Next Article in Journal / Special Issue
The Roles of Direct Recognition by Animal Lectins in Antiviral Immunity and Viral Pathogenesis
Previous Article in Journal
The Anti-Addiction Drug Ibogaine and the Heart: A Delicate Relation
Previous Article in Special Issue
Insecticidal Activity of Plant Lectins and Potential Application in Crop Protection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Human Lectins and Their Roles in Viral Infections

by
Christopher P. Mason
and
Alexander W. Tarr
*
School of Life Sciences and Biomedical Research Unit in Gastroenterology, Faculty of Medicine and Health Sciences, University of Nottingham, Nottingham NG7 2UH, UK
*
Author to whom correspondence should be addressed.
Molecules 2015, 20(2), 2229-2271; https://doi.org/10.3390/molecules20022229
Submission received: 17 November 2014 / Revised: 21 January 2015 / Accepted: 23 January 2015 / Published: 29 January 2015
(This article belongs to the Special Issue Lectins)

Abstract

:
Innate recognition of virus proteins is an important component of the immune response to viral pathogens. A component of this immune recognition is the family of lectins; pattern recognition receptors (PRRs) that recognise viral pathogen-associated molecular patterns (PAMPs) including viral glycoproteins. In this review we discuss the contribution of soluble and membrane-associated PRRs to immunity against virus pathogens, and the potential role of these molecules in facilitating virus replication. These processes are illustrated with examples of viruses including human immunodeficiency virus (HIV), hepatitis C virus (HCV) and Ebola virus (EBOV). We focus on the structure, function and genetics of the well-characterised C-type lectin mannose-binding lectin, the ficolins, and the membrane-bound CD209 proteins expressed on dendritic cells. The potential for lectin-based antiviral therapies is also discussed.

1. Introduction

Lectins are a diverse group of proteins broadly defined as non-immunoglobulin proteins that exhibit high avidity for glycoprotein- and/or glycolipid-associated carbohydrates, but display no enzymatic activity [1]. Genes encoding lectins been identified in all forms of life, including plants, animals, and viruses, an indication of their evolutionary conservation [1]. Indeed, phylogenetic studies indicated that the primitive immune system depended on lectin-protease-mediated opsonophagocytosis [2]. Lectins differ in tissue expression, ligand affinities, structure and function, and are classified by the phylogeny and primary and tertiary amino acid structures of their carbohydrate-recognition domains (CRD). However there are several inconsistencies in this classification system (reviewed in [3]).
Glycosylation is a form of post-translational modification in which glycans—monosaccharides or oligosaccharides—are glycosidically bonded to an organic molecule (reviewed in [4]). This modification plays an essential role in the expression and function of many proteins in eukaryotes and prokaryote cells, with roles in, for example, inter-compound interaction and pathogenic immune evasion. Eukaryotic and viral glycosylation occurs in the host cell endoplasmic reticulum (ER) and Golgi apparatus. Lectin-glycan interactions are generally achieved by hydrogen bonding and Van der Waals forces, and often depend on cations and multivalent interactions between multiple CRDs and multiple, clustered target glycans in order to achieve sufficient affinities for lectin activity [3].
The roles of human lectins include protein modulation, cell growth and homeostasis [4]. As glycoproteins are found on the surfaces of several pathogens to a diverse and widespread degree, some lectins act as pattern-recognition receptors (PRRs), recognising pathogen-associated molecular patterns (PAMPs)—including glycans and nucleic acid—related to invading microorganisms and malignant, apoptotic or dead host cells. This can lead to the induction of an immune response against the invading pathogen. However, the relationship between lectins and viruses is complex. In addition to immune evasion, glycosylation is essential for protein expression, assembly and entry steps in virus replication cycles [5], and many viruses have evolved mechanisms to exploit lectins to enhance infection.
This review focuses on human lectins and their roles during viral infections, concentrating on the well-described lectins mannose-binding lectin (MBL), ficolins and dendritic cell-specific ICAM-3 grabbing non-integrin (DC-SIGN). It highlights the different effects of lectins on viral infections and the consequences of genetic variation in lectin genes on susceptibility to virus infections.

2. The Complement Cascade

Complement contributes towards the initial defence against viral infections through a sequential protein activation cascade (reviewed in [6]). There are three known complement activation pathways. Each pathway converges at the formation of C3 convertase, which activates downstream complement factors to constitute the membrane attack complex (MAC). The MAC subsequently forms pores in the lipid membranes of pathogens and infected cells, causing osmotic lysis. Parallel to the cascade, some complement cleavage products mediate inflammation and opsonise pathogens, attracting phagocytes, encouraging antigen aggregation and preventing viral entry. To avoid potentially damaging, excessive complement activity, stringent regulation mechanisms have evolved, including complement factor cleavage and endocytotic shedding of MACs.
The alternate and classical complement pathways are triggered by foreign surface molecules and antigen:antibody complexes, respectively. The lectin pathway involves the binding of microbial surface carbohydrate moieties to serum lectins, which activates lectin-bound MBL-associated serine proteases (MASPs) and proteins (MAPs) (Figure 1).
Figure 1. The lectin pathway of complement activation. MBL and ficolins undergo conformational changes upon interaction with viral glycoproteins via glycan-associated mannose and N-acetylglucosamine residues, respectively. Sugars are labelled in accordance with reference [7]. This activates MASP-1 followed by MASP-2, which initiates a cleavage cascade of complement factors, with roles in opsonisation, inflammation and pathogen and infected cell lysis.
Figure 1. The lectin pathway of complement activation. MBL and ficolins undergo conformational changes upon interaction with viral glycoproteins via glycan-associated mannose and N-acetylglucosamine residues, respectively. Sugars are labelled in accordance with reference [7]. This activates MASP-1 followed by MASP-2, which initiates a cleavage cascade of complement factors, with roles in opsonisation, inflammation and pathogen and infected cell lysis.
Molecules 20 02229 g001

The MBL-Associated Serine Proteases

MASP-1, MASP-3 and MAP-1 are alternative splice products of the MASP1 gene [8,9], while MASP-2 and sMAP are encoded by the MASP2 gene [10]. MASP-1 and MASP-2 cleave C3 and C4, respectively, while C2 is cleaved by both [11]. MASP-2 can produce the C3 convertase C4bC2a and thus activate complement [6] (Figure 1).
Upon ligand binding, lectins undergo a conformational change that brings MASP serine protease domains within close proximity of each other, thus allowing proteolytic autoactivation [12,13,14,15] via the cleavage of an arginine-isoleucine bond in the serine protease domain [16].
MASP-2 activation appears to be primarily dependent on MASP-1, which may autoactivate then trans-activate the MASP-2 proenzyme [17]. This might be mediated either by heterodimeric MASP complexes, MBL-MASP-1-MASP-2 co-complexes, or separate MBL-MASP complexes [18,19]. If these complexes are separate, the binding sites of MASP-2 and MASP-1 and -3 on the lectin are likely to be within close proximity and overlapping, but not identical [20].
sMAP and MAP-1 are truncated MASP proteins lacking serine protease domains, and have putative roles in the modulation of MASP-associated complement activation [9,10,21]. MAP-1 may inhibit MASP-2 activation by the disruption of inter-MASP and lectin-MASP co-complexes [18,22].
The MASPs may also associate the lectin pathway of complement with the coagulation system [23,24]. MASP-3 has a putative regulatory function of MASP-2 activity [8], and—due to its involvement in the facial dysmorphic 3MC syndrome—roles for the protein in embryonic development have been hypothesised [25].

3. Mannose-Binding Lectin (MBL)

3.1. Genetics, Structure, Expression and Binding Specificities of MBL

MBL is a soluble, Ca2+-dependent protein of the collectin family–characterised as C-type lectins with collagenous domains–encoded by the MBL2 gene on human chromosome 10q11.2–10q21 [26]. MBL is primarily expressed in the liver and secreted into the blood, however lower expression has been detected elsewhere, including in mammalian muscle tissue and brain, often following immune challenge [27].
MBL monomers are 32 kDa in molecular weight and possess a typical collectin structure: an N-terminal cysteine-rich domain, a collagen-like domain (CLD) of approximately 20 Gly-Xaa-Yaa tandem repeats, a neck region and a CRD responsible for ligand binding (Figure 2a) [28]. These monomers form homotrimeric subunits that further oligomerise into trimeric to hexameric structures that can activate the complement cascade [29]. Trimeric and tetrameric MBL are the most common physiological configurations (Figure 3a) [30].
The CLDs form MBL trimers by hydrophobic interactions, as initiated by α-helical triple coiled coil formation of the neck domain and stabilised by inter- and intra-monomer bridges via 3 N-terminal cysteines [28,30,31]. Oligomeric MBL arranges into a sertiform structure, with trimeric subunits stretching out from the short, bundled N-terminal regions [30,32]. A short amino acid sequence after the first 7 N-terminal Gly-Xaa-Yaa repeats creates a “kink” bend in the CLD, however this is only seen in a minority of MBL molecules in vivo and does not influence MBL activity [30,32]. Significant flexibility exists at the CLD-neck and CLD-N-terminus regions, influencing CRD positioning, ligand specificity and MASP interaction [30,32].
MASP binding is centred around a conserved lysine occupying the Xaa position of a CLD repeat [33]. Ligand binding induces a “stretching” event which splays the MBL trimeric subunits and brings MASPs together to enable proteolytic autoactivation [12,13,14,15]. Trimeric/tetrameric MBL may represent the optimal configuration to accommodate MASP auto-activation [30,32].
MBL exhibits specificity for pairs of adjacent equatorial monosaccharide 3- and 4-hydroxyl groups, present in terminal mannose, N-acetylglucosamine (GlcNAc), N-acetylmannosamine and L-fucose oligosaccharides [34]. This is consistent with the presence of a conserved glutamate-proline-asparagine (EPN) motif in the CRD. MBL also binds phospholipids and nucleic acids, supporting a role in clearance of necrotic tissue [35,36]. MBL may employ higher order oligomers in order to achieve sufficient ligand binding via multivalent bonding [37].
Figure 2. Exon and monomeric structures of (a) MBL; (b) ficolins; (c) DC-SIGN. MBL and ficolins possess 3/2 N-terminal cysteines and a lysine in its CLD, important in oligomerisation and MASP/phagocyte interaction respectively. DC-SIGN contains N-terminal, cytoplasmic dileucine and tyrosine internalisation motifs.
Figure 2. Exon and monomeric structures of (a) MBL; (b) ficolins; (c) DC-SIGN. MBL and ficolins possess 3/2 N-terminal cysteines and a lysine in its CLD, important in oligomerisation and MASP/phagocyte interaction respectively. DC-SIGN contains N-terminal, cytoplasmic dileucine and tyrosine internalisation motifs.
Molecules 20 02229 g002
Figure 3. Active, oligomeric structures of (a) MBL; (b) M-/L-ficolin; (c) H-ficolin; (d) DC-SIGN. MBL and DC-SIGN possess C-terminal carbohydrate-recognition domains (CRD) whereas ficolins possess fibrinogen-like domains (FBG). MBL and ficolin possess collagen-like domains and consist of trimeric subunits. DC-SIGN possesses a neck region and transmembrane (TM) domain.
Figure 3. Active, oligomeric structures of (a) MBL; (b) M-/L-ficolin; (c) H-ficolin; (d) DC-SIGN. MBL and DC-SIGN possess C-terminal carbohydrate-recognition domains (CRD) whereas ficolins possess fibrinogen-like domains (FBG). MBL and ficolin possess collagen-like domains and consist of trimeric subunits. DC-SIGN possesses a neck region and transmembrane (TM) domain.
Molecules 20 02229 g003

3.2. MBL Interaction with Viruses

MBL interacts with several, but not all, viruses in a Ca2+-dependent manner. It interacts with several strains of human immunodeficiency virus-1 (HIV-1) via its spatially conserved, high mannose-type, N-linked glycosylated gp120 envelope glycoprotein [38]. While HIV-1 uses several mechanisms to evade adaptive immune responses—such as “glycan shielding” in which gp120 glycosylation site mutations prevent neutralising antibody binding but maintain cell receptor binding [39]—MBL is able to bind and neutralise diverse strains of HIV-1. However, MBL does not neutralise HIV-1 through complement activation, even at concentrations far exceeding physiological serum levels [40,41]. Therefore, MBL directly neutralises infection in complement-independent manners, such as opsonisation to enhance phagocytosis by DCs and macrophages, as observed for bacterial infections (Figure 4) [42]. The phagocyte cell surface receptor for MBL has not yet been identified, but a likely candidate is calreticulin—a protein-folding chaperone typically situated in the ER, with roles in antigen presentation—which may bind the MBL at the MASP-binding site [43].
Figure 4. The pro-viral and antiviral activities of serum MBL and ficolins. Both have antiviral activities through complement activation, opsonophagocytosis and direct spatial blocking of virus-receptor interactions and entry. Ficolins also enhance production of inflammatory cytokines and nitric oxide. MBL has specific pro-viral activities in Ebola virus and HIV-1 infections, through enhanced virus entry and neuronal apoptosis respectively.
Figure 4. The pro-viral and antiviral activities of serum MBL and ficolins. Both have antiviral activities through complement activation, opsonophagocytosis and direct spatial blocking of virus-receptor interactions and entry. Ficolins also enhance production of inflammatory cytokines and nitric oxide. MBL has specific pro-viral activities in Ebola virus and HIV-1 infections, through enhanced virus entry and neuronal apoptosis respectively.
Molecules 20 02229 g004
MBL-gp120 interaction also prevents HIV-1 interaction with cell entry inhibitors and inhibits trans-infection by direct spatial blocking [40,44]. The same viral glycoprotein-mediated complement-dependent/independent mechanisms are employed by MBL to neutralise influenza A virus (IAV) [45,46], hepatitis C virus (HCV) [47], severe acute respiratory syndrome coronavirus (SARS-CoV) [48,49], Dengue virus (DV) and West Nile virus (WNV) [50,51] infection in vitro. Furthermore, MBL can indirectly activate coagulation upon pathogen binding, perhaps via MASP-1 recruitment, as observed in HCV [52] and IAV [45] infection.

Viral Exploitation of MBL

Recombinant MBL has therapeutic potential against Ebola virus (EBOV) infection as it neutralises the virus in vitro and in vivo via complement activation [53], phagocytosis, and direct inhibition of glycoprotein interaction with the DC-SIGN/L-SIGN receptor [54]. However, MBL may have pro-viral effects, enhancing EBOV infection by mediating macropinocytosis in low complement conditions, perhaps via the C1QBP cell receptor (Figure 4) [55]. This may partially explain the high prevalence of allelic variants conferring low MBL levels, as in the context of EBOV lower MBL levels may prevent excessive infection [55]. MBL can also enhance HIV-1 infection of the brain. AIDS-associated dementia complex is characterised by neuronal cell death and cognitive deficits, however HIV-1 infects relatively few brain cells [56]. Instead, HIV-1 sheds gp120 glycoprotein, which is internalised via the CXCR4 receptor on neuronal cells, then bound and shuttled by intracellular MBL in vesicle complexes to the ER and Golgi apparatus, perhaps facilitating gp120-mediated apoptosis (Figure 4) [56,57].

3.3. MBL Variants

In addition to the wild-type MBL allele A, 3 MBL variants with single nucleotide polymorphisms (SNPs) in exon 1 are common. These—termed MBL-B, -C and -D—are products of amino acid substitutions in codons 54 [58], 57 [59] and 52 [60] which disrupt Gly-Xaa-Yaa repeats. These cause less stable CLDs, hindering the capacity for trimerisation, MASP-binding and complement activation [20]. The mutations are N-terminal to the CLD kink whereas the MASP-binding sites are C-terminal, indicating that the decreased affinity for MASPs is an indirect result of the destabilisation of the CLD [20].
Certain African and South American populations have low serum MBL levels as a result of high allele MBL-C and -B frequencies, respectively [61]. Frequency of the MBL-D allele is particularly high in African and Caucasian populations [60]. Nevertheless, approximately 10%–30% of the global population are MBL deficient [62]. Heterozygotes for the variant alleles exhibit characteristics of both wild-type and variant homozygotes, including different patterns of oligomerisation and MBL serum levels [63].
However, despite its polymorphisms, the MBL2 gene and its primate orthologues are highly homologous and conserved, emphasising the evolutionary and immune importance of the protein [64]. Several theories for the high frequencies of various MBL variants with significantly different levels and activities have been proposed, including founder effect, or selective advantages such as heterozygous advantage, decreased complement-mediated tissue damage and a role of MBL in the enhancement of infection [65].
MBL2 promoter SNPs −550G>C (H/L), −221C>G (X/Y) and +4C>T (P/Q) influence MBL serum concentration [61,66,67]. As a result of linkage disequilibrium, 7 common haplotypes have been described, each with differing serum levels of MBL ranging from <0.01 μg/mL to >5 μg/mL, with an average of 1.7 μg/mL [61,68]. Although alternative immune mechanisms often compensate for MBL defects [69], both excessively high and low serum levels are associated with increased susceptibility to both infectious and autoimmune disease [61].

