaVine D, Beutler B. 2011. Toll-like receptors. Curr Biol. 21:R488–R493. [PubMed] [Google Scholar]
819. Pétrilli V, Dostert C, Muruve DA, Tschopp J. 2007. The inflammasome: A danger sensing complex triggering innate immunity. Curr Opin Immunol. 19:615–622. [PubMed] [Google Scholar]
820. Martinon F, Mayor A, Tschopp J. 2009. The inflammasomes: Guardians of the body. Annu Rev Immunol. 27:229–265. [PubMed] [Google Scholar]
821. Davis BK, Wen H, Ting JP. 2011. The inflammasome NLRs in immunity, inflammation, and associated diseases. Annu Rev Immunol. 29:707–735. [PMC free article] [PubMed] [Google Scholar]
822. Caruso R, Warner N, Inohara N, Núñez G. 2014. NOD1 and NOD2: Signaling, host defense, and inflammatory disease. Immunity. 41:898–908. [PMC free article] [PubMed] [Google Scholar]
823. Kogelberg H, Feizi T. 2001. New structural insights into lectin-type proteins of the immune system. Curr Opin Struct Biol. 11:635–643. [PubMed] [Google Scholar]
824. Marshall AS, Gordon S. 2004. Commentary: C-type lectins on the macrophage cell surface—recent findings. Eur J Immunol. 34:18–24. [PubMed] [Google Scholar]
825. Takeuchi O, Akira S. 2010. Pattern recognition receptors and inflammation. Cell. 140:805–820. [PubMed] [Google Scholar]
826. Rabinovich GA, Croci DO. 2012. Regulatory circuits mediated by lectin–glycan interactions in autoimmunity and cancer. Immunity. 36:322–335. [PubMed] [Google Scholar]
827. Fujimoto Y, Tanaka K, Shimoyama A, Fukase K. 2010. Self and nonself recognition with bacterial and animal glycans, surveys by synthetic chemistry. Methods Enzymol. 478:323–342. [PubMed] [Google Scholar]
828. Mahla RS, Reddy MC, Prasad DV, Kumar H. 2013. Sweeten PAMPs: Role of sugar complexed PAMPs in innate immunity and vaccine biology. Front Immunol. 4:248. [PMC free article] [PubMed] [Google Scholar]
829. Silipo A, Erbs G, Shinya T, Dow JM, Parrilli M, Lanzetta R, Shibuya N, Newman MA, Molinaro A. 2010. Glyco-conjugates as elicitors or suppressors of plant innate immunity. Glycobiology. 20:406–419. [PubMed] [Google Scholar]
830. Tundup S, Srivastava L, Harn DA. 2012. Polarization of host immune responses by helminth-expressed glycans. Ann NY Acad Sci. 1253:E1–E13. [PubMed] [Google Scholar]
831. van Die I, Cummings RD. 2010. Glycan gimmickry by parasitic helminths: A strategy for modulating the host immune response. Glycobiology. 20:2–12. [PubMed] [Google Scholar]
832. Geijtenbeek TB, Van Vliet SJ, Engering A, ‘T Hart BA, Van Kooyk Y. 2004. Self- and nonself-recognition by C-type lectins on dendritic cells. Annu Rev Immunol. 22:33–54. [PubMed] [Google Scholar]
833. McGreal EP, Miller JL, Gordon S. 2005. Ligand recognition by antigen-presenting cell C-type lectin receptors. Curr Opin Immunol. 17:18–24. [PMC free article] [PubMed] [Google Scholar]
834. van Vliet SJ, van Liempt E, Saeland E, Aarnoudse CA, Appelmelk B, Irimura T, Geijtenbeek TB, Blixt O, Alvarez R, van Die I, et al.. 2005. Carbohydrate profiling reveals a distinctive role for the C-type lectin MGL in the recognition of helminth parasites and tumor antigens by dendritic cells. Int Immunol. 17:661–669. [PubMed] [Google Scholar]
835. Chung CH, Mirakhur B, Chan E, Le QT, Berlin J, Morse M, Murphy BA, Satinover SM, Hosen J, Mauro D, et al.. 2008. Cetuximab-induced anaphylaxis and IgE specific for galactose-alpha-1,3-galactose. N Engl J Med. 358:1109–1117. [PMC free article] [PubMed] [Google Scholar]
836. Ghaderi D, Taylor RE, Padler-Karavani V, Diaz S, Varki A. 2010. Implications of the presence of N-glycolylneuraminic acid in recombinant therapeutic glycoproteins. Nat Biotechnol. 28:863–867. [PMC free article] [PubMed] [Google Scholar]
837. Kawano T, Cui J, Koezuka Y, Toura I, Kaneko Y, Motoki K, Ueno H, Nakagawa R, Sato H, Kondo E et al.. 1997. CD1d-restricted and TCR-mediated activation of valpha14 NKT cells by glycosylceramides. Science. 278:1626–1629. [PubMed] [Google Scholar]
838. Burdin N, Brossay L, Koezuka Y, Smiley ST, Grusby MJ, Gui M, Taniguchi M, Hayakawa K, Kronenberg M. 1998. Selective ability of mouse CD1 to present glycolipids: Alpha-galactosylceramide specifically stimulates V alpha 14+ NK T lymphocytes. J Immunol. 161:3271–3281. [PubMed] [Google Scholar]
839. Zhou D, Mattner J, Cantu C, Schrantz N, Yin N, Gao Y, Sagiv Y, Hudspeth K, Wu YP, Yamashita T, et al.. 2004. Lysosomal glycosphingolipid recognition by NKT cells. Science. 306:1786–1789. [PubMed] [Google Scholar]
840. Pei B, Vela JL, Zajonc D, Kronenberg M. 2012. Interplay between carbohydrate and lipid in recognition of glycolipid antigens by natural killer T cells. Ann NY Acad Sci. 1253:68–79. [PMC free article] [PubMed] [Google Scholar]
841. Dellabona P, Consonni M, de Lalla C, Casorati G. 2015. Group 1 CD1-restricted T cells and the pathophysiological implications of self-lipid antigen recognition. Tissue Antigens. 86:393–405. [PubMed] [Google Scholar]
842. Zajonc DM, Girardi E. 2015. Recognition of microbial glycolipids by natural killer T cells. Front Immunol. 6:400. [PMC free article] [PubMed] [Google Scholar]
843. Geno KA, Gilbert GL, Song JY, Skovsted IC, Klugman KP, Jones C, Konradsen HB, Nahm MH. 2015. Pneumococcal capsules and their types: Past, present, and future. Clin Microbiol Rev. 28:871–899. [PMC free article] [PubMed] [Google Scholar]
844. Croucher NJ, Harris SR, Fraser C, Quail MA, Burton J, van der Linden M, McGee L, von Gottberg A, Song JH, Ko KS, et al.. 2011. Rapid pneumococcal evolution in response to clinical interventions. Science. 331:430–434. [PMC free article] [PubMed] [Google Scholar]
845. López R, García E. 2004. Recent trends on the molecular biology of pneumococcal capsules, lytic enzymes, and bacteriophage. FEMS Microbiol Rev. 28:553–580. [PubMed] [Google Scholar]
846. Steinbacher S, Baxa U, Miller S, Weintraub A, Seckler R, Huber R. 1996. Crystal structure of phage P22 tailspike protein complexed with Salmonella sp. O-antigen receptors. Proc Natl Acad Sci USA. 93:10584–10588. [PMC free article] [PubMed] [Google Scholar]
847. Schwarzer D, Buettner FF, Browning C, Nazarov S, Rabsch W, Bethe A, Oberbeck A, Bowman VD, Stummeyer K, Mühlenhoff M, et al.. 2012. A multivalent adsorption apparatus explains the broad host range of phage phi92: A comprehensive genomic and structural analysis. J Virol. 86:10384–10398. [PMC free article] [PubMed] [Google Scholar]
848. Javed MA, van Alphen LB, Sacher J, Ding W, Kelly J, Nargang C, Smith DF, Cummings RD, Szymanski CM. 2015. A receptor-binding protein of Campylobacter jejuni bacteriophage NCTC 12673 recognizes flagellin glycosylated with acetamidino-modified pseudaminic acid. Mol Microbiol. 95:101–115. [PubMed] [Google Scholar]
849. Shen Y, Barros M, Vennemann T, Gallagher DT, Yin Y, Linden SB, Heselpoth RD, Spencer DJ, Donovan DM, Moult J, et al.. 2016. A bacteriophage endolysin that eliminates intracellular streptococci. Elife. 5:e13152. [PMC free article] [PubMed] [Google Scholar]
850. Breitbart M. 2012. Marine viruses: Truth or dare. Ann Rev Mar Sci. 4:425–448. [PubMed] [Google Scholar]
851. Simpson DJ, Sacher JC, Szymanski CM. 2015. Exploring the interactions between bacteriophage-encoded glycan binding proteins and carbohydrates. Curr Opin Struct Biol. 34:69–77. [PubMed] [Google Scholar]
852. Rodríguez-Rubio L, Martínez B, Donovan DM, Rodríguez A, García P. 2013. Bacteriophage virion-associated peptidoglycan hydrolases: Potential new enzybiotics. Crit Rev Microbiol. 39:427–434. [PubMed] [Google Scholar]
853. Sutherland IW, Hughes KA, Skillman LC, Tait K. 2004. The interaction of phage and biofilms. FEMS Microbiol Lett. 232:1–6. [PubMed] [Google Scholar]
854. Simpson DJ, Sacher JC, Szymanski CM. 2016. Development of an assay for the identification of receptor binding proteins from bacteriophages. Viruses. 8, 10.3390/v8010017. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
855. Li E, Tabas I, Kornfeld S. 1978. The synthesis of complex-type oligosaccharides. I. Structure of the lipid-linked oligosaccharide precursor of the complex-type oligosaccharides of the vesicular stomatitis virus G protein. J Biol Chem. 253:7762–7770. [PubMed] [Google Scholar]
