Next Article in Journal
FOXA1: A Pioneer of Nuclear Receptor Action in Breast Cancer
Next Article in Special Issue
A Recurrent STAT5BN642H Driver Mutation in Feline Alimentary T Cell Lymphoma
Previous Article in Journal
Breast Cancer Drug Approvals Issued by EMA: A Review of Clinical Trials
Previous Article in Special Issue
The Cellular Prion Protein and the Hallmarks of Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Re-Expression of Poly/Oligo-Sialylated Adhesion Molecules on the Surface of Tumor Cells Disrupts Their Interaction with Immune-Effector Cells and Contributes to Pathophysiological Immune Escape

1
German Cancer Research Center, Toxicology and Chemotherapy Unit Heidelberg, 69120 Heidelberg, Germany
2
Department of Hematology, Faculty of Medicine, Tabriz University of Medical Sciences, Tabriz 5165665931, Iran
3
Stem Cells and Regenerative Medicine Unit at King Fahad Medical Research Centre, Jeddah 11211, Saudi Arabia
4
Cancer Gene Therapy Research Center (CGRC), Zanjan University of Medical Sciences, Zanjan 4513956184, Iran
5
Department of Biotechnology, Inland Norway University of Applied Sciences, 2418 Hamar, Norway
*
Author to whom correspondence should be addressed.
Cancers 2021, 13(20), 5203; https://doi.org/10.3390/cancers13205203
Submission received: 7 September 2021 / Revised: 11 October 2021 / Accepted: 12 October 2021 / Published: 16 October 2021
(This article belongs to the Special Issue Feature Paper from Journal Reviewers)

Abstract

:

Simple Summary

The immune system consists of various mechanisms contributing to the battle against cancer cells or hazardous pathogens. However, in cancer progression the immune system is often unable to eliminate neoplastic cells, although immune effector cells infiltrate the tumor environment. The current paper reviews the causes for this immune escape. Specifically, we comprehensively discuss various roles of sialic acids in this process. Specific focus is given to adhesion molecules re-expressed on membranes of tumor cells, which carry oligo- and polysialic acid chains. These carrier proteins loaded with sialic acids direct the interaction between immune effector and tumor cells and thus prevent the “kiss of death” between the latter and the former cells. We also discuss strategies suited to reduce the degree of sialic acid presence on the surface of tumor cells, which can be the basis for future therapeutic intervention.

Abstract

Glycans linked to surface proteins are the most complex biological macromolecules that play an active role in various cellular mechanisms. This diversity is the basis of cell–cell interaction and communication, cell growth, cell migration, as well as co-stimulatory or inhibitory signaling. Our review describes the importance of neuraminic acid and its derivatives as recognition elements, which are located at the outermost positions of carbohydrate chains linked to specific glycoproteins or glycolipids. Tumor cells, especially from solid tumors, mask themselves by re-expression of hypersialylated neural cell adhesion molecule (NCAM), neuropilin-2 (NRP-2), or synaptic cell adhesion molecule 1 (SynCAM 1) in order to protect themselves against the cytotoxic attack of the also highly sialylated immune effector cells. More particularly, we focus on α-2,8-linked polysialic acid chains, which characterize carrier glycoproteins such as NCAM, NRP-2, or SynCam-1. This characteristic property correlates with an aggressive clinical phenotype and endows them with multiple roles in biological processes that underlie all steps of cancer progression, including regulation of cell–cell and/or cell–extracellular matrix interactions, as well as increased proliferation, migration, reduced apoptosis rate of tumor cells, angiogenesis, and metastasis. Specifically, re-expression of poly/oligo-sialylated adhesion molecules on the surface of tumor cells disrupts their interaction with immune-effector cells and contributes to pathophysiological immune escape. Further, sialylated glycoproteins induce immunoregulatory cytokines and growth factors through interactions with sialic acid-binding immunoglobulin-like lectins. We describe the processes, which modulate the interaction between sialylated carrier glycoproteins and their ligands, and illustrate that sialic acids could be targets of novel therapeutic strategies for treatment of cancer and immune diseases.

Graphical Abstract

1. Introduction

In cancer progression, the immune system is often not capable of eliminating cancer cells, despite the presence of immune effector cells infiltrating the tumor microenvironment [1,2,3]. In this regard it is important to realize the reciprocal effects between cancer and immune cells, which are modeled to a large extend by carbohydrates present on membrane glycoproteins. Due to their high variability, they orchestrate not only the cellular regulation, but are also involved in processes, such as protein folding, intracellular transport, and immune cell polarization. Furthermore, glycoproteins play a key role in activating the immune system [4,5,6]. However, the poor prognosis of metastatic cancers has been correlated with overexpression of membrane glycoproteins on cancer cells, which are (poly-) sialylated [7,8,9]. In this respect, recent advances in glycobiology and cancer research have described a mechanistic role of sialic acid in tumor development and progression: aberrant sialylation of glycoproteins and glycolipids has been shown to mediate conditions such as increased tumor growth [10,11], inhibition of apoptosis [12,13], metastasis [14,15,16], resistance to therapy [17,18,19], and enhanced invasiveness of tumor cells [11,20]. Specifically, increased sialylation of metastasizing tumor cells [9,21] leads to altered adhesion and changes in transmembrane signaling [22,23]. Therefore, sialic acid has been repeatedly proposed as a possible therapeutic target against tumors [11,20,24,25].

2. Materials and Methods

2.1. Literature Search

This systematic review was carried out according to the Systematic Reviews and Meta-Analyzes (PRISMA) guidelines in order to define the pathophysiological roles of poly/oligo- sialylated adhesion molecules and their co-partners. As indicated in the PRISMA FLOW DIAGRAM, we did an extensive search of the PubMed database (Scheme 1). A complete list of the keywords that have been used to search for items are presented in the form of 17 complementary search boxes in File S1. The database search was done from 1995 to 2021. The search results were imported into EndNote and the duplicate entries removed based on title and year of publication. (Registration ID 277798).

2.2. Data Selection

We have included into the review all items that meet the following criteria:
(1)
The physiological and pathological roles of membrane adhesion molecules linked with oligo/poly-sialic acid glycosylation in tumor progression, apoptosis, metastasis, angiogenesis, migration, proliferation, and growth of tumors;
(2)
The biological roles of heterophilic and homophilic membrane adhesion molecules in neuronal and embryonic development and the development of certain neuronal diseases (such as Alzheimer's disease, Parkinson's disease, multiple sclerosis, and schizophrenia);
(3)
The role of sialic acid in the immune escape of tumors and pathogens (bacteria, viruses);
(4)
The role of sialic acid in electrostatic repulsion between immune effector and target cells (tumor or pathogens), which re-express membrane adhesion molecules linked with oligo/polysialic acid;
(5)
The role of sialic acid in differentiation of immune cells (T-cells), and in virus infection;
(6)
The role of sialic acid receptors/copartners (lectins such as siglecs or other adhesion molecules) in relation to the function of the immune cells;
(7)
The role of galectins, selectins, and kinases in cooperation with sialylated glycoproteins in or outside of cells.
Entries such as non-English articles, review articles, book chapters, case reports, letters to the editor, and responses were excluded from the study.

2.3. Legend to Prisma Flow Diagram

Records were identified by searching the PubMed database for keywords within the period of 1995–2021, which have been listed in the respective supplementary file (File S1). The original number of records was reduced by subtracting duplicates, which was followed by selecting those records that focus on the role of sialic acids (especially of poly- and oligo sialic acids) in various biological processes. The final selection was made by concentrating on those records that describe the role of adhesion molecules decorated with poly- and oligo sialic acids that play an important function in embryonal and neuronal development and can be re-expressed on tumor cells for immune escape.

3. Results and Discussion

3.1. Structure and Regulation of Sialic (N-Acetylneuraminic)-Acid

Membrane proteins are post-translationally modified by N- or O-glycosylation for improving their molecular stability, for receptor-ligand interactions and ensuing signal transduction [14,26]. Glycoprotein- and ganglioside modifications are differentiated by their terminal glycan structures, e.g., sialic acid- or heparan sulfate- derivatives, which have differential effects for most physiological and pathophysiological functions [27,28,29]. Almost all eukaryotic organisms are able to express sialic acids. This is made possible by sialyltransferases of the glycosyltransferase family 29 (CAZY GT_29), which are widespread in the Deuterostoma lineages and more rarely described in Protostoma, Viridiplantae, and various protist lineages, which have a common ancestor with eukaryotes. In fact, even some pathogenic microorganisms also contain sialic acids, e.g., bacteria, viruses, and parasites; they utilize cell surface sialic acids as ligands to attach to corresponding cell surface lectins, and to infect respective cells [30,31,32,33]. (Some Gram-negative bacteria synthesize sialic acids via an aldolase enzyme: they use a mannose derivative (Man-NAc) as a substrate and attach the three carbons of pyruvate to it, thus resulting in a sialic acid structure [34].) Sialic acid is a key component of glycoconjugates in glycoproteins/glycolipids, which commonly occur in cell membranes, cytoplasm, and glandular secretions that mediate cellular communication [24,35]. The position of sialic acid at the most distal part of non-reducing glycan structures qualifies this α-keto acid to play important roles in biological processes. Sialic acid formation involves more than 20 different sialyltransferases, which transfer sialic acid residues to glycoproteins [36] (Figure 1A,B).
The most common sialic acids are N-acetylneuraminic acid (Neu5Ac), N-glycolylneuraminic acid (Neu5Gc) [37,38], and the deaminated sialic acid derivative 2-keto-3-deoxy-D-glycero-D-galacto-nononic acid (KDN) [13,37,39], in which the fifth carbon atom is substituted by acetoamide/hydroxyacetoamide/hydroxyl groups [40]. Sialic acids (Neu5Ac) may occur in glyco-conjugates, linked as mono sialyl residues via α-2,3 or α-2,6 to galactose or N-acetyl galactosamine at the non-reducing terminal position of glycan chains on glycoproteins and glycolipids, whereas oligo- or polymeric forms of sialylation (oligoSia or polySia) render α-2,8 or α-2,9 glycosidic linkages [5,24,40,41,42]. Their structure and function depend on the linkage type between sialic acid residues and possible substitutions of sulfate/hydroxyl-groups in positions C4/7/8/9 by acetyl/lactoyl/methyl/sulfonyl or phosphonyl groups (Figure 1A,B).
Figure 1. Type of sialic acid linkages and their relation to various carrier molecules. (A): The structure of sialic acid (right), which is linked to galactose at positions α2,3 or α2,6, not only protect glycoproteins from uptake and degradation, but also prevent recognition of subterminal Gal and GalNAc residues by asialoglycoprotein receptors [14]. The sialic acid α2,3 linkage can be extended to a sialic acid chain by adding further sialic acid residues to position α2,8. The transfer of sialic acid residues involves more than 20 different sialyltransferases, the most important subgroup being C8 sialyltransferases (ST8Sia I–VI), which transfer a polysialic acid (PSA) chain via an α2,8-linkage to proteins. Sialic acid (Neu5Ac) may occur in glyco-conjugates in α2,3-Sia, α2,6-Sia, α2,8-oligosialylated (3–7), and α2,8-polysialylated (8–200) forms. Various groups, which are attached to sialic acid, are designated as R1–R9 (left). Most common sialic acids are N-acetylneuraminic acid (Neu5Ac)/N-glycolylneuraminic acid (Neu5Gc)/3-deoxy-non-2-ulosonic acid (KDN) [13,37], in which the fifth. carbon atom is substituted by acetoamide/hydroxyacetoamide/hydroxyl groups [40]. Their structure and function depend on the linkage type between sialic acid residues and possible substitutions of sulfate/hydroxyl-groups in positions C4/7/8/9 by acetyl/lactoyl/methyl/sulfonyl or phosphonyl groups. (B): Polysialic acid chains (with α2,8 linkages) can be attached to different carriers including O- or N-glycans, GPI anchored proteins, oligosphinglolipids, and phospholipids. The functionally most important polysialic acid chains have a α-helical conformation with dynamic function by their carboxyl groups being directed inwardly and the N-acetyl groups (PolyNeu5Ac and PolyNeu5Gc) outwardly. (C): The consequences of altered glycosylations are shown in part (C). The scheme indicates that tumor cells with altered glycosylation will show a series of distinctly changed biologic processes.
Figure 1. Type of sialic acid linkages and their relation to various carrier molecules. (A): The structure of sialic acid (right), which is linked to galactose at positions α2,3 or α2,6, not only protect glycoproteins from uptake and degradation, but also prevent recognition of subterminal Gal and GalNAc residues by asialoglycoprotein receptors [14]. The sialic acid α2,3 linkage can be extended to a sialic acid chain by adding further sialic acid residues to position α2,8. The transfer of sialic acid residues involves more than 20 different sialyltransferases, the most important subgroup being C8 sialyltransferases (ST8Sia I–VI), which transfer a polysialic acid (PSA) chain via an α2,8-linkage to proteins. Sialic acid (Neu5Ac) may occur in glyco-conjugates in α2,3-Sia, α2,6-Sia, α2,8-oligosialylated (3–7), and α2,8-polysialylated (8–200) forms. Various groups, which are attached to sialic acid, are designated as R1–R9 (left). Most common sialic acids are N-acetylneuraminic acid (Neu5Ac)/N-glycolylneuraminic acid (Neu5Gc)/3-deoxy-non-2-ulosonic acid (KDN) [13,37], in which the fifth. carbon atom is substituted by acetoamide/hydroxyacetoamide/hydroxyl groups [40]. Their structure and function depend on the linkage type between sialic acid residues and possible substitutions of sulfate/hydroxyl-groups in positions C4/7/8/9 by acetyl/lactoyl/methyl/sulfonyl or phosphonyl groups. (B): Polysialic acid chains (with α2,8 linkages) can be attached to different carriers including O- or N-glycans, GPI anchored proteins, oligosphinglolipids, and phospholipids. The functionally most important polysialic acid chains have a α-helical conformation with dynamic function by their carboxyl groups being directed inwardly and the N-acetyl groups (PolyNeu5Ac and PolyNeu5Gc) outwardly. (C): The consequences of altered glycosylations are shown in part (C). The scheme indicates that tumor cells with altered glycosylation will show a series of distinctly changed biologic processes.
Cancers 13 05203 g001
ST8Sia II and ST8Sia IV sialyltransferases specifically modify carbohydrate chains of certain neuronal cell adhesion molecules, such as NCAM or Neuropilin-2 and SynCam-1 [43,44]. They link up to 200 sialic acid residues by α-2,8-linkage (polysialic acid) and thus endow these molecules a multifunctional role in biological processes including all steps of cancer progression, e.g., cell–cell and/or cell–extracellular matrix interactions, increased proliferation, migration, and decreased apoptosis rate, as well as evasion from immune effector cells and metastasis of tumor cells [45].
The polysialic acid (PSA) chains, which are most flexible, are functionally most important; they have a α-helical conformation with dynamic characteristics caused by their carboxylic group (ionized at physiological pH), which are directed inwards and the N-acetyl groups (PolyNeu5Ac and PolyNeu5Gc) outwards [8,41]. The consequences of altered glycosylations are shown in Figure 1C.
In contrast to linear macromolecules, like proteins and nucleic acids, carbohydrates can form complex structures. This is based on their physico-chemical reactivity to form branches and their isomeric conformation [46]. When linked covalently to proteins or lipids, the resulting glycoproteins or glycolipids (gangliosides) are endowed with new functions of interaction, e.g., by forming bridges between proteins or proteins and glycolipids (Figure 1B). This complexity corresponds to a code with distinct geometric features that has the ability to transmit information with geometrically progressing proficiency [13,47].
Tumor cells are characterized by a general increase in total sialylation and expression of respective carrier molecules as compared to normal cells [27,48]. These changes in glycosylation have been termed a new hallmark [49,50,51] of malignant transformation. They are driven by altered gene expression, substrate availability, cellular environment, and protein conformation, which can be altered by mutations [27,42]. The degree of sialylation, their type of linkage as well as the length (mono, di, oligo, poly) of the resulting sialic acid chain will determine subsequent effects. The synthesis of PSA is mediated by two dominant polysialyltransferases (ST8Sia II and -IV), which transfer a poly/oligo sialic acid chain via a 2,8-linkage to proteins (resulting to 200 residues with helical form) [13,52].

3.2. Physiologic Role of Poly/Oligo-Sialylated Adhesion Molecules and Their Interaction with Growth Factors and Their Receptors

The biological role of sialic acids in relation to the evolutionary perspective of organisms is enormous. Given the breadth of the topic, it is not possible to cover individual topics in detail in this overview, but we briefly describe some areas in which sialic acids play an important role. These include physiological functions as homing of leukocytes via selectins; modulation of the immune system through interaction of sialylated pathogens with siglecs or other receptors located on immune cells; sialic acids as ligands for many microbes or viruses; influence of sialidases or sialic acid transferases on immune cell reactions; modulation of biophysical effects by factor H; suppression of immune cells by apoptosis-inducing sialylated glycoproteins; sialic acid during fertilization (egg-sperm interaction); and sialic acid to control embryonic and neuronal development and learning/cognitive functions [53,54].
The dynamic development of the nervous system depends on processes such as neural plasticity, migration and differentiation of neural precursor cells [55,56,57,58], and the formation of neuronal networks through regulation of genes implicated in neurite growth, guidance, and target recognition. Within this development, the cell numbers are controlled by regulating cell survival and apoptosis, and neural connections are strengthened by the fasciculation of neurites, and the formation, maturation, and plasticity of synapses [56,58,59,60]. All of the aforementioned procedures are regulated to a large extent by a few cell adhesion molecules (CAMs) with specific polySia glycosylation, which possess regulatory properties for homophilic and heterophilic interactions to other cells (trans), or to components of the extracellular matrix (ECM) (cis or trans) [61,62]. If poly/oligo-sialylated glycoproteins are expressed on both interaction partners, they will inhibit cell–cell and cell–extracellular matrix interactions by steric and repulsive hindrance, caused by their bulky poly/oligo-anionic configuration [63]. They are normally expressed in neural and synaptic cells [e.g., NCAM-1PSA, [64] neuropilin-2 (NRP-2PSA) [43] SynCAM-1 (CADM-1PSA) [65]], in various lymphoid tissues and activated B and T lymphocytes [e.g., C-C chemokine receptor type 7 (CCR7PSA or CD197PSA [43,66]], and in stem cell-derived microglia [e.g., E-selectin ligand-1 (ESL-1PSA)] [67]. Other glycoproteins characterized by poly/oligo-sialylation include the α subunit of the voltage-dependent sodium channel [68], the MUC1 protein from the serum of breast cancer patients [69], and CD36 (known as fatty acid translocase or scavenger receptor), which can also be found in human milk [70]. A large number of cells express also polysialyltransferases, which are capable of auto-polysialylation and transferring this PSA-chain to the aforementioned carrier glycoproteins [45,71].
The polysialic acid-associated adhesion molecules are able to interact with their homologous co-receptors and may thus regulate biological processes in the immune defense by homophilic and heterophilic interactions. This is illustrated by, e.g., NCAM-1PSA [55,72] its co-receptor NCAM-2 (RNCAM), SynCAM-1PSA (TSLC1, CADM-1), its respective co-receptors SynCAM-2/3 (CADM2/3 or TSLC2/3) [73], NRP-2PSA (Neuropilin-2 or VEGF165R2) [74,75,76], and its homologous co-receptor NRP-1 [77,78]. Studies on NCAM-2 NCAM-2/OCAM (olfactory cell adhesion molecule, expressed on most cells of the sensory smell system) demonstrate, that the protein is expressed in the human CNS as well as in other tissues [79]. There are three isoforms of NCAM2 located on the membrane, with molecular weights of 90, 115, and 125 kDa, respectively, the smallest isoform of which is attached to glycosylphosphatidylinositol (GPI) on membranes [79,80,81,82,83]. NCAM-2 has a high sequence identity with NCAM-1 [84,85]. The sequence identity between NCAM-1 and NCAM-2 is highest at the Ig1, Ig2, and Ig5 modules and in the cytoplasmic region. This high similarity suggests that the genes encoding NCAM-1 and NCAM-2 are paralogs.
Comparable to NCAM-1 and NCAM-2, the function of polysialylated NRP-2 and its proteoglycan coreceptor NRP-1, which promote binding of the class 3 semaphorins (SEMA3) to their receptor plexinA/B, involves axonal guidance and growth suppressive properties in tumor cells [86,87,88] (Figure 2a). SEMAs were first described as negative mediators of axon pathfinding that repel axons and collapse growth cones [89,90]. Similar to this mechanism, SEMA3F is able to prevent the angio-invasion of tumor cells by repelling epithelial cells [91,92]. This is achieved when SEMA3 forms a complex with both Plexin and NRP-1/-2 [93,94,95]. Besides inhibiting the invasion of blood vessels, the formation of this complex induces a signal, which results in a higher apoptosis rate and further confines cell growth, migration, metastasis, and angiogenesis [91,92,96,97,98]. As SEMA3s and VEGFRs share the same ligands (NRP-1/NRP-2), the binding of SEMA3s constitutes a competitive inhibition of VEGFR binding and thus a negative effect on tumor progression [99,100] (Figure 2a,b).
In this way, NRP1 and 2 interact with SEMA3 membrane-associated receptors located in the central nervous system (plexins) and their growth factors (class 3 semaphorins) [101,102,103]. In addition, they bind to a series of growth factor s and their receptors, including VEGF/VEGFR [75,76,104,105,106], TGF-β1/TGFβ1R [107,108], HGF/c-Met [109,110], PDGF/PDGFR [111,112,113], FGFs/FGFR [112,114,115,116,117], EGF/EGFR, BDNF/TRKA,B [118] as well as IGF/INSR [119]. These interactions favor the growth of potential tumor cells (Figure 3a,b).
Receptor tyrosine kinases (RTK) cannot be activated or induced solely by their respective receptor, as they require additional induction molecules that contribute to receptor activation and even signal transduction. Examples of RTK/co-receptor pairs include EGFR with E-cadherin, alternatively with NCAM-1, [121,122], CDCP1 [123], CD44 [124], L1-CAM (CD171) [125], SynCAM-1 [126], or TSP-1 [127]. Also included is FGFR with Syndecans [128], or, alternatively NCAM-1 [116,129,130,131], TSP-1 [127], as well as VEGFR-2/-3 with NRP-1/NRP-2, or CADM1, NCAM [132,133,134,135,136,137]. Interestingly, VEGFR-2 can also interact with VE-cadherin [138,139], and VEGFR-1 with NRP-1 [133,134,140,141,142,143]. Furthermore, PDGFR can interact with integrins, or NCAM, CADM1 [130,142], CD44 [144], Necl-5 [145], and TSP-1 [127]. Moreover, TGF-β can interact with syndecan-2 [146,147], NRP-1/NRP-2, and NCAM [108]. Further, there is a collaboration between c-Met and several co-receptors, e.g., α6β4 integrin [148], as well as between plexin A/B with NRP-1/NRP2 [43,96] and L-CAM [149]. The interaction of plexin A with NRPs has recently been described in detail [150]. Initially, an interaction is started between a plexin A glycoprotein and a neighboring NRP-1/NRP-2. Subsequently, the NRP–plexin A interaction will prompt the formation of a complex with two NRP-1/-2 and plexin proteins, respectively, which are complemented by binding to a Sema3 factor (Figure 2a). Afterwards, this complex is able to trimerize for intensive signal transduction [102,151].

3.3. Re-Expression of Polysialylated Adhesion Molecules in Cancer Progression

Besides the physiologic functions of poly/oligosialylated adhesion molecules, they have also pathophysiologic roles. In this regard, quality and quantity and not simply presence or absence play important roles not only in the neuronal and embryonal development, but also in cancer metastasis and some psychiatric diseases. In fact, re-expression of polysialylated adhesion molecules by tumor cells will lead to increased proliferation, plasticity, migration, angiogenesis, metastasis, and downregulation of important adhesion molecules.
From a mechanistical perspective, the negative charges between effector and target cells result from the opposing carboxylate groups of glycoproteins capped with sialic acid. They generate antiadhesive properties [8,56,152,153] by providing an electrostatic repulsive field, which regulates cell–cell interactions in trans, and among adhesion molecules on the cell surface in cis [154]. These sialic acids based physiologic mechanisms are also incorporated in pathophysiological traits and support proliferation, migration, and resistance of tumor cells [18,19]. High sialylation levels also protect tumor cells from complement-mediated lysis by preventing antibody binding [155,156] and from phagocytosis by immune cells [156,157].

3.4. Lectins Are Potential Co-Partners of Sialylated Glycoproteins

Proteins or glycoproteins with affinity for carbohydrates have been termed lectins, which can be found in plants, animals, viruses, and bacteria [8,158]. Lectins have important roles as intracellular, cell surface, or secreted molecules. Secreted lectins are incorporated to a varying degree into the extracellular matrix, where they regulate biologic processes between cells and their matrix by interactions with external glycans. For example, the viral lectin hemagglutinin is a structural protein of the influenza virus capsid, which can bind to sialic acid residues located on the surface of target cells. For cellular uptake, another structural virus protein, termed neuraminidase, cleaves the glycosidic bond to sialic acid residues, and thus liberates the virus for fusion with the target cell membrane [30,159,160]. In general, lectins can participate in pathological and physiological processes and have different interactions with the immune system, depending on their structure. Typically, lectins contain two or more binding sites for carbohydrate units of proteins, but some may have an oligomeric structure with multiple binding sites [161,162]. The affinity between lectins on the surface of one cell and carbohydrated proteins of neighboring cells is relatively weak, but the sum of interactions is strong due to the resulting sum effect [163].
A key function of lectins in mammals is to serve as adhesion molecules, which facilitate cell–cell interactions. As a result, signal transduction is triggered, including the activation of immune effector cells [164]. Some plant lectins can also serve as potent toxins. The mammalian lectins can be differentiated into classes based on their amino acid sequence and biochemical properties (Table 1). However, lectins in animals serve to interact with existing recognition sites of sialic acids, which initiate a broad range of biological processes and thus affect the complex roles of sialylated glycoproteins [38,165].
One large class are the C-type lectins (C for calcium-dependent). These proteins constitute a superfamily, and its members have in common a domain of 120 amino acids that are responsible for carbohydrate binding. In this structure, a calcium ion links a mannose residue to the lectin and this renders their interaction with co-partners to be calcium-dependent. C-type lectins recognize a vast array of ligands that regulate various physiological functions, including those in the innate and adaptive immune responses. Defects in these molecules lead to developmental and physiological abnormalities, as well as altered susceptibility to infectious and non-infectious diseases [166].
Selectins are members of this family, which are present on epithelial (E), lymphocytic (L), and platelet (P) cells, and by their carbohydrate binding facilitate the homing process of immune-system cells [167,168,169]. All three selectins (heterophilic CAMs) preferentially bind to sulfated and fuco-sialylated derivatives, e.g., PSGL-1 (P-selectin glycoprotein ligand-1), which is a mucin-type glycoprotein expressed on all white blood cells [170,171].

3.5. I-Type Lectins or Siglecs

Another important class are the I-Type lectins or siglecs (sialic acid-binding Ig-like lectins), which belong to the Ig superfamily. They are characterized by their specificity for sialic acids, which are attached to the terminal regions of cell-surface glycoconjugates. These type 1 transmembrane proteins comprise a sialic acid-binding N-terminal V-set domain, variable numbers of C2-set Ig ligand binding domains, a transmembrane region, and a cytosolic tail (Figure 4). Based on their sequence similarities and evolutionary conservation, two primary subsets of siglecs have been identified: the first subset includes siglec-1 (sialoadhesin), siglec-2, siglec-4 (MAG or myelin-associated glycoprotein), and siglec-15, all of which are well-conserved in mammals [172].
Siglec-1 (sialoadhesin), which lacks tyrosine-based motifs, plays a role as a positive regulator of the immune system and is a target for sialylated bacteria, enveloped viruses (e.g., HIV), and other sialylated pathogens [177,178].
The second subset of siglecs is designated to the CD33-related siglecs, which potentially inhibit immune response activation by ITIM-dependent signals. ITIM-containing CD33-related siglecs-3 and 5-12 siglecs are negative immunoregulators and endocytotic receptors. A subgroup of these CD33-related siglecs lack the intracellular ITIM but are associated with the adaptor protein DAP-12 (DNAX-activating protein), which contains the immunoreceptor tyrosine-based activating motif (ITAM) [179]. Additionally, CD33 induces apoptosis, and it may also enhance the production of anti-inflammatory cytokines and suppress the production of pro-inflammatory cytokines. According to a recent publication, the overexpression of siglec-9 in macrophages inhibits the production of pro-inflammatory cytokines such as TNFα and enhances the production of the anti-inflammatory cytokine IL-10 in an ITIM-dependent manner in response to toll-like receptor signaling [177,180]. With the exception of siglec-4, in humans, the CD33-related siglecs are expressed differently on various subsets of leucocytes, where they play an important role in the regulation of immune and inflammatory responses as discussed recently [181,182,183]. In this context it is noteworthy that siglec-3, which interacts with polysialylated NCAM-1 (CD56), is a key player in the plaque formation of Alzheimer’s disease [184,185,186,187,188,189,190].
NK and other effector cells express various siglecs, e.g., siglec-3, 7, 8, and 9 [191]. Their presence stabilizes the conformation of membrane glycoproteins in cis interactions with endogenous sialoconjugates at the cell surface of NK cells [192]. In addition, they are MHC class I-independent inhibitory receptors on immune effector cells, which—if stimulated—prevent the activation of these cells. However, the expression of corresponding ligands on target cells results in the inhibition of NK cell-mediated cytotoxicity by interacting with siglec-7 and 9 [191]. The outermost position of sialic acids during the glycosylation process implies the capping of galactose and related residues by sialic acids. As a result, these residues are not available for interaction with, e.g., galectins inducing apoptosis [24,193].
Siglecs often function as sensor for sialylated glycoproteins [194]. Based on their intracellular ITIM, they induce strong inhibitory signaling when binding to sialic acid [195]. Interestingly, this mechanism is also used by tumor cells and pathogens to escape the immune system, by adding sialic acid residues to their glycan structures, thus highlighting that the sialic acid–siglec interaction is key to the immune function against pathogens and cancer [196,197]. Thus, the, siglecs 14, 15, and 16 play a role in positive and negative immune regulation [161,181,198]. Siglec-14 has an arginine residue in its transmembrane region that is required for its association with DAP12 (ITAM-containing adapter), by which PI3K is recruited [161,181]. It is reported that Siglec-5 and Siglec-14 can be associated for delivering opposing signals via ITIM- and ITAM-dependent pathways, respectively. Another residue of arginine (the first immunoglobulin domain) required for sialic acid recognition by Siglec-5 and Siglec-14 is present in humans, suggesting that these two proteins work cooperatively and balance activating and inhibiting signal transmission through sialic acid recognition [161,181,182].