The Effect of MBL Variant Alleles on Viral Infection

MBL deficiency-associated alleles correlate with increased susceptibility to infectious diseases, however these relationships are complex (Table 1). Homozygotes and heterozygotes of the variant alleles increase risk of HIV-1 infection [70,71] and disease progression [72]. HIV-1 patients carrying the MBL-B allele had higher viral loads in their sera, likely as a result of decreased MBL-mediated viral elimination [73]. The -221 SNP has been correlated with increased risk of perinatal HIV-1 infection [74], whereas higher MBL levels conferred protection. Low serum level haplotypes of the −550 and −221 SNPs were associated with low CD4+ T-cell counts, higher viral loads [75] and accelerated disease progression [72]. In contrast, some studies found no association between variant alleles and susceptibility to HIV-1 infection [76] and disease progression [77,78]. Although MBL levels do not change throughout HIV-1 disease progression, MBL serum levels are elevated in HIV-1-infected patients and are correlated with good response to highly active antiretroviral therapy (HAART) [41,79]. Some studies suggest a protective effect against HIV-1 disease progression conferred by the variant alleles [80].
Table 1. Single nucleotide polymorphisms of the MBL2, FCN and CD209 genes with associations with virus infections. del = deletion; ins = insertion. +1 represents the A of the ATG translation start site for all genes except MBL2, where +1 represents the transcription start site, to comply with literature.
Table 1. Single nucleotide polymorphisms of the MBL2, FCN and CD209 genes with associations with virus infections. del = deletion; ins = insertion. +1 represents the A of the ATG translation start site for all genes except MBL2, where +1 represents the transcription start site, to comply with literature.
GenedbSNP (Alternative Name)Nucleotide PositionMajor AlleleMinor AlleleRegionAmino Acid MutationRelevance to Specific Virus Infections
MBL2rs11003125 (H/L)−550GCPromoter-HIV [74,75]
rs7096206 (X/Y)−221CGPromoter-HBV [81], HCV [82,83], HIV [72,75], HTLV [84], SARS-CoV [48]
rs7095891 (P/Q)+4CT5’ UTR-
rs5030737 (MBL-D)+223CTExon 1Arg52CysCMV [85,86], DV[50], HBV [87,88], HCV [82,83,89,90], HIV [70,72,80],
rs1800450 (MBL-B)+230GAExon 1Gly54AspCMV [85,86], DV [50], HBV [81,88,91], HCV [82,83,89,90], HIV [70,72,73,80], HTLV [92], SARS-CoV [48]
rs1800451 (MBL-C)+239GAExon 1Gly57GluCMV [85,86], DV [50], HBV [88], HCV [82,83,89,90], HIV [70,72,80]
FCN1rs2989727−1981GAPromoter-
rs10120023−542GAPromoter-
rs28909976−271-InsTPromoter-
rs10117466−144CAPromoter-Increased serum concentration [93]
rs10441778+1435GAExon 2Gly43AspLikely affects structure and oligomerisation [94]
ss76901539+3458GAExon 4Arg93GlnLikely affects structure and oligomerisation [94]
rs148649884+6658GAExon 8Ala218ThrReduced serum concentration, reduced ligand binding [93]
rs150625869+7895TCExon 9Ser268ProAbolished serum concentration [93]
rs1071583+7918GAExon 9-
ss76901546+7929GAExon 9Trp279STOPLikely affects structure and oligomerisation [94]
rs138055828+7959AGExon 9Ala289SerReduced serum concentration, reduced ligand binding [93]
ss76901547+8000GAExon 9Gly303SerLikely affects function [94]
FCN2rs3124952−986GAPromoter-Reduced serum concentration [94]; HBV [95]
rs3124953−602GAPromoter-Increased serum concentration [94]; HBV [95]
rs17514136−4AGPromoter-Increased serum concentration [94]; HBV [95]
ss76901565+4423CTExon 5Arg103CysLikely affects chemical and structural properties [94]
ss76901566+4526CTExon 5Thr137MetLikely affects chemical and structural properties [94]
ss76901570+4957GAExon 6 Arg147GlnLikely affects ligand binding [94]
ss76901571+4987GAExon 6 Arg157GlnLikely affects ligand binding [94]
rs17549193+6359CTExon 8Thr236MetReduced binding to GlcNAc [94] and PTX3 [96];
rs7851696+6424GTExon 8Ala258SerIncreased binding to GlcNAc [94]; CMV [86], HBV [95]
rs28357091+6443_44CTAExon 8Ala264fsTruncated protein [94]
FCN3rs28357092+1637CdelCExon 5Leu117fsTruncated protein [94]; Severe, recurrent respiratory and gastrointestinal infections [97,98,99]
ss76901551+1663AGExon 5Thr125AlaLikely affects function [94]
ss76901555+5543TCExon 8Val287AlaLikely affects function [94]
CD209rs4804803−336AGPromoter-DV [100,101,102], HCV [103], HIV [104,105,106], SARS-CoV [107]
rs11465366−201CAPromoter-HIV [104,106]
rs2287886−139TCPromoter-HIV [106,108]
rs41374747+660GAExon 4Arg198GlnHIV [104]
rs11465380+791CGExon 4Leu242ValHIV [104]
The MBL variants may also influence Dengue virus infection and disease progression [50]. MBL deficiency is also linked to cytomegalovirus (CMV) reactivation after lung or liver transplantation [85,86] and susceptibility to SARS-CoV [48], yet displayed no correlation with IAV H1N1 infection [109]. Correlations between MBL-B homozygotes and the −221 SNP with susceptibility to human T-cell lymphotropic virus (HTLV) infection have been observed [84,92].
Some studies suggested that MBL variants do not influence susceptibility to HCV infection [110,111] or disease progression [112], however others show an association with HCV infection, disease progression and treatment response [82,83,89,112]. One study associated MBL-B and -D with protection against HCV infection, while associating MBL-C with increased susceptibility [90]. MBL-B is associated with chronic hepatitis B disease progression [91], whereas MBL-D displayed a Caucasian-specific association with hepatitis B virus (HBV) persistence [87]. All variants and the −221 SNP may influence HBV-associated hepatitis, liver cirrhosis and hepatocellular carcinoma [81,88], and MBL serum levels have a role in perinatal HBV infection [113]. However, one study found no correlation between variants and chronic HBV infection [114].
When studying the association of polymorphisms with disease severity, the geographical population, ethnicity, disease severity, asymptomatic patients, age, route of transmission, study techniques and sample size must be taken into account. These confounding factors can, in part, explain discrepancies between studies.

4. Ficolins

4.1. Genetics, Structure, Expression and Binding Specificities of Ficolins

Three human ficolins have been described: L-ficolin [115], M-ficolin [116,117] and H-ficolin [118]. Several orthologues have been identified in genetically diverse lineages of animals, an indication of the ancient ancestral origins of these lectins. These include the invertebrate ascidians [119], and vertebrates such chickens [120], non-human primates [121] and pigs [116,122]. The human ficolins differ in several ways, for example in their localisation in the human body and their capacity to trigger an immune response.
The M-, L- and H-ficolin proteins are encoded by the FCN1, FCN2 and FCN3 genes, respectively, encoding polypeptides of 326, 313 and 299 amino acids, including the signal peptide (Figure 2b) (reviewed in [123]). The FCN1 and FCN2 genes are both situated on chromosome 9q34 whereas FCN3 is found on chromosome 1p36.11. The FCN2 and FCN3 genes consist of eight exons, whereas FCN1 comprises nine exons.
Each ficolin monomer comprises an N-terminal region with two functionally important cysteine residues, a CLD containing Gly-Xaa-Yaa repeats, a linker region and, characteristically, a C-terminal globular fibrinogen-like domain [124,125]. Like CRDs, the fibrinogen-like (FBG) domain recognises specific pathogen-associated carbohydrates, and the CLD is responsible for signalling to induce an immune response via MASP proteins [124,125]. Active, oligomeric L-ficolin and M-ficolin are dodecamers comprised of four homotrimeric subunits to form what has been labelled as a “bouquet” structure (Figure 3b), whereas H-ficolin is octadecameric (Figure 3c) [124,125,126]. Like MBL, ficolin homotrimers are stabilised by interactions between hydrophobic residues in the CLDs [31,127], and oligomerise by inter-monomer and -trimer disulphide bridges between the N-terminal cysteine residues [124,128].
Hepatocytes are the main site of expression and secretion of both L- and H- ficolin [125,129], although H-ficolin is also highly expressed in type II alveolar and bronchial epithelial cells [129]. Despite minor lung and blood expression, most M-ficolin is associated with the surface of peripheral blood leukocytes [117,130]. H-ficolin is the most abundant serum ficolin (median concentration of ~26 µg/mL; range 6–83 µg/mL) [131] followed by L-ficolin (median of 3.7–5.4 μg/mL; range ~1–13 µg/mL) [132,133] and M-ficolin (median of 1.07 µg/mL; range 0.28–4.05 µg/mL) [134].
All ficolins bind GlcNAc and N-acetylgalactosamine (GalNAc) [118,130,135]. H-ficolin also binds GalNAc and D-fucose, but not mannose and lactose [118,135]. M-ficolin also binds sialic acid [130]. These differing ligand specificities are conferred by sequence differences in the binding site—S1—near the Ca2+-binding site of the FBG [135]. In addition to S1, L-ficolin has 3 inner binding sites, S2–S4, which exhibit great structural plasticity, thus allowing the sites to accommodate a wide variety of ligands in both Ca2+-dependent and -independent ways such as phosphocholine moieties of bacterial teichoic acids, in addition to acetylated compounds [135,136]. For example, S2 is the major binding site for galactose and N-acetylcysteine, whereas S3 and S4 cooperate to bind (1,3)-β-d-glucan, among others [135].

4.2. The Roles of Ficolins in the Immune Response

Like MBL, ficolins indirectly activate the lytic complement pathway via MASP activation, induce phagocytosis by opsonisation, and stimulate the production and secretion of inflammatory cytokines and nitric oxide by macrophages [137]. Interaction with phagocytes is believed to be mediated by a functionally significant lysine in the CLD, at residues 57 and 47 for L- and H-ficolin respectively, which binds calreticulin on phagocyte cell surfaces [126]. The same residue is responsible for interaction with MASPs, therefore it is possible that the phagocytic and complement effects of L-ficolin are competitive [20,126]. L-ficolin may also clear apoptotic and necrotic host cells through the binding of apoptosis-associated ligands [138]. Furthermore, L-ficolin directly prevents viral entry into host cells [139].

4.3. Ficolin Interaction with Viruses

The role of ficolins in the clearance of several pathogens has become increasingly evident; however their role in viral clearance requires greater investigation. L-ficolin interacts with viruses via N-linked glycans on viral envelope glycoproteins [140,141,142]. L-ficolin binding of HCV triggers infected-cell lysis via C4 deposition, however L-ficolin interaction is abrogated if the HCV E2 glycoprotein is not glycosylated [141]. Biologically relevant levels of recombinant oligomeric L-ficolin, which displayed similar binding activity and structure to serum L-ficolin, neutralised HCV entry in a dose-dependent manner by preventing E2 interaction with cell surface lipoprotein receptor and scavenger receptor B1, which are important for HCV entry [139,143]. Monomeric L-ficolin can activate complement [141] but not inhibit HCV entry [139].
Human L-ficolin and porcine ficolin-α neutralise replication and infection of IAV in vivo [142] and porcine reproductive and respiratory syndrome virus in vitro [140], respectively. L-ficolin directly inhibits IAV entry and promotes complement-mediated lysis of IAV and infected cells [142]. IAV binds sialylated glycans on serum H-ficolin in the airway before viral entry, enabling H-ficolin mediated inhibition of IAV infectivity by direct blocking, viral aggregation and complement activation [144]. As yet unpublished research by Ren et al. implicates L-ficolin mediated complement activation following interaction with HIV-1 gp120 [145].
Few M-ficolin interactions with pathogens have been observed, despite its ability to activate complement [130]. M-ficolin inhibits IAV infection [144]. The majority of M-ficolin is monocyte and granulocyte membrane-bound, despite its lack of a transmembrane (TM) domain, and associates with sialylated membranes via its FBG domain [146]. A candidate receptor of M-ficolin is G-protein-coupled receptor 43 (GPCR43) which, upon M-ficolin-mediated pathogenic interaction, indirectly activates IL-8 production [147]. Serum M-ficolin binds sialic acid on capsulated Streptococcus agalactiae via its FBG and activates complement, however L- and H-ficolin were found not to interact with this pathogen [148].

4.4. Single Nucleotide Polymorphisms in FCN Genes

Hummelshøj et al. extensively described the SNPs of the highly polymorphic FCN genes (Table 1) [94,149]. In general, polymorphisms in the promoter regions of the ficolin genes are expected to affect gene regulation and protein concentration whereas coding region polymorphisms likely affect protein stability, modification, folding and activity, thus altering protein function. Non-synonymous substitutions alter protein activity, however non-synonymous mutations may influence mRNA processing and protein expression.
While polymorphisms have been identified in the FCN1 and FCN3 genes, several more significant SNPs have been identified in the FCN2 gene. The frequencies of FCN polymorphisms often differ between ethnicities, with some existing solely in a particular geographical population, more so in African populations [94,150]. This likely arose from distinct geographical selective pressures, such as genetically-determined and infectious diseases.
The FCN2 and FCN3 genes have three and two as yet undetected splicing variants, respectively [115,151]. The FCN1 gene contains 45 SNPs, nine of which are exclusive to African populations and eight of which are non-synonymous. Gly43Asp, Arg93Gln and Trp279STOP likely affect M-ficolin structure and oligomerisation, whereas Gly303Ser may affect M-ficolin function. The FCN3 gene showed 15 low frequency SNPs, none of which were found globally. Only Leu117fs, Thr125Ala—corresponding to FCN2 Thr137Met—and Val287Ala are predicted to affect H-ficolin function.
Of the 36 SNPs in the FCN2 gene, five significant SNPs have been identified. Promoter polymorphisms −986A>G, −602G>A and −4A>G affect serum levels of L-ficolin and exon 8 polymorphisms +6359C>T and +6424G>T, conferring Thr236Met and Ala258Ser respectively, in the FBG alter L-ficolin affinity to GlcNAc. Several of these and other polymorphisms were in strong linkage disequilibrium.
Additional FCN2 SNPs were detected and their effects hypothesised, however no associated phenotype has yet been described. The Arg147Gln and Arg157Gln mutations are found in the S2 and S3 binding sites respectively, and are therefore expected to affect ligand binding. Furthermore, the Arg103Cys and Thr137Met mutations are expected to affect the chemical and structural properties of L-ficolin. A rare frame shift mutation encoding Ala264fs has also been described, however a homozygote for this polymorphism has not been found, hence the physiological implications are unknown [149]. The rare frame shift SNP +1637delC of FCN3, correlating to Leu117fs, encodes a truncated H-ficolin protein that cannot be expressed [97]. This leads to full H-ficolin deficiency—in homozygotes, causing high levels of lung infection and disease, and severe necrotising enterocolitis [97,98,99].

4.4.1. The Significance of FCN Gene Single Nucleotide Polymorphisms in Viral Infections

There have been several clinical studies monitoring the part that ficolins play in disease outcome, typically focussing on one or both of two factors: ficolin gene polymorphisms (Table 1) and ficolin serum concentrations.
One FCN2 haplotype is associated with protection against HBV infection [95]. L-ficolin levels were higher in patients with acute rather than chronic HBV infection, suggesting that the protein is directly involved in immediate clearance of the virus, and influences subsequent liver disease [95]. The L-ficolin Ala258Ser mutation appears to confer a protective effect against CMV re-infection in liver transplantation when compared to wild-type L-ficolin [86].
The Thr236Met mutation reduced affinity towards pentraxin 3 (PTX3), a serum protein which enhances L-ficolin-mediated complement response to Aspergillus fumigatus, suggesting that FCN2 polymorphisms also alter affinity towards cooperative proteins and thus affect the immune response [96]. Interestingly, the +6424T SNP is associated with low serum levels of L-ficolin, yet also increased L-ficolin binding [133]. It has been hypothesised that this unexpected correlation is due to higher activity and “exhaustion” of the L-ficolin protein as a result of its increased binding affinity.
Although there are few investigations of the role of L-ficolin in viral infections, FCN2 polymorphisms were found to be significant in susceptibility to and disease severity of bacterial diseases, including cutaneous leishmaniasis [152], Mycobacterium leprae [153], Pseudomonas aeruginosa-associated bronchiectasis [154], and Streptococcus pygones-associated rheumatic fever and chronic rheumatic heart disease [155]. However, the SNPs were not associated with invasive pneumococcal disease [156] and other respiratory tract infections [157]. These studies have been relatively small-scale and geographically limited, and often did not measure the serum concentrations of L-ficolin to confirm the relationship of the polymorphisms and haplotypes with L-ficolin levels, therefore more rigorous larger scale studies would yield more reliable results.
M-ficolin SNPs have not yet been reported to have roles in viral infections, however the −144C SNP is associated with protection against M. leprae-associated leprosy, whereas −1981A, −271delT and −542G correlate with susceptibility, perhaps by altering transcription factor affinity [158]. The −144C SNP has been associated with increased M-ficolin serum levels [93]. The Ala218Thr and Asn289Ser non-synonymous mutations reduced serum levels and ligand binding activity, whereas Ser268Pro abolished serum levels [93]. The −1981A and +7918 SNPs were correlated with rheumatoid arthritis [159]. Interestingly, expression the FCN1 gene, among others, is up-regulated in chronic HCV patients who possess the CC genotype at the IL28B rs12979860 promoter SNP, which is associated with favourable response to pegylated interferon-α and ribavirin treatment [160,161]. This suggests a possible role of M-ficolin in the clearance of HCV, however further studies are needed. No studies have correlated H-ficolin SNPs with infectious diseases.

4.4.2. The Significance of Ficolin Serum Concentrations in Viral Infections

Serum ficolin levels are dependent on the expressed alleles; homozygotes for particular SNPs exhibit the highest or lowest levels whereas heterozygotes display intermediate levels of ficolin [149]. Specific FCN2 SNPs that are associated with low levels of L-ficolin tend to cause higher susceptibility to infection [152]. L-ficolin levels are significantly increased in the serum of HCV-infected individuals, and concentrations correlate with the severity of fibrosis [141]. Chronic HCV-associated liver damage also did not reduce the levels of L-ficolin expressed [139]. In chronic HCV-infected patients with abnormal alanine aminotransferase (ALT) levels, serum concentrations of L-ficolin correlated with ALT levels [162]. ALT is a known marker of fibrosis and inflammation [162]. After successful therapy, ALT and HCV RNA levels of these patients all decreased to normal values, followed by a decrease in L-ficolinlevels, suggesting a correlation of ALT and RNA levels with disease outcome, as a result of L-ficolin activity [149,162].
Higher L-ficolin serum concentrations also appear to confer protective effects against microorganism-induced inflammation in allergic respiratory disease [163]. L-ficolin levels were higher in acute severe cases of Plasmodium falciparum-based malaria, rather than mild cases [164]. Low H-ficolin serum levels correlate with fever and neutropenia in paediatric cancer patients treated with chemotherapy [131].

4.5. Cooperative Relationships between Lectins and Other Immune Proteins

A consequence of complement activation is the subsequent activation of the humoral immune system against a pathogen, thus enhancing the adaptive immune response and memory [6]. Lectins can also directly interact with immune components and enhance the antimicrobial response. MBL can bind the serum PRRs PTX3 and serum amyloid P component (SAP) via its CLD to promote complement activation and opsonophagocytosis of Candida albicans (Figure 5a) [165]. PTX3 required C1q to enhance activation of the classical complement pathway [165].
Natural antibodies are produced without prior exposure to infection or immunisation, and are important in the protection of individuals exposed to pathogens for first time, such as neonates [166]. They are able to initiate complement alone, however under mild acidosis and reduced calcium levels—conditions found at infection-inflammation sites—an additional binding site is exposed on the ficolin FBG to allow complex formation with natural immunoglobulin G (nIgG) [166,167]. This allows indirect nIgG-based phagocytosis via L-ficolin opsonisation, leading to a stronger immune response (Figure 5b) [166]. PTX3 interacts with L-ficolin to enhance its binding of Aspergillus fumigatus and its induction of C4 deposition [96]. Similarly, M-ficolin binds sialic acid on PTX3 in a Ca2+-dependent manner via its FBG domain [168,169]. M-ficolin:PTX3 complexes enhance phagocytosis of apoptotic and necrotic cells [169]. H-ficolin may also cooperate with PTX3 [144], however specific interaction between the two has not been reported [96,168].
Similar to nIgG:ficolin complexes, infection-inflammation conditions significantly increase interaction between the acute phase protein C-reactive protein (CRP) and L-ficolin, leading to a stronger classical- and lectin-mediated complement response against Pseudomonas aeruginosa [170]. A pH- and calcium-sensitive binding site on the ficolin FBG domain enables binding to CRP [171]. Later phase infection-inflammation conditions also enhance interaction between membrane GPCR43-associated M-ficolin and CRP, thus blocking M-ficolin binding of PAMPs and curtailing GPCR43-mediated IL-8 production, and allowing the restoration of homeostasis upon infection and injury [147,171].
Figure 5. The cooperative antimicrobial relationships of serum (a) MBL and (b) ficolins with other immune proteins. Complement activation and opsonophagocytosis can be enhanced by SAP and PTX3-C1q interaction with MBL, and PTX3, CRP-C1qrs and natural IgG interaction with ficolins.
Figure 5. The cooperative antimicrobial relationships of serum (a) MBL and (b) ficolins with other immune proteins. Complement activation and opsonophagocytosis can be enhanced by SAP and PTX3-C1q interaction with MBL, and PTX3, CRP-C1qrs and natural IgG interaction with ficolins.
Molecules 20 02229 g005

5. DC-SIGN

5.1. Genetics, Structure, Expression and Binding Specificities of DC-SIGN

Dendritic cells (DCs) express PRRs, such as DC-SIGN [172], which play a role in antigen capture and internalisation. This contributes to maturation and migration of DCs to secondary lymphoid organs where they present antigen to resting T-cells [173]. The DC-T-cell interaction is mediated by the transient interaction of DC-SIGN with ICAM-3 on the T-cell surface [174].
Human DC-SIGN—encoded by the CD209 gene—is a C-type lectin, the monomeric structure of which comprises a C-terminal, Ca2+-dependent CRD followed by a flexible neck region which forms an α-helical coiled coil upon trimerisation and which consists of 7.5 23-amino acid repeats involved in oligomerization [175,176]. A TM region follows, then a cytoplasmic N-terminal region responsible for internalisation via di-leucine and tyrosine motifs (Figure 2c) [175,177]. Adjacent to CD209 on chromosome 19p13 is the CD299 gene, encoding the membrane-bound C-type lectin L-SIGN (DC-SIGNR)—a protein that is highly homologous to DC-SIGN [178]. This protein performs similar roles to DC-SIGN, but differs in tissue distribution and expression, being expressed mainly in liver and lymph nodes.
Active, tetrameric DC-SIGN (Figure 3d) is organised on DC surfaces in nanoclusters, mediated by the neck region and permitting interactions with pathogens of a variety of sizes, as large as 300 nm across as seen with the measles virus (MV) [175,176]. The level of DC-SIGN expression is important to optimal virus binding, which partially explains the relatively more efficient HIV-1 binding exhibited by DCs rather than other DC-SIGN+ cells [179,180]. Furthermore, the virus-binding capacity of DC-SIGN is also governed by the cell type in which the virus replicated, as each cell type causes subtle changes in the glycosylation profile of viral glycoproteins [181].
DC-SIGN preferentially binds fucosylated and high mannose-type oligosaccharides using different binding sites, with differing avidity depending on the configuration of the mannose and fucose residues in the glycan ligand, as well as the presentation of the target molecule on the pathogen surface [175].