856. Welply JK, Shenbagamurthi P, Naider F, Park HR, Lennarz WJ. 1985. Active site-directed photoaffinity labeling and partial characterization of oligosaccharyltransferase. J Biol Chem. 260:6459–6465. [PubMed] [Google Scholar]
857. Bosch M, Trombetta S, Engstrom U, Parodi AJ. 1988. Characterization of dolichol diphosphate oligosaccharide: Protein oligosaccharyltransferase and glycoprotein-processing glucosidases occurring in trypanosomatid protozoa. J Biol Chem. 263:17360–17365. [PubMed] [Google Scholar]
858. Imperiali B, Shannon KL. 1991. Differences between Asn-Xaa-Thr-containing peptides: A comparison of solution conformation and substrate behavior with oligosaccharyltransferase. Biochemistry. 30:4374–4380. [PubMed] [Google Scholar]
859. Silberstein S, Gilmore R. 1996. Biochemistry, molecular biology, and genetics of the oligosaccharyltransferase. FASEB J. 10:849–858. [PubMed] [Google Scholar]
860. Parodi AJ. 1993. N-glycosylation in trypanosomatid protozoa. Glycobiology. 3:193–199. [PubMed] [Google Scholar]
861. Samuelson J, Robbins PW. 2015. Effects of N-glycan precursor length diversity on quality control of protein folding and on protein glycosylation. Semin Cell Dev Biol. 41:121–128. [PMC free article] [PubMed] [Google Scholar]
862. Parodi AJ, Mendelzon DH, Lederkremer GZ. 1983. Transient glucosylation of protein-bound Man9GlcNAc2, Man8GlcNAc2, and Man7GlcNAc2 in calf thyroid cells. A possible recognition signal in the processing of glycoproteins. J Biol Chem. 258:8260–8265. [PubMed] [Google Scholar]
863. Parodi AJ, Mendelzon DH, Lederkremer GZ, Martin BJ. 1984. Evidence that transient glucosylation of protein-linked Man9GlcNAc2, Man8GlcNAc2, and Man7GlcNAc2 occurs in rat liver and Phaseolus vulgaris cells. J Biol Chem. 259:6351–6357. [PubMed] [Google Scholar]
864. Hammond C, Helenius A. 1994. Quality control in the secretory pathway: Retention of a misfolded viral membrane glycoprotein involves cycling between the ER, intermediate compartment, and Golgi apparatus. J Cell Biol. 126:41–52. [PMC free article] [PubMed] [Google Scholar]
865. Hammond C, Braakman I, Helenius A. 1994. Role of N-linked oligosaccharide recognition, glucose trimming, and calnexin in glycoprotein folding and quality control. Proc Natl Acad Sci USA. 91:913–917. [PMC free article] [PubMed] [Google Scholar]
866. Kearse, KP, Williams DB, Singer A. 1994. Persistence of glucose residues on core oligosaccharides prevents association of TCR alpha and TCR beta proteins with calnexin and results specifically in accelerated degradation of nascent TCR alpha proteins within the endoplasmic reticulum. EMBO J. 13:3678–3686. [PMC free article] [PubMed] [Google Scholar]
867. Hebert DN, Foellmer B, Helenius A. 1995. Glucose trimming and reglucosylation determine glycoprotein association with calnexin in the endoplasmic reticulum. Cell. 81:425–433. [PubMed] [Google Scholar]
868. Pelletier MF, Marcil A, Sevigny G, Jakob CA, Tessier DC, Chevet E, Menard R, Bergeron JJM, Thomas DY. 2000. The heterodimeric structure of glucosidase II is required for its activity, solubility, and localization in vivo. Glycobiology. 10:815–827. [PubMed] [Google Scholar]
869. Petrescu AJ, Butters TD, Reinkensmeier G, Petrescu S, Platt FM, Dwek RA, Wormald MR. 1997. The solution NMR structure of glucosylated N-glycans involved in the early stages of glycoprotein biosynthesis and folding. EMBO J. 16:4302–4310. [PMC free article] [PubMed] [Google Scholar]
870. Jakob CA, Burda P, Roth J, Aebi M. 1998. Degradation of misfolded endoplasmic reticulum glycoproteins in Saccharomyces cerevisiae is determined by a specific oligosaccharide structure. J Cell Biol. 142:1223–1233. [PMC free article] [PubMed] [Google Scholar]
871. Jakob CA, Bodmer D, Spirig U, Battig P, Marcil A, Dignard D, Bergeron JJ, Thomas DY, Aebi M. 2001. Htm1p, a mannosidase-like protein, is involved in glycoprotein degradation in yeast. EMBO Rep. 2:423–430. [PMC free article] [PubMed] [Google Scholar]
872. Deprez P, Gautschi M, Helenius A. 2005. More than one glycan is needed for ER glucosidase II to allow entry of glycoproteins into the calnexin/calreticulin cycle. Mol Cell. 19:183–195. [PubMed] [Google Scholar]
873. Molinari M, Galli C, Vanoni O, Arnold SM, Kaufman RJ. 2005. Persistent glycoprotein misfolding activates the glucosidase II/UGT1-driven calnexin cycle to delay aggregation and loss of folding competence. Mol Cell. 20:503–512. [PubMed] [Google Scholar]
874. Spear ED, Ng DT. 2005. Single, context-specific glycans can target misfolded glycoproteins for ER-associated degradation. J Cell Biol. 169:73–82. [PMC free article] [PubMed] [Google Scholar]
875. Gauss R, Jarosch E, Sommer T, Hirsch C. 2006. A complex of Yos9p and the HRD ligase integrates endoplasmic reticulum quality control into the degradation machinery. Nat Cell Biol. 8:849–854. [PubMed] [Google Scholar]
876. Quan EM, Kamiya Y, Kamiya D, Denic V, Weibezahn J, Kato K, Weissman JS. 2008. Defining the glycan destruction signal for endoplasmic reticulum-associated degradation. Mol Cell. 32:870–877. [PMC free article] [PubMed] [Google Scholar]
877. Schallus T, Jaeckh C, Feher K, Palma AS, Liu Y, Simpson JC, Mackeen M, Stier G, Gibson TJ, Feizi T, et al.. 2008. Malectin: A novel carbohydrate-binding protein of the endoplasmic reticulum and a candidate player in the early steps of protein N-glycosylation. Mol Biol Cell. 19:3404–3414. [PMC free article] [PubMed] [Google Scholar]
878. Clerc S, Hirsch C, Oggier DM, Deprez P, Jakob C, Sommer T, Aebi M. 2009. Htm1 protein generates the N-glycan signal for glycoprotein degradation in the endoplasmic reticulum. J Cell Biol. 184:159–172. [PMC free article] [PubMed] [Google Scholar]
879. Helenius A, Aebi M. 2004. Roles of N-linked glycans in the endoplasmic reticulum. Annu Rev Biochem. 73:1019–1049. [PubMed] [Google Scholar]
880. Spiro RG. 2004. Role of N-linked polymannose oligosaccharides in targeting glycoproteins for endoplasmic reticulum-associated degradation. Cell Mol Life Sci. 61:1025–1041. [PMC free article] [PubMed] [Google Scholar]
881. Chantret I, Moore SE. 2008. Free oligosaccharide regulation during mammalian protein N-glycosylation. Glycobiology. 18:210–224. [PubMed] [Google Scholar]
882. Määttänen P, Gehring K, Bergeron JJ, Thomas DY. 2010. Protein quality control in the ER: the recognition of misfolded proteins. Semin Cell Dev Biol. 21:500–511. [PubMed] [Google Scholar]
883. Caramelo JJ, Parodi AJ. 2015. A sweet code for glycoprotein folding. FEBS Lett. 589:3379–3387. [PubMed] [Google Scholar]
884. Satoh T, Yamaguchi T, Kato K. 2015. Emerging structural insights into glycoprotein quality control coupled with N-glycan processing in the endoplasmic reticulum. Molecules. 20:2475–2491. [PMC free article] [PubMed] [Google Scholar]
885. Christianson JC, Shaler TA, Tyler RE, Kopito RR, Satoh T, Chen Y, Hu D, Hanashima S, Yamamoto K, Yamaguchi Y. 2008. OS-9 and GRP94 deliver mutant alpha1-antitrypsin to the Hrd1-SEL1L ubiquitin ligase complex for ERAD. Structural basis for oligosaccharide recognition of misfolded glycoproteins by OS-9 in ER-associated degradation. Nat Cell Biol Mol Cell. 10:272–282. [PMC free article] [PubMed] [Google Scholar]
886. Hosokawa N, Kamiya Y, Kato K. 2010. The role of MRH domain-containing lectins in ERAD. Glycobiology. 20:651–660. [PubMed] [Google Scholar]
887. Gauss R, Kanehara K, Carvalho P, Ng DT, Aebi M. 2011. A complex of Pdi1p and the mannosidase Htm1p initiates clearance of unfolded glycoproteins from the endoplasmic reticulum. Mol Cell. 42:782–793. [PubMed] [Google Scholar]
888. Fujimori T, Kamiya Y, Nagata K, Kato K, Hosokawa N. 2013. Endoplasmic reticulum lectin XTP3-B inhibits endoplasmic reticulum-associated degradation of a misfolded alpha1-antitrypsin variant. FEBS J. 280:1563–1575. [PubMed] [Google Scholar]
889. Lecca MR, Wagner U, Patrignani A, Berger EG, Hennet T. 2005. Genome-wide analysis of the unfolded protein response in fibroblasts from congenital disorders of glycosylation type-I patients. FASEB J. 19:240–242. [PubMed] [Google Scholar]