3.6. Sialylation of Check Point Receptors

Studies from anti-inflammatory check point receptors (e.g., CTLA-4 [199,200], PD-1 [200,201], and TIM3 [200,202]) show that an altered degree of sialylation (Figure 1C) influences the interaction with respective co-partners and their proper function [51,203,204,205].
The same is true for pro-inflammatory immune checkpoint receptors, such as the TNF receptor superfamily including 1-4-BB, [206] as well as the two stimulatory immune checkpoint proteins CD28 [207] and ICOS (B7-H2) belonging to the B7-CD28 superfamily [205,208,209,210].
It is known that the extracellular domain of ICOS carries three putative N-glycosylation sites [211]. Glycan modification of ICOS is essential for correct folding, trafficking to the cell surface, and ligand binding activity [212]. Other checkpoint molecules like PD-1 and its ligand programmed death ligand 1 (PD-L1, B7-H1) also carry N-glycosylation sites (four in PD-1 and one in PD-L1). Carbohydrate modification of PD-1 is not required for ligand binding, but determines the orientation of PD-1 in vivo and is therefore able to influence its complex formation with PD-L1 [213,214]. PD-L2 is a transmembrane protein expressed in normal tissues to inhibit the activity of T-cells and prevent autoimmunity. PD-L1 is commonly upregulated on the surface of tumor cells, binds to PD-1 expressed on tumor-infiltrating lymphocytes, and eventually causes T-cell tolerance [215,216].
CTLA-4 contains two glycosylation sites in its extracellular domain with low glycan heterogeneity and high production of tetra-antennary N-glycan structures [217]. Low N-glycosylation of CTLA-4 at the cell surface reduces T cell hyperactivity in autoimmunity [197,218].

3.7. Sialylated Glycans in Tumor Cells Prevent Galectin Induced Apoptosis, Autophagy, and Cluster Formation

Galectins (S-type lectins) can induce apoptosis, autophagy, and cluster formation of glycoproteins [140,219]. They specifically bind to β-galactosides, such as N-acetyl-lactosamine (Galβ1-3GlcNAc or Galβ1-4GlcNAc), which is linked by N- or O- glycosylation, but this interaction is impeded by increased sialylation of the corresponding glycoproteins present on tumor cells, which reduce their interaction. Other apoptosis-inducing lectins include, e.g., C-type lectins and annexins [162]. Galectins have three different forms, including dimeric (homodimeric galectin 1, 2, 5, 7, 10, 11, 14, 15), tandem (heterodimeric galectin 4, 6, 8, 9, 12), or chimeric structures (galectin-3 exists in monomeric form or, associated via the non-lectin domain, in a pentameric complex). Their stability and ability to bind carbohydrates depends on disulphide bonds. Galectins have a broad variety of functions, including mediation of cell–cell interactions, cell–matrix adhesion, and transmembrane signaling for, e.g., apoptosis regulation [219,220,221]. Galectins participate in controlling the positive and negative selection of T cells in the thymus, as they prevent the circulation of T cells, which would recognize self-antigens by interacting with, e.g., CD43, CD45, or CD7 and then become self-reactive [222,223,224]. Following an immune response to eradicate the excess T cells, epithelial thymic cells secrete both galectin 1 [225,226] and galectin 9, thereby mediating apoptosis in activated or infected T cells [227,228]. Interaction of apoptosis inducing glycoproteins (e.g., CD43, CD 45, CD7, etc.) with galectin 1, galectin 3, or galectin 9 will further initiate different intracellular death pathways (Figure 5a,b) [229,230,231].
The function of many glycoproteins of the cell surface depends on their cluster formation [e.g., integrins [232,233], CD45 [234], TNFR, TRAILR, EGFR [29,235], PECAM [236], and Fas death receptor (CD95) [12], but this function is lost following their high degree sialylation.
The degree of cell surface sialylation is regulated by numerous enzymes, including (i) enzymes that control the synthesis and distribution of the activated sialic acid substrate (i.e., cytidine-5′-monophospho-N-acetylneuraminic acid (CMP-sialic acid)), (ii) the sialyltransferases, and (iii) the sialidases (neuraminidases or neu1-4). In all cancers, altered expression and activity of both sialyltransferases and sialidases is observed [237,238]. These enzymes are typically found in subcellular compartments. The sialyltransferases are located in Golgi and the sialidases in lysosomes, endosomes, and membranes. In all cases, it has been shown that sialylation levels are higher in tumor cells [49,239,240]. Similarly, in macrophages, Fas, and TRAIL receptors, which show increased N-glycan sialylation, they prevent the formation of clusters and thus the binding of apoptosis-inducing ligands [209].
Figure 5. (a) Regulatory function of polysialylated glycoproteins in conjunction with adhesion molecules in NKT- effector and non-tumor cells. NKT-effector cells show a high degree of sialylation, which is precondition for their properties, including migration, mobility, and reduced adhesion. The upper part shows examples of glycoproteins present on NKT effector cells carrying either polysialic acid residues (NCAMPSA) or α2,3/α2,6 linked sialic acid residues (CD45, CD43, CD7, CD2/LFA2, integrins), inhibitory receptors (siglecs, PD1, TIM3), T-cell receptors (CD3, CD8) as well as sialic acid/heparin sulfate-binding adhesion molecules (CD44, and L-selectin). On normal cells (bottom part), these glycoproteins are opposed by ligands/receptors, which allow interaction with the effector cells, including PD-L1/2, HLA, abc, LFA-3, MADCAM-1, VCAM-1, ICAM-1, CD44s, E-cadherin, NCAMPSA-neg, L1-CAM, low sialylated CD7, CD43, and CD45. The normal or down-regulated levels of sialylation will promote target- effector cell interactions, binding of galectins to β-galactosides of glycoproteins, as well as facilitate apoptosis induction by binding of galectin 3 or other ligands to apoptosis inducing receptors (e.g., CD43, CD7, CD45) [241,242]. (b) Regulatory function of polysialylated glycoproteins in conjunction with adhesion molecules in tumor and NK-effector cells. NK-effector cells show a high degree of sialylation, which is precondition for their properties, including migration, mobility, and reduced adhesion. The upper part shows examples of glycoproteins present on NK effector cells carrying either polysialic acid residues (NCAMPSA) or α2,3/α2,6 linked sialic acid residues (CD45, CD43, CD7, CD2/LFA2, integrins), inhibitory receptors (siglecs, PD1, TIM3, NKGA/B, KIR receptors), as well as sialic acid/heparin sulfate-binding adhesion molecules (CD44 and L-selectin). On tumor cells (bottom part) these glycoproteins are opposed by ligands/receptors, which do not allow interaction with the effector cells, because they are highly sialylated. In addition, many polysialylated adhesion molecules are upregulated (including NCAM, NRP2, SynCAM1, mucins (e.g., MAdCAM-1, CEACAM-1), N-cadherin, L1-CAM, fucosialylated CD44v6) [243]. Inhibitory ligands (PDL1/2, HLA-E/G) show increased affinity for their co-partners on effector cells and thus inhibit the function of the effector cells. High sialylation of CD45, CD7, and CD43 inhibits the binding to galectins and thus protects against apoptosis induction of tumor cells [242].
Figure 5. (a) Regulatory function of polysialylated glycoproteins in conjunction with adhesion molecules in NKT- effector and non-tumor cells. NKT-effector cells show a high degree of sialylation, which is precondition for their properties, including migration, mobility, and reduced adhesion. The upper part shows examples of glycoproteins present on NKT effector cells carrying either polysialic acid residues (NCAMPSA) or α2,3/α2,6 linked sialic acid residues (CD45, CD43, CD7, CD2/LFA2, integrins), inhibitory receptors (siglecs, PD1, TIM3), T-cell receptors (CD3, CD8) as well as sialic acid/heparin sulfate-binding adhesion molecules (CD44, and L-selectin). On normal cells (bottom part), these glycoproteins are opposed by ligands/receptors, which allow interaction with the effector cells, including PD-L1/2, HLA, abc, LFA-3, MADCAM-1, VCAM-1, ICAM-1, CD44s, E-cadherin, NCAMPSA-neg, L1-CAM, low sialylated CD7, CD43, and CD45. The normal or down-regulated levels of sialylation will promote target- effector cell interactions, binding of galectins to β-galactosides of glycoproteins, as well as facilitate apoptosis induction by binding of galectin 3 or other ligands to apoptosis inducing receptors (e.g., CD43, CD7, CD45) [241,242]. (b) Regulatory function of polysialylated glycoproteins in conjunction with adhesion molecules in tumor and NK-effector cells. NK-effector cells show a high degree of sialylation, which is precondition for their properties, including migration, mobility, and reduced adhesion. The upper part shows examples of glycoproteins present on NK effector cells carrying either polysialic acid residues (NCAMPSA) or α2,3/α2,6 linked sialic acid residues (CD45, CD43, CD7, CD2/LFA2, integrins), inhibitory receptors (siglecs, PD1, TIM3, NKGA/B, KIR receptors), as well as sialic acid/heparin sulfate-binding adhesion molecules (CD44 and L-selectin). On tumor cells (bottom part) these glycoproteins are opposed by ligands/receptors, which do not allow interaction with the effector cells, because they are highly sialylated. In addition, many polysialylated adhesion molecules are upregulated (including NCAM, NRP2, SynCAM1, mucins (e.g., MAdCAM-1, CEACAM-1), N-cadherin, L1-CAM, fucosialylated CD44v6) [243]. Inhibitory ligands (PDL1/2, HLA-E/G) show increased affinity for their co-partners on effector cells and thus inhibit the function of the effector cells. High sialylation of CD45, CD7, and CD43 inhibits the binding to galectins and thus protects against apoptosis induction of tumor cells [242].
Cancers 13 05203 g005aCancers 13 05203 g005b
This will result in the overexpression of the STGal-family members on tumor cells, which are then characterized by increased sialylation (all linkages) and will escape the immune system by blocking multiple signaling pathways leading to apoptosis (e.g., galectin receptors, TNFR1, Fas, TRAIL, etc.) [240,244].
TNFα and FasL, the ligands for TNFR1, TRAILR, and Fas, are mainly expressed by immune cells, which are also a rich source of galectins. The latter are characterized by their binding to receptors (e.g., CD43, CD7, CD29, CD45, and others), which can also induce apoptosis [230,245] (Figure 6a,b). This outcome, however, is prevented in tumor cells as they express sialoproteins or gangliosides, which protect cancer cells from immune effector cells via interaction with inhibitory receptors (e.g., siglecs).

3.8. Polysialylation of Glycoproteins Generates Diverse Functions

Sialic acid plays also an important role in the transport of proteins, amino acids, and ions into cancer cells [38,246,247]. Surface glycoproteins of dendritic cells (DCs) are mostly sialylated. It is suggested that sialylated glycoproteins in DCs have an impact on several aspects of DC biology. Blocking of sialic acid expression in human monocyte-derived DCs (moDCs) by a synthetic, fluorinated sialic acid, which, as a potent mimetic blocks sialic acid expression, enhances the responsiveness of moDCs to toll-like receptor (TLR) stimulation [248]. Additionally, sialidase treatment of DCs improves the efficacy of antigen presentation of DCs, which may be used for vaccines in anti-cancer immunotherapy [66,249,250].
The ensuing sialylation effects include blocking of galectin interactions with galactose residues, which causes changes in protein conformation and stabilization of the spatial distances on cell membranes. This results in altered cellular properties, characterized by anti-adhesion (Figure 5b), cell flow behavior in suspension, metastasis, but also cellular and synaptic plasticity, necessary for cell differentiation and migration, cell growth, and development (e.g., NRP-2/EGFR-complex, Figure 2b) [3,8,18,251,252,253].
Additionally, interactions between sialylated glycoproteins or glycolipids and their ligands (e.g., lectins) regulate important intra- and extracellular signal transduction pathways. These include modulation of the affinity between growth factors and their receptors with consequences for cell differentiation, as well as cell growth and development (Figure 2a,b). Specifically, this interaction alters intra- and extracellular communication related to the initiation of signal transduction for activating the immune system against cancer cells and viral or bacterial pathogens [8]. Mechanistically, the adhesive carrier proteins are located in the area of unsialylated growth factor receptors, which are clustered by galectins and exert an intrinsic capacity to interact with these receptors [251].
The additional presence of PSA chains allows for the regulation of signal transduction via influencing the access of ligands to their receptors, thus supporting/regulating the cellular growth and survival as well as angiogenesis [8,117] (Figure 2b). These effects are consequences of functional imbalances between sialic acid residues located on carrier proteins and their lectin or lectin-like binding partners. This ultimately leads to the escape of pathogens (parasitic, bacterial, and viral) and/or tumor cells from the innate/adaptive immune system or even in hyper-reactivity associated with autoimmune and neuronal diseases [196]. Regarding the latter diseases, only a short hint to a vast array of publications seems appropriate in the context of this review. In fact, numerous publications have described that imbalances in sialic acid distribution and the degree of sialylated glycoproteins in conjunction with their co-partners are causal factors for certain neuronal diseases, such as Alzheimer’s disease [85,188,189,247,254], Parkinson’s disease [255,256], multiple sclerosis [257,258,259], and schizophrenia [260]. It is well known that the absence of sialic acid and galactose residues or the upregulation of certain lectins, e.g., siglec-1 in body cells, leads to various autoimmune diseases and tissue inflammation by activation of immune effector cells [261]. T-cell populations have their own specific surface glycosylation profiles, which are largely responsible for the different susceptibilities of individual subpopulation cells to galectin-1 or galectin-9 -induced cell death [262,263]. The main factor for selection and deletion of T cells during development is galectin-1 induced cell death, which is crucial for the development of distinct subpopulations in the thymus. Low level sialylation of glycans on the surface of CD4/CD8 double positive thymocytes renders this population susceptible, whereas high level sialylation of glycans confers resistance to galectin-1 binding and ensuing apoptosis [225,264]. Low level sialylation of CD43, CD45, and CD7 on T-cell surface allows binding of galectins, and initiates p53/ARF-dependent apoptosis during the process of differentiation [241,265].
Chien et al. published that the differentiation of Th1, Th2, and T-reg cell populations strongly depend on the availability of glucosamine as a substrate for N-glycosylation of CD25, which will impair Th1, Th2 and T-reg populations, but will also enhance Th17 differentiation [203]. The fact that different forms of glycosylation will determine a cell’s fate can be concluded from the following observation: incorrect N-glycosylation of CD25 results in altered IL-2 signaling, which plays an important role in the regulation of T-cell differentiation [203,266,267]. In addition, the core-2 branching of CD45 is associated with increased susceptibility to galectin 1 mediated T cell death [229,268]. However, sialylation of CD45 N-glycans by ST6Gal I inhibits its recognition by galectins, thus preventing the clustering of this receptor tyrosine phosphatase at the cell surface, and rendering the T-cells resistant to galectin-1 mediated cell death [269,270,271]. Various glycosyltransferases of the Golgi apparatus are responsible for posttranslationally modifying the N- and O-glycans, which are crucial for TCRs association with other glycoproteins (MHC-I or II plus antigens) and thus generate the TCR signal transduction and receptor internalization by endocytosis [269]. In summary, TCR, CD28, and CD45 each have multiple N-linked glycosylated sites that interact with endogenous galectin 1 and 3 [225,272].

3.9. Proteins with the HNK-1 Epitope Serve a Function Similar to Poly/Oligo-Sialylated Glycoproteins

The human natural killer-1 (HNK-1) epitope is a trisaccharide moiety with a 3′-sulfated glucuronic acid at the non-reducing terminal (HSO3–3GlcAβ1–3Galβ1–4GlcNAc-R) and is predominantly expressed in the central nervous system [273]. An important function of this epitope is to facilitate the migration of neural crest cells in the nervous system [274,275]. Proteins involved in the guidance and growth of axons include CAMs [276,277]. The NCAM glycoproteins are important examples of this group. They are a family (e.g., NCAM1, NCAM2) with different splice forms (i.e., different length) and roles. For its function, NCAM can be dimerized in cis- or trans-positions. In addition, cis-dimerized NCAM can form a tetramer by homophilic interaction in trans. For dimerization in cis, the Ig1/2 modules interact with each other, for tetramerization, they interact with Ig5 and Ig4, whereas Ig3 interacts with a corresponding Ig3 module [278]. Similar to NCAM1, NCAM-2 may be a marker of certain types of cancer, including human prostate cancer [84,279,280,281]. NCAM-2 is heavily glycosylated on its Ig5 module, which contains four of the eight possible N-glycosylation sites of NCAM-2. Accordingly, glycosylations on NCAM-2 play a role in axon guidance and target recognition [86]. The degree of glycosylation in the Ig5 module probably determines a possible modulation and function of NCAM-2 [282]. Similar to NCAM-2 [283,284], NCAM-1 (except its 180 kDa isoform) carries a HNK-1 carbohydrate structure at the Ig5 module [59,85] (see below).
As HNK1-carbohydrates are also known to modulate cell–cell interactions, the simultaneous presence of more than one carbohydrate epitope may reflect a new mechanism involved in the fine-tuning of NCAM functions. The HNK-1 carbohydrate consists of a small epitope of 3’-sulfated glucuronic acid, which is linked to a lactosaminyl residue that is involved in the homophilic binding of NCAM. [285]. However, the HNK-1 epitope lacks the ability of NCAMPSA to influence the pathways associated with memory consolidation. HNK-1 carbohydrate expression is regulated by the brain-derived neurotrophic factor (BDNF) and the tyrosine kinase receptor (TRKR), which both transmit signals in regenerating motor nerves and promote functional recovery after peripheral nerve repair [286,287].
The extracellular matrix molecule tenascin-R (TNR) also carries the human HNK-1 epitope, which inhibits postsynaptic GABA receptors [288]. Interactions of the HNK-1 epitope with chondroitin sulfate proteoglycans enhance neuronal cell adhesion and neurite outgrowth [287]. The HNK-1 epitope (also referred to as L2 or CD57) [289] is also found in a number of other neural cell adhesion molecules, including L1 [290], close homolog of L1 (CHL1) [291,292] myelin-associated glycoprotein (MAG) [110,290,293], melanoma cell adhesion molecule (MCAM) [293], and contactin [153,294].
Remarkably, tumor cells can use proteins with the HNK-1 epitope as support for tumor progression including migration and metastasis. Proteoglycans like NRP-1 or CD44V6 interact with certain integrins (e.g., integrin β1) and tyrosine kinase receptors [295,296,297] modified polyglycosaminoglycan chains (GAG) bearing chondroitin or heparan-sulfate. This modified version of GAG can contain an HNK-1 epitope and is thus able to promote or inhibit cancer cell growth, survival, and invasion [298,299].

3.10. Polysialylated Glycoproteins Are Co-Receptors for Growth Factors and Their Receptors

An important aspect of α2,8 linked poly/oligosialylated N-glycans of cell adhesion molecules is their ability to form complexes with growth factors and their receptors, thus modulating the respective signal intensity. Besides NCAM-1, the polysialylated carrier glycoproteins include NRP-2 [300,301], and SynCAM-1 [302,303], which both take part in synaptic plasticity of memory consolidation, differentiation, migration, and growth, as well as in cancer progression [300,304,305]. NRP-2PSA is a member of the neuropilin family [75,300], which is expressed on T/DC cells as well as in the central nervous system [306]. NRP-1 and NRP-2 are up to 44% homologous. The sequence homology of human and murine NRP-2 is around 94% [307]. NRP-2 has two major splice variants, which are categorized as NRP-2a and NRP-2b. NRP-2PSA is upregulated in a number of tumor types including osteosarcoma [308], melanoma [309], breast cancer [310], lung cancer [103,311], brain tumors [312,313], colorectal cancer [107,314], pancreatic cancer [315,316,317], myeloid leukemia [132], saliva adenoid cystic carcinoma [300], infantile hemangioma [318], as well as ovary- [100], bladder- [319], and prostate cancers [320]
The neuropilins are key receptors within the vascular, nervous, and immune systems and can be found on endothelial cells, neuronal axons, and regulatory T cells, respectively. They serve as co-receptors for the plexins in semaphorin binding on neuronal and vascular endothelial cells and for the VEGFRs in VEGF binding on vascular and lymphatic endothelial cells. Hence, they regulate the initiation and coordination of cell signaling by semaphorins and VEGFs [133,321].
Syn-CAM-1PSA is another adhesion molecule with an important function in embryonic development [302,303,322], which is re-expressed in T-cell leukemia [323,324], hepatocellular [325], and lung carcinomas [326]. Over-expression of these polysialylated glycoproteins in cancer correlates, as for NCAM-1PSA and NRP-1PSA, with an aggressive clinical phenotype. SynCAM-1PSA is expressed on activated NK/CD8+ T cells and on dendritic cells. The presence of NRP-2PSA, NCAM-1PSA, Syn-CAM-1PSA will change the balance of sialic acids on the cell surface and consequently reduce the inter-membrane adhesion between tumor and effector cells. This is achieved through steric hindrance caused by a high density of negative charges that contribute to the hydrated volume of polysialylated glycoproteins [43,327,328].

3.11. Polysialylated NCAM-1, NRP-2, and CADM-1 Potentiate Cell Growth Signaling in Tumor Cells and Increase Tumor Progression

The adhesion molecules NCAM-1, NRP-2, and CADM-1 can gain specific surplus functions by the addition of a PSA chain. With the aid of such a PSA chain they can stimulate various growth factor receptors in tumor cells. For example, polysialylated NCAM-1-stimulated migration and proliferation were paralleled by activation of the FGFR and its downstream signaling components. Moreover, NCAM- and FGF-2-mediated FGFR1-signaling in the tumor microenvironment of esophageal cancer stimulated the survival and migration of tumor-associated macrophages and cancer cells [29,309,329].
Epidermal growth factor receptor (EGFR) is a heavily glycosylated transmembrane receptor tyrosine kinase. Upon EGF-binding, EGFR undergoes conformational changes to dimerize, resulting in kinase activation and autophosphorylation and downstream signaling. Increased expression of ST6GalI in cancer cell lines enhances the α2,6-sialylation of epidermal growth factor receptor (EGFR), which increases its tyrosine kinase activity and the phosphorylation of its targets [6,330]. It has been shown that sialylation of epidermal growth factor receptor increases signaling activity and inhibits gefitinib-induced cell death [6].
Cell growth signaling in tumor cells and their progression depend on polysialylated NCAM-1, NRP-2, CADM-1, and some another membrane sialylated tumor proteins (see below as well as Figure 2a,b and Figure 3a,b) [298,331]. Another growth factor receptor, which is modulated in its activity by the presence of PSA chains on tumor associated neuropilins, is VEGFR. A complex formed between VEGFs, VEGFRs, and NRP-1/NRP-2 promotes tumor progression. In line with this, enzymatic removal of PSA from the cell surface led to reduced proliferation and activity of the extracellular signal-regulated kinase (ERK), thus inducing neuronal differentiation of neuroblastoma cells [332].
The respective ligand, human VEGF has five related genes, i.e., VEGF-A, -B, -C, -D, and placental growth factor (PlGF). All isoforms share the same binding sites for receptor tyrosine kinases [333,334]. VEGF-A ligands are bivalent and bind two monomers of their related receptors, i.e., VEGFR1 and/or VEGFR2. The isoforms are capable of binding and cleaving proteoglycans of the extracellular matrix [333] or proteoglycans located on the cell surface (HSPGs) [333]. Heparin and HS increase the affinity of VEGF-forms for VEGFRs and NRPs [134,335]. However, VEGF-A165a forms a ‘bridge’ between VEGFR2 and NRP1/2 by binding both receptors simultaneously [134,336]. Further, VEGF-A and VEGF-C induce the interaction of NRP2 with VEGFR-2 [134,135]. Knockdown of NRP2 expression potently inhibits human endothelial cell migration induced by VEGF-A and VEGF-C [135]. Binding of VEGF to dimerized VEGFR2/3 via Gal-1 or Gal-3 induces angiogenesis [337,338,339]. Tumor cells, which express VEGFRs with α2,6-linked sialic acid residues, are unable to induce angiogenesis, as VEGVRs are capped with sialic acid cannot bind to Gal-1- or Gal-3, which would otherwise favor dimerization of VEGFRs [251,340]. Another mechanism leading to similar results can be observed when polysialylated NCAM or NRP-2 form a complex with VEGFRs via additional factors [341]. This will prevent angiogenesis, but promote growth [135,224], cell signal transduction, and initiation of migration and metastasis via VEGFR-3 [18,134,305,338] (Figure 2b).
Islamov and colleagues have shown that co-transfection of VEGF with GDNF and NCAM, or VEGF with ANG and NCAM, or NCAM plus VEGF, or NCAM plus GDNF into transgenic mice (used as amyotrophic lateral sclerosis model), increased the life-span of the rodents. The results suggest that both approaches enhance the synaptic plasticity (up-regulation of PSD95 and synaptophysin), and support the proliferation, migration, and myelinization of neuron-glial antigen 2 (NG2; chondroitin sulfate proteoglycan 4) positive glia cells [342,343,344]. In contrast, it was shown that abolishment of NCAM in transgenic Rip1Tag2 mice, which are a model for pancreatic β-cell carcinogenesis, induces tumor metastases by upregulating lymph angiogenesis [137]. These authors, however, did not provide data on the sialylation status of NCAM. As has been described, some tumor cell lines express NCAM without polysialylation [345].
In summary, the polysialic acid-associated adhesion molecules are involved not only in cell–matrix and cell–cell adhesion and thus control [135,346] growth, migration, metastasis [304,347], vascular permeability [347,348], and block apoptosis in cancer stem cells [17,305,349,350], but also take part also in cell signal transduction [18,134,305]. Additionally, in a complex with growth factor receptors, these molecules control tumor progression [77,304] and regulate angiogenesis.
The important role of Ig-CAMs is revealed by their stimulation of tyrosine kinase receptors for epidermal growth factor (EGF), fibroblast growth factor (FGF), and nerve growth factor (NGF) [351,352].
Functions of polysialylated NCAM-1, NRP-2, and CADM-1 in tumor growth and tumor progression influence homo- and heterophilic interactions, which define key intracellular signaling pathways, e.g., by complex formation with FGFR-1, ERK1/2, FAK, and c-Met/ALK [45,353] as has been described recently in neurologic articles. In addition, data from the last fifteen years indicate that tumor cells re-express polysialylated glycoproteins in an analogous manner to bacteria that express polySia [33,354]. Both bacteria and tumor cells profit from the fact that they become resistant to the immune system and therapeutic approaches [14,196,355,356].
The functionality of growth factor receptors is closely linked to the clustering and dimerization of their subunits [118] (Figure 2b and Figure 3a,b). Further, ligand binding leads to receptor dimerization, which induces phosphorylation of the kinase domain and thus its activation. The signaling of these dimers, which respond to respective growth factors, will be amplified by association with oligo- and polysialylated glycoproteins, which are upregulated in tumor cells and function as co-receptors. Different receptors utilize different dimerization/activation strategies. For example, PDGF is a dimer, which cross-links two cell surface PDGF receptor monomers. In contrast, the binding of EGF to its receptor induces a conformational change, which promotes dimerization. Furthermore, FGF is complexed by heparin and crosslinks two FGF monomers. In the case of insulin, the receptor is already dimerized on the cell surface; ligand binding causes a conformational change and autophosphorylation [357].
NCAM specifically binds the glial cell-derived neurotrophic factor (GDNF) via its third immunoglobulin (Ig) domain in the plasma membrane. When NCAM and GDNF receptor GFRα1 are co-expressed, a complex is established via GDNF, which inhibits the homophilic cell adhesion of NCAM [358] and potentiates the role of GDNF in the survival and differentiation of neurons, as well as in malignant neurogenic tumors.
The interaction of brain-derived neurotrophic factor (BDNF) with tropomyosin receptor kinase A/B (TrkA/B) results in the dimerization of the receptor [359,360]. BDNF binds directly to polysialic acid, which caps two N-linked glycans on the Ig5 domain of NCAM and can be transferred to TrkA/B via direct contact or dynamic movement. This complex of BDNF with PSA upregulates growth and/or survival of neuroblastoma cells [118].
The transmembrane proto-oncogene tyrosine-protein kinase ROS is encoded by the ROS1 (c-ros oncogene) gene, which originally was discovered as a homologue of the transforming sequence of the avian sarcoma RNA virus UR2 [361]. The structure of ROS1 is similar to the human anaplastic lymphoma kinases ALK and LTK [362,363,364,365]. ROS1 is highly expressed in a variety of tumor cell lines including non–small-cell lung cancer (NSCLC), cholangiocarcinoma, and glioblastoma [361,366]. Abnormal ROS1 kinase activity leads to activated downstream signaling components of several oncogenic pathways including PLCγ, STAT3, PI3K/AKT, VAV3, and MAPK/ERK [367,368,369,370]. When the ROS1 kinase domain is fused to the ligand binding domain of EGFR or TRKA, and is stimulated on the cell membrane by a corresponding growth factor, activation of various combinations of previously noted pathway signaling components [367,371,372] will follow. ROS1 gene fusions (rearrangements) were first identified in a human glioblastoma cell line [368,373,374]. They have been further identified in solid tumors including inflammatory myo-fibroblastic tumor [375,376], cholangiocarcinoma [377], ovarian cancer [378], gastric cancer [379], colorectal cancer [380], angiosarcoma [381], spitzoid melanoma [382], and NSCLC [383,384,385,386]. Other fusion partners include surface proteins such as the novel solute carrier family 34-member 2 gene (SLC34A2), which was described in HCC78 cells [387]. Sequencing of lung cancer tissue revealed the presence of 14 different ROS1 fusion partner genes, including CD74, [366,383,388,389], SLC34A2, [366,383,387], syndecan 4 gene (SDC4) [366,383,388], ezrin gene (EZR) [388,390,391], fused in glioblastoma gene (FIG) [389,392], tropomyosin 3 gene (TPM3) [383,388], leucine-rich repeats and immunoglobulin-like domains 3 gene (LRIG3) [388], gene of KDELR2 (KDEL endoplasmic reticulum protein retention receptor 2 gene) [393], coiled-coil domain containing 6 gene (CCDC6) [394], moesin gene (MSN) [383,386], transmembrane protein 106B gene MEM106B) [365], tumor protein D52 like 1 gene (TPD52L1) [385], clathrin heavy chain gene (CLTC) [384], as well as LIM domain and actin binding 1 gene (LIMA1) [383,395]. Out of this list, CD74-ROS1 occurs most frequently in NSCLC. The break point of ROS1 gene-exons for fusions fusion partner genes are not at the same position [381]. All of the possible breakpoints in ROS1 allow the resulting fusion to preserve the ROS1 kinase domain while also retaining the transmembrane domain of fused ROS1 [388,396] (Figure 3b).