5.2. Exploitation of DC-SIGN by Viruses

DC-SIGN is exploited by several pathogens for host cell binding and entry. The current, generally accepted model of DC-SIGN-mediated HIV-1 infection utilises DC-SIGN+ cells, such as DCs, to transport virions from sites of HIV-1 exposure—at the mucosal membranes or bloodstream—to CD4+ T-cell targets in the lymphoid tissues (reviewed in [182]). In addition to the primary HIV-1 receptor CD4 and CCR5 or CXCR4 co-receptors [183], DC-SIGN also recognises HIV-1 gp120 [172], resulting in the activation of downstream processes. Most HIV-1 virions are shuttled to the proteasome, aided by the interaction of the cytoskeletal phosphoprotein LSP1 with HIV-1-bound DC-SIGN, where it is degraded for MHC class II presentation to T-cells [184,185]. Alternatively, HIV-1 exploits the DC-SIGN signalosome—comprising DC-SIGN, LSP1, KSR1 and CNK—to activate Raf-1 and modulate cytokine response, and to enhance NF-κB-mediated transcription of the HIV-1 genome [186]. Independently, HIV-1-bound DC-SIGN modulates TLR-induced cytokine and HIV-1 genome expression (Figure 6) [186].
Figure 6. The role of DC-SIGN in HIV-1 infection. DC-SIGN neutralises HIV-1 infection through increased DC-SIGN signalosome-mediated cytokine production and degradation of virions. DC-SIGN then aids dendritic cell-T-cell interaction through transient ICAM-3 binding, thus allowing antigen presentation to T-cells to enhance the immune response. However, HIV-1 exploits DC-SIGN to increase dendritic cell apoptosis via ASK-1, enhance viral replication via the DC-SIGN signalosome and to evade the immune response in specialised non-lysosomal endosomes. DC-SIGN also enhances HIV-1 trans-infection of T-cells.
Figure 6. The role of DC-SIGN in HIV-1 infection. DC-SIGN neutralises HIV-1 infection through increased DC-SIGN signalosome-mediated cytokine production and degradation of virions. DC-SIGN then aids dendritic cell-T-cell interaction through transient ICAM-3 binding, thus allowing antigen presentation to T-cells to enhance the immune response. However, HIV-1 exploits DC-SIGN to increase dendritic cell apoptosis via ASK-1, enhance viral replication via the DC-SIGN signalosome and to evade the immune response in specialised non-lysosomal endosomes. DC-SIGN also enhances HIV-1 trans-infection of T-cells.
Molecules 20 02229 g006
DC-SIGN binding to HIV-1 can also enhance CD4-gp120 glycoprotein interactions and CCR5-mediated entry [187,188]. As a result, DC-SIGN-bound, intact HIV-1 virions are non-fusogenically internalised into a specialised, non-lysosomal, low-pH endosome, where virions are able to maintain stability and infectivity for several days without the need for replication [177,189]. Whether an HIV-1 virion is degraded or maintained in a DC is dependent on several factors, including the N-linked glycan composition of gp120 [190].
DC-internalised, stable HIV-1 is can be transferred to T-cells [177]. Upon DC-T-cell contact, protected HIV-1 virions, adhesion molecules and receptors rapidly accumulate at the DC-T-cell interface, enabling HIV-1 transfer across the infectious synapse and gp120 binding to T-cell receptors [191]. HIV-1 Nef may prevent DC-SIGN-bound HIV-1 internalisation to enhance trans-infection [192,193]. Trans-infection is more likely to occur in more mature DCs, which are more efficient at protecting the virus by endocytosis whereas immature DCs are more efficient at degradation [184,191]. In contrast, HIV-1 binding to DC-SIGN may slow DC maturation by—for example—reducing expression of CD86 and MHC class II, in order to prevent immune response and prime the cell for trans-infection [194]. Other lectins are implicated in HIV-1 trans-infection, such as the mature DC-expressed sialic acid-binding Ig-like lectin 1 (Siglec-1), which binds sialyllactose-containing gangliosides on the HIV-1 surface [195]. Interestingly, a feature of acquired immunodeficiency syndrome (AIDS) is the gradual depletion of DC levels, increasing the risk of opportunistic infection [196]. This is a result of gp120-DC-SIGN interaction in certain conditions to prime the DC for apoptosis, through excessive activation of the pro-apoptotic kinase ASK-1 [196].
Several other viruses utilise DC-SIGN for different purposes. For example, many viruses use DC-SIGN as an attachment factor and for trans-infection, such as MV [197], EBOV [198], SARS-CoV [199], IAV [200], HCV [201] and CMV [202]. Others use DC-SIGN for fusion and internalisation, such as human herpesvirus-8 [203] and HCV [204]. L-SIGN is similarly exploited by many viruses as a glycoprotein-mediated attachment and internalisation receptor, as is observed with HIV-1 [205], HCV [201], SARS-CoV [206], Marburg virus [206] and EBOV [198]. However, DC-SIGN and L-SIGN differ in viral interaction and polymorphisms, as the neck region of L-SIGN is highly variable and polymorphic whereas DC-SIGN is generally conserved [207]. Furthermore, L-SIGN binds and promotes infection of certain viruses more efficiently than DC-SIGN, for example WNV [208].

5.3. DC-SIGN Variants

Several DC-SIGN isoforms exist as a result of alternative splicing, resulting in both membrane-bound and soluble forms with alternative cytoplasmic and CRD regions, missing TM domains and variable neck domain repeat regions [209]. Relatively little is known of the physiological and immune roles of soluble DC-SIGN (sDC-SIGN). Tetrameric sDC-SIGN has been detected in bodily fluids and the cytoplasm of some DC-SIGN+ cells, however it is unknown how the protein is secreted [209,210]. Interestingly, CMV interacts with sDC-SIGN via its gB glycoprotein, in the same manner as its interaction with membrane-bound DC-SIGN, possibly to enhance viral uptake by DCs [210].
Considerable purifying selection pressure has been exerted on the CD209 region encoding the DC-SIGN neck region, preventing the accumulation of polymorphisms in the global population and maintaining the typical neck region size of 7.5 repeats [207]. In fact, the CD209 gene only displays 2% variant heterozygosity, and the prototypic 7.5 neck region repeat allele is present in 99% of the global population [207]. This length and the tetrameric structure constrain the CRDs and partially influence specificity for certain viruses and glycan structures with certain geometric configurations [175]. Despite this conservation, eight low frequency CD209 alleles encoding DC-SIGN proteins with two to 10 neck region repeats have been documented [207]. Different DC-SIGN isoforms can be found on the same cell surface and can affect DC-SIGN multimerisation, and therefore could, in theory, influence susceptibility to HIV-1 infection [211].

5.4. The Significance of DC-SIGN Variants in Viral Infection

The Leu232Val and Arg198Gln mutations of DC-SIGN enhanced HIV-1 capture and trans-infection [104]. While some studies have identified no significant DC-SIGN exon 4 neck region SNPs in the context of HIV-1 [212] it has been observed that DC-SIGN variants with less than five neck region repeats are rare but are more frequent in Chinese populations, and are associated with protection against HIV-1 transmission [213]. Heterozygotes for variants with atypical numbers of neck region repeats above five were also associated with protection against HIV-1 infection [214]. CMV has been observed to interact with sDC-SIGN via its gB glycoprotein, in the same manner as its interaction with membrane-bound DC-SIGN, possibly to enhance viral uptake by DCs [210].
CD209 promoter SNPs have been implicated in resistance and susceptibility to several infectious diseases, however such correlations are often limited to particular geographical populations. The −336G SNP and AA genotype confer protection against DV-based dengue fever [100,101], perhaps due to reduced transcription factor binding and activity [107], decreased susceptibility of DCs to infection [101] and the suppression of symptoms [102]. The −336 GG/AG genotypes, on the other hand, increase susceptibility to dengue haemorrhagic fever as a result of higher DC-SIGN expression and immune activity against DV [100]. The −139G allele is associated with protection against HIV-1 infection [108], whereas −336C is implicated in higher susceptibility to parenteral HIV-1 infection [105]. The −139, −201 and −336 SNPs are also implicated in perinatal transmission of HIV-1 [104,106]. The −336 SNP influenced liver disease severity in HCV patients [103] and the −336 AG/GG genotypes pre-empted improved prognosis for SARS-CoV patients [107].

6. Additional Lectins with Pro-Viral and Antiviral Properties

In addition to DC-SIGN, several membrane-bound lectins are exploited by viruses to enhance infection, such as HIV-1 exploitation of the C-type lectin mannose receptor [215,216] and Siglec-1 [195], however there are significantly less soluble lectins involved in enhancing infection. Nevertheless, a few candidates, in addition to MBL, have been identified.
The soluble pulmonary collectins SP-A and SP-D both inhibit IAV haemagglutinin activity by viral aggregation and enhancement of the neutrophil immune response [217]. SP-A also has roles in enhancing phagocytosis via the host SP-A receptor 210, which may be exploited by IAV to cause excessive lung inflammation [218]. Furthermore, SP-D can bind and neutralise several strains of HIV-1 through virion agglutination and spatial blocking of gp120-CD4 interaction [219]. However, SP-D enhances HIV-1 trans-infection from DCs to T-cells [220]. Similarly, SP-A prevents HIV-1 cis-infection of CD4+ T-cells, yet enhances cis-infection of DCs and trans-infection of T-cells [221]. SP-A also enhances entry and fusion of respiratory syncytial virus (RSV) [222].
Galectins are soluble lectins of varying structure and oligomerisation that exhibit specificity for glycans containing β-galactoside [3]. Galectin-1 binds both HIV-1 gp120 glycans—through its CRD—and human cell CD4 receptor, thus allowing the protein to directly stabilise and cross-link HIV-1-CD4 interaction and enhance viral attachment and entry [223]. Similarly, galectin-9 cross-links HIV-1 interaction with cell surface-associated protein disulphide isomerase to enhance fusion and entry [224]. Galectin-3, on the other hand, intracellularly bridges HIV-1 Gag p6 interaction with ALG-2-interacting protein X (Alix) to promote HIV-1 budding [225].
Membrane-associated lectins are generally described to be exploited by viruses for entry. However, some can have significant roles in viral clearance. The C-type lectin langerin is expressed on the surface of Langerhans cells, an epithelial DC subset, and binds the mannose-containing glycans of gp120 [226]. Unlike DC-SIGN, langerin mediates internalisation into Birbeck granules for viral degradation and antigen presentation while efficiently preventing HIV-1 transmission [226]. Langerin is also responsible for the capture, but not internalisation, of MV [227].

7. Lectin Therapy for Viral Infections

7.1. Soluble Lectin Therapy

Virus-associated glycans are emerging as potential targets for antiviral therapy [7]. Preliminary clinical trials of regular MBL replacement therapy for MBL-deficient patients using plasma-purified MBL have been attempted, and resulted in normal, long-term complement activation and opsonisation activities with no obvious adverse or autoimmune effects [228]. The production and purification of safe, active and functional therapeutic MBL is feasible, but requires optimisation [229]. Therapy using recombinant MBL avoids ethical issues and allows cost-effective large-scale production [230], and phase I trials proved recombinant MBL to be tolerable, safe and effective in the restoration of MBL activity in MBL-deficient patients, with mild to no adverse effects [231].
Due to its complex structure, recombinant active MBL is complicated and expensive to produce [232], whereas recombinant chimaeric lectins (RCLs) can exhibit superior protective potency, cost-effectiveness and safety [217,230,231,233]. The antiviral efficacy of RCLs consisting of the structurally similar lectins MBL and L-ficolin has been studied [232,234]. Like L-ficolin, active MBL/L-ficolin RCLs formed dodecamers, and RCL2 (L-FCN/MBL76) exhibited stronger MBL CRD-mediated ligand binding, likely due to enhanced CRD flexibility [232]. RCL2 possessed a MASP- and calreticulin-binding site consisting of both MBL and L-ficolin fragments, and displayed strong complement activation and opsonophagocytic properties [232]. Physiological levels of RCL2 neutralised EBOV pseudotype infection, and moderately neutralised Hendra and Nipah virus infection [232]. RCL2 and RCL3 (L-FCN/MBL64)—which possesses the MBL MASP-binding site and kink—exhibited anti-IAV activity in vitro by enhancing complement activation, inducing viral aggregation, inhibiting IAV envelope glycoprotein activities and, as a therapeutic advantage, exhibiting a reduced association with MASP-1 to diminish coagulation system activities [235]. RCL3 demonstrated more efficient anti-IAV effects in vivo by maintaining a cytokine and inflammatory balance that is more advantageous for the host [234].
Similar therapeutic approaches for other lectins have been researched. For example, specific mutagenic engineering of human SP-D enhanced IAV binding and clearance and murine survival in vivo [236]. Porcine SP-D neutralises a wider range of IAV infections more potently than recombinant human SP-D in vitro and ex vivo, however it may be immunogenic in humans, therefore further development is necessary [237].
Although ficolins have not been used clinically as antiviral therapies, given the promise of MBL therapy there are is the potential for therapeutic administration of ficolins for virus infections. Due to the essential and conserved nature of N-linked glycosylation sites in the life cycles of many enveloped viruses, acetylated sugars are potential targets for direct entry inhibition and viral clearance by L-ficolin administration. Indeed, there are no documented HCV strains resistant to the antiviral effects of the ficolins [238]. Ficolins could be used in combination with antibody therapy. However, the degree to which antibodies interact with these complement components varies. For example, MBL prevents, rather than enhances, the HIV-1 neutralising activity of the 2G12 antibody [239]. Moreover, the importance of complement in the protection against infectious disease varies with each pathogen and each strain. For example, HCV genotypes with more heavily glycosylated E1E2 glycoproteins appear to be more susceptible to lectin-mediated neutralisation [47].
Interference of the immune system, in particular the complement system, is extremely complex and may lead to potentially harmful and excessive immune activity, therefore strict caution and testing should be imposed [240]. Passive immunotherapy does not actively trigger an immune response to fight infection, and avoids excessive stimulation of the complement system [139]. Therefore L-ficolin could solely act as an entry inhibitor, after site-directed mutagenesis at the Lys-57 residue of the FBG domain, which is responsible for MASP interaction and interaction with phagocytic receptors [126].

7.2. Therapy Using Xenogeneic Lectins

An alternative approach is the use of carbohydrate-binding agents (CBA), which interact with viral glycoproteins in order to inhibit DC-SIGN interaction. These tend to be xenogeneic, such as the algal and cyanobacterial lectins griffithsin (GRFT), cyanovirin-N (CV-N), scytovirin (SVN), Oscillatoria agardhii agglutinin (OAA) and the synthetic antibiotic pradimicin S, which act as CBA inhibitors of HIV-1 interaction with DC-SIGN in vitro, in vivo and ex vivo [241,242,243,244,245,246]. These lectins therefore have potential uses in antiviral microbicides and have elicited non-toxic and non-immunogenic HIV-1 neutralisation in mammals in vitro and in vivo [241,244,247]. OAA and the hybrid OPA molecule (a synthetic chimera with the Pseudomonas fluorescens agglutinin) are notable as uniquely they interact with oligosaccharide Man-9, rather than terminal mannoses of Man-8/9 [248].
GRFT also has antiviral activity against HCV [249], SARS-CoV and several coronaviruses in vitro [250]. In vivo, GRFT reduced viral titres and pathology in Japanese encephalitis-infected, HCV-infected and SARS-CoV-infected mice when administered intraperitoneally, subcutaneously and intranasally, respectively [250,251,252]. CV-N inhibited entry by HCV [253], MV and human herpesvirus 6 [254] in vitro and prevented further infection and mortality of IAV-infected mice and ferrets when administered intranasally, however it was ineffective when administered subcutaneously [255]. Nevertheless, several CBAs have small viral target ranges, likely as a result of diverse viral glycoproteins. For example, certain CBAs inhibited HIV-1 and HCV infection, yet were ineffective against Herpes simplex virus, vesicular stomatitis virus, RSV, parainfluenza virus-3 [256], adenovirus type 5, CMV, herpesvirus type 1 [257] or vaccinia virus [254].
Mitogenic activity, which affects cytokine expression, and a short plasma half-life are characteristic of non-human immune proteins used in the human milieu [243,258]. To overcome this, a linker-extended and PEGylated CV-N was produced in order to extend the protein’s half-life, dampen its immunogenicity and cytotoxicity, and yet maintain its antiviral activity [258].
GRFT elicits a far greater anti-HIV-1 effect than CV-N and SVN, and does not cause significant mitogenic, inflammatory, cytotoxic or irritant effects [259,260]. Despite the possibility that GRFT is not sufficiently stable for therapeutic administration [260], subcutaneously-administered GRFT remained active in mice for several days after injection, with minimal adverse effects [259].

8. Conclusions

With their ancient origins and multiple physiological functions, lectins are part of the evolutionary arms race between the human immune system and infecting viruses. Lectins have evolved the ability to bind diverse ligands and limit viral infections through a range of immune activities. Viruses have evolved highly glycosylated proteins that are able bind lectins to enhance their attachment, entry and transmission, yet can evade lectin-mediated neutralisation. Despite viral exploitation, lectins have the potential to prime the immune system and enhance viral clearance. This emerging field has already yielded novel antiviral therapies, but our understanding of the specificity of these lectins and the virus-host dynamics remains incomplete. Improved understanding of these interactions will aid our development of these proteins as antiviral therapeutics.

Acknowledgments

The authors are funded by the Biomedical Research Unit in Gastrointestinal and Liver Diseases at Nottingham University Hospitals NHS Trust and the University of Nottingham.

Author Contributions

CPM and AWT contributed to the preparation of this review article.