890. Hirayama H, Hosomi A, Suzuki T. 2015. Physiological and molecular functions of the cytosolic peptide:N-glycanase. Semin Cell Dev Biol. 41:110–120. [PubMed] [Google Scholar]
891. Zhu Y, Liu TW, Cecioni S, Eskandari R, Zandberg WF, Vocadlo DJ. 2015. O-GlcNAc occurs cotranslationally to stabilize nascent polypeptide chains. Nat Chem Biol. 11:319–325. [PubMed] [Google Scholar]
892. Itin C, Roche AC, Monsigny M, Hauri HP. 1996. ERGIC-53 is a functional mannose-selective and calcium-dependent human homologue of leguminous lectins. Mol Biol Cell. 7:483–493. [PMC free article] [PubMed] [Google Scholar]
893. Zhang B, Cunningham MA, Nichols WC, Bernat JA, Seligsohn U, Pipe SW, McVey JH, Schulte-Overberg U, de Bosch NB, Ruiz-Saez A, et al.. 2003. Bleeding due to disruption of a cargo-specific ER-to-Golgi transport complex. Nat Genet. 34:220–225. [PubMed] [Google Scholar]
894. Everett LA, Cleuren AC, Khoriaty RN, Ginsburg D. 2014. Murine coagulation factor VIII is synthesized in endothelial cells. Blood. 123:3697–3705. [PMC free article] [PubMed] [Google Scholar]
895. Zhang B, Zheng C, Zhu M, Tao J, Vasievich MP, Baines A, Kim J, Schekman R, Kaufman RJ, Ginsburg D. 2011. Mice deficient in LMAN1 exhibit FV and FVIII deficiencies and liver accumulation of alpha1-antitrypsin. Blood. 118:3384–3391. [PMC free article] [PubMed] [Google Scholar]
896. Hara-Kuge S, Seko A, Yamashita K. 2003. Carbohydrate recognition of vesicular integral protein of 36 kDa (ViP36) in intracellular transport of newly synthesized glycoproteins. Methods Enzymol. 363:525–532. [PubMed] [Google Scholar]
897. Nufer O, Mitrovic S, Hauri HP. 2003. Profile-based data base scanning for animal L-type lectins and characterization of VIPL, a novel VIP36-like endoplasmic reticulum protein. J Biol Chem. 278:15886–15896. [PubMed] [Google Scholar]
898. Yamaguchi D, Kawasaki N, Matsuo I, Totani K, Tozawa H, Matsumoto N, Ito Y, Yamamoto K. 2007. VIPL has sugar-binding activity specific for high-mannose-type N-glycans, and glucosylation of the alpha1,2 mannotriosyl branch blocks its binding. Glycobiology. 17:1061–1069. [PubMed] [Google Scholar]
899. Kamiya Y, Kamiya D, Yamamoto K, Nyfeler B, Hauri HP, Kato K. 2008. Molecular basis of sugar recognition by the human L-type lectins ERGIC-53, VIPL, and VIP36. J Biol Chem. 283:1857–1861. [PubMed] [Google Scholar]
900. Reiterer V, Nyfeler B, Hauri HP. 2010. Role of the lectin VIP36 in post-ER quality control of human alpha1-antitrypsin. Traffic. 11:1044–1055. [PubMed] [Google Scholar]
901. Arshad N, Ballal S, Visweswariah SS. 2013. Site-specific N-linked glycosylation of receptor guanylyl cyclase C regulates ligand binding, ligand-mediated activation and interaction with vesicular integral membrane protein 36, VIP36. J Biol Chem. 288:3907–3917. [PMC free article] [PubMed] [Google Scholar]
902. Grewal PK, Uchiyama S, Ditto D, Varki N, Le DT, Nizet V, Marth JD. 2008. The Ashwell receptor mitigates the lethal coagulopathy of sepsis. Nat Med. 14:648–655. [PMC free article] [PubMed] [Google Scholar]
903. Grewal PK, Aziz PV, Uchiyama S, Rubio GR, Lardone RD, Le D, Varki NM, Nizet V, Marth JD. 2013. Inducing host protection in pneumococcal sepsis by preactivation of the Ashwell-Morell receptor. Proc Natl Acad Sci USA. 110:20218–20223. [PMC free article] [PubMed] [Google Scholar]
904. Simpson SD, Ashford DA, Harvey DJ, Bowles DJ. 1998. Short chain oligogalacturonides induce ethylene production and expression of the gene encoding aminocyclopropane 1-carboxylic acid oxidase in tomato plants. Glycobiology. 8:579–583. [PubMed] [Google Scholar]
905. Dumville JC, Fry SC. 2000. Uronic acid-containing oligosaccharins: Their biosynthesis, degradation and signalling roles in non-diseased plant tissues. Plant Physiol Biochem. 38:125–140. [Google Scholar]
906. Ridley BL, O'Neill MA, Mohnen D. 2001. Pectins: Structure, biosynthesis, and oligogalacturonide-related signaling. Phytochemistry. 57:929–967. [PubMed] [Google Scholar]
907. Gough C, Cullimore J. 2011. Lipo-chitooligosaccharide signaling in endosymbiotic plant-microbe interactions. Mol Plant Microbe Interact. 24:867–878. [PubMed] [Google Scholar]
908. Murray JD. 2011. Invasion by invitation: Rhizobial infection in legumes. Mol Plant Microbe Interact. 24:631–639. [PubMed] [Google Scholar]
909. Oldroyd GE, Murray JD, Poole PS, Downie JA. 2011. The rules of engagement in the legume-rhizobial symbiosis. Annu Rev Genet. 45:119–144. [PubMed] [Google Scholar]
910. Gough C, Jacquet C. 2013. Nod factor perception protein carries weight in biotic interactions. Trends Plant Sci. 18:566–574. [PubMed] [Google Scholar]
911. Roberts NJ, Morieri G, Kalsi G, Rose A, Stiller J, Edwards A, Xie F, Gresshoff PM, Oldroyd GE, Downie JA, et al.. 2013. Rhizobial and mycorrhizal symbioses in Lotus japonicus require lectin nucleotide phosphohydrolase, which acts upstream of calcium signaling. Plant Physiol. 161:556–567. [PMC free article] [PubMed] [Google Scholar]
912. Liang Y, Cao Y, Tanaka K, Thibivilliers S, Wan J, Choi J, Kang C, Qiu J, Stacey G. 2013. Nonlegumes respond to rhizobial Nod factors by suppressing the innate immune response. Science. 341:1384–1387. [PubMed] [Google Scholar]
913. Taylor KR, Trowbridge JM, Rudisill JA, Termeer CC, Simon JC, Gallo RL. 2004. Hyaluronan fragments stimulate endothelial recognition of injury through TLR4. J Biol Chem. 279:17079–17084. [PubMed] [Google Scholar]
914. Jiang D, Liang J, Noble PW. 2011. Hyaluronan as an immune regulator in human diseases. Physiol Rev. 91:221–264. [PMC free article] [PubMed] [Google Scholar]
915. Muto J, Morioka Y, Yamasaki K, Kim M, Garcia A, Carlin AF, Varki A, Gallo RL. 2014. Hyaluronan digestion controls DC migration from the skin. J Clin Invest. 124:1309–1319. [PMC free article] [PubMed] [Google Scholar]
916. Hauselmann I, Borsig L. 2014. Altered tumor-cell glycosylation promotes metastasis. Front Oncol. 4:28. [PMC free article] [PubMed] [Google Scholar]
917. Impellizzeri D, Cuzzocrea S. 2014. Targeting selectins for the treatment of inflammatory diseases. Expert Opin Ther Targets. 18:55–67. [PubMed] [Google Scholar]
918. McEver RP. 2015. Selectins: Initiators of leucocyte adhesion and signalling at the vascular wall. Cardiovasc Res. 107:331–339. [PMC free article] [PubMed] [Google Scholar]
919. Bochner BS. 2016. “Siglec”ting the allergic response for therapeutic targeting. Glycobiology. 26:546–552. [PMC free article] [PubMed] [Google Scholar]
920. Dykstra B, Lee J, Mortensen LJ, Yu H, Wu ZL, Lin CP, Rossi DJ, Sackstein R. 2016. Glycoengineering of E-selectin ligands by intracellular versus extracellular fucosylation differentially affects osteotropism of human mesenchymal stem cells. Stem Cells, 10.1002/stem.2435. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
921. Sackstein R. 2016. Fulfilling Koch's postulates in glycoscience: HCELL, GPS and translational glycobiology. Glycobiology. 26:560–570. [PMC free article] [PubMed] [Google Scholar]
922. Telen MJ, Wun T, McCavit TL, De Castro LM, Krishnamurti L, Lanzkron S, Hsu LL, Smith WR, Rhee S, Magnani JL, et al.. 2015. Randomized phase 2 study of GMI-1070 in SCD: Reduction in time to resolution of vaso-occlusive events and decreased opioid use. Blood. 125:2656–2664. [PMC free article] [PubMed] [Google Scholar]
923. Stähli BE, Tardif JC, Carrier M, Gallo R, Emery RW, Robb S, Cournoyer D, Blondeau L, Johnson D, Mann J, et al.. 2016. Effects of P-selectin antagonist inclacumab in patients undergoing coronary artery bypass graft surgery: SELECT-CABG trial. J Am Coll Cardiol. 67:344–346. [PubMed] [Google Scholar]
924. Benkerrou M, Delarche C, Brahimi L, Fay M, Vilmer E, Elion J, Gougerot-Pocidalo MA, Elbim C. 2002. Hydroxyurea corrects the dysregulated L-selectin expression and increased H2O2 production of polymorphonuclear neutrophils from patients with sickle cell anemia. Blood. 99:2297–2303. [PubMed] [Google Scholar]
925. Matsui NM, Varki A, Embury SH. 2002. Heparin inhibits the flow adhesion of sickle red blood cells to P-selectin. Blood. 100:3790–3796. [PubMed] [Google Scholar]