3.12. Characterization of NCAMPSA

The polysialylated form of NCAM plays a well-known role in brain development and neural plasticity [56,397]. NCAMPSA is an important glycoprotein of the brain, as it makes up around 30% of the weight of the embryonic brain and later still up to 10% of the adult brain [398]. In fact, polysialylated NCAM is present in embryonic phases and during the regeneration of adult neurons. However, it is strongly downregulated in adult neurons [399,400,401]. Interestingly, polysialylation is lost in adult brain tissue, except for areas with cognitive tasks. PolySia expression continues to be detectable in restricted regions only, including the hippocampus, olfactory bulb, amygdala, prefrontal cortex, and hypothalamus, in which neurogenesis and neural modeling continues throughout life [57,60,322]. In addition, NCAMPSA is expressed on monocytes, NK, and NKT cells.
Re-expression of polysialylated NCAM is found in neural diseases as well as in many neoplasms [402,403]. In the latter type of diseases, presence of NCAMPSA is related to an aggressive clinical phenotype and to metastasis [14,15,404]. For example, NCAMPSA is highly re-expressed in many solid tumors (e.g., colorectal carcinoma, small cell lung cancer [20,405], melanoma [406,407], breast cancer [42], neuroblastoma [408], and other brain tumors such as astrocytoma [409], etc.), as well as in a variety of hematological malignancies including acute myeloid leukemia, acute promyelocytic leukemia [404,410,411], acute lymphoblastic leukemia [412,413,414], NK-cell leukemia, B- and γδ T-cell lymphomas [415,416], etc.
The expression level of polysialylated, oligosialylated, or sialofucosylated adhesion molecules and the quantitative expression of carrier glycoproteins are the main regulators of adhesion and polarization between tumor and immune effector cells [9,40]. They inhibit cell–cell and cell–matrix interactions [305,417]. The polysialic acid-associated adhesion molecules mask the cell surface by their repulsive effect and downregulate other adhesion molecules, for example, ICAM and E-cadherin [121,345,418,419]. E-Cadherin and ICAM inhibit the migration of cells. Basically, both proteins are adhesion molecules, but E-cadherin is part of cell–cell junctions, and its presence therefore inhibits the separation of a cell from the environment. In slight contrast, ICAM binds to integrins and is partly responsible for the interaction between effector and target cells. A downregulation of E-Cadherin and ICAM through re-expression of NCAM-PSA or other PSA carrier adhesion molecules would therefore be expected to facilitate migration and metastasis of tumor cells.
For cell separation from muscle cluster tissue, downregulation of E/N-cadherin and upregulation of PSA are sufficient, thus reducing adhesion [420]. Further, they block the cytotoxic interaction of effector cells, and by siglecs can even induce apoptosis in neighboring effector cells. These combined effects are responsible for the immune escape of tumor cells expressing polysialylated adhesion molecules.
Tumor cells benefit from α2,8 poly/oligosialylated carrier proteins as their surface is masked and their apoptosis signaling is thus blocked (Figure 6b) [337,421]. Further, this results in a physical anti-adhesion effect between the effector and target cells. Moreover, the tumor cells may influence the effector cells by inhibiting their cytotoxic function via inhibitory ligands (siglecs or checkpoint proteins; Figure 4 and Figure 5b) or by induction of pro-apoptotic molecules (e.g., tumors upregulate CD70, which is a ligand for the apoptosis inducing protein CD27 on lymphocytes) [183,422]. In addition, in tumor and effector cells, the presence of α2,8-linked poly/oligosialic acid residues prevents homophilic interaction of adhesion molecules, as well as receptor ligand binding, e.g., for FAS and TNFR with FAS-L, galectins, and TNFα, which otherwise would induce apoptosis in these cells. However, their presence can elicit apoptosis in neighboring cells via trans-homophilic and heterophilic interactions [209,423] (Figure 6b).
The respective tumor cells seem to gain stem-cell-like properties including increased migratory and metastatic potential as well as a higher drug resistance to chemotherapy treatment [17,19]. They are masked towards effector cells (immune escape) and become resistant to anticancer treatment as the uptake of negatively charged drugs will be reduced [424]. In addition, hyper-sialylation of respective adhesion molecules inhibits binding to specific galectins (Gal 3 and Gal 9), which would induce apoptosis in the tumor cells [9,421]. In addition to this external mechanism, an intracellular mechanism is based on, e.g., sialylation of the Fas receptor, which prevents the initiation of the death receptor complex [220,423].

3.13. NCAM-1PSA, NRP-2PSA or SynCAM-1PSA Downregulate the Expression of Adhesion Molecules

A major obstacle to an endogenous anti-tumor immune response is poor infiltration of TILs into the tumor mass. The migration of immune cells into tumors can be hindered by many factors, such as an impaired chemokine expression in the tumor environment, the reduction of adhesion molecules in tumor cells, tumor- related endothelial cells, or leukocytes, and can be responsible for a defective monitoring of the immune system.
Cytotoxicity studies involving co-incubation of polyclonal natural killer (NK) or clonal NK-92 (NK tumor cell line) effector cells with NCAM-deficient or NCAM-transfected tumor cells showed strong reduction in NK-mediated lysis of tumor cells overexpressing NCAMPSA [345]. This can be explained by NCAMPSA overexpression-induced reduction of adhesion molecule expression, e.g., ICAM or E-cadherin [23,345,425]. This reduction, in turn, will inhibit the interaction and polarization of target and effector cells [345].
An effective mechanism of the tumor associated endothelial cells is the down regulation of adhesion molecules such as ICAM-1/2, VCAM-1, E-selectin, P-selectin, and MAdCAM-1 to prevent immune cell trafficking into the tumor site [426,427,428,429] (Figure 5a,b).
Moreover, the interaction of the PSA chains with inhibitory siglecs will block the effector cell activity or may induce apoptosis in the respective immune cells. In contrast, tumor cells being NCAM-deficient or expressing NCAM without polysialylation and cytoplasmic domain showed high susceptibility to NK lysis [345]. This effect was corroborated by an experiment in which the interaction of NCAMPSA tumor cells with NK cells was blocked by an NCAM-specific monoclonal antibody, which abolished the anti-adhesive properties of the PSA chains and thus normalized the NK cell mediated lysis of the target cells expressing NCAMPSA [345]. These two mechanisms highlight the key role of NCAM in the immune escape of tumor cells expressing NCAMPSA.
NCAMPSA interacts in heterophilic form in cis or trans with a series of surface glycoproteins including chondroitin sulfate, heparan sulfate proteoglycans [430], L1-CAM [431,432], E/N-cadherins [419,433], integrins [13,22], DC-SIGN [434], growth factor receptors [131], and their ligands. Heterophilic interaction of NCAMPSA with the glycoproteins noted above can modulate their interaction intensity with corresponding proteins from opposing cells, without affecting their intrinsic binding properties [55,433,435].
Furthermore, the interaction of PSA chains located on carrier glycoproteins with the noted glycoproteins can change their function, which then contributes to the progression of many cancer types [153,436,437]. Heterophilic and homophilic interactions of NCAM at the cell membrane in cis and trans depend on the presence of PSA. In comparison with CD2/LFA-3 and LFA-1/ICAM-1 interactions, which are involved in NK cell cytotoxicity [438], NCAMPSA on NK cells can negatively influence NK cell cytotoxicity against tumor cells expressing carrier glycoproteins decorated with PSA [345]. Proof of this fact can be found in the observation that antibodies against LFA1, LFA3, and ICAM-1 block tumor cell lysis caused by NK cell mediated cytotoxicity [438,439]. Anti-NCAM mAbs added to peripheral blood mononuclear cells (PBMC) inhibit cytokine-induced killer cells (CIK), which indicates an essential role of NCAM molecules expressed on NK cells in either alloantigen recognition or delivery of accessory signals to CD8+ T cell precursors [440].
An example for this modulation has been described for DC-SIGN (CD209), which interacts with weakly polysialylated NCAM in cis or trans. DC-SIGN on macrophages interacts with mannose type carbohydrates, being part of pathogen-associated molecular patterns commonly found on viruses, bacteria, and fungi [441]. This interaction induces phagocytosis of pathogens. DC-SIGN is a C-type lectin and has high affinity for ICAM3 (CD50) and DC-SIGNR (CD299) [442], but a low affinity to weakly polysialylated NCAM-1 [434]. DC-SIGN may bind various microorganisms by recognizing high-mannose-containing glycoproteins on their envelopes. It further functions as a receptor for several viruses such as HIV, Lassa, Ebola, and Hepatitis C [443,444,445]. Binding to DC-SIGN can promote HIV and hepatitis C viruses to infect T-cells via dendritic cells [444,446]. Lysis of HIV infected dendritic cells by polysialylated NCAMdim NK cells will be increased by the presence of an anti-DC-SIGN antibody, which inhibits the interaction between DC-SIGN and its ligand, and thus favors the interaction between DC-SIGN and NCAM-1 [434]. This in turn will decrease the repulsion from adhesion molecules in trans (neuropilin 2 on the surface of dendritic cells and NCAM on the surface of NK cells) and thus increase the cytotoxicity of these NK-cells [447]. Expression of NRP-2 is up-regulated in dendritic cells during maturation, coincident with increased expression of ST8Sia IV [447]. Additionally, activated effector cells of the adaptive immune system express concomitantly or alternatively polysialylated NRP-2, SynCAM-1 (CADM1), or polysialylated NCAM-1 [448].
Therefore, the presence of polysialylated glycoproteins on the surface of target and effector cells prevents interaction between dendritic cells and T lymphocytes. Removing PSA from NRP-2 or blocking NRP-2 by specific antibodies promotes interaction between dendritic and T cells [447]. Similar to this procedure, using an anti-NCAMPSA antibody modulates NK-mediated lysis of target cells expressing polysialylated NCAM [345].
Another example is given by the cis interaction of NCAMPSA with E-cadherin, which reduces the homophilic E-cadherin-mediated cell adhesion in trans [418,419] and thus promotes increased cell migration of pancreatic cancer cells. Cadherins are transmembrane proteins that mediate cell–cell adhesion, depending on the site of expression [419,449,450]. E-cadherin is instrumental in controlling cell polarity and organization of epithelial cells during embryonic development [436,449,451,452]. N-cadherin is normally down regulated during development and is absent in regenerated adult nerve fibers [453].
Cancer cells, which re-express polysialylated adhesion molecules like NCAM-1, NRP2, or SynCAM-1 by oncogenic K-ras will not only downregulate E-cadherin but also prevent homophilic interaction of E-cadherin-mediated cell adhesion in trans and thus promote metastasis [349,361,418,454,455]. Enzymatic removal of PSA from NCAMPSA or blocking polysialylation leads to increased E-cadherin-mediated cell–cell aggregation and decreased cell migration [418,419].
Mechanisms contributing to immune evasion are summarized below. As stated above, adhesion molecules including ICAM and E-cadherin are downregulated in tumor cells in response to the re-expression of NCAM1, NRP3, or SynCAM1, and therefore the polarization between tumor infiltrating lymphocytes (TILs) and solid tumor cells is interrupted [3,355]. TILs contain chemoattracted infiltrating immune effector cells expressing high levels of granzyme B and other highly sialylated markers, for instance mucin-like leukosialin/CD43, CD34, or high sialylated lamp-family such as Macrosialin (CD68) and/or polysialylated adhesions molecules such as NCAM (CD56), NPR-2, and SynCAM-1. These immune effector cells include CD8(+) T cells, CD56(+) NK cells, CD56(+) NKT cells, and CD68+ macrophages [3,355,456,457]. The consequence of interactions, resulting from effector cells (expressing NCAM-1PSA) and tumor cells (re-expressing NCAM-1PSA) is an anti-adhesion effect caused by electrostatic repulsion.

3.14. Dual Role of NCAM-1PSA

NCAM-1PSA has a key role in neuronal synapse development, but its re-expression in tumor cells promotes progression. NCAMPSA contains a heparin-binding domain, which after binding to heparan sulfate is vital for its developmental role in synapses, correspondingly, the binding of heparin to this site inhibits NCAM-1 polysialylation and thus prevents migration, invasion, and transition to resistant tumor cells [247,458,459].
Enzymatic removal of PSA or of heparan sulfates from neuronal cells diminished the formation of synapses on respective neurons, suggesting that interaction of NCAM-1PSA with heparan sulfate proteoglycans, e.g., CD44, is necessary for their function. In line with this, transfection of NCAM-1PSA-deficient neurons with an NCAM-1 isoform carrying PSA stimulated synapses formation on NCAM-1 isoform-expressing neurons [460].
L1-CAM (CD171) belongs to the immunoglobulin superfamily and is widely present on the cell surface of post-mitotic neurons, on axons of post-migratory neurons, and on glial cells [461]. Physiologically, L1-CAM is moderately expressed on the surface of immune cells, at moderate to low density on normal cells, and at intermediate density on many human tumor cells, including melanoma and neuroblastoma, carcinomas from lung, kidney, and skin, and monocytic leukemias [408,461,462,463]. In vertebrates, the family of L1-CAM (cell adhesion molecules) includes four structurally related transmembrane proteins: L1, close homolog of L1 (CHL1), NrCAM, and neurofascin. L1-CAM is a sialic acid-binding lectin and together with NCAM plays a role as signaling coreceptors in neuronal migration and process outgrowth [461,464]. L1 is an enhancer of integrin-mediated cell migration [465]. NCAM-180 and L1-CAMs are co-receptors of integrin- and GDNF receptor signal transduction [466]. Extracellular domains of L1-CAM have six Ig-like C2-type domains, five fibronectin type-3 domains with 21 potential N-linked glycosylation sites. L1-CAM can have homophilic interaction or heterophilic association with neurocan, phosphocan, laminin, integrins [αVβ3, αIIβ3, and α5β1] [467,468], CD9 [469], CD24 [464,470], NCAM-1 [461], neuropilins [454,471,472], FGFRs [473], and proteoglycans containing chondroitin sulphate, etc. [431,455].
Cell adhesion molecule L1 disrupts E-cadherin-containing adherens junctions and increases scattering and motility of MCF7 breast cancer cells [474,475]. The cytoplasmic domain of L1-CAM interacts intracellularly with ankyrin and kinases [476]. L1-CAM plays a role in kidney morphogenesis, lymph node architecture, T cell co-stimulation, neurohistogenesis, and homotypic interaction, and thus plays an important role in nervous system development, including neuronal migration, cell to cell adhesion, and differentiation [408,435,477]. NCAM-1-transfected cells, depending on polysialylation, alter their L1-CAM expression. This aspect hints to the directive role of NCAM-1 towards L1-CAM expression [345,478]. The complex formation between polysialylated NCAM-1 variants and L1-CAM in cis is the basis of synaptic plasticity, memory formation, and damage repair in the central nervous system (CNS) [55,461,479]. An NCAMPSA/L1-CAM or NRP-2 PSA/L1-CAM interaction may significantly influence the role of L1-CAM regarding the tumor cells’ adhesion or migration [480]. In summary, regulation of L1-CAM, E-cadherin, and ICAM-1 by re-expressed NCAM-1 PSA, NRP-2 PSA, or Syn-CAM PSA plays a crucial role in the immune escape mechanism of advanced carcinomas.

3.15. Perspectives and Therapeutic Potential

Conventional cancer therapies such as surgery, chemotherapy, and radiotherapy failed to cure most types of cancer, especially when the disease has reached a metastatic state. Therefore, new cancer treatment methods are desperately needed. One of these could be immunotherapy, which promotes the activation of the immune system against demasked tumor cells [481,482].
The pathophysiological mechanisms outlined above suggest a treatment strategy that will interfere with tumor cell properties resulting from hyper-sialylation. This concept can be realized by reducing the density of sialic acid residues present on tumor cells. Reduced expression of PSA on cancer cells will result in the demasking of cell surface glycans characterized by sulfated oligosaccharides (e.g., the HNK-1 epitope) [483], which can act as tumor suppressors [484] and thus support a therapeutic strategy that is based on a sensitization of cancers to the immune system [248].
For this purpose, multifunctional antibodies can be used, which on the one hand mark re-expressed oligo- or polysialylated carrier proteins on tumor cells [356] and link them to infiltrating effector cells or oncolytic viruses and also show sialidase activity for removal of sialic acid residues, or carry a lectin-like peptide for neutralizing the anti-adhesive properties of sialic acid residues [8]. On the other hand, nanoparticles/vesicles are equipped with a tumor targeting device and are loaded with factors, which knock down components of the process leading to oligo- or poly-sialylation. Finally, viral vectors infecting tumor cells can be used, which cause re-expression of sialidases or knockdown of oligo/poly-sialic acid transferases [248].
In our opinion, the most effective method against the immune escape of tumor cells is based on reducing the link between carrier molecules and sialic acids. This may be achieved by blocking sialylation via either directly suppressing the activity of sialic acid transferases, or with competitive inhibition by presenting an alternative substrate competing for the same binding site. This treatment does not only interfere with a single protein or pathway, but targets a mechanism, which is common to metastasizing tumor cells and immune escape.

4. Conclusions

Sialic acids (neuraminic acids), which are located at outermost positions of carbohydrate chains linked to specific glycoproteins or glycolipids play a central role in immune regulation. In addition, changes in protein sialylation of cell surface are one of the most important characteristics of tumor cells. The most important form is formation of poly-sialic acid chains (α-2,8-linkages), which are attached to carbohydrate scaffolds of cell adhesion molecules by ST8Sia II and ST8Sia IV sialyltransferases. This characteristic property correlates with an aggressive clinical phenotype and endows them with multiple roles in biological processes that underlie all steps of cancer progression, including regulation of cell-cell and/or cell-extracellular matrix interactions, as well as increased proliferation, migration, reduced apoptosis rate of tumor cells, angiogenesis and metastasis.
Furthermore, upregulation of poly- and oligosialylated carrier proteins triggers down-regulation of important adhesion molecules as e.g., of ICAM-1 and E-cadherin. The respective tumor cells gain stem cell like properties including increased migratory and metastatic potential as well as drug resistance. They are masked towards effector cells (immune escape) and become resistant to anticancer treatment because of reduced drug uptake as well as loss of apoptosis induction.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cancers13205203/s1, File S1: Information for PRIZMA search boxes.