Abbreviations

CRD
carbohydrate-recognition domain
PRR
pattern-recognition receptor
PAMP
pathogen-associated molecular pattern
MBL
mannose-binding lectin
DC-SIGN
dendritic cell-specific ICAM-3 grabbing non-integrin
MAC
membrane attack complex
MASP
MBL-associated serine protease
MAP
MBL-associated protein
CLD
collagen-like domain
GlcNAc
N-acetylglucosamine
SARS-CoV
severe acute respiratory syndrome coronavirus
DV
Dengue virus
WNV
West Nile virus
EBOV
Ebola virus
HAART
highly active antiretroviral therapy
CMV
cytomegalovirus
HTLV
human T-cell lymphotropic virus
FBG
fibrinogen-like
GalNAc
N-acetylgalactosamine
GPCR43
G-protein-coupled receptor 43
PTX3
pentraxin 3
SAP
serum amyloid P component
nIgG
natural immunoglobulin G
CRP
C-reactive protein
DC
dendritic cell
MV
measles virus
Siglec-1
sialic acid-binding Ig-like lectin 1
sDC-SIGN
soluble DC-SIGN
RCL
recombinant chimaeric lectin
CBA
carbohydrate-binding agent
GRFT
Griffithsin
CV-N
Cyanovirin
SVN
Scytovirin
OAA
Oscillatoria agardhii agglutinin