926. Embury SH, Matsui NM, Ramanujam S, Mayadas TN, Noguchi CT, Diwan BA, Mohandas N, Cheung AT. 2004. The contribution of endothelial cell P-selectin to the microvascular flow of mouse sickle erythrocytes in vivo. Blood. 104:3378–3385. [PubMed] [Google Scholar]
927. Wood K, Russell J, Hebbel RP, Grange DN. 2004. Differential expression of E- and P-selectin in the microvasculature of sickle cell transgenic mice. Microcirculation. 11:377–385. [PubMed] [Google Scholar]
928. Chang J, Patton JT, Sarkar A, Ernst B, Magnani JL, Frenette PS. 2010. GMI-1070, a novel pan-selectin antagonist, reverses acute vascular occlusions in sickle cell mice. Blood. 116:1779–1786. [PMC free article] [PubMed] [Google Scholar]
929. Wun T, Styles L, DeCastro L, Telen MJ, Kuypers F, Cheung A, Kramer W, Flanner H, Rhee S, Magnani JL, et al.. 2014. Phase 1 study of the E-selectin inhibitor GMI 1070 in patients with sickle cell anemia. PLoS One. 9:e101301. [PMC free article] [PubMed] [Google Scholar]
930. Mehta NR, Nguyen T, Bullen JW, Griffin JW, Schnaar RL. 2010. Myelin-associated glycoprotein (MAG) protects neurons from acute toxicity using a ganglioside-dependent mechanism. ACS Chem Neurosci. 1:215–222. [PMC free article] [PubMed] [Google Scholar]
931. Schnaar RL, Gerardy-Schahn R, Hildebrandt H. 2014. Sialic acids in the brain: Gangliosides and polysialic acid in nervous system development, stability, disease, and regeneration. Physiol Rev. 94:461–518. [PMC free article] [PubMed] [Google Scholar]
932. Schwardt O, Kelm S, Ernst B. 2015. SIGLEC-4 (MAG) antagonists: From the natural carbohydrate epitope to glycomimetics. Top Curr Chem. 367:151–200. [PubMed] [Google Scholar]
933. Sun J, Shaper NL, Itonori S, Heffer-Lauc M, Sheikh KA, Schnaar RL. 2004. Myelin-associated glycoprotein (Siglec-4) expression is progressively and selectively decreased in the brains of mice lacking complex gangliosides. Glycobiology. 14:851–857. [PubMed] [Google Scholar]
934. Yoo SW, Motari MG, Susuki K, Prendergast J, Mountney A, Hurtado A, Schnaar RL. 2015. Sialylation regulates brain structure and function. FASEB J. 29:3040–3053. [PMC free article] [PubMed] [Google Scholar]
935. de La Motte CA, Hascall VC, Calabro A, Yen-Lieberman B, Strong SA. 1999. Mononuclear leukocytes preferentially bind via CD44 to hyaluronan on human intestinal mucosal smooth muscle cells after virus infection or treatment with poly(I.C). J Biol Chem. 274:30747–30755. [PubMed] [Google Scholar]
936. Zhuo L, Kanamori A, Kannagi R, Itano N, Wu J, Hamaguchi M, Ishiguro N, Kimata K. 2006. SHAP potentiates the CD44-mediated leukocyte adhesion to the hyaluronan substratum. J Biol Chem. 281:20303–20314. [PubMed] [Google Scholar]
937. Ren J, Hascall VC, Wang A. 2009. Cyclin D3 mediates synthesis of a hyaluronan matrix that is adhesive for monocytes in mesangial cells stimulated to divide in hyperglycemic medium. J Biol Chem. 284:16621–16632. [PMC free article] [PubMed] [Google Scholar]
938. Wassarman PM, Litscher ES. 2001. Towards the molecular basis of sperm and egg interaction during mammalian fertilization. Cells Tissues Organs. 168:36–45. [PubMed] [Google Scholar]
939. Diekman AB. 2003. Glycoconjugates in sperm function and gamete interactions: How much sugar does it take to sweet-talk the egg. Cell Mol Life Sci. 60:298–308. [PMC free article] [PubMed] [Google Scholar]
940. Pang PC, Chiu PC, Lee CL, Chang LY, Panico M, Morris HR, Haslam SM, Khoo KH, Clark GF, Yeung WS, et al.. 2011. Human sperm binding is mediated by the sialyl-Lewis(x) oligosaccharide on the zona pellucida. Science. 333:1761–1764. [PubMed] [Google Scholar]
941. Clark GF. 2014. A role for carbohydrate recognition in mammalian sperm-egg binding. Biochem Biophys Res Commun. 450:1195–1203. [PubMed] [Google Scholar]
942. Tecle E, Gagneux P. 2015. Sugar-coated sperm: Unraveling the functions of the mammalian sperm glycocalyx. Mol Reprod Dev. 82:635–650. [PMC free article] [PubMed] [Google Scholar]
943. Ghaderi D, Springer SA, Ma F, Cohen M, Secrest P, Taylor RE, Varki A, Gagneux P. 2011. Sexual selection by female immunity against paternal antigens can fix loss of function alleles. Proc Natl Acad Sci USA. 108:17743–17748. [PMC free article] [PubMed] [Google Scholar]
944. Genbacev OD, Prakobphol A, Foulk RA, Krtolica AR, Ilic D, Singer MS, Yang ZQ, Kiessling LL, Rosen SD, Fisher SJ. 2003. Trophoblast L-selectin-mediated adhesion at the maternal-fetal interface. Science. 299:405–408. [PubMed] [Google Scholar]
945. Brinkman-Van der Linden EC, Hurtado-Ziola N, Hayakawa T, Wiggleton L, Benirschke K, Varki A, Varki N. 2007. Human-specific expression of Siglec-6 in the placenta. Glycobiology. 17:922–931. [PubMed] [Google Scholar]
946. Aplin JD, Jones CJ. 2012. Fucose, placental evolution and the glycocode. Glycobiology. 22:470–478. [PubMed] [Google Scholar]
947. Yang WH, Aziz PV, Heithoff DM, Mahan MJ, Smith JW, Marth JD. 2015. An intrinsic mechanism of secreted protein aging and turnover. Proc Natl Acad Sci USA. 112:13657–13662. [PMC free article] [PubMed] [Google Scholar]
948. Jankovic MM, Milutinovic BS. 2008. Glycoforms of CA125 antigen as a possible cancer marker. Cancer Biomark. 4:35–42. [PubMed] [Google Scholar]
949. Gupta D, Lis CG. 2009. Role of CA125 in predicting ovarian cancer survival—A review of the epidemiological literature. J Ovarian Res. 2:13. [PMC free article] [PubMed] [Google Scholar]
950. Akita K, Yoshida S, Ikehara Y, Shirakawa S, Toda M, Inoue M, Kitawaki J, Nakanishi H, Narimatsu H, Nakada H. 2012. Different levels of sialyl-Tn antigen expressed on MUC16 in patients with endometriosis and ovarian cancer. Int J Gynecol Cancer. 22:531–538. [PubMed] [Google Scholar]
951. Scarà S, Bottoni P, Scatena R. 2015. CA 19-9: Biochemical and clinical aspects. Adv Exp Med Biol. 867:247–260. [PubMed] [Google Scholar]
952. Jacobs IJ, Menon U, Ryan A, Gentry-Maharaj A, Burnell M, Kalsi JK, Amso NN, Apostolidou S, Benjamin E, Cruickshank D, et al.. 2016. Ovarian cancer screening and mortality in the UK Collaborative Trial of Ovarian Cancer Screening (UKCTOCS): A randomised controlled trial. Lancet. 387:945–956. [PMC free article] [PubMed] [Google Scholar]
953. Wang HY, Hsieh CH, Wen CN, Wen YH, Chen CH, Lu JJ. 2016. Cancers screening in an asymptomatic population by using multiple tumour markers. PLoS One. 11:e0158285. [PMC free article] [PubMed] [Google Scholar]
954. MacArthur JM, Bishop JR, Stanford KI, Wang L, Bensadoun A, Witztum JL, Esko JD. 2007. Liver heparan sulfate proteoglycans mediate clearance of triglyceride-rich lipoproteins independently of LDL receptor family members. J Clin Invest. 117:153–164. [PMC free article] [PubMed] [Google Scholar]
955. Williams KJ, Chen K. 2010. Recent insights into factors affecting remnant lipoprotein uptake. Curr Opin Lipidol. 21:218–228. [PubMed] [Google Scholar]
956. Moreth K, Brodbeck R, Babelova A, Gretz N, Spieker T, Zeng-Brouwers J, Pfeilschifter J, Young MF, Schaefer RM, Schaefer L. 2010. The proteoglycan biglycan regulates expression of the B cell chemoattractant CXCL13 and aggravates murine lupus nephritis. J Clin Invest. 120:4251–4272. [PMC free article] [PubMed] [Google Scholar]
957. Song R, Ao L, Zhao KS, Zheng D, Venardos N, Fullerton DA, Meng X. 2014. Soluble biglycan induces the production of ICAM-1 and MCP-1 in human aortic valve interstitial cells through TLR2/4 and the ERK1/2 pathway. Inflamm Res. 63:703–710. [PMC free article] [PubMed] [Google Scholar]
958. Mora-Montes HM, Bates S, Netea MG, Castillo L, Brand A, Buurman ET, Diaz-Jimenez DF, Jan Kullberg B, Brown AJ, Odds FC, et al.. 2010. A multifunctional mannosyltransferase family in Candida albicans determines cell wall mannan structure and host–fungus interactions. J Biol Chem. 285:12087–12095. [PMC free article] [PubMed] [Google Scholar]
959. Gow NA, van de Veerdonk FL, Brown AJ, Netea MG. 2012. Candida albicans morphogenesis and host defence: Discriminating invasion from colonization. Nat Rev Microbiol. 10:112–122. [PMC free article] [PubMed] [Google Scholar]
960. Rappleye CA, Eissenberg LG, Goldman WE. 2007. Histoplasma capsulatum alpha-(1,3)-glucan blocks innate immune recognition by the beta-glucan receptor. Proc Natl Acad Sci USA. 104:1366–1370. [PMC free article] [PubMed] [Google Scholar]