Author Contributions

M.J. designed the review and established the concept. M.J., M.S.M. and M.R.B. wrote the review. M.G. and A.K. supervised the PRIZMA search and updated the literature, M.J. designed the figures and F.M. assisted in this process, Proof reading was done by all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors thank Katharina Marstaller (Medical Faculty, Karl-Ruprechts University, Heidelberg) for intensive corrections of the English language and Steffen Schmitt (Core Facility ‘Flow Cytometry’, DKFZ) for carefully reading the manuscript. They also thank Andrea Heppert for her invaluable help regarding the PubMed search.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lanitis, E.; Dangaj, D.; Irving, M.; Coukos, G. Mechanisms regulating T-cell infiltration and activity in solid tumors. Ann. Oncol. 2017, 28, xii18–xii32. [Google Scholar] [CrossRef]
  2. Gajewski, T.F.; Schreiber, H.; Fu, Y.X. Innate and adaptive immune cells in the tumor microenvironment. Nat. Immunol. 2013, 14, 1014–1022. [Google Scholar] [CrossRef] [Green Version]
  3. Bull, C.; Boltje, T.J.; Balneger, N.; Weischer, S.M.; Wassink, M.; van Gemst, J.J.; Bloemendal, V.R.; Boon, L.; van der Vlag, J.; Heise, T.; et al. Sialic Acid Blockade Suppresses Tumor Growth by Enhancing T-cell-Mediated Tumor Immunity. Cancer Res. 2018, 78, 3574–3588. [Google Scholar] [CrossRef] [Green Version]
  4. Sun, L.; Middleton, D.R.; Wantuch, P.L.; Ozdilek, A.; Avci, F.Y. Carbohydrates as T-cell antigens with implications in health and disease. Glycobiology 2016, 26, 1029–1040. [Google Scholar] [CrossRef]
  5. Rabinovich, G.A.; van Kooyk, Y.; Cobb, B.A. Glycobiology of immune responses. Ann. N. Y. Acad. Sci. 2012, 1253, 1–15. [Google Scholar] [CrossRef] [PubMed]
  6. Reily, C.; Stewart, T.J.; Renfrow, M.B.; Novak, J. Glycosylation in health and disease. Nat. Rev. Nephrol. 2019, 15, 346–366. [Google Scholar] [CrossRef]
  7. Tian, Y.; Esteva, F.J.; Song, J.; Zhang, H. Altered expression of sialylated glycoproteins in breast cancer using hydrazide chemistry and mass spectrometry. Mol. Cell. Proteom. 2012, 11, M111.011403. [Google Scholar] [CrossRef] [Green Version]
  8. Colley, K.J.; Kitajima, K.; Sato, C. Polysialic acid: Biosynthesis, novel functions and applications. Crit. Rev. Biochem. Mol. Biol. 2014, 49, 498–532. [Google Scholar] [CrossRef]
  9. Rodrigues, E.; Macauley, M.S. Hypersialylation in Cancer: Modulation of Inflammation and Therapeutic Opportunities. Cancers 2018, 10, 207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. De Freitas Junior, J.C.; Carvalho, S.; Dias, A.M.; Oliveira, P.; Cabral, J.; Seruca, R.; Oliveira, C.; Morgado-Diaz, J.A.; Reis, C.A.; Pinho, S.S. Insulin/IGF-I signaling pathways enhances tumor cell invasion through bisecting GlcNAc N-glycans modulation. an interplay with E-cadherin. PLoS ONE 2013, 8, e81579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Munkley, J.; Scott, E. Targeting Aberrant Sialylation to Treat Cancer. Medicines 2019, 6, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Swindall, A.F.; Bellis, S.L. Sialylation of the Fas death receptor by ST6Gal-I provides protection against Fas-mediated apoptosis in colon carcinoma cells. J. Biol. Chem. 2011, 286, 22982–22990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Varki, A. Multiple changes in sialic acid biology during human evolution. Glycoconj. J. 2009, 26, 231–245. [Google Scholar] [CrossRef] [PubMed]
  14. Varki, A. Biological roles of glycans. Glycobiology 2017, 27, 3–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Cavallaro, U.; Dejana, E. Adhesion molecule signalling: Not always a sticky business. Nat. Rev. Mol. Cell Biol. 2011, 12, 189–197. [Google Scholar] [CrossRef]
  16. Pinho, S.S.; Figueiredo, J.; Cabral, J.; Carvalho, S.; Dourado, J.; Magalhaes, A.; Gartner, F.; Mendonfa, A.M.; Isaji, T.; Gu, J.; et al. E-cadherin and adherens-junctions stability in gastric carcinoma: Functional implications of glycosyltransferases involving N-glycan branching biosynthesis, N-acetylglucosaminyltransferases III and V. BBA-Gen. Subj. 2013, 1830, 2690–2700. [Google Scholar] [CrossRef]
  17. Mallard, B.W.; Tiralongo, J. Cancer stem cell marker glycosylation: Nature, function and significance. Glycoconj. J. 2017, 34, 441–452. [Google Scholar] [CrossRef]
  18. Ferreira, I.G.; Pucci, M.; Venturi, G.; Malagolini, N.; Chiricolo, M.; Dall’Olio, F. Glycosylation as a Main Regulator of Growth and Death Factor Receptors Signaling. Int. J. Mol. Sci. 2018, 19, 580. [Google Scholar] [CrossRef] [Green Version]
  19. Very, N.; Lefebvre, T.; El Yazidi-Belkoura, I. Drug resistance related to aberrant glycosylation in colorectal cancer. Oncotarget 2018, 9, 1380–1402. [Google Scholar] [CrossRef] [Green Version]
  20. Fossella, F.; McCann, J.; Tolcher, A.; Xie, H.; Hwang, L.L.; Carr, C.; Berg, K.; Fram, R. Phase II trial of BB-10901 (huN901-DM1) given weekly for four consecutive weeks every 6 weeks in patients with relapsed SCLC and CD56-positive small cell carcinoma. J. Clin. Oncol. 2005, 23 (Suppl. 16), 7159. [Google Scholar] [CrossRef]
  21. Laubli, H.; Borsig, L. Altered Cell Adhesion and Glycosylation Promote Cancer Immune Suppression and Metastasis. Front. Immunol. 2019, 10, 2120. [Google Scholar] [CrossRef] [Green Version]
  22. Bassaganas, S.; Carvalho, S.; Dias, A.M.; Perez-Garay, M.; Ortiz, M.R.; Figueras, J.; Reis, C.A.; Pinho, S.S.; Peracaula, R. Pancreatic cancer cell glycosylation regulates cell adhesion and invasion through the modulation of alpha2beta1 integrin and E-cadherin function. PLoS ONE 2014, 9, e98595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ohtsubo, K.; Marth, J.D. Glycosylation in cellular mechanisms of health and disease. Cell 2006, 126, 855–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Varki, A. Glycan-based interactions involving vertebrate sialic-acid-recognizing proteins. Nature 2007, 446, 1023–1029. [Google Scholar] [CrossRef] [PubMed]
  25. Scott, E.; Munkley, J. Glycans as Biomarkers in Prostate Cancer. Int. J. Mol. Sci. 2019, 20, 1389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Sola, R.J.; Griebenow, K. Effects of glycosylation on the stability of protein pharmaceuticals. J. Pharm. Sci. 2009, 98, 1223–1245. [Google Scholar] [CrossRef] [Green Version]
  27. Pearce, O.M.; Laubli, H. Sialic acids in cancer biology and immunity. Glycobiology 2016, 26, 111–128. [Google Scholar] [CrossRef] [Green Version]
  28. Zhang, Q.; Li, Z.; Wang, Y.; Zheng, Q.; Li, J. Mass spectrometry for protein sialoglycosylation. Mass Spectrom. Rev. 2018, 37, 652–680. [Google Scholar] [CrossRef]
  29. Liu, Y.C.; Yen, H.Y.; Chen, C.Y.; Chen, C.H.; Cheng, P.F.; Juan, Y.H.; Chen, C.H.; Khoo, K.H.; Yu, C.J.; Yang, P.C.; et al. Sialylation and fucosylation of epidermal growth factor receptor suppress its dimerization and activation in lung cancer cells. Proc. Natl. Acad. Sci. USA 2011, 108, 11332–11337. [Google Scholar] [CrossRef] [Green Version]
  30. Lenman, A.; Liaci, A.M.; Liu, Y.; Frangsmyr, L.; Frank, M.; Blaum, B.S.; Chai, W.; Podgorski, I.I.; Harrach, B.; Benko, M.; et al. Polysialic acid is a cellular receptor for human adenovirus 52. Proc. Natl. Acad. Sci. USA 2018, 115, E4264–E4273. [Google Scholar] [CrossRef] [Green Version]
  31. Tong, J.; Fu, Y.; Meng, F.; Kruger, N.; Valentin-Weigand, P.; Herrler, G. The Sialic Acid Binding Activity of Human Parainfluenza Virus 3 and Mumps Virus Glycoproteins Enhances the Adherence of Group B Streptococci to HEp-2 Cells. Front. Cell Infect. Mi. 2018, 8, 280. [Google Scholar] [CrossRef] [Green Version]
  32. Zlatina, K.; Galuska, S.P. Polysialic Acid Modulates Only the Antimicrobial Properties of Distinct Histones. ACS Omega 2019, 4, 1601–1610. [Google Scholar] [CrossRef]
  33. Freiberger, F.; Claus, H.; Gunzel, A.; Oltmann-Norden, I.; Vionnet, J.; Muhlenhoff, M.; Vogel, U.; Vann, W.F.; Gerardy-Schahn, R.; Stummeyer, K. Biochemical characterization of a Neisseria meningitidis polysialyltransferase reveals novel functional motifs in bacterial sialyltransferases. Mol. Microbiol. 2007, 65, 1258–1275. [Google Scholar] [CrossRef] [Green Version]
  34. Yu, H.; Chen, X. Aldolase-catalyzed synthesis of beta-D-galp-(1-->9)-D-KDN: A novel acceptor for sialyltransferases. Org. Lett. 2006, 8, 2393–2396. [Google Scholar] [CrossRef]
  35. Ghosh, S. Sialic acid and biology of life: An introduction. In Sialic Acids and Sialoglycoconjugates in the Biology of Life, Health and Disease; Academic Press: Cambridge, MA, USA, 2020; Volume 1. [Google Scholar]
  36. Kooner, A.S.; Yu, H.; Chen, X. Synthesis of N-Glycolylneuraminic Acid (Neu5Gc) and Its Glycosides. Front. Immunol. 2019, 10, 2004. [Google Scholar] [CrossRef]
  37. Schauer, R. Sialic acids as regulators of molecular and cellular interactions. Curr. Opin. Struc. Biol. 2009, 19, 507–514. [Google Scholar] [CrossRef] [PubMed]
  38. Schauer, R.; Kamerling, J.P. Exploration of the Sialic Acid World. Adv. Carbohydr. Chem. Biochem. 2018, 75, 1–213. [Google Scholar] [CrossRef] [PubMed]
  39. Hawsawi, M. Exploring the Scope and Limitations of the Oxidative Deamination of N-Acetyl Neuraminic Acid; Wayne State University: Detroit, MI, USA, 2020. [Google Scholar]
  40. Zhang, R.; Loers, G.; Schachner, M.; Boelens, R.; Wienk, H.; Siebert, S.; Eckert, T.; Kraan, S.; Rojas-Macias, M.A.; Lutteke, T.; et al. Molecular Basis of the Receptor Interactions of Polysialic Acid (polySia), polySia Mimetics, and Sulfated Polysaccharides. ChemMedChem 2016, 11, 990–1002. [Google Scholar] [CrossRef]
  41. Zhou, G.P.; Huang, R.B.; Troy, F.A., 2nd. 3D structural conformation and functional domains of polysialyltransferase ST8Sia IV required for polysialylation of neural cell adhesion molecules. Protein Pept. Lett. 2015, 22, 137–148. [Google Scholar] [CrossRef]
  42. Teoh, S.T.; Ogrodzinski, M.P.; Ross, C.; Hunter, K.W.; Lunt, S.Y. Sialic Acid Metabolism: A Key Player in Breast Cancer Metastasis Revealed by Metabolomics. Front. Oncol. 2018, 8, 174. [Google Scholar] [CrossRef]
  43. Muhlenhoff, M.; Rollenhagen, M.; Werneburg, S.; Gerardy-Schahn, R.; Hildebrandt, H. Polysialic acid: Versatile modification of NCAM, SynCAM 1 and neuropilin-2. Neurochem. Res. 2013, 38, 1134–1143. [Google Scholar] [CrossRef] [PubMed]
  44. Elkashef, S.M.; Allison, S.J.; Sadiq, M.; Basheer, H.A.; Ribeiro Morais, G.; Loadman, P.M.; Pors, K.; Falconer, R.A. Polysialic acid sustains cancer cell survival and migratory capacity in a hypoxic environment. Sci. Rep. 2016, 6, 33026. [Google Scholar] [CrossRef] [Green Version]
  45. Falconer, R.A.; Errington, R.J.; Shnyder, S.D.; Smith, P.J.; Patterson, L.H. Polysialyltransferase: A new target in metastatic cancer. Curr. Cancer Drug. Tar. 2012, 12, 925–939. [Google Scholar] [CrossRef]
  46. Kronewitter, S.R.; Marginean, I.; Cox, J.T.; Zhao, R.; Hagler, C.D.; Shukla, A.K.; Carlson, T.S.; Adkins, J.N.; Camp, D.G., 2nd; Moore, R.J.; et al. Polysialylated N-glycans identified in human serum through combined developments in sample preparation, separations, and electrospray ionization-mass spectrometry. Anal. Chem. 2014, 86, 8700–8710. [Google Scholar] [CrossRef] [Green Version]
  47. RodrIguez, E.; Schetters, S.T.T.; van Kooyk, Y. The tumour glyco-code as a novel immune checkpoint for immunotherapy. Nat. Rev. Immunol. 2018, 18, 204–211. [Google Scholar] [CrossRef] [PubMed]
  48. Stowell, S.R.; Ju, T.; Cummings, R.D. Protein glycosylation in cancer. Annu. Rev. Pathol. 2015, 10, 473–510. [Google Scholar] [CrossRef] [Green Version]
  49. Vajaria, B.N.; Begum, R.; Patel, P.S. Clinical Significance of Glycosylation Changes in Oral Cancer. Glycobiology 2015, 25, 1300–1301. [Google Scholar]
  50. Munkley, J.; Elliott, D.J. Hallmarks of glycosylation in cancer. Oncotarget 2016, 7, 35478–35489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Blanas, A.; Sahasrabudhe, N.M.; Rodriguez, E.; van Kooyk, Y.; van Vliet, S.J. Fucosylated Antigens in Cancer: An Alliance toward Tumor Progression, Metastasis, and Resistance to Chemotherapy. Front. Oncol. 2018, 8, 39. [Google Scholar] [CrossRef]
  52. Gong, L.; Zhou, X.; Yang, J.; Jiang, Y.; Yang, H. Effects of the regulation of polysialyltransferase ST8SiaII on the invasiveness and metastasis of small cell lung cancer cells. Oncol. Rep. 2017, 37, 131–138. [Google Scholar] [CrossRef] [Green Version]
  53. Sato, C.; Kitajima, K. Sialic Acids in Neurology. In Advances in Carbohydrate Chemistry and Biochemistry; Academic Press: Cambridge, MA, USA, 2019; Volume 76, pp. 1–64. [Google Scholar]
  54. Bruses, J.L.; Rutishauser, U. Roles, regulation, and mechanism of polysialic acid function during neural development. Biochimie 2001, 83, 635–643. [Google Scholar] [CrossRef]
  55. Bonfanti, L. PSA-NCAM in mammalian structural plasticity and neurogenesis. Prog. Neurobiol. 2006, 80, 129–164. [Google Scholar] [CrossRef]
  56. Burgess, A.; Wainwright, S.R.; Shihabuddin, L.S.; Rutishauser, U.; Seki, T.; Aubert, I. Polysialic acid regulates the clustering, migration, and neuronal differentiation of progenitor cells in the adult hippocampus. Dev. Neurobiol. 2008, 68, 1580–1590. [Google Scholar] [CrossRef] [PubMed]
  57. Rutishauser, U. Polysialic acid in the plasticity of the developing and adult vertebrate nervous system. Nat. Rev. Neurosci. 2008, 9, 26–35. [Google Scholar] [CrossRef] [PubMed]
  58. Rutishauser, U.; El Maarouf, A. Polysialic Acid in the CNS: Plasticity and Repair. Glycobiology 2008, 18, 944. [Google Scholar]
  59. Mehrabian, M.; Hildebrandt, H.; Schmitt-Ulms, G. NCAM1 Polysialylation: The Prion Protein’s Elusive Reason for Being? ASN Neuro 2016, 8, 1759091416679074. [Google Scholar] [CrossRef] [Green Version]
  60. Boutin, C.; Schmitz, B.; Cremer, H.; Diestel, S. NCAM expression induces neurogenesis in vivo. Eur. J. Neurosci. 2009, 30, 1209–1218. [Google Scholar] [CrossRef] [PubMed]
  61. Gibson, N.J. Cell adhesion molecules in context: CAM function depends on the neighborhood. Cell Adh. Migr. 2011, 5, 48–51. [Google Scholar] [CrossRef] [Green Version]
  62. Tang, H.; Chang, H.; Dong, Y.; Guo, L.; Shi, X.; Wu, Y.; Huang, Y.; He, Y. Architecture of cell-cell adhesion mediated by sidekicks. Proc. Natl. Acad. Sci. USA 2018, 115, 9246–9251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Villringer, S.; Madl, J.; Sych, T.; Manner, C.; Imberty, A.; Romer, W. Lectin-mediated protocell crosslinking to mimic cell-cell junctions and adhesion. Sci. Rep. 2018, 8, 1932. [Google Scholar] [CrossRef]
  64. Baldwin, K.T.; Eroglu, C. Molecular mechanisms of astrocyte-induced synaptogenesis. Curr. Opin. Neurobiol. 2017, 45, 113–120. [Google Scholar] [CrossRef] [PubMed]
  65. Bhide, G.P. Biophysical and Biochemical Determinants of Protein-Specific Polysialylation. Ph.D Thesis, University of Illinois at Chicago, Chicago, IL, USA, 2017. [Google Scholar]
  66. Kiermaier, E.; Moussion, C.; Veldkamp, C.T.; Gerardy-Schahn, R.; de Vries, I.; Williams, L.G.; Chaffee, G.R.; Phillips, A.J.; Freiberger, F.; Imre, R.; et al. Polysialylation controls dendritic cell trafficking by regulating chemokine recognition. Science 2016, 351, 186–190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Werneburg, S.; Buettner, F.F.; Erben, L.; Mathews, M.; Neumann, H.; Muhlenhoff, M.; Hildebrandt, H. Polysialylation and lipopolysaccharide-induced shedding of E-selectin ligand-1 and neuropilin-2 by microglia and THP-1 macrophages. Glia 2016, 64, 1314–1330. [Google Scholar] [CrossRef] [PubMed]
  68. Fujioka, Y.; Nishide, S.; Ose, T.; Suzuki, T.; Kato, I.; Fukuhara, H.; Fujioka, M.; Horiuchi, K.; Satoh, A.O.; Nepal, P. A sialylated voltage-dependent Ca2+ channel binds hemagglutinin and mediates influenza a virus entry into mammalian cells. Cell Host Microbe 2018, 23, 809–818.e805. [Google Scholar] [CrossRef] [Green Version]
  69. Jing, X.; Liang, H.; Hao, C.; Yang, X.; Cui, X. Overexpression of MUC1 predicts poor prognosis in patients with breast cancer. Oncol. Rep. 2019, 41, 801–810. [Google Scholar] [CrossRef]
  70. Yabe, U.; Sato, C.; Matsuda, T.; Kitajima, K. Polysialic acid in human milk. CD36 is a new member of mammalian polysialic acid-containing glycoprotein. J. Biol. Chem. 2003, 278, 13875–13880. [Google Scholar] [CrossRef] [Green Version]
  71. Angata, K.; Chan, D.; Thibault, J.; Fukuda, M. Molecular dissection of the ST8Sia IV polysialyltransferase. Distinct domains are required for neural cell adhesion molecule recognition and polysialylation. J. Biol. Chem. 2004, 279, 25883–25890. [Google Scholar] [CrossRef] [Green Version]
  72. Cao, L.; Wang, X.; Yang, J.; Guo, J.; Li, X.; Yang, X.; Tan, Z.; Guan, F. NCAM and attached polysialic acid affect behaviors of breast epithelial cells through differential signaling pathways. Res. Square 2020. (Under Revision). [Google Scholar]
  73. Rollenhagen, M.; Kuckuck, S.; Ulm, C.; Hartmann, M.; Galuska, S.P.; Geyer, R.; Geyer, H.; Muhlenhoff, M. Polysialylation of the synaptic cell adhesion molecule 1 (SynCAM 1) depends exclusively on the polysialyltransferase ST8SiaII in vivo. J. Biol. Chem. 2012, 287, 35170–35180. [Google Scholar] [CrossRef] [Green Version]
  74. Ellis, L.M. The role of neuropilins in cancer. Mol. Cancer Ther. 2006, 5, 1099–1107. [Google Scholar] [CrossRef] [Green Version]
  75. Pellet-Many, C.; Frankel, P.; Jia, H.; Zachary, I. Neuropilins: Structure, function and role in disease. Biochem. J. 2008, 411, 211–226. [Google Scholar] [CrossRef] [Green Version]
  76. Bhide, G.P.; Fernandes, N.R.; Colley, K.J. Sequence Requirements for Neuropilin-2 Recognition by ST8SiaIV and Polysialylation of Its O-Glycans. J. Biol. Chem. 2016, 291, 9444–9457. [Google Scholar] [CrossRef] [Green Version]
  77. Sulpice, E.; Plouet, J.; Berge, M.; Allanic, D.; Tobelem, G.; Merkulova-Rainon, T. Neuropilin-1 and neuropilin-2 act as coreceptors, potentiating proangiogenic activity. Blood 2008, 111, 2036–2045. [Google Scholar] [CrossRef]
  78. Grun, D.; Adhikary, G.; Eckert, R.L. VEGF-A acts via neuropilin-1 to enhance epidermal cancer stem cell survival and formation of aggressive and highly vascularized tumors. Oncogene 2016, 35, 4379–4387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Winther, M.; Berezin, V.; Walmod, P.S. NCAM2/OCAM/RNCAM: Cell adhesion molecule with a role in neuronal compartmentalization. Int. J. Biochem. Cell Biol. 2012, 44, 441–446. [Google Scholar] [CrossRef] [PubMed]
  80. Leshchyns’ka, I.; Liew, H.T.; Shepherd, C.; Halliday, G.M.; Stevens, C.H.; Ke, Y.D.; Ittner, L.M.; Sytnyk, V. Abeta-dependent reduction of NCAM2-mediated synaptic adhesion contributes to synapse loss in Alzheimer’s disease. Nat. Commun. 2015, 6, 8836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Parcerisas, A.; Pujadas, L.; Ortega-Gasco, A.; Perello-Amoros, B.; Viais, R.; Hino, K.; Figueiro-Silva, J.; La Torre, A.; Trullas, R.; Simo, S.; et al. NCAM2 Regulates Dendritic and Axonal Differentiation through the Cytoskeletal Proteins MAP2 and 14-3-3. Cereb. Cortex 2020, 30, 3781–3799. [Google Scholar] [CrossRef] [PubMed]
  82. Rasmussen, K.K.; Falkesgaard, M.H.; Winther, M.; Roed, N.K.; Quistgaard, C.L.; Teisen, M.N.; Edslev, S.M.; Petersen, D.L.; Aljubouri, A.; Christensen, C.; et al. NCAM2 Fibronectin type-III domains form a rigid structure that binds and activates the Fibroblast Growth Factor Receptor. Sci. Rep. 2018, 8, 8957. [Google Scholar] [CrossRef] [PubMed]
  83. Sheng, L.; Leshchyns’ka, I.; Sytnyk, V. Neural Cell Adhesion Molecule 2 (NCAM2)-Induced c-Src-Dependent Propagation of Submembrane Ca2+ Spikes Along Dendrites Inhibits Synapse Maturation. Cereb. Cortex 2019, 29, 1439–1459. [Google Scholar] [CrossRef]
  84. Kulahin, N.; Walmod, P.S. The neural cell adhesion molecule NCAM2/OCAM/RNCAM, a close relative to NCAM. Adv. Exp. Med. Biol. 2010, 663, 403–420. [Google Scholar] [CrossRef]
  85. Kim, W.; Watanabe, H.; Lomoio, S.; Tesco, G. Spatiotemporal processing of neural cell adhesion molecules 1 and 2 by BACE1 in vivo. J. Biol. Chem. 2021, 296, 100372. [Google Scholar] [CrossRef]
  86. Niland, S.; Eble, J.A. Neuropilins in the context of tumor vasculature. Int. J. Mol. Sci. 2019, 20, 639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Neufeld, G.; Kessler, O. The semaphorins: Versatile regulators of tumour progression and tumour angiogenesis. Nat. Rev. Cancer 2008, 8, 632–645. [Google Scholar] [CrossRef] [PubMed]
  88. Gaur, P.; Bielenberg, D.R.; Samuel, S.; Bose, D.; Zhou, Y.; Gray, M.J.; Dallas, N.A.; Fan, F.; Xia, L.; Lu, J.; et al. Role of class 3 semaphorins and their receptors in tumor growth and angiogenesis. Clin. Cancer Res. 2009, 15, 6763–6770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Khare, N.; Fascetti, N.; DaRocha, S.; Chiquet-Ehrismann, R.; Baumgartner, S. Expression patterns of two new members of the Semaphorin family in Drosophila suggest early functions during embryogenesis. Mech. Dev. 2000, 91, 393–397. [Google Scholar] [CrossRef]
  90. Xiao, W. Class 5 Semaphorins Mediate Synapse Elimination and Activity-Dependent Synaptic Plasticity in Hippocampal Neurons. Ph.D Thesis, University of British Columbia, Vancouver, BC, Canada, 2017. [Google Scholar]
  91. Mucka, P.; Levonyak, N.; Geretti, E.; Zwaans, B.M.M.; Li, X.; Adini, I.; Klagsbrun, M.; Adam, R.M.; Bielenberg, D.R. Inflammation and Lymphedema Are Exacerbated and Prolonged by Neuropilin 2 Deficiency. Am. J. Pathol. 2016, 186, 2803–2812. [Google Scholar] [CrossRef] [Green Version]
  92. Li, X.; Chen, Q.; Yin, D.; Shi, S.; Yu, L.; Zhou, S.; Chen, E.; Zhou, Z.; Shi, Y.; Fan, J.; et al. Novel role of semaphorin 3A in the growth and progression of hepatocellular carcinoma. Oncol. Rep. 2017, 37, 3313–3320. [Google Scholar] [CrossRef] [Green Version]
  93. Nakamura, F.; Kalb, R.G.; Strittmatter, S.M. Molecular basis of semaphorin-mediated axon guidance. J. Neurobiol. 2000, 44, 219–229. [Google Scholar] [CrossRef]
  94. Tamagnone, L.; Comoglio, P.M. Signalling by semaphorin receptors: Cell guidance and beyond. Trends Cell Biol. 2000, 10, 377–383. [Google Scholar] [CrossRef]
  95. Iragavarapu-Charyulu, V.; Wojcikiewicz, E.; Urdaneta, A. Semaphorins in angiogenesis and autoimmune diseases: Therapeutic targets? Front. Immunol. 2020, 11, 346. [Google Scholar] [CrossRef]
  96. Geretti, E.; Klagsbrun, M. Neuropilins: Novel targets for anti-angiogenesis therapies. Cell Adh. Migr. 2007, 1, 56–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Christensen, C.; Ambartsumian, N.; Gilestro, G.; Thomsen, B.; Comoglio, P.; Tamagnone, L.; Guldberg, P.; Lukanidin, E. Proteolytic processing converts the repelling signal Sema3E into an inducer of invasive growth and lung metastasis. Cancer Res. 2005, 65, 6167–6177. [Google Scholar] [CrossRef] [Green Version]
  98. Tian, L.; Coletti, D.; Li, Z.L. Angiotensin II induces the exocytosis of galectin-3 via integrin alphav/AKT/NF-kappaB signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7183. [Google Scholar] [CrossRef] [PubMed]
  99. Bielenberg, D.R.; Hida, Y.; Shimizu, A.; Kaipainen, A.; Kreuter, M.; Kim, C.C.; Klagsbrun, M. Semaphorin 3F, a chemorepulsant for endothelial cells, induces a poorly vascularized, encapsulated, nonmetastatic tumor phenotype. J. Clin. Investig. 2004, 114, 1260–1271. [Google Scholar] [CrossRef] [Green Version]
  100. Osada, R.; Horiuchi, A.; Kikuchi, N.; Ohira, S.; Ota, M.; Katsuyama, Y.; Konishi, I. Expression of semaphorins, vascular endothelial growth factor, and their common receptor neuropilins and alleic loss of semaphorin locus in epithelial ovarian neoplasms: Increased ratio of vascular endothelial growth factor to semaphorin is a poor prognostic factor in ovarian carcinomas. Hum. Pathol. 2006, 37, 1414–1425. [Google Scholar] [CrossRef] [PubMed]
  101. Gu, C.; Rodriguez, E.R.; Reimert, D.V.; Shu, T.; Fritzsch, B.; Richards, L.J.; Kolodkin, A.L.; Ginty, D.D. Neuropilin-1 conveys semaphorin and VEGF signaling during neural and cardiovascular development. Dev. Cell 2003, 5, 45–57. [Google Scholar] [CrossRef] [Green Version]
  102. Giger, R.J.; Cloutier, J.-F.; Sahay, A.; Prinjha, R.K.; Levengood, D.V.; Moore, S.E.; Pickering, S.; Simmons, D.; Rastan, S.; Walsh, F.S.; et al. Neuropilin-2 Is Required In Vivo for Selective Axon Guidance Responses to Secreted Semaphorins. Neuron 2000, 25, 29–41. [Google Scholar] [CrossRef] [Green Version]
  103. Lantuejoul, S.; Constantin, B.; Drabkin, H.; Brambilla, C.; Roche, J.; Brambilla, E. Expression of VEGF, semaphorin SEMA3F, and their common receptors neuropilins NP1 and NP2 in preinvasive bronchial lesions, lung tumours, and cell lines. J. Pathol. 2003, 200, 336–347. [Google Scholar] [CrossRef]
  104. Wang, X.; Zhang, W.; Cheever, T.; Schwarz, V.; Opperman, K.; Hutter, H.; Koepp, D.; Chen, L. The C. elegans L1CAM homologue LAD-2 functions as a coreceptor in MAB-20/Sema2 mediated axon guidance. J. Cell Biol. 2008, 180, 233–246. [Google Scholar] [CrossRef] [Green Version]
  105. Matkar, P.N.; Jong, E.D.; Ariyagunarajah, R.; Prud’homme, G.J.; Singh, K.K.; Leong-Poi, H. Jack of many trades: Multifaceted role of neuropilins in pancreatic cancer. Cancer Med. 2018, 7, 5036–5046. [Google Scholar] [CrossRef]
  106. Strubl, S.; Schubert, U.; Kühnle, A.; Rebl, A.; Ahmadvand, N.; Fischer, S.; Preissner, K.T.; Galuska, S.P. Polysialic acid is released by human umbilical vein endothelial cells (HUVEC) in vitro. Cell Biosci. 2018, 8, 64. [Google Scholar] [CrossRef]
  107. Grandclement, C.; Pallandre, J.R.; Valmary Degano, S.; Viel, E.; Bouard, A.; Balland, J.; Remy-Martin, J.P.; Simon, B.; Rouleau, A.; Boireau, W.; et al. Neuropilin-2 expression promotes TGF-beta1-mediated epithelial to mesenchymal transition in colorectal cancer cells. PLoS ONE 2011, 6, e20444. [Google Scholar] [CrossRef] [Green Version]
  108. Glinka, Y.; Stoilova, S.; Mohammed, N.; Prud’homme, G.J. Neuropilin-1 exerts co-receptor function for TGF-beta-1 on the membrane of cancer cells and enhances responses to both latent and active TGF-beta. Carcinogenesis 2011, 32, 613–621. [Google Scholar] [CrossRef] [Green Version]
  109. Matsushita, A.; Gotze, T.; Korc, M. Hepatocyte growth factor-mediated cell invasion in pancreatic cancer cells is dependent on neuropilin-1. Cancer Res. 2007, 67, 10309–10316. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, M.; Theis, T.; Kabat, M.; Loers, G.; Agre, L.A.; Schachner, M. Functions of Small Organic Compounds that Mimic the HNK-1 Glycan. Int. J. Mol. Sci. 2020, 21, 7018. [Google Scholar] [CrossRef] [PubMed]
  111. Zhang, H.; Vutskits, L.; Calaora, V.; Durbec, P.; Kiss, J.Z. A role for the polysialic acid-neural cell adhesion molecule in PDGF-induced chemotaxis of oligodendrocyte precursor cells. J. Cell Sci. 2004, 117, 93–103. [Google Scholar] [CrossRef] [Green Version]
  112. Hinek, A.; Bodnaruk, T.D.; Bunda, S.; Wang, Y.; Liu, K. Neuraminidase-1, a subunit of the cell surface elastin receptor, desialylates and functionally inactivates adjacent receptors interacting with the mitogenic growth factors PDGF-BB and IGF-2. Am. J. Pathol. 2008, 173, 1042–1056. [Google Scholar] [CrossRef] [Green Version]
  113. Muhl, L.; Folestad, E.B.; Gladh, H.; Wang, Y.; Moessinger, C.; Jakobsson, L.; Eriksson, U. Neuropilin 1 binds PDGF-D and is a co-receptor in PDGF-D–PDGFRβ signaling. J. Cell Sci. 2017, 130, 1365–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Francavilla, C.; Cattaneo, P.; Berezin, V.; Bock, E.; Ami, D.; de Marco, A.; Christofori, G.; Cavallaro, U. The binding of NCAM to FGFR1 induces a specific cellular response mediated by receptor trafficking. J. Cell Biol. 2009, 187, 1101–1116. [Google Scholar] [CrossRef]
  115. Christensen, C.; Berezin, V.; Bock, E. Neural cell adhesion molecule differentially interacts with isoforms of the fibroblast growth factor receptor. Neuroreport 2011, 22, 727–732. [Google Scholar] [CrossRef] [PubMed]
  116. Ono, S.; Hane, M.; Kitajima, K.; Sato, C. Novel regulation of fibroblast growth factor 2 (FGF2)-mediated cell growth by polysialic acid. J. Biol. Chem. 2012, 287, 3710–3722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Cirovic, S.; Vjestica, J.; Mueller, C.A.; Tatic, S.; Vasiljevic, J.; Milenkovic, S.; Mueller, G.A.; Markovic-Lipkovski, J. NCAM and FGFR1 coexpression and colocalization in renal tumors. Int. J. Clin. Exp. Pathol. 2014, 7, 1402–1414. [Google Scholar]