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Loris, R. Principles of structures of animal and plant lectins. Biochim. Biophys. Acta 2002, 1572, 198–208. [Google Scholar] [CrossRef] [PubMed]
  2. Sekine, H.; Kenjo, A.; Azumi, K.; Ohi, G.; Takahashi, M.; Kasukawa, R.; Ichikawa, N.; Nakata, M.; Mizuochi, T.; Matsushita, M.; et al. An ancient lectin-dependent complement system in an ascidian: Novel lectin isolated from the plasma of the solitary ascidian, Halocynthia roretzi. J. Immunol. 2001, 167, 4504–4510. [Google Scholar] [CrossRef] [PubMed]
  3. Varki, A.; Etzler, M.E.; Cummings, R.D.; Esko, J.D. Discovery and Classification of Glycan-Binding Proteins. In Essentials of Glycobiology, 2nd ed.; Varki, A., Cummings, R.D., Esko, J.D., Freeze, H.H., Stanley, P., Bertozzi, C.R., Hart, G.W., Etzler, M.E., Eds.; Cold Spring Harbor: Laurel Hollow, NY, USA, 2009. [Google Scholar]
  4. Ghazarian, H.; Idoni, B.; Oppenheimer, S.B. A glycobiology review: Carbohydrates, lectins and implications in cancer therapeutics. Acta Histochem. 2011, 113, 236–247. [Google Scholar] [CrossRef] [PubMed]
  5. Vigerust, D.J.; Shepherd, V.L. Virus glycosylation: Role in virulence and immune interactions. Trends Microbiol. 2007, 15, 211–218. [Google Scholar] [CrossRef] [PubMed]
  6. Sarma, J.V.; Ward, P.A. The complement system. Cell Tissue Res. 2011, 343, 227–235. [Google Scholar] [CrossRef] [PubMed]
  7. Dalziel, M.; Crispin, M.; Scanlan, C.N.; Zitzmann, N.; Dwek, R.A. Emerging principles for the therapeutic exploitation of glycosylation. Science 2014, 343, 1235681. [Google Scholar] [CrossRef] [PubMed]
  8. Dahl, M.R.; Thiel, S.; Matsushita, M.; Fujita, T.; Willis, A.C.; Christensen, T.; Vorup-Jensen, T.; Jensenius, J.C. MASP-3 and its association with distinct complexes of the mannan-binding lectin complement activation pathway. Immunity 2001, 15, 127–135. [Google Scholar] [CrossRef] [PubMed]
  9. Skjoedt, M.O.; Hummelshoj, T.; Palarasah, Y.; Honore, C.; Koch, C.; Skjodt, K.; Garred, P. A novel mannose-binding lectin/ficolin-associated protein is highly expressed in heart and skeletal muscle tissues and inhibits complement activation. J. Biol. Chem. 2010, 285, 8234–8243. [Google Scholar] [CrossRef] [PubMed]
  10. Stover, C.M.; Thiel, S.; Thelen, M.; Lynch, N.J.; Vorup-Jensen, T.; Jensenius, J.C.; Schwaeble, W.J. Two constituents of the initiation complex of the mannan-binding lectin activation pathway of complement are encoded by a single structural gene. J. Immunol. 1999, 162, 3481–3490. [Google Scholar] [PubMed]
  11. Matsushita, M.; Thiel, S.; Jensenius, J.C.; Terai, I.; Fujita, T. Proteolytic activities of two types of mannose-binding lectin-associated serine protease. J. Immunol. 2000, 165, 2637–2642. [Google Scholar] [CrossRef] [PubMed]
  12. Dong, M.; Xu, S.; Oliveira, C.L.; Pedersen, J.S.; Thiel, S.; Besenbacher, F.; Vorup-Jensen, T. Conformational changes in mannan-binding lectin bound to ligand surfaces. J. Immunol. 2007, 178, 3016–3022. [Google Scholar] [CrossRef] [PubMed]
  13. Feinberg, H.; Uitdehaag, J.C.; Davies, J.M.; Wallis, R.; Drickamer, K.; Weis, W.I. Crystal structure of the CUB1-EGF-CUB2 region of mannose-binding protein associated serine protease-2. EMBO J. 2003, 22, 2348–2359. [Google Scholar] [CrossRef] [PubMed]
  14. Gál, P.; Harmat, V.; Kocsis, A.; Bian, T.; Barna, L.; Ambrus, G.; Vegh, B.; Balczer, J.; Sim, R.B.; Naray-Szabo, G.; et al. A true autoactivating enzyme. Structural insight into mannose-binding lectin-associated serine protease-2 activations. J. Biol. Chem. 2005, 280, 33435–33444. [Google Scholar] [CrossRef]
  15. Gingras, A.R.; Girija, U.V.; Keeble, A.H.; Panchal, R.; Mitchell, D.A.; Moody, P.C.; Wallis, R. Structural basis of mannan-binding lectin recognition by its associated serine protease MASP-1: Implications for complement activation. Structure 2011, 19, 1635–1643. [Google Scholar] [CrossRef] [PubMed]
  16. Ambrus, G.; Gal, P.; Kojima, M.; Szilagyi, K.; Balczer, J.; Antal, J.; Graf, L.; Laich, A.; Moffatt, B.E.; Schwaeble, W.; et al. Natural substrates and inhibitors of mannan-binding lectin-associated serine protease-1 and -2: A study on recombinant catalytic fragments. J. Immunol. 2003, 170, 1374–1382. [Google Scholar] [CrossRef]
  17. Megyeri, M.; Harmat, V.; Major, B.; Vegh, A.; Balczer, J.; Heja, D.; Szilagyi, K.; Datz, D.; Pal, G.; Zavodszky, P.; et al. Quantitative characterization of the activation steps of mannan-binding lectin (MBL)-associated serine proteases (MASPs) points to the central role of MASP-1 in the initiation of the complement lectin pathway. J. Biol. Chem. 2013, 288, 8922–8934. [Google Scholar] [CrossRef]
  18. Degn, S.E.; Jensen, L.; Olszowski, T.; Jensenius, J.C.; Thiel, S. Co-complexes of MASP-1 and MASP-2 associated with the soluble pattern-recognition molecules drive lectin pathway activation in a manner inhibitable by MAp44. J. Immunol. 2013, 191, 1334–1345. [Google Scholar] [CrossRef] [PubMed]
  19. Parej, K.; Hermann, A.; Donath, N.; Zavodszky, P.; Gal, P.; Dobo, J. Dissociation and re-association studies on the interaction domains of mannan-binding lectin (MBL)-associated serine proteases, MASP-1 and MASP-2, provide evidence for heterodimer formation. Mol. Immunol. 2014, 59, 1–9. [Google Scholar] [CrossRef] [PubMed]
  20. Wallis, R.; Shaw, J.M.; Uitdehaag, J.; Chen, C.B.; Torgersen, D.; Drickamer, K. Localization of the serine protease-binding sites in the collagen-like domain of mannose-binding protein: Indirect effects of naturally occurring mutations on protease binding and activation. J. Biol. Chem. 2004, 279, 14065–14073. [Google Scholar] [CrossRef] [PubMed]
  21. Iwaki, D.; Kanno, K.; Takahashi, M.; Endo, Y.; Lynch, N.J.; Schwaeble, W.J.; Matsushita, M.; Okabe, M.; Fujita, T. Small mannose-binding lectin-associated protein plays a regulatory role in the lectin complement pathway. J. Immunol. 2006, 177, 8626–8632. [Google Scholar] [CrossRef] [PubMed]
  22. Rosbjerg, A.; Munthe-Fog, L.; Garred, P.; Skjoedt, M.O. Heterocomplex formation between MBL/ficolin/CL-11-associated serine protease-1 and -3 and MBL/ficolin/CL-11-associated protein-1. J. Immunol. 2014, 192, 4352–4360. [Google Scholar] [CrossRef] [PubMed]
  23. Gulla, K.C.; Gupta, K.; Krarup, A.; Gal, P.; Schwaeble, W.J.; Sim, R.B.; O’Connor, C.D.; Hajela, K. Activation of mannan-binding lectin-associated serine proteases leads to generation of a fibrin clot. Immunology 2010, 129, 482–495. [Google Scholar] [CrossRef] [PubMed]
  24. Takahashi, K.; Chang, W.C.; Takahashi, M.; Pavlov, V.; Ishida, Y.; la Bonte, L.; Shi, L.; Fujita, T.; Stahl, G.L.; van Cott, E.M. Mannose-binding lectin and its associated proteases (MASPs) mediate coagulation and its deficiency is a risk factor in developing complications from infection, including disseminated intravascular coagulation. Immunobiology 2011, 216, 96–102. [Google Scholar] [CrossRef] [PubMed]
  25. Sirmaci, A.; Walsh, T.; Akay, H.; Spiliopoulos, M.; Sakalar, Y.B.; Hasanefendioglu-Bayrak, A.; Duman, D.; Farooq, A.; King, M.C.; Tekin, M. MASP1 mutations in patients with facial, umbilical, coccygeal, and auditory findings of Carnevale, Malpuech, OSA, and Michels syndromes. Am. J. Hum. Genet. 2010, 87, 679–686. [Google Scholar] [CrossRef] [PubMed]
  26. Sastry, K.; Herman, G.A.; Day, L.; Deignan, E.; Bruns, G.; Morton, C.C.; Ezekowitz, R.A. The human mannose-binding protein gene. Exon structure reveals its evolutionary relationship to a human pulmonary surfactant gene and localization to chromosome 10. J. Exp. Med. 1989, 170, 1175–1189. [Google Scholar] [CrossRef] [PubMed]
  27. Singh, K.K.; Nathamu, S.; Adame, A.; Alire, T.U.; Dumaop, W.; Gouaux, B.; Moore, D.J.; Masliah, E.; HIV Neurobehavioral Research Center Group. Expression of mannose binding lectin in HIV-1-infected brain: Implications for HIV-related neuronal damage and neuroAIDS. Neurobehav HIV Med. 2011, 3, 41–52. [Google Scholar] [CrossRef]
  28. Jensen, P.H.; Weilguny, D.; Matthiesen, F.; McGuire, K.A.; Shi, L.; Hojrup, P. Characterization of the oligomer structure of recombinant human mannan-binding lectin. J. Biol. Chem. 2005, 280, 11043–11051. [Google Scholar] [CrossRef] [PubMed]
  29. Yokota, Y.; Arai, T.; Kawasaki, T. Oligomeric structures required for complement activation of serum mannan-binding proteins. J. Biochem. 1995, 117, 414–419. [Google Scholar] [CrossRef] [PubMed]
  30. Jensenius, H.; Klein, D.C.; van Hecke, M.; Oosterkamp, T.H.; Schmidt, T.; Jensenius, J.C. Mannan-binding lectin: Structure, oligomerization, and flexibility studied by atomic force microscopy. J. Mol. Biol. 2009, 391, 246–259. [Google Scholar] [CrossRef] [PubMed]
  31. Sheriff, S.; Chang, C.Y.; Ezekowitz, R.A. Human mannose-binding protein carbohydrate recognition domain trimerizes through a triple alpha-helical coiled-coil. Nat. Struct. Biol. 1994, 1, 789–794. [Google Scholar] [CrossRef] [PubMed]
  32. Miller, A.; Phillips, A.; Gor, J.; Wallis, R.; Perkins, S.J. Near-planar solution structures of mannose-binding lectin oligomers provide insight on activation of lectin pathway of complement. J. Biol. Chem. 2012, 287, 3930–3945. [Google Scholar] [CrossRef] [PubMed]
  33. Teillet, F.; Lacroix, M.; Thiel, S.; Weilguny, D.; Agger, T.; Arlaud, G.J.; Thielens, N.M. Identification of the site of human mannan-binding lectin involved in the interaction with its partner serine proteases: The essential role of Lys55. J. Immunol. 2007, 178, 5710–5716. [Google Scholar] [CrossRef] [PubMed]
  34. Weis, W.I.; Drickamer, K.; Hendrickson, W.A. Structure of a C-type mannose-binding protein complexed with an oligosaccharide. Nature 1992, 360, 127–134. [Google Scholar] [CrossRef] [PubMed]
  35. Kilpatrick, D.C. Phospholipid-binding activity of human mannan-binding lectin. Immunol. Lett. 1998, 61, 191–195. [Google Scholar] [CrossRef] [PubMed]
  36. Palaniyar, N.; Nadesalingam, J.; Clark, H.; Shih, M.J.; Dodds, A.W.; Reid, K.B. Nucleic acid is a novel ligand for innate, immune pattern recognition collectins surfactant proteins A and D and mannose-binding lectin. J. Biol. Chem. 2004, 279, 32728–32736. [Google Scholar] [CrossRef] [PubMed]
  37. Teillet, F.; Dublet, B.; Andrieu, J.P.; Gaboriaud, C.; Arlaud, G.J.; Thielens, N.M. The two major oligomeric forms of human mannan-binding lectin: Chemical characterization, carbohydrate-binding properties, and interaction with MBL-associated serine proteases. J. Immunol. 2005, 174, 2870–2877. [Google Scholar] [CrossRef] [PubMed]
  38. Saifuddin, M.; Hart, M.L.; Gewurz, H.; Zhang, Y.; Spear, G.T. Interaction of mannose-binding lectin with primary isolates of human immunodeficiency virus type 1. J. Gen. Virol. 2000, 81, 949–955. [Google Scholar] [PubMed]
  39. Wei, X.; Decker, J.M.; Wang, S.; Hui, H.; Kappes, J.C.; Wu, X.; Salazar-Gonzalez, J.F.; Salazar, M.G.; Kilby, J.M.; Saag, M.S.; et al. Antibody neutralization and escape by HIV-1. Nature 2003, 422, 307–312. [Google Scholar]
  40. Ying, H.; Ji, X.; Hart, M.L.; Gupta, K.; Saifuddin, M.; Zariffard, M.R.; Spear, G.T. Interaction of mannose-binding lectin with HIV type 1 is sufficient for virus opsonization but not neutralization. AIDS Res. Hum. Retrovir. 2004, 20, 327–335. [Google Scholar] [CrossRef] [PubMed]
  41. Senaldi, G.; Davies, E.T.; Mahalingam, M.; Lu, J.; Pozniak, A.; Peakman, M.; Reid, K.B.; Vergani, D. Circulating levels of mannose binding protein in human immunodeficiency virus infection. J. Infect. 1995, 31, 145–148. [Google Scholar] [CrossRef] [PubMed]
  42. Jack, D.L.; Lee, M.E.; Turner, M.W.; Klein, N.J.; Read, R.C. Mannose-binding lectin enhances phagocytosis and killing of Neisseria meningitidis by human macrophages. J. Leukoc. Biol. 2005, 77, 328–336. [Google Scholar] [CrossRef] [PubMed]
  43. Pagh, R.; Duus, K.; Laursen, I.; Hansen, P.R.; Mangor, J.; Thielens, N.; Arlaud, G.J.; Kongerslev, L.; Hojrup, P.; Houen, G. The chaperone and potential mannan-binding lectin (MBL) co-receptor calreticulin interacts with MBL through the binding site for MBL-associated serine proteases. FEBS J. 2008, 275, 515–526. [Google Scholar] [CrossRef] [PubMed]
  44. Spear, G.T.; Zariffard, M.R.; Xin, J.; Saifuddin, M. Inhibition of DC-SIGN-mediated trans infection of T cells by mannose-binding lectin. Immunology 2003, 110, 80–85. [Google Scholar] [CrossRef] [PubMed]
  45. Chang, W.C.; White, M.R.; Moyo, P.; McClear, S.; Thiel, S.; Hartshorn, K.L.; Takahashi, K. Lack of the pattern recognition molecule mannose-binding lectin increases susceptibility to influenza A virus infection. BMC Immunol. 2010, 11, 64. [Google Scholar] [CrossRef] [PubMed]
  46. Kase, T.; Suzuki, Y.; Kawai, T.; Sakamoto, T.; Ohtani, K.; Eda, S.; Maeda, A.; Okuno, Y.; Kurimura, T.; Wakamiya, N. Human mannan-binding lectin inhibits the infection of influenza A virus without complement. Immunology 1999, 97, 385–392. [Google Scholar] [CrossRef] [PubMed]
  47. Brown, K.S.; Keogh, M.J.; Owsianka, A.M.; Adair, R.; Patel, A.H.; Arnold, J.N.; Ball, J.K.; Sim, R.B.; Tarr, A.W.; Hickling, T.P. Specific interaction of hepatitis C virus glycoproteins with mannan binding lectin inhibits virus entry. Protein Cell 2010, 1, 664–674. [Google Scholar] [CrossRef] [PubMed]
  48. Ip, W.K.; Chan, K.H.; Law, H.K.; Tso, G.H.; Kong, E.K.; Wong, W.H.; To, Y.F.; Yung, R.W.; Chow, E.Y.; Au, K.L.; et al. Mannose-binding lectin in severe acute respiratory syndrome coronavirus infection. J. Infect. Dis. 2005, 191, 1697–1704. [Google Scholar] [CrossRef] [PubMed]
  49. Zhou, Y.; Lu, K.; Pfefferle, S.; Bertram, S.; Glowacka, I.; Drosten, C.; Pohlmann, S.; Simmons, G. A single asparagine-linked glycosylation site of the severe acute respiratory syndrome coronavirus spike glycoprotein facilitates inhibition by mannose-binding lectin through multiple mechanisms. J. Virol. 2010, 84, 8753–8764. [Google Scholar] [CrossRef] [PubMed]
  50. Avirutnan, P.; Hauhart, R.E.; Marovich, M.A.; Garred, P.; Atkinson, J.P.; Diamond, M.S. Complement-mediated neutralization of dengue virus requires mannose-binding lectin. MBio 2011, 2. [Google Scholar] [CrossRef] [PubMed]
  51. Fuchs, A.; Lin, T.Y.; Beasley, D.W.; Stover, C.M.; Schwaeble, W.J.; Pierson, T.C.; Diamond, M.S. Direct complement restriction of flavivirus infection requires glycan recognition by mannose-binding lectin. Cell Host Microbe 2010, 8, 186–195. [Google Scholar] [CrossRef] [PubMed]
  52. El Saadany, S.A.; Ziada, D.H.; Farrag, W.; Hazaa, S. Fibrosis severity and mannan-binding lectin (MBL)/MBL-associated serine protease 1 (MASP-1) complex in HCV-infected patients. Arab. J. Gastroenterol. 2011, 12, 68–73. [Google Scholar]
  53. Michelow, I.C.; Lear, C.; Scully, C.; Prugar, L.I.; Longley, C.B.; Yantosca, L.M.; Ji, X.; Karpel, M.; Brudner, M.; Takahashi, K.; et al. High-dose mannose-binding lectin therapy for Ebola virus infection. J. Infect. Dis. 2011, 203, 175–179. [Google Scholar] [CrossRef] [PubMed]
  54. Ji, X.; Olinger, G.G.; Aris, S.; Chen, Y.; Gewurz, H.; Spear, G.T. Mannose-binding lectin binds to Ebola and Marburg envelope glycoproteins, resulting in blocking of virus interaction with DC-SIGN and complement-mediated virus neutralization. J. Gen. Virol. 2005, 86, 2535–2542. [Google Scholar] [CrossRef] [PubMed]
  55. Brudner, M.; Karpel, M.; Lear, C.; Chen, L.; Yantosca, L.M.; Scully, C.; Sarraju, A.; Sokolovska, A.; Zariffard, M.R.; Eisen, D.P.; et al. Lectin-dependent enhancement of Ebola virus infection via soluble and transmembrane C-type lectin receptors. PLoS One 2013, 8, e60838. [Google Scholar] [CrossRef] [PubMed]
  56. Bachis, A.; Aden, S.A.; Nosheny, R.L.; Andrews, P.M.; Mocchetti, I. Axonal transport of human immunodeficiency virus type 1 envelope protein glycoprotein 120 is found in association with neuronal apoptosis. J. Neurosci. 2006, 26, 6771–6780. [Google Scholar] [CrossRef] [PubMed]
  57. Teodorof, C.; Divakar, S.; Soontornniyomkij, B.; Achim, C.L.; Kaul, M.; Singh, K.K. Intracellular mannose binding lectin mediates subcellular trafficking of HIV-1 gp120 in neurons. Neurobiol. Dis. 2014, 69, 54–64. [Google Scholar] [CrossRef] [PubMed]
  58. Sumiya, M.; Super, M.; Tabona, P.; Levinsky, R.J.; Arai, T.; Turner, M.W.; Summerfield, J.A. Molecular basis of opsonic defect in immunodeficient children. Lancet 1991, 337, 1569–1570. [Google Scholar] [CrossRef] [PubMed]
  59. Lipscombe, R.J.; Sumiya, M.; Hill, A.V.; Lau, Y.L.; Levinsky, R.J.; Summerfield, J.A.; Turner, M.W. High frequencies in African and non-African populations of independent mutations in the mannose binding protein gene. Hum. Mol. Genet. 1992, 1, 709–715. [Google Scholar] [CrossRef] [PubMed]
  60. Madsen, H.O.; Garred, P.; Kurtzhals, J.A.; Lamm, L.U.; Ryder, L.P.; Thiel, S.; Svejgaard, A. A new frequent allele is the missing link in the structural polymorphism of the human mannan-binding protein. Immunogenetics 1994, 40, 37–44. [Google Scholar] [CrossRef] [PubMed]
  61. Madsen, H.O.; Satz, M.L.; Hogh, B.; Svejgaard, A.; Garred, P. Different molecular events result in low protein levels of mannan-binding lectin in populations from southeast Africa and South America. J. Immunol. 1998, 161, 3169–3175. [Google Scholar] [PubMed]
  62. Chalmers, J.D.; McHugh, B.J.; Doherty, C.; Smith, M.P.; Govan, J.R.; Kilpatrick, D.C.; Hill, A.T. Mannose-binding lectin deficiency and disease severity in non-cystic fibrosis bronchiectasis: A prospective study. Lancet Respir. Med. 2013, 1, 224–232. [Google Scholar] [CrossRef] [PubMed]
  63. Lipscombe, R.J.; Sumiya, M.; Summerfield, J.A.; Turner, M.W. Distinct physicochemical characteristics of human mannose binding protein expressed by individuals of differing genotype. Immunology 1995, 85, 660–667. [Google Scholar] [PubMed]
  64. Verga Falzacappa, M.V.; Segat, L.; Puppini, B.; Amoroso, A.; Crovella, S. Evolution of the mannose-binding lectin gene in primates. Genes Immun. 2004, 5, 653–661. [Google Scholar]
  65. Garred, P.; Larsen, F.; Seyfarth, J.; Fujita, R.; Madsen, H.O. Mannose-binding lectin and its genetic variants. Genes Immun. 2006, 7, 85–94. [Google Scholar] [CrossRef] [PubMed]
  66. Soltani, A.; RahmatiRad, S.; Pourpak, Z.; Alizadeh, Z.; Saghafi, S.; HajiBeigi, B.; Zeidi, M.; Farazmand, A. Polymorphisms and serum level of mannose-binding lectin: An Iranian survey. Iran. J. Allergy Asthma Immunol. 2014, 13, 428–432. [Google Scholar] [PubMed]
  67. Lee, S.G.; Yum, J.S.; Moon, H.M.; Kim, H.J.; Yang, Y.J.; Kim, H.L.; Yoon, Y.; Lee, S.; Song, K. Analysis of mannose-binding lectin 2 (MBL2) genotype and the serum protein levels in the Korean population. Mol. Immunol. 2005, 42, 969–977. [Google Scholar] [CrossRef] [PubMed]
  68. Thiel, S.; Bjerke, T.; Hansen, D.; Poulsen, L.K.; Schiotz, P.O.; Jensenius, J.C. Ontogeny of human mannan-binding protein, a lectin of the innate immune system. Pediatr. Allergy Immunol. 1995, 6, 20–23. [Google Scholar] [CrossRef] [PubMed]
  69. Roos, A.; Garred, P.; Wildenberg, M.E.; Lynch, N.J.; Munoz, J.R.; Zuiverloon, T.C.; Bouwman, L.H.; Schlagwein, N.; Fallaux van den Houten, F.C.; Faber-Krol, M.C.; et al. Antibody-mediated activation of the classical pathway of complement may compensate for mannose-binding lectin deficiency. Eur. J. Immunol. 2004, 34, 2589–2598. [Google Scholar] [CrossRef] [PubMed]
  70. Garred, P.; Madsen, H.O.; Balslev, U.; Hofmann, B.; Pedersen, C.; Gerstoft, J.; Svejgaard, A. Susceptibility to HIV infection and progression of AIDS in relation to variant alleles of mannose-binding lectin. Lancet 1997, 349, 236–240. [Google Scholar] [CrossRef] [PubMed]
  71. Prohaszka, Z.; Thiel, S.; Ujhelyi, E.; Szlavik, J.; Banhegyi, D.; Fust, G. Mannan-binding lectin serum concentrations in HIV-infected patients are influenced by the stage of disease. Immunol. Lett. 1997, 58, 171–175. [Google Scholar] [CrossRef] [PubMed]
  72. Catano, G.; Agan, B.K.; Kulkarni, H.; Telles, V.; Marconi, V.C.; Dolan, M.J.; Ahuja, S.K. Independent effects of genetic variations in mannose-binding lectin influence the course of HIV disease: The advantage of heterozygosity for coding mutations. J. Infect. Dis. 2008, 198, 72–80. [Google Scholar] [CrossRef] [PubMed]
  73. Vallinoto, A.C.; Menezes-Costa, M.R.; Alves, A.E.; Machado, L.F.; de Azevedo, V.N.; Souza, L.L.; Ishak Mde, O.; Ishak, R. Mannose-binding lectin gene polymorphism and its impact on human immunodeficiency virus 1 infection. Mol. Immunol. 2006, 43, 1358–1362. [Google Scholar] [CrossRef] [PubMed]
  74. Boniotto, M.; Crovella, S.; Pirulli, D.; Scarlatti, G.; Spano, A.; Vatta, L.; Zezlina, S.; Tovo, P.A.; Palomba, E.; Amoroso, A. Polymorphisms in the MBL2 promoter correlated with risk of HIV-1 vertical transmission and AIDS progression. Genes Immun. 2000, 1, 346–348. [Google Scholar] [CrossRef] [PubMed]
  75. Vallinoto, A.C.; Muto, N.A.; Alves, A.E.; Machado, L.F.; Azevedo, V.N.; Souza, L.L.; Ishak, M.O.; Ishak, R. Characterization of polymorphisms in the mannose-binding lectin gene promoter among human immunodeficiency virus 1 infected subjects. Mem. Inst. Oswaldo Cruz. 2008, 103, 645–649. [Google Scholar] [CrossRef] [PubMed]
  76. Malik, S.; Arias, M.; di Flumeri, C.; Garcia, L.F.; Schurr, E. Absence of association between mannose-binding lectin gene polymorphisms and HIV-1 infection in a Colombian population. Immunogenetics 2003, 55, 49–52. [Google Scholar] [PubMed]
  77. Lian, Y.C.; Della-Negra, M.; Rutz, R.; Ferriani, V.; de Moraes Vasconcelos, D.; da Silva Duarte, A.J.; Kirschfink, M.; Grumach, A.S. Immunological analysis in paediatric HIV patients at different stages of the disease. Scand. J. Immunol. 2004, 60, 615–624. [Google Scholar] [CrossRef] [PubMed]
  78. Nielsen, S.L.; Andersen, P.L.; Koch, C.; Jensenius, J.C.; Thiel, S. The level of the serum opsonin, mannan-binding protein in HIV-1 antibody-positive patients. Clin. Exp. Immunol. 1995, 100, 219–222. [Google Scholar] [CrossRef] [PubMed]
  79. Heggelund, L.; Mollnes, T.E.; Ueland, T.; Christophersen, B.; Aukrust, P.; Froland, S.S. Mannose-binding lectin in HIV infection: Relation to disease progression and highly active antiretroviral therapy. J. Acquir. Immune Defic. Syndr. 2003, 32, 354–361. [Google Scholar] [CrossRef] [PubMed]
  80. Maas, J.; de Roda Husman, A.M.; Brouwer, M.; Krol, A.; Coutinho, R.; Keet, I.; van Leeuwen, R.; Schuitemaker, H. Presence of the variant mannose-binding lectin alleles associated with slower progression to AIDS. Amsterdam Cohort Study. AIDS 1998, 12, 2275–2280. [Google Scholar] [CrossRef] [PubMed]
  81. Chong, W.P.; To, Y.F.; Ip, W.K.; Yuen, M.F.; Poon, T.P.; Wong, W.H.; Lai, C.L.; Lau, Y.L. Mannose-binding lectin in chronic hepatitis B virus infection. Hepatology 2005, 42, 1037–1045. [Google Scholar] [CrossRef] [PubMed]
  82. Halla, M.C.; do Carmo, R.F.; Silva Vasconcelos, L.R.; Pereira, L.B.; Moura, P.; de Siqueira, E.R.; Pereira, L.M.; Mendonca Cavalcanti Mdo, S. Association of hepatitis C virus infection and liver fibrosis severity with the variants alleles of MBL2 gene in a Brazilian population. Hum. Immunol. 2010, 71, 883–887. [Google Scholar] [CrossRef] [PubMed]
  83. Alves Pedroso, M.L.; Boldt, A.B.; Pereira-Ferrari, L.; Steffensen, R.; Strauss, E.; Jensenius, J.C.; Ioshii, S.O.; Messias-Reason, I. Mannan-binding lectin MBL2 gene polymorphism in chronic hepatitis C: Association with the severity of liver fibrosis and response to interferon therapy. Clin. Exp. Immunol. 2008, 152, 258–264. [Google Scholar]
  84. Alves, A.E.; Hermes, R.B.; Tamegao-Lopes, B.; Machado, L.F.; Azevedo, V.N.; Ishak, M.O.; Ishak, R.; Lemos, J.A.; Vallinoto, A.C. Polymorphism in the promoter region of the mannose-binding lectin gene among human T-cell lymphotropic virus infected subjects. Mem. Inst. Oswaldo Cruz. 2007, 102, 991–994. [Google Scholar] [CrossRef]
  85. Kwakkel-van Erp, J.M.; Paantjens, A.W.; van Kessel, D.A.; Grutters, J.C.; van den Bosch, J.M.; van de Graaf, E.A.; Otten, H.G. Mannose-binding lectin deficiency linked to cytomegalovirus (CMV) reactivation and survival in lung transplantation. Clin. Exp. Immunol. 2011, 165, 410–416. [Google Scholar]
  86. De Rooij, B.J.; van der Beek, M.T.; van Hoek, B.; Vossen, A.C.; Rogier Ten Hove, W.; Roos, A.; Schaapherder, A.F.; Porte, R.J.; van der Reijden, J.J.; Coenraad, M.J.; et al. Mannose-binding lectin and ficolin-2 gene polymorphisms predispose to cytomegalovirus (re)infection after orthotopic liver transplantation. J. Hepatol. 2011, 55, 800–807. [Google Scholar]
  87. Thomas, H.C.; Foster, G.R.; Sumiya, M.; McIntosh, D.; Jack, D.L.; Turner, M.W.; Summerfield, J.A. Mutation of gene of mannose-binding protein associated with chronic hepatitis B viral infection. Lancet 1996, 348, 1417–1419. [Google Scholar] [CrossRef] [PubMed]
  88. Xu, H.D.; Zhao, M.F.; Wan, T.H.; Song, G.Z.; He, J.L.; Chen, Z. Association between Mannose-binding lectin gene polymorphisms and hepatitis B virus infection: A meta-analysis. PLoS One 2013, 8, e75371. [Google Scholar] [CrossRef] [PubMed]
  89. Koutsounaki, E.; Goulielmos, G.N.; Koulentaki, M.; Choulaki, C.; Kouroumalis, E.; Galanakis, E. Mannose-binding lectin MBL2 gene polymorphisms and outcome of hepatitis C virus-infected patients. J. Clin. Immunol. 2008, 28, 495–500. [Google Scholar] [CrossRef] [PubMed]
  90. Abdelaal, A.; Mossad, N.; Abdel Hafez, H.; Elsayed, N. Mannose-binding lectin exon 1 polymorphisms in Egyptian patients with chronic hepatitis C virus infection. Comp. Clin. Pathol. 2014, 23, 1339–1342. [Google Scholar] [CrossRef]
  91. Yuen, M.F.; Lau, C.S.; Lau, Y.L.; Wong, W.M.; Cheng, C.C.; Lai, C.L. Mannose binding lectin gene mutations are associated with progression of liver disease in chronic hepatitis B infection. Hepatology 1999, 29, 1248–1251. [Google Scholar] [CrossRef] [PubMed]
  92. Pontes, G.S.; Tamegao-Lopes, B.; Machado, L.F.; Azevedo, V.N.; Ishak, M.O.; Ishak, R.; Lemos, J.A.; Vallinoto, A.C. Characterization of mannose-binding lectin gene polymorphism among human T-cell lymphotropic virus 1 and 2-infected asymptomatic subjects. Hum. Immunol. 2005, 66, 892–896. [Google Scholar] [CrossRef] [PubMed]
  93. Ammitzbøll, C.G.; Kjaer, T.R.; Steffensen, R.; Stengaard-Pedersen, K.; Nielsen, H.J.; Thiel, S.; Bogsted, M.; Jensenius, J.C. Non-synonymous polymorphisms in the FCN1 gene determine ligand-binding ability and serum levels of M-ficolin. PLoS One 2012, 7, e50585. [Google Scholar] [CrossRef] [PubMed]
  94. Hummelshøj, T.; Munthe-Fog, L.; Madsen, H.O.; Garred, P. Functional SNPs in the human ficolin (FCN) genes reveal distinct geographical patterns. Mol. Immunol. 2008, 45, 2508–2520. [Google Scholar] [CrossRef] [PubMed]
  95. Hoang, T.V.; Toan, N.L.; Song le, H.; Ouf, E.A.; Bock, C.T.; Kremsner, P.G.; Kun, J.F.; Velavan, T.P. Ficolin-2 levels and FCN2 haplotypes influence hepatitis B infection outcome in Vietnamese patients. PLoS One 2011, 6, e28113. [Google Scholar] [CrossRef] [PubMed]
  96. Ma, Y.J.; Doni, A.; Hummelshoj, T.; Honore, C.; Bastone, A.; Mantovani, A.; Thielens, N.M.; Garred, P. Synergy between ficolin-2 and pentraxin 3 boosts innate immune recognition and complement deposition. J. Biol. Chem. 2009, 284, 28263–28275. [Google Scholar] [CrossRef] [PubMed]
  97. Munthe-Fog, L.; Hummelshoj, T.; Ma, Y.J.; Hansen, B.E.; Koch, C.; Madsen, H.O.; Skjodt, K.; Garred, P. Characterization of a polymorphism in the coding sequence of FCN3 resulting in a Ficolin-3 (Hakata antigen) deficiency state. Mol. Immunol. 2008, 45, 2660–2666. [Google Scholar] [CrossRef] [PubMed]
  98. Munthe-Fog, L.; Hummelshoj, T.; Honore, C.; Madsen, H.O.; Permin, H.; Garred, P. Immunodeficiency associated with FCN3 mutation and ficolin-3 deficiency. N. Engl. J. Med. 2009, 360, 2637–2644. [Google Scholar] [CrossRef] [PubMed]
  99. Schlapbach, L.J.; Thiel, S.; Kessler, U.; Ammann, R.A.; Aebi, C.; Jensenius, J.C. Congenital H-ficolin deficiency in premature infants with severe necrotising enterocolitis. Gut 2011, 60, 1438–1439. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, L.; Chen, R.F.; Liu, J.W.; Lee, I.K.; Lee, C.P.; Kuo, H.C.; Huang, S.K.; Yang, K.D. DC-SIGN (CD209) Promoter -336 A/G polymorphism is associated with dengue hemorrhagic fever and correlated to DC-SIGN expression and immune augmentation. PLoS Negl. Trop. Dis. 2011, 5, e934. [Google Scholar] [CrossRef] [PubMed]
  101. Sakuntabhai, A.; Turbpaiboon, C.; Casademont, I.; Chuansumrit, A.; Lowhnoo, T.; Kajaste-Rudnitski, A.; Kalayanarooj, S.M.; Tangnararatchakit, K.; Tangthawornchaikul, N.; Vasanawathana, S.; et al. A variant in the CD209 promoter is associated with severity of dengue disease. Nat. Genet. 2005, 37, 507–513. [Google Scholar] [CrossRef] [PubMed]
  102. Oliveira, L.F.; Lima, C.P.; Azevedo Rdo, S.; Mendonca, D.S.; Rodrigues, S.G.; Carvalho, V.L.; Pinto, E.V.; Maia, A.L.; Maia, M.H.; Vasconcelos, J.M.; et al. Polymorphism of DC-SIGN (CD209) promoter in association with clinical symptoms of dengue fever. Viral Immunol. 2014, 27, 245–249. [Google Scholar] [CrossRef] [PubMed]
  103. Ryan, E.J.; Dring, M.; Ryan, C.M.; McNulty, C.; Stevenson, N.J.; Lawless, M.W.; Crowe, J.; Nolan, N.; Hegarty, J.E.; O’Farrelly, C. Variant in CD209 promoter is associated with severity of liver disease in chronic hepatitis C virus infection. Hum. Immunol. 2010, 71, 829–832. [Google Scholar] [CrossRef] [PubMed]
  104. Boily-Larouche, G.; Milev, M.P.; Zijenah, L.S.; Labbe, A.C.; Zannou, D.M.; Humphrey, J.H.; Ward, B.J.; Poudrier, J.; Mouland, A.J.; Cohen, E.A.; et al. Naturally-occurring genetic variants in human DC-SIGN increase HIV-1 capture, cell-transfer and risk of mother-to-child transmission. PLoS One 2012, 7, e40706. [Google Scholar] [CrossRef] [PubMed]
  105. Martin, M.P.; Lederman, M.M.; Hutcheson, H.B.; Goedert, J.J.; Nelson, G.W.; van Kooyk, Y.; Detels, R.; Buchbinder, S.; Hoots, K.; Vlahov, D.; et al. Association of DC-SIGN promoter polymorphism with increased risk for parenteral, but not mucosal, acquisition of human immunodeficiency virus type 1 infection. J. Virol. 2004, 78, 14053–14056. [Google Scholar] [CrossRef] [PubMed]
  106. Da Silva, R.C.; Segat, L.; Zanin, V.; Arraes, L.C.; Crovella, S. Polymorphisms in DC-SIGN and L-SIGN genes are associated with HIV-1 vertical transmission in a Northeastern Brazilian population. Hum. Immunol. 2012, 73, 1159–1165. [Google Scholar] [CrossRef] [PubMed]
  107. Chan, K.Y.; Xu, M.S.; Ching, J.C.; So, T.M.; Lai, S.T.; Chu, C.M.; Yam, L.Y.; Wong, A.T.; Chung, P.H.; Chan, V.S.; et al. CD209 (DC-SIGN) -336A>G promoter polymorphism and severe acute respiratory syndrome in Hong Kong Chinese. Hum. Immunol. 2010, 71, 702–707. [Google Scholar] [CrossRef] [PubMed]
  108. Kagone, T.S.; Bisseye, C.; Meda, N.; Testa, J.; Pietra, V.; Kania, D.; Yonli, A.T.; Compaore, T.R.; Nikiema, J.B.; de Souza, C.; et al. A variant of DC-SIGN gene promoter associated with resistance to HIV-1 in serodiscordant couples in Burkina Faso. Asian Pac. J. Trop. Med. 2014, 7S1, S93–S96. [Google Scholar] [CrossRef]
  109. Eisen, D.P.; Marshall, C.; Dean, M.M.; Sasadeusz, J.; Richards, M.; Buising, K.; Cheng, A.; Johnson, P.D.; Barr, I.G.; McBryde, E.S. No association between mannose-binding lectin deficiency and H1N1 2009 infection observed during the first season of this novel pandemic influenza virus. Hum. Immunol. 2011, 72, 1091–1094. [Google Scholar] [CrossRef] [PubMed]
  110. Brown, E.E.; Zhang, M.; Zarin-Pass, R.; Bernig, T.; Tseng, F.C.; Xiao, N.; Yeager, M.; Edlin, B.R.; Chanock, S.J.; O’Brien, T.R. MBL2 and hepatitis C virus infection among injection drug users. BMC Infect. Dis. 2008, 8, 57. [Google Scholar] [CrossRef] [PubMed]
  111. Vallinoto, A.C.; da Silva, R.F.; Hermes, R.B.; Amaral, I.S.; Miranda, E.C.; Barbosa, M.S.; Moia Lde, J.; Conde, S.R.; Soares Mdo, C.; Lemos, J.A.; et al. Mannose-binding lectin gene polymorphisms are not associated with susceptibility to hepatitis C virus infection in the Brazilian Amazon region. Hum. Immunol. 2009, 70, 754–757. [Google Scholar] [CrossRef] [PubMed]
  112. Esmat, S.; Omran, D.; Sleem, G.A.; Rashed, L. Serum mannan-binding lectin in egyptian patients with chronic hepatitis C: Its relation to disease progression and response to treatment. Hepat. Mon. 2012, 12, 259–264. [Google Scholar] [CrossRef] [PubMed]
  113. Wu, Y.; Zhou, Q.; Wang, H.; Tian, T.; Zhu, Q.; Wang, H.; Bai, X.; Yang, X.; Wang, Z.; Dong, M. Potential role of mannose-binding lectin in intrauterine transmission of hepatitis B virus. Jpn. J. Infect. Dis. 2013, 66, 391–393. [Google Scholar] [PubMed]
  114. Hӧhler, T.; Wunschel, M.; Gerken, G.; Schneider, P.M.; Meyer zum Buschenfelde, K.H.; Rittner, C. No association between mannose-binding lectin alleles and susceptibility to chronic hepatitis B virus infection in German patients. Exp. Clin. Immunogenet. 1998, 15, 130–133. [Google Scholar] [CrossRef] [PubMed]
  115. Endo, Y.; Sato, Y.; Matsushita, M.; Fujita, T. Cloning and characterization of the human lectin P35 gene and its related gene. Genomics 1996, 36, 515–521. [Google Scholar] [CrossRef] [PubMed]
  116. Ichijo, H.; Ronnstrand, L.; Miyagawa, K.; Ohashi, H.; Heldin, C.H.; Miyazono, K. Purification of transforming growth factor-beta 1 binding proteins from porcine uterus membranes. J. Biol. Chem. 1991, 266, 22459–22464. [Google Scholar] [PubMed]
  117. Lu, J.; Le, Y.; Kon, O.L.; Chan, J.; Lee, S.H. Biosynthesis of human ficolin, an Escherichia coli-binding protein, by monocytes: Comparison with the synthesis of two macrophage-specific proteins, C1q and the mannose receptor. Immunology 1996, 89, 289–294. [Google Scholar] [CrossRef] [PubMed]
  118. Sugimoto, R.; Yae, Y.; Akaiwa, M.; Kitajima, S.; Shibata, Y.; Sato, H.; Hirata, J.; Okochi, K.; Izuhara, K.; Hamasaki, N. Cloning and characterization of the Hakata antigen, a member of the ficolin/opsonin p35 lectin family. J. Biol. Chem. 1998, 273, 20721–20727. [Google Scholar] [CrossRef] [PubMed]
  119. Kenjo, A.; Takahashi, M.; Matsushita, M.; Endo, Y.; Nakata, M.; Mizuochi, T.; Fujita, T. Cloning and characterization of novel ficolins from the solitary ascidian, Halocynthia roretzi. J. Biol. Chem. 2001, 276, 19959–19965. [Google Scholar] [CrossRef] [PubMed]
  120. Lynch, N.J.; Khan, S.U.; Stover, C.M.; Sandrini, S.M.; Marston, D.; Presanis, J.S.; Schwaeble, W.J. Composition of the lectin pathway of complement in Gallus gallus: Absence of mannan-binding lectin-associated serine protease-1 in birds. J. Immunol. 2005, 174, 4998–5006. [Google Scholar] [CrossRef] [PubMed]
  121. Hummelshøj, T.; Nissen, J.; Munthe-Fog, L.; Koch, C.; Bertelsen, M.F.; Garred, P. Allelic lineages of the ficolin genes (FCNs) are passed from ancestral to descendant primates. PLoS One 2011, 6, e28187. [Google Scholar] [CrossRef] [PubMed]
  122. Ichijo, H.; Hellman, U.; Wernstedt, C.; Gonez, L.J.; Claesson-Welsh, L.; Heldin, C.H.; Miyazono, K. Molecular cloning and characterization of ficolin, a multimeric protein with fibrinogen- and collagen-like domains. J. Biol. Chem. 1993, 268, 14505–14513. [Google Scholar] [PubMed]
  123. Garred, P.; Honore, C.; Ma, Y.J.; Rorvig, S.; Cowland, J.; Borregaard, N.; Hummelshoj, T. The genetics of ficolins. J. Innate Immun. 2010, 2, 3–16. [Google Scholar] [CrossRef] [PubMed]
  124. Hummelshøj, T.; Thielens, N.M.; Madsen, H.O.; Arlaud, G.J.; Sim, R.B.; Garred, P. Molecular organization of human Ficolin-2. Mol. Immunol. 2007, 44, 401–411. [Google Scholar] [CrossRef] [PubMed]
  125. Matsushita, M.; Endo, Y.; Taira, S.; Sato, Y.; Fujita, T.; Ichikawa, N.; Nakata, M.; Mizuochi, T. A novel human serum lectin with collagen- and fibrinogen-like domains that functions as an opsonin. J. Biol. Chem. 1996, 271, 2448–2454. [Google Scholar] [CrossRef] [PubMed]
  126. Lacroix, M.; Dumestre-Perard, C.; Schoehn, G.; Houen, G.; Cesbron, J.Y.; Arlaud, G.J.; Thielens, N.M. Residue Lys57 in the collagen-like region of human L-ficolin and its counterpart Lys47 in H-ficolin play a key role in the interaction with the mannan-binding lectin-associated serine proteases and the collectin receptor calreticulin. J. Immunol. 2009, 182, 456–465. [Google Scholar] [CrossRef] [PubMed]
  127. Weis, W.I.; Drickamer, K. Trimeric structure of a C-type mannose-binding protein. Structure 1994, 2, 1227–1240. [Google Scholar] [CrossRef] [PubMed]
  128. Ohashi, T.; Erickson, H.P. The disulfide bonding pattern in ficolin multimers. J. Biol. Chem. 2004, 279, 6534–6539. [Google Scholar] [CrossRef] [PubMed]
  129. Akaiwa, M.; Yae, Y.; Sugimoto, R.; Suzuki, S.O.; Iwaki, T.; Izuhara, K.; Hamasaki, N. Hakata antigen, a new member of the ficolin/opsonin p35 family, is a novel human lectin secreted into bronchus/alveolus and bile. J. Histochem. Cytochem. 1999, 47, 777–786. [Google Scholar] [CrossRef] [PubMed]
  130. Liu, Y.; Endo, Y.; Iwaki, D.; Nakata, M.; Matsushita, M.; Wada, I.; Inoue, K.; Munakata, M.; Fujita, T. Human M-ficolin is a secretory protein that activates the lectin complement pathway. J. Immunol. 2005, 175, 3150–3156. [Google Scholar] [CrossRef] [PubMed]
  131. Schlapbach, L.J.; Aebi, C.; Hansen, A.G.; Hirt, A.; Jensenius, J.C.; Ammann, R.A. H-ficolin serum concentration and susceptibility to fever and neutropenia in paediatric cancer patients. Clin. Exp. Immunol. 2009, 157, 83–89. [Google Scholar] [CrossRef] [PubMed]
  132. Kilpatrick, D.C.; Fujita, T.; Matsushita, M. P35, an opsonic lectin of the ficolin family, in human blood from neonates, normal adults, and recurrent miscarriage patients. Immunol. Lett. 1999, 67, 109–112. [Google Scholar] [CrossRef] [PubMed]
  133. Munthe-Fog, L.; Hummelshoj, T.; Hansen, B.E.; Koch, C.; Madsen, H.O.; Skjodt, K.; Garred, P. The impact of FCN2 polymorphisms and haplotypes on the Ficolin-2 serum levels. Scand. J. Immunol. 2007, 65, 383–392. [Google Scholar] [CrossRef] [PubMed]
  134. Wittenborn, T.; Thiel, S.; Jensen, L.; Nielsen, H.J.; Jensenius, J.C. Characteristics and biological variations of M-ficolin, a pattern recognition molecule, in plasma. J. Innate Immun. 2010, 2, 167–180. [Google Scholar] [CrossRef] [PubMed]
  135. Garlatti, V.; Belloy, N.; Martin, L.; Lacroix, M.; Matsushita, M.; Endo, Y.; Fujita, T.; Fontecilla-Camps, J.C.; Arlaud, G.J.; Thielens, N.M.; et al. Structural insights into the innate immune recognition specificities of L- and H-ficolins. EMBO J. 2007, 26, 623–633. [Google Scholar] [CrossRef] [PubMed]
  136. Vassal-Stermann, E.; Lacroix, M.; Gout, E.; Laffly, E.; Pedersen, C.M.; Martin, L.; Amoroso, A.; Schmidt, R.R.; Zahringer, U.; Gaboriaud, C.; et al. Human L-Ficolin Recognizes Phosphocholine Moieties of Pneumococcal Teichoic Acid. J. Immunol. 2014. [Google Scholar] [CrossRef]
  137. Luo, F.; Sun, X.; Wang, Y.; Wang, Q.; Wu, Y.; Pan, Q.; Fang, C.; Zhang, X.L. Ficolin-2 Defends against Virulent Mycobacteria Tuberculosis Infection in Vivo, and Its Insufficiency Is Associated with Infection in Humans. PLoS One 2013, 8, e73859. [Google Scholar] [CrossRef] [PubMed]
  138. Jensen, M.L.; Honore, C.; Hummelshoj, T.; Hansen, B.E.; Madsen, H.O.; Garred, P. Ficolin-2 recognizes DNA and participates in the clearance of dying host cells. Mol. Immunol. 2007, 44, 856–865. [Google Scholar] [CrossRef] [PubMed]
  139. Hamed, M.R.; Brown, R.J.; Zothner, C.; Urbanowicz, R.A.; Mason, C.P.; Krarup, A.; McClure, C.P.; Irving, W.L.; Ball, J.K.; Harris, M.; et al. Recombinant human L-ficolin directly neutralizes hepatitis C virus entry. J. Innate Immun. 2014, 6, 676–684. [Google Scholar] [CrossRef] [PubMed]
  140. Keirstead, N.D.; Lee, C.; Yoo, D.; Brooks, A.S.; Hayes, M.A. Porcine plasma ficolin binds and reduces infectivity of porcine reproductive and respiratory syndrome virus (PRRSV) in vitro. Antivir. Res. 2008, 77, 28–38. [Google Scholar] [CrossRef] [PubMed]
  141. Liu, J.; Ali, M.A.; Shi, Y.; Zhao, Y.; Luo, F.; Yu, J.; Xiang, T.; Tang, J.; Li, D.; Hu, Q.; et al. Specifically binding of L-ficolin to N-glycans of HCV envelope glycoproteins E1 and E2 leads to complement activation. Cell. Mol. Immunol. 2009, 6, 235–244. [Google Scholar] [CrossRef] [PubMed]
  142. Pan, Q.; Chen, H.; Wang, F.; Jeza, V.T.; Hou, W.; Zhao, Y.; Xiang, T.; Zhu, Y.; Endo, Y.; Fujita, T.; et al. L-ficolin binds to the glycoproteins hemagglutinin and neuraminidase and inhibits influenza A virus infection both in vitro and in vivo. J. Innate Immun. 2012, 4, 312–324. [Google Scholar] [CrossRef] [PubMed]
  143. Zhao, Y.; Ren, Y.; Zhang, X.; Zhao, P.; Tao, W.; Zhong, J.; Li, Q.; Zhang, X.L. Ficolin-2 inhibits hepatitis C virus infection, whereas apolipoprotein E3 mediates viral immune escape. J. Immunol. 2014, 193, 783–796. [Google Scholar] [CrossRef] [PubMed]
  144. Verma, A.; White, M.; Vathipadiekal, V.; Tripathi, S.; Mbianda, J.; Ieong, M.; Qi, L.; Taubenberger, J.K.; Takahashi, K.; Jensenius, J.C.; et al. Human H-ficolin inhibits replication of seasonal and pandemic influenza A viruses. J. Immunol. 2012, 189, 2478–2487. [Google Scholar] [CrossRef] [PubMed]
  145. Ren, Y.; Ding, Q.; Zhang, X. Ficolins and infectious diseases. Virol. Sin. 2014, 29, 25–32. [Google Scholar] [CrossRef] [PubMed]
  146. Honore, C.; Rorvig, S.; Hummelshoj, T.; Skjoedt, M.O.; Borregaard, N.; Garred, P. Tethering of Ficolin-1 to cell surfaces through recognition of sialic acid by the fibrinogen-like domain. J. Leukoc. Biol. 2010, 88, 145–158. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, J.; Yang, L.; Ang, Z.; Yoong, S.L.; Tran, T.T.; Anand, G.S.; Tan, N.S.; Ho, B.; Ding, J.L. Secreted M-ficolin anchors onto monocyte transmembrane G protein-coupled receptor 43 and cross talks with plasma C-reactive protein to mediate immune signaling and regulate host defense. J. Immunol. 2010, 185, 6899–6910. [Google Scholar] [CrossRef] [PubMed]
  148. Kjaer, T.R.; Hansen, A.G.; Sorensen, U.B.; Nielsen, O.; Thiel, S.; Jensenius, J.C. Investigations on the pattern recognition molecule M-ficolin: Quantitative aspects of bacterial binding and leukocyte association. J. Leukoc. Biol. 2011, 90, 425–437. [Google Scholar] [CrossRef] [PubMed]
  149. Hummelshøj, T.; Munthe-Fog, L.; Madsen, H.O.; Fujita, T.; Matsushita, M.; Garred, P. Polymorphisms in the FCN2 gene determine serum variation and function of Ficolin-2. Hum. Mol. Genet. 2005, 14, 1651–1658. [Google Scholar] [CrossRef] [PubMed]
  150. Herpers, B.L.; Immink, M.M.; de Jong, B.A.; van Velzen-Blad, H.; de Jongh, B.M.; van Hannen, E.J. Coding and non-coding polymorphisms in the lectin pathway activator L-ficolin gene in 188 Dutch blood bank donors. Mol. Immunol. 2006, 43, 851–855. [Google Scholar] [CrossRef] [PubMed]
  151. Garred, P.; Honore, C.; Ma, Y.J.; Munthe-Fog, L.; Hummelshoj, T. MBL2, FCN1, FCN2 and FCN3-The genes behind the initiation of the lectin pathway of complement. Mol. Immunol. 2009, 46, 2737–2744. [Google Scholar] [CrossRef] [PubMed]
  152. Assaf, A.; Hoang, T.V.; Faik, I.; Aebischer, T.; Kremsner, P.G.; Kun, J.F.; Velavan, T.P. Genetic evidence of functional ficolin-2 haplotype as susceptibility factor in cutaneous leishmaniasis. PLoS One 2012, 7, e34113. [Google Scholar] [CrossRef] [PubMed]
  153. De Messias-Reason, I.; Kremsner, P.G.; Kun, J.F. Functional haplotypes that produce normal ficolin-2 levels protect against clinical leprosy. J. Infect. Dis. 2009, 199, 801–804. [Google Scholar] [CrossRef] [PubMed]
  154. Chalmers, J.D.; Kilpatrick, D.C.; McHugh, B.J.; Smith, M.P.; Govan, J.R.W.; Doherty, C.; Matsushita, M.; Hart, S.P.; Sethi, T.; Hill, A.T. T2 Single nucleotide polymorphisms in the ficolin-2 gene predispose to Pseudomonas aeruginosa infection and disease severity in non-cystic fibrosis bronchiectasis. Thorax 2011, 66, A1–A2. [Google Scholar] [CrossRef] [PubMed]
  155. Messias-Reason, I.J.; Schafranski, M.D.; Kremsner, P.G.; Kun, J.F. Ficolin 2 (FCN2) functional polymorphisms and the risk of rheumatic fever and rheumatic heart disease. Clin. Exp. Immunol. 2009, 157, 395–399. [Google Scholar] [CrossRef] [PubMed]
  156. Chapman, S.J.; Vannberg, F.O.; Khor, C.C.; Segal, S.; Moore, C.E.; Knox, K.; Day, N.P.; Davies, R.J.; Crook, D.W.; Hill, A.V. Functional polymorphisms in the FCN2 gene are not associated with invasive pneumococcal disease. Mol. Immunol. 2007, 44, 3267–3270. [Google Scholar] [CrossRef] [PubMed]
  157. Ruskamp, J.M.; Hoekstra, M.O.; Postma, D.S.; Kerkhof, M.; Bottema, R.W.; Koppelman, G.H.; Rovers, M.M.; Wijga, A.H.; de Jongste, J.C.; Brunekreef, B.; et al. Exploring the role of polymorphisms in ficolin genes in respiratory tract infections in children. Clin. Exp. Immunol. 2009, 155, 433–440. [Google Scholar] [CrossRef] [PubMed]
  158. Boldt, A.B.; Sanchez, M.I.; Stahlke, E.R.; Steffensen, R.; Thiel, S.; Jensenius, J.C.; Prevedello, F.C.; Mira, M.T.; Kun, J.F.; Messias-Reason, I.J. Susceptibility to leprosy is associated with M-ficolin polymorphisms. J. Clin. Immunol. 2013, 33, 210–219. [Google Scholar] [CrossRef] [PubMed]
  159. Vander Cruyssen, B.; Nuytinck, L.; Boullart, L.; Elewaut, D.; Waegeman, W.; van Thielen, M.; de Meester, E.; Lebeer, K.; Rossau, R.; de Keyser, F. Polymorphisms in the ficolin 1 gene (FCN1) are associated with susceptibility to the development of rheumatoid arthritis. Rheumatology (Oxford) 2007, 46, 1792–1795. [Google Scholar] [CrossRef]
  160. Urban, T.J.; Thompson, A.J.; Bradrick, S.S.; Fellay, J.; Schuppan, D.; Cronin, K.D.; Hong, L.; McKenzie, A.; Patel, K.; Shianna, K.V.; et al. IL28B genotype is associated with differential expression of intrahepatic interferon-stimulated genes in patients with chronic hepatitis C. Hepatology 2010, 52, 1888–1896. [Google Scholar] [CrossRef] [PubMed]
  161. McCarthy, J.J.; Li, J.H.; Thompson, A.; Suchindran, S.; Lao, X.Q.; Patel, K.; Tillmann, H.L.; Muir, A.J.; McHutchison, J.G. Replicated association between an IL28B gene variant and a sustained response to pegylated interferon and ribavirin. Gastroenterology 2010, 138, 2307–2314. [Google Scholar] [CrossRef] [PubMed]
  162. Hu, Y.L.; Luo, F.L.; Fu, J.L.; Chen, T.L.; Wu, S.M.; Zhou, Y.D.; Zhang, X.L. Early increased ficolin-2 concentrations are associated with severity of liver inflammation and efficacy of anti-viral therapy in chronic hepatitis C patients. Scand. J. Immunol. 2013, 77, 144–150. [Google Scholar] [CrossRef] [PubMed]
  163. Cedzynski, M.; Atkinson, A.P.; St Swierzko, A.; MacDonald, S.L.; Szala, A.; Zeman, K.; Buczylko, K.; Bak-Romaniszyn, L.; Wiszniewska, M.; Matsushita, M.; et al. L-ficolin (ficolin-2) insufficiency is associated with combined allergic and infectious respiratory disease in children. Mol. Immunol. 2009, 47, 415–419. [Google Scholar] [CrossRef] [PubMed]
  164. Faik, I.; Oyedeji, S.I.; Idris, Z.; de Messias-Reason, I.J.; Lell, B.; Kremsner, P.G.; Kun, J.F. Ficolin-2 levels and genetic polymorphisms of FCN2 in malaria. Hum. Immunol. 2011, 72, 74–79. [Google Scholar] [CrossRef] [PubMed]
  165. Ma, Y.J.; Doni, A.; Skjoedt, M.O.; Honore, C.; Arendrup, M.; Mantovani, A.; Garred, P. Heterocomplexes of mannose-binding lectin and the pentraxins PTX3 or serum amyloid P component trigger cross-activation of the complement system. J. Biol. Chem. 2011, 286, 3405–3417. [Google Scholar] [CrossRef] [PubMed]
  166. Panda, S.; Zhang, J.; Tan, N.S.; Ho, B.; Ding, J.L. Natural IgG antibodies provide innate protection against ficolin-opsonized bacteria. EMBO J. 2013, 32, 2905–2919. [Google Scholar] [CrossRef] [PubMed]
  167. Panda, S.; Zhang, J.; Yang, L.; Anand, G.S.; Ding, J.L. Molecular interaction between natural IgG and ficolin—Mechanistic insights on adaptive-innate immune crosstalk. Sci. Rep. 2014, 4, 3675. [Google Scholar] [CrossRef] [PubMed]
  168. Gout, E.; Moriscot, C.; Doni, A.; Dumestre-Perard, C.; Lacroix, M.; Perard, J.; Schoehn, G.; Mantovani, A.; Arlaud, G.J.; Thielens, N.M. M-ficolin interacts with the long pentraxin PTX3: A novel case of cross-talk between soluble pattern-recognition molecules. J. Immunol. 2011, 186, 5815–5822. [Google Scholar] [CrossRef] [PubMed]
  169. Ma, Y.J.; Doni, A.; Romani, L.; Jurgensen, H.J.; Behrendt, N.; Mantovani, A.; Garred, P. Ficolin-1-PTX3 complex formation promotes clearance of altered self-cells and modulates IL-8 production. J. Immunol. 2013, 191, 1324–1333. [Google Scholar] [CrossRef] [PubMed]
  170. Zhang, J.; Koh, J.; Lu, J.; Thiel, S.; Leong, B.S.; Sethi, S.; He, C.Y.; Ho, B.; Ding, J.L. Local inflammation induces complement crosstalk which amplifies the antimicrobial response. PLoS Pathog. 2009, 5, e1000282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Zhang, J.; Yang, L.; Anand, G.S.; Ho, B.; Ding, J.L. Pathophysiological condition changes the conformation of a flexible FBG-related protein, switching it from pathogen-recognition to host-interaction. Biochimie 2011, 93, 1710–1719. [Google Scholar] [CrossRef] [PubMed]
  172. Geijtenbeek, T.B.; Kwon, D.S.; Torensma, R.; van Vliet, S.J.; van Duijnhoven, G.C.; Middel, J.; Cornelissen, I.L.; Nottet, H.S.; KewalRamani, V.N.; Littman, D.R.; et al. DC-SIGN, a dendritic cell-specific HIV-1-binding protein that enhances trans-infection of T cells. Cell 2000, 100, 587–597. [Google Scholar] [CrossRef] [PubMed]
  173. Ganguly, D.; Haak, S.; Sisirak, V.; Reizis, B. The role of dendritic cells in autoimmunity. Nat. Rev. Immunol. 2013, 13, 566–577. [Google Scholar] [CrossRef] [PubMed]
  174. Geijtenbeek, T.B.; Torensma, R.; van Vliet, S.J.; van Duijnhoven, G.C.; Adema, G.J.; van Kooyk, Y.; Figdor, C.G. Identification of DC-SIGN, a novel dendritic cell-specific ICAM-3 receptor that supports primary immune responses. Cell 2000, 100, 575–585. [Google Scholar] [CrossRef] [PubMed]
  175. Mitchell, D.A.; Fadden, A.J.; Drickamer, K. A novel mechanism of carbohydrate recognition by the C-type lectins DC-SIGN and DC-SIGNR. Subunit organization and binding to multivalent ligands. J. Biol. Chem. 2001, 276, 28939–28945. [Google Scholar] [CrossRef] [PubMed]
  176. Manzo, C.; Torreno-Pina, J.A.; Joosten, B.; Reinieren-Beeren, I.; Gualda, E.J.; Loza-Alvarez, P.; Figdor, C.G.; Garcia-Parajo, M.F.; Cambi, A. The neck region of the C-type lectin DC-SIGN regulates its surface spatiotemporal organization and virus-binding capacity on antigen-presenting cells. J. Biol. Chem. 2012, 287, 38946–38955. [Google Scholar] [CrossRef] [PubMed]
  177. Kwon, D.S.; Gregorio, G.; Bitton, N.; Hendrickson, W.A.; Littman, D.R. DC-SIGN-mediated internalization of HIV is required for trans-enhancement of T cell infection. Immunity 2002, 16, 135–144. [Google Scholar] [CrossRef]
  178. Soilleux, E.J.; Barten, R.; Trowsdale, J. DC-SIGN; a related gene, DC-SIGNR; and CD23 form a cluster on 19p13. J. Immunol. 2000, 165, 2937–2942. [Google Scholar] [CrossRef] [PubMed]
  179. Baribaud, F.; Pohlmann, S.; Leslie, G.; Mortari, F.; Doms, R.W. Quantitative expression and virus transmission analysis of DC-SIGN on monocyte-derived dendritic cells. J. Virol. 2002, 76, 9135–9142. [Google Scholar] [CrossRef] [PubMed]
  180. Pӧhlmann, S.; Baribaud, F.; Lee, B.; Leslie, G.J.; Sanchez, M.D.; Hiebenthal-Millow, K.; Munch, J.; Kirchhoff, F.; Doms, R.W. DC-SIGN interactions with human immunodeficiency virus type 1 and 2 and simian immunodeficiency virus. J. Virol. 2001, 75, 4664–4672. [Google Scholar] [CrossRef] [PubMed]
  181. Lin, G.; Simmons, G.; Pohlmann, S.; Baribaud, F.; Ni, H.; Leslie, G.J.; Haggarty, B.S.; Bates, P.; Weissman, D.; Hoxie, J.A.; et al. Differential N-linked glycosylation of human immunodeficiency virus and Ebola virus envelope glycoproteins modulates interactions with DC-SIGN and DC-SIGNR. J. Virol. 2003, 77, 1337–1346. [Google Scholar] [CrossRef] [PubMed]
  182. Haase, A.T. Targeting early infection to prevent HIV-1 mucosal transmission. Nature 2010, 464, 217–223. [Google Scholar] [CrossRef] [PubMed]
  183. Xiao, X.; Kinter, A.; Broder, C.C.; Dimitrov, D.S. Interactions of CCR5 and CXCR4 with CD4 and gp120 in human blood monocyte-derived dendritic cells. Exp. Mol. Pathol. 2000, 68, 133–138. [Google Scholar] [CrossRef] [PubMed]
  184. Moris, A.; Nobile, C.; Buseyne, F.; Porrot, F.; Abastado, J.P.; Schwartz, O. DC-SIGN promotes exogenous MHC-I-restricted HIV-1 antigen presentation. Blood 2004, 103, 2648–2654. [Google Scholar] [CrossRef] [PubMed]
  185. Smith, A.L.; Ganesh, L.; Leung, K.; Jongstra-Bilen, J.; Jongstra, J.; Nabel, G.J. Leukocyte-specific protein 1 interacts with DC-SIGN and mediates transport of HIV to the proteasome in dendritic cells. J. Exp. Med. 2007, 204, 421–430. [Google Scholar] [CrossRef] [PubMed]
  186. Gringhuis, S.I.; van der Vlist, M.; van den Berg, L.M.; den Dunnen, J.; Litjens, M.; Geijtenbeek, T.B. HIV-1 exploits innate signaling by TLR8 and DC-SIGN for productive infection of dendritic cells. Nat. Immunol. 2010, 11, 419–426. [Google Scholar] [CrossRef] [PubMed]
  187. Hijazi, K.; Wang, Y.; Scala, C.; Jeffs, S.; Longstaff, C.; Stieh, D.; Haggarty, B.; Vanham, G.; Schols, D.; Balzarini, J.; et al. DC-SIGN increases the affinity of HIV-1 envelope glycoprotein interaction with CD4. PLoS One 2011, 6, e28307. [Google Scholar] [CrossRef] [PubMed]
  188. Lee, B.; Leslie, G.; Soilleux, E.; O’Doherty, U.; Baik, S.; Levroney, E.; Flummerfelt, K.; Swiggard, W.; Coleman, N.; Malim, M.; et al. cis Expression of DC-SIGN allows for more efficient entry of human and simian immunodeficiency viruses via CD4 and a coreceptor. J. Virol. 2001, 75, 12028–12038. [Google Scholar] [CrossRef] [PubMed]
  189. Miyauchi, K.; Kim, Y.; Latinovic, O.; Morozov, V.; Melikyan, G.B. HIV enters cells via endocytosis and dynamin-dependent fusion with endosomes. Cell 2009, 137, 433–444. [Google Scholar] [CrossRef] [PubMed]
  190. Van Montfort, T.; Eggink, D.; Boot, M.; Tuen, M.; Hioe, C.E.; Berkhout, B.; Sanders, R.W. HIV-1 N-glycan composition governs a balance between dendritic cell-mediated viral transmission and antigen presentation. J. Immunol. 2011, 187, 4676–4685. [Google Scholar] [CrossRef] [PubMed]
  191. McDonald, D.; Wu, L.; Bohks, S.M.; KewalRamani, V.N.; Unutmaz, D.; Hope, T.J. Recruitment of HIV and its receptors to dendritic cell-T cell junctions. Science 2003, 300, 1295–1297. [Google Scholar] [CrossRef] [PubMed]
  192. Sol-Foulon, N.; Moris, A.; Nobile, C.; Boccaccio, C.; Engering, A.; Abastado, J.P.; Heard, J.M.; van Kooyk, Y.; Schwartz, O. HIV-1 Nef-induced upregulation of DC-SIGN in dendritic cells promotes lymphocyte clustering and viral spread. Immunity 2002, 16, 145–155. [Google Scholar] [CrossRef] [PubMed]
  193. Cavrois, M.; Neidleman, J.; Kreisberg, J.F.; Greene, W.C. In vitro derived dendritic cells trans-infect CD4 T cells primarily with surface-bound HIV-1 virions. PLoS Pathog. 2007, 3, e4. [Google Scholar] [CrossRef] [PubMed]
  194. Hodges, A.; Sharrocks, K.; Edelmann, M.; Baban, D.; Moris, A.; Schwartz, O.; Drakesmith, H.; Davies, K.; Kessler, B.; McMichael, A.; et al. Activation of the lectin DC-SIGN induces an immature dendritic cell phenotype triggering Rho-GTPase activity required for HIV-1 replication. Nat. Immunol. 2007, 8, 569–577. [Google Scholar] [CrossRef] [PubMed]
  195. Izquierdo-Useros, N.; Lorizate, M.; Puertas, M.C.; Rodriguez-Plata, M.T.; Zangger, N.; Erikson, E.; Pino, M.; Erkizia, I.; Glass, B.; Clotet, B.; et al. Siglec-1 is a novel dendritic cell receptor that mediates HIV-1 trans-infection through recognition of viral membrane gangliosides. PLoS Biol. 2012, 10, e1001448. [Google Scholar] [CrossRef] [PubMed]
  196. Chen, Y.; Hwang, S.L.; Chan, V.S.; Chung, N.P.; Wang, S.R.; Li, Z.; Ma, J.; Lin, C.W.; Hsieh, Y.J.; Chang, K.P.; et al. Binding of HIV-1 gp120 to DC-SIGN promotes ASK-1-dependent activation-induced apoptosis of human dendritic cells. PLoS Pathog. 2013, 9, e1003100. [Google Scholar] [CrossRef] [PubMed]
  197. De Witte, L.; de Vries, R.D.; van der Vlist, M.; Yuksel, S.; Litjens, M.; de Swart, R.L.; Geijtenbeek, T.B. DC-SIGN and CD150 have distinct roles in transmission of measles virus from dendritic cells to T-lymphocytes. PLoS Pathog. 2008, 4, e1000049. [Google Scholar] [CrossRef] [PubMed]
  198. Alvarez, C.P.; Lasala, F.; Carrillo, J.; Muniz, O.; Corbi, A.L.; Delgado, R. C-type lectins DC-SIGN and L-SIGN mediate cellular entry by Ebola virus in cis and in trans. J. Virol. 2002, 76, 6841–6844. [Google Scholar] [CrossRef] [PubMed]
  199. Yang, Z.Y.; Huang, Y.; Ganesh, L.; Leung, K.; Kong, W.P.; Schwartz, O.; Subbarao, K.; Nabel, G.J. pH-dependent entry of severe acute respiratory syndrome coronavirus is mediated by the spike glycoprotein and enhanced by dendritic cell transfer through DC-SIGN. J. Virol. 2004, 78, 5642–5650. [Google Scholar] [CrossRef] [PubMed]
  200. Wang, S.F.; Huang, J.C.; Lee, Y.M.; Liu, S.J.; Chan, Y.J.; Chau, Y.P.; Chong, P.; Chen, Y.M. DC-SIGN mediates avian H5N1 influenza virus infection in cis and in trans. Biochem. Biophys. Res. Commun. 2008, 373, 561–566. [Google Scholar] [CrossRef] [PubMed]
  201. Pӧhlmann, S.; Zhang, J.; Baribaud, F.; Chen, Z.; Leslie, G.J.; Lin, G.; Granelli-Piperno, A.; Doms, R.W.; Rice, C.M.; McKeating, J.A. Hepatitis C virus glycoproteins interact with DC-SIGN and DC-SIGNR. J. Virol. 2003, 77, 4070–4080. [Google Scholar] [CrossRef] [PubMed]
  202. Halary, F.; Amara, A.; Lortat-Jacob, H.; Messerle, M.; Delaunay, T.; Houles, C.; Fieschi, F.; Arenzana-Seisdedos, F.; Moreau, J.F.; Dechanet-Merville, J. Human cytomegalovirus binding to DC-SIGN is required for dendritic cell infection and target cell trans-infection. Immunity 2002, 17, 653–664. [Google Scholar] [CrossRef] [PubMed]
  203. Rappocciolo, G.; Hensler, H.R.; Jais, M.; Reinhart, T.A.; Pegu, A.; Jenkins, F.J.; Rinaldo, C.R. Human herpesvirus 8 infects and replicates in primary cultures of activated B lymphocytes through DC-SIGN. J. Virol. 2008, 82, 4793–4806. [Google Scholar] [CrossRef] [PubMed]
  204. Ludwig, I.S.; Lekkerkerker, A.N.; Depla, E.; Bosman, F.; Musters, R.J.; Depraetere, S.; van Kooyk, Y.; Geijtenbeek, T.B. Hepatitis C virus targets DC-SIGN and L-SIGN to escape lysosomal degradation. J. Virol. 2004, 78, 8322–8332. [Google Scholar] [CrossRef] [PubMed]
  205. Pӧhlmann, S.; Soilleux, E.J.; Baribaud, F.; Leslie, G.J.; Morris, L.S.; Trowsdale, J.; Lee, B.; Coleman, N.; Doms, R.W. DC-SIGNR, a DC-SIGN homologue expressed in endothelial cells, binds to human and simian immunodeficiency viruses and activates infection in trans. Proc. Natl. Acad. Sci. USA 2001, 98, 2670–2675. [Google Scholar] [CrossRef] [PubMed]
  206. Marzi, A.; Gramberg, T.; Simmons, G.; Moller, P.; Rennekamp, A.J.; Krumbiegel, M.; Geier, M.; Eisemann, J.; Turza, N.; Saunier, B.; et al. DC-SIGN and DC-SIGNR interact with the glycoprotein of Marburg virus and the S protein of severe acute respiratory syndrome coronavirus. J. Virol. 2004, 78, 12090–12095. [Google Scholar] [CrossRef] [PubMed]
  207. Barreiro, L.B.; Patin, E.; Neyrolles, O.; Cann, H.M.; Gicquel, B.; Quintana-Murci, L. The heritage of pathogen pressures and ancient demography in the human innate-immunity CD209/CD209L region. Am. J. Hum. Genet. 2005, 77, 869–886. [Google Scholar] [CrossRef] [PubMed]
  208. Davis, C.W.; Nguyen, H.Y.; Hanna, S.L.; Sanchez, M.D.; Doms, R.W.; Pierson, T.C. West Nile virus discriminates between DC-SIGN and DC-SIGNR for cellular attachment and infection. J. Virol 2006, 80, 1290–1301. [Google Scholar] [CrossRef] [PubMed]
  209. Mummidi, S.; Catano, G.; Lam, L.; Hoefle, A.; Telles, V.; Begum, K.; Jimenez, F.; Ahuja, S.S.; Ahuja, S.K. Extensive repertoire of membrane-bound and soluble dendritic cell-specific ICAM-3-grabbing nonintegrin 1 (DC-SIGN1) and DC-SIGN2 isoforms. Inter-individual variation in expression of DC-SIGN transcripts. J. Biol. Chem. 2001, 276, 33196–33212. [Google Scholar] [CrossRef] [PubMed]
  210. Plazolles, N.; Humbert, J.M.; Vachot, L.; Verrier, B.; Hocke, C.; Halary, F. Pivotal advance: The promotion of soluble DC-SIGN release by inflammatory signals and its enhancement of cytomegalovirus-mediated cis-infection of myeloid dendritic cells. J. Leukoc. Biol. 2011, 89, 329–342. [Google Scholar] [CrossRef] [PubMed]
  211. Serrano-Gómez, D.; Sierra-Filardi, E.; Martinez-Nunez, R.T.; Caparros, E.; Delgado, R.; Munoz-Fernandez, M.A.; Abad, M.A.; Jimenez-Barbero, J.; Leal, M.; Corbi, A.L. Structural requirements for multimerization of the pathogen receptor dendritic cell-specific ICAM3-grabbing non-integrin (CD209) on the cell surface. J. Biol. Chem. 2008, 283, 3889–3903. [Google Scholar] [CrossRef] [PubMed]
  212. Alagarasu, K.; Selvaraj, P.; Swaminathan, S.; Raghavan, S.; Narendran, G.; Narayanan, P.R. CCR2, MCP-1, SDF-1a & DC-SIGN gene polymorphisms in HIV-1 infected patients with & without tuberculosis. Indian J. Med. Res. 2009, 130, 444–450. [Google Scholar] [PubMed]
  213. Zhang, J.; Zhang, X.; Fu, J.; Bi, Z.; Arheart, K.L.; Barreiro, L.B.; Quintana-Murci, L.; Pahwa, S.; Liu, H. Protective role of DC-SIGN (CD209) neck-region alleles with <5 repeat units in HIV-1 transmission. J. Infect. Dis. 2008, 198, 68–71. [Google Scholar]
  214. Liu, H.; Hwangbo, Y.; Holte, S.; Lee, J.; Wang, C.; Kaupp, N.; Zhu, H.; Celum, C.; Corey, L.; McElrath, M.J.; et al. Analysis of genetic polymorphisms in CCR5, CCR2, stromal cell-derived factor-1, RANTES, and dendritic cell-specific intercellular adhesion molecule-3-grabbing nonintegrin in seronegative individuals repeatedly exposed to HIV-1. J. Infect. Dis. 2004, 190, 1055–1058. [Google Scholar] [CrossRef] [PubMed]
  215. Nguyen, D.G.; Hildreth, J.E. Involvement of macrophage mannose receptor in the binding and transmission of HIV by macrophages. Eur. J. Immunol. 2003, 33, 483–493. [Google Scholar] [CrossRef] [PubMed]
  216. Trujillo, J.R.; Rogers, R.; Molina, R.M.; Dangond, F.; McLane, M.F.; Essex, M.; Brain, J.D. Noninfectious entry of HIV-1 into peripheral and brain macrophages mediated by the mannose receptor. Proc. Natl. Acad. Sci. USA 2007, 104, 5097–5102. [Google Scholar] [CrossRef] [PubMed]
  217. Hartshorn, K.L.; Sastry, K.N.; Chang, D.; White, M.R.; Crouch, E.C. Enhanced anti-influenza activity of a surfactant protein D and serum conglutinin fusion protein. Am. J. Physiol. Lung Cell. Mol. Physiol. 2000, 278, L90–L98. [Google Scholar] [PubMed]
  218. Yang, L.; Hu, S.; Chroneos, Z. Targeting of the surfactant protein A receptor SP-R210L variant by influenza A virus in macrophages. (INM8P.440). J. Immunol. 2014, 192, 124–126. [Google Scholar]
  219. Pandit, H.; Gopal, S.; Sonawani, A.; Yadav, A.K.; Qaseem, A.S.; Warke, H.; Patil, A.; Gajbhiye, R.; Kulkarni, V.; Al-Mozaini, M.A.; et al. Surfactant protein D inhibits HIV-1 infection of target cells via interference with gp120-CD4 interaction and modulates pro-inflammatory cytokine production. PLoS One 2014, 9, e102395. [Google Scholar] [CrossRef] [PubMed]
  220. Madsen, J.; Gaiha, G.D.; Palaniyar, N.; Dong, T.; Mitchell, D.A.; Clark, H.W. Surfactant Protein D modulates HIV infection of both T-cells and dendritic cells. PLoS One 2013, 8, e59047. [Google Scholar] [CrossRef] [PubMed]
  221. Gaiha, G.D.; Dong, T.; Palaniyar, N.; Mitchell, D.A.; Reid, K.B.; Clark, H.W. Surfactant protein A binds to HIV and inhibits direct infection of CD4+ cells, but enhances dendritic cell-mediated viral transfer. J. Immunol. 2008, 181, 601–609. [Google Scholar] [CrossRef] [PubMed]
  222. Sano, H.; Nagai, K.; Tsutsumi, H.; Kuroki, Y. Lactoferrin and surfactant protein A exhibit distinct binding specificity to F protein and differently modulate respiratory syncytial virus infection. Eur. J. Immunol. 2003, 33, 2894–2902. [Google Scholar] [CrossRef] [PubMed]
  223. St-Pierre, C.; Manya, H.; Ouellet, M.; Clark, G.F.; Endo, T.; Tremblay, M.J.; Sato, S. Host-soluble galectin-1 promotes HIV-1 replication through a direct interaction with glycans of viral gp120 and host CD4. J. Virol. 2011, 85, 11742–11751. [Google Scholar] [CrossRef] [PubMed]
  224. Bi, S.; Hong, P.W.; Lee, B.; Baum, L.G. Galectin-9 binding to cell surface protein disulfide isomerase regulates the redox environment to enhance T-cell migration and HIV entry. Proc. Natl. Acad. Sci. USA 2011, 108, 10650–10655. [Google Scholar] [CrossRef] [PubMed]
  225. Wang, S.F.; Tsao, C.H.; Lin, Y.T.; Hsu, D.K.; Chiang, M.L.; Lo, C.H.; Chien, F.C.; Chen, P.; Arthur Chen, Y.M.; Chen, H.Y.; et al. Galectin-3 promotes HIV-1 budding via association with Alix and Gag p6. Glycobiology 2014, 24, 1022–1035. [Google Scholar] [CrossRef] [PubMed]
  226. De Witte, L.; Nabatov, A.; Pion, M.; Fluitsma, D.; de Jong, M.A.; de Gruijl, T.; Piguet, V.; van Kooyk, Y.; Geijtenbeek, T.B. Langerin is a natural barrier to HIV-1 transmission by Langerhans cells. Nat. Med. 2007, 13, 367–371. [Google Scholar] [CrossRef] [PubMed]
  227. Van der Vlist, M.; de Witte, L.; de Vries, R.D.; Litjens, M.; de Jong, M.A.; Fluitsma, D.; de Swart, R.L.; Geijtenbeek, T.B. Human Langerhans cells capture measles virus through Langerin and present viral antigens to CD4(+) T cells but are incapable of cross-presentation. Eur. J. Immunol. 2011, 41, 2619–2631. [Google Scholar] [CrossRef] [PubMed]
  228. Valdimarsson, H. Infusion of plasma-derived mannan-binding lectin (MBL) into MBL-deficient humans. Biochem. Soc. Trans. 2003, 31, 768–769. [Google Scholar] [CrossRef] [PubMed]
  229. Laursen, I. Mannan-binding lectin (MBL) production from human plasma. Biochem. Soc. Trans. 2003, 31, 758–762. [Google Scholar] [CrossRef] [PubMed]
  230. Jensenius, J.C.; Jensen, P.H.; McGuire, K.; Larsen, J.L.; Thiel, S. Recombinant mannan-binding lectin (MBL) for therapy. Biochem. Soc. Trans. 2003, 31, 763–767. [Google Scholar] [CrossRef] [PubMed]
  231. Petersen, K.A.; Matthiesen, F.; Agger, T.; Kongerslev, L.; Thiel, S.; Cornelissen, K.; Axelsen, M. Phase I safety, tolerability, and pharmacokinetic study of recombinant human mannan-binding lectin. J. Clin. Immunol. 2006, 26, 465–475. [Google Scholar] [CrossRef] [PubMed]
  232. Michelow, I.C.; Dong, M.; Mungall, B.A.; Yantosca, L.M.; Lear, C.; Ji, X.; Karpel, M.; Rootes, C.L.; Brudner, M.; Houen, G.; et al. A novel L-ficolin/mannose-binding lectin chimeric molecule with enhanced activity against Ebola virus. J. Biol. Chem. 2010, 285, 24729–24739. [Google Scholar] [CrossRef] [PubMed]
  233. White, M.R.; Crouch, E.; Chang, D.; Sastry, K.; Guo, N.; Engelich, G.; Takahashi, K.; Ezekowitz, R.A.; Hartshorn, K.L. Enhanced antiviral and opsonic activity of a human mannose-binding lectin and surfactant protein D chimera. J. Immunol. 2000, 165, 2108–2115. [Google Scholar] [CrossRef] [PubMed]
  234. Takahashi, K.; Moyo, P.; Chigweshe, L.; Chang, W.C.; White, M.R.; Hartshorn, K.L. Efficacy of recombinant chimeric lectins, consisting of mannose binding lectin and L-ficolin, against influenza A viral infection in mouse model study. Virus Res. 2013, 178, 495–501. [Google Scholar] [CrossRef] [PubMed]
  235. Chang, W.C.; Hartshorn, K.L.; White, M.R.; Moyo, P.; Michelow, I.C.; Koziel, H.; Kinane, B.T.; Schmidt, E.V.; Fujita, T.; Takahashi, K. Recombinant chimeric lectins consisting of mannose-binding lectin and L-ficolin are potent inhibitors of influenza A virus compared with mannose-binding lectin. Biochem. Pharmacol. 2011, 81, 388–395. [Google Scholar] [CrossRef] [PubMed]
  236. Crouch, E.; Nikolaidis, N.; McCormack, F.X.; McDonald, B.; Allen, K.; Rynkiewicz, M.J.; Cafarella, T.M.; White, M.; Lewnard, K.; Leymarie, N.; et al. Mutagenesis of surfactant protein D informed by evolution and x-ray crystallography enhances defenses against influenza A virus in vivo. J. Biol. Chem. 2011, 286, 40681–40692. [Google Scholar] [CrossRef] [PubMed]
  237. Hillaire, M.L.; van Eijk, M.; Vogelzang-van Trierum, S.E.; Fouchier, R.A.; Osterhaus, A.D.; Haagsman, H.P.; Rimmelzwaan, G.F. Recombinant porcine surfactant protein D inhibits influenza A virus replication ex vivo. Virus Res. 2014, 181, 22–26. [Google Scholar] [CrossRef] [PubMed]
  238. Tarr, A.W.; Urbanowicz, R.A.; Ball, J.K. The role of humoral innate immunity in hepatitis C virus infection. Viruses 2012, 4, 1–27. [Google Scholar] [CrossRef] [PubMed]
  239. Marzi, A.; Mitchell, D.A.; Chaipan, C.; Fisch, T.; Doms, R.W.; Carrington, M.; Desrosiers, R.C.; Pohlmann, S. Modulation of HIV and SIV neutralization sensitivity by DC-SIGN and mannose-binding lectin. Virology 2007, 368, 322–330. [Google Scholar] [CrossRef] [PubMed]
  240. Boross, P.; Leusen, J.H. Boosting antibody therapy with complement. Blood 2012, 119, 5945–5947. [Google Scholar] [CrossRef] [PubMed]
  241. Alexandre, K.B.; Gray, E.S.; Mufhandu, H.; McMahon, J.B.; Chakauya, E.; O’Keefe, B.R.; Chikwamba, R.; Morris, L. The lectins griffithsin, cyanovirin-N and scytovirin inhibit HIV-1 binding to the DC-SIGN receptor and transfer to CD4(+) cells. Virology 2012, 423, 175–186. [Google Scholar] [CrossRef] [PubMed]
  242. Balzarini, J.; Francois, K.O.; van Laethem, K.; Hoorelbeke, B.; Renders, M.; Auwerx, J.; Liekens, S.; Oki, T.; Igarashi, Y.; Schols, D. Pradimicin S, a highly soluble nonpeptidic small-size carbohydrate-binding antibiotic, is an anti-HIV drug lead for both microbicidal and systemic use. Antimicrob. Agents Chemother. 2010, 54, 1425–1435. [Google Scholar] [CrossRef] [PubMed]
  243. Buffa, V.; Stieh, D.; Mamhood, N.; Hu, Q.; Fletcher, P.; Shattock, R.J. Cyanovirin-N potently inhibits human immunodeficiency virus type 1 infection in cellular and cervical explant models. J. Gen. Virol. 2009, 90, 234–243. [Google Scholar] [CrossRef] [PubMed]
  244. Tsai, C.C.; Emau, P.; Jiang, Y.; Agy, M.B.; Shattock, R.J.; Schmidt, A.; Morton, W.R.; Gustafson, K.R.; Boyd, M.R. Cyanovirin-N inhibits AIDS virus infections in vaginal transmission models. AIDS Res. Hum. Retrovir. 2004, 20, 11–18. [Google Scholar] [CrossRef] [PubMed]
  245. Whitley, M.J.; Furey, W.; Kollipara, S.; Gronenborn, A.M. Burkholderia oklahomensis agglutinin is a canonical two-domain OAA-family lectin: Structures, carbohydrate binding and anti-HIV activity. FEBS J. 2013, 280, 2056–2067. [Google Scholar] [CrossRef] [PubMed]
  246. Ferir, G.; Huskens, D.; Noppen, S.; Koharudin, L.M.; Gronenborn, A.M.; Schols, D. Broad anti-HIV activity of the Oscillatoria agardhii agglutinin homologue lectin family. J. Antimicrob. Chemother. 2014, 69, 2746–2758. [Google Scholar] [CrossRef] [PubMed]
  247. Liu, X.; Lagenaur, L.A.; Simpson, D.A.; Essenmacher, K.P.; Frazier-Parker, C.L.; Liu, Y.; Tsai, D.; Rao, S.S.; Hamer, D.H.; Parks, T.P.; et al. Engineered vaginal lactobacillus strain for mucosal delivery of the human immunodeficiency virus inhibitor cyanovirin-N. Antimicrob. Agents Chemother. 2006, 50, 3250–3259. [Google Scholar] [CrossRef] [PubMed]
  248. Koharudin, L.M.; Furey, W.; Gronenborn, A.M. Novel fold and carbohydrate specificity of the potent anti-HIV cyanobacterial lectin from Oscillatoria agardhii. J. Biol. Chem. 2011, 286, 1588–1597. [Google Scholar] [CrossRef] [PubMed]
  249. Meuleman, P.; Albecka, A.; Belouzard, S.; Vercauteren, K.; Verhoye, L.; Wychowski, C.; Leroux-Roels, G.; Palmer, K.E.; Dubuisson, J. Griffithsin has antiviral activity against hepatitis C virus. Antimicrob. Agents Chemother. 2011, 55, 5159–5167. [Google Scholar] [CrossRef] [PubMed]
  250. O’Keefe, B.R.; Giomarelli, B.; Barnard, D.L.; Shenoy, S.R.; Chan, P.K.; McMahon, J.B.; Palmer, K.E.; Barnett, B.W.; Meyerholz, D.K.; Wohlford-Lenane, C.L.; et al. Broad-spectrum in vitro activity and in vivo efficacy of the antiviral protein griffithsin against emerging viruses of the family Coronaviridae. J. Virol. 2010, 84, 2511–2521. [Google Scholar] [CrossRef] [PubMed]
  251. Ishag, H.Z.; Li, C.; Huang, L.; Sun, M.X.; Wang, F.; Ni, B.; Malik, T.; Chen, P.Y.; Mao, X. Griffithsin inhibits Japanese encephalitis virus infection in vitro and in vivo. Arch. Virol. 2013, 158, 349–358. [Google Scholar] [CrossRef] [PubMed]
  252. Takebe, Y.; Saucedo, C.J.; Lund, G.; Uenishi, R.; Hase, S.; Tsuchiura, T.; Kneteman, N.; Ramessar, K.; Tyrrell, D.L.; Shirakura, M.; et al. Antiviral lectins from red and blue-green algae show potent in vitro and in vivo activity against hepatitis C virus. PLoS One 2013, 8, e64449. [Google Scholar] [CrossRef] [PubMed]
  253. Helle, F.; Wychowski, C.; Vu-Dac, N.; Gustafson, K.R.; Voisset, C.; Dubuisson, J. Cyanovirin-N inhibits hepatitis C virus entry by binding to envelope protein glycans. J. Biol. Chem. 2006, 281, 25177–25183. [Google Scholar] [CrossRef] [PubMed]
  254. Dey, B.; Lerner, D.L.; Lusso, P.; Boyd, M.R.; Elder, J.H.; Berger, E.A. Multiple antiviral activities of cyanovirin-N: Blocking of human immunodeficiency virus type 1 gp120 interaction with CD4 and coreceptor and inhibition of diverse enveloped viruses. J. Virol. 2000, 74, 4562–4569. [Google Scholar] [CrossRef] [PubMed]
  255. Smee, D.F.; Bailey, K.W.; Wong, M.H.; O’Keefe, B.R.; Gustafson, K.R.; Mishin, V.P.; Gubareva, L.V. Treatment of influenza A (H1N1) virus infections in mice and ferrets with cyanovirin-N. Antivir. Res. 2008, 80, 266–271. [Google Scholar] [CrossRef] [PubMed]
  256. Bertaux, C.; Daelemans, D.; Meertens, L.; Cormier, E.G.; Reinus, J.F.; Peumans, W.J.; van Damme, E.J.; Igarashi, Y.; Oki, T.; Schols, D.; et al. Entry of hepatitis C virus and human immunodeficiency virus is selectively inhibited by carbohydrate-binding agents but not by polyanions. Virology 2007, 366, 40–50. [Google Scholar] [CrossRef] [PubMed]
  257. Boyd, M.R.; Gustafson, K.R.; McMahon, J.B.; Shoemaker, R.H.; O’Keefe, B.R.; Mori, T.; Gulakowski, R.J.; Wu, L.; Rivera, M.I.; Laurencot, C.M.; et al. Discovery of cyanovirin-N, a novel human immunodeficiency virus-inactivating protein that binds viral surface envelope glycoprotein gp120: Potential applications to microbicide development. Antimicrob. Agents Chemother. 1997, 41, 1521–1530. [Google Scholar] [PubMed]
  258. Chen, J.; Huang, D.; Chen, W.; Guo, C.; Wei, B.; Wu, C.; Peng, Z.; Fan, J.; Hou, Z.; Fang, Y.; et al. Linker-extended native cyanovirin-N facilitates PEGylation and potently inhibits HIV-1 by targeting the glycan ligand. PLoS One 2014, 9, e86455. [Google Scholar] [CrossRef] [PubMed]
  259. Barton, C.; Kouokam, J.C.; Lasnik, A.B.; Foreman, O.; Cambon, A.; Brock, G.; Montefiori, D.C.; Vojdani, F.; McCormick, A.A.; O’Keefe, B.R.; et al. Activity of and effect of subcutaneous treatment with the broad-spectrum antiviral lectin griffithsin in two laboratory rodent models. Antimicrob. Agents Chemother. 2014, 58, 120–127. [Google Scholar] [CrossRef] [PubMed]
  260. Xiong, S.; Fan, J.; Kitazato, K. The antiviral protein cyanovirin-N: The current state of its production and applications. Appl. Microbiol. Biotechnol. 2010, 86, 805–812. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Mason, C.P.; Tarr, A.W. Human Lectins and Their Roles in Viral Infections. Molecules 2015, 20, 2229-2271. https://doi.org/10.3390/molecules20022229

AMA Style

Mason CP, Tarr AW. Human Lectins and Their Roles in Viral Infections. Molecules. 2015; 20(2):2229-2271. https://doi.org/10.3390/molecules20022229

Chicago/Turabian Style

Mason, Christopher P., and Alexander W. Tarr. 2015. "Human Lectins and Their Roles in Viral Infections" Molecules 20, no. 2: 2229-2271. https://doi.org/10.3390/molecules20022229

Article Metrics

Back to TopTop