961. Varki A. 2011. b. Since there are PAMPs and DAMPs, there must be SAMPs? Glycan “self-associated molecular patterns” dampen innate immunity, but pathogens can mimic them. Glycobiology. 21:1121–1124. [PMC free article] [PubMed] [Google Scholar]
962. Carlin AF, Uchiyama S, Chang YC, Lewis AL, Nizet V, Varki A. 2009. Molecular mimicry of host sialylated glycans allows a bacterial pathogen to engage neutrophil Siglec-9 and dampen the innate immune response. Blood. 113:3333–3336. [PMC free article] [PubMed] [Google Scholar]
963. Chen GY, Tang J, Zheng P, Liu Y. 2009. CD24 and Siglec-10 selectively repress tissue damage-induced immune responses. Science. 323:1722–1725. [PMC free article] [PubMed] [Google Scholar]
964. Liu Y, Chen GY, Zheng P. 2011. Sialoside-based pattern recognitions discriminating infections from tissue injuries. Curr Opin Immunol. 23:41–45. [PMC free article] [PubMed] [Google Scholar]
965. Chen GY, Brown NK, Zheng P, Liu Y. 2014. Siglec-G/10 in self-nonself discrimination of innate and adaptive immunity. Glycobiology. 24:800–806. [PMC free article] [PubMed] [Google Scholar]
966. Blaum BS, Hannan JP, Herbert AP, Kavanagh D, Uhrin D, Stehle T. 2015. Structural basis for sialic acid-mediated self-recognition by complement factor H. Nat Chem Biol. 11:77–82. [PubMed] [Google Scholar]
967. Lewis LA, Gulati S, Burrowes E, Zheng B, Ram S, Rice PA. 2015. α-2,3-Sialyltransferase expression level impacts the kinetics of lipooligosaccharide sialylation, complement resistance, and the ability of Neisseria gonorrhoeae to colonize the murine genital tract. MBio. 6, 10.1128/mBio.02465-14. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
968. Ram S, Shaughnessy J, DeOliveira RB, Lewis LA, Gulati S, Rice PA. 2016. Utilizing complement evasion strategies to design complement-based antibacterial immunotherapeutics: Lessons from the pathogenic Neisseriae. Immunobiology. 221:1110–1123. [PMC free article] [PubMed] [Google Scholar]
969. Hyvärinen S, Meri S, Jokiranta TS. 2016. Disturbed sialic acid recognition on endothelial cells and platelets in complement attack causes atypical hemolytic uremic syndrome. Blood. 127:2701–2710. [PubMed] [Google Scholar]
970. Secundino I, Lizcano A, Roupé KM, Wang X, Cole JN, Olson J, Ali SR, Dahesh S, Amayreh LK, Henningham A, et al.. 2016. Host and pathogen hyaluronan signal through human siglec-9 to suppress neutrophil activation. J Mol Med (Berl). 94:219–233. [PMC free article] [PubMed] [Google Scholar]
971. Shilova N, Navakouski M, Khasbiullina N, Blixt O, Bovin N. 2012. Printed glycan array: Antibodies as probed in undiluted serum and effects of dilution. Glycoconj J. 29:87–91. [PubMed] [Google Scholar]
972. Shilova N, Huflejt ME, Vuskovic M, Obukhova P, Navakouski M, Khasbiullina N, Pazynina G, Galanina O, Bazhenov A, Bovin N. 2013. Natural antibodies against sialoglycans. Top Curr Chem. 366:169–181. [PubMed] [Google Scholar]
973. Muthana SM, Xia L, Campbell CT, Zhang Y, Gildersleeve JC. 2015. Competition between serum IgG, IgM, and IgA anti-glycan antibodies. PLoS One. 10:e0119298. [PMC free article] [PubMed] [Google Scholar]
974. Fötisch K, Vieths S. 2001. N- and O-linked oligosaccharides of allergenic glycoproteins. Glycoconj J. 18:373–390. [PubMed] [Google Scholar]
975. Wilson IBH, Zeleny R, Kolarich D, Staudacher E, Stroop CJM, Kamerling JP, Altmann F. 2001. Analysis of Asn-linked glycans from vegetable foodstuffs: Widespread occurrence of Lewis a, core alpha1,3-linked fucose and xylose substitutions. Glycobiology. 11:261–274. [PubMed] [Google Scholar]
976. Faveeuw C, Mallevaey T, Paschinger K, Wilson IB, Fontaine J, Mollicone R, Oriol R, Altmann F, Lerouge P, Capron M, et al.. 2003. Schistosome N-glycans containing core alpha3-fucose and core beta2-xylose epitopes are strong inducers of Th2 responses in mice. Eur J Immunol. 33:1271–1281. [PubMed] [Google Scholar]
977. Van Remoortere A, Bank CM, Nyame AK, Cummings RD, Deelder AM, Van Die I. 2003. Schistosoma mansoni-infected mice produce antibodies that cross-react with plant, insect, and mammalian glycoproteins and recognize the truncated biantennaryN-glycan Man3GlcNAc2-R. Glycobiology. 13:217–225. [PubMed] [Google Scholar]
978. Commins SP, James HR, Kelly LA, Pochan SL, Workman LJ, Perzanowski MS, Kocan KM, Fahy JV, Nganga LW, Ronmark E, et al.. 2011. The relevance of tick bites to the production of IgE antibodies to the mammalian oligosaccharide galactose-alpha-1,3-galactose. J Allergy Clin Immunol. 127:1286-93.e6. [PMC free article] [PubMed] [Google Scholar]
979. Mullins RJ, James H, Platts-Mills TA, Commins S. 2012. Relationship between red meat allergy and sensitization to gelatin and galactose-alpha-1,3-galactose. J Allergy Clin Immunol. 129:1334–1342.e1. [PMC free article] [PubMed] [Google Scholar]
980. Lutz AJ, Li P, Estrada JL, Sidner RA, Chihara RK, Downey SM, Burlak C, Wang ZY, Reyes LM, Ivary B, et al.. 2013. Double knockout pigs deficient in N-glycolylneuraminic acid and Galactose alpha-1,3-Galactose reduce the humoral barrier to xenotransplantation. Xenotransplantation. 20:27–35. [PubMed] [Google Scholar]
981. Burlak C, Paris LL, Lutz AJ, Sidner RA, Estrada J, Li P, Tector M, Tector AJ. 2014. Reduced binding of human antibodies to cells from GGTA1/CMAH KO pigs. Am J Transplant. 14:1895–1900. [PMC free article] [PubMed] [Google Scholar]
982. Wang ZY, Burlak C, Estrada JL, Li P, Tector MF, Tector AJ. 2014. Erythrocytes from GGTA1/CMAH knockout pigs: Implications for xenotransfusion and testing in non-human primates. Xenotransplantation. 21:376–384. [PMC free article] [PubMed] [Google Scholar]
983. Cooper DK. 2016. Modifying the sugar icing on the transplantation cake. Glycobiology. 26:571–581. [PMC free article] [PubMed] [Google Scholar]
984. Samraj AN, Laubli H, Varki N, Varki A. 2014. Involvement of a non-human sialic acid in human cancer. Front Oncol. 4:33. [PMC free article] [PubMed] [Google Scholar]
985. Alisson-Silva F, Kawanishi K, Varki A. 2016. Human risk of diseases associated with red meat intake: Analysis of current theories and proposed role for metabolic incorporation of a non-human sialic acid. Mol Aspects Med. 51:16–30. [PMC free article] [PubMed] [Google Scholar]
986. Crisp A, Boschetti C, Perry M, Tunnacliffe A, Micklem G. 2015. Expression of multiple horizontally acquired genes is a hallmark of both vertebrate and invertebrate genomes. Genome Biol. 16:50. [PMC free article] [PubMed] [Google Scholar]
987. Michon F, Brisson JR, Dell A, Kasper DL, Jennings HJ. 1988. Multiantennary group-specific polysaccharide of group B Streptococcus. Biochemistry. 27:5341–5351. [PubMed] [Google Scholar]
988. Penner JL, Aspinall GO. 1997. Diversity of lipopolysaccharide structures in Campylobacter jejuni. J Infect Dis. 176(Suppl. 2):135–S138. [PubMed] [Google Scholar]
989. Prendergast MM, Lastovica AJ, Moran AP. 1998. Lipopolysaccharides from Campylobacter jejuni O:41 strains associated with Guillain-Barre syndrome exhibit mimicry of GM1 ganglioside. Infect Immun. 66:3649–3655. [PMC free article] [PubMed] [Google Scholar]
990. Willison HJ, Yuki N. 2002. Peripheral neuropathies and anti-glycolipid antibodies. Brain. 125:2591–2625. [PubMed] [Google Scholar]
991. Kaida K, Ariga T, Yu RK. 2009. Antiganglioside antibodies and their pathophysiological effects on Guillain-Barre syndrome and related disorders—a review. Glycobiology. 19:676–692. [PMC free article] [PubMed] [Google Scholar]
992. Avril T, Wagner ER, Willison HJ, Crocker PR. 2006. Sialic acid-binding immunoglobulin-like lectin 7 mediates selective recognition of sialylated glycans expressed on Campylobacter jejuni lipooligosaccharides. Infect Immun. 74:4133–4141. [PMC free article] [PubMed] [Google Scholar]
993. Heikema AP, Bergman MP, Richards H, Crocker PR, Gilbert M, Samsom JN, van Wamel WJ, Endtz HP, van Belkum A. 2010. Characterization of the specific interaction between sialoadhesin and sialylated Campylobacter jejuni lipooligosaccharides. Infect Immun. 78:3237–3246. [PMC free article] [PubMed] [Google Scholar]