  118. Kanato, Y.; Kitajima, K.; Sato, C. Direct binding of polysialic acid to a brain-derived neurotrophic factor depends on the degree of polymerization. Glycobiology 2008, 18, 1044–1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Monzo, H.J.; Park, T.I.; Dieriks, B.V.; Jansson, D.; Faull, R.L.; Dragunow, M.; Curtis, M.A. Insulin and IGF1 modulate turnover of polysialylated neural cell adhesion molecule (PSA-NCAM) in a process involving specific extracellular matrix components. J. Neurochem. 2013, 126, 758–770. [Google Scholar] [CrossRef]
  120. Lynch, C.C.; Vargo-Gogola, T.; Martin, M.D.; Fingleton, B.; Crawford, H.C.; Matrisian, L.M. Matrix metalloproteinase 7 mediates mammary epithelial cell tumorigenesis through the ErbB4 receptor. Cancer Res. 2007, 67, 6760–6767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Chen, D.; Wu, Z.; Luo, L.J.; Huang, X.; Qian, W.Q.; Wang, H.; Li, S.H.; Liu, J. E-cadherin maintains the activity of neural stem cells and inhibits the migration. Int. J. Clin. Exp. Pathol. 2015, 8, 14247–14251. [Google Scholar]
  122. Qian, X.; Karpova, T.; Sheppard, A.M.; McNally, J.; Lowy, D.R. E-cadherin-mediated adhesion inhibits ligand-dependent activation of diverse receptor tyrosine kinases. EMBO J. 2004, 23, 1739–1748. [Google Scholar] [CrossRef] [Green Version]
  123. Alajati, A.; Guccini, I.; Pinton, S.; Garcia-Escudero, R.; Bernasocchi, T.; Sarti, M.; Montani, E.; Rinaldi, A.; Montemurro, F.; Catapano, C.; et al. Interaction of CDCP1 with HER2 enhances HER2-driven tumorigenesis and promotes trastuzumab resistance in breast cancer. Cell Rep. 2015, 11, 564–576. [Google Scholar] [CrossRef]
  124. Morath, I.; Jung, C.; Leveque, R.; Linfeng, C.; Toillon, R.A.; Warth, A.; Orian-Rousseau, V. Differential recruitment of CD44 isoforms by ErbB ligands reveals an involvement of CD44 in breast cancer. Oncogene 2018, 37, 1472–1484. [Google Scholar] [CrossRef]
  125. Donier, E.; Gomez-Sanchez, J.A.; Grijota-Martinez, C.; Lakoma, J.; Baars, S.; Garcia-Alonso, L.; Cabedo, H. L1CAM binds ErbB receptors through Ig-like domains coupling cell adhesion and neuregulin signalling. PLoS ONE 2012, 7, e40674. [Google Scholar] [CrossRef] [Green Version]
  126. Kawano, S.; Ikeda, W.; Kishimoto, M.; Ogita, H.; Takai, Y. Silencing of ErbB3/ErbB2 signaling by immunoglobulin-like Necl-2. J. Biol. Chem. 2009, 284, 23793–23805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Resovi, A.; Pinessi, D.; Chiorino, G.; Taraboletti, G. Current understanding of the thrombospondin-1 interactome. Matrix Biol. 2014, 37, 83–91. [Google Scholar] [CrossRef] [PubMed]
  128. Latko, M.; Czyrek, A.; Porebska, N.; Kucinska, M.; Otlewski, J.; Zakrzewska, M.; Opalinski, L. Cross-Talk between Fibroblast Growth Factor Receptors and Other Cell Surface Proteins. Cells 2019, 8, 455. [Google Scholar] [CrossRef] [Green Version]
  129. Cavallaro, U.; Francavilla, C.; Loeffler, S.; Christofori, G. The NCAM/FGFR signaling complex: A novel player in metastatic dissemination. Clin. Exp. Metastas 2007, 24, 241–242. [Google Scholar]
  130. Eggers, K.; Werneburg, S.; Schertzinger, A.; Abeln, M.; Schiff, M.; Scharenberg, M.A.; Burkhardt, H.; Muhlenhoff, M.; Hildebrandt, H. Polysialic acid controls NCAM signals at cell-cell contacts to regulate focal adhesion independent from FGF receptor activity. J. Cell Sci. 2011, 124, 3279–3291. [Google Scholar] [CrossRef] [Green Version]
  131. Zecchini, S.; Bombardelli, L.; Decio, A.; Bianchi, M.; Mazzarol, G.; Sanguineti, F.; Aletti, G.; Maddaluno, L.; Berezin, V.; Bock, E.; et al. The adhesion molecule NCAM promotes ovarian cancer progression via FGFR signalling. EMBO Mol. Med. 2011, 3, 480–494. [Google Scholar] [CrossRef]
  132. Vales, A.; Kondo, R.; Aichberger, K.J.; Mayerhofer, M.; Kainz, B.; Sperr, W.R.; Sillaber, C.; Jager, U.; Valent, P. Myeloid leukemias express a broad spectrum of VEGF receptors including neuropilin-1 (NRP-1) and NRP-2. Leuk. Lymphoma 2007, 48, 1997–2007. [Google Scholar] [CrossRef]
  133. Guo, H.F.; Vander Kooi, C.W. Neuropilin Functions as an Essential Cell Surface Receptor. J. Biol. Chem. 2015, 290, 29120–29126. [Google Scholar] [CrossRef] [Green Version]
  134. Sarabipour, S.; Mac Gabhann, F. VEGF-A121a binding to Neuropilins—A concept revisited. Cell Adh. Migr. 2018, 12, 204–214. [Google Scholar] [CrossRef]
  135. Favier, B.; Alam, A.; Barron, P.; Bonnin, J.; Laboudie, P.; Fons, P.; Mandron, M.; Herault, J.P.; Neufeld, G.; Savi, P.; et al. Neuropilin-2 interacts with VEGFR-2 and VEGFR-3 and promotes human endothelial cell survival and migration. Blood 2006, 108, 1243–1250. [Google Scholar] [CrossRef]
  136. Hamerlik, P.; Lathia, J.D.; Rasmussen, R.; Wu, Q.; Bartkova, J.; Lee, M.; Moudry, P.; Bartek, J., Jr.; Fischer, W.; Lukas, J.; et al. Autocrine VEGF-VEGFR2-Neuropilin-1 signaling promotes glioma stem-like cell viability and tumor growth. J. Exp. Med. 2012, 209, 507–520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Crnic, I.; Strittmatter, K.; Cavallaro, U.; Kopfstein, L.; Jussila, L.; Alitalo, K.; Christofori, G. Loss of neural cell adhesion molecule induces tumor metastasis by up-regulating lymphangiogenesis. Cancer Res. 2004, 64, 8630–8638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Hrgovic, I.; Doll, M.; Pinter, A.; Kaufmann, R.; Kippenberger, S.; Meissner, M. Histone deacetylase inhibitors interfere with angiogenesis by decreasing endothelial VEGFR-2 protein half-life in part via a VE-cadherin-dependent mechanism. Exp. Dermatol. 2017, 26, 194–201. [Google Scholar] [CrossRef]
  139. Lampugnani, M.G.; Orsenigo, F.; Gagliani, M.C.; Tacchetti, C.; Dejana, E. Vascular endothelial cadherin controls VEGFR-2 internalization and signaling from intracellular compartments. J. Cell Biol. 2006, 174, 593–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Shintani, Y.; Takashima, S.; Asano, Y.; Kato, H.; Liao, Y.; Yamazaki, S.; Tsukamoto, O.; Seguchi, O.; Yamamoto, H.; Fukushima, T.; et al. Glycosaminoglycan modification of neuropilin-1 modulates VEGFR2 signaling. EMBO J. 2006, 25, 3045–3055. [Google Scholar] [CrossRef] [Green Version]
  141. Stanton, M.J.; Dutta, S.; Zhang, H.; Polavaram, N.S.; Leontovich, A.A.; Honscheid, P.; Sinicrope, F.A.; Tindall, D.J.; Muders, M.H.; Datta, K. Autophagy control by the VEGF-C/NRP-2 axis in cancer and its implication for treatment resistance. Cancer Res. 2013, 73, 160–171. [Google Scholar] [CrossRef] [Green Version]
  142. Koch, S. Neuropilin signalling in angiogenesis. Biochem. Soc. Trans. 2012, 40, 20–25. [Google Scholar] [CrossRef] [Green Version]
  143. Raimondi, C. Neuropilin-1 enforces extracellular matrix signalling via ABL1 to promote angiogenesis. Biochem. Soc. Trans. 2014, 42, 1429–1434. [Google Scholar] [CrossRef]
  144. Jastrzebski, K.; Zdzalik-Bielecka, D.; Maminska, A.; Kalaidzidis, Y.; Hellberg, C.; Miaczynska, M. Multiple routes of endocytic internalization of PDGFRbeta contribute to PDGF-induced STAT3 signaling. J. Cell Sci. 2017, 130, 577–589. [Google Scholar] [CrossRef] [Green Version]
  145. Ueda, Y.; Kedashiro, S.; Maruoka, M.; Mizutani, K.; Takai, Y. Roles of the third Ig-like domain of Necl-5/PVR and the fifth Ig-like domain of the PDGF receptor in its signaling. Genes Cells 2018, 23, 214–224. [Google Scholar] [CrossRef]
  146. Gopal, S. Syndecans in inflammation at a glance. Front. Immunol. 2020, 11, 227. [Google Scholar] [CrossRef] [Green Version]
  147. Chen, L.; Klass, C.; Woods, A. Syndecan-2 regulates transforming growth factor-beta signaling. J. Biol. Chem. 2004, 279, 15715–15718. [Google Scholar] [CrossRef] [Green Version]
  148. Gomes, C.; Osorio, H.; Pinto, M.T.; Campos, D.; Oliveira, M.J.; Reis, C.A. Expression of ST3GAL4 leads to SLe(x) expression and induces c-Met activation and an invasive phenotype in gastric carcinoma cells. PLoS ONE 2013, 8, e66737. [Google Scholar] [CrossRef]
  149. Horiguchi, H.; Tsujimoto, H.; Shinomiya, N.; Matsumoto, Y.; Sugasawa, H.; Yamori, T.; Miyazaki, H.; Saitoh, D.; Kishi, Y.; Ueno, H. A Potential Role of Adhesion Molecules on Lung Metastasis Enhanced by Local Inflammation. Anticancer Res. 2020, 40, 6171–6178. [Google Scholar] [CrossRef]
  150. Zhang, X.; Shao, S.; Li, L. Characterization of Class-3 Semaphorin Receptors, Neuropilins and Plexins, as Therapeutic Targets in a Pan-Cancer Study. Cancers 2020, 12, 1816. [Google Scholar] [CrossRef]
  151. Janssen, B.J.; Malinauskas, T.; Weir, G.A.; Cader, M.Z.; Siebold, C.; Jones, E.Y. Neuropilins lock secreted semaphorins onto plexins in a ternary signaling complex. Nat. Struct. Mol. Biol 2012, 19, 1293–1299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Pinho, S.S.; Reis, C.A. Glycosylation in cancer: Mechanisms and clinical implications. Nat. Rev. Cancer 2015, 15, 540–555. [Google Scholar] [CrossRef] [PubMed]
  153. Kleene, R.; Schachner, M. Glycans and neural cell interactions. Nat. Rev. Neurosci. 2004, 5, 195–208. [Google Scholar] [CrossRef] [PubMed]
  154. Xie, Y.; Chen, S.; Li, Q.; Sheng, Y.; Alvarez, M.R.; Reyes, J.; Xu, G.; Solakyildirim, K.; Lebrilla, C.B. Glycan–protein cross-linking mass spectrometry reveals sialic acid-mediated protein networks on cell surfaces. Chem. Sci. 2021, 12, 8767–8777. [Google Scholar] [CrossRef]
  155. Buhe, V.; Loisel, S.; Pers, J.O.; Le Ster, K.; Berthou, C.; Youinou, P. Updating the physiology, exploration and disease relevance of complement factor H. Int. J. Immunopath. Pharmacol. 2010, 23, 397–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Karlstetter, M.; Kopatz, J.; Aslanidis, A.; Shahraz, A.; Caramoy, A.; Linnartz-Gerlach, B.; Lin, Y.; Luckoff, A.; Fauser, S.; Duker, K.; et al. Polysialic acid blocks mononuclear phagocyte reactivity, inhibits complement activation, and protects from vascular damage in the retina. EMBO Mol. Med. 2017, 9, 154–166. [Google Scholar] [CrossRef]
  157. Meesmann, H.M.; Fehr, E.M.; Kierschke, S.; Herrmann, M.; Bilyy, R.; Heyder, P.; Blank, N.; Krienke, S.; Lorenz, H.M.; Schiller, M. Decrease of sialic acid residues as an eat-me signal on the surface of apoptotic lymphocytes. J. Cell Sci. 2010, 123, 3347–3356. [Google Scholar] [CrossRef] [Green Version]
  158. Gohari, B.; Abu-Zahra, N. Polyethersulfone Membranes Prepared with 3-Aminopropyltriethoxysilane Modified Alumina Nanoparticles for Cu(II) Removal from Water. ACS Omega 2018, 3, 10154–10162. [Google Scholar] [CrossRef] [Green Version]
  159. Jarahian, M.; Watzl, C.; Fournier, P.; Arnold, A.; Djandji, D.; Zahedi, S.; Cerwenka, A.; Paschen, A.; Schirrmacher, V.; Momburg, F. Activation of natural killer cells by newcastle disease virus hemagglutinin-neuraminidase. J. Virol. 2009, 83, 8108–8121. [Google Scholar] [CrossRef] [Green Version]
  160. Jarahian, M.; Fiedler, M.; Cohnen, A.; Djandji, D.; Hammerling, G.J.; Gati, C.; Cerwenka, A.; Turner, P.C.; Moyer, R.W.; Watzl, C.; et al. Modulation of NKp30- and NKp46-mediated natural killer cell responses by poxviral hemagglutinin. PLoS Pathog. 2011, 7, e1002195. [Google Scholar] [CrossRef] [PubMed]
  161. Crocker, P.R.; Paulson, J.C.; Varki, A. Siglecs and their roles in the immune system. Nat. Rev. Immunol. 2007, 7, 255–266. [Google Scholar] [CrossRef] [PubMed]
  162. Yau, T.; Dan, X.; Ng, C.C.; Ng, T.B. Lectins with potential for anti-cancer therapy. Molecules 2015, 20, 3791–3810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Santos, A.; Carvalho, F.C.; Roque-Barreira, M.C.; Bueno, P.R. Impedance-derived electrochemical capacitance spectroscopy for the evaluation of lectin-glycoprotein binding affinity. Biosens. Bioelectron. 2014, 62, 102–105. [Google Scholar] [CrossRef]
  164. Barrow, A.D.; Colonna, M. Exploiting NK Cell Surveillance Pathways for Cancer Therapy. Cancers 2019, 11, 55. [Google Scholar] [CrossRef] [Green Version]
  165. Mazalovska, M.; Kouokam, J.C. Lectins as Promising Therapeutics for the Prevention and Treatment of HIV and Other Potential Coinfections. Biomed. Res. Int. 2018, 2018, 3750646. [Google Scholar] [CrossRef]
  166. Brown, G.D.; Willment, J.A.; Whitehead, L. C-type lectins in immunity and homeostasis. Nat. Rev. Immunol. 2018, 18, 374–389. [Google Scholar] [CrossRef] [PubMed]
  167. Barthel, S.R.; Gavino, J.D.; Descheny, L.; Dimitroff, C.J. Targeting selectins and selectin ligands in inflammation and cancer. Expert Opin. Ther. Tar. 2007, 11, 1473–1491. [Google Scholar] [CrossRef]
  168. Zeisig, R.; Stahn, R.; Wenzel, K.; Behrens, D.; Fichtner, I. Effect of sialyl Lewis X-glycoliposomes on the inhibition of E-selectin-mediated tumour cell adhesion in vitro. BBA-Gen. Subj. 2004, 1660, 31–40. [Google Scholar] [CrossRef] [Green Version]
  169. Barthel, S.R.; Gavino, J.D.; Wiese, G.K.; Jaynes, J.M.; Siddiqui, J.; Dimitroff, C.J. Analysis of glycosyltransferase expression in metastatic prostate cancer cells capable of rolling activity on microvascular endothelial (E)-selectin. Glycobiology 2008, 18, 806–817. [Google Scholar] [CrossRef] [PubMed]
  170. Silva-Filho, A.F.; Sena, W.L.B.; Lima, L.R.A.; Carvalho, L.V.N.; Pereira, M.C.; Santos, L.G.S.; Santos, R.V.C.; Tavares, L.B.; Pitta, M.G.R.; Rego, M. Glycobiology Modifications in Intratumoral Hypoxia: The Breathless Side of Glycans Interaction. Cell Physiol. Biochem. 2017, 41, 1801–1829. [Google Scholar] [CrossRef]
  171. Silva, M.; Videira, P.A.; Sackstein, R. E-Selectin Ligands in the Human Mononuclear Phagocyte System: Implications for Infection, Inflammation, and Immunotherapy. Front. Immunol. 2017, 8, 1878. [Google Scholar] [CrossRef] [Green Version]
  172. Bornhofft, K.F.; Goldammer, T.; Rebl, A.; Galuska, S.P. Siglecs: A journey through the evolution of sialic acid-binding immunoglobulin-type lectins. Dev. Comp. Immunol. 2018, 86, 219–231. [Google Scholar] [CrossRef]
  173. Macauley, M.S.; Crocker, P.R.; Paulson, J.C. Siglec-mediated regulation of immune cell function in disease. Nat. Rev. Immunol. 2014, 14, 653–666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Tsubata, T. Ligand Recognition Determines the Role of Inhibitory B Cell Co-receptors in the Regulation of B Cell Homeostasis and Autoimmunity. Front. Immunol. 2018, 9, 2276. [Google Scholar] [CrossRef]
  175. Baba, Y.; Matsumoto, M.; Kurosaki, T. Calcium signaling in B cells: Regulation of cytosolic Ca2+ increase and its sensor molecules, STIM1 and STIM2. Mol. Immunol. 2014, 62, 339–343. [Google Scholar] [CrossRef] [Green Version]
  176. Gerlach, J.; Ghosh, S.; Jumaa, H.; Reth, M.; Wienands, J.; Chan, A.C.; Nitschke, L. B cell defects in SLP65/BLNK-deficient mice can be partially corrected by the absence of CD22, an inhibitory coreceptor for BCR signaling. Eur. J. Immunol. 2003, 33, 3418–3426. [Google Scholar] [CrossRef]
  177. Crocker, P.R.; Redelinghuys, P. Siglecs as positive and negative regulators of the immune system. Biochem. Soc. Trans. 2008, 36, 1467–1471. [Google Scholar] [CrossRef] [PubMed]
  178. Ram, S.; Gulati, S.; Lewis, L.A.; Chakraborti, S.; Zheng, B.; DeOliveira, R.B.; Reed, G.W.; Cox, A.D.; Li, J.; St Michael, F.; et al. A Novel Sialylation Site on Neisseria gonorrhoeae Lipooligosaccharide Links Heptose II Lactose Expression with Pathogenicity. Infect. Immun. 2018, 86, e00285-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Dong, J.; Li, J.; Liu, S.M.; Feng, X.Y.; Chen, S.; Chen, Y.B.; Zhang, X.S. CD33(+)/p-STAT1(+) double-positive cell as a prognostic factor for stage IIIa gastric cancer. Med. Oncol. 2013, 30, 442. [Google Scholar] [CrossRef] [Green Version]
  180. Ishikawa, J.; Kano, F.; Ando, Y.; Hibi, H.; Yamamoto, A. Monocyte chemoattractant protein-1 and secreted ectodomain of sialic acid-binding Ig-like lectin-9 enhance bone regeneration by inducing M2 macrophages. J. Oral Maxillofac. Surg. Med. Pathol. 2019, 31, 169–174. [Google Scholar] [CrossRef]
  181. McMillan, S.J.; Crocker, P.R. CD33-related sialic-acid-binding immunoglobulin-like lectins in health and disease. Carbohyd. Res. 2008, 343, 2050–2056. [Google Scholar] [CrossRef]
  182. Pillai, S.; Netravali, I.A.; Cariappa, A.; Mattoo, H. Siglecs and immune regulation. Annu. Rev. Immunol. 2012, 30, 357–392. [Google Scholar] [CrossRef] [Green Version]
  183. Fraschilla, I.; Pillai, S. Viewing Siglecs through the lens of tumor immunology. Immunol. Rev. 2017, 276, 178–191. [Google Scholar] [CrossRef] [Green Version]
  184. Kakio, A.; Yano, Y.; Takai, D.; Kuroda, Y.; Matsumoto, O.; Kozutsumi, Y.; Matsuzaki, K. Interaction between amyloid beta-protein aggregates and membranes. J. Pept. Sci. 2004, 10, 612–621. [Google Scholar] [CrossRef]
  185. Cowan, C.B.; Patel, D.A.; Good, T.A. Exploring the mechanism of beta-amyloid toxicity attenuation by multivalent sialic acid polymers through the use of mathematical models. J. Theor. Biol. 2009, 258, 189–197. [Google Scholar] [CrossRef] [Green Version]
  186. Hollingworth, P.; Harold, D.; Sims, R.; Gerrish, A.; Lambert, J.C.; Carrasquillo, M.M.; Abraham, R.; Hamshere, M.L.; Pahwa, J.S.; Moskvina, V.; et al. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat. Genet. 2011, 43, 429–435. [Google Scholar] [CrossRef] [Green Version]
  187. Malik, M.; Simpson, J.F.; Parikh, I.; Wilfred, B.R.; Fardo, D.W.; Nelson, P.T.; Estus, S. CD33 Alzheimer’s risk-altering polymorphism, CD33 expression, and exon 2 splicing. J. Neurosci. 2013, 33, 13320–13325. [Google Scholar] [CrossRef] [Green Version]
  188. Malik, M.; Chiles, J., 3rd; Xi, H.S.; Medway, C.; Simpson, J.; Potluri, S.; Howard, D.; Liang, Y.; Paumi, C.M.; Mukherjee, S.; et al. Genetics of CD33 in Alzheimer’s disease and acute myeloid leukemia. Hum. Mol. Genet. 2015, 24, 3557–3570. [Google Scholar] [CrossRef] [Green Version]
  189. Griciuc, A.; Serrano-Pozo, A.; Parrado, A.R.; Lesinski, A.N.; Asselin, C.N.; Mullin, K.; Hooli, B.; Choi, S.H.; Hyman, B.T.; Tanzi, R.E. Alzheimer’s disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron 2013, 78, 631–643. [Google Scholar] [CrossRef] [Green Version]
  190. Heneka, M.T.; Golenbock, D.T.; Latz, E. Innate immunity in Alzheimer’s disease. Nat. Immunol. 2015, 16, 229–236. [Google Scholar] [CrossRef] [PubMed]
  191. Jandus, C.; Boligan, K.F.; Chijioke, O.; Liu, H.; Dahlhaus, M.; Demoulins, T.; Schneider, C.; Wehrli, M.; Hunger, R.E.; Baerlocher, G.M.; et al. Interactions between Siglec-7/9 receptors and ligands influence NK cell-dependent tumor immunosurveillance. J. Clin. Investig. 2014, 124, 1810–1820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Nicoll, G.; Avril, T.; Lock, K.; Furukawa, K.; Bovin, N.; Crocker, P.R. Ganglioside GD3 expression on target cells can modulate NK cell cytotoxicity via siglec-7-dependent and -independent mechanisms. Eur. J. Immunol. 2003, 33, 1642–1648. [Google Scholar] [CrossRef] [PubMed]
  193. Kitajima, K. Structural Diversity and Evolution of Sialic Acids. Trends Glycosci. Glycotechnol. 2019, 31, SE18–SE20. [Google Scholar] [CrossRef]
  194. Ali, S.R.; Fong, J.J.; Carlin, A.F.; Busch, T.D.; Linden, R.; Angata, T.; Areschoug, T.; Parast, M.; Varki, N.; Murray, J.; et al. Siglec-5 and Siglec-14 are polymorphic paired receptors that modulate neutrophil and amnion signaling responses to group B Streptococcus. J. Exp. Med. 2014, 211, 1231–1242. [Google Scholar] [CrossRef]
  195. Avril, T.; Attrill, H.; Zhang, J.; Raper, A.; Crocker, P.R. Negative regulation of leucocyte functions by CD33-related siglecs. Biochem. Soc. Trans. 2006, 34, 1024–1027. [Google Scholar] [CrossRef] [Green Version]
  196. Lubbers, J.; Rodriguez, E.; van Kooyk, Y. Modulation of Immune Tolerance via Siglec-Sialic Acid Interactions. Front. Immunol. 2018, 9, 2807. [Google Scholar] [CrossRef] [Green Version]
  197. Stanczak, M.A.; Siddiqui, S.S.; Trefny, M.P.; Thommen, D.S.; Boligan, K.F.; von Gunten, S.; Tzankov, A.; Tietze, L.; Lardinois, D.; Heinzelmann-Schwarz, V.; et al. Self-associated molecular patterns mediate cancer immune evasion by engaging Siglecs on T cells. J. Clin. Investig. 2018, 128, 4912–4923. [Google Scholar] [CrossRef] [Green Version]
  198. Duan, S.; Paulson, J.C. Siglecs as immune cell checkpoints in disease. Annu. Rev. Immunol. 2020, 38, 365–395. [Google Scholar] [CrossRef] [Green Version]
  199. Kolar, P.; Knieke, K.; Hegel, J.K.; Quandt, D.; Burmester, G.R.; Hoff, H.; Brunner-Weinzierl, M.C. CTLA-4 (CD152) controls homeostasis and suppressive capacity of regulatory T cells in mice. Arthritis Rheum. 2009, 60, 123–132. [Google Scholar] [CrossRef]
  200. Syn, N.L.; Teng, M.W.L.; Mok, T.S.K.; Soo, R.A. De-novo and acquired resistance to immune checkpoint targeting. Lancet. Oncol. 2017, 18, e731–e741. [Google Scholar] [CrossRef]
  201. Philips, G.K.; Atkins, M. Therapeutic uses of anti-PD-1 and anti-PD-L1 antibodies. Int. Immunol. 2015, 27, 39–46. [Google Scholar] [CrossRef] [Green Version]
  202. Liu, J.F.; Wu, L.; Yang, L.L.; Deng, W.W.; Mao, L.; Wu, H.; Zhang, W.F.; Sun, Z.J. Blockade of TIM3 relieves immunosuppression through reducing regulatory T cells in head and neck cancer. J. Exp. Clin. Canc. Res. 2018, 37, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Chien, M.W.; Lin, M.H.; Huang, S.H.; Fu, S.H.; Hsu, C.Y.; Yen, B.L.; Chen, J.T.; Chang, D.M.; Sytwu, H.K. Glucosamine Modulates T Cell Differentiation through Down-regulating N-Linked Glycosylation of CD25. J. Biol. Chem. 2015, 290, 29329–29344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Chen, J.T.; Chen, C.H.; Ku, K.L.; Hsiao, M.; Chiang, C.P.; Hsu, T.L.; Chen, M.H.; Wong, C.H. Glycoprotein B7-H3 overexpression and aberrant glycosylation in oral cancer and immune response. Proc. Natl. Acad. Sci. USA 2015, 112, 13057–13062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Pereira, M.S.; Alves, I.; Vicente, M.; Campar, A.; Silva, M.C.; Padrao, N.A.; Pinto, V.; Fernandes, A.; Dias, A.M.; Pinho, S.S. Glycans as Key Checkpoints of T Cell Activity and Function. Front. Immunol. 2018, 9, 2754. [Google Scholar] [CrossRef] [PubMed]
  206. Mittler, R.S.; Foell, J.; McCausland, M.; Strahotin, S.; Niu, L.; Bapat, A.; Hewes, L.B. Anti-CD137 antibodies in the treatment of autoimmune disease and cancer. Immunol. Res. 2004, 29, 197–208. [Google Scholar] [CrossRef]
  207. Eastwood, D.; Findlay, L.; Poole, S.; Bird, C.; Wadhwa, M.; Moore, M.; Burns, C.; Thorpe, R.; Stebbings, R. Monoclonal antibody TGN1412 trial failure explained by species differences in CD28 expression on CD4+ effector memory T-cells. Br. J. Pharmacol. 2010, 161, 512–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Sun, R.; Kim, A.M.J.; Lim, S.-O. Glycosylation of Immune Receptors in Cancer. Cells 2021, 10, 1100. [Google Scholar] [CrossRef]
  209. Liu, Z.; Swindall, A.F.; Kesterson, R.A.; Schoeb, T.R.; Bullard, D.C.; Bellis, S.L. ST6Gal-I regulates macrophage apoptosis via alpha2-6 sialylation of the TNFR1 death receptor. J. Biol. Chem. 2011, 286, 39654–39662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Pardoll, D.M. The blockade of immune checkpoints in cancer immunotherapy. Nat. Rev. Cancer 2012, 12, 252–264. [Google Scholar] [CrossRef] [Green Version]
  211. Aicher, A.; Hayden-Ledbetter, M.; Brady, W.A.; Pezzutto, A.; Richter, G.; Magaletti, D.; Buckwalter, S.; Ledbetter, J.A.; Clark, E.A. Characterization of human inducible costimulator ligand expression and function. J. Immunol. 2000, 164, 4689–4696. [Google Scholar] [CrossRef] [Green Version]
  212. Rujas, E.; Cui, H.; Sicard, T.; Semesi, A.; Julien, J.-P. Structural characterization of the ICOS/ICOS-L immune complex reveals high molecular mimicry by therapeutic antibodies. Nat. Commun. 2020, 11, 5066. [Google Scholar] [CrossRef]
  213. Sun, L.; Li, C.-W.; Chung, E.M.; Yang, R.; Kim, Y.-S.; Park, A.H.; Lai, Y.-J.; Yang, Y.; Wang, Y.-H.; Liu, J. Targeting glycosylated PD-1 induces potent antitumor immunity. Cancer Res. 2020, 80, 2298–2310. [Google Scholar]
  214. Lin, D.Y.; Tanaka, Y.; Iwasaki, M.; Gittis, A.G.; Su, H.P.; Mikami, B.; Okazaki, T.; Honjo, T.; Minato, N.; Garboczi, D.N. The PD-1/PD-L1 complex resembles the antigen-binding Fv domains of antibodies and T cell receptors. Proc. Natl. Acad. Sci. USA 2008, 105, 3011–3016. [Google Scholar] [CrossRef] [Green Version]
  215. Zak, K.M.; Grudnik, P.; Magiera, K.; Domling, A.; Dubin, G.; Holak, T.A. Structural Biology of the Immune Checkpoint Receptor PD-1 and Its Ligands PD-L1/PD-L2. Structure 2017, 25, 1163–1174. [Google Scholar] [CrossRef]
  216. Dong, Y.; Sun, Q.; Zhang, X. PD-1 and its ligands are important immune checkpoints in cancer. Oncotarget 2017, 8, 2171–2186. [Google Scholar] [CrossRef] [Green Version]
  217. Redelinghuys, P.; Antonopoulos, A.; Liu, Y.; Campanero-Rhodes, M.A.; McKenzie, E.; Haslam, S.M.; Dell, A.; Feizi, T.; Crocker, P.R. Early murine T-lymphocyte activation is accompanied by a switch from N-Glycolyl- to N-acetyl-neuraminic acid and generation of ligands for siglec-E. J. Biol. Chem. 2011, 286, 34522–34532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Kavvoura, F.K.; Ioannidis, J.P. CTLA-4 gene polymorphisms and susceptibility to type 1 diabetes mellitus: A HuGE Review and meta-analysis. Am. J. Epidemiol. 2005, 162, 3–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Martinez Allo, V.C.; Hauk, V.; Sarbia, N.; Pinto, N.A.; Croci, D.O.; Dalotto-Moreno, T.; Morales, R.M.; Gatto, S.G.; Manselle Cocco, M.N.; Stupirski, J.C.; et al. Suppression of age-related salivary gland autoimmunity by glycosylation-dependent galectin-1-driven immune inhibitory circuits. Proc. Natl. Acad. Sci. USA 2020, 117, 6630–6639. [Google Scholar] [CrossRef] [PubMed]
  220. Elola, M.T.; Wolfenstein-Todel, C.; Troncoso, M.F.; Vasta, G.R.; Rabinovich, G.A. Galectins: Matricellular glycan-binding proteins linking cell adhesion, migration, and survival. Cell Mol. Life Sci. 2007, 64, 1679–1700. [Google Scholar] [CrossRef] [PubMed]
  221. Chou, F.C.; Chen, H.Y.; Kuo, C.C.; Sytwu, H.K. Role of Galectins in Tumors and in Clinical Immunotherapy. Int. J. Mol. Sci. 2018, 19, 430. [Google Scholar] [CrossRef] [Green Version]
  222. Vasta, G.R.; Feng, C.; Gonzalez-Montalban, N.; Mancini, J.; Yang, L.; Abernathy, K.; Frost, G.; Palm, C. Functions of galectins as ‘self/non-self’-recognition and effector factors. Pathog. Dis. 2017, 75, ftx046. [Google Scholar] [CrossRef] [Green Version]
  223. Wan, L.; Yang, R.Y.; Liu, F.T. Galectin-12 in Cellular Differentiation, Apoptosis and Polarization. Int. J. Mol. Sci. 2018, 19, 176. [Google Scholar] [CrossRef] [Green Version]
  224. Wang, J.; Huang, Y.; Zhang, J.; Xing, B.; Xuan, W.; Wang, H.; Huang, H.; Yang, J.; Tang, J. NRP-2 in tumor lymphangiogenesis and lymphatic metastasis. Cancer Lett. 2018, 418, 176–184. [Google Scholar] [CrossRef]
  225. Galvan, M.; Tsuboi, S.; Fukuda, M.; Baum, L.G. Expression of a specific glycosyltransferase enzyme regulates T cell death mediated by galectin-1. J. Biol. Chem. 2000, 275, 16730–16737. [Google Scholar] [CrossRef] [Green Version]
  226. Garin, M.I.; Chu, C.C.; Golshayan, D.; Cernuda-Morollon, E.; Wait, R.; Lechler, R.I. Galectin-1: A key effector of regulation mediated by CD4+CD25+ T cells. Blood 2007, 109, 2058–2065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Sturm, A.; Lensch, M.; Andre, S.; Kaltner, H.; Wiedenmann, B.; Rosewicz, S.; Dignass, A.U.; Gabius, H.J. Human galectin-2: Novel inducer of T cell apoptosis with distinct profile of caspase activation. J. Immunol. 2004, 173, 3825–3837. [Google Scholar] [CrossRef] [Green Version]
  228. Seki, M.; Oomizu, S.; Sakata, K.M.; Sakata, A.; Arikawa, T.; Watanabe, K.; Ito, K.; Takeshita, K.; Niki, T.; Saita, N.; et al. Galectin-9 suppresses the generation of Th17, promotes the induction of regulatory T cells, and regulates experimental autoimmune arthritis. Clin. Immunol. 2008, 127, 78–88. [Google Scholar] [CrossRef]
  229. Clark, M.C.; Baum, L.G. T cells modulate glycans on CD43 and CD45 during development and activation, signal regulation, and survival. Ann. N. Y. Acad. Sci. 2012, 1253, 58–67. [Google Scholar] [CrossRef] [Green Version]
  230. Kadaja-Saarepuu, L.; Laos, S.; Jaager, K.; Viil, J.; Balikova, A.; Looke, M.; Hansson, G.C.; Maimets, T. CD43 promotes cell growth and helps to evade FAS-mediated apoptosis in non-hematopoietic cancer cells lacking the tumor suppressors p53 or ARF. Oncogene 2008, 27, 1705–1715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Earl, L.A.; Bi, S.; Baum, L.G. N- and O-glycans modulate galectin-1 binding, CD45 signaling, and T cell death. J. Biol. Chem. 2010, 285, 2232–2244. [Google Scholar] [CrossRef] [Green Version]
