Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-24T11:32:56.461Z Has data issue: false hasContentIssue false

Comparisons of N-glycans across invertebrate phyla

Published online by Cambridge University Press:  03 May 2019

Katharina Paschinger
Affiliation:
Department für Chemie, Universität für Bodenkultur, 1190 Wien, Austria
Iain B. H. Wilson*
Affiliation:
Department für Chemie, Universität für Bodenkultur, 1190 Wien, Austria
*
Author for correspondence: Iain B. H. Wilson, E-mail: iain.wilson@boku.ac.at

Abstract

Many invertebrates are either parasites themselves or vectors involved in parasite transmission; thereby, the interactions of parasites with final or intermediate hosts are often mediated by glycans. Therefore, it is of interest to compare the glycan structures or motifs present across invertebrate species. While a typical vertebrate modification such as sialic acid is rare in lower animals, antennal and core modifications of N-glycans are highly varied and range from core fucose, galactosylated fucose, fucosylated galactose, methyl groups, glucuronic acid and sulphate through to addition of zwitterionic moieties (phosphorylcholine, phosphoethanolamine and aminoethylphosphonate). Only in some cases are the enzymatic bases and the biological function of these modifications known. We are indeed still in the phase of discovering invertebrate glycomes primarily using mass spectrometry, but molecular biology and microarraying techniques are complementary to the determination of novel glycan structures and their functions.

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © Cambridge University Press 2019

Introduction

The co- or post-translational addition of glycans to proteins takes various forms in all kingdoms of life (Varki, Reference Varki2011); amongst the most common is N-glycosylation, by which asparagine residues are modified. In eukaryotes, most commonly a Glc3Man9GlcNAc2 precursor is transferred from dolichol to proteins in the endoplasmic reticulum (Aebi, Reference Aebi2013); however, some protists utilise shorter precursors or even do not N-glycosylate at all (Samuelson et al., Reference Samuelson, Banerjee, Magnelli, Cui, Kelleher, Gilmore and Robbins2005). The fates of protein-linked N-glycans are varied and depend on the types of glycosidases and glycosyltransferases expressed in the Golgi apparatus. It is this variability that makes glycan analysis a challenge, as so many possibilities occur by which N-glycans are trimmed and then built up again.

Other than the first steps in the endoplasmic reticulum, the final size and form of N-glycans differ between protists, fungi, plants and animals (whether invertebrate or vertebrate), although some modifications are found in more than one of these groups of organisms. Unlike plants whose N-glycomes are similar from mosses through to Arabidopsis, there is high variability between non-vertebrate eukaryotes (Schiller et al., Reference Schiller, Hykollari, Yan, Paschinger and Wilson2012). Here, we will concentrate on primarily structural aspects of invertebrate N-glycans, not only due to the parasitological relevance (as many invertebrates are either hosts, vectors or themselves parasites), but also because only recently have mass spectrometric analyses revealed a previously unrealised range of modifications, some of which are shared with O- and lipid-linked glycans. A few years ago, one would probably have read that invertebrates only produce oligomannosidic (Man5–9GlcNAc2) and paucimannosidic (Man1–4GlcNAc2Fuc0–2) N-glycans (Williams et al., Reference Williams, Wormald, Dwek, Rademacher, Parker and Roberts1991); this may be due to insensitive methods and low expectations, but it is now known that even complete glycomes of some mammalian cell types are dominated by oligomannosidic forms present within the secretory pathway (Hamouda et al., Reference Hamouda, Kaup, Ullah, Berger, Sandig, Tauber and Blanchard2014).

Oligomannosidic N-glycans

Even within the glycans containing primarily mannose residues (hence oligomannosidic or high mannose), there is variation arising from the different orders of processing by so-called class I α1,2-mannosidases, also in parasitic metazoa. Most eukaryotes have multiple forms of these α1,2-mannosidases (Wilson, Reference Wilson2012), which also include enzymes known as EDEMs (ER degradation-enhancing α-mannosidases) acting as part of the quality control pathway in the endoplasmic reticulum. The result is that there are multiple isomers of oligomannosidic structures (e.g. three isomers of glycans with the composition Man8GlcNAc2; Fig. 1) just depending on which mannosidase acts first on particular terminal mannose residues; the final product is the ‘Golgi’ isomer of Man5GlcNAc2. These structures can be differentiated by, e.g. RP-HPLC in combination with MS/MS and thus it is only appropriate to annotate specific isomers based on such information; for instance, a Hex8HexNAc2 structure could also be Glc1Man7GlcNAc2 and not necessarily one of three typical forms of Man8GlcNAc2. Oligomannosidic glycans may also be the ‘final’ processed forms in the cases where protein folding prevents a specific glycosylation site from being accessible to enzymes in the Golgi apparatus (Thaysen-Andersen and Packer, Reference Thaysen-Andersen and Packer2012). On the basis of the universality of oligomannosidic glycans in metazoa, it is not surprising that these glycans have been observed in a wide range of invertebrates including trematodes, nematodes, molluscs and insects.

Fig. 1. Simplified biosynthetic scheme for N-linked glycans in animals. Starting with the Glc3Man9GlcNAc2 precursor, various glycosidases result in different isomers of oligomannosidic glycans with the maximal degree of processing by class I mannosidases yielding Man5GlcNAc2. This is the substrate for N-acetylglucosaminyltransferase I (GlcNAc-TI) which generates a ‘hybrid’ structure which can be further modified by the action of Golgi mannosidase II, GlcNAc-TII and Golgi hexosaminidase. The maximum number of antennae (three or four) depends on the presence of GlcNAc-TIV and GlcNAc-TV; example hybrid, pseudohybrid, paucimannosidic and tri-/tetra-antennary glycans are shown as known from various model, host, vector or parasitic invertebrates. For simplicity, fucosylation and other modifications are not included. Glycans are depicted according to the Standard Nomenclature for Glycans (see also box).

Hybrid and pseudohybrid N-glycans

The classical ‘hybrid’ structure is a ‘Golgi-type’ Man5GlcNAc2 modified on the ‘lower’ α1,3-mannose by β1,2-specific N-acetylglucosaminyltransferase I (GlcNAc-TI; encoded by the mammalian MGAT1 gene and its homologues in multicellular eukaryotes, including parasitic invertebrates) to yield Man5GlcNAc3 (Fig. 1) which may be the substrate for further modification. This is a key intermediate in N-glycan biosynthesis (Fig. 1) in terms of the routes to processed structures as well as in biological terms (Schachter, Reference Schachter2010), as ablation of this gene results in large glycomic shifts as well as a range of phenotypes (Shi et al., Reference Shi, Tan and Schachter2006; Sarkar et al., Reference Sarkar, Iliadi, Leventis, Schachter and Boulianne2010; Yan et al., Reference Yan, Wang, Schachter, Jin, Wilson and Paschinger2018b), most dramatically, the embryonic lethal phenotype in mammals (Ioffe and Stanley, Reference Ioffe and Stanley1994; Metzler et al., Reference Metzler, Gertz, Sarkar, Schachter, Schrader and Marth1994).

Once GlcNAc-TI has acted, Golgi mannosidase II will remove one or two of the ‘upper’ mannose residues and also represents a potential biosynthetic bottleneck (Paschinger et al., Reference Paschinger, Hackl, Gutternigg, Kretschmer-Lubich, Stemmer, Jantsch, Lochnit and Wilson2006). If there is no transfer by GlcNAc-TII thereafter to the α1,6-mannose, then the glycan remains in a hybrid state (i.e. sharing aspects of oligomannosidic and complex structures); also even if GlcNAc-TIV (β1,4-specific) modifies the α1,3-mannose, then the glycan is still classified as being hybrid (Kornfeld and Kornfeld, Reference Kornfeld and Kornfeld1985). The lower arm β1,2- and β1,4-GlcNAc residues on hybrid glycans can be modified in different ways; elongation by β1,4-N-acetylgalactosamine and β1,3- or β1,4-galactose are known in insects, molluscs, nematodes and trematodes, whether these be host or parasitic organisms (Nyame et al., Reference Nyame, Smith, Damian and Cummings1989; Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013, Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015; Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015). If however, GlcNAc-TII acts and then the ‘lower’ arm β1,2-GlcNAc, transferred by GlcNAc-TI, is removed by a Golgi hexosaminidase such as fdl (fused lobes) in insects or HEX-2 in nematodes (Gutternigg et al., Reference Gutternigg, Kretschmer-Lubich, Paschinger, Rendić, Hader, Geier, Ranftl, Jantsch, Lochnit and Wilson2007b; Geisler and Jarvis, Reference Geisler and Jarvis2012), then the resulting glycans can be referred to as ‘pseudohybrid’. Such structures are also found in protist parasites lacking GlcNAc-TI, but having GlcNAc-TII-like enzyme activities (Paschinger et al., Reference Paschinger, Hykollari, Razzazi-Fazeli, Greenwell, Leitsch, Walochnik and Wilson2012b; Damerow et al., Reference Damerow, Rodrigues, Wu, Güther, Mehlert and Ferguson2014). The core of hybrid glycans in animals can also be modified, most commonly by α1,6-fucose.

Paucimannosidic N-glycans

The term ‘paucimannosidic’ glycans was introduced to cover those glycans which have been processed serially by GlcNAc-TI, Golgi mannosidase II and a Golgi hexosaminidase to result in Man3–4GlcNAc2 (Gutternigg et al., Reference Gutternigg, Kretschmer-Lubich, Paschinger, Rendić, Hader, Geier, Ranftl, Jantsch, Lochnit and Wilson2007b). Such structures are well known in invertebrates and plants, but also occur due to the action of acidic glycosidases on glycoproteins in the secretory granules in some mammalian cells (Loke et al., Reference Loke, Ostergaard, Heegaard, Packer and Thaysen-Andersen2017). A significant portion of paucimannosidic glycans are core fucosylated and carry the ‘mammalian-like’ α1,6-fucose and the ‘plant-like’ α1,3-fucose either alone or in combination on the reducing-terminal (proximal) GlcNAc of the core region of the N-glycan as found first on bee venom glycoproteins (Kubelka et al., Reference Kubelka, Altmann, Staudacher, Tretter, März, Hård, Kamerling and Vliegenthart1993). In nematodes, the second (distal) GlcNAc can also be modified (Haslam et al., Reference Haslam, Coles, Munn, Smith, Smith, Morris and Dell1996; Hanneman et al., Reference Hanneman, Rosa, Ashline and Reinhold2006).