994. Bax M, Kuijf ML, Heikema AP, van Rijs W, Bruijns SC, Garcia-Vallejo JJ, Crocker PR, Jacobs BC, van Vliet SJ, van Kooyk Y. 2011. Campylobacter jejuni lipooligosaccharides modulate dendritic cell-mediated T cell polarization in a sialic acid linkage-dependent manner. Infect Immun. 79:2681–2689. [PMC free article] [PubMed] [Google Scholar]
995. Heikema AP, Jacobs BC, Horst-Kreft D, Huizinga R, Kuijf ML, Endtz HP, Samsom JN, van Wamel WJ. 2013. Siglec-7 specifically recognizes Campylobacter jejuni strains associated with oculomotor weakness in Guillain-Barre syndrome and Miller Fisher syndrome. Clin Microbiol Infect. 19:E106–E112. [PubMed] [Google Scholar]
996. Saint-Cyr MJ, Guyard-Nicodème M, Messaoudi S, Chemaly M, Cappelier JM, Dousset X, Haddad N. 2016. Recent advances in screening of anti-campylobacter activity in probiotics for use in poultry. Front Microbiol. 7:553. [PMC free article] [PubMed] [Google Scholar]
997. Khatua B, Bhattacharya K, Mandal C. 2012. Sialoglycoproteins adsorbed by Pseudomonas aeruginosa facilitate their survival by impeding neutrophil extracellular trap through siglec-9. J Leukoc Biol. 91:641–655. [PubMed] [Google Scholar]
998. Freire-de-Lima L, Fonseca LM, Oeltmann T, Mendonça-Previato L, Previato JO. 2015. The trans-sialidase, the major Trypanosoma cruzi virulence factor: Three decades of studies. Glycobiology. 25:1142–1149. [PubMed] [Google Scholar]
999. Johnston JW, Zaleski A, Allen S, Mootz JM, Armbruster D, Gibson BW, Apicella MA, Munson RSJ. 2007. Regulation of sialic acid transport and catabolism in Haemophilus influenzae. Mol Microbiol. 66:26–39. [PubMed] [Google Scholar]
1000. Parsons NJ, Patel PV, Tan EL, Andrade JR, Nairn CA, Goldner M, Cole JA, Smith H. 1988. Cytidine 5’-monophospho-N-acetyl neuraminic acid and a low molecular weight factor from human blood cells induce lipopolysaccharide alteration in gonococci when conferring resistance to killing by human serum. Microb Pathog. 5:303–309. [PubMed] [Google Scholar]
1001. Vimr ER. 2013. Unified theory of bacterial sialometabolism: How and why bacteria metabolize host sialic acids. ISRN Microbiol. 2013:816713. [PMC free article] [PubMed] [Google Scholar]
1002. Cress BF, Englaender JA, He W, Kasper D, Linhardt RJ, Koffas MA. 2014. Masquerading microbial pathogens: Capsular polysaccharides mimic host-tissue molecules. FEMS Microbiol Rev. 38:660–697. [PMC free article] [PubMed] [Google Scholar]
1003. Wessels MR, Moses AE, Goldberg JB, DiCesare TJ. 1991. Hyaluronic acid capsule is a virulence factor for mucoid group A streptococci. Proc Natl Acad Sci USA. 88:8317–8321. [PMC free article] [PubMed] [Google Scholar]
1004. DeAngelis PL, Gunay NS, Toida T, Mao WJ, Linhardt RJ. 2002. Identification of the capsular polysaccharides of Type D and F Pasteurella multocida as unmodified heparin and chondroitin, respectively. Carbohydr Res. 337:1547–1552. [PubMed] [Google Scholar]
1005. DeAngelis PL. 2002. Microbial glycosaminoglycan glycosyltransferases. Glycobiology. 12:9R–16R. [PubMed] [Google Scholar]
1006. Spath GF, Lye LF, Segawa H, Sacks DL, Turco SJ, Beverley SM. 2003. Persistence without pathology in phosphoglycan-deficient Leishmania major. Science. 301:1241–1243. [PubMed] [Google Scholar]
1007. Kamhawi S, Ramalho-Ortigao M, Pham VM, Kumar S, Lawyer PG, Turco SJ, Barillas-Mury C, Sacks DL, Valenzuela JG. 2004. A role for insect galectins in parasite survival. Cell. 119:329–341. [PubMed] [Google Scholar]
1008. Christianson HC, Belting M. 2014. Heparan sulfate proteoglycan as a cell-surface endocytosis receptor. Matrix Biol. 35:51–55. [PubMed] [Google Scholar]
1009. Stewart MD, Sanderson RD. 2014. Heparan sulfate in the nucleus and its control of cellular functions. Matrix Biol. 35:56–59. [PMC free article] [PubMed] [Google Scholar]
1010. Thacker BE, Xu D, Lawrence R, Esko JD. 2014. Heparan sulfate 3-O-sulfation: A rare modification in search of a function. Matrix Biol. 35:60–72. [PMC free article] [PubMed] [Google Scholar]
1011. Zhang X, Wang B, Li JP. 2014. Implications of heparan sulfate and heparanase in neuroinflammation. Matrix Biol. 35:174–181. [PubMed] [Google Scholar]
1012. Habuchi H, Nagai N, Sugaya N, Atsumi F, Stevens RL, Kimata K. 2007. Mice deficient in heparan sulfate 6-O-sulfotransferase-1 exhibit defective heparan sulfate biosynthesis, abnormal placentation, and late embryonic lethality. J Biol Chem. 282:15578–15588. [PubMed] [Google Scholar]
1013. Irie F, Badie-Mahdavi H, Yamaguchi Y. 2012. Autism-like socio-communicative deficits and stereotypies in mice lacking heparan sulfate. Proc Natl Acad Sci USA. 109:5052–5056. [PMC free article] [PubMed] [Google Scholar]
1014. Petridis AK, El MA, Rutishauser U. 2004. Polysialic acid regulates cell contact-dependent neuronal differentiation of progenitor cells from the subventricular zone. Dev Dyn. 230:675–684. [PubMed] [Google Scholar]
1015. Weinhold B, Seidenfaden R, Rockle I, Muhlenhoff M, Schertzinger F, Conzelmann S, Marth JD, Gerardy-Schahn R, Hildebrandt H. 2005. Genetic ablation of polysialic acid causes severe neurodevelopmental defects rescued by deletion of the neural cell adhesion molecule. J Biol Chem. 280:42971–42977. [PubMed] [Google Scholar]
1016. Burgess A, Wainwright SR, Shihabuddin LS, Rutishauser U, Seki T, Aubert I. 2008. Polysialic acid regulates the clustering, migration, and neuronal differentiation of progenitor cells in the adult hippocampus. Dev Neurobiol. 68:1580–1590. [PubMed] [Google Scholar]
1017. Franceschini I, Angata K, Ong E, Hong A, Doherty P, Fukuda M. 2001. Polysialyltransferase ST8Sia II (STX) polysialylates all of the major isoforms of NCAM and facilitates neurite outgrowth. Glycobiology. 11:231–239. [PubMed] [Google Scholar]
1018. Brocco MA, Frasch AC. 2006. Interfering polysialyltransferase ST8SiaII/STX mRNA inhibits neurite growth during early hippocampal development. FEBS Lett. 580:4723–4726. [PubMed] [Google Scholar]
1019. Takahashi K, Mitoma J, Hosono M, Shiozaki K, Sato C, Yamaguchi K, Kitajima K, Higashi H, Nitta K, Shima H, et al.. 2012. Sialidase NEU4 hydrolyzes polysialic acids of neural cell adhesion molecules and negatively regulates neurite formation by hippocampal neurons. J Biol Chem. 287:14816–14826. [PMC free article] [PubMed] [Google Scholar]
1020. Fewou SN, Ramakrishnan H, Bussow H, Gieselmann V, Eckhardt M. 2007. Down-regulation of polysialic acid is required for efficient myelin formation. J Biol Chem. 282:16700–16711. [PubMed] [Google Scholar]
1021. Jakovcevski I, Mo Z, Zecevic N. 2007. Down-regulation of the axonal polysialic acid-neural cell adhesion molecule expression coincides with the onset of myelination in the human fetal forebrain. Neuroscience. 149:328–337. [PMC free article] [PubMed] [Google Scholar]
1022. Koutsoudaki PN, Hildebrandt H, Gudi V, Skripuletz T, Skuljec J, Stangel M. 2010. Remyelination after cuprizone induced demyelination is accelerated in mice deficient in the polysialic acid synthesizing enzyme St8siaIV. Neuroscience. 171:235–244. [PubMed] [Google Scholar]
1023. Muller D, Djebbara-Hannas Z, Jourdain P, Vutskits L, Durbec P, Rougon C, Kiss JZ. 2000. Brain-derived neurotrophic factor restores long-term potentiation in polysialic acid-neural cell adhesion molecule-deficient hippocampus. Proc Natl Acad Sci USA. 97:4315–4320. [PMC free article] [PubMed] [Google Scholar]
1024. Vutskits L, Djebbara-Hannas Z, Zhang H, Paccaud JP, Durbec P, Rougon G, Muller D, Kiss JZ. 2001. PSA-NCAM modulates BDNF-dependent survival and differentiation of cortical neurons. Eur J Neurosci. 13:1391–1402. [PubMed] [Google Scholar]
1025. Kanato Y, Kitajima K, Sato C. 2008. Direct binding of polysialic acid to a brain-derived neurotrophic factor depends on the degree of polymerization. Glycobiology. 18:1044–1053. [PubMed] [Google Scholar]
1026. Muller D, Wang C, Skibo G, Toni N, Cremer H, Calaora V, Rougon G, Kiss JZ. 1996. PSA-NCAM is required for activity-induced synaptic plasticity. Neuron. 17:413–422. [PubMed] [Google Scholar]