  232. Woodard-Grice, A.V.; McBrayer, A.C.; Wakefield, J.K.; Zhuo, Y.; Bellis, S.L. Proteolytic shedding of ST6Gal-I by BACE1 regulates the glycosylation and function of alpha4beta1 integrins. J. Biol. Chem. 2008, 283, 26364–26373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Pretzlaff, R.K.; Xue, V.W.; Rowin, M.E. Sialidase treatment exposes the beta1-integrin active ligand binding site on HL60 cells and increases binding to fibronectin. Cell Adhes. Commun. 2000, 7, 491–500. [Google Scholar] [CrossRef] [PubMed]
  234. Chang, V.T.; Fernandes, R.A.; Ganzinger, K.A.; Lee, S.F.; Siebold, C.; McColl, J.; Jönsson, P.; Palayret, M.; Harlos, K.; Coles, C.H. Initiation of T cell signaling by CD45 segregation at‘close contacts’. Nat. Immunol. 2016, 17, 574–582. [Google Scholar] [CrossRef] [Green Version]
  235. Stroop, C.J.; Weber, W.; Gerwig, G.J.; Nimtz, M.; Kamerling, J.P.; Vliegenthart, J.F. Characterization of the carbohydrate chains of the secreted form of the human epidermal growth factor receptor. Glycobiology 2000, 10, 901–917. [Google Scholar] [CrossRef] [Green Version]
  236. Kitazume, S.; Imamaki, R.; Kurimoto, A.; Ogawa, K.; Kato, M.; Yamaguchi, Y.; Tanaka, K.; Ishida, H.; Ando, H.; Kiso, M.; et al. 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. 2014, 289, 27604–27613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Dall’Olio, F.; Malagolini, N.; Trinchera, M.; Chiricolo, M. Sialosignaling: Sialyltransferases as engines of self-fueling loops in cancer progression. BBA-Gen. Subj. 2014, 1840, 2752–2764. [Google Scholar] [CrossRef] [Green Version]
  238. Cheung, I.Y.; Vickers, A.; Cheung, N.K. Sialyltransferase STX (ST8SiaII): A novel molecular marker of metastatic neuroblastoma. Int. J. Cancer 2006, 119, 152–156. [Google Scholar] [CrossRef]
  239. Cazet, A.; Julien, S.; Bobowski, M.; Burchell, J.; Delannoy, P. Tumour-associated carbohydrate antigens in breast cancer. Breast Cancer Res. 2010, 12, 204. [Google Scholar] [CrossRef] [Green Version]
  240. Schultz, M.J.; Swindall, A.F.; Bellis, S.L. Regulation of the metastatic cell phenotype by sialylated glycans. Cancer Metast Rev. 2012, 31, 501–518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Garcia, G.G.; Berger, S.B.; Sadighi Akha, A.A.; Miller, R.A. Age-associated changes in glycosylation of CD43 and CD45 on mouse CD4 T cells. Eur. J. Immunol. 2005, 35, 622–631. [Google Scholar] [CrossRef] [PubMed]
  242. Stillman, B.N.; Hsu, D.K.; Pang, M.; Brewer, C.F.; Johnson, P.; Liu, F.T.; Baum, L.G. Galectin-3 and galectin-1 bind distinct cell surface glycoprotein receptors to induce T cell death. J. Immunol. 2006, 176, 778–789. [Google Scholar] [CrossRef] [Green Version]
  243. Ma, L.; Dong, L.; Chang, P. CD44v6 engages in colorectal cancer progression. Cell Death Dis. 2019, 10, 30. [Google Scholar] [CrossRef]
  244. Zhao, Y.; Sato, Y.; Isaji, T.; Fukuda, T.; Matsumoto, A.; Miyoshi, E.; Gu, J.; Taniguchi, N. Branched N-glycans regulate the biological functions of integrins and cadherins. FEBS J. 2008, 275, 1939–1948. [Google Scholar] [CrossRef]
  245. Fukumori, T.; Takenaka, Y.; Yoshii, T.; Kim, H.R.; Hogan, V.; Inohara, H.; Kagawa, S.; Raz, A. CD29 and CD7 mediate galectin-3-induced type II T-cell apoptosis. Cancer Res. 2003, 63, 8302–8311. [Google Scholar] [PubMed]
  246. Baycin-Hizal, D.; Gottschalk, A.; Jacobson, E.; Mai, S.; Wolozny, D.; Zhang, H.; Krag, S.S.; Betenbaugh, M.J. Physiologic and pathophysiologic consequences of altered sialylation and glycosylation on ion channel function. Biochem. Biophys. Res. Commun. 2014, 453, 243–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Sato, C.; Kitajima, K. Polysialylation and disease. Mol. Asp. Med. 2020, 79, 100892. [Google Scholar] [CrossRef] [PubMed]
  248. Bull, C.; Collado-Camps, E.; Kers-Rebel, E.D.; Heise, T.; Sondergaard, J.N.; den Brok, M.H.; Schulte, B.M.; Boltje, T.J.; Adema, G.J. Metabolic sialic acid blockade lowers the activation threshold of moDCs for TLR stimulation. Immunol. Cell Biol. 2017, 95, 408–415. [Google Scholar] [CrossRef] [PubMed]
  249. Silva, M.; Silva, Z.; Marques, G.; Ferro, T.; Goncalves, M.; Monteiro, M.; van Vliet, S.J.; Mohr, E.; Lino, A.C.; Fernandes, A.R.; et al. Sialic acid removal from dendritic cells improves antigen cross-presentation and boosts anti-tumor immune responses. Oncotarget 2016, 7, 41053–41066. [Google Scholar] [CrossRef]
  250. Bull, C.; Heise, T.; Adema, G.J.; Boltje, T.J. Sialic Acid Mimetics to Target the Sialic Acid-Siglec Axis. Trends Biochem. Sci. 2016, 41, 519–531. [Google Scholar] [CrossRef] [PubMed]
  251. Cerliani, J.P.; Blidner, A.G.; Toscano, M.A.; Croci, D.O.; Rabinovich, G.A. Translating the ‘Sugar Code’ into Immune and Vascular Signaling Programs. Trends Biochem. Sci. 2017, 42, 255–273. [Google Scholar] [CrossRef]
  252. Adams, R.; Brown, E.; Brown, L.; Butler, R.; Falk, S.; Fisher, D.; Kaplan, R.; Quirke, P.; Richman, S.; Samuel, L.; et al. Inhibition of EGFR, HER2, and HER3 signalling in patients with colorectal cancer wild-type for BRAF, PIK3CA, KRAS, and NRAS (FOCUS4-D): A phase 2–3 randomised trial. Lancet Gastroenterol. Hepatol. 2018, 3, 162–171. [Google Scholar] [CrossRef] [Green Version]
  253. Wittmann, P.; Grubinger, M.; Groger, C.; Huber, H.; Sieghart, W.; Peck-Radosavljevic, M.; Mikulits, W. Neuropilin-2 induced by transforming growth factor-beta augments migration of hepatocellular carcinoma cells. BMC Cancer 2015, 15, 909. [Google Scholar] [CrossRef] [Green Version]
  254. Patel, D.A.; Henry, J.E.; Good, T.A. Attenuation of beta-amyloid-induced toxicity by sialic-acid-conjugated dendrimers: Role of sialic acid attachment. Brain Res. 2007, 1161, 95–105. [Google Scholar] [CrossRef] [Green Version]
  255. Malicdan, M.C.; Noguchi, S.; Nonaka, I.; Hayashi, Y.K.; Nishino, I. A Gne knockout mouse expressing human GNE D176V mutation develops features similar to distal myopathy with rimmed vacuoles or hereditary inclusion body myopathy. Hum. Mol. Genet. 2007, 16, 2669–2682. [Google Scholar] [CrossRef] [Green Version]
  256. Murray, H.C.; Low, V.F.; Swanson, M.E.; Dieriks, B.V.; Turner, C.; Faull, R.L.; Curtis, M.A. Distribution of PSA-NCAM in normal, Alzheimer’s and Parkinson’s disease human brain. Neuroscience 2016, 330, 359–375. [Google Scholar] [CrossRef] [PubMed]
  257. Cvetko, A.; Kifer, D.; Gornik, O.; Klarić, L.; Visser, E.; Lauc, G.; Wilson, J.F.; Štambuk, T. Glycosylation alterations in multiple sclerosis show increased proinflammatory potential. Biomedicines 2020, 8, 410. [Google Scholar] [CrossRef]
  258. Nishino, I.; Wu, S.W.; Ibarra, C.A.; Malicdan, M.C.V.; Murayama, K.; Noguchi, S.; Hayashi, Y.K.; Nonaka, I. Central core disease is due to RYR1 mutations in more than 90% of patients. Brain 2006, 129, 1470–1480. [Google Scholar]
  259. Oizumi, H.; Hayashita-Kinoh, H.; Hayakawa, H.; Arai, H.; Furuya, T.; Ren, Y.R.; Yasuda, T.; Seki, T.; Mizuno, Y.; Mochizuki, H. Alteration in the differentiation-related molecular expression in the subventricular zone in a mouse model of Parkinson’s disease. Neurosci. Res. 2008, 60, 15–21. [Google Scholar] [CrossRef]
  260. Albrecht, A.; Stork, O. Are NCAM deficient mice an animal model for schizophrenia? Front. Behav. Neurosci. 2012, 6, 43. [Google Scholar] [CrossRef] [Green Version]
  261. McHugh, J. Autoimmunity: Glycoengineering has therapeutic potential. Nat. Rev. Rheumatol. 2018, 14, 121. [Google Scholar] [CrossRef] [PubMed]
  262. Opferman, J.T. Apoptosis in the development of the immune system. Cell Death Differ. 2008, 15, 234–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Kovacs-Solyom, F.; Blasko, A.; Fajka-Boja, R.; Katona, R.L.; Vegh, L.; Novak, J.; Szebeni, G.J.; Krenacs, L.; Uher, F.; Tubak, V.; et al. Mechanism of tumor cell-induced T-cell apoptosis mediated by galectin-1. Immunol. Lett. 2010, 127, 108–118. [Google Scholar] [CrossRef]
  264. Toscano, M.A.; Bianco, G.A.; Ilarregui, J.M.; Croci, D.O.; Correale, J.; Hernandez, J.D.; Zwirner, N.W.; Poirier, F.; Riley, E.M.; Baum, L.G.; et al. Differential glycosylation of TH1, TH2 and TH-17 effector cells selectively regulates susceptibility to cell death. Nat. Immunol. 2007, 8, 825–834. [Google Scholar] [CrossRef]
  265. Bi, S.; Baum, L.G. Sialic acids in T cell development and function. BBA-Gen. Subj. 2009, 1790, 1599–1610. [Google Scholar] [CrossRef]
  266. Liao, W.; Lin, J.X.; Wang, L.; Li, P.; Leonard, W.J. Modulation of cytokine receptors by IL-2 broadly regulates differentiation into helper T cell lineages. Nat. Immunol. 2011, 12, 551–559. [Google Scholar] [CrossRef] [Green Version]
  267. Chen, Q.; Kim, Y.C.; Laurence, A.; Punkosdy, G.A.; Shevach, E.M. IL-2 controls the stability of Foxp3 expression in TGF-beta-induced Foxp3+ T cells in vivo. J. Immunol. 2011, 186, 6329–6337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Earl, L.A.; Baum, L.G. CD45 glycosylation controls T-cell life and death. Immunol. Cell Biol. 2008, 86, 608–615. [Google Scholar] [CrossRef] [PubMed]
  269. Marth, J.D.; Grewal, P.K. Mammalian glycosylation in immunity. Nat. Rev. Immunol. 2008, 8, 874–887. [Google Scholar] [CrossRef] [Green Version]
  270. Gascoigne, N.R. T-cell differentiation: MHC class I’s sweet tooth lost on maturity. Curr. Biol. 2002, 12, R99–R101. [Google Scholar] [CrossRef] [Green Version]
  271. Dings, R.P.M.; Miller, M.C.; Griffin, R.J.; Mayo, K.H. Galectins as Molecular Targets for Therapeutic Intervention. Int. J. Mol. Sci. 2018, 19, 905. [Google Scholar] [CrossRef] [Green Version]
  272. Grigorian, A.; Torossian, S.; Demetriou, M. T-cell growth, cell surface organization, and the galectin-glycoprotein lattice. Immunol. Rev. 2009, 230, 232–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Kizuka, Y.; Oka, S. Regulated expression and neural functions of human natural killer-1 (HNK-1) carbohydrate. Cell Mol. Life Sci. 2012, 69, 4135–4147. [Google Scholar] [CrossRef] [Green Version]
  274. Morise, J.; Kizuka, Y.; Yabuno, K.; Tonoyama, Y.; Hashii, N.; Kawasaki, N.; Manya, H.; Miyagoe-Suzuki, Y.; Takeda, S.i.; Endo, T. Structural and biochemical characterization of O-mannose-linked human natural killer-1 glycan expressed on phosphacan in developing mouse brains. Glycobiology 2014, 24, 314–324. [Google Scholar] [CrossRef] [Green Version]
  275. Yu, R.K.; Yanagisawa, M. Glycobiology of neural stem cells: Functional aspects. J. Neurochem. 2006, 5, 415–423. [Google Scholar]
  276. Ribeiro, M.; Levay, K.; Yon, B.; Ayupe, A.C.; Salgueiro, Y.; Park, K.K. Neural cadherin plays distinct roles for neuronal survival and axon growth under different regenerative conditions. Eneuro 2020, 7. [Google Scholar] [CrossRef]
  277. Hamlin, J.A.; Fang, H.; Schwob, J.E. Differential expression of the mammalian homologue of fasciclin II during olfactory development in vivo and in vitro. J. Comp. Neurol. 2004, 474, 438–452. [Google Scholar] [CrossRef] [PubMed]
  278. Winther, M.; Walmod, P.S. Neural cell adhesion molecules belonging to the family of leucine-rich repeat proteins. Cell Adhes. Mol. 2014, 315–395. [Google Scholar]
  279. Takahashi, S.; Kato, K.; Nakamura, K.; Nakano, R.; Kubota, K.; Hamada, H. Neural cell adhesion molecule 2 as a target molecule for prostate and breast cancer gene therapy. Cancer Sci. 2011, 102, 808–814. [Google Scholar] [CrossRef] [PubMed]
  280. Edwards, S.; Campbell, C.; Flohr, P.; Shipley, J.; Giddings, I.; Te-Poele, R.; Dodson, A.; Foster, C.; Clark, J.; Jhavar, S.; et al. Expression analysis onto microarrays of randomly selected cDNA clones highlights HOXB13 as a marker of human prostate cancer. Brit. J. Cancer 2005, 92, 376–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Nelson, E.A.; Walker, S.R.; Li, W.; Liu, X.S.; Frank, D.A. Identification of human STAT5-dependent gene regulatory elements based on interspecies homology. J. Biol. Chem. 2006, 281, 26216–26224. [Google Scholar] [CrossRef] [Green Version]
  282. Cirielli, V.; Cima, L.; Chindemi, C.; Danzi, O.; Ghimenton, C.; Eccher, A.; Mauriello, S.; Bortolotti, F.; De Leo, D.; Brunelli, M. Cortical Expression of the Polysialylated Isoform of the Neural Cell Adhesion Molecule on Brain Tissue to Recognize Drug-Related Death: An Exploratory Analysis. Am. J. Forensic Med. Pathol. 2018, 39, 8–13. [Google Scholar] [CrossRef]
  283. Wuhrer, M.; Geyer, H.; von der Ohe, M.; Gerardy-Schahn, R.; Schachner, M.; Geyer, R. Localization of defined carbohydrate epitopes in bovine polysialylated NCAM. Biochimie 2003, 85, 207–218. [Google Scholar] [CrossRef]
  284. Ong, E.; Suzuki, M.; Belot, F.; Yeh, J.C.; Franceschini, I.; Angata, K.; Hindsgaul, O.; Fukuda, M. Biosynthesis of HNK-1 glycans on O-linked oligosaccharides attached to the neural cell adhesion molecule (NCAM): The requirement for core 2 beta 1,6-N-acetylglucosaminyltransferase and the muscle-specific domain in NCAM. J. Biol. Chem. 2002, 277, 18182–18190. [Google Scholar] [CrossRef] [Green Version]
  285. Liedtke, S.; Geyer, H.; Wuhrer, M.; Geyer, R.; Frank, G.; Gerardy-Schahn, R.; Zahringer, U.; Schachner, M. Characterization of N-glycans from mouse brain neural cell adhesion molecule. Glycobiology 2001, 11, 373–384. [Google Scholar] [CrossRef]
  286. Eberhardt, K.A.; Irintchev, A.; Al-Majed, A.A.; Simova, O.; Brushart, T.M.; Gordon, T.; Schachner, M. BDNF/TrkB signaling regulates HNK-1 carbohydrate expression in regenerating motor nerves and promotes functional recovery after peripheral nerve repair. Exp. Neurol. 2006, 198, 500–510. [Google Scholar] [CrossRef] [PubMed]
  287. Ferrer-Ferrer, M.; Dityatev, A. Shaping Synapses by the Neural Extracellular Matrix. Front. Neuroanat. 2018, 12, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Bukalo, O.; Schachner, M.; Dityatev, A. Hippocampal metaplasticity induced by deficiency in the extracellular matrix glycoprotein tenascin-R. J. Neurosci. 2007, 27, 6019–6028. [Google Scholar] [CrossRef]
  289. Saito, H.; Nakao, Y.; Takayama, S.; Toyama, Y.; Asou, H. Specific expression of an HNK-1 carbohydrate epitope and NCAM on femoral nerve Schwann cells in mice. Neurosci. Res. 2005, 53, 314–322. [Google Scholar] [CrossRef]
  290. Irintchev, A.; Schachner, M. The injured and regenerating nervous system: Immunoglobulin superfamily members as key players. Neuroscientist. 2012, 18, 452–466. [Google Scholar] [CrossRef]
  291. Makhina, T.; Loers, G.; Schulze, C.; Ueberle, B.; Schachner, M.; Kleene, R. Extracellular GAPDH binds to L1 and enhances neurite outgrowth. Mol. Cell. Neurosci. 2009, 41, 206–218. [Google Scholar] [CrossRef]
  292. Hillenbrand, R.; Molthagen, M.; Montag, D.; Schachner, M. The close homologue of the neural adhesion molecule L1 (CHL1): Patterns of expression and promotion of neurite outgrowth by heterophilic interactions. Eur. J. Neurosci. 1999, 11, 813–826. [Google Scholar] [CrossRef]
  293. Franceschini, I.; Vitry, S.; Padilla, F.; Casanova, P.; Tham, T.N.; Fukuda, M.; Rougon, G.; Durbec, P.; Dubois-Dalcq, M. Migrating and myelinating potential of neural precursors engineered to overexpress PSA-NCAM. Mol. Cell. Neurosci. 2004, 27, 151–162. [Google Scholar] [CrossRef] [PubMed]
  294. Nakamura, A.; Morise, J.; Yabuno-Nakagawa, K.; Hashimoto, Y.; Takematsu, H.; Oka, S. Site-specific HNK-1 epitope on alternatively spliced fibronectin type-III repeats in tenascin-C promotes neurite outgrowth of hippocampal neurons through contactin-1. PLoS ONE 2019, 14, e0210193. [Google Scholar] [CrossRef]
  295. Valdembri, D.; Caswell, P.T.; Anderson, K.I.; Schwarz, J.P.; Konig, I.; Astanina, E.; Caccavari, F.; Norman, J.C.; Humphries, M.J.; Bussolino, F.; et al. Neuropilin-1/GIPC1 signaling regulates alpha5beta1 integrin traffic and function in endothelial cells. PLoS Biol. 2009, 7, e25. [Google Scholar] [CrossRef]
  296. Robinson, S.D.; Reynolds, L.E.; Kostourou, V.; Reynolds, A.R.; da Silva, R.G.; Tavora, B.; Baker, M.; Marshall, J.F.; Hodivala-Dilke, K.M. Alphav beta3 integrin limits the contribution of neuropilin-1 to vascular endothelial growth factor-induced angiogenesis. J. Biol. Chem. 2009, 284, 33966–33981. [Google Scholar] [CrossRef] [Green Version]
  297. Chen, C.; Zhao, S.; Karnad, A.; Freeman, J.W. The biology and role of CD44 in cancer progression: Therapeutic implications. J. Hematol. Oncol. 2018, 11, 64. [Google Scholar] [CrossRef] [Green Version]
  298. Uniewicz, K.A.; Ori, A.; Ahmed, Y.A.; Yates, E.A.; Fernig, D.G. Characterisation of the interaction of neuropilin-1 with heparin and a heparan sulfate mimetic library of heparin-derived sugars. Peerj 2014, 2, e461. [Google Scholar] [CrossRef] [Green Version]
  299. Soares, M.A.; Teixeira, F.C.; Fontes, M.; Areas, A.L.; Leal, M.G.; Pavao, M.S.; Stelling, M.P. Heparan Sulfate Proteoglycans May Promote or Inhibit Cancer Progression by Interacting with Integrins and Affecting Cell Migration. Biomed. Res. Int. 2015, 2015, 453801. [Google Scholar] [CrossRef] [Green Version]
  300. Cai, Y.; Wang, R.; Zhao, Y.F.; Jia, J.; Sun, Z.J.; Chen, X.M. Expression of Neuropilin-2 in salivary adenoid cystic carcinoma: Its implication in tumor progression and angiogenesis. Pathol. Res. Pract. 2010, 206, 793–799. [Google Scholar] [CrossRef]
  301. Tu, D.G.; Chang, W.W.; Jan, M.S.; Tu, C.W.; Lu, Y.C.; Tai, C.K. Promotion of metastasis of thyroid cancer cells via NRP-2-mediated induction. Oncol. Lett. 2016, 12, 4224–4230. [Google Scholar] [CrossRef] [Green Version]
  302. Dewan, M.Z.; Takamatsu, N.; Hidaka, T.; Hatakeyama, K.; Nakahata, S.; Fujisawa, J.; Katano, H.; Yamamoto, N.; Morishita, K. Critical role for TSLC1 expression in the growth and organ infiltration of adult T-cell leukemia cells in vivo. J. Virol. 2008, 82, 11958–11963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Galuska, S.P.; Rollenhagen, M.; Kaup, M.; Eggers, K.; Oltmann-Norden, I.; Schiff, M.; Hartmann, M.; Weinhold, B.; Hildebrandt, H.; Geyer, R.; et al. Synaptic cell adhesion molecule SynCAM 1 is a target for polysialylation in postnatal mouse brain. Proc. Natl. Acad. Sci. USA 2010, 107, 10250–10255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Caunt, M.; Mak, J.; Liang, W.C.; Stawicki, S.; Pan, Q.; Tong, R.K.; Kowalski, J.; Ho, C.; Reslan, H.B.; Ross, J.; et al. Blocking neuropilin-2 function inhibits tumor cell metastasis. Cancer Cell 2008, 13, 331–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Prud’homme, G.J. Cancer stem cells and novel targets for antitumor strategies. Curr. Pharm. Des. 2012, 18, 2838–2849. [Google Scholar] [CrossRef]
  306. Schellenburg, S.; Schulz, A.; Poitz, D.M.; Muders, M.H. Role of neuropilin-2 in the immune system. Mol. Immunol. 2017, 90, 239–244. [Google Scholar] [CrossRef]
  307. Rossignol, M.; Gagnon, M.L.; Klagsbrun, M. Genomic organization of human neuropilin-1 and neuropilin-2 genes: Identification and distribution of splice variants and soluble isoforms. Genomics 2000, 70, 211–222. [Google Scholar] [CrossRef] [PubMed]
  308. Handa, A.; Tokunaga, T.; Tsuchida, T.; Lee, Y.H.; Kijima, H.; Yamazaki, H.; Ueyama, Y.; Fukuda, H.; Nakamura, M. Neuropilin-2 expression affects the increased vascularization and is a prognostic factor in osteosarcoma. Int. J. Oncol. 2000, 17, 291–295. [Google Scholar] [CrossRef] [PubMed]
  309. Li, J.; Yang, R.; Yang, H.; Chen, S.; Wang, L.; Li, M.; Yang, S.; Feng, Z.; Bi, J. NCAM regulates the proliferation, apoptosis, autophagy, EMT, and migration of human melanoma cells via the Src/Akt/mTOR/cofilin signaling pathway. J. Cell Biochem. 2020, 121, 1192–1204. [Google Scholar] [CrossRef]
  310. Yasuoka, H.; Kodama, R.; Tsujimoto, M.; Yoshidome, K.; Akamatsu, H.; Nakahara, M.; Inagaki, M.; Sanke, T.; Nakamura, Y. Neuropilin-2 expression in breast cancer: Correlation with lymph node metastasis, poor prognosis, and regulation of CXCR4 expression. Bmc. Cancer 2009, 9, 220. [Google Scholar] [CrossRef] [Green Version]
  311. Nasarre, P.; Gemmill, R.M.; Potiron, V.A.; Roche, J.; Lu, X.; Baron, A.E.; Korch, C.; Garrett-Mayer, E.; Lagana, A.; Howe, P.H.; et al. Neuropilin-2 Is upregulated in lung cancer cells during TGF-beta1-induced epithelial-mesenchymal transition. Cancer Res. 2013, 73, 7111–7121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. Rieger, J.; Wick, W.; Weller, M. Human malignant glioma cells express semaphorins and their receptors, neuropilins and plexins. Glia 2003, 42, 379–389. [Google Scholar] [CrossRef] [PubMed]
  313. Fakhari, M.; Pullirsch, D.; Abraham, D.; Paya, K.; Hofbauer, R.; Holzfeind, P.; Hofmann, M.; Aharinejad, S. Selective upregulation of vascular endothelial growth factor receptors neuropilin-1 and -2 in human neuroblastoma. Cancer-Am. Cancer Soc. 2002, 94, 258–263. [Google Scholar] [CrossRef]
  314. Hansel, D.E.; Wilentz, R.E.; Yeo, C.J.; Schulick, R.D.; Montgomery, E.; Maitra, A. Expression of neuropilin-1 in high-grade dysplasia, invasive cancer, and metastases of the human gastrointestinal tract. Am. J. Surg. Pathol. 2004, 28, 347–356. [Google Scholar] [CrossRef] [PubMed]
  315. Dallas, N.A.; Gray, M.J.; Xia, L.; Fan, F.; van Buren, G., 2nd; Gaur, P.; Samuel, S.; Lim, S.J.; Arumugam, T.; Ramachandran, V.; et al. Neuropilin-2-mediated tumor growth and angiogenesis in pancreatic adenocarcinoma. Clin. Cancer Res. 2008, 14, 8052–8060. [Google Scholar] [CrossRef] [Green Version]
  316. Cohen, T.; Herzog, Y.; Brodzky, A.; Greenson, J.K.; Eldar, S.; Gluzman-Poltorak, Z.; Neufeld, G.; Resnick, M.B. Neuropilin-2 is a novel marker expressed in pancreatic islet cells and endocrine pancreatic tumours. J. Pathol. 2002, 198, 77–82. [Google Scholar] [CrossRef]
  317. Parikh, A.A.; Liu, W.B.; Fan, F.; Stoeltzing, O.; Reinmuth, N.; Bruns, C.J.; Bucana, C.D.; Evans, D.B.; Ellis, L.M. Expression and regulation of the novel vascular endothelial growth factor receptor neuropilin-1 by epidermal growth factor in human pancreatic carcinoma. Cancer-Am. Cancer Soc. 2003, 98, 720–729. [Google Scholar] [CrossRef] [PubMed]
  318. Calicchio, M.L.; Collins, T.; Kozakewich, H.P. Identification of signaling systems in proliferating and involuting phase infantile hemangiomas by genome-wide transcriptional profiling. Am. J. Pathol. 2009, 174, 1638–1649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Schulz, A.; Gorodetska, I.; Behrendt, R.; Fuessel, S.; Erdmann, K.; Foerster, S.; Datta, K.; Mayr, T.; Dubrovska, A.; Muders, M.H. Linking NRP2 with EMT and chemoradioresistance in bladder cancer. Front. Oncol. 2020, 9, 1461. [Google Scholar] [CrossRef] [PubMed]
  320. Polavaram, N.S.; Dutta, S.; Islam, R.; Bag, A.K.; Roy, S.; Poitz, D.; Karnes, J.; Hofbauer, L.C.; Kohli, M.; Costello, B.A. Tumor-and osteoclast-derived NRP2 in prostate cancer bone metastases. Bone Res. 2021, 9, 1–16. [Google Scholar] [CrossRef]
  321. Butti, R.; Kumar, T.V.; Nimma, R.; Kundu, G.C. Impact of semaphorin expression on prognostic characteristics in breast cancer. Breast Cancer 2018, 10, 79–88. [Google Scholar] [CrossRef] [Green Version]
  322. Hildebrandt, H.; Dityatev, A. Polysialic acid in brain development and synaptic plasticity. SialoGlyco Chem. Biol. I 2013, 366, 55–96. [Google Scholar]
  323. Windisch, R.; Pirschtat, N.; Kellner, C.; Chen-Wichmann, L.; Lausen, J.; Humpe, A.; Krause, D.S.; Wichmann, C. Oncogenic deregulation of cell adhesion molecules in leukemia. Cancers 2019, 11, 311. [Google Scholar] [CrossRef] [Green Version]
  324. Sasaki, H.; Nishikata, I.; Shiraga, T.; Akamatsu, E.; Fukami, T.; Hidaka, T.; Kubuki, Y.; Okayama, A.; Hamada, K.; Okabe, H.; et al. Overexpression of a cell adhesion molecule, TSLC1, as a possible molecular marker for acute-type adult T-cell leukemia. Blood 2005, 105, 1204–1213. [Google Scholar] [CrossRef]
  325. Qin, L.; Zhu, W.; Xu, T.; Hao, Y.; Zhang, Z.; Tian, Y.; Yang, D. Effect of TSLC1 gene on proliferation, invasion and apoptosis of human hepatocellular carcinoma cell line HepG2. J. Huazhong Univ. Sci. Technol. 2007, 27, 535–537. [Google Scholar] [CrossRef] [PubMed]
  326. Usami, Y.; Ito, A.; Ohnuma, K.; Fuku, T.; Komori, T.; Yokozaki, H. Tumor suppressor in lung cancer-1 as a novel ameloblast adhesion molecule and its downregulation in ameloblastoma. Pathol. Int. 2007, 57, 68–75. [Google Scholar] [CrossRef]
  327. Werneburg, S.; Buettner, F.F.; Muhlenhoff, M.; Hildebrandt, H. Polysialic acid modification of the synaptic cell adhesion molecule SynCAM 1 in human embryonic stem cell-derived oligodendrocyte precursor cells. Stem. Cell Res. 2015, 14, 339–346. [Google Scholar] [CrossRef] [Green Version]
  328. Fogel, A.I.; Akins, M.R.; Krupp, A.J.; Stagi, M.; Stein, V.; Biederer, T. SynCAMs organize synapses through heterophilic adhesion. J. Neurosci. 2007, 27, 12516–12530. [Google Scholar] [CrossRef] [Green Version]
  329. Takase, N.; Koma, Y.; Urakawa, N.; Nishio, M.; Arai, N.; Akiyama, H.; Shigeoka, M.; Kakeji, Y.; Yokozaki, H. NCAM- and FGF-2-mediated FGFR1 signaling in the tumor microenvironment of esophageal cancer regulates the survival and migration of tumor-associated macrophages and cancer cells. Cancer Lett. 2016, 380, 47–58. [Google Scholar] [CrossRef] [Green Version]
  330. Britain, C.M.; Holdbrooks, A.T.; Anderson, J.C.; Willey, C.D.; Bellis, S.L. Sialylation of EGFR by the ST6Gal-I sialyltransferase promotes EGFR activation and resistance to gefitinib-mediated cell death. J. Ovarian Res. 2018, 11, 12. [Google Scholar] [CrossRef] [Green Version]
  331. Karpanen, T.; Heckman, C.A.; Keskitalo, S.; Jeltsch, M.; Ollila, H.; Neufeld, G.; Tamagnone, L.; Alitalo, K. Functional interaction of VEGF-C and VEGF-D with neuropilin receptors. FASEB J. 2006, 20, 1462–1472. [Google Scholar] [CrossRef] [Green Version]
  332. Seidenfaden, R.; Krauter, A.; Schertzinger, F.; Gerardy-Schahn, R.; Hildebrandt, H. Polysialic acid directs tumor cell growth by controlling heterophilic neural cell adhesion molecule interactions. Mol. Cell Biol. 2003, 23, 5908–5918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Vempati, P.; Popel, A.S.; Mac Gabhann, F. Extracellular regulation of VEGF: Isoforms, proteolysis, and vascular patterning. Cytokine Growth Factor Rev. 2014, 25, 1–19. [Google Scholar] [CrossRef] [Green Version]
  334. Grunewald, F.S.; Prota, A.E.; Giese, A.; Ballmer-Hofer, K. Structure-function analysis of VEGF receptor activation and the role of coreceptors in angiogenic signaling. BBA-Gen. Subj. 2010, 1804, 567–580. [Google Scholar] [CrossRef] [PubMed]
  335. Shraga-Heled, N.; Kessler, O.; Prahst, C.; Kroll, J.; Augustin, H.; Neufeld, G. Neuropilin-1 and neuropilin-2 enhance VEGF121 stimulated signal transduction by the VEGFR-2 receptor. FASEB J. 2007, 21, 915–926. [Google Scholar] [CrossRef] [PubMed]
  336. Ellis, L.M.; Rosen, L.; Gordon, M.S. Overview of anti-VEGF therapy and angiogenesis. Part 1: Angiogenesis inhibition in solid tumor malignancies. Clin. Adv. Hematol. Oncol. 2006, 4 (Suppl. 1–10). [Google Scholar]
  337. Elola, M.T.; Blidner, A.G.; Ferragut, F.; Bracalente, C.; Rabinovich, G.A. Assembly, organization and regulation of cell-surface receptors by lectin-glycan complexes. Biochem. J. 2015, 469, 1–16. [Google Scholar] [CrossRef]
  338. Domigan, C.K.; Ziyad, S.; Iruela-Arispe, M.L. Canonical and noncanonical vascular endothelial growth factor pathways: New developments in biology and signal transduction. Arter. Thromb. Vasc. Biol. 2015, 35, 30–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  339. Basagiannis, D.; Zografou, S.; Murphy, C.; Fotsis, T.; Morbidelli, L.; Ziche, M.; Bleck, C.; Mercer, J.; Christoforidis, S. VEGF induces signalling and angiogenesis by directing VEGFR2 internalisation through macropinocytosis. J. Cell Sci. 2016, 129, 4091–4104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  340. Croci, D.O.; Cerliani, J.P.; Pinto, N.A.; Morosi, L.G.; Rabinovich, G.A. Regulatory role of glycans in the control of hypoxia-driven angiogenesis and sensitivity to anti-angiogenic treatment. Glycobiology 2014, 24, 1283–1290. [Google Scholar] [CrossRef] [Green Version]
  341. Povlsen, G.K.; Berezin, V.; Bock, E. Neural cell adhesion molecule-180-mediated homophilic binding induces epidermal growth factor receptor (EGFR) down-regulation and uncouples the inhibitory function of EGFR in neurite outgrowth. J. Neurochem. 2008, 104, 624–639. [Google Scholar] [CrossRef] [PubMed]