Both proximal core α1,3-fucose and substitution of the β-mannose by β1,2-xylose (see Fig. 2 for example structures) are immunogenic in mammals and antibodies raised against plant and invertebrate glycoproteins often recognise these epitopes, the best known example of which is anti-horseradish peroxidase (anti-HRP); both structural elements are epitopes for IgE or IgG in parasite-infected animals or children as well as in individuals allergic to plant pollen, food or insect venom, although the clinical relevance is controversial (van Die et al., Reference van Die, Gomord, Kooyman, van der Berg, Cummings and Vervelde1999; Altmann, Reference Altmann2007; Paschinger et al., Reference Paschinger, Rendić and Wilson2009; Brzezicka et al., Reference Brzezicka, Echeverria, Serna, van Diepen, Hokke and Reichardt2015; Amoah et al., Reference Amoah, Asuming-Brempong, Obeng, Versteeg, Larbi, Aryeetey, Platts-Mills, Mari, Brzezicka, Gyan, Mutocheluh, Boakye, Reichardt, van Ree, Hokke, van Diepen and Yazdanbakhsh2018). While core α1,3-fucose is widespread in invertebrates, xylosylation of N-glycans is known from gastropods and, in a stage-specific manner, Schistosoma spp. (Khoo et al., Reference Khoo, Chatterjee, Caulfield, Morris and Dell1997; Gutternigg et al., Reference Gutternigg, Bürgmayr, Pöltl, Rudolf and Staudacher2007a; Lehr et al., Reference Lehr, Frank, Natsuka, Geyer, Beuerlein, Doenhoff, Hase and Geyer2010; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015). This is interesting as some gastropods (specifically snails such as Biomphalaria glabrata) are intermediate hosts for schistosomes. The activities of core-modifying fucosyl- and xylosyltransferases have been detected in extracts of various species, but only for core α1,3/α1,6-difucosylation have relevant genes been identified and recombinant forms of the enzymes characterised (Fabini et al., Reference Fabini, Freilinger, Altmann and Wilson2001; Paschinger et al., Reference Paschinger, Staudacher, Stemmer, Fabini and Wilson2005; Rendić et al., Reference Rendić, Klaudiny, Stemmer, Schmidt, Paschinger and Wilson2007; Kurz et al., Reference Kurz, King, Dinglasan, Paschinger and Wilson2016).

Fig. 2. Example N-glycans from invertebrates. Structures are depicted from either parasitic or free-living organisms, whereby some of the latter are hosts or vectors for parasites. Some types of structures are species- or class-specific, but others are found in more than one phylum. Only a non-exhaustive selection of core and antennal epitopes is shown in the inset: core difucosylation, core ‘GalFuc’, Lewis X (LeX), fucosylated and non-fucosylated LacdiNAc (LDN) and blood group A (BGA). (A) The bisecting and distal core modifications found in the free-living C. elegans are indicated by pink boxes; (B) free-living C. elegans, the necromenic P. pacificus and the parasites H. contortus, H. polygyrus and O. dentatum express di- and/or tri-fucosylated cores with species-specific galactosylation and methylation; (C) varying antennal modifications are found in all nematodes as well as the cestodes E. granulosus and T. crassiceps, (D) while filarial species have up to four long antennae including D. immitis, which has in addition glucuronylated structures; (E) galactosylated core fucose (GalFuc) is found in many invertebrates, sometimes in substituted form; (F, G and H) selected complex glycans from larvae of different insect phyla; (I) selected S. mansoni N-glycan modifications which are partly stage-specific; (J, K and L) selected gastropod and bivalve glycans, including those of Crassostrea virginica, B. glabrata, Volvarina rubella and Mytilus edulis. Note that some modifications, such as core β-mannosylation, are at low abundance in the relevant glycomes. Glycans are depicted according to the Standard Nomenclature for Glycans; undefined hexoses/N-acetylhexosamines are shown as white circles/squares. Me, methyl; MAEP, N-methyl-aminoethylphosphonate; PC, phosphorylcholine; PE, phosphoethanolamine (2-aminoethylphosphate); S, sulphate. Broken lines,±or brackets indicate structure-, species- or stage-dependent variations in these elements.

Modified N-glycan cores

In addition to fucosylation and xylosylation, some invertebrates attach further monosaccharide units to the basic paucimannosidic core. Recently joining the list of core modifications alongside galactosylation of core α1,6-fucose (‘GalFuc’), first detected in squid and then in keyhole limpet, planaria and nematodes (Takahashi et al., Reference Takahashi, Masuda, Hiraki, Yoshihara, Huang, Khoo and Kato2004; Wuhrer et al., Reference Wuhrer, Robijn, Koeleman, Balog, Geyer, Deelder and Hokke2004; Titz et al., Reference Titz, Butschi, Henrissat, Fan, Hennet, Razzazi-Fazeli, Hengartner, Wilson, Künzler and Aebi2009; Paschinger et al., Reference Paschinger, Razzazi-Fazeli, Furukawa and Wilson2011; Subramanian et al., Reference Subramanian, Babu, Palakodeti and Subramanian2018), are α-galactosylation of the proximal and distal core α1,3-fucose residues (Yan et al., Reference Yan, Vanbeselaere, Jin, Blaukopf, Wols, Wilson and Paschinger2018a), elongation of the GalFuc unit by galactose (Wuhrer et al., Reference Wuhrer, Robijn, Koeleman, Balog, Geyer, Deelder and Hokke2004; Subramanian et al., Reference Subramanian, Babu, Palakodeti and Subramanian2018), fucose (Yan et al., Reference Yan, Jin, Wilson and Paschinger2015b), phosphorylcholine or methylaminoethylphosphonate (Eckmair et al., Reference Eckmair, Jin, Abed-Navandi and Paschinger2016), β-mannosylation of the proximal GlcNAc (Eckmair et al., Reference Eckmair, Jin, Abed-Navandi and Paschinger2016; Hykollari et al., Reference Hykollari, Malzl, Eckmair, Vanbeselaere, Scheidl, Jin, Karlsson, Wilson and Paschinger2018) and the galactosylation of the core β-mannose to form a bisected structure, which can also be modified by methylated or nonmethylated fucose (Yan et al., Reference Yan, Brecker, Jin, Titz, Dragosits, Karlsson, Jantsch, Wilson and Paschinger2015a) (Fig. 2). While the latter bisecting modifications have only been found in the non-parasitic nematode Caenorhabditis elegans, the zwitterionic modifications of the GalFuc have been detected uniquely in a marine gastropod; however, galactosylation of the proximal α1,6-fucose and distal α1,3-fucose residues has also been found in the parasitic nematodes Oesophagostomum dentatum and Haemonchus contortus (Paschinger and Wilson, Reference Paschinger and Wilson2015; Sutov, Reference Sutov2016; Jiménez-Castells et al., Reference Jiménez-Castells, Vanbeselaere, Kohlhuber, Ruttkowski, Joachim and Paschinger2017).

Some of the reason for the apparent restriction in what is found might be methodological. For instance, the presence of α-galactose on the proximal α1,3-fucose was only detected in C. elegans when using hydrazine or the newly-developed PNGase Ar enzyme to release the N-glycans, whereby the maximal degree of core fucosylation in this worm (five fucoses) was only found after hydrazinolysis (Yan et al., Reference Yan, Vanbeselaere, Jin, Blaukopf, Wols, Wilson and Paschinger2018a). Only in the case of O. dentatum can we say that these modifications are absent, since hydrazinolysis was also performed with samples from this organism and MS/MS did not reveal any glycan with the relevant fragmentation pattern (Jiménez-Castells et al., Reference Jiménez-Castells, Vanbeselaere, Kohlhuber, Ruttkowski, Joachim and Paschinger2017). On the other hand, H. contortus glycans were only ever analysed after ‘classical’ PNGase F and A digestion and so it can only be speculated as to whether it shares more complex cores with C. elegans.

The enzymatic basis for only some of these modifications is known. Three core-modifying α-fucosyltransferases (FUT-1, FUT-6 and FUT-8) are known from C. elegans as is the α1,6-fucose-modifying GALT-1 galactosyltransferase from the same organism (Paschinger et al., Reference Paschinger, Rendić, Lochnit, Jantsch and Wilson2004, Reference Paschinger, Staudacher, Stemmer, Fabini and Wilson2005; Titz et al., Reference Titz, Butschi, Henrissat, Fan, Hennet, Razzazi-Fazeli, Hengartner, Wilson, Künzler and Aebi2009; Yan et al., Reference Yan, Serna, Reichardt, Paschinger and Wilson2013). The in vitro activity data is complemented by glycomic studies on mutants showing the absence of the relevant epitopes (Butschi et al., Reference Butschi, Titz, Wälti, Olieric, Paschinger, Nöbauer, Guo, Seeberger, Wilson, Aebi, Hengartner and Künzler2010; Yan et al., Reference Yan, Jin, Wilson and Paschinger2015b). Some of these glyco-mutants have altered susceptibility to nematoxic fungal lectins (Butschi et al., Reference Butschi, Titz, Wälti, Olieric, Paschinger, Nöbauer, Guo, Seeberger, Wilson, Aebi, Hengartner and Künzler2010; Schubert et al., Reference Schubert, Bleuler-Martinez, Butschi, Walti, Egloff, Stutz, Yan, Wilson, Hengartner, Aebi, Allain and Künzler2012), which are also toxic to H. contortus (Heim et al., Reference Heim, Hertzberg, Butschi, Bleuler-Martinez, Aebi, Deplazes, Künzler and Stefanic2015). For all the other modifications around the core, e.g. the addition of various α-galactose residues or of bisecting β-galactose in C. elegans we have no clues as to which enzymes may be responsible. The same lack of knowledge applies to β-mannosylation of the proximal core GlcNAc in molluscs and insects.

Complex N-glycans

The definition ‘complex N-glycan’ is based on the knowledge of mammalian glycosylation and refers to glycans with at least one GlcNAc modifying both α-mannose residues of the trimannosyl core. Thus, both GlcNAc-TI and -TII (MGAT1 and MGAT2) have acted and these can be supplemented by GlcNAc-TIV, GlcNAc-TV and in some species ‘GlcNAc-TVI’ (Schachter, Reference Schachter1986). The other common N-acetylglucosaminyltransferase, GlcNAc-TIII, is a bisecting enzyme found in vertebrates. The result of the action of these various enzymes (see Fig. 1) is the various bi-, tri- and tetra-antennary glycans (even penta-antennary in birds and fish), which are well known from the serum glycomes of mammals.

It may come as a surprise that even relatively primitive animals have tri- or tetra-antennary N-glycans as found in Hydra, molluscs, insects and nematodes (Kang et al., Reference Kang, Cummings and McCall1993; Morelle et al., Reference Morelle, Haslam, Olivier, Appleton, Morris and Dell2000; Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013, Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015; Sahadevan et al., Reference Sahadevan, Antonopoulos, Haslam, Dell, Ramaswamy and Babu2014; Eckmair et al., Reference Eckmair, Jin, Abed-Navandi and Paschinger2016). The exact nature of the tri-antennary glycans varies, as GlcNAc-TIV products occur in molluscs and insects (Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013, Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015), but GlcNAc-TV acts in glycan biosynthesis in a number of nematodes such as C. elegans and Pristionchus pacificus (Yan et al., Reference Yan, Wilson and Paschinger2015c). Other nematodes, though, do have both GlcNAc-TIV and -TV homologues and so can have up to four branches on their N-glycans as found in filarial species or in Trichinella (Haslam et al., Reference Haslam, Houston, Harnett, Reason, Morris and Dell1999; Kang et al., Reference Kang, Cummings and McCall1993; Morelle et al., Reference Morelle, Haslam, Olivier, Appleton, Morris and Dell2000; Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019).

Amongst trematodes, triantennary glycans have been long established to exist in S. mansoni males (Nyame et al., Reference Nyame, Smith, Damian and Cummings1989); it has also been suggested that up to four branches may also be on N-glycans of S. mansoni eggs or in Opisthorchis viverrini (Talabnin et al., Reference Talabnin, Aoki, Saichua, Wongkham, Kaewkes, Boons, Sripa and Tiemeyer2013; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015); however, as in another trematode Fasciola hepatica (McVeigh et al., Reference McVeigh, Cwiklinski, Garcia-Campos, Mulcahy, O'Neill, Maule and Dalton2018), BLAST searching of the available genomes only shows an obvious GlcNAc-TV homologue and none of GlcNAc-TIV (unpublished data). However, only for C. elegans GlcNAc-TI and GlcNAc-TII is there in vitro evidence from recombinant enzymes to verify the predicted activities (Chen et al., Reference Chen, Tan, Reinhold, Spence and Schachter2002), while C. elegans GlcNAc-TV has been shown to complement a relevant Chinese hamster ovary mutant cell line in terms of lectin sensitivity (Warren et al., Reference Warren, Krizius, Roy, Culotti and Dennis2002); the activity of an invertebrate GlcNAc-TIV has still to be proven.