1027. Eckhardt M, Bukalo O, Chazal G, Wang LH, Goridis C, Schachner M, Gerardy-Schahn R, Cremer H, Dityatev A. 2000. Mice deficient in the polysialyltransferase ST8SialV/PST-1 allow discrimination of the roles of neural cell adhesion molecule protein and polysialic acid in neural development and synaptic plasticity. J Neurosci. 20:5234–5244. [PMC free article] [PubMed] [Google Scholar]
1028. Kochlamazashvili G, Senkov O, Grebenyuk S, Robinson C, Xiao MF, Stummeyer K, Gerardy-Schahn R, Engel AK, Feig L, Semyanov A, et al.. 2010. Neural cell adhesion molecule-associated polysialic acid regulates synaptic plasticity and learning by restraining the signaling through GluN2B-containing NMDA receptors. J Neurosci. 30:4171–4183. [PMC free article] [PubMed] [Google Scholar]
1029. Cremer H, Lange R, Christoph A, Plomann M, Vopper G, Roes J, Brown R, Baldwin S, Kraemer P, Scheff S, et al.. 1994. Inactivation of the N-CAM gene in mice results in size reduction of the olfactory bulb and deficits in spatial learning. Nature. 367:455–459. [PubMed] [Google Scholar]
1030. Becker CG, Artola A, Gerardy-Schahn R, Decker T, Welzl H, Schachner M. 1996. The polysialic acid modification of the neural cell adhesion molecule is involved in spatial learning and hippocampal long-term potentiation. J Neurosci Res. 45:143–152. [PubMed] [Google Scholar]
1031. Senkov O, Sun M, Weinhold B, Gerardy-Schahn R, Schachner M, Dityatev A. 2006. Polysialylated neural cell adhesion molecule is involved in induction of long-term potentiation and memory acquisition and consolidation in a fear-conditioning paradigm. J Neurosci. 26:10888–109898. [PMC free article] [PubMed] [Google Scholar]
1032. Markram K, Lopez Fernandez MA, Abrous DN, Sandi C. 2007. Amygdala upregulation of NCAM polysialylation induced by auditory fear conditioning is not required for memory formation, but plays a role in fear extinction. Neurobiol Learn Mem. 87:573–582. [PubMed] [Google Scholar]
1033. Szele FG, Chesselet MF. 1996. Cortical lesions induce an increase in cell number and PSA-NCAM expression in the subventricular zone of adult rats. J Comp Neurol. 368:439–454. [PubMed] [Google Scholar]
1034. Hayashi T, Seki T, Sato K, Iwai M, Zhang WR, Manabe Y, Abe K. 2001. Expression of polysialylated neural cell adhesion molecule in rat brain after transient middle cerebral artery occlusion. Brain Res. 907:130–133. [PubMed] [Google Scholar]
1035. Franz CK, Rutishauser U, Rafuse VF. 2005. Polysialylated neural cell adhesion molecule is necessary for selective targeting of regenerating motor neurons. J Neurosci. 25:2081–2091. [PMC free article] [PubMed] [Google Scholar]
1036. Ghosh M, Tuesta LM, Puentes R, Patel S, Melendez K, El Maarouf A, Rutishauser U, Pearse DD. 2012. Extensive cell migration, axon regeneration, and improved function with polysialic acid-modified Schwann cells after spinal cord injury. Glia. 60:979–992. [PMC free article] [PubMed] [Google Scholar]
1037. Barbeau D, Liang JJ, Robitalille Y, Quirion R, Srivastava LK. 1995. Decreased expression of the embryonic form of the neural cell adhesion molecule in schizophrenic brains. Proc Natl Acad Sci USA. 92:2785–2789. [PMC free article] [PubMed] [Google Scholar]
1038. Arai M, Yamada K, Toyota T, Obata N, Haga S, Yoshida Y, Nakamura K, Minabe Y, Ujike H, Sora I, et al.. 2005. Association between polymorphisms in the promoter region of the sialyltransferase 8B (SIAT8B) gene and schizophrenia. Biol Psychiatry. 59:652-659. [PubMed] [Google Scholar]
1039. Tao R, Li C, Zheng Y, Qin W, Zhang J, Li X, Xu Y, Shi YY, Feng G, He L. 2007. Positive association between SIAT8B and schizophrenia in the Chinese Han population. Schizophr Res. 90:108–114. [PubMed] [Google Scholar]
1040. Hildebrandt H, Muhlenhoff M, Oltmann-Norden I, Rockle I, Burkhardt H, Weinhold B, Gerardy-Schahn R. 2009. Imbalance of neural cell adhesion molecule and polysialyltransferase alleles causes defective brain connectivity. Brain. 132:2831–2838. [PubMed] [Google Scholar]
1041. Isomura R, Kitajima K, Sato C. 2011. Structural and functional impairments of polysialic acid by a mutated polysialyltransferase found in schizophrenia. J Biol Chem. 286:21535–21545. [PMC free article] [PubMed] [Google Scholar]
1042. Gilabert-Juan J, Varea E, Guirado R, Blasco-Ibanez JM, Crespo C, Nacher J. 2012. Alterations in the expression of PSA-NCAM and synaptic proteins in the dorsolateral prefrontal cortex of psychiatric disorder patients. Neurosci Lett. 530:97–102. [PubMed] [Google Scholar]
1043. Sairanen M, O'Leary OF, Knuuttila JE, Castren E. 2007. Chronic antidepressant treatment selectively increases expression of plasticity-related proteins in the hippocampus and medial prefrontal cortex of the rat. Neuroscience. 144:368–374. [PubMed] [Google Scholar]
1044. Varea E, Guirado R, Gilabert-Juan J, Marti U, Castillo-Gomez E, Blasco-Ibanez JM, Crespo C, Nacher J. 2012. Expression of PSA-NCAM and synaptic proteins in the amygdala of psychiatric disorder patients. J Psychiatr Res. 46:189–197. [PubMed] [Google Scholar]
1045. Maheu ME, Davoli MA, Turecki G, Mechawar N. 2013. Amygdalar expression of proteins associated with neuroplasticity in major depression and suicide. J Psychiatr Res. 47:384–390. [PubMed] [Google Scholar]
1046. Barker JM, Torregrossa MM, Taylor JR. 2012. Low prefrontal PSA-NCAM confers risk for alcoholism-related behavior. Nat Neurosci. 15:1356–1358. [PMC free article] [PubMed] [Google Scholar]
1047. Mikkonen M, Soininen H, Kalvianen R, Tapiola T, Ylinen A, Vapalahti M, Paljarvi L, Pitkanen A. 1998. Remodeling of neuronal circuitries in human temporal lobe epilepsy: Increased expression of highly polysialylated neural cell adhesion molecule in the hippocampus and the entorhinal cortex. Ann Neurol. 44:923–934. [PubMed] [Google Scholar]
1048. Pekcec A, Weinhold B, Gerardy-Schahn R, Potschka H. 2010. Polysialic acid affects pathophysiological consequences of status epilepticus. Neuroreport. 21:549–553. [PubMed] [Google Scholar]
1049. Calandreau L, Marquez C, Bisaz R, Fantin M, Sandi C. 2010. Differential impact of polysialyltransferase ST8SiaII and ST8SiaIV knockout on social interaction and aggression. Genes Brain Behav. 9:958–967. [PubMed] [Google Scholar]
1050. Laine RA. 1994. A calculation of all possible oligosaccharide isomers both branched and linear yields 1.05 x 1012 structures for a reducing hexasaccharide: The Isomer Barrier to development of single-method saccharide sequencing or synthesis systems. Glycobiology. 4:759–767. [PubMed] [Google Scholar]
1051. Galili U. 2005. The alpha-gal epitope and the anti-Gal antibody in xenotransplantation and in cancer immunotherapy. Immunol Cell Biol. 83:674–686. [PubMed] [Google Scholar]
1052. Chou HH, Takematsu H, Diaz S, Iber J, Nickerson E, Wright KL, Muchmore EA, Nelson DL, Warren ST, Varki A. 1998. A mutation in human CMP-sialic acid hydroxylase occurred after the Homo-Pan divergence. Proc Natl Acad Sci USA. 95:11751–11756. [PMC free article] [PubMed] [Google Scholar]
1053. Springer SA, Diaz SL, Gagneux P. 2014. Parallel evolution of a self-signal: Humans and new world monkeys independently lost the cell surface sugar Neu5Gc. Immunogenetics. 66:671–674. [PMC free article] [PubMed] [Google Scholar]
1054. Schauer R, Srinivasan GV, Coddeville B, Zanetta JP, Guerardel Y. 2009. Low incidence of N-glycolylneuraminic acid in birds and reptiles and its absence in the platypus. Carbohydr Res. 344:1494–1500. [PubMed] [Google Scholar]
1055. Varki A, Gagneux P. 2009. Human-specific evolution of sialic acid targets: Explaining the malignant malaria mystery. Proc Natl Acad Sci USA. 106:14739–14740. [PMC free article] [PubMed] [Google Scholar]
1056. Moore SEH. 1999. Oligosaccharide transport: Pumping waste from the ER into lysosomes. Trends Cell Biol. 9:441–446. [PubMed] [Google Scholar]
1057. Seko A, Kitajima K, Inoue Y, Inoue S. 1991. Peptide:N-glycosidase activity found in the early embryos of Oryzias latipes (Medaka fish). The first demonstration of the occurrence of peptide:N-glycosidase in animal cells and its implication for the presence of a de-N-glycosylation system in living organisms. J Biol Chem. 266:22110–22114. [PubMed] [Google Scholar]
1058. Seino J, Fujihira H, Nakakita SI, Masahara-Negishi Y, Miyoshi E, Hirabayashi J, Suzuki T. 2016. Occurrence of free sialyl oligosaccharides related to N-glycans (sialyl FNGs) in animal sera. Glycobiology. 10.1093/glycob/cww048. [PubMed] [CrossRef] [Google Scholar]
1059. Roseman DS, Baenziger JU. 2000. Molecular basis of lutropin recognition by the mannose/GalNAc-4-SO4 receptor. Proc Natl Acad Sci USA. 97:9949–9954. [PMC free article] [PubMed] [Google Scholar]
1060. Norgard-Sumnicht KE, Varki NM, Varki A. 1993. Calcium-dependent heparin-like ligands for L-selectin in nonlymphoid endothelial cells. Science. 261:480–483. [PubMed] [Google Scholar]
1061. Norgard-Sumnicht KE, Varki A. 1995. Endothelial heparan sulfate proteoglycans that bind to L-selectin have glucosamine residues with unsubstituted amino groups. J Biol Chem. 270:12012–12024. [PubMed] [Google Scholar]
1062. Talaga ML, Fan N, Fueri AL, Brown RK, Bandyopadhyay P, Dam TK. 2016. Multitasking human lectin galectin-3 interacts with sulfated glycosaminoglycans and chondroitin sulfate proteoglycans. Biochemistry. 55:4541–4551. [PubMed] [Google Scholar]
1063. Hsu KL, Pilobello KT, Mahal LK. 2006. Analyzing the dynamic bacterial glycome with a lectin microarray approach. Nat Chem Biol. 2:153–157. [PubMed] [Google Scholar]
1064. Propheter DC, Hsu KL, Mahal LK. 2011. Recombinant lectin microarrays for glycomic analysis. Methods Mol Biol. 723:67–77. [PubMed] [Google Scholar]
1065. Rakus JF, Mahal LK. 2011. New technologies for glycomic analysis: Toward a systematic understanding of the glycome. Annu Rev Anal Chem (Palo Alto Calif). 4:367–392. [PubMed] [Google Scholar]
1066. Ribeiro JP, Mahal LK. 2013. Dot by dot: Analyzing the glycome using lectin microarrays. Curr Opin Chem Biol. 17:827–831. [PMC free article] [PubMed] [Google Scholar]
1067. Fearon DT. 1978. Regulation by membrane sialic acid of beta1H-dependent decay-dissociation of amplification C3 convertase of the alternative complement pathway. Proc Natl Acad Sci USA. 75:1971–1975. [PMC free article] [PubMed] [Google Scholar]
1068. Pangburn MK, Muller-Eberhard HJ. 1978. Complement C3 convertase: Cell surface restriction of beta1H control and generation of restriction on neuraminidase-treated cells. Proc Natl Acad Sci USA. 75:2416–2420. [PMC free article] [PubMed] [Google Scholar]
1069. Sun Y, Senger K, Baginski TK, Mazloom A, Chinn Y, Pantua H, Hamidzadeh K, Ramani SR, Luis E, Tom I, et al.. 2012. Evolutionarily conserved paired immunoglobulin-like receptor alpha (PILRalpha) domain mediates its interaction with diverse sialylated ligands. J Biol Chem. 287:15837–15850. [PMC free article] [PubMed] [Google Scholar]
1070. Kitazume S, Imamaki R, Ogawa K, Taniguchi N. 2014. a. Sweet role of platelet endothelial cell adhesion molecule in understanding angiogenesis. Glycobiology. 24:1260–1264. [PubMed] [Google Scholar]
1071. Kitazume S, Imamaki R, Kurimoto A, Ogawa K, Kato M, Yamaguchi Y, Tanaka K, Ishida H, Ando H, Kiso M, et al.. 2014. b. Interaction of platelet endothelial cell adhesion molecule (PECAM) with alpha2,6-sialylated glycan regulates its cell surface residency and anti-apoptotic role. J Biol Chem. 289:27604–27613. [PMC free article] [PubMed] [Google Scholar]
1072. Gout E, Garlatti V, Smith DF, Lacroix M, Dumestre-Perard C, Lunardi T, Martin L, Cesbron JY, Arlaud GJ, Gaboriaud C, et al.. 2010. Carbohydrate recognition properties of human ficolins: Glycan array screening reveals the sialic acid binding specificity of M-ficolin. J Biol Chem. 285:6612–6622. [PMC free article] [PubMed] [Google Scholar]
1073. Kiessling LL, Grim JC. 2013. Glycopolymer probes of signal transduction. Chem Soc Rev. 42:4476–4491. [PMC free article] [PubMed] [Google Scholar]
1074. Woods RJ. 1995. Three-dimensional structures of oligosaccharides. Curr Opin Struct Biol. 5:591–598. [PubMed] [Google Scholar]
1075. Brisson JR, Uhrinova S, Woods RJ, van der Zwan M, Jarrell HC, Paoletti LC, Kasper DL, Jennings HJ. 1997. NMR and molecular dynamics studies of the conformational epitope of the type III group B Streptococcus capsular polysaccharide and derivatives. Biochemistry. 36:3278–3292. [PubMed] [Google Scholar]
1076. Marki A, Esko JD, Pries AR, Ley K. 2015. Role of the endothelial surface layer in neutrophil recruitment. J Leukoc Biol. 98:503–515. [PMC free article] [PubMed] [Google Scholar]
1077. Dimitroff CJ, Lee JY, Fuhlbrigge RC, Sackstein R. 2000. A distinct glycoform of CD44 is an L-selectin ligand on human hematopoietic cells. Proc Natl Acad Sci USA. 97:13841–13846. [PMC free article] [PubMed] [Google Scholar]
1078. Dimitroff CJ, Lee JY, Rafii S, Fuhlbrigge RC, Sackstein R. 2001. CD44 is a major E-selectin ligand on human hematopoietic progenitor cells. J Cell Biol. 153:1277–1286. [PMC free article] [PubMed] [Google Scholar]
1079. Crick F. 1970. Central dogma of molecular biology. Nature. 227:561–563. [PubMed] [Google Scholar]
1080. Agre P, Bertozzi C, Bissell M, Campbell KP, Cummings RD, Desai UR, Estes M, Flotte T, Fogleman G, Gage F, et al.. 2016. Training the next generation of biomedical investigators in glycosciences. J Clin Invest. 126:405–408. [PMC free article] [PubMed] [Google Scholar]
1081. Feizi T, Mulloy B. 2003. Carbohydrates and glycoconjugates. Glycomics: The new era of carbohydrate biology. Curr Opin Struct Biol. 13:602–604. [PubMed] [Google Scholar]
1082. North SJ, Hitchen PG, Haslam SM, Dell A. 2009. Mass spectrometry in the analysis of N-linked and O-linked glycans. Curr Opin Struct Biol. 19:498–506. [PMC free article] [PubMed] [Google Scholar]
1083. Hart GW, Copeland RJ. 2010. Glycomics hits the big time. Cell. 143:672–676. [PMC free article] [PubMed] [Google Scholar]
1084. Tao N, Wu S, Kim J, An HJ, Hinde K, Power ML, Gagneux P, German JB, Lebrilla CB. 2011. Evolutionary glycomics: characterization of milk oligosaccharides in primates. J Proteome Res. 10:1548–1557. [PMC free article] [PubMed] [Google Scholar]
1085. Aoki-Kinoshita KF. 2013. Using databases and web resources for glycomics research. Mol Cell Proteomics. 12:1036–1045. [PMC free article] [PubMed] [Google Scholar]
1086. Wells L, Hart GW. 2013. Glycomics: Building upon proteomics to advance glycosciences. Mol Cell Proteomics. 12:833–835. [PMC free article] [PubMed] [Google Scholar]
1087. Zoldos V, Horvat T, Lauc G. 2013. Glycomics meets genomics, epigenomics and other high throughput omics for system biology studies. Curr Opin Chem Biol. 17:34–40. [PubMed] [Google Scholar]
1088. Cummings RD, Pierce JM. 2014. The challenge and promise of glycomics. Chem Biol. 21:1–15. [PMC free article] [PubMed] [Google Scholar]
1089. Levery SB, Steentoft C, Halim A, Narimatsu Y, Clausen H, Vakhrushev SY. 2015. Advances in mass spectrometry driven O-glycoproteomics. Biochim Biophys Acta. 1850:33–42. [PubMed] [Google Scholar]
1090. Hsu KL, Gildersleeve JC, Mahal LK. 2008. A simple strategy for the creation of a recombinant lectin microarray. Mol Biosyst. 4:654–662. [PMC free article] [PubMed] [Google Scholar]
1091. Gambetta MC, Oktaba K, Muller J. 2009. Essential role of the glycosyltransferase sxc/Ogt in polycomb repression. Science. 325:93–96. [PubMed] [Google Scholar]
1092. Sinclair DA, Syrzycka M, Macauley MS, Rastgardani T, Komljenovic I, Vocadlo DJ, Brock HW, Honda BM. 2009. Drosophila O-GlcNAc transferase (OGT) is encoded by the Polycomb group (PcG) gene, super sex combs (sxc). Proc Natl Acad Sci USA. 106:13427–13432. [PMC free article] [PubMed] [Google Scholar]
1093. Manzi AE, Higa HH, Diaz S, Varki A. 1994. Intramolecular self-cleavage of polysialic acid. J Biol Chem. 269:23617–23624. [PubMed] [Google Scholar]
1094. Varki A. 2013. Omics: Account for the “dark matter” of biology. Nature. 497:565. [PubMed] [Google Scholar]
1095. Glycosciences Committee on Assessing the Importance and Impact of Glycomics and Technology. 2012. Transforming Glycoscience: A Roadmap for the Future. A National Research Council Report. Washington, DC: The National Academies Press. [PubMed] [Google Scholar]

Articles from Glycobiology are provided here courtesy of Oxford University Press

-