  342. Islamov, R.R.; Izmailov, A.A.; Sokolov, M.E.; Fadeev, P.O.; Bashirov, F.V.; Eremeev, A.A.; Shaymardanova, G.F.; Shmarov, M.M.; Naroditskiy, B.S.; Chelyshev, Y.A.; et al. Evaluation of direct and cell-mediated triple-gene therapy in spinal cord injury in rats. Brain Res. Bull. 2017, 132, 44–52. [Google Scholar] [CrossRef] [PubMed]
  343. Islamov, R.R.; Rizvanov, A.A.; Mukhamedyarov, M.A.; Salafutdinov, I.I.; Garanina, E.E.; Fedotova, V.Y.; Solovyeva, V.V.; Mukhamedshina, Y.O.; Safiullov, Z.Z.; Izmailov, A.A.; et al. Symptomatic improvement, increased life-span and sustained cell homing in amyotrophic lateral sclerosis after transplantation of human umbilical cord blood cells genetically modified with adeno-viral vectors expressing a neuro-protective factor and a neural cell adhesion molecule. Curr. Gene Ther. 2015, 15, 266–276. [Google Scholar] [CrossRef]
  344. Izmailov, A.A.; Povysheva, T.V.; Bashirov, F.V.; Sokolov, M.E.; Fadeev, F.O.; Garifulin, R.R.; Naroditsky, B.S.; Logunov, D.Y.; Salafutdinov, I.I.; Chelyshev, Y.A.; et al. Spinal Cord Molecular and Cellular Changes Induced by Adenoviral Vector- and Cell-Mediated Triple Gene Therapy after Severe Contusion. Front. Pharm. 2017, 8, 813. [Google Scholar] [CrossRef] [PubMed]
  345. Jarahian, M.; Watzl, C.; Issa, Y.; Altevogt, P.; Momburg, F. Blockade of natural killer cell-mediated lysis by NCAM140 expressed on tumor cells. Int. J. Cancer 2007, 120, 2625–2634. [Google Scholar] [CrossRef]
  346. Bonfanti, L.; Theodosis, D.T. Polysialic acid and activity-dependent synapse remodeling. Cell Adh. Migr. 2009, 3, 43–50. [Google Scholar] [CrossRef] [Green Version]
  347. Plein, A.; Fantin, A.; Ruhrberg, C. Neuropilin regulation of angiogenesis, arteriogenesis, and vascular permeability. Microcirculation 2014, 21, 315–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  348. Wee, P.; Wang, Z. Epidermal Growth Factor Receptor Cell Proliferation Signaling Pathways. Cancers 2017, 9, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  349. Zhou, P.; Hu, J.; Wang, X.; Wang, J.; Zhang, Y.; Wang, C. Epidermal growth factor receptor expression affects proliferation and apoptosis in non-small cell lung cancer cells via the extracellular signal-regulated kinase/microRNA 200a signaling pathway. Oncol. Lett. 2018, 15, 5201–5207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  350. Berger, R.P.; Sun, Y.H.; Kulik, M.; Lee, J.K.; Nairn, A.V.; Moremen, K.W.; Pierce, M.; Dalton, S. ST8SIA4-Dependent Polysialylation is Part of a Developmental Program Required for Germ Layer Formation from Human Pluripotent Stem Cells. Stem Cells 2016, 34, 1742–1752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  351. Li, X.; Young, N.M.; Tropp, S.; Hu, D.; Xu, Y.; Hallgrimsson, B.; Marcucio, R.S. Quantification of shape and cell polarity reveals a novel mechanism underlying malformations resulting from related FGF mutations during facial morphogenesis. Hum. Mol. Genet. 2013, 22, 5160–5172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Colombo, F.; Meldolesi, J. L1-CAM and N-CAM: From Adhesion Proteins to Pharmacological Targets. Trends Pharm. Sci. 2015, 36, 769–781. [Google Scholar] [CrossRef] [PubMed]
  353. Li, J.; Dai, G.; Cheng, Y.B.; Qi, X.; Geng, M.Y. Polysialylation promotes neural cell adhesion molecule-mediated cell migration in a fibroblast growth factor receptor-dependent manner, but independent of adhesion capability. Glycobiology 2011, 21, 1010–1018. [Google Scholar] [CrossRef] [Green Version]
  354. Steenbergen, S.M.; Vimr, E.R. Chromatographic analysis of the Escherichia coli polysialic acid capsule. Methods Mol. Biol. 2013, 966, 109–120. [Google Scholar] [CrossRef] [Green Version]
  355. Bull, C.; den Brok, M.H.; Adema, G.J. Sweet escape: Sialic acids in tumor immune evasion. BBA-Gen. Subj. 2014, 1846, 238–246. [Google Scholar] [CrossRef]
  356. Roy, S.; Bag, A.K.; Singh, R.K.; Talmadge, J.E.; Batra, S.K.; Datta, K. Multifaceted Role of Neuropilins in the Immune System: Potential Targets for Immunotherapy. Front. Immunol. 2017, 8, 1228. [Google Scholar] [CrossRef] [Green Version]
  357. Lemmon, M.A.; Schlessinger, J. Cell signaling by receptor tyrosine kinases. Cell 2010, 141, 1117–1134. [Google Scholar] [CrossRef] [Green Version]
  358. Sjostrand, D.; Ibanez, C.F. Insights into GFRalpha1 regulation of neural cell adhesion molecule (NCAM) function from structure-function analysis of the NCAM/GFRalpha1 receptor complex. J. Biol. Chem. 2008, 283, 13792–13798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  359. Sandhya, V.K.; Raju, R.; Verma, R.; Advani, J.; Sharma, R.; Radhakrishnan, A.; Nanjappa, V.; Narayana, J.; Somani, B.L.; Mukherjee, K.K.; et al. A network map of BDNF/TRKB and BDNF/p75NTR signaling system. J. Cell Commun. Signal. 2013, 7, 301–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  360. Zhang, L.; Zhao, H.; Zhang, X.; Chen, L.; Zhao, X.; Bai, X.; Zhang, J. Nobiletin protects against cerebral ischemia via activating the p-Akt, p-CREB, BDNF and Bcl-2 pathway and ameliorating BBB permeability in rat. Brain Res. Bull. 2013, 96, 45–53. [Google Scholar] [CrossRef] [PubMed]
  361. Liu, Z.; Ren, Z.; Zhang, J.; Chuang, C.C.; Kandaswamy, E.; Zhou, T.; Zuo, L. Role of ROS and Nutritional Antioxidants in Human Diseases. Front. Physiol. 2018, 9, 477. [Google Scholar] [CrossRef] [Green Version]
  362. Muramatsu, T. Midkine and pleiotrophin: Two related proteins involved in development, survival, inflammation and tumorigenesis. J. Biochem. 2002, 132, 359–371. [Google Scholar] [CrossRef]
  363. Palmer, R.H.; Vernersson, E.; Grabbe, C.; Hallberg, B. Anaplastic lymphoma kinase: Signalling in development and disease. Biochem. J. 2009, 420, 345–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Berge, E.M.; Doebele, R.C. Targeted therapies in non-small cell lung cancer: Emerging oncogene targets following the success of epidermal growth factor receptor. Semin. Oncol. 2014, 41, 110–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  365. Ou, S.H.; Tan, J.; Yen, Y.; Soo, R.A. ROS1 as a ‘druggable’ receptor tyrosine kinase: Lessons learned from inhibiting the ALK pathway. Expert Rev. Anticancer. Ther. 2012, 12, 447–456. [Google Scholar] [CrossRef]
  366. Davies, K.D.; Le, A.T.; Theodoro, M.F.; Skokan, M.C.; Aisner, D.L.; Berge, E.M.; Terracciano, L.M.; Cappuzzo, F.; Incarbone, M.; Roncalli, M.; et al. Identifying and targeting ROS1 gene fusions in non-small cell lung cancer. Clin. Cancer Res. 2012, 18, 4570–4579. [Google Scholar] [CrossRef] [Green Version]
  367. Acquaviva, J.; Wong, R.; Charest, A. The multifaceted roles of the receptor tyrosine kinase ROS in development and cancer. BBA-Gen. Subj. 2009, 1795, 37–52. [Google Scholar] [CrossRef]
  368. Charest, A.; Wilker, E.W.; McLaughlin, M.E.; Lane, K.; Gowda, R.; Coven, S.; McMahon, K.; Kovach, S.; Feng, Y.; Yaffe, M.B.; et al. ROS fusion tyrosine kinase activates a SH2 domain-containing phosphatase-2/phosphatidylinositol 3-kinase/mammalian target of rapamycin signaling axis to form glioblastoma in mice. Cancer Res. 2006, 66, 7473–7481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  369. Chiarle, R.; Voena, C.; Ambrogio, C.; Piva, R.; Inghirami, G. The anaplastic lymphoma kinase in the pathogenesis of cancer. Nat. Rev. Cancer 2008, 8, 11–23. [Google Scholar] [CrossRef] [PubMed]
  370. Chiarle, R.; Simmons, W.J.; Cai, H.; Dhall, G.; Zamo, A.; Raz, R.; Karras, J.G.; Levy, D.E.; Inghirami, G. Stat3 is required for ALK-mediated lymphomagenesis and provides a possible therapeutic target. Nat. Med. 2005, 11, 623–629. [Google Scholar] [CrossRef]
  371. Soda, M.; Takada, S.; Takeuchi, K.; Ishikawa, Y.; Sugiyama, Y.; Mano, H. Analysis of a Mouse Model for EML4-ALK-Positive Lung Cancer. Am. J. Respir. Crit. Care Med. 2009, 179, A2693. [Google Scholar]
  372. Nguyen, K.T.; Zong, C.S.; Uttamsingh, S.; Sachdev, P.; Bhanot, M.; Le, M.T.; Chan, J.L.; Wang, L.H. The role of phosphatidylinositol 3-kinase, rho family GTPases, and STAT3 in Ros-induced cell transformation. J. Biol. Chem. 2002, 277, 11107–11115. [Google Scholar] [CrossRef] [Green Version]
  373. Charest, A.; Lane, K.; McMahon, K.; Park, J.; Preisinger, E.; Conroy, H.; Housman, D. Fusion of FIG to the receptor tyrosine kinase ROS in a glioblastoma with an interstitial del(6)(q21q21). Genes Chromosomes Cancer 2003, 37, 58–71. [Google Scholar] [CrossRef] [PubMed]
  374. Stransky, N.; Cerami, E.; Schalm, S.; Kim, J.L.; Lengauer, C. The landscape of kinase fusions in cancer. Nat. Commun. 2014, 5, 4846. [Google Scholar] [CrossRef] [Green Version]
  375. Lovly, C.M.; Gupta, A.; Lipson, D.; Otto, G.; Brennan, T.; Chung, C.T.; Borinstein, S.C.; Ross, J.S.; Stephens, P.J.; Miller, V.A.; et al. Inflammatory myofibroblastic tumors harbor multiple potentially actionable kinase fusions. Cancer Discov. 2014, 4, 889–895. [Google Scholar] [CrossRef] [Green Version]
  376. Yamamoto, H.; Yoshida, A.; Taguchi, K.; Kohashi, K.; Hatanaka, Y.; Yamashita, A.; Mori, D.; Oda, Y. ALK, ROS1 and NTRK3 gene rearrangements in inflammatory myofibroblastic tumours. Histopathology 2016, 69, 72–83. [Google Scholar] [CrossRef]
  377. Gu, T.L.; Deng, X.; Huang, F.; Tucker, M.; Crosby, K.; Rimkunas, V.; Wang, Y.; Deng, G.; Zhu, L.; Tan, Z.; et al. Survey of tyrosine kinase signaling reveals ROS kinase fusions in human cholangiocarcinoma. PLoS ONE 2011, 6, e15640. [Google Scholar] [CrossRef] [Green Version]
  378. Birch, A.H.; Arcand, S.L.; Oros, K.K.; Rahimi, K.; Watters, A.K.; Provencher, D.; Greenwood, C.M.; Mes-Masson, A.M.; Tonin, P.N. Chromosome 3 anomalies investigated by genome wide SNP analysis of benign, low malignant potential and low grade ovarian serous tumours. PLoS ONE 2011, 6, e28250. [Google Scholar] [CrossRef]
  379. Lee, J.; Lee, S.E.; Kang, S.Y.; Do, I.G.; Lee, S.; Ha, S.Y.; Cho, J.; Kang, W.K.; Jang, J.; Ou, S.H.; et al. Identification of ROS1 rearrangement in gastric adenocarcinoma. Cancer-Am. Cancer Soc. 2013, 119, 1627–1635. [Google Scholar] [CrossRef]
  380. Aisner, D.L.; Nguyen, T.T.; Paskulin, D.D.; Le, A.T.; Haney, J.; Schulte, N.; Chionh, F.; Hardingham, J.; Mariadason, J.; Tebbutt, N.; et al. ROS1 and ALK fusions in colorectal cancer, with evidence of intratumoral heterogeneity for molecular drivers. Mol. Cancer Res. 2014, 12, 111–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  381. Giacomini, C.P.; Sun, S.; Varma, S.; Shain, A.H.; Giacomini, M.M.; Balagtas, J.; Sweeney, R.T.; Lai, E.; Del Vecchio, C.A.; Forster, A.D.; et al. Breakpoint analysis of transcriptional and genomic profiles uncovers novel gene fusions spanning multiple human cancer types. PLoS Genet. 2013, 9, e1003464. [Google Scholar] [CrossRef] [PubMed]
  382. Wiesner, T.; He, J.; Yelensky, R.; Esteve-Puig, R.; Botton, T.; Yeh, I.; Lipson, D.; Otto, G.; Brennan, K.; Murali, R.; et al. Kinase fusions are frequent in Spitz tumours and spitzoid melanomas. Nat. Commun. 2014, 5, 3116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  383. Shaw, A.; Riley, G.J.; Bang, Y.J.; Kim, D.W.; Camidge, D.R.; Varella-Garcia, M.; Lafrate, A.J.; Shapiro, G.; Winter, M.; Usari, T.; et al. Crizotinib in advanced ROS1-rearranged non-small cell lung cancer (NSCLC): Updated results from PROFILE 1001. Ann. Oncol. 2016, 27, vi418. [Google Scholar] [CrossRef]
  384. Collisson, E.A.; Campbell, J.D.; Brooks, A.N.; Berger, A.H.; Lee, W.; Chmielecki, J.; Beer, D.G.; Cope, L.; Creighton, C.J.; Danilova, L.; et al. Comprehensive molecular profiling of lung adenocarcinoma. Nature 2014, 511, 543–550. [Google Scholar] [CrossRef]
  385. Zhu, V.W.; Upadhyay, D.; Schrock, A.B.; Gowen, K.; Ali, S.M.; Ou, S.H. TPD52L1-ROS1, a new ROS1 fusion variant in lung adenosquamous cell carcinoma identified by comprehensive genomic profiling. Lung Cancer 2016, 97, 48–50. [Google Scholar] [CrossRef] [PubMed]
  386. Farago, A.F.; Zheng, Z.L.; Muzikansky, A.; Gainor, J.F.; Iafrate, A.J.; Engelman, J.A.; Le, L.P.; Shaw, A.T. Clinical implementation of anchored multiplex PCR with targeted next-generation sequencing for detection of ALK, ROS1, RET and NTRK1 fusions in non-small cell lung carcinoma. J. Clin. Oncol. 2015, 33, 8095. [Google Scholar] [CrossRef]
  387. Rikova, K.; Guo, A.; Zeng, Q.; Possemato, A.; Yu, J.; Haack, H.; Nardone, J.; Lee, K.; Reeves, C.; Li, Y.; et al. Global survey of phosphotyrosine signaling identifies oncogenic kinases in lung cancer. Cell 2007, 131, 1190–1203. [Google Scholar] [CrossRef] [Green Version]
  388. Takeuchi, K.; Soda, M.; Togashi, Y.; Suzuki, R.; Sakata, S.; Hatano, S.; Asaka, R.; Hamanaka, W.; Ninomiya, H.; Uehara, H.; et al. RET, ROS1 and ALK fusions in lung cancer. Nat. Med. 2012, 18, 378–381. [Google Scholar] [CrossRef] [PubMed]
  389. Rimkunas, V.M.; Crosby, K.E.; Li, D.; Hu, Y.; Kelly, M.E.; Gu, T.L.; Mack, J.S.; Silver, M.R.; Zhou, X.; Haack, H. Analysis of receptor tyrosine kinase ROS1-positive tumors in non-small cell lung cancer: Identification of a FIG-ROS1 fusion. Clin. Cancer Res. 2012, 18, 4449–4457. [Google Scholar] [CrossRef] [Green Version]
  390. Al-Saraireh, Y.M.; Sutherland, M.; Springett, B.R.; Freiberger, F.; Ribeiro Morais, G.; Loadman, P.M.; Errington, R.J.; Smith, P.J.; Fukuda, M.; Gerardy-Schahn, R.; et al. Pharmacological inhibition of polysialyltransferase ST8SiaII modulates tumour cell migration. PLoS ONE 2013, 8, e73366. [Google Scholar] [CrossRef] [Green Version]
  391. Yoshida, A.; Kohno, T.; Tsuta, K.; Wakai, S.; Arai, Y.; Shimada, Y.; Asamura, H.; Furuta, K.; Shibata, T.; Tsuda, H. ROS1-rearranged lung cancer: A clinicopathologic and molecular study of 15 surgical cases. Am. J. Surg. Pathol. 2013, 37, 554–562. [Google Scholar] [CrossRef]
  392. Suehara, Y.; Arcila, M.; Wang, L.; Hasanovic, A.; Ang, D.; Ito, T.; Kimura, Y.; Drilon, A.; Guha, U.; Rusch, V.; et al. Identification of KIF5B-RET and GOPC-ROS1 fusions in lung adenocarcinomas through a comprehensive mRNA-based screen for tyrosine kinase fusions. Clin. Cancer Res. 2012, 18, 6599–6608. [Google Scholar] [CrossRef] [Green Version]
  393. Govindan, R.; Ding, L.; Griffith, M.; Subramanian, J.; Dees, N.D.; Kanchi, K.L.; Maher, C.A.; Fulton, R.; Fulton, L.; Wallis, J.; et al. Genomic landscape of non-small cell lung cancer in smokers and never-smokers. Cell 2012, 150, 1121–1134. [Google Scholar] [CrossRef] [Green Version]
  394. Seo, J.S.; Ju, Y.S.; Lee, W.C.; Shin, J.Y.; Lee, J.K.; Bleazard, T.; Lee, J.; Jung, Y.J.; Kim, J.O.; Shin, J.Y.; et al. The transcriptional landscape and mutational profile of lung adenocarcinoma. Genome Res. 2012, 22, 2109–2119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  395. Gainor, J.F.; Shaw, A.T. Novel targets in non-small cell lung cancer: ROS1 and RET fusions. Oncologist 2013, 18, 865–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  396. Lin, J.J.; Shaw, A.T. Recent Advances in Targeting ROS1 in Lung Cancer. J. Thorac. Oncol. 2017, 12, 1611–1625. [Google Scholar] [CrossRef] [Green Version]
  397. Hayashi, Y.; Jinnou, H.; Sawamoto, K.; Hitoshi, S. Adult neurogenesis and its role in brain injury and psychiatric diseases. J. Neurochem. 2018, 147, 584–594. [Google Scholar] [CrossRef] [Green Version]
  398. Quartu, M.; Serra, M.P.; Boi, M.; Ibba, V.; Melis, T.; Del Fiacco, M. Polysialylated-neural cell adhesion molecule (PSA-NCAM) in the human trigeminal ganglion and brainstem at prenatal and adult ages. BMC Neurosci. 2008, 9, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  399. Coviello, S.; Benedetti, B.; Jakubecova, D.; Belles, M.; Klimczak, P.; Gramuntell, Y.; Couillard-Despres, S.; Nacher, J. PSA Depletion Induces the Differentiation of Immature Neurons in the Piriform Cortex of Adult Mice. Int. J. Mol. Sci. 2021, 22, 5733. [Google Scholar] [CrossRef] [PubMed]
  400. Varea, E.; Castillo-Gomez, E.; Gomez-Climent, M.A.; Guirado, R.; Blasco-Ibanez, J.M.; Crespo, C.; Martinez-Guijarro, F.J.; Nacher, J. Differential evolution of PSA-NCAM expression during aging of the rat telencephalon. Neurobiol. Aging 2009, 30, 808–818. [Google Scholar] [CrossRef] [PubMed]
  401. Aonurm-Helm, A.; Jaako, K.; Jürgenson, M.; Zharkovsky, A. Pharmacological approach for targeting dysfunctional brain plasticity: Focus on neural cell adhesion molecule (NCAM). Pharmacol. Res. 2016, 113, 731–738. [Google Scholar] [CrossRef] [PubMed]
  402. Gattenlohner, S.; Stuhmer, T.; Leich, E.; Reinhard, M.; Etschmann, B.; Volker, H.U.; Rosenwald, A.; Serfling, E.; Bargou, R.C.; Ertl, G.; et al. Specific detection of CD56 (NCAM) isoforms for the identification of aggressive malignant neoplasms with progressive development. Am. J. Pathol. 2009, 174, 1160–1171. [Google Scholar] [CrossRef] [Green Version]
  403. Galuska, C.E.; Lutteke, T.; Galuska, S.P. Is Polysialylated NCAM Not Only a Regulator during Brain Development But also during the Formation of Other Organs? Biology 2017, 6, 27. [Google Scholar] [CrossRef] [Green Version]
  404. Ono, T.; Takeshita, A.; Kishimoto, Y.; Kiyoi, H.; Okada, M.; Yamauchi, T.; Emi, N.; Horikawa, K.; Matsuda, M.; Shinagawa, K.; et al. Expression of CD56 is an unfavorable prognostic factor for acute promyelocytic leukemia with higher initial white blood cell counts. Cancer Sci. 2014, 105, 97–104. [Google Scholar] [CrossRef]
  405. Murray, N.; Salgia, R.; Fossella, F.V. Targeted molecules in small cell lung cancer. Semin. Oncol. 2004, 31, 106–111. [Google Scholar] [CrossRef]
  406. Rushing, E.C.; Stine, M.J.; Hahn, S.J.; Shea, S.; Eller, M.S.; Naif, A.; Khanna, S.; Westra, W.H.; Jungbluth, A.A.; Busam, K.J.; et al. Neuropilin-2: A novel biomarker for malignant melanoma? Hum. Pathol. 2012, 43, 381–389. [Google Scholar] [CrossRef] [Green Version]
  407. Vossen, L.I.; Markovsky, E.; Eldar-Boock, A.; Tschiche, H.R.; Wedepohl, S.; Pisarevsky, E.; Satchi-Fainaro, R.; Calderon, M. PEGylated dendritic polyglycerol conjugate targeting NCAM-expressing neuroblastoma: Limitations and challenges. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 1169–1179. [Google Scholar] [CrossRef]
  408. Wachowiak, R.; Krause, M.; Mayer, S.; Peukert, N.; Suttkus, A.; Muller, W.C.; Lacher, M.; Meixensberger, J.; Nestler, U. Increased L1CAM (CD171) levels are associated with glioblastoma and metastatic brain tumors. Medicine 2018, 97, e12396. [Google Scholar] [CrossRef] [PubMed]
  409. Ardizzone, A.; Scuderi, S.A.; Giuffrida, D.; Colarossi, C.; Puglisi, C.; Campolo, M.; Cuzzocrea, S.; Esposito, E.; Paterniti, I. Role of Fibroblast Growth Factors Receptors (FGFRs) in Brain Tumors, Focus on Astrocytoma and Glioblastoma. Cancers 2020, 12, 3825. [Google Scholar] [CrossRef] [PubMed]
  410. Gattenloehner, S.; Chuvpilo, S.; Langebrake, C.; Reinhardt, D.; Muller-Hermelink, H.K.; Serfling, E.; Vincent, A.; Marx, A. Novel RUNX1 isoforms determine the fate of acute myeloid leukemia cells by controlling CD56 expression. Blood 2007, 110, 2027–2033. [Google Scholar] [CrossRef] [Green Version]
  411. Montesinos, P.; Rayon, C.; Vellenga, E.; Brunet, S.; Gonzalez, J.; Gonzalez, M.; Holowiecka, A.; Esteve, J.; Bergua, J.; Gonzalez, J.D.; et al. Clinical significance of CD56 expression in patients with acute promyelocytic leukemia treated with all-trans retinoic acid and anthracycline-based regimens. Blood 2011, 117, 1799–1805. [Google Scholar] [CrossRef] [Green Version]
  412. Paietta, E.; Neuberg, D.; Richards, S.; Bennett, J.M.; Han, L.; Racevskis, J.; Dewald, G.; Rowe, J.M.; Wiernik, P.H. Rare adult acute lymphocytic leukemia with CD56 expression in the ECOG experience shows unexpected phenotypic and genotypic heterogeneity. Am. J. Hematol. 2001, 66, 189–196. [Google Scholar] [CrossRef]
  413. Fischer, L.; Gokbuget, N.; Schwartz, S.; Burmeister, T.; Rieder, H.; Bruggemann, M.; Hoelzer, D.; Thiel, E. CD56 expression in T-cell acute lymphoblastic leukemia is associated with non-thymic phenotype and resistance to induction therapy but no inferior survival after risk-adapted therapy. Haematologica 2009, 94, 224–229. [Google Scholar] [CrossRef] [PubMed]
  414. Wielgat, P.; Rogowski, K.; Niemirowicz-Laskowska, K.; Car, H. Sialic acid-siglec axis as molecular checkpoints targeting of immune system: Smart players in pathology and conventional therapy. Int. J. Mol. Sci. 2020, 21, 4361. [Google Scholar] [CrossRef]
  415. Rose, M.G.; Berliner, N. T-cell large granular lymphocyte leukemia and related disorders. Oncologist 2004, 9, 247–258. [Google Scholar] [CrossRef]
  416. Kawasaki, T.; Suzuki, M.; Sato, A.; Yashima-Abo, A.; Satoh, T.; Kato, R.; Kato, Y.; Obara, W.; Shimoyama, T.; Ishida, Y. Neural cell adhesion molecule (CD56)-positive B cell lymphoma of the urinary bladder. J. Clin. Pathol. 2016, 69, 89–92. [Google Scholar] [CrossRef]
  417. Guan, F.; Wang, X.; He, F. Promotion of cell migration by neural cell adhesion molecule (NCAM) is enhanced by PSA in a polysialyltransferase-specific manner. PLoS ONE 2015, 10, e0124237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  418. Lehembre, F.; Yilmaz, M.; Wicki, A.; Schomber, T.; Strittmatter, K.; Ziegler, D.; Kren, A.; Went, P.; Derksen, P.W.; Berns, A.; et al. NCAM-induced focal adhesion assembly: A functional switch upon loss of E-cadherin. EMBO J. 2008, 27, 2603–2615. [Google Scholar] [CrossRef] [Green Version]
  419. Schreiber, S.C.; Giehl, K.; Kastilan, C.; Hasel, C.; Muhlenhoff, M.; Adler, G.; Wedlich, D.; Menke, A. Polysialylated NCAM represses E-cadherin-mediated cell-cell adhesion in pancreatic tumor cells. Gastroenterology 2008, 134, 1555–1566. [Google Scholar] [CrossRef] [PubMed]
  420. Sun, Z.; Parrish, A.R.; Hill, M.A.; Meininger, G.A. N-cadherin, A Vascular Smooth Muscle Cell–Cell Adhesion Molecule: Function and Signaling for Vasomotor Control. Microcirculation 2014, 21, 208–218. [Google Scholar] [CrossRef]
  421. Cagnoni, A.J.; Perez Saez, J.M.; Rabinovich, G.A.; Marino, K.V. Turning-Off Signaling by Siglecs, Selectins, and Galectins: Chemical Inhibition of Glycan-Dependent Interactions in Cancer. Front. Oncol. 2016, 6, 109. [Google Scholar] [CrossRef] [Green Version]
  422. Ge, H.; Mu, L.; Jin, L.; Yang, C.; Chang, Y.E.; Long, Y.; DeLeon, G.; Deleyrolle, L.; Mitchell, D.A.; Kubilis, P.S.; et al. Tumor associated CD70 expression is involved in promoting tumor migration and macrophage infiltration in GBM. Int. J. Cancer 2017, 141, 1434–1444. [Google Scholar] [CrossRef] [Green Version]
  423. Micheau, O. Regulation of TNF-Related Apoptosis-Inducing Ligand Signaling by Glycosylation. Int. J. Mol. Sci. 2018, 19, 715. [Google Scholar] [CrossRef] [Green Version]
  424. Beatty, G.L.; Gladney, W.L. Immune escape mechanisms as a guide for cancer immunotherapy. Clin. Cancer Res. 2015, 21, 687–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  425. Takeuchi, H.; Haltiwanger, R.S. Significance of glycosylation in Notch signaling. Biochem. Biophys. Res. Commun. 2014, 453, 235–242. [Google Scholar] [CrossRef] [Green Version]
  426. Wu, A.A.; Drake, V.; Huang, H.S.; Chiu, S.; Zheng, L. Reprogramming the tumor microenvironment: Tumor-induced immunosuppressive factors paralyze T cells. Oncoimmunology 2015, 4, e1016700. [Google Scholar] [CrossRef] [PubMed]
  427. Klein, D. The Tumor Vascular Endothelium as Decision Maker in Cancer Therapy. Front. Oncol. 2018, 8, 367. [Google Scholar] [CrossRef] [PubMed]
  428. Weishaupt, C.; Munoz, K.N.; Buzney, E.; Kupper, T.S.; Fuhlbrigge, R.C. T-cell distribution and adhesion receptor expression in metastatic melanoma. Clin. Cancer Res. 2007, 13, 2549–2556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  429. Afanasiev, O.K.; Nagase, K.; Simonson, W.; Vandeven, N.; Blom, A.; Koelle, D.M.; Clark, R.; Nghiem, P. Vascular E-selectin expression correlates with CD8 lymphocyte infiltration and improved outcome in Merkel cell carcinoma. J. Investig. Dermatol. 2013, 133, 2065–2073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  430. Iozzo, R.V.; Schaefer, L. Proteoglycan form and function: A comprehensive nomenclature of proteoglycans. Matrix Biol. 2015, 42, 11–55. [Google Scholar] [CrossRef]
  431. Haspel, J.; Grumet, M. The L1CAM extracellular region: A multi-domain protein with modular and cooperative binding modes. Front. Biosci. 2003, 8, s1210–s1225. [Google Scholar] [CrossRef] [Green Version]
  432. Pollerberg, G.E.; Thelen, K.; Theiss, M.O.; Hochlehnert, B.C. The role of cell adhesion molecules for navigating axons: Density matters. Mech. Dev. 2013, 130, 359–372. [Google Scholar] [CrossRef]
  433. Fujimoto, I.; Bruses, J.L.; Rutishauser, U. Regulation of cell adhesion by polysialic acid. Effects on cadherin, immunoglobulin cell adhesion molecule, and integrin function and independence from neural cell adhesion molecule binding or signaling activity. J. Biol. Chem. 2001, 276, 31745–31751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  434. Nabatov, A.A.; Raginov, I.S. The DC-SIGN-CD56 interaction inhibits the anti-dendritic cell cytotoxicity of CD56 expressing cells. Infect. Agents Cancer 2015, 10, 49. [Google Scholar] [CrossRef]
  435. Chung, K.Y.; Leung, K.M.; Lin, C.C.; Tam, K.C.; Hao, Y.L.; Taylor, J.S.; Chan, S.O. Regionally specific expression of L1 and sialylated NCAM in the retinofugal pathway of mouse embryos. J. Comp. Neurol. 2004, 471, 482–498. [Google Scholar] [CrossRef]
  436. Trouillas, J.; Daniel, L.; Guigard, M.P.; Tong, S.; Gouvernet, J.; Jouanneau, E.; Jan, M.; Perrin, G.; Fischer, G.; Tabarin, A.; et al. Polysialylated neural cell adhesion molecules expressed in human pituitary tumors and related to extrasellar invasion. J. Neurosurg. 2003, 98, 1084–1093. [Google Scholar] [CrossRef]
  437. Fernandez-Briera, A.; Garcia-Parceiro, I.; Cuevas, E.; Gil-Martin, E. Effect of human colorectal carcinogenesis on the neural cell adhesion molecule expression and polysialylation. Oncology-Basel 2010, 78, 196–204. [Google Scholar] [CrossRef] [PubMed]
  438. Binder, C.; Cvetkovski, F.; Sellberg, F.; Berg, S.; Paternina Visbal, H.; Sachs, D.H.; Berglund, E.; Berglund, D. CD2 Immunobiology. Front. Immunol. 2020, 11, 1090. [Google Scholar] [CrossRef] [PubMed]
  439. Hamai, A.; Meslin, F.; Benlalam, H.; Jalil, A.; Mehrpour, M.; Faure, F.; Lecluse, Y.; Vielh, P.; Avril, M.F.; Robert, C.; et al. ICAM-1 has a critical role in the regulation of metastatic melanoma tumor susceptibility to CTL lysis by interfering with PI3K/AKT pathway. Cancer Res. 2008, 68, 9854–9864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  440. Valgardsdottir, R.; Capitanio, C.; Texido, G.; Pende, D.; Cantoni, C.; Pesenti, E.; Rambaldi, A.; Golay, J.; Introna, M. Direct involvement of CD56 in cytokine-induced killer-mediated lysis of CD56+ hematopoietic target cells. Exp. Hematol. 2014, 42, 1013–1021.e1011. [Google Scholar] [CrossRef]
  441. McGreal, E.P.; Miller, J.L.; Gordon, S. Ligand recognition by antigen-presenting cell C-type lectin receptors. Curr. Opin. Immunol. 2005, 17, 18–24. [Google Scholar] [CrossRef] [PubMed]
  442. Khoo, U.S.; Chan, K.Y.; Chan, V.S.; Lin, C.L. DC-SIGN and L-SIGN: The SIGNs for infection. J. Mol. Med. 2008, 86, 861–874. [Google Scholar] [CrossRef] [PubMed]
  443. Leger, P.; Tetard, M.; Youness, B.; Cordes, N.; Rouxel, R.N.; Flamand, M.; Lozach, P.Y. Differential Use of the C-Type Lectins L-SIGN and DC-SIGN for Phlebovirus Endocytosis. Traffic 2016, 17, 639–656. [Google Scholar] [CrossRef]
  444. Lozach, P.Y.; Amara, A.; Bartosch, B.; Virelizier, J.L.; Arenzana-Seisdedos, F.; Cosset, F.L.; Altmeyer, R. C-type lectins L-SIGN and DC-SIGN capture and transmit infectious hepatitis C virus pseudotype particles. J. Biol. Chem. 2004, 279, 32035–32045. [Google Scholar] [CrossRef] [Green Version]
  445. Lozach, P.Y.; Kuhbacher, A.; Meier, R.; Mancini, R.; Bitto, D.; Bouloy, M.; Helenius, A. DC-SIGN as a receptor for phleboviruses. Cell Host Microbe 2011, 10, 75–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  446. Goncalves, A.R.; Moraz, M.L.; Pasquato, A.; Helenius, A.; Lozach, P.Y.; Kunz, S. Role of DC-SIGN in Lassa virus entry into human dendritic cells. J. Virol. 2013, 87, 11504–11515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  447. Curreli, S.; Arany, Z.; Gerardy-Schahn, R.; Mann, D.; Stamatos, N.M. Polysialylated neuropilin-2 is expressed on the surface of human dendritic cells and modulates dendritic cell-T lymphocyte interactions. J. Biol. Chem. 2007, 282, 30346–30356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  448. Abel, A.M.; Yang, C.; Thakar, M.S.; Malarkannan, S. Natural Killer Cells: Development, Maturation, and Clinical Utilization. Front. Immunol. 2018, 9, 1869. [Google Scholar] [CrossRef] [Green Version]