Subsequent to the initial transfer of up to four non-reducing terminal GlcNAc residues, further elongation events can occur and these are extremely variable and, in non-vertebrates, include substitutions with β1,3-galactose, β1,4-N-acetylgalactosamine, α1,4-N-acetylgalactosamine or fucose as well as anionic, zwitterionic or methyl groups (Fig. 2). The typical mammalian form of galactosylation (β1,4) is not so widespread in lower animals in general, but it can, e.g. be found in Schistosoma spp. (Khoo et al., Reference Khoo, Chatterjee, Caulfield, Morris and Dell1997; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015); it can only be distinguished from β1,3-galactosylation by use of specific galactosidases and, if amounts allow, by NMR spectroscopy or GC-MS methods. Antennal GlcNAc residues modified with β1,3Gal or even β1,3Galβ1,4GalNAc are found on N-glycans from, e.g. mosquitoes acting as intermediate hosts for parasites and viruses (Kurz et al., Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015) and β1,3-Gal is also found in the oyster Crassostrea virginica, which is a host for the Perkinsus marinus protist parasite (Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013). On the other hand, GalNAcβ1,4GlcNAc (LacdiNAc) is a known motif from various insects and nematodes (see also the section on fucosylated antennae below) and a stage-specific bias in its expression is known from trematode parasites (Talabnin et al., Reference Talabnin, Aoki, Saichua, Wongkham, Kaewkes, Boons, Sripa and Tiemeyer2013; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015). Longer chito-based (GlcNAcβ1,4GlcNAc) antennae are a feature of filarial nematodes as well as of H. contortus and O. dentatum (Haslam et al., Reference Haslam, Houston, Harnett, Reason, Morris and Dell1999; Sutov, Reference Sutov2016; Jiménez-Castells et al., Reference Jiménez-Castells, Vanbeselaere, Kohlhuber, Ruttkowski, Joachim and Paschinger2017; Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019).

For some of these terminal modifications, the relevant enzymes have been identified and characterised in recombinant form, such as β1,4-N-acetylgalactosaminyltransferases from C. elegans and Trichoplusia ni, a β1,3-galactosyltransferase from the honeybee and an α1,4-N-acetylgalactosaminyltransferase from Drosophila (Kawar et al., Reference Kawar, van Die and Cummings2002; Mucha et al., Reference Mucha, Domlatil, Lochnit, Rendić, Paschinger, Hinterkörner, Hofinger, Kosma and Wilson2004; Vadaie and Jarvis, Reference Vadaie and Jarvis2004; Ichimiya et al., Reference Ichimiya, Maeda, Sakamura, Kanazawa, Nishihara and Kimura2015). However, although some relevant enzyme activities have been detected in crude extracts, the identities of relevant genes in parasites are yet to be established.

Antennally fucosylated N-glycans

Fucose as a deoxyhexose rather than a standard hexose may well, due to its chemical properties, be pre-destined to act as a recognition element. Indeed, fucose is the basis for mammalian histo-blood group antigens such as ABO and Lewis motifs (Fig. 2). Fucosylated LacNAc (Lex) and LacdiNAc (LDNF) epitopes are well known from S. mansoni (Khoo et al., Reference Khoo, Chatterjee, Caulfield, Morris and Dell1997; Wuhrer et al., Reference Wuhrer, Koeleman, Deelder and Hokke2006; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015) and may contribute to the lectin-dependent immunomodulatory activity of secreted schistosome proteins (Wilbers et al., Reference Wilbers, Westerhof, van Noort, Obieglo, Driessen, Everts, Gringhuis, Schramm, Goverse, Smant, Bakker, Smits, Yazdanbakhsh, Schots and Hokke2017). Also, some nematodes (e.g. Dictyocaulus viviparus, Trichuris suis or H. contortus) and insects (e.g. the honeybee) express these epitopes (Kubelka et al., Reference Kubelka, Altmann, Staudacher, Tretter, März, Hård, Kamerling and Vliegenthart1993; Haslam et al., Reference Haslam, Coles, Morris and Dell2000; Paschinger and Wilson, Reference Paschinger and Wilson2015; Wilson and Paschinger, Reference Wilson and Paschinger2016), while fucosylated chito-oligomers are a feature of the antennae of some N-glycans from Dirofilaria immitis (Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019) and Fucα1,3GlcNAc as a terminal motif is also known from the cestode Taenia crassiceps (Lee et al., Reference Lee, Dissanayake, Panico, Morris, Dell and Haslam2005). Less familiar may be the occurrence of blood group A on oyster glycans (Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013), which are probable ligands for noroviruses in the marine environment, but which are also recognised by the oyster's own galectins (Feng et al., Reference Feng, Ghosh, Amin, Giomarelli, Shridhar, Banerjee, Fernandez-Robledo, Bianchet, Wang, Wilson and Vasta2013). Interestingly, though, these galectins also mediate entry of P. marinus into oyster haemocytes, despite the apparent lack of blood group antigens on the parasite.

Generally, fucose on glycan antennae is unsubstituted, but branched fucose (i.e. disubstituted) is known in some molluscs (Zhou et al., Reference Zhou, Hanneman, Chasteen and Reinhold2013; Eckmair et al., Reference Eckmair, Jin, Abed-Navandi and Paschinger2016) and fucosylated fucose (Fucα1,2Fucα1,3) occurs in S. mansoni (Jang-Lee et al., Reference Jang-Lee, Curwen, Ashton, Tissot, Mathieson, Panico, Dell, Wilson and Haslam2007; Smit et al., Reference Smit, van Diepen, Nguyen, Wuhrer, Hoffmann, Deelder and Hokke2015) (Fig. 2). Interestingly, the various fucosylated antennal modifications of S. mansoni are epitopes for various natural and monoclonal antibodies (van Remoortere et al., Reference van Remoortere, Hokke, van Dam, van Die, Deelder and van den Eijnden2000; van Diepen et al., Reference van Diepen, Smit, van Egmond, Kabatereine, Pinot de Moira, Dunne and Hokke2012) and may mediate interactions of parasitic proteins with cells of the host immune system (Meevissen et al., Reference Meevissen, Driessen, Smits, Versteegh, van Vliet, van Kooyk, Schramm, Deelder, Haas, Yazdanbakhsh and Hokke2012). The schistosome genome encodes a number of fucosyltransferases, but only one has proven enzymatic activity in recombinant form, specifically as a LeX synthase (Mickum et al., Reference Mickum, Rojsajjakul, Yu and Cummings2016b). Other defined invertebrate Lewis-type fucosyltransferases include the FucTC from the honeybee and a mosquito (Kurz et al., Reference Kurz, King, Dinglasan, Paschinger and Wilson2016; Rendić et al., Reference Rendić, Klaudiny, Stemmer, Schmidt, Paschinger and Wilson2007).

Methylated N-glycans

Substitution of glycans by methyl groups is known in bacteria, plants and invertebrates. In the case of N-glycans from mollusc, planaria and free-living or parasitic nematodes, examples include methylation of mannose, fucose, galactose and N-acetylgalactosamine residues (van Kuik et al., Reference van Kuik, Sijbesma, Kamerling, Vliegenthart and Wood1986, Reference Van Kuik, Sijbesma, Kamerling, Vliegenthart and Wood1987b; Gutternigg et al., Reference Gutternigg, Bürgmayr, Pöltl, Rudolf and Staudacher2007a; Paschinger et al., Reference Paschinger, Razzazi-Fazeli, Furukawa and Wilson2011; Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013; Hewitson et al., Reference Hewitson, Nguyen, van Diepen, Smit, Koeleman, McSorley, Murray, Maizels and Hokke2016; Jiménez-Castells et al., Reference Jiménez-Castells, Vanbeselaere, Kohlhuber, Ruttkowski, Joachim and Paschinger2017; Yan et al., Reference Yan, Vanbeselaere, Jin, Blaukopf, Wols, Wilson and Paschinger2018a) (Fig. 2). If analysing the glycans using standard permethylation conditions, such natural methyl groups are lost; thus, perdeuteromethylation has to be employed (Wohlschlager et al., Reference Wohlschlager, Butschi, Grassi, Sutov, Gauss, Hauck, Schmieder, Knobel, Titz, Dell, Haslam, Hengartner, Aebi and Künzler2014). For standard exoglycosidase sequencing, methylation normally prevents removal of a residue, but the methylated GalNAc on oyster glycans could be removed with chicken α-N-acetylgalactosaminidase (Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013), while methylated α1,2- or α1,3-fucose residues on nematode glycans can be partially or fully released by hydrofluoric acid treatment (Yan et al., Reference Yan, Vanbeselaere, Jin, Blaukopf, Wols, Wilson and Paschinger2018a). The type of methylation can also vary within a species, as methylation of mannose was more common in male O. dentatum parasites as opposed to the methylfucose residues found in the female (Jiménez-Castells et al., Reference Jiménez-Castells, Vanbeselaere, Kohlhuber, Ruttkowski, Joachim and Paschinger2017).

Glucuronylated and sialylated N-glycans

By separating neutral from anionic glycans early in the analyses, we have been able to find glucuronic acid on the termini of N-glycans from a number of species, including mosquitoes, moths and the honeybee, as well as a marine snail (Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013, Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015; Eckmair et al., Reference Eckmair, Jin, Abed-Navandi and Paschinger2016; Stanton et al., Reference Stanton, Hykollari, Eckmair, Malzl, Dragosits, Palmberger, Wang, Wilson and Paschinger2017; Hykollari et al., Reference Hykollari, Malzl, Eckmair, Vanbeselaere, Scheidl, Jin, Karlsson, Wilson and Paschinger2018) (Fig. 2). Like methylated hexose residues, the presence of glucuronic acid results in a mass increment of 176 Da, but GlcA-containing glycans can be detected by negative mode mass spectrometry (Hykollari et al., Reference Hykollari, Paschinger, Eckmair and Wilson2017). Using permethylation, others also detected glucuronic acid on N-glycans of Drosophila (Aoki and Tiemeyer, Reference Aoki and Tiemeyer2010), whereas we have also used glucuronidases to help prove its occurrence on oligosaccharide structures from other insects (Stanton et al., Reference Stanton, Hykollari, Eckmair, Malzl, Dragosits, Palmberger, Wang, Wilson and Paschinger2017; Hykollari et al., Reference Hykollari, Malzl, Eckmair, Vanbeselaere, Scheidl, Jin, Karlsson, Wilson and Paschinger2018). Except for Dirofilaria immitis (Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019), there are no reports to date of GlcA on N-glycans of nematodes or trematodes, but glycosaminoglycan chains and O-glycans from these species do contain this residue (Palaima et al., Reference Palaima, Leymarie, Stroud, Mizanur, Hodgkin, Gravato-Nobre, Costello and Cipollo2010; Vanbeselaere et al., Reference Vanbeselaere, Yan, Joachim, Paschinger and Wilson2018), including the circulating anodic antigen of S. mansoni (Bergwerff et al., Reference Bergwerff, Van Dam, Rotmans, Deelder, Kamerling and Vliegenthart1994).

As glucuronic acid is a major component of glycosaminoglycans and these are known to play roles in host-pathogen interactions (Pinzon-Ortiz et al., Reference Pinzon-Ortiz, Friedman, Esko and Sinnis2001; Armistead et al., Reference Armistead, Wilson, van Kuppevelt and Dinglasan2011), one can speculate that glucuronic acid on N-glycans may be another ligand involved in, e.g. Plasmodium transmission by mosquitoes. The role of glucuronylation of Dirofilaria N-glycans is also unclear. The actual transfer of glucuronic acid to N-glycans has not been proven for any invertebrate glucuronyltransferase, other than for two enzymes of broad specificity from Drosophila (Kim et al., Reference Kim, Tsuchida, Lincecum, Kitagawa, Bernfield and Sugahara2003).

In terms of sialylation, for which there is no hint in most invertebrates, its occurrence in insects has been controversial. Other than mass spectrometric studies on N-glycans from Drosophila embryos (Aoki et al., Reference Aoki, Perlman, Lim, Cantu, Wells and Tiemeyer2007; Frappaolo et al., Reference Frappaolo, Sechi, Kumagai, Robinson, Fraschini, Karimpour-Ghahnavieh, Belloni, Piergentili, Tiemeyer, Tiemeyer and Giansanti2017), there is no firm proof to date for sialylation in any other insect; this is despite genome sequencing typically indicating the presence of one sialyltransferase homologue per insect species, some of which have proven in vitro activities (Koles et al., Reference Koles, Irvine and Panin2004; Kajiura et al., Reference Kajiura, Hamaguchi, Mizushima, Misaki and Fujiyama2015). Higher up the evolutionary tree, however, there is good evidence for sialic acids on the O-glycans of Echinodermata (Miyata et al., Reference Miyata, Sato, Kumita, Toriyama, Vacquier and Kitajima2006). Lectin binding data, although suggestive, is too ambiguous to be considered proof of the presence of sialic acid in unknown glycomes, as the ‘summarised’ specificities of many lectins are probably a simplification, but also contamination must be considered if detecting sialylation in glycans of a parasite derived from a mammalian host.

Sulphated and phosphorylated N-glycans

Another surprisingly widespread anionic modification is sulphate, which results in signals in negative mode mass MS and a Δm/z of 80 mass units. Thereby, for many instruments, sulphate cannot be differentiated from phosphate; however, some very high resolution mass spectrometers can be used to distinguish these. Other proofs include the ionisation of phosphate in both positive and negative mode or the susceptibility of phosphate (and not of sulphate) to hydrofluoric acid or phosphatase treatments (Hykollari et al., Reference Hykollari, Paschinger, Eckmair and Wilson2017). By pre-separating neutral and anionic glycans prior to off-line LC-MS, we have detected sulphate in marine molluscs (including oyster) and in insects (including mosquitoes). On the other hand, standard permethylation procedures will result in loss of sulphated glycans, but modified solid phase extraction methods are compatible with subsequent detection of permethylated sulphated glycans as performed with mosquito or royal jelly N-glycans (Kurz et al., Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015; Hykollari et al., Reference Hykollari, Malzl, Eckmair, Vanbeselaere, Scheidl, Jin, Karlsson, Wilson and Paschinger2018).

Sulphation of invertebrate N-glycans may occur at different positions, e.g. of mannose or core fucose in arthropods or of galactose as in oyster (van Kuik et al., Reference van Kuik, Breg, Kolsteeg, Kamerling and Vliegenthart1987a; Kurz et al., Reference Kurz, Jin, Hykollari, Gregorich, Giomarelli, Vasta, Wilson and Paschinger2013, Reference Kurz, Aoki, Jin, Karlsson, Tiemeyer, Wilson and Paschinger2015) (Fig. 2), but we have yet to definitely prove sulphation in a parasite. Others have detected phosphorylation of mannose residues in F. hepatica (Ravida et al., Reference Ravida, Aldridge, Driessen, Heus, Hokke and O'Neill2016). The mannose-6-phosphorylation system known for trafficking of lysosomal enzymes in vertebrates is not proven in any invertebrate; strangely, though, a mannose phosphorylation mediated by a homologue of the relevant GlcNAc-1-phosphotransferase enzyme is found in an amoeba (Qian et al., Reference Qian, West and Kornfeld2010). There is no information regarding any N-glycan-modifying sulpho- or phosphotransferase from any invertebrate.

Zwitterionic N-glycans

Phosphodiester and phosphonate modifications such as phosphorylcholine, phosphoethanolamine and aminoethylphosphonate may be familiar to many from bacterial lipopolysaccharides and glycosylphosphatidylinositol anchors or related molecules, but have been reported on a number of invertebrate N-, O- and lipid-linked glycans. While detection of these modifications is incompatible with permethylation procedures, they can all be released with hydrofluoric acid (HF) and so some earlier reports for their presence were based partly on detection of permethylated forms of ‘stripped’ glycans as well as of perdeuteroacetylated structures without HF treatment (Haslam et al., Reference Haslam, Houston, Harnett, Reason, Morris and Dell1999; Morelle et al., Reference Morelle, Haslam, Olivier, Appleton, Morris and Dell2000). However, when conducting more ‘native’ mass spectrometric analyses, phosphorylcholine (PC; Δm/z 165 mass units) ionises very well in positive mode and is a widespread modification of nematode N-glycans (Hanneman et al., Reference Hanneman, Rosa, Ashline and Reinhold2006; Pöltl et al., Reference Pöltl, Kerner, Paschinger and Wilson2007; Paschinger and Wilson, Reference Paschinger and Wilson2015; Hewitson et al., Reference Hewitson, Nguyen, van Diepen, Smit, Koeleman, McSorley, Murray, Maizels and Hokke2016; Wilson and Paschinger, Reference Wilson and Paschinger2016; Jiménez-Castells et al., Reference Jiménez-Castells, Vanbeselaere, Kohlhuber, Ruttkowski, Joachim and Paschinger2017; Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019), but has also been found in a cestode (Echinococcus granulosus) and more recently on moth N-glycans (Paschinger et al., Reference Paschinger, Gonzalez-Sapienza and Wilson2012a; Stanton et al., Reference Stanton, Hykollari, Eckmair, Malzl, Dragosits, Palmberger, Wang, Wilson and Paschinger2017) (Fig. 2).

Phosphoethanolamine (PE; Δm/z 123), aminoethylphosphonate (AEP; Δm/z 107) and methylaminophosphonate (MEAP; Δm/z 121) are detected in both positive and negative modes (Paschinger and Wilson, Reference Paschinger and Wilson2016). PE is found on N-glycans of royal jelly, AEP on those of a locust glycoprotein and MEAP on the antennae and core regions of N-glycans from a marine snail (Hård et al., Reference Hård, Van Doorn, Thomas-Oates, Kamerling and Van der Horst1993; Eckmair et al., Reference Eckmair, Jin, Abed-Navandi and Paschinger2016) (Fig. 2); other reports have shown PC, PE and MEAP on glycolipids or O-glycans of various invertebrates, including Ascaris suum (Hayashi and Matsubara, Reference Hayashi and Matsubara1989; Sugita et al., Reference Sugita, Fujii, Inagaki, Suzuki, Hayata and Hori1992; Lochnit et al., Reference Lochnit, Dennis, Ulmer and Geyer1998; Seppo et al., Reference Seppo, Moreland, Schweingruber and Tiemeyer2000; Maes et al., Reference Maes, Garenaux, Strecker, Leroy, Wieruszeski, Brassart and Guerardel2005; Urai et al., Reference Urai, Nakamura, Uzawa, Baba, Taniguchi, Seki and Ushida2009).

PC and PE are ligands for pentraxins and so binding of Echinococcus Ag5 or of Dirofilaria glycans to C-reactive protein or of royal jelly N-glycans to serum amyloid P have been shown (Paschinger et al., Reference Paschinger, Gonzalez-Sapienza and Wilson2012a; Hykollari et al., Reference Hykollari, Malzl, Eckmair, Vanbeselaere, Scheidl, Jin, Karlsson, Wilson and Paschinger2018; Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019). On the other hand, PC modifications of glycoconjugates are associated with immunomodulation; a well-known example of this being the ES-62 excretory-secretory protein from the filarial worm Acanthocheilonema viteae (Pineda et al., Reference Pineda, Lumb, Harnett and Harnett2014). The biosynthesis of zwitterionic N-glycans remains unresolved, other than a requirement for the prior action of GlcNAc-TI in C. elegans (Houston et al., Reference Houston, Sutharsan, Steiger, Schachter and Harnett2008), but comparisons with pathways in bacteria and fungi may help in the future to decipher the molecular basis for these reactions.

N-glycan arrays

Glycans mediate function when they can be recognised and glycan arrays have become an established method for determining which proteins can bind them. However, other than S. mansoni (van Diepen et al., Reference van Diepen, Smit, van Egmond, Kabatereine, Pinot de Moira, Dunne and Hokke2012; Mickum et al., Reference Mickum, Prasanphanich, Song, Dorabawila, Mandalasi, Lasanajak, Luyai, Secor, Wilkins, Van Die, Smith, Nyame, Cummings and Rivera-Marrero2016a), studies using natural structures are in their relative infancy for invertebrates, but pools or fractions of natural N-glycans from royal jelly, Dirofilaria and C. elegans have been tested recently in an immobilised format with pentraxins, selected antibodies or standard lectins (Hykollari et al., Reference Hykollari, Malzl, Eckmair, Vanbeselaere, Scheidl, Jin, Karlsson, Wilson and Paschinger2018; Jankowska et al., Reference Jankowska, Parsons, Song, Smith, Cummings and Cipollo2018; Martini et al., Reference Martini, Eckmair, Neupert, Štefanić, Jin, Garg, Jiménez-Castells, Hykollari, Yan, Venco, Varón Silva, Wilson and Paschinger2019). The bias in the literature towards schistosome arrays is probably due to a number of factors, such as availability of the various stages of the life-cycle and of monoclonal antibodies as well as three decades of relevant glycomic research. Thus, it has been possible to construct arrays of N-, O- and lipid-linked glycans derived from different stages of the schistosome life-cycle and screen them, e.g. with antibodies or antisera (van Diepen et al., Reference van Diepen, Smit, van Egmond, Kabatereine, Pinot de Moira, Dunne and Hokke2012, Reference van Diepen, van der Plas, Kozak, Royle, Dunne and Hokke2015; Yang et al., Reference Yang, Li, Brzezicka, Reichardt, Wilson, van Diepen and Hokke2017, Reference Yang, van Diepen, Brzezicka, Reichardt and Hokke2018). Otherwise, some anti-helminth antibody responses have been tested against the primarily mammalian array of the Consortium for Functional Glycomics, remodelled glycans or conjugates with shorter saccharides to identify potential protective or diagnostic epitopes (van Stijn et al., Reference van Stijn, van den Broek, Vervelde, Alvarez, Cummings, Tefsen and van Die2009; Aranzamendi et al., Reference Aranzamendi, Tefsen, Jansen, Chiumiento, Bruschi, Kortbeek, Smith, Cummings, Pinelli and Van Die2011; Luyai et al., Reference Luyai, Heimburg-Molinaro, Prasanphanich, Mickum, Lasanajak, Song, Nyame, Wilkins, Rivera-Marrero, Smith, Van Die, Secor and Cummings2014).

Another option is to use chemoenzymatic synthesis to replicate natural glycostructural motifs and so some structures akin or identical to those of schistosomes or nematodes have been prepared, in part with our defined C. elegans FUT-1, FUT-6 and FUT-8 core fucosyltransferases (Yan et al., Reference Yan, Serna, Reichardt, Paschinger and Wilson2013). The resulting synthetic arrays, which can also be studied in parallel to natural arrays, have been probed with, e.g. human lectins, anti-Schistosoma monoclonal antibodies or with the sera of Schistosoma-infected humans or macaques (Brzezicka et al., Reference Brzezicka, Echeverria, Serna, van Diepen, Hokke and Reichardt2015; Yang et al., Reference Yang, Li, Brzezicka, Reichardt, Wilson, van Diepen and Hokke2017, Reference Yang, van Diepen, Brzezicka, Reichardt and Hokke2018; Echeverria et al., Reference Echeverria, Serna, Achilli, Vives, Pham, Thepaut, Hokke, Fieschi and Reichardt2018). Thereby, some detailed insights into recognised structures can be obtained; for instance, antibodies recognising fucosylated antennae may correlate with the stage of parasite infection, while the presence of xylose or the exact antennal N-glycan configuration may have a negative role on lectin binding or be associated with skewed IgG subtype reactivity.

Conclusion

With this brief summary of the different categories of N-glycan modifications in invertebrates, we hope the reader will appreciate the great glycomic variety. Comparing parasitic, non-parasitic and host species is still far from complete; thus, it is still difficult to state whether certain N-glycans or epitopes are themselves hallmarks for parasitism or tropism. It may well be that each parasite has adopted aspects of its ancestors' or its hosts' glycomic capacity in order to fill a specific patho-ecological niche. On the other hand, knowledge about the glycomic status of hosts for recombinant protein production (e.g. insect cell lines) is important before, or may even aid, their use as factories for production of vaccines against parasites. In any case, only carefully performed glycomics can yield the deepest knowledge about invertebrate glycans, including exclusion of host glycans from the analyses, and is a pre-requisite for binding and other functional studies. Here, one challenge is to isolate sufficient natural glycans from parasites or related species or to recreate the structures in vitro. Another is to identify relevant ‘glycozyme’ genes, which will allow more recombinant glycosyltransferases to be used in chemoenzymatic synthesis and, as CRISPR/Cas9-based genetic engineering is beginning to be used in metazoan parasites (Gang et al., Reference Gang, Castelletto, Bryant, Yang, Mancuso, Lopez, Pellegrini and Hallem2017; McVeigh and Maule, Reference McVeigh and Maule2019), enable the switching on/off of certain glycosylation pathways. Indeed, a mix of analytical, biological and chemical tools will certainly prove valuable in the future to not only define the binding partners of specific glycans, but to predict their wider evolutionary occurrence and determine their function in host-parasite interactions.

Author ORCIDs

Katharina Paschinger, 0000-0002-3594-7136; Iain B. H. Wilson, 0000-0001-8996-1518.

Financial support

Work on our laboratory relevant to this review has been funded by the Austrian Science Fund (FWF; grants P21946 and P25058 to K.P. and P23922 and P29466 to I.B.H.W.) and the European Union (Glycopar EU FP7 Marie Curie Initial Training Network; PITN-GA-2013-608295).

Conflict of interest

None.

Ethical standards

Not applicable.

References

Aebi, M (2013) N-linked protein glycosylation in the ER. Biochimica et Biophysica Acta 1833, 24302437.Google Scholar
Altmann, F (2007) The role of protein glycosylation in allergy. International Archives of Allergy and Immunology 142, 99115.Google Scholar
Amoah, AS, Asuming-Brempong, EK, Obeng, BB, Versteeg, SA, Larbi, IA, Aryeetey, Y, Platts-Mills, TAE, Mari, A, Brzezicka, K, Gyan, BA, Mutocheluh, M, Boakye, DA, Reichardt, NC, van Ree, R, Hokke, CH, van Diepen, A and Yazdanbakhsh, M (2018) Identification of dominant anti-glycan IgE responses in school children by glycan microarray. Journal of Allergy and Clinical Immunology 141, 11301133.Google Scholar
Aoki, K and Tiemeyer, M (2010) The glycomics of glycan glucuronylation in Drosophila melanogaster. Methods in Enzymology 480, 297321.Google Scholar
Aoki, K, Perlman, M, Lim, JM, Cantu, R, Wells, L and Tiemeyer, M (2007) Dynamic developmental elaboration of N-linked glycan complexity in the Drosophila melanogaster embryo. Journal of Biological Chemistry 282, 91279142.Google Scholar
Aranzamendi, C, Tefsen, B, Jansen, M, Chiumiento, L, Bruschi, F, Kortbeek, T, Smith, DF, Cummings, RD, Pinelli, E and Van Die, I (2011) Glycan microarray profiling of parasite infection sera identifies the LDNF glycan as a potential antigen for serodiagnosis of trichinellosis. Experimental Parasitology 129, 221226.Google Scholar
Armistead, JS, Wilson, IBH, van Kuppevelt, TH and Dinglasan, RR (2011) A role for heparan sulfate proteoglycans in Plasmodium falciparum sporozoite invasion of anopheline mosquito salivary glands. Biochemical Journal 438, 475483.Google Scholar
Bergwerff, AA, Van Dam, GJ, Rotmans, JP, Deelder, AM, Kamerling, JP and Vliegenthart, JFG (1994) The immunologically reactive part of immunopurified circulating anodic antigen from Schistosoma mansoni is a threonine- linked polysaccharide consisting of →6)-(β-D-GlcpA-(1→3))-β-D-GalpNAc-(1→ repeating units. Journal of Biological Chemistry 269, 3151031517.Google Scholar
Brzezicka, K, Echeverria, B, Serna, S, van Diepen, A, Hokke, CH and Reichardt, NC (2015) Synthesis and microarray-assisted binding studies of core xylose and fucose containing N-glycans. ACS Chemical Biology 10, 12901302.Google Scholar
Butschi, A, Titz, A, Wälti, M, Olieric, V, Paschinger, K, Nöbauer, K, Guo, X, Seeberger, PH, Wilson, IBH, Aebi, M, Hengartner, M and Künzler, M (2010) Caenorhabditis elegans N-glycan core β-galactoside confers sensitivity towards nematotoxic fungal galectin CGL2. PLOS Pathogens 6, e1000717.Google Scholar
Chen, S, Tan, J, Reinhold, VN, Spence, AM and Schachter, H (2002) UDP-N-acetylglucosamine:α-3-D-mannoside β-1,2-N-acetylglucosaminyltransferase I and UDP-N-acetylglucosamine:α-6-D-mannoside β-1,2-N-acetylglucosaminyltransferase II in Caenorhabditis elegans. Biochimica et Biophysica Acta 1573, 271279.Google Scholar
Damerow, M, Rodrigues, JA, Wu, D, Güther, ML, Mehlert, A and Ferguson, MAJ (2014) Identification and functional characterization of a highly divergent N-acetylglucosaminyltransferase I (TbGnTI) in Trypanosoma brucei. Journal of Biological Chemistry 289, 93289339.Google Scholar
Echeverria, B, Serna, S, Achilli, S, Vives, C, Pham, J, Thepaut, M, Hokke, CH, Fieschi, F and Reichardt, NC (2018) Chemoenzymatic synthesis of N-glycan positional isomers and evidence for branch selective binding by monoclonal antibodies and human C-type lectin receptors. ACS Chemical Biology 13, 22692279.Google Scholar
Eckmair, B, Jin, C, Abed-Navandi, D and Paschinger, K (2016) Multi-step fractionation and mass spectrometry reveals zwitterionic and anionic modifications of the N- and O-glycans of a marine snail. Molecular and Cellular Proteomics 15, 573597.Google Scholar
Fabini, G, Freilinger, A, Altmann, F and Wilson, IBH (2001) Identification of core α1,3-fucosylated glycans and the requisite fucosyltransferase in Drosophila melanogaster. Potential basis of the neural anti-horseradish peroxidase epitope. Journal of Biological Chemistry 276, 2805828067.Google Scholar
Feng, C, Ghosh, A, Amin, MN, Giomarelli, B, Shridhar, S, Banerjee, A, Fernandez-Robledo, JA, Bianchet, MA, Wang, LX, Wilson, IBH and Vasta, GR (2013) The galectin CvGal1 from the eastern oyster (Crassostrea virginica) binds to blood group A oligosaccharides on the hemocyte surface. Journal of Biological Chemistry 288, 2439424409.Google Scholar
Frappaolo, A, Sechi, S, Kumagai, T, Robinson, S, Fraschini, R, Karimpour-Ghahnavieh, A, Belloni, G, Piergentili, R, Tiemeyer, KH, Tiemeyer, M and Giansanti, MG (2017) COG7 deficiency in Drosophila generates multifaceted developmental, behavioral and protein glycosylation phenotypes. Journal of Cell Science 130, 36373649.Google Scholar
Gang, SS, Castelletto, ML, Bryant, AS, Yang, E, Mancuso, N, Lopez, JB, Pellegrini, M and Hallem, EA (2017) Targeted mutagenesis in a human-parasitic nematode. PLOS Pathogens 13, e1006675.Google Scholar
Geisler, C and Jarvis, DL (2012) Substrate specificities and intracellular distributions of three N-glycan processing enzymes functioning at a key branch point in the insect N-glycosylation pathway. Journal of Biological Chemistry 287, 70847097.Google Scholar
Gutternigg, M, Bürgmayr, S, Pöltl, G, Rudolf, J and Staudacher, E (2007 a) Neutral N-glycan patterns of the gastropods Limax maximus, Cepaea hortensis, Planorbarius corneus, Arianta arbustorum and Achatina fulica. Glycoconjugate Journal 24, 475489.Google Scholar
Gutternigg, M, Kretschmer-Lubich, D, Paschinger, K, Rendić, D, Hader, J, Geier, P, Ranftl, R, Jantsch, V, Lochnit, G and Wilson, IBH (2007b) Biosynthesis of truncated N-linked oligosaccharides results from non-orthologous hexosaminidase-mediated mechanisms in nematodes, plants and insects. Journal of Biological Chemistry 282, 2782527840.Google Scholar
Hamouda, H, Kaup, M, Ullah, M, Berger, M, Sandig, V, Tauber, R and Blanchard, V (2014) Rapid analysis of cell surface N-glycosylation from living cells using mass spectrometry. Journal of Proteome Research 13, 61446151.Google Scholar
Hanneman, AJ, Rosa, JC, Ashline, D and Reinhold, V (2006) Isomer and glycomer complexities of core GlcNAcs in Caenorhabditis elegans. Glycobiology 16, 874890.Google Scholar
Hård, K, Van Doorn, JM, Thomas-Oates, JE, Kamerling, JP and Van der Horst, DJ (1993) Structure of the Asn-linked oligosaccharides of apolipophorin III from the insect Locusta migratoria. Carbohydrate- linked 2-aminoethylphosphonate as a constituent of a glycoprotein. Biochemistry 32, 766775.Google Scholar
Haslam, SM, Coles, GC, Munn, EA, Smith, TS, Smith, HF, Morris, HR and Dell, A (1996) Haemonchus contortus glycoproteins contain N-linked oligosaccharides with novel highly fucosylated core structures. Journal of Biological Chemistry 271, 3056130570.Google Scholar
Haslam, SM, Houston, KM, Harnett, W, Reason, AJ, Morris, HR and Dell, A (1999) Structural studies of N-glycans of filarial parasites. Conservation of phosphorylcholine-substituted glycans among species and discovery of novel chito-oligomers. Journal of Biological Chemistry 274, 2095320960.Google Scholar
Haslam, SM, Coles, GC, Morris, HR and Dell, A (2000) Structural characterisation of the N-glycans of Dictyocaulus viviparus: discovery of the Lewisx structure in a nematode. Glycobiology 10, 223229.Google Scholar
Hayashi, A and Matsubara, T (1989) A New homolog of phosphonoglycosphingolipid, N-methylaminoethylphosphonyltrigalactosylceramide. Biochimica Et Biophysica Acta 1006, 8996.Google Scholar
Heim, C, Hertzberg, H, Butschi, A, Bleuler-Martinez, S, Aebi, M, Deplazes, P, Künzler, M and Stefanic, S (2015) Inhibition of Haemonchus contortus larval development by fungal lectins. Parasites & Vectors 8, 425.Google Scholar
Hewitson, JP, Nguyen, DL, van Diepen, A, Smit, CH, Koeleman, CA, McSorley, HJ, Murray, J, Maizels, RM and Hokke, CH (2016) Novel O-linked methylated glycan antigens decorate secreted immunodominant glycoproteins from the intestinal nematode Heligmosomoides polygyrus. International Journal for Parasitology 46, 157170.Google Scholar
Houston, KM, Sutharsan, R, Steiger, CN, Schachter, H and Harnett, W (2008) Gene inactivation confirms the identity of enzymes involved in nematode phosphorylcholine-N-glycan synthesis. Molecular and Biochemical Parasitology 157, 8891.Google Scholar
Hykollari, A, Paschinger, K, Eckmair, B and Wilson, IBH (2017) Analysis of invertebrate and protist N-glycans. Methods in Molecular Biology 1503, 167184.Google Scholar
Hykollari, A, Malzl, D, Eckmair, B, Vanbeselaere, J, Scheidl, P, Jin, C, Karlsson, NG, Wilson, IBH and Paschinger, K (2018) Isomeric separation and recognition of anionic and zwitterionic N-glycans from royal jelly glycoproteins. Molecular & Cellular Proteomics 17, 21772196.Google Scholar
Ichimiya, T, Maeda, M, Sakamura, S, Kanazawa, M, Nishihara, S and Kimura, Y (2015) Identification of β1,3-galactosyltransferases responsible for biosynthesis of insect complex-type N-glycans containing a T-antigen unit in the honeybee. Glycoconjugate Journal 32, 141151.Google Scholar
Ioffe, E and Stanley, P (1994) Mice lacking N-acetylglucosaminyltransferase I activity die at mid-gestation, revealing an essential role for complex or hybrid N-linked carbohydrates. Proceedings of the National Academy of Sciences of the United States of America 91, 728732.Google Scholar
Jang-Lee, J, Curwen, RS, Ashton, PD, Tissot, B, Mathieson, W, Panico, M, Dell, A, Wilson, RA and Haslam, SM (2007) Glycomics analysis of Schistosoma mansoni egg and cercarial secretions. Molecular & Cellular Proteomics 6, 14851499.Google Scholar
Jankowska, E, Parsons, LM, Song, X, Smith, DF, Cummings, RD and Cipollo, JF (2018) A comprehensive Caenorhabditis elegans N-glycan shotgun array. Glycobiology 28, 223232.Google Scholar
Jiménez-Castells, C, Vanbeselaere, J, Kohlhuber, S, Ruttkowski, B, Joachim, A and Paschinger, K (2017) Gender and developmental specific N-glycomes of the porcine parasite Oesophagostomum dentatum. Biochimica et Biophysica Acta 1861, 418430.Google Scholar
Kajiura, H, Hamaguchi, Y, Mizushima, H, Misaki, R and Fujiyama, K (2015) Sialylation potentials of the silkworm, Bombyx mori; b. mori possesses an active α2,6-sialyltransferase. Glycobiology 25, 14411453.Google Scholar
Kang, S, Cummings, RD and McCall, JW (1993) Characterization of the N-linked oligosaccharides in glycoproteins synthesized by microfilariae of Dirofilaria immitis. Journal of Parasitology 79, 815828.Google Scholar
Kawar, Z, van Die, I and Cummings, RD (2002) Molecular cloning and enzymatic characterisation of a UDP-GalNAc:GlcNAcb-R β1,4-N-acetylgalactosaminyltransferase from Caenorhabditis elegans. Journal of Biological Chemistry 277, 3492434932.Google Scholar
Khoo, K-H, Chatterjee, D, Caulfield, JP, Morris, HR and Dell, A (1997) Structural mapping of the glycans from the egg glycoproteins of Schistosoma mansoni and Schistosoma japonicum: identification of novel core structures and terminal sequences. Glycobiology 7, 663677.Google Scholar
Kim, B-T, Tsuchida, K, Lincecum, J, Kitagawa, H, Bernfield, M and Sugahara, K (2003) Identification and characterization of three Drosophila melanogaster glucuronyltransferases responsible for the synthesis of the conserved glycosaminoglycan-protein linkage region of proteoglycans. Two novel homologs exhibit broad specificity toward oligosaccharides from proteoglycans, glycoproteins, and glycosphingolipids. Journal of Biological Chemistry 278, 91169124.Google Scholar
Koles, K, Irvine, KD and Panin, VM (2004) Functional characterization of Drosophila sialyltransferase. Journal of Biological Chemistry 279, 43464357.Google Scholar
Kornfeld, R and Kornfeld, S (1985) Assembly of asparagine-linked oligosaccharides. Annual Review of Biochemistry 54, 631664.Google Scholar
Kubelka, V, Altmann, F, Staudacher, E, Tretter, V, März, L, Hård, K, Kamerling, JP and Vliegenthart, JFG (1993) Primary structures of the N-linked carbohydrate chains from honeybee venom phospholipase A2. European Journal of Biochemistry 213, 11931204.Google Scholar
Kurz, S, Jin, C, Hykollari, A, Gregorich, D, Giomarelli, B, Vasta, GR, Wilson, IBH and Paschinger, K (2013) Haemocytes and plasma of the eastern oyster (Crassostrea virginica) display a diverse repertoire of sulphated and blood group A-modified N-glycans. Journal of Biological Chemistry 288, 2441024428.Google Scholar
Kurz, S, Aoki, K, Jin, C, Karlsson, NG, Tiemeyer, M, Wilson, IBH and Paschinger, K (2015) Targetted release and fractionation reveal glucuronylated and sulphated N- and O-glycans in larvae of dipteran insects. Journal of Proteomics 126, 172188.Google Scholar
Kurz, S, King, JG, Dinglasan, RR, Paschinger, K and Wilson, IBH (2016) The fucomic potential of mosquitoes: fucosylated N-glycan epitopes and their cognate fucosyltransferases. Insect Biochemistry and Molecular Biology 68, 5263.Google Scholar
Lee, JJ, Dissanayake, S, Panico, M, Morris, HR, Dell, A and Haslam, SM (2005) Mass spectrometric characterisation of Taenia crassiceps metacestode N-glycans. Molecular and Biochemical Parasitology 143, 245249.Google Scholar
Lehr, T, Frank, S, Natsuka, S, Geyer, H, Beuerlein, K, Doenhoff, MJ, Hase, S and Geyer, R (2010) N-Glycosylation patterns of hemolymph glycoproteins from Biomphalaria glabrata strains expressing different susceptibility to Schistosoma mansoni infection. Experimental Parasitology 126, 592602.Google Scholar
Lochnit, G, Dennis, RD, Ulmer, AJ and Geyer, R (1998) Structural elucidation and monokine-inducing activity of two biologically active zwitterionic glycosphingolipids derived from the porcine parasitic nematode Ascaris suum. Journal of Biological Chemistry 273, 466474.Google Scholar
Loke, I, Ostergaard, O, Heegaard, NHH, Packer, NH and Thaysen-Andersen, M (2017) Paucimannose-rich N-glycosylation of spatiotemporally regulated human neutrophil elastase modulates Its immune functions. Molecular & Cellular Proteomics 16, 15071527.Google Scholar
Luyai, AE, Heimburg-Molinaro, J, Prasanphanich, NS, Mickum, ML, Lasanajak, Y, Song, X, Nyame, AK, Wilkins, P, Rivera-Marrero, CA, Smith, DF, Van Die, I, Secor, WE and Cummings, RD (2014) Differential expression of anti-glycan antibodies in schistosome-infected humans, rhesus monkeys and mice. Glycobiology 24, 602618.Google Scholar
Maes, E, Garenaux, E, Strecker, G, Leroy, Y, Wieruszeski, JM, Brassart, C and Guerardel, Y (2005) Major O-glycans from the nest of Vespula germanica contain phospho-ethanolamine. Carbohydrate Research 340, 18521858.Google Scholar
Martini, F, Eckmair, B, Neupert, C, Štefanić, S, Jin, C, Garg, M, Jiménez-Castells, C, Hykollari, A, Yan, S, Venco, L, Varón Silva, D, Wilson, IBH and Paschinger, K (2019) Highly modified and immunoactive N-glycans of the canine heartworm. Nature Communications 10, 75.Google Scholar
McVeigh, P and Maule, AG (2019) Can CRISPR help in the fight against parasitic worms? Elife 8, e44382. doi: 10.7554/eLife.44382.Google Scholar
McVeigh, P, Cwiklinski, K, Garcia-Campos, A, Mulcahy, G, O'Neill, SM, Maule, AG and Dalton, JP (2018) In silico analyses of protein glycosylating genes in the helminth Fasciola hepatica (liver fluke) predict protein-linked glycan simplicity and reveal temporally-dynamic expression profiles. Scientific Reports 8, 11700.Google Scholar
Meevissen, MH, Driessen, NN, Smits, HH, Versteegh, R, van Vliet, SJ, van Kooyk, Y, Schramm, G, Deelder, AM, Haas, H, Yazdanbakhsh, M and Hokke, CH (2012) Specific glycan elements determine differential binding of individual egg glycoproteins of the human parasite Schistosoma mansoni by host C-type lectin receptors. International Journal for Parasitology 42, 269277.Google Scholar
Metzler, M, Gertz, A, Sarkar, M, Schachter, H, Schrader, JW and Marth, JD (1994) Complex asparagine-linked oligosaccharides are required for morphogenic events during post-implantation development. The EMBO Journal 13, 20562065.Google Scholar
Mickum, ML, Prasanphanich, NS, Song, X, Dorabawila, N, Mandalasi, M, Lasanajak, Y, Luyai, A, Secor, WE, Wilkins, PP, Van Die, I, Smith, DF, Nyame, AK, Cummings, RD and Rivera-Marrero, CA (2016a) Identification of antigenic glycans from Schistosoma mansoni by using a shotgun egg glycan microarray. Infection and Immunity 84, 13711386.Google Scholar
Mickum, ML, Rojsajjakul, T, Yu, Y and Cummings, RD (2016b) Schistosoma mansoni α1,3-fucosyltransferase-F generates the Lewis X antigen. Glycobiology 26, 270285.Google Scholar
Miyata, S, Sato, C, Kumita, H, Toriyama, M, Vacquier, VD and Kitajima, K (2006) Flagellasialin: a novel sulfated α2,9-linked polysialic acid glycoprotein of sea urchin sperm flagella. Glycobiology 16, 12291241.Google Scholar
Morelle, W, Haslam, SM, Olivier, V, Appleton, JA, Morris, HR and Dell, A (2000) Phosphorylcholine-containing N-glycans of Trichinella spiralis: identification of multiantennary lacdiNAc structures. Glycobiology 10, 941950.Google Scholar
Mucha, J, Domlatil, J, Lochnit, G, Rendić, D, Paschinger, K, Hinterkörner, G, Hofinger, A, Kosma, P and Wilson, IBH (2004) The Drosophila melanogaster homologue of the humna histo-blood group Pk gene encodes a glycolipid-modifying α1,4-N-acetylgalactosaminyltransferase. Biochemical Journal 382, 6774.Google Scholar
Nyame, K, Smith, DF, Damian, RT and Cummings, RD (1989) Complex-type asparagine-linked oligosaccharides in glycoproteins synthesized by Schistosoma mansoni adult males contain terminal beta-linked N-acetylgalactosamine. Journal of Biological Chemistry 264, 32353243.Google Scholar
Palaima, E, Leymarie, N, Stroud, D, Mizanur, RM, Hodgkin, J, Gravato-Nobre, MJ, Costello, CE and Cipollo, JF (2010) The Caenorhabditis elegans bus-2 mutant reveals a new class of O-glycans affecting bacterial resistance. Journal of Biological Chemistry 285, 1766217672.Google Scholar
Paschinger, K and Wilson, IBH (2015) Two types of galactosylated fucose motifs are present on N-glycans of Haemonchus contortus. Glycobiology 25, 585590.Google Scholar
Paschinger, K and Wilson, IBH (2016) Analysis of zwitterionic and anionic N-linked glycans from invertebrates and protists by mass spectrometry. Glycoconjugate Journal 33, 273283.Google Scholar
Paschinger, K, Rendić, D, Lochnit, G, Jantsch, V and Wilson, IBH (2004) Molecular basis of anti-horseradish peroxidase staining in Caenorhabditis elegans. Journal of Biological Chemistry 279, 4958849598.Google Scholar
Paschinger, K, Staudacher, E, Stemmer, U, Fabini, G and Wilson, IBH (2005) Fucosyltransferase substrate specificity and the order of fucosylation in invertebrates. Glycobiology 15, 463474.Google Scholar
Paschinger, K, Hackl, M, Gutternigg, M, Kretschmer-Lubich, D, Stemmer, U, Jantsch, V, Lochnit, G and Wilson, IBH (2006) A deletion in the Golgi α-mannosidase II gene of Caenorhabditis elegans results in unexpected non-wild type N-glycan structures. Journal of Biological Chemistry 281, 2826528277.Google Scholar
Paschinger, K, Rendić, D and Wilson, IBH (2009) Revealing the anti-HRP epitope in Drosophila and Caenorhabditis. Glycoconjugate Journal 26, 385395.Google Scholar
Paschinger, K, Razzazi-Fazeli, E, Furukawa, K and Wilson, IBH (2011) Presence of galactosylated core fucose on N-glycans in the planaria Dugesia japonica. Journal of Mass Spectrometry 46, 561567.Google Scholar
Paschinger, K, Gonzalez-Sapienza, GG and Wilson, IBH (2012a) Mass spectrometric analysis of the immunodominant glycan epitope of Echinococcus granulosus antigen Ag5. International Journal for Parasitology 42, 279285.Google Scholar
Paschinger, K, Hykollari, A, Razzazi-Fazeli, E, Greenwell, P, Leitsch, D, Walochnik, J and Wilson, IBH (2012b) The N-glycans of Trichomonas vaginalis contain variable core and antennal modifications. Glycobiology 22, 300313.Google Scholar
Pineda, MA, Lumb, F, Harnett, MM and Harnett, W (2014) ES-62, a therapeutic anti-inflammatory agent evolved by the filarial nematode Acanthocheilonema viteae. Molecular and Biochemical Parasitology 194, 18.Google Scholar
Pinzon-Ortiz, C, Friedman, J, Esko, J and Sinnis, P (2001) The binding of the circumsporozoite protein to cell surface heparan sulfate proteoglycans is required for plasmodium sporozoite attachment to target cells. Journal of Biological Chemistry 276, 2678426791.Google Scholar
Pöltl, G, Kerner, D, Paschinger, K and Wilson, IBH (2007) N-Glycans of the porcine nematode parasite Ascaris suum are modified with phosphorylcholine and core fucose residues. The FEBS Journal 274, 714726.Google Scholar
Qian, Y, West, CM and Kornfeld, S (2010) UDP-GlcNAc:Glycoprotein N-acetylglucosamine-1-phosphotransferase mediates the initial step in the formation of the methylphosphomannosyl residues on the high mannose oligosaccharides of Dictyostelium discoideum glycoproteins. Biochemical and Biophysical Research Communications 393, 678681.Google Scholar
Ravida, A, Aldridge, AM, Driessen, NN, Heus, FA, Hokke, CH and O'Neill, SM (2016) Fasciola hepatica surface coat glycoproteins contain mannosylated and phosphorylated N-glycans and exhibit immune modulatory properties independent of the mannose receptor. PLOS Neglected Tropical Diseases 10, e0004601.Google Scholar
Rendić, D, Klaudiny, J, Stemmer, U, Schmidt, J, Paschinger, K and Wilson, IBH (2007) Towards abolition of immunogenic structures in insect cells: characterization of a honey-bee (Apis mellifera) multi-gene family reveals both an allergy-related core α1,3-fucosyltransferase and the first insect Lewis-histo-blood-group-related antigen-synthesizing enzyme. Biochemical Journal 402, 105115.Google Scholar
Sahadevan, S, Antonopoulos, A, Haslam, SM, Dell, A, Ramaswamy, S and Babu, P (2014) Unique, polyfucosylated glycan-receptor interactions are essential for regeneration of Hydra magnipapillata. ACS Chemical Biology 9, 147155.Google Scholar
Samuelson, J, Banerjee, S, Magnelli, P, Cui, J, Kelleher, DJ, Gilmore, R and Robbins, PW (2005) The diversity of dolichol-linked precursors to Asn-linked glycans likely results from secondary loss of sets of glycosyltransferases. Proceedings of the National Academy of Sciences of the United States of America 102, 15481553.Google Scholar
Sarkar, M, Iliadi, KG, Leventis, PA, Schachter, H and Boulianne, GL (2010) Neuronal expression of Mgat1 rescues the shortened life span of Drosophila mgat11 null mutants and increases life span. Proceedings of the National Academy of Sciences of the United States of America 107, 96779682.Google Scholar
Schachter, H (1986) Biosynthetic controls that determine the branching and microheterogeneity of protein-bound oligosaccharides. Biochemistry and Cell Biology 64, 163181.Google Scholar
Schachter, H (2010) Mgat1-dependent N-glycans are essential for the normal development of both vertebrate and invertebrate metazoans. Seminars in Cell and Developmental Biology 21, 609615.Google Scholar
Schiller, B, Hykollari, A, Yan, S, Paschinger, K and Wilson, IBH (2012) Complicated N-linked glycans in simple organisms. Biological Chemistry. (Hoppe Seyler) 393, 661673.Google Scholar
Schubert, M, Bleuler-Martinez, S, Butschi, A, Walti, MA, Egloff, P, Stutz, K, Yan, S, Wilson, IBH, Hengartner, MO, Aebi, M, Allain, FH and Künzler, M (2012) Plasticity of the β-trefoil protein fold in the recognition and control of invertebrate predators and parasites by a fungal defence system. PLOS Pathogens 8, e1002706.Google Scholar
Seppo, A, Moreland, M, Schweingruber, H and Tiemeyer, M (2000) Zwitterionic and acidic glycosphingolipids of the Drosophila melanogaster embryo. European Journal of Biochemistry 267, 35493558.Google Scholar
Shi, H, Tan, J and Schachter, H (2006) N-glycans are involved in the response of Caenorhabditis elegans to bacterial pathogens. Methods in Enzymology 417, 359389.Google Scholar
Smit, CH, van Diepen, A, Nguyen, DL, Wuhrer, M, Hoffmann, KF, Deelder, AM and Hokke, CH (2015) Glycomic analysis of life stages of the human parasite Schistosoma mansoni reveals developmental expression profiles of functional and antigenic glycan motifs. Molecular & Cellular Proteomics 14, 17501769.Google Scholar
Stanton, R, Hykollari, A, Eckmair, B, Malzl, D, Dragosits, M, Palmberger, D, Wang, P, Wilson, IBH and Paschinger, K (2017) The underestimated N-glycomes of lepidopteran species. Biochimica et Biophysica Acta 1861, 699714.Google Scholar
Subramanian, SP, Babu, P, Palakodeti, D and Subramanian, R (2018) Identification of multiple isomeric core chitobiose-modified high-mannose and paucimannose N-glycans in the planarian Schmidtea mediterranea. Journal of Biological Chemistry 293, 67076720.Google Scholar
Sugita, M, Fujii, H, Inagaki, F, Suzuki, M, Hayata, C and Hori, T (1992) Polar glycosphingolipids in annelida. A novel series of glycosphingolipids containing choline phosphate from the earthworm, Pheretima hilgendorf. Journal of Biological Chemistry 267, 2259522598.Google Scholar
Sutov, G (2016) Glycomic studies of parasitic nematodes (PhD thesis). Imperial College, London.Google Scholar
Takahashi, N, Masuda, K, Hiraki, K, Yoshihara, K, Huang, H-H, Khoo, K-H and Kato, K (2004) N-glycan structures of squid rhodopsin. Existence of the α1-3 and α1-6 difucosylated innermost GlcNAc residue in a molluscan glycoprotein. European Journal of Biochemistry 270, 26272632.Google Scholar
Talabnin, K, Aoki, K, Saichua, P, Wongkham, S, Kaewkes, S, Boons, GJ, Sripa, B and Tiemeyer, M (2013) Stage-specific expression and antigenicity of glycoprotein glycans isolated from the human liver fluke, Opisthorchis viverrini. International Journal for Parasitology 43, 3750.Google Scholar
Thaysen-Andersen, M and Packer, NH (2012) Site-specific glycoproteomics confirms that protein structure dictates formation of N-glycan type, core fucosylation and branching. Glycobiology 22, 14401452.Google Scholar
Titz, A, Butschi, A, Henrissat, B, Fan, YY, Hennet, T, Razzazi-Fazeli, E, Hengartner, MO, Wilson, IBH, Künzler, M and Aebi, M (2009) Molecular basis for galactosylation of core fucose residues in invertebrates: identification of Caenorhabditis elegans N-glycan core α1,6-fucoside β1,4-galactosyltransferase GALT-1 as a member of a novel glycosyltransferase family. Journal of Biological Chemistry 284, 3622336233.Google Scholar
Urai, M, Nakamura, T, Uzawa, J, Baba, T, Taniguchi, K, Seki, H and Ushida, K (2009) Structural analysis of O-glycans of mucin from jellyfish (Aurelia aurita) containing 2-aminoethylphosphonate. Carbohydrate Research 344, 21822187.Google Scholar
Vadaie, N and Jarvis, DL (2004) Molecular cloning and functional characterization of a Lepidopteran insect β4-N-acetylgalactosaminyltransferase with broad substrate specificity, a functional role in glycoprotein biosynthesis, and a potential functional role in glycolipid biosynthesis. Journal of Biological Chemistry 279, 3350133508.Google Scholar
van Die, I, Gomord, V, Kooyman, FNJ, van der Berg, TK, Cummings, RD and Vervelde, L (1999) Core α1→3-fucose is a common modification of N-glycans in parasitic helminths and constitutes an important epitope for IgE from Haemonchus contortus infected sheep. FEBS Letters 463, 189193.Google Scholar
van Diepen, A, Smit, CH, van Egmond, L, Kabatereine, NB, Pinot de Moira, A, Dunne, DW and Hokke, CH (2012) Differential anti-glycan antibody responses in Schistosoma mansoni-infected children and adults studied by shotgun glycan microarray. PLOS Neglected Tropical Diseases 6, e1922.Google Scholar
van Diepen, A, van der Plas, AJ, Kozak, RP, Royle, L, Dunne, DW and Hokke, CH (2015) Development of a Schistosoma mansoni shotgun O-glycan microarray and application to the discovery of new antigenic schistosome glycan motifs. International Journal for Parasitology 45, 465475.Google Scholar
van Kuik, JA, Sijbesma, RP, Kamerling, JP, Vliegenthart, JFG and Wood, EJ (1986) Primary structure of a low-molecular-mass N-linked oligosaccharide from hemocyanin of Lymnaea stagnalis. 3-O-methyl-D-mannose as a constituent of the xylose-containing core structure in an animal glycoprotein. European Journal of Biochemistry 160, 621625.Google Scholar
van Kuik, JA, Breg, J, Kolsteeg, CEM, Kamerling, JP and Vliegenthart, JFG (1987a) Primary structure of the acidic carbohydrate chain of hemocyanin from Panulirus interruptus. FEBS Letters 221, 150154.Google Scholar
Van Kuik, JA, Sijbesma, RP, Kamerling, JP, Vliegenthart, JF and Wood, EJ (1987b) Primary structure determination of seven novel N-linked carbohydrate chains derived from hemocyanin of Lymnaea stagnalis. 3-O-methyl-D-galactose and N-acetyl-D-galactosamine as constituents of xylose-containing N-linked oligosaccharides in an animal glycoprotein. European Journal of Biochemistry 169, 399411.Google Scholar
van Remoortere, A, Hokke, CH, van Dam, GJ, van Die, I, Deelder, AM and van den Eijnden, DH (2000) Various stages of schistosoma express Lewis(x), LacdiNAc, GalNAcβ1-4 (Fucα1-3)GlcNAc and GalNAcβ1-4(Fucα1-2Fucα1-3)GlcNAc carbohydrate epitopes: detection with monoclonal antibodies that are characterized by enzymatically synthesized neoglycoproteins. Glycobiology 10, 601609.Google Scholar
van Stijn, CM, van den Broek, M, Vervelde, L, Alvarez, RA, Cummings, RD, Tefsen, B and van Die, I (2009) Vaccination-induced IgG response to Galα1-3GalNAc glycan epitopes in lambs protected against Haemonchus contortus challenge infection. International Journal for Parasitology 40, 215222.Google Scholar
Vanbeselaere, J, Yan, S, Joachim, A, Paschinger, K and Wilson, IBH (2018) The parasitic nematode Oesophagostomum dentatum synthesizes unusual glycosaminoglycan-like O-glycans. Glycobiology 28, 474481.Google Scholar
Varki, A (2011) Evolutionary forces shaping the Golgi glycosylation machinery: why cell surface glycans are universal to living cells. Cold Spring Harbor Perspectives in Biology 3, a005462.Google Scholar
Warren, CE, Krizius, A, Roy, PJ, Culotti, JG and Dennis, JW (2002) The C. elegans gene, gly-2, can rescue the N-acetylglucosaminyltransferase V mutation of Lec4 cells. Journal of Biological Chemistry 277, 2282922838.Google Scholar
Wilbers, RH, Westerhof, LB, van Noort, K, Obieglo, K, Driessen, NN, Everts, B, Gringhuis, SI, Schramm, G, Goverse, A, Smant, G, Bakker, J, Smits, HH, Yazdanbakhsh, M, Schots, A and Hokke, CH (2017) Production and glyco-engineering of immunomodulatory helminth glycoproteins in plants. Scientific Reports 7, 45910.Google Scholar
Williams, PJ, Wormald, MR, Dwek, RA, Rademacher, TW, Parker, GF and Roberts, DR (1991) Characterisation of oligosaccharides from Drosophila melanogaster glycoproteins. Biochimica et Biophysica Acta 1075, 146153.Google Scholar
Wilson, IBH (2012) The class I α1,2-mannosidases of Caenorhabditis elegans. Glycoconjugate Journal 29, 173179.Google Scholar
Wilson, IBH and Paschinger, K (2016) Sweet secrets of a therapeutic worm: mass spectrometric N-glycomic analysis of Trichuris suis. Analytical and Bioanalytical Chemistry 408, 461471.Google Scholar
Wohlschlager, T, Butschi, A, Grassi, P, Sutov, G, Gauss, R, Hauck, D, Schmieder, SS, Knobel, M, Titz, A, Dell, A, Haslam, SM, Hengartner, MO, Aebi, M and Künzler, M (2014) Methylated glycans as conserved targets of animal and fungal innate defense. Proceedings of the National Academy of Sciences of the United States of America 111, E2787E2796.Google Scholar
Wuhrer, M, Robijn, ML, Koeleman, CA, Balog, CI, Geyer, R, Deelder, AM and Hokke, CH (2004) A novel Gal(β1-4)Gal(β1-4)Fuc(α1-6)-core modification attached to the proximal N-acetylglucosamine of keyhole limpet haemocyanin (KLH) N-glycans. Biochemical Journal 378, 625632.Google Scholar
Wuhrer, M, Koeleman, CA, Deelder, AM and Hokke, CH (2006) Repeats of LacdiNAc and fucosylated LacdiNAc on N-glycans of the human parasite Schistosoma mansoni. The FEBS Journal 273, 347361.Google Scholar
Yan, S, Serna, S, Reichardt, NC, Paschinger, K and Wilson, IBH (2013) Array-assisted characterization of a fucosyltransferase required for the biosynthesis of complex core modifications of nematode N-glycans. Journal of Biological Chemistry 288, 2101521028.Google Scholar
Yan, S, Brecker, L, Jin, C, Titz, A, Dragosits, M, Karlsson, N, Jantsch, V, Wilson, IBH and Paschinger, K (2015 a) Bisecting galactose as a feature of N-glycans of wild-type and mutant Caenorhabditis elegans. Molecular & Cellular Proteomics 14, 21112125.Google Scholar
Yan, S, Jin, C, Wilson, IBH and Paschinger, K (2015 b) Comparisons of Caenorhabditis fucosyltransferase mutants reveal a multiplicity of isomeric N-glycan structures. Journal of Proteome Research 14, 52915305.Google Scholar
Yan, S, Wilson, IBH and Paschinger, K (2015 c) Comparison of RP-HPLC modes to analyse the N-glycome of the free-living nematode Pristionchus pacificus. Electrophoresis 36, 13141329.Google Scholar
Yan, S, Vanbeselaere, J, Jin, C, Blaukopf, M, Wols, F, Wilson, IBH and Paschinger, K (2018 a) Core richness of N-glycans of Caenorhabditis elegans: a case study on chemical and enzymatic release. Analytical Chemistry 90, 928935.Google Scholar
Yan, S, Wang, H, Schachter, H, Jin, C, Wilson, IBH and Paschinger, K (2018b) Ablation of N-acetylglucosaminyltransferases in Caenorhabditis induces expression of unusual intersected and bisected N-glycans. Biochimica et Biophysica Acta 1862, 21912203.Google Scholar
Yang, YY, Li, XH, Brzezicka, K, Reichardt, NC, Wilson, RA, van Diepen, A and Hokke, CH (2017) Specific anti-glycan antibodies are sustained during and after parasite clearance in Schistosoma japonicum-infected rhesus macaques. PLOS Neglected Tropical Diseases 11, e0005339.Google Scholar
Yang, YYM, van Diepen, A, Brzezicka, K, Reichardt, NC and Hokke, CH (2018) Glycan microarray-assisted identification of IgG subclass targets in schistosomiasis. Front Immunol 9, 2331.Google Scholar
Zhou, H, Hanneman, AJ, Chasteen, ND and Reinhold, VN (2013) Anomalous N-glycan structures with an internal fucose branched to GlcA and GlcN residues isolated from a mollusk shell-forming fluid. Journal of Proteome Research 12, 45474555.Google Scholar
Figure 0

Fig. 1. Simplified biosynthetic scheme for N-linked glycans in animals. Starting with the Glc3Man9GlcNAc2 precursor, various glycosidases result in different isomers of oligomannosidic glycans with the maximal degree of processing by class I mannosidases yielding Man5GlcNAc2. This is the substrate for N-acetylglucosaminyltransferase I (GlcNAc-TI) which generates a ‘hybrid’ structure which can be further modified by the action of Golgi mannosidase II, GlcNAc-TII and Golgi hexosaminidase. The maximum number of antennae (three or four) depends on the presence of GlcNAc-TIV and GlcNAc-TV; example hybrid, pseudohybrid, paucimannosidic and tri-/tetra-antennary glycans are shown as known from various model, host, vector or parasitic invertebrates. For simplicity, fucosylation and other modifications are not included. Glycans are depicted according to the Standard Nomenclature for Glycans (see also box).

Figure 1

Fig. 2. Example N-glycans from invertebrates. Structures are depicted from either parasitic or free-living organisms, whereby some of the latter are hosts or vectors for parasites. Some types of structures are species- or class-specific, but others are found in more than one phylum. Only a non-exhaustive selection of core and antennal epitopes is shown in the inset: core difucosylation, core ‘GalFuc’, Lewis X (LeX), fucosylated and non-fucosylated LacdiNAc (LDN) and blood group A (BGA). (A) The bisecting and distal core modifications found in the free-living C. elegans are indicated by pink boxes; (B) free-living C. elegans, the necromenic P. pacificus and the parasites H. contortus, H. polygyrus and O. dentatum express di- and/or tri-fucosylated cores with species-specific galactosylation and methylation; (C) varying antennal modifications are found in all nematodes as well as the cestodes E. granulosus and T. crassiceps, (D) while filarial species have up to four long antennae including D. immitis, which has in addition glucuronylated structures; (E) galactosylated core fucose (GalFuc) is found in many invertebrates, sometimes in substituted form; (F, G and H) selected complex glycans from larvae of different insect phyla; (I) selected S. mansoni N-glycan modifications which are partly stage-specific; (J, K and L) selected gastropod and bivalve glycans, including those of Crassostrea virginica, B. glabrata, Volvarina rubella and Mytilus edulis. Note that some modifications, such as core β-mannosylation, are at low abundance in the relevant glycomes. Glycans are depicted according to the Standard Nomenclature for Glycans; undefined hexoses/N-acetylhexosamines are shown as white circles/squares. Me, methyl; MAEP, N-methyl-aminoethylphosphonate; PC, phosphorylcholine; PE, phosphoethanolamine (2-aminoethylphosphate); S, sulphate. Broken lines,±or brackets indicate structure-, species- or stage-dependent variations in these elements.