  449. Gumbiner, B.M. Regulation of cadherin-mediated adhesion in morphogenesis. Nat. Rev. Mol. Cell Biol. 2005, 6, 622–634. [Google Scholar] [CrossRef]
  450. Shapiro, L.; Fannon, A.M.; Kwong, P.D.; Thompson, A.; Lehmann, M.S.; Grubel, G.; Legrand, J.F.; Als-Nielsen, J.; Colman, D.R.; Hendrickson, W.A. Structural basis of cell-cell adhesion by cadherins. Nature 1995, 374, 327–337. [Google Scholar] [CrossRef]
  451. Marie, P.J.; Hay, E.; Modrowski, D.; Revollo, L.; Mbalaviele, G.; Civitelli, R. Cadherin-mediated cell-cell adhesion and signaling in the skeleton. Calcif. Tissue Int. 2014, 94, 46–54. [Google Scholar] [CrossRef] [Green Version]
  452. Takei, R.; Suzuki, D.; Hoshiba, T.; Nagaoka, M.; Seo, S.J.; Cho, C.S.; Akaike, T. Role of E-cadherin molecules in spheroid formation of hepatocytes adhered on galactose-carrying polymer as an artificial asialoglycoprotein model. Biotechnol. Lett. 2005, 27, 1149–1156. [Google Scholar] [CrossRef] [PubMed]
  453. Miyamoto, Y.; Sakane, F.; Hashimoto, K. N-cadherin-based adherens junction regulates the maintenance, proliferation, and differentiation of neural progenitor cells during development. Cell Adh. Migr. 2015, 9, 183–192. [Google Scholar] [CrossRef] [Green Version]
  454. Liu, T.; Guo, Z.; Yang, Q.; Sad, S.; Jennings, H.J. Biochemical engineering of surface alpha 2-8 polysialic acid for immunotargeting tumor cells. J. Biol. Chem. 2000, 275, 32832–32836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  455. Crossin, K.L.; Krushel, L.A. Cellular signaling by neural cell adhesion molecules of the immunoglobulin superfamily. Dev. Dynam. 2000, 218, 260–279. [Google Scholar] [CrossRef]
  456. Pruneri, G.; Vingiani, A.; Denkert, C. Tumor infiltrating lymphocytes in early breast cancer. Breast 2018, 37, 207–214. [Google Scholar] [CrossRef] [Green Version]
  457. Krpina, K.; Babarovic, E.; Spanjol, J.; Dordevic, G.; Maurer, T.; Jonjic, N. Correlation of tumor-associated macrophages and NK cells with bladder cancer size and T stage in patients with solitary low-grade urothelial carcinoma. Wien. Klin. Wochenschr. 2016, 128, 248–252. [Google Scholar] [CrossRef]
  458. Briercheck, E.L.; Trotta, R.; Chen, L.; Hartlage, A.S.; Cole, J.P.; Cole, T.D.; Mao, C.; Banerjee, P.P.; Hsu, H.T.; Mace, E.M.; et al. PTEN is a negative regulator of NK cell cytolytic function. J. Immunol. 2015, 194, 1832–1840. [Google Scholar] [CrossRef] [Green Version]
  459. Peng, L.X.; Liu, X.H.; Lu, B.; Liao, S.M.; Zhou, F.; Huang, J.M.; Chen, D.; Troy, F.A., II; Zhou, G.P.; Huang, R.B. The Inhibition of Polysialyltranseferase ST8SiaIV Through Heparin Binding to Polysialyltransferase Domain (PSTD). Med. Chem. 2019, 15, 486–495. [Google Scholar] [CrossRef] [PubMed]
  460. Dityatev, A.; Bukalo, O.; Schachner, M. Modulation of synaptic transmission and plasticity by cell adhesion and repulsion molecules. Neuron Glia Biol. 2008, 4, 197–209. [Google Scholar] [CrossRef] [PubMed]
  461. Schmid, R.S.; Maness, P.F. L1 and NCAM adhesion molecules as signaling coreceptors in neuronal migration and process outgrowth. Curr. Opin. Neurobiol. 2008, 18, 245–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  462. Schmid, R.S.; Pruitt, W.M.; Maness, P.F. A MAP kinase-signaling pathway mediates neurite outgrowth on L1 and requires Src-dependent endocytosis. J. Neurosci. 2000, 20, 4177–4188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  463. Angiolini, F.; Cavallaro, U. The Pleiotropic Role of L1CAM in Tumor Vasculature. Int. J. Mol. Sci. 2017, 18, 254. [Google Scholar] [CrossRef] [Green Version]
  464. Kleene, R.; Yang, H.; Kutsche, M.; Schachner, M. The neural recognition molecule L1 is a sialic acid-binding lectin for CD24, which induces promotion and inhibition of neurite outgrowth. J. Biol. Chem. 2001, 276, 21656–21663. [Google Scholar] [CrossRef] [Green Version]
  465. Buhusi, M.; Midkiff, B.R.; Gates, A.M.; Richter, M.; Schachner, M.; Maness, P.F. Close homolog of L1 is an enhancer of integrin-mediated cell migration. J. Biol. Chem. 2003, 278, 25024–25031. [Google Scholar] [CrossRef] [Green Version]
  466. Takei, K.; Chan, T.A.; Wang, F.S.; Deng, H.; Rutishauser, U.; Jay, D.G. The neural cell adhesion molecules L1 and NCAM-180 act in different steps of neurite outgrowth. J. Neurosci. 1999, 19, 9469–9479. [Google Scholar] [CrossRef]
  467. Mechtersheimer, S.; Gutwein, P.; Agmon-Levin, N.; Stoeck, A.; Oleszewski, M.; Riedle, S.; Postina, R.; Fahrenholz, F.; Fogel, M.; Lemmon, V.; et al. Ectodomain shedding of L1 adhesion molecule promotes cell migration by autocrine binding to integrins. J. Cell Biol. 2001, 155, 661–673. [Google Scholar] [CrossRef] [Green Version]
  468. Kiefel, H.; Bondong, S.; Pfeifer, M.; Schirmer, U.; Erbe-Hoffmann, N.; Schafer, H.; Sebens, S.; Altevogt, P. EMT-associated up-regulation of L1CAM provides insights into L1CAM-mediated integrin signalling and NF-kappaB activation. Carcinogenesis 2012, 33, 1919–1929. [Google Scholar] [CrossRef] [PubMed]
  469. Schmidt, C.; Künemund, V.; Wintergerst, E.S.; Schmitz, B.; Schachner, M. CD9 of mouse brain is implicated in neurite outgrowth and cell migration in vitro and is associated with the α6/β1 integrin and the neural adhesion molecule L1. J. Neurosci. Res. 1996, 43, 12–31. [Google Scholar] [CrossRef] [PubMed]
  470. Moulla, A.; Miliaras, D.; Sioga, A.; Kaidoglou, A.; Economou, L. The immunohistochemical expression of CD24 and CD171 adhesion molecules in borderline ovarian tumors. Pol. J. Pathol. 2013, 64, 180–184. [Google Scholar] [CrossRef] [PubMed]
  471. Stoeck, A.; Schlich, S.; Issa, Y.; Gschwend, V.; Wenger, T.; Herr, I.; Marme, A.; Bourbie, S.; Altevogt, P.; Gutwein, P. L1 on ovarian carcinoma cells is a binding partner for Neuropilin-1 on mesothelial cells. Cancer Lett. 2006, 239, 212–226. [Google Scholar] [CrossRef] [PubMed]
  472. Castellani, V.; De Angelis, E.; Kenwrick, S.; Rougon, G. Cis and trans interactions of L1 with neuropilin-1 control axonal responses to semaphorin 3A. EMBO J. 2002, 21, 6348–6357. [Google Scholar] [CrossRef] [Green Version]
  473. Mohanan, V.; Temburni, M.K.; Kappes, J.C.; Galileo, D.S. L1CAM stimulates glioma cell motility and proliferation through the fibroblast growth factor receptor. Clin. Exp. Metastasis 2013, 30, 507–520. [Google Scholar] [CrossRef]
  474. Shtutman, M.; Levina, E.; Ohouo, P.; Baig, M.; Roninson, I.B. Cell adhesion molecule L1 disrupts E-cadherin-containing adherens junctions and increases scattering and motility of MCF7 breast carcinoma cells. Cancer Res. 2006, 66, 11370–11380. [Google Scholar] [CrossRef] [Green Version]
  475. Nagasundaram, M.; Horstkorte, R.; Gnanapragassam, V.S. Sialic Acid Metabolic Engineering of Breast Cancer Cells Interferes with Adhesion and Migration. Molecules 2020, 25, 2632. [Google Scholar] [CrossRef]
  476. Kamiguchi, H.; Lemmon, V. A neuronal form of the cell adhesion molecule L1 contains a tyrosine-based signal required for sorting to the axonal growth cone. J. Neurosci. 1998, 18, 3749–3756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  477. Anderson, H.J.; Galileo, D.S. Small-molecule inhibitors of FGFR, integrins and FAK selectively decrease L1CAM-stimulated glioblastoma cell motility and proliferation. Cell Oncol. 2016, 39, 229–242. [Google Scholar] [CrossRef] [PubMed]
  478. Sytnyk, V.; Leshchyns’ka, I.; Schachner, M. Neural Cell Adhesion Molecules of the Immunoglobulin Superfamily Regulate Synapse Formation, Maintenance, and Function. Trends Neurosci. 2017, 40, 295–308. [Google Scholar] [CrossRef] [PubMed]
  479. Angiolini, F.; Belloni, E.; Giordano, M.; Campioni, M.; Forneris, F.; Paronetto, M.P.; Lupia, M.; Brandas, C.; Pradella, D.; Di Matteo, A.; et al. A novel L1CAM isoform with angiogenic activity generated by NOVA2-mediated alternative splicing. eLife 2019, 8, e44305. [Google Scholar] [CrossRef]
  480. He, Z. Crossed wires: L1 and neuropilin interactions. Neuron 2000, 27, 191–193. [Google Scholar] [CrossRef] [Green Version]
  481. McNutt, M. Cancer immunotherapy. Science 2013, 342, 1417. [Google Scholar] [CrossRef]
  482. Rosenberg, S.A. Progress in human tumour immunology and immunotherapy. Nature 2001, 411, 380–384. [Google Scholar] [CrossRef]
  483. Morise, J.; Takematsu, H.; Oka, S. The role of human natural killer-1 (HNK-1) carbohydrate in neuronal plasticity and disease. Biochim. Et Biophys. Acta-Gen. Subj. 2017, 1861, 2455–2461. [Google Scholar] [CrossRef] [PubMed]
  484. Suzuki-Anekoji, M.; Suzuki, M.; Kobayashi, T.; Sato, Y.; Nakayama, J.; Suzuki, A.; Bao, X.; Angata, K.; Fukuda, M. HNK-1 glycan functions as a tumor suppressor for astrocytic tumor. J. Biol. Chem. 2011, 286, 32824–32833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Prisma Flow Diagram.
Scheme 1. Prisma Flow Diagram.
Cancers 13 05203 sch001
Figure 2. (a) Polysialylated glycoproteins or proteoglycans have a regulatory function in neuronal development. Association of multifunctional, polysialylated glycoproteins (e.g., NRP-2) or proteoglycans (e.g., poly-glucosaminylated NRP-1) with growth factor receptors (e.g., plexins) regulates several functions in neuronal cell development including apoptosis (induction), proliferation (inhibition), and migration (inhibition). Shown are the intra-, transmembrane-, and extracellular domains, as well as potential glycosylation sites (PSA). (b) Polysialylated glycoproteins or proteoglycans regulate various functions in association with growth factor receptors in tumor cells. Association of multifunctional, polysialylated glycoproteins (e.g., NRP-2) or proteoglycans (e.g., polyglucosaminylated NRP-1 with growth factor receptors (e.g., VEGFR1/2/3) modulates tumor progression/immune escape. Complex formation by polysialylated neuropilin receptor NRP-2 and its co-receptor NRP-1 with VEGF receptor via galectin induced stabilization of the complex will foster tumor progression and immune escape. Ligands of the various receptors/complexes are listed in the green boxes above the transmembrane proteins. Their signaling results in effects as shown in the boxes below the respective proteins. The respective signals lead to changes in cell survival, lymph-angiogenesis, migration, cytoskeleton rearrangement, invasion, EMT, and metastasis, as indicated in boxes below the lipid membrane. These interactions are crucial for vascularization and angiogenesis. Finally, galectin-1 interacts with neuropilin-1/vascular endothelial growth factor-2 (NPR-1/VEGFR-2) complexes to advance endothelial cells migration.
Figure 2. (a) Polysialylated glycoproteins or proteoglycans have a regulatory function in neuronal development. Association of multifunctional, polysialylated glycoproteins (e.g., NRP-2) or proteoglycans (e.g., poly-glucosaminylated NRP-1) with growth factor receptors (e.g., plexins) regulates several functions in neuronal cell development including apoptosis (induction), proliferation (inhibition), and migration (inhibition). Shown are the intra-, transmembrane-, and extracellular domains, as well as potential glycosylation sites (PSA). (b) Polysialylated glycoproteins or proteoglycans regulate various functions in association with growth factor receptors in tumor cells. Association of multifunctional, polysialylated glycoproteins (e.g., NRP-2) or proteoglycans (e.g., polyglucosaminylated NRP-1 with growth factor receptors (e.g., VEGFR1/2/3) modulates tumor progression/immune escape. Complex formation by polysialylated neuropilin receptor NRP-2 and its co-receptor NRP-1 with VEGF receptor via galectin induced stabilization of the complex will foster tumor progression and immune escape. Ligands of the various receptors/complexes are listed in the green boxes above the transmembrane proteins. Their signaling results in effects as shown in the boxes below the respective proteins. The respective signals lead to changes in cell survival, lymph-angiogenesis, migration, cytoskeleton rearrangement, invasion, EMT, and metastasis, as indicated in boxes below the lipid membrane. These interactions are crucial for vascularization and angiogenesis. Finally, galectin-1 interacts with neuropilin-1/vascular endothelial growth factor-2 (NPR-1/VEGFR-2) complexes to advance endothelial cells migration.
Cancers 13 05203 g002aCancers 13 05203 g002b
Figure 3. (a) Highly sialylated glycoproteins stimulate human tyrosine kinase receptors. A series of human tyrosine kinase (growth factor) receptors including (from left to right) PTK7 (pseudo tyrosine kinase, receptor7, CCK4), VEGFR1/2/3 (vascular endothelial growth factor receptor 1/2/3), PDGFRα/β ( platelet derived growth factor receptor α/β; CSF1R, KIT, FLK2), FGFR1/2/3/4 (fibroblast growth factor receptor 1/2/3/4), c-Met (tyrosine-protein kinase Met or hepatocyte growth factor receptor (HGFR); Ron, Sea), TrKA/B (tropomyosin receptor kinase A/B), RORα/β (RAR (retinoic acid receptor)-related orphan receptor-α/β), MuSK (muscle-specific kinase), TieR1/2 (tyrosine kinase with immunoglobulin-like and EGF-like domains 1/2, angiopoietin receptor), and AXL (AXL receptor tyrosine kinase). Published interaction partners are listed above the respective growth factor receptors. These interactions arguably influence cellular properties, including proliferation, cell survival, differentiation, apoptosis, metastasis, migration, and angiogenesis, as listed below the growth factor receptors. The bottom part describes symbols used to characterize the growth factor receptors. (b) Highly sialylated glycoproteins stimulate human tyrosine kinase receptors. A series of human tyrosine kinase (growth factor) receptors including (from left to right) ErbB-1 (EGFR, epithelial growth factor receptor), ErbB-2/3/4 (HER2/3/4) [120], INSR (insulin receptor; IGF1R, insulin-like growth factor 1 receptor); ALK (anaplastic lymphoma kinase; LTK, leukocyte tyrosine kinase), DDR1/2 (discoidin domain receptor 1/2), Ros (proto-oncogene receptor tyrosine kinase), TGF-βR (transforming growth factor-β receptor), and plexin A, B, C, D. Published interaction partners are listed above the respective growth factor receptors. These interactions arguably influence cellular properties, including proliferation, cell survival, differentiation, apoptosis, metastasis, migration, and angiogenesis, as listed below the growth factor receptors. The bottom part describes symbols used to characterize the growth factor receptors.
Figure 3. (a) Highly sialylated glycoproteins stimulate human tyrosine kinase receptors. A series of human tyrosine kinase (growth factor) receptors including (from left to right) PTK7 (pseudo tyrosine kinase, receptor7, CCK4), VEGFR1/2/3 (vascular endothelial growth factor receptor 1/2/3), PDGFRα/β ( platelet derived growth factor receptor α/β; CSF1R, KIT, FLK2), FGFR1/2/3/4 (fibroblast growth factor receptor 1/2/3/4), c-Met (tyrosine-protein kinase Met or hepatocyte growth factor receptor (HGFR); Ron, Sea), TrKA/B (tropomyosin receptor kinase A/B), RORα/β (RAR (retinoic acid receptor)-related orphan receptor-α/β), MuSK (muscle-specific kinase), TieR1/2 (tyrosine kinase with immunoglobulin-like and EGF-like domains 1/2, angiopoietin receptor), and AXL (AXL receptor tyrosine kinase). Published interaction partners are listed above the respective growth factor receptors. These interactions arguably influence cellular properties, including proliferation, cell survival, differentiation, apoptosis, metastasis, migration, and angiogenesis, as listed below the growth factor receptors. The bottom part describes symbols used to characterize the growth factor receptors. (b) Highly sialylated glycoproteins stimulate human tyrosine kinase receptors. A series of human tyrosine kinase (growth factor) receptors including (from left to right) ErbB-1 (EGFR, epithelial growth factor receptor), ErbB-2/3/4 (HER2/3/4) [120], INSR (insulin receptor; IGF1R, insulin-like growth factor 1 receptor); ALK (anaplastic lymphoma kinase; LTK, leukocyte tyrosine kinase), DDR1/2 (discoidin domain receptor 1/2), Ros (proto-oncogene receptor tyrosine kinase), TGF-βR (transforming growth factor-β receptor), and plexin A, B, C, D. Published interaction partners are listed above the respective growth factor receptors. These interactions arguably influence cellular properties, including proliferation, cell survival, differentiation, apoptosis, metastasis, migration, and angiogenesis, as listed below the growth factor receptors. The bottom part describes symbols used to characterize the growth factor receptors.
Cancers 13 05203 g003
Figure 4. Specific relations of siglecs with their sialylated ligands and resulting functions: The important inhibitory siglecs are shown with their intracellular-, transmembrane-, and extracellular domains (bottom part). The Siglecs‘ roles depend on their specific binding domain (V-Set Ig domain) as well as on intracellular inhibitory motives (GRB2, ITIM-like, ITIM, FYN kinase phosphorylation site; see legend). Cells, which express the various siglecs are given below the respective symbols (Mac: macrophage, DC: dendritic cell, Mon: monocyte, MyPr: myeloid precursor, Neu: neutrophilic cell, B: B-cell, Trophob: trophoblast cell, NK: natural killer cell, Eos: eosinophilic cell, Bas: basophilic cell, Mast: mast cell, Epithel: epithelial cell). Examples of ligands binding to the siglecs are indicated above their symbols. Gp120 of HIV: glycoprotein 120 from HIV capsid, SR-BI: Scavenger receptor class B type 1, PRRSV: porcine reproductive and respiratory syndrome virus, MUC1: Mucin 1, CD43: semaphorin, sialyl-lactose containing gangliosides, Siaα2,3 > Siaα2,6: Preferred sialic acid linkage type of respective binding partners. Further ligands of other siglecs: BCR: B-cell receptor, CD45, IgM, GD1: ganglioside 1, GT1b: ganglioside, carbohydrates from selected gram-negative bacteria (e.g., Neisseria meningitidis, Campylobacter jejuni, Pseudomonas aeruginosa, E. coli K1, etc.) and some gram-positive bacteria (e.g., Streptococcus spp.) that express sialic acid. In addition, the preferential sialic acid linkages are indicated. The role of siglecs is indicated in the blue boxes (bottom). The cytoplasmic domain of inhibitory siglecs is characterized by ITIMs, which recruit and activate phosphatases. In B-cells, sialic acid binding to siglec-2 or siglec-G/10 will recruit the tyrosine phosphatase SHP-1 to their cytoplasmic ITIM and inhibits BCR signal transduction [173]. These phosphatases then negatively regulate B cell antigen receptor signaling and thus antibody production [174,175,176].
Figure 4. Specific relations of siglecs with their sialylated ligands and resulting functions: The important inhibitory siglecs are shown with their intracellular-, transmembrane-, and extracellular domains (bottom part). The Siglecs‘ roles depend on their specific binding domain (V-Set Ig domain) as well as on intracellular inhibitory motives (GRB2, ITIM-like, ITIM, FYN kinase phosphorylation site; see legend). Cells, which express the various siglecs are given below the respective symbols (Mac: macrophage, DC: dendritic cell, Mon: monocyte, MyPr: myeloid precursor, Neu: neutrophilic cell, B: B-cell, Trophob: trophoblast cell, NK: natural killer cell, Eos: eosinophilic cell, Bas: basophilic cell, Mast: mast cell, Epithel: epithelial cell). Examples of ligands binding to the siglecs are indicated above their symbols. Gp120 of HIV: glycoprotein 120 from HIV capsid, SR-BI: Scavenger receptor class B type 1, PRRSV: porcine reproductive and respiratory syndrome virus, MUC1: Mucin 1, CD43: semaphorin, sialyl-lactose containing gangliosides, Siaα2,3 > Siaα2,6: Preferred sialic acid linkage type of respective binding partners. Further ligands of other siglecs: BCR: B-cell receptor, CD45, IgM, GD1: ganglioside 1, GT1b: ganglioside, carbohydrates from selected gram-negative bacteria (e.g., Neisseria meningitidis, Campylobacter jejuni, Pseudomonas aeruginosa, E. coli K1, etc.) and some gram-positive bacteria (e.g., Streptococcus spp.) that express sialic acid. In addition, the preferential sialic acid linkages are indicated. The role of siglecs is indicated in the blue boxes (bottom). The cytoplasmic domain of inhibitory siglecs is characterized by ITIMs, which recruit and activate phosphatases. In B-cells, sialic acid binding to siglec-2 or siglec-G/10 will recruit the tyrosine phosphatase SHP-1 to their cytoplasmic ITIM and inhibits BCR signal transduction [173]. These phosphatases then negatively regulate B cell antigen receptor signaling and thus antibody production [174,175,176].
Cancers 13 05203 g004
Figure 6. (a) Low-sialylation effects in apoptotic signaling: Reduced (or lack of) sialylation permits cluster formation of apoptosis inducing receptors via galectins, which results into strong apoptotic signaling. Apoptosis-inducing receptors include: sTRAIL1/2 (DR4, DR5), TNFαR, Fas/CD95. By interaction with these receptors, Gal 3 induces apoptosis in the absence of sialylation/glucosamination (e.g., integrin α4β7/VLD-4 in conjunction with CD7, or complex of CD43 and CD45 with CD7). However, cytosolic Gal 3 will block the downstream signaling cascade leading to apoptosis. The bottom part lists some of the related signaling cascades. Cluster formation of apoptosis-inducing receptors is mediated by Gal 1, which bridges the subunits of a trimeric cluster. This is exemplified for CD95 (top left). (b) Sialylation effects in apoptotic signaling: Upregulated sialylation in tumor cells inhibits cluster formation of neighboring apoptosis-inducing receptors (e.g., sTRAIL1/2 (DR4, DR5), TNFαR, Fas/CD95) and results in poor apoptotic signaling and subsequently in immune escape. Increased α2,3-Sia or α2,6-sialylation of glycoproteins (e.g., integrin α4β7/VLD-4 in conjunction with CD7, or complex of CD43 and CD45 with CD7) will inhibit complex formation and binding between extracellular Gal 3 or Gal 1 and galactose residues, thereby preventing apoptosis signaling. Cluster formation of apoptosis-inducing receptors is inhibited by increased sialylation; therefore Gal 1 cannot bridge the subunits of trimeric clusters. This is exemplified for CD95 (top left).
Figure 6. (a) Low-sialylation effects in apoptotic signaling: Reduced (or lack of) sialylation permits cluster formation of apoptosis inducing receptors via galectins, which results into strong apoptotic signaling. Apoptosis-inducing receptors include: sTRAIL1/2 (DR4, DR5), TNFαR, Fas/CD95. By interaction with these receptors, Gal 3 induces apoptosis in the absence of sialylation/glucosamination (e.g., integrin α4β7/VLD-4 in conjunction with CD7, or complex of CD43 and CD45 with CD7). However, cytosolic Gal 3 will block the downstream signaling cascade leading to apoptosis. The bottom part lists some of the related signaling cascades. Cluster formation of apoptosis-inducing receptors is mediated by Gal 1, which bridges the subunits of a trimeric cluster. This is exemplified for CD95 (top left). (b) Sialylation effects in apoptotic signaling: Upregulated sialylation in tumor cells inhibits cluster formation of neighboring apoptosis-inducing receptors (e.g., sTRAIL1/2 (DR4, DR5), TNFαR, Fas/CD95) and results in poor apoptotic signaling and subsequently in immune escape. Increased α2,3-Sia or α2,6-sialylation of glycoproteins (e.g., integrin α4β7/VLD-4 in conjunction with CD7, or complex of CD43 and CD45 with CD7) will inhibit complex formation and binding between extracellular Gal 3 or Gal 1 and galactose residues, thereby preventing apoptosis signaling. Cluster formation of apoptosis-inducing receptors is inhibited by increased sialylation; therefore Gal 1 cannot bridge the subunits of trimeric clusters. This is exemplified for CD95 (top left).
Cancers 13 05203 g006aCancers 13 05203 g006b
Table 1. Overview of mammalian lectin families, their ligands, subcellular location, and functions with examples.
Table 1. Overview of mammalian lectin families, their ligands, subcellular location, and functions with examples.
Lectin FamilySaccharide LigandsSubcellular LocationExamples of FunctionsExample-Proteins
C-Type-Lectin
Calcium-dependent
Man6Ps (a), Gal (b), N-acetylgalactosamine (Sialic acid), Fuc (c), etc.Cell membrane, extracellular Cell adhesion, (selectins), Glycoprotein clearance, Innate immunity (Collectins)DC-SIGN (d), Dectin1, NKG2D (e), REG3 (f) proteins, L-Selectin, MMR (g), MBL (h)-protein, Tetranectin, PKD1 (i), Thrombomodulin, Attractin (DPPT-L), Human Macrophage Galactose-Type Lectin (MGL) (j)
Siglecs (k) (I-Type-Lectin)Sialic acidCell membraneMolecular and cell recognitionSiglec1, Siglec2, Siglec4 & Siglecs 3-16
F-Type Lectin (fucose binding lectins)Fuc terminiExtracellularInnate immunityL-fucose binding proteins
F-box LectinGlcNAc2CytoplasmDegradation of misfolded proteinsβ-TRCP1 and β-TRCP2 (l) proteins
L-Type-LectinMan6Ps, Gal, Sialic acid, Fuc, etc.ER (m), ERGIC (n), GolgiProtein sorting in the ERERGIC-53, ERGL (o), VIP36 (p), and VIPL (q)
M-Type-LectinMan8PsERER-associated degradationEDEM1 (r), EDEM2, EDEM3
P-Type-LectinMan6Ps and othersSecretory pathwayPost-Golgi glycoprotein traffickingMannose 6-phosphate receptor (M6Ps)
R-Type-LectinMan6Ps, Gal, Sialic acid, Xyl (s)Golgi, Cell membraneEnzyme targeting, hormone turnoverMannose receptor family e.g., DEC-205
S-Type-Lectin (Galectins)β-GalactosidesCytoplasm, extracellularCell surface crosslinkingGalectin-1, -2, -5, -7, -10, -11, -14 and -15
X-type-Lectin (Intelectins)Gal, galactofuranose, pentoseCell membrane, extracellularInnate immunity, fertilization and embryogenesisIntelectin-1 and 2
FicolinsSialic acid, GlcNAc, GalNAcCell membrane, extracellularInnate immunityH-ficolin and M-ficolin
Calnexin familyGlc1Man9ERProtein sorting in the ERCalnexin, calmegin, calreticulin
Chitinase like-LectinChito (t)-oligosaccharidesExtracellularCollagen metabolismChitinase 3-like 1
(a) Mannose-6-phosphate; (b) Galactose; (c) Fucose; (d) Dendritic cell-specific intercellular adhesion molecule-3-grabbing non-integrin (DC-SIGN); (e) Natural killer group 2D (NKG2D); (f) Regenerating islet-derived protein 3 (REG3); (g) DNA mismatch repair (MMR) proteins; (h) Mannose-binding lectin (MBL); (i) Polycystic kidney disease 1 (PKD1); (j) Human Macrophage Galactose-Type Lectin (MGL) is an oligomeric type II transmembrane protein, which is expressed on macrophages, monocytes, and dendritic cells and activates these cells. MGL binds terminal α- & β-linked GalNAc residues on glycoproteins, glycolipids, and bacterial LPS, including Tn antigen and GalNAcβ1-4GlcNAc-R (LDN) antigens. (k) Sialic acid-binding Ig-like lectins (siglec); (l) β-Transducin repeat containing protein (β-TRCP); (m) Endoplasmic reticulum; (n) ER-Golgi intermediate compartment (ERGIC); (o) Lectin, mannose binding 1 like (LMAN1L, ERGL); (p) Vesicular integral-membrane protein of 36 kDa (VIP36); (q) VIP36-ligand; (r) ER degradation-enhancing alpha-mannosidase-like protein (EDEM); (s) D-Xylose (t) Chitin or Chitosan.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jarahian, M.; Marofi, F.; Maashi, M.S.; Ghaebi, M.; Khezri, A.; Berger, M.R. Re-Expression of Poly/Oligo-Sialylated Adhesion Molecules on the Surface of Tumor Cells Disrupts Their Interaction with Immune-Effector Cells and Contributes to Pathophysiological Immune Escape. Cancers 2021, 13, 5203. https://doi.org/10.3390/cancers13205203

AMA Style

Jarahian M, Marofi F, Maashi MS, Ghaebi M, Khezri A, Berger MR. Re-Expression of Poly/Oligo-Sialylated Adhesion Molecules on the Surface of Tumor Cells Disrupts Their Interaction with Immune-Effector Cells and Contributes to Pathophysiological Immune Escape. Cancers. 2021; 13(20):5203. https://doi.org/10.3390/cancers13205203

Chicago/Turabian Style

Jarahian, Mostafa, Faroogh Marofi, Marwah Suliman Maashi, Mahnaz Ghaebi, Abdolrahman Khezri, and Martin R. Berger. 2021. "Re-Expression of Poly/Oligo-Sialylated Adhesion Molecules on the Surface of Tumor Cells Disrupts Their Interaction with Immune-Effector Cells and Contributes to Pathophysiological Immune Escape" Cancers 13, no. 20: 5203. https://doi.org/10.3390/cancers13205203

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop