Next Article in Journal
Adrenal Insufficiency following Stereotactic Ablative Radiotherapy (SAbR) of Adrenal Gland Metastases
Next Article in Special Issue
Enhancing the Efficacy of Breast Cancer Immunotherapy Using a Smac-Armed Oncolytic Virus
Previous Article in Journal
Role of Percutaneous Endoscopic Gastrostomy for the Nutrition of Head and Neck Cancer Patients before and up to 6 Months after Cancer Treatment
Previous Article in Special Issue
Intratumoral Delivery of Interleukin 9 via Oncolytic Vaccinia Virus Elicits Potent Antitumor Effects in Tumor Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Receptors and Host Factors for Enterovirus Infection: Implications for Cancer Therapy

by
Olga N. Alekseeva
1,†,
Le T. Hoa
2,†,
Pavel O. Vorobyev
1,
Dmitriy V. Kochetkov
1,
Yana D. Gumennaya
1,
Elizaveta R. Naberezhnaya
1,
Denis O. Chuvashov
1,
Alexander V. Ivanov
1,*,
Peter M. Chumakov
1 and
Anastasia V. Lipatova
1,*
1
Engelhardt Institute of Molecular Biology, Russian Academy of Sciences, 119991 Moscow, Russia
2
Department of Molecular Microbiology and Immunology, Norris Comprehensive Cancer Center, Keck School of Medicine, University of Southern California, Los Angeles, CA 90033, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2024, 16(18), 3139; https://doi.org/10.3390/cancers16183139
Submission received: 8 May 2024 / Revised: 29 August 2024 / Accepted: 6 September 2024 / Published: 12 September 2024
(This article belongs to the Special Issue Oncolytic Viruses: A Key Step toward Cancer Immunotherapy)

Abstract

:

Simple Summary

Enteroviruses are considered to be promising oncolytic agents towards a variety of human cancers. Enteroviruses comprise polioviruses, coxsackieviruses, and echoviruses. Their efficacy depends on their ability to infect respective tumor cells. As non-enveloped viruses, they can enter cells via different receptors, many of which are expressed in tumors at higher levels. However, understanding the precise role of individual receptors in virus entry is complicated, and often requires the development of knockdown/knockout models. In this review we summarize current concepts about the roles of individual receptors in the entry of enteroviruses into cells, as well as the impact of host RNA-sensing machinery that can activate interferon signaling in response to viral infections. Several other host cell factors are also discussed.

Abstract

Enteroviruses, with their diverse clinical manifestations ranging from mild or asymptomatic infections to severe diseases such as poliomyelitis and viral myocarditis, present a public health threat. However, they can also be used as oncolytic agents. This review shows the intricate relationship between enteroviruses and host cell factors. Enteroviruses utilize specific receptors and coreceptors for cell entry that are critical for infection and subsequent viral replication. These receptors, many of which are glycoproteins, facilitate virus binding, capsid destabilization, and internalization into cells, and their expression defines virus tropism towards various types of cells. Since enteroviruses can exploit different receptors, they have high oncolytic potential for personalized cancer therapy, as exemplified by the antitumor activity of certain enterovirus strains including the bioselected non-pathogenic Echovirus type 7/Rigvir, approved for melanoma treatment. Dissecting the roles of individual receptors in the entry of enteroviruses can provide valuable insights into their potential in cancer therapy. This review discusses the application of gene-targeting techniques such as CRISPR/Cas9 technology to investigate the impact of the loss of a particular receptor on the attachment of the virus and its subsequent internalization. It also summarizes the data on their expression in various types of cancer. By understanding how enteroviruses interact with specific cellular receptors, researchers can develop more effective regimens of treatment, offering hope for more targeted and efficient therapeutic strategies.

1. Introduction

Enteroviruses are small RNA viruses with non-enveloped virions and a single-strand positive RNA genome comprising approximately 7400 nucleotides [1]. They have attracted much attention due to a wide array of clinical manifestations in humans, ranging from mild symptoms to severe diseases like poliomyelitis and viral myocarditis [2]. Several enteroviruses also belong to a group of zoonotic infections [3]. Their simple, structured genomes are the perfect models for the investigation of fundamental virology and molecular biology [4]. Enteroviruses exhibit remarkable genetic flexibility through frequent mutations and recombination, allowing further investigation of viral evolution and adaptation [5].
Enteroviruses are classified into fifteen species based on their structural similarity (Enteroviruses A to L, rhinoviruses A, B, and C); however, many of them retain their original names, like Echovirus 1 and Coxsackievirus A7 [1,6,7,8,9,10,11]. Most enterovirus strains are non-pathogenic. This could be exemplified by the isolation of these viruses from the stool of healthy children or even the usage of certain strains as live enterovirus vaccines (LEVs) [12] for non-specific prevention of viral infections [13,14]. Importantly, enteroviruses can also exhibit oncolytic activity [15]. The most prominent example is the bioselected strain of a non-pathogenic Echovirus type 7/Rigvir that was approved for melanoma treatment, although currently its license is suspended [16,17]. Several other enteroviruses, including the non-pathogenic Coxsackievirus A21 strain CAVATAK and the recombinant poliovirus PVSRIPO, are also currently being evaluated as anticancer agents [18,19,20].
The sensitivity of cells to enteroviruses partly depends on the ability of the agents to exploit specific receptors and coreceptors for viral entry (Table 1) [21,22]. These receptors facilitate virus binding and capsid destabilization, while coreceptors assist in virus binding and interaction with the receptor or any other uncoating factors [23]. Most receptors are glycoproteins, and their carbohydrate components affect interaction with virus particles [7,24]. Many enteroviruses can bind to several alternative cellular receptors for internalization and subsequent replication (Figure 1), thus ensuring the infection of various types of cells. For this reason, enteroviruses have attracted special attention as agents for personalized cancer therapy, as the tumors that do not respond to treatment with one oncolytic enterovirus could still be sensitive to another enterovirus strain that enters cells via an alternative mechanism. So, the development of anticancer agents based on enteroviruses requires an understanding of the mechanisms of their entry into tumor cells, and thus, of the usage of specific cell receptors.
One of the major directions in the development of more effective oncolytic viruses is the enhancement of their oncoselectivity by introducing modifications into their genomes. These modifications include the deletion of viral genes that confer pathogenicity of these infectious agents, alteration of structural proteins to potentiate binding of virions to tumor cells, and expression of factors that enhance the clearance of tumor cells by the host immune system [25,26,27,28,29]. Such approaches are generally applicable to viruses with large genomes including herpesviruses, poxviruses, and rhabdoviruses (particularly to vesicular stomatitis virus) [30,31,32]. However, in the case of enteroviruses, whose genome encodes a peptide of just 2000 amino acids, this approach is much less efficient, as deletions/insertions profoundly reduce virus fitness. Nevertheless, there are a few successful examples. The first one is Rigvir that was obtained by passaging Echovirus 7 on melanoma cells with a suppressed metabolism [33]. Such passaging resulted in several mutations in core proteins that affected the efficiency of virus entry and replication. Another example is the bioselected coxsackievirus B6 obtained by our group that can effectively replicate in various types of tumor cells [34]. Finally, we recently reported the adaptation of Echovirus 1 to glioblastoma and melanoma cells [35]. However, the problem of attenuation of virus internalization remains mostly unexplored.
This review aims to present the current knowledge on the interaction of various enteroviruses with host cell receptors. As non-enveloped viruses may use different receptors, we specifically analyzed literature data on the attachment and internalization of enteroviruses into cells with downregulated expression of specific receptors using CRISPR/Cas9 technology and other approaches. Furthermore, we discuss the existing data on the expression of these receptors in various tumors. So, this review should broaden our understanding of the mechanisms of enterovirus infection and contribute to the development of more efficient oncolytic enteroviruses.
Table 1. Cell receptors used by enteroviruses and mechanisms of their internalization.
Table 1. Cell receptors used by enteroviruses and mechanisms of their internalization.
ReceptorVirusMechanism of EntryReferences
PVRPoliovirus types 1, 2, 3Receptor-mediated endocytosis[36,37,38,39,40,41]
FCGRTEchovirus 1, 3, 6, 7, 9, 11, 13, 14, 18, 25, 26, 30, Coxsackievirus A9, EV-B85Caveolar-mediated endocytosis/receptor-mediated endocytosis[42,43,44]
SCARB2EV71, Coxsackievirus A7, A10 A14, A16Clathrin-mediated endocytosis[45,46,47,48]
IntegrinsEchovirus 1, 5, 8, 9, 22, Coxsackievirus A9, B1Caveolar-mediated endocytosis[49,50,51,52,53]
KREMEN1Coxsackievirus A10,Caveolin-dependent mechanism[54,55]
ICAM-1Coxsackievirus A21Receptor-mediated endocytosis[56,57]
CARCoxsackievirus B1, B2, B3, B4, B5, B6Receptor-mediated endocytosis[58,59,60,61,62]
DAFEchoviruses 6, 7, 11, 12, 20, 21, 70, and D68 (Probable), Coxsackievirus A21, B1, B3, B5Receptor-mediated endocytosis[63,64,65,66,67,68,69,70,71,72]

2. Poliovirus Receptor

The poliovirus receptor (PVR/CD155), a transmembrane immunoglobulin-like glycoprotein, plays a crucial role in cell-to-cell interaction. PVR interacts with other members of the nectin family as well as with IgSF molecules (TIGIT, CD226, CD96), modulating the immune response [73,74,75]. This interaction activates intracellular signaling, leading to IL-10 production by dendritic cells [76,77].
PVR comprises three extracellular Ig-like domains: a V-type domain and two C2 domains [78]. The N-terminal Ig-like domain (D1) is essential for the attachment of poliovirus particles [79,80,81].
PVR binds to poliovirus, interacting with the capsid’s “canyon” region [82,83]. It leads to changes in virion structure and the consequent release of myristoylated VP4 molecules and RNA. This process involves a “pocket factor” lipid molecule in the canyon, facilitating further capsid transformations and direct binding to the cell membrane [6]. The PVR binds a virion with an oblique angle, allowing its D1 domain to enter the canyon and connect with the capsid surface near the center of the icosahedral asymmetric unit [37]. The PVR-binding site is composed of residues of VP1, VP2, and VP3 capsid proteins. On the capsid surface, PVR leaves a distinctive imprint consisting of three separate patches, two of which resemble the attachment pattern of an intercellular adhesion molecule-1 on rhinovirus.
RNA penetration into the cytosol occurs at neutral pH, thus being independent of endosomal internalization [38]. However, internalization traits might vary depending on cell type. For example, in brain endothelial cells, poliovirus binding to PVR induces the rearrangement of the cytoskeleton required for caveolin-dependent entry of the virus [84].
PVR is absolutely crucial for poliovirus, as a pooled sgRNA CRISPR/Cas9-mediated gene knockout in HeLa cells was shown to have led to a complete resistance to poliovirus type 1 infection [85]. This was later confirmed in rhabdomyosarcoma cells (RD) [86]. At the same time, knockout of the PVR gene does not affect the sensitivity of cells to other enteroviruses, indicating that PVR is strictly specific for polioviruses [21]. This is supported by the well-known fact that poliovirus infection typically occurs only in human and primate cells, whereas it can also replicate in rat cells expressing human PVR [87]. So, in the case of resistance of tumor cells to poliovirus, other oncolytic enteroviruses could be used.

3. Intercellular Adhesion Receptor 1

Intercellular Adhesion Molecule 1 (ICAM-1), a member of the immunoglobulin superfamily, comprises 505 amino acids [62,74]. Due to tissue-specific glycosylation, this protein has an apparent molecular weight ranging from 76 to 114 kDa. ICAM-1 consists of five extracellular Ig-like domains, a transmembrane region, and a 28 amino acid cytoplasmic domain [62]. ICAM-1, present on the surface of endothelial and epithelial cells, fibroblasts, and specific hematopoietic cells [75,76,77], acts as a ligand for leukocyte integrins and a receptor for fibrinogen [74,78]. ICAM-1 facilitates the entry of Coxsackievirus A21, Enterovirus 71, and other enteroviruses [21,87], including rhinoviruses [88]. In addition, ICAM-1 is exploited by rhinoviruses [88]. Virions bind ICAM-1 by the “canyon” groove of their capsid, causing the release of the “pocket factor” and subsequent release of viral RNA [88]. After attachment to ICAM-1, enteroviruses are internalized via endocytosis. A noteworthy point is that ICAM-1 is also engaged in cell-to-cell virus spread [89].
The formation of the CVA21/ICAM-1 complex involves the interaction of the ICAM-1 D1 domain with specific regions of the VP1, VP2, and VP3 viral proteins [57]. The D-E loop, B-C loop, and G-F loop of the ICAM-1 D1 domain form contacts with the VP1 βG and βH strands, as well as with the VP1 G-H loop. In addition, the βG strand of the ICAM-1 D1 domain interacts with the VP1 E-F loop and the VP3 G-H loop. This is further enhanced by the connection of the βD and βE strands of the same ICAM-1 domain with the puff region of VP2. Poliovirus has a different structure of the VP2 puff region and the VP1 G-H loop, making interaction with ICAM-1 impossible.

4. Scavenger Receptor B2

Scavenger Receptor Class B Member 2 (SCARB2), also known as the Lysosomal Integral Membrane Protein II (LIMP-2) or CD36-like protein-2, is another receptor for enteroviruses. It is recognized by the coxsackievirus species A7, A14, and A16, as well as by EV71 [48,90,91]. This member of the CD36-like scavenger receptor family is a glycosylated type III transmembrane protein comprising 478 amino acids [91,92,93]. SCARB2 consists of an N-terminal transmembrane domain, an extracellular domain, a C-terminal transmembrane domain, and a cytoplasmic tail [94].
Mutations leading to SCARB2 inactivation in mammals confer heavy syndromes, including peripheral neuropathy and renal failure syndrome, characterized by myoclonus [95]. Deletion of its gene is related to Gaucher disease [96]. SCARB2 controls cell metabolism, as its deletion results in reduced lipid storage and a disbalance between glycolysis and oxidative phosphorylation [97]. In addition, it plays a role in endosome biogenesis and cholesterol trafficking [94]. SCARB2 expression in neural tissues explains EV71 neurotoxicity [48,91,98]. The protein has ten N-glycosylation sites, though glycosylation is not essential for interacting with EV71 [48,99]. Amino acid residues 142–204 of SCARB2 are critical for EV71 attachment and cellular infection [48,99]. Models of the EV71-SCARB2 complex demonstrate that the hydrophobic pocket of the EV71 VP1 capsid protein is positioned next to the lipid-transfer tunnel of SCARB2 [90]. In this proposed binding arrangement, helices α4 and α5 from SCARB2′s viral binding domain are inserted into the EV71 capsid canyon, flanked by helices α7 and α15 in proximity of the virus five-fold and three-fold axes, respectively. This alignment correctly positions the conserved residues such as N102, D219, and K285 of VP1, and R182 and D183 of VP3 for interaction with domain III of the SCARB2 protein. Upon virus binding to SCARB2, EV71 undergoes structural changes facilitated by interaction with other receptors like PVR, CXADR, or ICAM-1. Virus RNA is released via the “lipid channel” in acidified endosomes, leading to VP4 release and A-particle formation, which is essential for virus replication [90,100].
SCARB2 is one of the key receptors for EV71, as both attachment and replication of this virus are significantly reduced after SCARB2 knockout [100,101,102]. SCARB2 importance is further exemplified by the increased susceptibility to EV71 of initially non-permissive mouse cells with ectopic SCARB2 expression [103]. CRISPR/Cas9-generated knock-in mice expressing human SCARB2 gain typical severe symptoms of EV71 infection [46,104,105]. A genome-wide CRISPR/Cas9 screen confirmed that SCARB2 is essential for EV71 infection but also identified SLC35B2 and B3GAT3 as additional host factors [106]. This work revealed that SCARB2 activity as a virus receptor depends on its sulfatation, with SLC35B2 regulating this process. Additionally, SCARB2 plays an important role in Coxsackievirus A10 (CVA10) infection, as siRNA-mediated suppression of SCARB2 expression in RD cells was shown to lead to a marked decrease in CVA10 VP1 expression after infection [47].

5. Integrins

Integrins are transmembrane proteins essential for cell adhesion and binding to various ligands, including extracellular matrix proteins and various plasma proteins [107]. Integrins are composed of α and β subunits, with vertebrates expressing eighteen α-subunits and eight β-subunits [107,108,109]. The subunits, consisting of about 1000 and 750 amino acid residues, are composed of multiple domains: an α-helical transmembrane region and an unstructured cytoplasmic tail.
The activity of α-subunits depends on chelating magnesium and calcium ions at specific sites. At the same time, the extracellular region of β-subunits has a cysteine-rich domain and binds bivalent cations [110]. The N-terminal region of integrin interacts with matrix glycoproteins that display an RGD motif, typical for such ligands as fibrinogen and fibronectin. The cytoplasmic domain is connected to the cytoskeleton via talin and vinculin proteins [107,108,109]. Expressed on numerous types of cells, integrins facilitate gene expression, signaling, cell proliferation, and differentiation [107,110].
Enteroviruses, such as Echovirus 1, 9, and Coxsackievirus A9, utilize integrins as receptors. The α2ß1 integrin (Very Late Antigen 2, VLA-2) binds Echovirus 1 through an RGD-independent mechanism [111], while avß3 and avß6 integrins serve as receptors for both Coxsackievirus A9 and Echovirus 9, with Coxsackievirus A9 requiring the RGD motif for interaction [111,112,113]. Other enteroviruses also utilize these integrins, despite an absence of the RGD motif [79]. The α2I domain binds within the canyon on the EV1 surface, forming extensive contacts with the outer canyon wall [80]. In such a docked position, both the N- and C-termini are exposed to solvent rather than to the virus surface. The α3 helix of the α2I domain and its connecting loops interact with the VP2 capsid protein of one protomer, while the opposite terminus contacts VP3 of a neighboring protomer.
The interaction of integrins with enteroviruses leads to virion structural rearrangement, resulting in the formation of A-particles. For instance, Echovirus 1 interaction with VLA-2 triggers caveolin-dependent endocytosis, with subsequent virion uncoating occurring in caveosomes [114]. The internalization signal is likely to be the virion-induced clustering of integrins that activates a cascade of protein kinases [6,81]. However, internalization mechanisms might vary depending on cell type, underlying different sensitivities to various enterovirus strains.

6. Coxsackievirus and Adenovirus Receptor

The Coxsackievirus and Adenovirus Receptor (CAR) is a transmembrane protein that belongs to the CTX family of the immunoglobulin superfamily (IgSF). It plays a significant role in cell adhesion [6,115,116,117,118,119,120,121]. The complete cellular function of CAR is not entirely understood yet. However, this receptor is involved in cell adhesion and recognition, particularly in homophilic interactions between epithelial cells [121]. CAR is abundantly present in normal and tumor tissues of various organs such as the heart, brain, stomach, prostate gland, and liver [10,115,120]. During embryogenesis, CAR expression is high in the central and peripheral nervous systems [118,119,120], skeletal muscles [121], and different types of epithelia [116,117].
Both human and mouse CAR proteins have an approximate molecular weight of 46 kDa and feature two disulfide bonds [119]. Human CAR, encoded by the CXADR gene, comprises 365 amino acids. It is organized into two extracellular immunoglobulin-like domains, a transmembrane domain, and an intracellular tail [115]. The D1 domain of CAR, which is capable of dimerization, is crucial for virus binding and facilitating intercellular adhesion [122]. Transcription and alternative splicing of the CXADR gene results in the expression of the two mRNA isoforms that encode structurally similar proteins that differ in the C-terminal PDZ segments [115,116,123].
CAR is the primary receptor for Coxsackie B viruses (CV-B1 to B6) and adenoviruses (types 2 and 5) [117,122,124]. Cardiac-specific CAR-knockout mice showed a significant reduction in the replication rate of Coxsackievirus B3 as well as of heart tissue damage caused by the infection. Moreover, knockout of the CXADR gene has demonstrated a loss of sensitivity to Coxsackievirus B5, confirming the vital role of this protein in cellular entry [125].
CAR interacts with viruses in the virion canyon area, triggering A-particle formation [126]. However, the uncoating process varies depending on the viral strain and cell type. For instance, Coxsackievirus B3 forms A-particles in acidified, clathrin-coated endosomes [127], whereas Coxsackievirus B4 is internalized in a clathrin-independent manner [126].

7. Decay Accelerating Factor

The Decay Acceleration Factor (DAF/CD55) is a member of a complement regulatory protein superfamily [128,129,130,131]. This 70 kDa protein comprises several short consensus repeat (SCR) domains [130,131]. DAF is anchored to the cell membrane via interaction with phosphatidyl-inositol groups. It is found on erythrocytes, platelets, and lymphocytes, as well as on various tumor and transformed cell lines [132]. DAF protects cells from complement-mediated lysis [128]. Besides, it is implicated in tumor progression, as it is highly expressed in cancer stem cells (i.e., tumor-initiating cells) and confers resistance to anticancer therapy [133].
DAF is recognized by numerous enteroviruses, including strains of Coxsackieviruses A21, B1, B3, and B5, or various echoviruses [67,68,69,71,134,135]. In particular, it is the main receptor for Rigvir [33]. The mechanism of virion binding to DAF is quite distinctive from other receptors, as it interacts not with the canyon region of a virus particle but stretches around the virion’s surface by forming an arch [6]. For this, DAF uses various SCR combinations and surface areas for Echoviruses 7, 12, and Coxsackievirus B3 [63]. The DAF/CD55 receptor interacts with Echovirus 6 mainly through its SCR3 and SCR4 domains [23]. In SCR4, an arginine residue at position 214 (R214) forms bonds with the T154 residue of the viral VP2 core protein, while aspartate D240 interacts with VP2′s threonine T163. The SCR3 domain interacts with both VP1 and VP2 core proteins through multiple amino acid residues.
The interaction of virions with the DAF receptor induces DAF clustering, activation of c-Abl tyrosine kinase, and subsequent cytoskeleton reorganization. This facilitates the movement of the virus–DAF complex towards CAR. Additionally, the Fyn protein kinase, a member of the Src family, enhances the access of a virus to CAR, subsequent formation of A-particles, and virion uncoating by the phosphorylation of caveolin [136]. It is also tempting to speculate that these clusters are formed on lipid rafts, as these rafts were shown to be involved in DAF-mediated internalization of EV11 [137]. However, it is worth noting that the specific internalization process depends on the viral strain and the cell type.
However, presence of DAF alone in many cases is not sufficient for viral infection, as it does not trigger virion uncoating. Instead, it acts rather as a co-receptor with other receptors on cell surface. For example, the entry of coxsackieviruses A also requires ICAM-1 [69,71] and possibly other viral receptors such as KREMEN 1. For coxsackieviruses B, CAR is essential [138]. According to current concepts, DAF ensures that viral particles concentrate on the cell surface to increase the probability of their interaction with the primary uncoating receptor. While CAR is located in tight junctions and less accessible in the intestinal epithelium [139], apical DAF protein could be more exposed on basal and lateral membranes of tumor cells [140]. This could resemble the case of receptors for hepatitis C virus: CD81 and SR-B1, acting as receptors on the basolateral membrane, are needed to transfer bound virions to claudin 1 and occludin receptors at the tight junctions [141,142]. It should be also discussed that DAF may act as a primary receptor for some strains of enteroviruses. This is supported by several findings for coxsackieviruses. First, overexpression of both DAF and ICAM1 diminishes virus replication levels, compared to overexpression of any single receptor [143]. Second, the bioselection of Coxsackievirus A21 in DAF-expressing but ICAM-1-negative cells by Johansson et al. yielded a strain that rapidly infects ICAM-1-deficient tumor cells [144]. However, it is worth noting that neither of these two studies evaluated the possibility of other co-receptors.
Nevertheless, it remains clear that DAF is one of the key factors for the lytic infection of coxsackieviruses [143,144]. This allows us to propose that the efficacy of oncolysis by coxsackieviruses may be modulated by drugs that decrease DAF expression, and non-steroid anti-inflammatory drugs (NSAID) in particular [145].

8. Kringle Containing Transmembrane Protein 1

Another receptor of coxsackieviruses is the KREMEN protein 1 (KRM1) [54]. Together with the KRM2 paralog, it is a high-affinity receptor for the Dickkopf proteins (DKKs), which are known inhibitors of the canonical Wnt signaling pathway, crucial for embryonic development [146,147,148]. KRM1 and KRM2 are type I transmembrane proteins with an ectodomain consisting of KR, WSC, and CUB domains, essential for inhibiting Wnt signaling [149,150].
KRM1 is specifically utilized as an entry receptor by CVA10, distinguishing it from other enteroviruses like EV71 and CVA16 [54]. KRM1 can bind directly to the virus canyon, triggering the release of the pocket factor and consequent changes in the viral capsid with a release of the viral genome [54,84]. This interaction also initiates viral uncoating by altering the structure of viral particles [55,151].
KRM1 binds across two adjacent asymmetric units (ASUs) via its KR and WSC domains, positioned above the canyon [151]. In the first ASU (VP1.1), interaction with KRM1 occurs through the EF loop (K161 and T163) and the GH gating loop (T209, S217, and T219) of VP1, creating binding patches on both sides of the canyon. The C-terminus of VP1.2 enhances the stability of the interaction by contacting the WSC domain of the receptor. Additionally, the KR domain interacts with the VP2.1 EF loop protrusion (138 to 143) and facilitates the formation of strong ionic bonds. The side chains of W94 and W106 in the KR domain form π-cation interactions with K140 of VP2.1. N140 and Y165 establish additional hydrogen bond and hydrophobic interactions with VP3.2.
Insights into the host genetic factors influencing virus resistance were gained through insertional mutagenesis, creating a large mutagenized HAP1 cell pool. KRM1 was identified in cells resistant to CVA10 as a protein playing an essential role in the infection. It was confirmed by generating KRM1-deficient HAP1 and HeLa cells using TALEN and CRISPR techniques [143]. This resistance was observed without significant response changes to siRNA-mediated KRM1 knockdown, indicating that the resistance phenotype is indeed due to the absence of KREMEN1 [47]. Interestingly, cell lines lacking KRM1 remain permissive to other enteroviruses like poliovirus type 1 and CVB1. Finally, we should mention that KRM1 is abundantly expressed in normal tissues, whereas its expression is much lower or even absent in many, if not most tumor cell lines [146]. This might limit the usage of enteroviruses that rely on KRM1.

9. Neonatal Fc Receptor

The Fc receptor and transporter for the IgG (FCGRT)–α-chain FcRn, also known as the Brambell receptor, is a versatile protein that, among its various functions, ensures the entry of Enterovirus B strains that do not use DAF/CD55 [23,43]. The FCGRT gene encodes FcRn that mediates the transfer of maternal gamma globulins to neonates, thus providing passive immunity during the early stages of immune system development [152]. Structurally, FcRn is a complex of an alpha chain encoded by the FCGRT gene (consisting of three extracellular domains, a transmembrane domain, and a short cytoplasmic tail) and the beta-2 microglobulin (B2M) [153,154]. FcRn also plays a role in protecting gamma globulin and albumin from degradation by binding to these proteins, thus prolonging their half-life in circulation and ensuring its functional efficacy [155,156,157]. Functionality of this receptor partially depends on glycosylation [158,159].
Elevated levels of FcRn have been reported for several types of tumors, including breast, ovarian, and lung cancers, where its expression has been associated with poor prognosis and disease progression [153,160,161,162,163]. FcRn can influence tumor evasion and metastasis [164]. It also modulates the activity of immune cells like NK cells and macrophages through IgG interactions, enhancing processes like antibody-dependent cell-mediated cytotoxicity (ADCC) and antibody-dependent cellular phagocytosis (ADCP), required for the elimination of tumor cells [165,166,167].
The role of FcRn as an enterovirus receptor was discovered in a study of Echovirus 6 (strain SJZ-366) via a CRISPR knockout screen in RD cells targeting genes of membrane proteins using an sgRNA library [99]. FcRn interacts with enteroviruses such as Echovirus 30 and enhances virus internalization after initial attachment to surface receptors like DAF or CAR [44]. However, other groups reported that DAF may not be involved in the FcRn-mediated entry of echoviruses [43]. According to current knowledge, FcRn is the key receptor for the entry of various types of echoviruses [168].
FcRn binds to Echovirus 6 virus via polar interactions with the α2 and α3 helices in the α2 domain, creating two interfaces on each side of the canyon [23]. Key residues like Q139, R140, Q143, and Q144 in the α2–α3 junction form hydrogen bond networks with the VP1, VP2, and VP3 subunits. Residues D145 and K146 interact in the canyon, forming bonds with VP1 subunit residues Y75, D91, and Q99, possibly crucial for transporting the “pocket factor”.
In summary, FcRn’s array of functions spans from its traditional role in immunity to its involvement in enterovirus infection and influence on immune cell functions, illustrating its importance in both viral pathogenesis and immune response mechanisms.

10. RIG-I and MDA5

The innate antiviral response activates RIG-I-like receptor (RLR) signaling, critically affecting virus entry [169]. Pattern recognition receptors detect molecular signatures associated with pathogens and initiate the production of type I and III interferons (IFN) to counteract viral invasion [170,171,172,173,174]. Two key sensors of viral RNA in human cells are the retinoic acid-inducible gene I (RIG-I) and melanoma differentiation-associated gene 5 (MDA5) [175]. Both of them contain two N-terminal caspase activation and recruitment domains (CARDs), a C-terminal regulatory/repressive domain, and a central DExD/H-box ATPase/helicase domain that detects viral RNA [176]. When enteroviruses infect the host cell, they release their RNA genome into the cytoplasm [177]. RIG-I identifies various features of viral RNA such as the triphosphate fragment on its 5′-terminus (5′-ppp) from de novo RNA synthesis and short double-stranded RNA (dsRNA) regions commonly found in single-stranded viral genomes [178]. When RIG-I binds to viral RNA, it undergoes conformational changes that expose its Caspase Activation and Recruitment Domains (CARD) motifs. This conformational shift enables RIG-I to interact with another crucial cellular component—the Mitochondrial Antiviral Signaling Protein (MAVS) (Figure 2) [179,180].
Different enteroviruses cleave MDA5, MAVS, and RIG-I in infected cells using their proteases: 2Apro digests MDA5 and MAVS, while 3Cpro targets RIG-I [181]. These cleavage events explain the downregulation of the IFN-α/β response in enterovirus-infected cells. In CVB3-infected cells, TBK1 phosphorylation and subsequent activation are suppressed, suggesting that this pathway is inhibited upstream of TBK1 [182]. The non-structural 2Cpro of CVA6 suppresses the production of IFN-β by targeting MDA5 and RIG-I to lysosomal proteases. This function is shared by EV71 and CVB3 but not by CVA16, suggesting that the latter may have an alternative way to stimulate virus replication. Also, 3Cpro of EV71 can inhibit IFN-β activation and RIG-I during viral infection. However, EV71 3Cpro does not exhibit inhibitory activity against MDA5 [183,184]. Structured RNA elements in the genome of CVB3 are recognized by MDA5 during infection and activate RIG-I if viral RNA retains 5′-ppp. The 5′ppp-containing cloverleaf (CL) RNA structure is a potent inducer of RIG-I, triggering a string antiviral response that includes the induction of classical interferon-stimulated genes as well as type III interferon and proinflammatory cytokines and chemokines like TNFα, IL-1α/β, IL-6, and CXCL-10 [185]. Interestingly, RIG-1 polymorphisms are linked to an increased risk of EV71-induced HFMD [186,187].
Toll-like receptor 3 (TLR3), located on the endosome membrane, is the third main sensor of viral double-stranded RNA. When bound to dsRNA, TLR3 associates with the Toll/interleukin-1 receptor (TIR)-containing adapter inducing IFN-β (TRIF or TICAM1). It activates IRF3 and NF-κB, triggering the activation of antiviral defense and inflammatory response in the infected cells. The TLR3/TICAM1 pathway is crucial for protecting the immune system against picornavirus infections. TICAM-1−/− and human PVR—transgenic mice were more susceptible to poliovirus infection than the wild-type and MAVS−/− transgenic mice [188]. Accordingly, TICAM-1−/− transgenic mice exhibited the lowest levels of type I IFN mRNA in spleen dendritic cells among the mice with knockouts of various dsRNA sensors. Recent research indicates that the antiviral response to poliovirus infection includes pathways mediated by MDA5, TLR3, and the TLR adaptor MyD88, with TLR3 being the most profoundly upregulated [189].
RIG-I activity is regulated by the inhibition of Ubiquitin-Specific-Processing Protease (CYLD) expression, as well as by miR-526a. EV71 infection specifically enhances miR-526a expression in macrophages through an interferon-regulatory factor (IRF)-dependent mechanism. Virus-induced miR-526a activation and suppression of CYLD are both blocked by EV71 3Cpro, which is required for effective infection; ectopic miR-526a expression downregulates EV71 replication [190]. CVB3 activates another microRNA (miR-30a) to facilitate its replication. This is achieved through modulation of the interferon response, as miR-30a is a potent negative regulator of IFN type I signaling, acting on the tripartite motif protein 25 (TRIM25) [191]. Overexpression of TRIM25 markedly inhibits CVB3 replication, while TRIM25 knockdown leads to an increase in virus RNA production and titer and expression of the VP1 protein. By inhibiting TRIM25 expression and TRIM25-mediated ubiquitination of RIG-I, miR-30a reduces IFN-β production and activation. This may represent a mechanism by which CVB3 evades the host’s innate immune response [191,192,193].

11. MAVS

MAVS is a common adaptor molecule of RIG-I/MDA5, thus being essential for type I IFNs [194,195]. Its C-terminal transmembrane domain confers localization on the outer membrane of mitochondria [196]. Viral infection triggers the activation of RIG-I/MDA5, which, in turn, leads to the recruitment of MAVS through tandem Activation and Recruitment Domains (CARDs). Subsequently, MAVS connects RIG-I/MDA5 to IKK and TBK1/IKKε, activating IRF3 and NFκB—the two crucial components of the pathway that induce IFN in response to viral infection [197,198]. The interaction between RIG-I and MAVS activates several signaling pathways, including NF-κB and IRF [199,200]. These events initiate the production of type I interferons [201].
RNA viruses such as Sendai virus (SeV) and VSV trigger the activation of TRIM21, interacting with MAVS and catalyzing MAVS polyubiquitination, activating IRF3, and suppressing viral infections [202]. When mice are infected with CVB3, TRIM21 expression is increased [203]. Assessment of viral kinetics revealed that the cleavage of MAVS is a significant predictor of outcomes of infections. A common MAVS mutation that eliminates the protease-targeted region, ironically, increases rates of CVB3 infection [204]. Despite type I IFN signaling induced by CVB3 infection in HeLa cells, antiviral cytokines cannot restrict CVB3 replication. This can be rescued by the treatment of cells with exogenous MAVS protein before or after CVB3 exposure for enhancement of the innate immune response and prevention of virus spread and consequent cell damage [205].
Overexpression of MAVS in RD cells can induce both autophagy and apoptosis [206]. MAVS-induced apoptosis requires the inhibition of the extracellular signal-regulated kinase (ERK) signaling pathway and activation of c-Jun N-terminal kinase (JNK) signaling. Additionally, suppression of autophagy by 3-methyladenine (3-MA) enhances MAVS-induced apoptosis by increasing the cleavage of caspase-3 and poly(ADP-ribose) polymerase (PARP). Rapamycin-induced autophagy suppresses apoptosis by MAVS overexpression. Many viruses target MAVS for evasion from the innate immune response. One of the first described examples was the cleavage of MAVS by hepatitis C virus serine protease [207]. Later, such cleavage was shown for EV71, Echovirus 7, and CVB5, that target the Gly209, Gly251, and Gly265 residues of this cellular protein [208]. Other enterovirus strains cleave MAVS at different sites, as exemplified by CVB3 targeting the Gln148 residue [209]. Coxsackievirus A16 (CA16) may also suppress MAVS expression for efficient replication [206].

12. Sialic Acids

Sialic acids are nine-carbon carboxylated sugars. They are commonly found as moieties of glycoproteins and glycolipids at the surface of mammalian cells [210]. They play a significant role in the mediation of cell–cell contacts, the immune response, and host–pathogen interaction [211,212]. Concerning enteroviruses, the presence or absence of specific sialic acids on the surface of host cells can affect virus attachment, entry, and therefore virus tropism [5,213,214,215,216].
Many enteroviruses, including coxsackieviruses [217] and echoviruses, use sialic acid-containing molecules as cell surface receptors [218]. The ability of enteroviruses to bind sialic acids is primarily conferred by the pocket factor [177]. Some enterovirus strains have adapted to recognize variants of specific sialic acids for a more efficient infection of target host species/tissues [219]. The interaction between enteroviruses and sialic acids depends on evolutionary dynamics. Over time, the virus may undergo genetic changes to adapt to different sialic acid receptors, which might affect virus pathogenicity and tissue tropism [220,221]. This evolution may have significant implications for the emergence of new strains and the potential for interspecies transmission [222].
Understanding the role of sialic acids in enterovirus infection may have therapeutic implications. Targeting the interaction of viruses with sialic acids, such as blocking the binding site on the viral capsid, may be a potential strategy for developing antiviral drugs [5]. However, this field remains largely unexplored, warranting additional studies.

13. PLA2G16

PLA2G16, also known as the group XVI phospholipase A2 or AdPLA (lipid-specific phospholipase A2), is an enzyme belonging to the phospholipase A2 (PLA2) family [223]. PLA2 enzymes play a significant role in lipid metabolism [224], inflammation [58], and the regulation of various cellular processes. PLA2G16 has received attention in the context of enterovirus infection due to its potential role in the replication and pathogenesis of several enteroviruses [54,225].
PLA2G16 appears to be involved in virus reproduction by promoting the formation of membranes or vesicles that serve as platforms for viral RNA replication, possibly by modifying cellular lipids and their recruitment to replication organelles [225]. Moreover, this protein ensures the delivery of viral RNA into the cytoplasm [226]. In contrast, it cannot act as a viral receptor or a factor of virion trafficking into the cell. A genome-wide forward screen using Haplobank pointed to PLA2G16 as a factor essential for the cytotoxicity of rhinovirus A—a pathogen responsible for the common cold [227]. PLA2G16 knockdown reduced the migration and invasion of p53-null cells, while its overexpression enhanced their invasion. Mutant p53 elevated PLA2G16 levels in mouse and human osteosarcoma cells, suggesting PLA2G16 as a downstream target of p53 [228]. Beggen et al. reported the bioselection of an EV71 variant in PLA2G16-knockout cells; the mutations in the viral capsid protein reduced the dependency of the pathogen on PLA2G16, although at a cost of reduced thermostability [225].
In addition to its significance for enterovirus replication, PLA2G16 participates in unrelated cellular processes like lipid metabolism, adipocyte function, and stress responses. Understanding the specific mechanisms by which PLA2G16 promotes viral replication and pathogenesis in enterovirus infections remains an active research area. It is also important to note that PLA2G16′s involvement may vary across different enterovirus serotypes and strains [226].

14. Expression of the Receptors in Tumors

14.1. Poliovirus Receptor

Poliovirus receptor is a protein that is ubiquitously expressed in a variety of cells and tissues, albeit mostly at moderate levels [229,230]. However, many types of tumors exhibit elevated PVR expression compared to normal tissues. They include colorectal [231,232], prostate [233], renal [233], pancreatic [233,234], and esophagus small cell carcinomas. Higher levels of the receptor were also reported for non-small cell lung [235] and gastric [236] cancers, melanoma [237], liver cancer [238], and specifically cholangiocarcinoma [239], glioblastoma [240], bladder cancer [238,241] and its muscle-invasive type [242], and breast cancer [243] including its triple-negative variants [244]. PVR is likely to be overexpressed in neuroectodermal tumors, as it is controlled by the Sonic Hedgehog signaling pathway [245]. Indeed, several groups demonstrated elevated levels of PVR/CD155 in medulloblastoma [246,247]. Induction of PVR/CD155 is not a feature of advanced tumors only: increased expression could be already observed at a stage of precancerous lesions (at least in the stomach [236]) and early-stage tumors [248]. Importantly, the development of cancer in many cases is accompanied by the presence of soluble PVR/CD155 that can serve as a biomarker [244,249].
Increased expression contributes to the motility and invasiveness of tumor cells, leading to higher recurrence rates and, therefore, to poor prognosis, as described for different types of tumors [233,235,239,241,250,251,252,253]. In addition, high expression of PVR was noted as a negative prognosis factor for multiple myeloma [254]. High levels of the receptor are also associated with enhanced angiogenesis [252]. For triple-negative breast cancer, high PVR expression is a feature of tumor cells with the mesenchymal phenotype [255]. However, at least for medulloblastoma, no differences in PVR expression between primary tumors and metastases were shown [247]. PVR/CD155 also contributes to cancer progression by downregulating the immune response to tumor cells. For example, its presence on the surface of tumor cells leads to the inactivation of melanoma-specific T-cells [237] and suppression of NK-cells [256] that could have eliminated cancer cells. In case of melanoma, PVR expression was shown to confer resistance to anti-PD1 immunotherapy [257]. High expression of PVR can also trigger the degradation of CD226 on the surface of CD8+ T-cells, also leading to an immunosuppressed phenotype [258].

14.2. ICAM-1

ICAM-1 is expressed in various types of tumors. Most reports discuss high levels of ICAM-1 in lung cancer, specifically in its triple-negative variant [259,260,261] as well as its metastasis into lungs [262]. Its expression is also elevated in thyroid carcinomas [263,264,265], although not in all tumors [266]. Eighty-three per cent of clear cell renal carcinomas [267] and half of gastric tumors [268] are ICAM-1-positive, although metastasis of the latter into liver leads to enhanced expression in almost all cases. A similar pattern was reported for colon [269] and bladder [270] cancer: expression of ICAM-1 is associated with invasiveness and poor prognosis. The opposite dependence between ICAM-1 expression and invasiveness was described for melanoma, in which the presence of this receptor on tumor cells is associated with lower aggressiveness and a higher patient survival rate [271,272]. Indeed, metastatic derivatives of melanoma cells are negative for ICAM-1 [271].
ICAM-1 is also found in tongue squamous cell carcinoma, at least in its invasive front area [273]. Not much is known about the status of ICAM-1 in lung tumors, but most, if not all, non-small cell lung cancer cell lines express this receptor [274]. In contrast, cell lines corresponding to small-cell lung cancer are predominantly ICAM-1-negative. Bladder cancer is characterized by varied levels of ICAM-1 and, as a consequence, different sensitivities of tumors to Coxsackievirus type A [275]. In contrast, ICAM-1 is not expressed in retinoblastoma [276]. It is also reduced in ovarian adenocarcinoma, which is achieved via methylation of its promoter [277].
There are also several reports showing an increase in ICAM-1 expression in cancer cells after treatment with pharmacological agents or experimental approaches, thus providing a clue to the enhancement of oncolysis. In B-cell lymphoma cells, the increase could be triggered by metformin [278], while in osteosarcomas, by interleukin 6 [279].

14.3. SCARB2

SCARB2 was cloned in 1991 and described as LIMP2 [280]. This protein is localized on the ER and plasma membrane, at least in Cos cells [280], as well as on the membranes of lysosomes in various types of cells [281]. SCARB2 mediates cholesterol efflux [282] and, together with SR-BI, participates in the influx of high-density lipoproteins [283]. Again, both SR-BI and SR-BII mediate HCV entry [284].
SCARB2 is highly expressed in the spleen, at somewhat lower levels in liver, colon, and adrenal gland, and at low levels in the heart, lung, and thymus [283]. No expression was found in the kidneys and pancreas [283]. Evaluation of its levels in tumor cell lines revealed high expression in cervix carcinoma HeLa cells, melanoma 14Mel and C32 lines, and promonocytic leukemia U973 cells. Moderate expression was noted in promyelocytic (HL60) and erythro-(HEL) leukemia cell lines, whereas very low levels of expression were described for a Jurkat cell line derived from a patient with T-cell leukemia [281]. In patient samples, elevated expression compared to normal tissues was reported for glioblastoma [285], breast cancer [286], and lymph node-positive oral squamous carcinoma [287], with high levels correlating with poor prognosis. Unfortunately, information about its expression in other types of cancer is still missing.

14.4. CAR

CAR is expressed in a wide array of normal tissues, while in tumors its levels remain heterogenous [288]. It is highly expressed in ovarian cancer [289,290,291], squamous cell and small cell lung cancer [292,293,294,295], oral squamous cell carcinoma [296], thyroid [297], prostate [298,299], and hormone-sensitive breast cancer [300,301,302], with levels often higher that in the respective normal tissues. CAR is also often found in soft tissue and bone sarcomas [296,303,304,305], although contrary data are also present [306]. Specifically, in prostate tumors CAR is found on the cell membrane, indicating that it should be fully functional [298]. This implies that these types of cancer could be treated with viruses that are internalized by this receptor. However, in the case of ovarian carcinoma, the highest levels are observed in early-stage tumors, while late-stage disease is characterized by a decline in CAR levels [307]. Nevertheless, the recurrent tumors are still receptor-positive, indicating that oncolytic viruses could still be effective [290], pending that an increase in soluble CAR receptors from an advanced tumor will not diminish the levels of the virus [291]. Endometrial sarcomas, another type of gynecological cancer, are in 50% of cases CAR-negative, although some positive tumors exhibit high levels of expression of this receptor [308]. Heterogenous expression was also described for head and neck cancer [309,310]. In non-small cell lung cancer, the expression of CAR is a feature of cancer stem cells [294] that are tumor-initiating cells and the cells with the tumor-resistant phenotype. This also points to the respective oncolytic viruses as a possible treatment for the disease. However, as tumors also produce soluble CAR, this can lead to reduced efficiency of oncolysis due to virus neutralization [311].
In contrast, pronouncedly decreased CAR expression was reported for bladder [312,313] and prostate [288] cancer, as well as in various gastrointestinal tumors, including esophagus [314,315], gastric [316], pancreatic [314], liver [314,317], and colon [288,314,318,319] cancers. However, CAR expression is elevated in Barrett’s esophagus, though in advanced tumors the receptor is no longer detected [320]. In bladder cancer and astrocytomas, CAR levels continue to decline with tumor progression [312,321], and in many bladder tumors and cell lines CAR is no longer detected [313,322,323]. Most advanced astrocytomas (glioblastomas) [321,324,325] and invasive gastric tumors [316] do not express the CAR receptor. No CAR expression was detected in B16 melanoma cell lines [326].
As decreased expression of the CAR receptor results from the methylation of its promoter, the inhibitors of histone deacetylases restore/enhance its levels [301,327]. However, hypoxia is another factor that reduces CAR levels, at least in prostate, gastric. and colon cancer [328]. In estrogen receptor-positive breast cancer cells, CAR expression could also be increased by treatment with estradiol [302], while in glioblastomas and cervical cancer CAR levels are reduced during treatment with dexamethasone [329].

14.5. DAF

DAF has heterogenous expression in various types of cancer. For example, it is not expressed in ductal carcinoma [330], meningiomas and astrocytomas [331,332], neuroblastomas [333], and gliomas [334], and many cases of non-Hodgkin lymphoma [335] but is abundant in acute lymphoblastic (ALL) [336] and chronic myeloblastic (CML) [336] leukemias and in many cell lines derived from Burkitt’s lymphoma [337]. Another group of researchers nevertheless noted that its expression in ALL and CML is decreased compared to cells from healthy donors [338]. Increased expression of DAF has been reported for oral carcinoma [339] and Barrett’s esophagus that precedes esophagus carcinoma [340,341]. Tumors of the gastrointestinal tract are generally positive for DAF, as noted for colon [342], pancreatic [342,343], and gastric [342,344] cancers, with the exception of KATO III gastric cancer [342]. This does not imply that DAF is present in all tumors of the abovementioned types of cancer: for colon cancer, approximately 30% of tumors remain negative [132], including metastasis of liver cancer [345]. But, there is a clear increase in DAF expression in tumors compared to normal tissues and benign tumors [132,346]. A similar increase was also reported for gastric [347] and pancreatic [343] cancer, as well as osteosarcoma [347], cholangiocarcinoma [348], gallbladder cancer [349], and oral squamous cell carcinoma [350]. In the case of gastric cancer, this increase is attributed to Helicobacter pylori and specifically to its CagA antigen [351,352].
Thyroid cancer is another type of malignancy that is associated with elevated DAF expression [353,354]. DAF expression was also reported for breast [355] and renal [356] cancer, head and neck squamous cell [357,358] and nasopharyngeal [359] carcinomas, as well as in various gynecological cancers: cervix squamous cell carcinomas [360], endometrial cancer [361,362,363], and uterine serous carcinomas [364]. The only exception is ovarian cancer, for which DAF expression is decreased [361]. A noteworthy point is that the increase in DAF expression in endometrial cancer is observed in early-stage tumors, whereas at advanced stages the expression of the receptor diminishes [362]. Finally, DAF is expressed in lung cancer cell lines [365].
Importantly, high DAF expression has been associated with invasiveness and poor prognosis, as exemplified for colon [343,366], cervix [360], gallbladder [99], and nasopharyngeal [359] cancer.
It should be noted that in some types of cancers, DAF is not expressed in all cells within a tumor. In neuroblastomas, DAF is found only in a minor population of tumor cells, with significant co-staining with HIF2α pointing to hypoxia as one of the mechanisms of induction of this receptor [367]. Similar heterogeneity in expression within a tumor was reported for endometrial cancer [368]. In gliomas, the weak staining was noted not in tumor cells but in endothelia [334], while in cervix carcinomas the DAF is found in stromal cells adjacent to tumor cells [369].

14.6. KRM1

There are only scarce data about KRM1 expression in various types of tumors and cancer cell lines. The Protein Atlas suggests that this gene has moderate-to-low expression in almost all cell lines investigated, with the highest rates of its transcript registered in the Rhabdomyosarcoma Ch30 cell line, gastric cardia OE19 cells, hepatocarcinoma HepG2, and Sk-Mel-30 melanoma cells [229,230]. The same database confirms the presence of the protein in Ch30 cells, as well as in the H2OS osteosarcoma cell line, albeit with nuclear localization. This gene is efficiently transcribed in the lung adenocarcinoma A549 cell line, although other lung cancer cells may exhibit much lower expression levels [370]. Similarly, KRM1 expression was shown for prostate and breast cancer cell lines (such as PC3, MCF7, or T47D) whereas such cell lines as PC3 and MDA-MB-231 are characterized by significantly lower levels of this gene [371]. Sumia et al. reported the reduced expression of the KRM1 gene in breast invasive carcinoma, colon carcinoma, head and neck squamous cell carcinoma, kidney renal carcinomas, prostate adenocarcinoma, and thyroid carcinoma, compared to normal tissues [372]. The same team showed elevated expression in hepatocellular carcinoma and cholangiocarcinoma, as well as in lung squamous cell carcinoma. Elevated expression was also reported in stromal cells from patients with multiple myeloma, compared to stromal cells from healthy patients [373].

14.7. FcRn

FcRn is expressed at high levels in various organs and tissues. The Protein Atlas reports that the highest levels are in the liver, spleen, and gastrointestinal tract, while the lowest are in the retina and cerebellum [229,230]. However, its expression is significantly diminished in various types of cancer. Lung, bladder, ovarian, and cervix cancer, as well as head and neck squamous cell carcinomas are FcRn-negative [374]. This gene is not expressed in most breast tumors, including triple-negative breast cancer, and also not in pancreatic and renal tumors. FcRn expression is detectable in 50% of cases of colorectal cancer [374] or in 40% of endometrial cancer. Non-small cell lung cancer is also mostly negative for the expression of this gene, although FcRn-positive cases are considered to have better survival rates [375]. Moreover, the detection in such tumors is often found not in cancer cells but in resident and tumor-infiltrating immune cells [163]. In addition, FcRn is expressed in dendritic cells that are involved in the protection against colorectal cancer [153]. FcRn expression is elevated in pilocytic astrocytomas—the benign tumors with low growth rates [376]. FcRn expression was reported in hepatocellular carcinoma cell lines (HepG2 and Snu-475) [377], as well as in THP1, Jurkat, and U937 cells that were derived from patients with acute monocytic leukemia, T-cell leukemia, and histiocytic lymphoma, respectively [378]. In contrast, various breast and prostate cancer cell lines are FcRn-negative [164].

15. Conclusions

The intricate interplay between enteroviruses and host cell factors offers a unique window into the mechanisms of viral infection and its potential exploitation for therapeutic purposes, particularly in cancer therapy. The specificity of interaction of enteroviruses with cellular receptors such as PVR, DAF, ICAM-1, and SCARB2 not only determines the tropism and pathogenicity of these viruses, but also unveils a promising avenue for oncolytic virus therapy. The ability of certain enterovirus strains to selectively target and lyse tumor cells emphasizes their therapeutic potential as anticancer agents. This could be exemplified by the use of Rigvir in clinical practice for treatment of melanoma as well as by ongoing trials of other enteroviruses.
Evaluation of the expression of these receptors in tumors gives clues to the choice of oncolytic virus for its treatment. SCARB2-expressing hepatocellular carcinoma cells or ICAM-1-positive triple-negative breast cancer cells could be targeted by EV71 or Coxsackievirus A, while PVR-expressing glioblastomas could be targeted by polioviruses. Special emphasis should be given to DAF, as this receptor is required for fast and lytic echovirus or coxsackievirus (A or B) infections. In the latter case, oncolytic therapy should probably not be applied to patients on non-steroid anti-inflammatory drugs, as they can diminish expression of the DAF receptor. Moreover, the influence of antitumor agents on the expression of receptors for enteroviruses merit further studies. We should also mention that tumors could possibly be treated by a composition of enteroviruses, as tumors are heterogeneous and could be composed of cells with different patterns of receptor expression.
Another question that requires future study is the exploration of whether enteroviruses can establish non-lytic infection, as there are several reports of their non-lytic replication [2,379]. Similar reports of high-level infection in tumor cells in the absence of signs of cytotoxicity also exist for such viruses as SARS-CoV-2 [380,381]. Such infection would not only be inefficient but may also impose threat to cancer patients. An investigation into mechanisms by which the virus may evade oncolytic [action and establish continuous replication is definitely required.
As we continue to unravel the complexities of enterovirus biology and its interaction with host cell factors, it is imperative that we also focus on the translational aspects of this research. Collaborative efforts at the interface of virology, molecular biology, immunology, and clinical oncology are essential to ensure host-cell usage of full therapeutic potential of enteroviruses against cancer. As a part of a broader oncolytic virus arsenal, enterovirus-based therapies offer hope for cancer patients, opening a new era of targeted, effective, and innovative treatments.

Author Contributions

Conceptualization, O.N.A., A.V.I., P.M.C. and A.V.I.; methodology, O.N.A., A.V.I. and A.V.L.; data curation, O.N.A., L.T.H., P.O.V., D.V.K., Y.D.G., E.R.N., D.O.C., A.V.I. and A.V.L.; writing—original draft preparation, O.N.A., A.V.I., P.M.C. and A.V.L.; writing—review and editing, O.N.A., L.T.H., P.O.V., D.V.K., Y.D.G., E.R.N., D.O.C., A.V.I., P.M.C. and A.V.L.; visualization, O.N.A.; supervision, A.V.L.; project administration, P.M.C. and A.V.L.; funding acquisition, P.M.C. and P.O.V. All authors have read and agreed to the published version of the manuscript.

Funding

The project was financially supported by the Russian Science Foundation (grant no. 23-14-00370). The part of the project dedicated to the analysis of echovirus receptors was supported by the Russian Science Foundation (grant no. 19-74-10086). The antiviral innate response was analyzed with support of the Ministry of Science and Higher Education of the Russian Federation (Agreement No. 075-15-2019-1660).

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Vorobyev, P.O.; Babaeva, F.E.; Panova, A.V.; Shakiba, J.; Kravchenko, S.K.; Soboleva, A.V.; Lipatova, A.V. Oncolytic Viruses in the Therapy of Lymphoproliferative Diseases. Mol. Biol. 2022, 56, 684–695. [Google Scholar] [CrossRef] [PubMed]
  2. Genoni, A.; Canducci, F.; Rossi, A.; Broccolo, F.; Chumakov, K.; Bono, G.; Salerno-Uriarte, J.; Salvatoni, A.; Pugliese, A.; Toniolo, A. Revealing enterovirus infection in chronic human disorders: An integrated diagnostic approach. Sci. Rep. 2017, 7, 5013. [Google Scholar] [CrossRef] [PubMed]
  3. Fieldhouse, J.K.; Wang, X.; Mallinson, K.A.; Tsao, R.W.; Gray, G.C. A systematic review of evidence that enteroviruses may be zoonotic. Emerg. Microbes Infect. 2018, 7, 164. [Google Scholar] [CrossRef]
  4. Wang, S.H.; Wang, K.; Zhao, K.; Hua, S.C.; Du, J. The Structure, Function, and Mechanisms of Action of Enterovirus Non-structural Protein 2C. Front. Microbiol. 2020, 11, 615965. [Google Scholar] [CrossRef]
  5. Baggen, J.; Thibaut, H.J.; Strating, J.; van Kuppeveld, F.J.M. The life cycle of non-polio enteroviruses and how to target it. Nat. Rev. Microbiol. 2018, 16, 368–381. [Google Scholar] [CrossRef]
  6. Bergelson, J.M.; Coyne, C.B. Picornavirus entry. Adv. Exp. Med. Biol. 2013, 790, 24–41. [Google Scholar] [CrossRef]
  7. Mercer, J.; Schelhaas, M.; Helenius, A. Virus entry by endocytosis. Annu. Rev. Biochem. 2010, 79, 803–833. [Google Scholar] [CrossRef]
  8. Hinshaw, J.E. Dynamin and its role in membrane fission. Annu. Rev. Cell Dev. Biol. 2000, 16, 483–519. [Google Scholar] [CrossRef]
  9. Mercer, J.; Helenius, A. Gulping rather than sipping: Macropinocytosis as a way of virus entry. Curr. Opin. Microbiol. 2012, 15, 490–499. [Google Scholar] [CrossRef]
  10. Takei, K.; Haucke, V. Clathrin-mediated endocytosis: Membrane factors pull the trigger. Trends Cell Biol. 2001, 11, 385–391. [Google Scholar] [CrossRef]
  11. Sinclair, W.; Omar, M. Enteroviruses; StatPearls Publishing: Treasure Island, FL, USA, 2023. Available online: https://www.ncbi.nlm.nih.gov/books/NBK562330/ (accessed on 28 April 2024).
  12. Voroshilova, M.K. Potential use of nonpathogenic enteroviruses for control of human disease. Prog. Med. Virol. 1989, 36, 191–202. [Google Scholar] [PubMed]
  13. Chumakov, K.; Benn, C.S.; Aaby, P.; Kottilil, S.; Gallo, R. Can existing live vaccines prevent COVID-19? Science 2020, 368, 1187–1188. [Google Scholar] [CrossRef] [PubMed]
  14. Chumakov, M.P.; Voroshilova, M.K.; Antsupova, A.S.; Boiko, V.M.; Blinova, M.I.; Priimyagi, L.S.; Rodin, V.I.; Seibil, V.B.; Sinyak, K.M.; Smorodintsev, A.A.; et al. Live enterovirus vaccines for the urgent nonspecific prevention of mass respiratory diseases during autumn and winter epidemics of influenza and acute respiratory diseses. J. Microbiol. Epidemiol. Immunol. 1992, 11, 37–40. [Google Scholar]
  15. Chumakov, P.M.; Morozova, V.V.; Babkin, I.V.; Baikov, I.K.; Netesov, S.V.; Tikunova, N.V. Oncolytic enteroviruses. Mol. Biol. 2012, 46, 712–725. [Google Scholar] [CrossRef]
  16. Donina, S.; Strele, I.; Proboka, G.; Auzins, J.; Alberts, P.; Jonsson, B.; Venskus, D.; Muceniece, A. Adapted ECHO-7 virus Rigvir immunotherapy (oncolytic virotherapy) prolongs survival in melanoma patients after surgical excision of the tumour in a retrospective study. Melanoma Res. 2015, 25, 421–426. [Google Scholar] [CrossRef]
  17. Babiker, H.M.; Riaz, I.B.; Husnain, M.; Borad, M.J. Oncolytic virotherapy including Rigvir and standard therapies in malignant melanoma. Oncolytic Virotherapy 2017, 6, 11–18. [Google Scholar] [CrossRef]
  18. Bradley, S.; Jakes, A.D.; Harrington, K.; Pandha, H.; Melcher, A.; Errington-Mais, F. Applications of coxsackievirus A21 in oncology. Oncolytic Virotherapy 2014, 3, 47–55. [Google Scholar] [CrossRef]
  19. Yla-Pelto, J.; Tripathi, L.; Susi, P. Therapeutic Use of Native and Recombinant Enteroviruses. Viruses 2016, 8, 57. [Google Scholar] [CrossRef]
  20. Desjardins, A.; Gromeier, M.; Herndon, J.E., II; Beaubier, N.; Bolognesi, D.P.; Friedman, A.H.; Friedman, H.S.; McSherry, F.; Muscat, A.M.; Nair, S.; et al. Recurrent Glioblastoma Treated with Recombinant Poliovirus. N. Engl. J. Med. 2018, 379, 150–161. [Google Scholar] [CrossRef]
  21. Hoa, L.T. Development of Diagnoatic Panels of Knockout Cells for Functional Classification of Enteroviruses. Ph.D. Thesis, Moscow Institute of Physics and Technology, Moscow, Russia, 2020. [Google Scholar]
  22. Lipatova, A.V. Non-Pathogenic Strain of Human Enterovirus, Related to Coxsackievirus B5 as a Model for Study Viral Oncolysis. Ph.D. Thesis, Engelhardt Institute of Molecular Biology (EIMB), RAS, Moscow, Russia, 2017. [Google Scholar]
  23. Zhao, X.; Zhang, G.; Liu, S.; Chen, X.; Peng, R.; Dai, L.; Qu, X.; Li, S.; Song, H.; Gao, Z.; et al. Human Neonatal Fc Receptor Is the Cellular Uncoating Receptor for Enterovirus B. Cell 2019, 177, 1553–1565.e1516. [Google Scholar] [CrossRef]
  24. Flint, S.J.; Enquist, L.W.; Racaniello, V.R.; Rall, G.F.; Skalka, A.M. Principal of Virology; ASM Press: Washington, DC, USA, 2015; Volume 1. [Google Scholar]
  25. Kanai, R.; Zaupa, C.; Sgubin, D.; Antoszczyk, S.J.; Martuza, R.L.; Wakimoto, H.; Rabkin, S.D. Effect of gamma34.5 deletions on oncolytic herpes simplex virus activity in brain tumors. J. Virol. 2012, 86, 4420–4431. [Google Scholar] [CrossRef] [PubMed]
  26. Nakano, K.; Asano, R.; Tsumoto, K.; Kwon, H.; Goins, W.F.; Kumagai, I.; Cohen, J.B.; Glorioso, J.C. Herpes simplex virus targeting to the EGF receptor by a gD-specific soluble bridging molecule. Mol. Ther. 2005, 11, 617–626. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, M.; Yang, C.S.; Guo, W.; Tang, J.; Huang, Q.; Feng, S.; Jiang, A.; Xu, X.; Jiang, G.; Liu, Y.Q. A novel fiber chimeric conditionally replicative adenovirus-Ad5/F35 for tumor therapy. Cancer Biol. Ther. 2017, 18, 833–840. [Google Scholar] [CrossRef]
  28. Loya, S.M.; Zhang, X. Enhancing the bystander killing effect of an oncolytic HSV by arming it with a secretable apoptosis activator. Gene Ther. 2015, 22, 237–246. [Google Scholar] [CrossRef]
  29. Xie, S.; Fan, W.; Yang, C.; Lei, W.; Pan, H.; Tong, X.; Wu, Y.; Wang, S. Beclin1-armed oncolytic Vaccinia virus enhances the therapeutic efficacy of R-CHOP against lymphoma in vitro and in vivo. Oncol. Rep. 2021, 45, 987–996. [Google Scholar] [CrossRef]
  30. Chouljenko, D.V.; Ding, J.; Lee, I.F.; Murad, Y.M.; Bu, X.; Liu, G.; Delwar, Z.; Sun, Y.; Yu, S.; Samudio, I.; et al. Induction of Durable Antitumor Response by a Novel Oncolytic Herpesvirus Expressing Multiple Immunomodulatory Transgenes. Biomedicines 2020, 8, 484. [Google Scholar] [CrossRef]
  31. Bourgeois-Daigneault, M.C.; Roy, D.G.; Falls, T.; Twumasi-Boateng, K.; St-Germain, L.E.; Marguerie, M.; Garcia, V.; Selman, M.; Jennings, V.A.; Pettigrew, J.; et al. Oncolytic vesicular stomatitis virus expressing interferon-gamma has enhanced therapeutic activity. Mol. Ther. Oncolytics 2016, 3, 16001. [Google Scholar] [CrossRef]
  32. Liu, Z.; Ravindranathan, R.; Li, J.; Kalinski, P.; Guo, Z.S.; Bartlett, D.L. CXCL11-Armed oncolytic poxvirus elicits potent antitumor immunity and shows enhanced therapeutic efficacy. Oncoimmunology 2016, 5, e1091554. [Google Scholar] [CrossRef]
  33. Hietanen, E.; Koivu, M.K.A.; Susi, P. Cytolytic Properties and Genome Analysis of Rigvir((R)) Oncolytic Virotherapy Virus and Other Echovirus 7 Isolates. Viruses 2022, 14, 525. [Google Scholar] [CrossRef]
  34. Svyatchenko, V.A.; Ternovoy, V.A.; Kiselev, N.N.; Demina, A.V.; Loktev, V.B.; Netesov, S.V.; Chumakov, P.M. Bioselection of coxsackievirus B6 strain variants with altered tropism to human cancer cell lines. Arch. Virol. 2017, 162, 3355–3362. [Google Scholar] [CrossRef]
  35. Alekseeva, O.; Gumennaya, Y.; Naberezhnaya, E.; Kushchenko, A.; Dmitriev, S.; Chumakov, P.; Lipatova, A. B37. An integrated approach reveals oncolytic activity and novel receptor requirements of bioselected echovirus variants. In Proceedings of the Viruses 2024—A World of Viruses, Barcelona, Spain, 14–16 February 2024; p. 189. [Google Scholar]
  36. Zhang, P.; Mueller, S.; Morais, M.C.; Bator, C.M.; Bowman, V.D.; Hafenstein, S.; Wimmer, E.; Rossmann, M.G. Crystal structure of CD155 and electron microscopic studies of its complexes with polioviruses. Proc. Natl. Acad. Sci. USA 2008, 105, 18284–18289. [Google Scholar] [CrossRef] [PubMed]
  37. Belnap, D.M.; McDermott, B.M., Jr.; Filman, D.J.; Cheng, N.; Trus, B.L.; Zuccola, H.J.; Racaniello, V.R.; Hogle, J.M.; Steven, A.C. Three-dimensional structure of poliovirus receptor bound to poliovirus. Proc. Natl. Acad. Sci. USA 2000, 97, 73–78. [Google Scholar] [CrossRef] [PubMed]
  38. Brandenburg, B.; Lee, L.Y.; Lakadamyali, M.; Rust, M.J.; Zhuang, X.; Hogle, J.M. Imaging poliovirus entry in live cells. PLoS Biol. 2007, 5, e183. [Google Scholar] [CrossRef]
  39. Zhand, S.; Hosseini, S.M.; Tabarraei, A.; Moradi, A.; Saeidi, M. Analysis of poliovirus receptor, CD155 expression in different human colorectal cancer cell lines: Implications for poliovirus virotherapy. J. Cancer Res. Ther. 2019, 15, 61–67. [Google Scholar] [CrossRef]
  40. Bowers, J.R.; Readler, J.M.; Sharma, P.; Excoffon, K. Poliovirus Receptor: More than a simple viral receptor. Virus Res. 2017, 242, 1–6. [Google Scholar] [CrossRef]
  41. Racaniello, V.R. Early events in poliovirus infection: Virus-receptor interactions. Proc. Natl. Acad. Sci. USA 1996, 93, 11378–11381. [Google Scholar] [CrossRef]
  42. Jackson, W.T.; Coyne, C.B. Enteroviruses: Omics, Molecular Biology, and Control; Caister Academic Press: Poole, UK, 2018. [Google Scholar]
  43. Chen, X.; Qu, X.; Liu, C.; Zhang, Y.; Zhang, G.; Han, P.; Duan, Y.; Li, Q.; Wang, L.; Ruan, W.; et al. Human FcRn Is a Two-in-One Attachment-Uncoating Receptor for Echovirus 18. mBio 2022, 13, e0116622. [Google Scholar] [CrossRef]
  44. Vandesande, H.; Laajala, M.; Kantoluoto, T.; Ruokolainen, V.; Lindberg, A.M.; Marjomaki, V. Early Entry Events in Echovirus 30 Infection. J. Virol. 2020, 94, 10–1128. [Google Scholar] [CrossRef]
  45. Lin, Y.W.; Lin, H.Y.; Tsou, Y.L.; Chitra, E.; Hsiao, K.N.; Shao, H.Y.; Liu, C.C.; Sia, C.; Chong, P.; Chow, Y.H. Human SCARB2-mediated entry and endocytosis of EV71. PLoS ONE 2012, 7, e30507. [Google Scholar] [CrossRef]
  46. Zhou, S.; Liu, Q.; Wu, X.; Chen, P.; Wu, X.; Guo, Y.; Liu, S.; Liang, Z.; Fan, C.; Wang, Y. A safe and sensitive enterovirus A71 infection model based on human SCARB2 knock-in mice. Vaccine 2016, 34, 2729–2736. [Google Scholar] [CrossRef]
  47. Yu, S.L.; Chung, N.H.; Lin, Y.C.; Liao, Y.A.; Chen, Y.C.; Chow, Y.H. Human SCARB2 Acts as a Cellular Associator for Helping Coxsackieviruses A10 Infection. Viruses 2023, 15, 932. [Google Scholar] [CrossRef] [PubMed]
  48. Yamayoshi, S.; Iizuka, S.; Yamashita, T.; Minagawa, H.; Mizuta, K.; Okamoto, M.; Nishimura, H.; Sanjoh, K.; Katsushima, N.; Itagaki, T.; et al. Human SCARB2-dependent infection by coxsackievirus A7, A14, and A16 and enterovirus 71. J. Virol. 2012, 86, 5686–5696. [Google Scholar] [CrossRef] [PubMed]
  49. Marjomaki, V.; Pietiainen, V.; Matilainen, H.; Upla, P.; Ivaska, J.; Nissinen, L.; Reunanen, H.; Huttunen, P.; Hyypia, T.; Heino, J. Internalization of echovirus 1 in caveolae. J. Virol. 2002, 76, 1856–1865. [Google Scholar] [CrossRef] [PubMed]
  50. Pulli, T.; Koivunen, E.; Hyypia, T. Cell-surface interactions of echovirus 22. J. Biol. Chem. 1997, 272, 21176–21180. [Google Scholar] [CrossRef]
  51. Merilahti, P.; Koskinen, S.; Heikkila, O.; Karelehto, E.; Susi, P. Endocytosis of integrin-binding human picornaviruses. Adv. Virol. 2012, 2012, 547530. [Google Scholar] [CrossRef]
  52. Israelsson, S.; Gullberg, M.; Jonsson, N.; Roivainen, M.; Edman, K.; Lindberg, A.M. Studies of Echovirus 5 interactions with the cell surface: Heparan sulfate mediates attachment to the host cell. Virus Res. 2010, 151, 170–176. [Google Scholar] [CrossRef]
  53. Bergelson, J.M.; St John, N.; Kawaguchi, S.; Chan, M.; Stubdal, H.; Modlin, J.; Finberg, R.W. Infection by echoviruses 1 and 8 depends on the alpha 2 subunit of human VLA-2. J. Virol. 1993, 67, 6847–6852. [Google Scholar] [CrossRef]
  54. Staring, J.; van den Hengel, L.G.; Raaben, M.; Blomen, V.A.; Carette, J.E.; Brummelkamp, T.R. KREMEN1 Is a Host Entry Receptor for a Major Group of Enteroviruses. Cell Host Microbe 2018, 23, 636–643.e635. [Google Scholar] [CrossRef]
  55. Zhao, Y.; Zhou, D.; Ni, T.; Karia, D.; Kotecha, A.; Wang, X.; Rao, Z.; Jones, E.Y.; Fry, E.E.; Ren, J.; et al. Hand-foot-and-mouth disease virus receptor KREMEN1 binds the canyon of Coxsackie Virus A10. Nat. Commun. 2020, 11, 38. [Google Scholar] [CrossRef]
  56. Xiao, C.; Bator, C.M.; Bowman, V.D.; Rieder, E.; He, Y.; Hebert, B.; Bella, J.; Baker, T.S.; Wimmer, E.; Kuhn, R.J.; et al. Interaction of coxsackievirus A21 with its cellular receptor, ICAM-1. J. Virol. 2001, 75, 2444–2451. [Google Scholar] [CrossRef]
  57. Xiao, C.; Bator-Kelly, C.M.; Rieder, E.; Chipman, P.R.; Craig, A.; Kuhn, R.J.; Wimmer, E.; Rossmann, M.G. The crystal structure of coxsackievirus A21 and its interaction with ICAM-1. Structure 2005, 13, 1019–1033. [Google Scholar] [CrossRef] [PubMed]
  58. Selinka, H.C.; Wolde, A.; Sauter, M.; Kandolf, R.; Klingel, K. Virus-receptor interactions of coxsackie B viruses and their putative influence on cardiotropism. Med. Microbiol. Immunol. 2004, 193, 127–131. [Google Scholar] [CrossRef] [PubMed]
  59. Fu, Y.; Xiong, S. Exosomes mediate Coxsackievirus B3 transmission and expand the viral tropism. PLoS Pathog. 2023, 19, e1011090. [Google Scholar] [CrossRef] [PubMed]
  60. Excoffon, K. The coxsackievirus and adenovirus receptor: Virological and biological beauty. FEBS Lett. 2020, 594, 1828–1837. [Google Scholar] [CrossRef]
  61. He, Y.; Chipman, P.R.; Howitt, J.; Bator, C.M.; Whitt, M.A.; Baker, T.S.; Kuhn, R.J.; Anderson, C.W.; Freimuth, P.; Rossmann, M.G. Interaction of coxsackievirus B3 with the full length coxsackievirus-adenovirus receptor. Nat. Struct. Biol. 2001, 8, 874–878. [Google Scholar] [CrossRef]
  62. Crowell, R.L.; Landau, B.J.; Philipson, L. The early interaction of coxsackievirus B3 with HeLa cells. Proc. Soc. Exp. Biol. Med. 1971, 137, 1082–1088. [Google Scholar] [CrossRef]
  63. Plevka, P.; Hafenstein, S.; Harris, K.G.; Cifuente, J.O.; Zhang, Y.; Bowman, V.D.; Chipman, P.R.; Bator, C.M.; Lin, F.; Medof, M.E.; et al. Interaction of decay-accelerating factor with echovirus 7. J. Virol. 2010, 84, 12665–12674. [Google Scholar] [CrossRef]
  64. Hafenstein, S.; Bowman, V.D.; Chipman, P.R.; Bator Kelly, C.M.; Lin, F.; Medof, M.E.; Rossmann, M.G. Interaction of decay-accelerating factor with coxsackievirus B3. J. Virol. 2007, 81, 12927–12935. [Google Scholar] [CrossRef]
  65. Karnauchow, T.M.; Tolson, D.L.; Harrison, B.A.; Altman, E.; Lublin, D.M.; Dimock, K. The HeLa cell receptor for enterovirus 70 is decay-accelerating factor (CD55). J. Virol. 1996, 70, 5143–5152. [Google Scholar] [CrossRef]
  66. Shieh, J.T.; Bergelson, J.M. Interaction with decay-accelerating factor facilitates coxsackievirus B infection of polarized epithelial cells. J. Virol. 2002, 76, 9474–9480. [Google Scholar] [CrossRef]
  67. Newcombe, N.G.; Beagley, L.G.; Christiansen, D.; Loveland, B.E.; Johansson, E.S.; Beagley, K.W.; Barry, R.D.; Shafren, D.R. Novel role for decay-accelerating factor in coxsackievirus A21-mediated cell infectivity. J. Virol. 2004, 78, 12677–12682. [Google Scholar] [CrossRef] [PubMed]
  68. He, Y.; Lin, F.; Chipman, P.R.; Bator, C.M.; Baker, T.S.; Shoham, M.; Kuhn, R.J.; Medof, M.E.; Rossmann, M.G. Structure of decay-accelerating factor bound to echovirus 7: A virus-receptor complex. Proc. Natl. Acad. Sci. USA 2002, 99, 10325–10329. [Google Scholar] [CrossRef] [PubMed]
  69. Shafren, D.R.; Dorahy, D.J.; Ingham, R.A.; Burns, G.F.; Barry, R.D. Coxsackievirus A21 binds to decay-accelerating factor but requires intercellular adhesion molecule 1 for cell entry. J. Virol. 1997, 71, 4736–4743. [Google Scholar] [CrossRef]
  70. Rezaikin, A.V.; Novoselov, A.V.; Sergeev, A.G.; Fadeyev, F.A.; Lebedev, S.V. Two clusters of mutations map distinct receptor-binding sites of echovirus 11 for the decay-accelerating factor (CD55) and for canyon-binding receptors. Virus Res. 2009, 145, 74–79. [Google Scholar] [CrossRef] [PubMed]
  71. Shafren, D.R.; Bates, R.C.; Agrez, M.V.; Herd, R.L.; Burns, G.F.; Barry, R.D. Coxsackieviruses B1, B3, and B5 use decay accelerating factor as a receptor for cell attachment. J. Virol. 1995, 69, 3873–3877. [Google Scholar] [CrossRef]
  72. Bhella, D.; Goodfellow, I.G.; Roversi, P.; Pettigrew, D.; Chaudhry, Y.; Evans, D.J.; Lea, S.M. The structure of echovirus type 12 bound to a two-domain fragment of its cellular attachment protein decay-accelerating factor (CD 55). J. Biol. Chem. 2004, 279, 8325–8332. [Google Scholar] [CrossRef]
  73. Solomon, B.L.; Garrido-Laguna, I. TIGIT: A novel immunotherapy target moving from bench to bedside. Cancer Immunol. Immunother. 2018, 67, 1659–1667. [Google Scholar] [CrossRef]
  74. Yu, X.; Harden, K.; Gonzalez, L.C.; Francesco, M.; Chiang, E.; Irving, B.; Tom, I.; Ivelja, S.; Refino, C.J.; Clark, H.; et al. The surface protein TIGIT suppresses T cell activation by promoting the generation of mature immunoregulatory dendritic cells. Nat. Immunol. 2009, 10, 48–57. [Google Scholar] [CrossRef]
  75. Zhou, X.; Du, J.; Wang, H.; Chen, C.; Jiao, L.; Cheng, X.; Zhou, X.; Chen, S.; Gou, S.; Zhao, W.; et al. Repositioning liothyronine for cancer immunotherapy by blocking the interaction of immune checkpoint TIGIT/PVR. Cell Commun. Signal. 2020, 18, 142. [Google Scholar] [CrossRef]
  76. Coyne, C.B.; Kim, K.S.; Bergelson, J.M. Poliovirus entry into human brain microvascular cells requires receptor-induced activation of SHP-2. EMBO J. 2007, 26, 4016–4028. [Google Scholar] [CrossRef]
  77. Stengel, K.F.; Harden-Bowles, K.; Yu, X.; Rouge, L.; Yin, J.; Comps-Agrar, L.; Wiesmann, C.; Bazan, J.F.; Eaton, D.L.; Grogan, J.L. Structure of TIGIT immunoreceptor bound to poliovirus receptor reveals a cell-cell adhesion and signaling mechanism that requires cis-trans receptor clustering. Proc. Natl. Acad. Sci. USA 2012, 109, 5399–5404. [Google Scholar] [CrossRef] [PubMed]
  78. Koike, S.; Horie, H.; Ise, I.; Okitsu, A.; Yoshida, M.; Iizuka, N.; Takeuchi, K.; Takegami, T.; Nomoto, A. The poliovirus receptor protein is produced both as membrane-bound and secreted forms. EMBO J. 1990, 9, 3217–3224. [Google Scholar] [CrossRef] [PubMed]
  79. Ylipaasto, P.; Eskelinen, M.; Salmela, K.; Hovi, T.; Roivainen, M. Vitronectin receptors, alpha v integrins, are recognized by several non-RGD-containing echoviruses in a continuous laboratory cell line and also in primary human Langerhans’ islets and endothelial cells. J. Gen. Virol. 2010, 91, 155–165. [Google Scholar] [CrossRef]
  80. Xing, L.; Huhtala, M.; Pietiainen, V.; Kapyla, J.; Vuorinen, K.; Marjomaki, V.; Heino, J.; Johnson, M.S.; Hyypia, T.; Cheng, R.H. Structural and functional analysis of integrin alpha2I domain interaction with echovirus 1. J. Biol. Chem. 2004, 279, 11632–11638. [Google Scholar] [CrossRef]
  81. Pietiäinen, V.; Marjomäki, V.; Upla, P.; Pelkmans, L.; Helenius, A.; Hyypiä, T. Echovirus 1 endocytosis into caveosomes requires lipid rafts, dynamin II, and signaling events. Mol. Biol. Cell 2004, 15, 4911–4925. [Google Scholar] [CrossRef]
  82. Grant, R.A.; Hiremath, C.N.; Filman, D.J.; Syed, R.; Andries, K.; Hogle, J.M. Structures of poliovirus complexes with anti-viral drugs: Implications for viral stability and drug design. Curr. Biol. 1994, 4, 784–797. [Google Scholar] [CrossRef]
  83. Strauss, M.; Filman, D.J.; Belnap, D.M.; Cheng, N.; Noel, R.T.; Hogle, J.M. Nectin-like interactions between poliovirus and its receptor trigger conformational changes associated with cell entry. J. Virol. 2015, 89, 4143–4157. [Google Scholar] [CrossRef]
  84. Hogle, J.M. Poliovirus cell entry: Common structural themes in viral cell entry pathways. Annu. Rev. Microbiol. 2002, 56, 677–702. [Google Scholar] [CrossRef]
  85. Kim, H.S.; Lee, K.; Bae, S.; Park, J.; Lee, C.K.; Kim, M.; Kim, E.; Kim, M.; Kim, S.; Kim, C.; et al. CRISPR/Cas9-mediated gene knockout screens and target identification via whole-genome sequencing uncover host genes required for picornavirus infection. J. Biol. Chem. 2017, 292, 10664–10671. [Google Scholar] [CrossRef]
  86. Nandi, S.S.; Sawant, S.; Gohil, T.; Lambe, U.; Sangal, L.; Patel, D.; Krishnasamy, K.; Ghoshal, U.; Harvey, P.; Deshpande, J. Poliovirus nonpermissive CD155 knockout cells derived from RD cell line for handling poliovirus potentially infectious materials in virology laboratories. J. Med. Virol. 2022, 94, 4901–4909. [Google Scholar] [CrossRef]
  87. Sosnovtseva, A.O.; Lipatova, A.V.; Grinenko, N.F.; Baklaushev, V.P.; Chumakov, P.M.; Chekhonin, V.P. Sensitivity of C6 Glioma Cells Carrying the Human Poliovirus Receptor to Oncolytic Polioviruses. Bull. Exp. Biol. Med. 2016, 161, 821–825. [Google Scholar] [CrossRef] [PubMed]
  88. van den Braak, W.J.P.; Monica, B.; Limpens, D.; Rockx-Brouwer, D.; de Boer, M.; Oosterhoff, D. Construction of a Vero Cell Line Expressing Human ICAM1 for the Development of Rhinovirus Vaccines. Viruses 2022, 14, 2235. [Google Scholar] [CrossRef] [PubMed]
  89. Shukla, S.D.; Shastri, M.D.; Vanka, S.K.; Jha, N.K.; Dureja, H.; Gupta, G.; Chellappan, D.V.K.; Oliver, B.G.; Dua, K.; Walters, E.H. Targeting intercellular adhesion molecule-1 (ICAM-1) to reduce rhinovirus-induced acute exacerbations in chronic respiratory diseases. Inflammopharmacology 2022, 30, 725–735. [Google Scholar] [CrossRef] [PubMed]
  90. Dang, M.; Wang, X.; Wang, Q.; Wang, Y.; Lin, J.; Sun, Y.; Li, X.; Zhang, L.; Lou, Z.; Wang, J.; et al. Molecular mechanism of SCARB2-mediated attachment and uncoating of EV71. Protein Cell 2014, 5, 692–703. [Google Scholar] [CrossRef]
  91. Yamayoshi, S.; Koike, S. Identification of a human SCARB2 region that is important for enterovirus 71 binding and infection. J. Virol. 2011, 85, 4937–4946. [Google Scholar] [CrossRef]
  92. Kuronita, T.; Eskelinen, E.L.; Fujita, H.; Saftig, P.; Himeno, M.; Tanaka, Y. A role for the lysosomal membrane protein LGP85 in the biogenesis and maintenance of endosomal and lysosomal morphology. J. Cell Sci. 2002, 115, 4117–4131. [Google Scholar] [CrossRef]
  93. Eskelinen, E.L.; Tanaka, Y.; Saftig, P. At the acidic edge: Emerging functions for lysosomal membrane proteins. Trends Cell Biol. 2003, 13, 137–145. [Google Scholar] [CrossRef]
  94. Heybrock, S.; Kanerva, K.; Meng, Y.; Ing, C.; Liang, A.; Xiong, Z.J.; Weng, X.; Ah Kim, Y.; Collins, R.; Trimble, W.; et al. Lysosomal integral membrane protein-2 (LIMP-2/SCARB2) is involved in lysosomal cholesterol export. Nat. Commun. 2019, 10, 3521. [Google Scholar] [CrossRef]
  95. Berkovic, S.F.; Dibbens, L.M.; Oshlack, A.; Silver, J.D.; Katerelos, M.; Vears, D.F.; Lullmann-Rauch, R.; Blanz, J.; Zhang, K.W.; Stankovich, J.; et al. Array-based gene discovery with three unrelated subjects shows SCARB2/LIMP-2 deficiency causes myoclonus epilepsy and glomerulosclerosis. Am. J. Hum. Genet. 2008, 82, 673–684. [Google Scholar] [CrossRef]
  96. Velayati, A.; DePaolo, J.; Gupta, N.; Choi, J.H.; Moaven, N.; Westbroek, W.; Goker-Alpan, O.; Goldin, E.; Stubblefield, B.K.; Kolodny, E.; et al. A mutation in SCARB2 is a modifier in Gaucher disease. Hum. Mutat. 2011, 32, 1232–1238. [Google Scholar] [CrossRef]
  97. Zou, Y.; Pei, J.; Wang, Y.; Chen, Q.; Sun, M.; Kang, L.; Zhang, X.; Zhang, L.; Gao, X.; Lin, Z. The Deficiency of SCARB2/LIMP-2 Impairs Metabolism via Disrupted mTORC1-Dependent Mitochondrial OXPHOS. Int. J. Mol. Sci. 2022, 23, 8634. [Google Scholar] [CrossRef] [PubMed]
  98. Fujita, H.; Takata, Y.; Kono, A.; Tanaka, Y.; Takahashi, T.; Himeno, M.; Kato, K. Isolation and sequencing of a cDNA clone encoding the 85 kDa human lysosomal sialoglycoprotein (hLGP85) in human metastatic pancreas islet tumor cells. Biochem. Biophys. Res. Commun. 1992, 184, 604–611. [Google Scholar] [CrossRef] [PubMed]
  99. Chen, P.; Song, Z.; Qi, Y.; Feng, X.; Xu, N.; Sun, Y.; Wu, X.; Yao, X.; Mao, Q.; Li, X.; et al. Molecular determinants of enterovirus 71 viral entry: Cleft around GLN-172 on VP1 protein interacts with variable region on scavenge receptor B 2. J. Biol. Chem. 2012, 287, 6406–6420. [Google Scholar] [CrossRef] [PubMed]
  100. Zhou, D.; Zhao, Y.; Kotecha, A.; Fry, E.E.; Kelly, J.T.; Wang, X.; Rao, Z.; Rowlands, D.J.; Ren, J.; Stuart, D.I. Unexpected mode of engagement between enterovirus 71 and its receptor SCARB2. Nat. Microbiol. 2019, 4, 414–419. [Google Scholar] [CrossRef]
  101. Zhang, X.; Yang, P.; Wang, N.; Zhang, J.; Li, J.; Guo, H.; Yin, X.; Rao, Z.; Wang, X.; Zhang, L. The binding of a monoclonal antibody to the apical region of SCARB2 blocks EV71 infection. Protein Cell 2017, 8, 590–600. [Google Scholar] [CrossRef]
  102. Ke, X.; Li, C.; Luo, D.; Wang, T.; Liu, Y.; Tan, Z.; Du, M.; He, Z.; Wang, H.; Zheng, Z.; et al. Metabolic labeling of enterovirus 71 with quantum dots for the study of virus receptor usage. J. Nanobiotechnol. 2021, 19, 295. [Google Scholar] [CrossRef]
  103. Yamayoshi, S.; Yamashita, Y.; Li, J.; Hanagata, N.; Minowa, T.; Takemura, T.; Koike, S. Scavenger receptor B2 is a cellular receptor for enterovirus 71. Nat. Med. 2009, 15, 798–801. [Google Scholar] [CrossRef]
  104. Jin, Y.; Sun, T.; Zhou, G.; Li, D.; Chen, S.; Zhang, W.; Li, X.; Zhang, R.; Yang, H.; Duan, G. Pathogenesis study of enterovirus 71 using a novel human SCARB2 knock-in mouse model. mSphere 2021, 6, 10–1128. [Google Scholar] [CrossRef]
  105. Lin, Y.-W.; Yu, S.-L.; Shao, H.-Y.; Lin, H.-Y.; Liu, C.-C.; Hsiao, K.-N.; Chitra, E.; Tsou, Y.-L.; Chang, H.-W.; Sia, C. Human SCARB2 transgenic mice as an infectious animal model for enterovirus 71. PLoS ONE 2013, 8, e57591. [Google Scholar] [CrossRef]
  106. Guo, D.; Yu, X.; Wang, D.; Li, Z.; Zhou, Y.; Xu, G.; Yuan, B.; Qin, Y.; Chen, M. SLC35B2 acts in a dual role in the host sulfation required for EV71 infection. J. Virol. 2022, 96, e02042-e21. [Google Scholar] [CrossRef]
  107. Hynes, R.O. Integrins: Versatility, modulation, and signaling in cell adhesion. Cell 1992, 69, 11–25. [Google Scholar] [CrossRef] [PubMed]
  108. Barczyk, M.; Carracedo, S.; Gullberg, D. Integrins. Cell Tissue Res. 2010, 339, 269–280. [Google Scholar] [CrossRef] [PubMed]
  109. Arnaout, M.A.; Goodman, S.L.; Xiong, J.P. Structure and mechanics of integrin-based cell adhesion. Curr. Opin. Cell Biol. 2007, 19, 495–507. [Google Scholar] [CrossRef]
  110. Campbell, I.D.; Humphries, M.J. Integrin structure, activation, and interactions. Cold Spring Harb. Perspect. Biol. 2011, 3, a004994. [Google Scholar] [CrossRef]
  111. King, S.L.; Kamata, T.; Cunningham, J.A.; Emsley, J.; Liddington, R.C.; Takada, Y.; Bergelson, J.M. Echovirus 1 interaction with the human very late antigen-2 (integrin alpha2beta1) I domain. Identification of two independent virus contact sites distinct from the metal ion-dependent adhesion site. J. Biol. Chem. 1997, 272, 28518–28522. [Google Scholar] [CrossRef]
  112. Williams, C.H.; Kajander, T.; Hyypia, T.; Jackson, T.; Sheppard, D.; Stanway, G. Integrin alpha v beta 6 is an RGD-dependent receptor for coxsackievirus A9. J. Virol. 2004, 78, 6967–6973. [Google Scholar] [CrossRef]
  113. Triantafilou, M.; Wilson, K.M.; Triantafilou, K. Identification of Echovirus 1 and coxsackievirus A9 receptor molecules via a novel flow cytometric quantification method. Cytometry 2001, 43, 279–289. [Google Scholar] [CrossRef]
  114. de Bournonville, S.; Vangrunderbeeck, S.; Kerckhofs, G. Contrast-Enhanced MicroCT for Virtual 3D Anatomical Pathology of Biological Tissues: A Literature Review. Contrast Media Mol. Imaging 2019, 2019, 8617406. [Google Scholar] [CrossRef]
  115. Coyne, C.B.; Bergelson, J.M. CAR: A virus receptor within the tight junction. Adv. Drug Deliv. Rev. 2005, 57, 869–882. [Google Scholar] [CrossRef]
  116. Raschperger, E.; Thyberg, J.; Pettersson, S.; Philipson, L.; Fuxe, J.; Pettersson, R.F. The coxsackie- and adenovirus receptor (CAR) is an in vivo marker for epithelial tight junctions, with a potential role in regulating permeability and tissue homeostasis. Exp. Cell Res. 2006, 312, 1566–1580. [Google Scholar] [CrossRef]
  117. Tomko, R.P.; Xu, R.; Philipson, L. HCAR and MCAR: The human and mouse cellular receptors for subgroup C adenoviruses and group B coxsackieviruses. Proc. Natl. Acad. Sci. USA 1997, 94, 3352–3356. [Google Scholar] [CrossRef] [PubMed]
  118. Ellis, B.L.; Potts, P.R.; Porteus, M.H. Creating higher titer lentivirus with caffeine. Hum. Gene Ther. 2011, 22, 93–100. [Google Scholar] [CrossRef] [PubMed]
  119. Honda, T.; Saitoh, H.; Masuko, M.; Katagiri-Abe, T.; Tominaga, K.; Kozakai, I.; Kobayashi, K.; Kumanishi, T.; Watanabe, Y.D.G.; Odani, S.; et al. The coxsackievirus-adenovirus receptor protein as a cell adhesion molecule in the developing mouse brain. Mol. Brain Res. 2000, 77, 19–28. [Google Scholar] [CrossRef]
  120. Hotta, Y.; Honda, T.; Naito, M.; Kuwano, R. Developmental distribution of coxsackie virus and adenovirus receptor localized in the nervous system. Brain Research. Dev. Brain Res. 2003, 143, 1–13. [Google Scholar] [CrossRef]
  121. Nalbantoglu, J.; Pari, G.; Karpati, G.; Holland, P.M.C. Expression of the primary coxsackie and adenovirus receptor is downregulated during skeletal muscle maturation and limits the efficacy of adenovirus-mediated gene delivery to muscle cells. Hum. Gene Ther. 1999, 10, 1009–1019. [Google Scholar] [CrossRef]
  122. Bergelson, J.M.; Cunningham, J.A.; Droguett, G.; Kurt-Jones, E.A.; Krithivas, A.; Hong, J.S.; Horwitz, M.S.; Crowell, R.L.; Finberg, R.W. Isolation of a common receptor for Coxsackie B viruses and adenoviruses 2 and 5. Science 1997, 275, 1320–1323. [Google Scholar] [CrossRef]
  123. Raschperger, E.; Engstrom, U.; Pettersson, R.F.; Fuxe, J. CLMP, a novel member of the CTX family and a new component of epithelial tight junctions. J. Biol. Chem. 2004, 279, 796–804. [Google Scholar] [CrossRef]
  124. Carson, S.D.; Chapman, N.N.; Tracy, S.M. Purification of the putative coxsackievirus B receptor from HeLa cells. Biochem. Biophys. Res. Commun. 1997, 233, 325–328. [Google Scholar] [CrossRef]
  125. Lipatova, A.V.; Le, T.H.; Sosnovtseva, A.O.; Babaeva, F.E.; Kochetkov, D.V.; Chumakov, P.M. Relationship between Cell Receptors and Tumor Cell Sensitivity to Oncolytic Enteroviruses. Bull. Exp. Biol. Med. 2018, 166, 58–62. [Google Scholar] [CrossRef]
  126. Triantafilou, K.; Triantafilou, M. Lipid-raft-dependent Coxsackievirus B4 internalization and rapid targeting to the Golgi. Virology 2004, 326, 6–19. [Google Scholar] [CrossRef]
  127. Chung, S.K.; Kim, J.Y.; Kim, I.B.; Park, S.I.; Paek, K.H.; Nam, J.H. Internalization and trafficking mechanisms of coxsackievirus B3 in HeLa cells. Virology 2005, 333, 31–40. [Google Scholar] [CrossRef] [PubMed]
  128. Lublin, D.M.; Atkinson, J.P. Decay-accelerating factor: Biochemistry, molecular biology, and function. Annu. Rev. Immunol. 1989, 7, 35–58. [Google Scholar] [CrossRef] [PubMed]
  129. Nicholson-Weller, A.; Burge, J.; Fearon, D.T.; Weller, P.F.; Austen, K.F. Isolation of a human erythrocyte membrane glycoprotein with decay-accelerating activity for C3 convertases of the complement system. J. Immunol. 1982, 129, 184–189. [Google Scholar] [CrossRef] [PubMed]
  130. Lukacik, P.; Roversi, P.; White, J.; Esser, D.; Smith, G.P.; Billington, J.; Williams, P.A.; Rudd, P.M.; Wormald, M.R.; Harvey, D.J.; et al. Complement regulation at the molecular level: The structure of decay-accelerating factor. Proc. Natl. Acad. Sci. USA 2004, 101, 1279–1284. [Google Scholar] [CrossRef]
  131. Low, M.G. Biochemistry of the glycosyl-phosphatidylinositol membrane protein anchors. Biochem. J. 1987, 244, 1–13. [Google Scholar] [CrossRef]
  132. Koretz, K.; Bruderlein, S.; Henne, C.; Moller, P. Decay-accelerating factor (DAF, CD55) in normal colorectal mucosa, adenomas and carcinomas. Br. J. Cancer 1992, 66, 810–814. [Google Scholar] [CrossRef]
  133. Bharti, R.; Dey, G.; Lin, F.; Lathia, J.; Reizes, O. CD55 in cancer: Complementing functions in a non-canonical manner. Cancer Lett. 2022, 551, 215935. [Google Scholar] [CrossRef]
  134. Clarkson, N.A.; Kaufman, R.; Lublin, D.M.; Ward, T.; Pipkin, P.A.; Minor, P.D.; Evans, D.J.; Almond, J.W. Characterization of the echovirus 7 receptor: Domains of CD55 critical for virus binding. J. Virol. 1995, 69, 5497–5501. [Google Scholar] [CrossRef]
  135. Shafren, D.R. Viral cell entry induced by cross-linked decay-accelerating factor. J. Virol. 1998, 72, 9407–9412. [Google Scholar] [CrossRef]
  136. Coyne, C.B.; Bergelson, J.M. Virus-induced Abl and Fyn kinase signals permit coxsackievirus entry through epithelial tight junctions. Cell 2006, 124, 119–131. [Google Scholar] [CrossRef]
  137. Stuart, A.D.; Eustace, H.E.; McKee, T.A.; Brown, T.D. A novel cell entry pathway for a DAF-using human enterovirus is dependent on lipid rafts. J. Virol. 2002, 76, 9307–9322. [Google Scholar] [CrossRef] [PubMed]
  138. Wang, Q.; Yang, Q.; Liu, C.; Wang, G.; Song, H.; Shang, G.; Peng, R.; Qu, X.; Liu, S.; Cui, Y.; et al. Molecular basis of differential receptor usage for naturally occurring CD55-binding and -nonbinding coxsackievirus B3 strains. Proc. Natl. Acad. Sci. USA 2022, 119, e2118590119. [Google Scholar] [CrossRef] [PubMed]
  139. Cohen, C.J.; Shieh, J.T.; Pickles, R.J.; Okegawa, T.; Hsieh, J.T.; Bergelson, J.M. The coxsackievirus and adenovirus receptor is a transmembrane component of the tight junction. Proc. Natl. Acad. Sci. USA 2001, 98, 15191–15196. [Google Scholar] [CrossRef] [PubMed]
  140. Inoue, T.; Yamakawa, M.; Takahashi, T. Expression of complement regulating factors in gastric cancer cells. Mol. Pathol. 2002, 55, 193–199. [Google Scholar] [CrossRef]
  141. Bartosch, B.; Vitelli, A.; Granier, C.; Goujon, C.; Dubuisson, J.; Pascale, S.; Scarselli, E.; Cortese, R.; Nicosia, A.; Cosset, F.L. Cell entry of hepatitis C virus requires a set of co-receptors that include the CD81 tetraspanin and the SR-B1 scavenger receptor. J. Biol. Chem. 2003, 278, 41624–41630. [Google Scholar] [CrossRef]
  142. Liu, S.; Yang, W.; Shen, L.; Turner, J.R.; Coyne, C.B.; Wang, T. Tight junction proteins claudin-1 and occludin control hepatitis C virus entry and are downregulated during infection to prevent superinfection. J. Virol. 2009, 83, 2011–2014. [Google Scholar] [CrossRef]
  143. Newcombe, N.G.; Johansson, E.S.; Au, G.; Lindberg, A.M.; Barry, R.D.; Shafren, D.R. Enterovirus capsid interactions with decay-accelerating factor mediate lytic cell infection. J. Virol. 2004, 78, 1431–1439. [Google Scholar] [CrossRef]
  144. Johansson, E.S.; Xing, L.; Cheng, R.H.; Shafren, D.R. Enhanced cellular receptor usage by a bioselected variant of coxsackievirus a21. J. Virol. 2004, 78, 12603–12612. [Google Scholar] [CrossRef]
  145. Holla, V.R.; Wang, D.; Brown, J.R.; Mann, J.R.; Katkuri, S.; DuBois, R.N. Prostaglandin E2 regulates the complement inhibitor CD55/decay-accelerating factor in colorectal cancer. J. Biol. Chem. 2005, 280, 476–483. [Google Scholar] [CrossRef]
  146. Nakamura, T.; Nakamura, T.; Matsumoto, K. The functions and possible significance of Kremen as the gatekeeper of Wnt signalling in development and pathology. J. Cell Mol. Med. 2008, 12, 391–408. [Google Scholar] [CrossRef]
  147. Nakamura, T.; Aoki, S.; Kitajima, K.; Takahashi, T.; Matsumoto, K.; Nakamura, T. Molecular cloning and characterization of Kremen, a novel kringle-containing transmembrane protein. Biochim. Biophys. Acta (BBA)-Gene Struct. Expr. 2001, 1518, 63–72. [Google Scholar] [CrossRef]
  148. Mao, B.; Wu, W.; Davidson, G.; Marhold, J.; Li, M.; Mechler, B.M.; Delius, H.; Hoppe, D.; Stannek, P.; Walter, C. Kremen proteins are Dickkopf receptors that regulate Wnt/β-catenin signalling. Nature 2002, 417, 664–667. [Google Scholar] [CrossRef] [PubMed]
  149. Romero, A.; Romão, M.J.; Varela, P.F.; Kölln, I.; Dias, J.M.; Carvalho, A.V.L.; Sanz, L.; Töpfer-Petersen, E.; Calvete, J.J. The crystal structures of two spermadhesins reveal the CUB domain fold. Nat. Struct. Biol. 1997, 4, 783–788. [Google Scholar] [CrossRef] [PubMed]
  150. Verna, J.; Lodder, A.; Lee, K.; Vagts, A.; Ballester, R. A family of genes required for maintenance of cell wall integrity and for the stress response in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 1997, 94, 13804–13809. [Google Scholar] [CrossRef] [PubMed]
  151. Cui, Y.; Peng, R.; Song, H.; Tong, Z.; Qu, X.; Liu, S.; Zhao, X.; Chai, Y.; Wang, P.; Gao, G.F.; et al. Molecular basis of Coxsackievirus A10 entry using the two-in-one attachment and uncoating receptor KRM1. Proc. Natl. Acad. Sci. USA 2020, 117, 18711–18718. [Google Scholar] [CrossRef]
  152. Andersen, J.T.; Sandlie, I. The versatile MHC class I-related FcRn protects IgG and albumin from degradation: Implications for development of new diagnostics and therapeutics. Drug Metab. Pharmacokinet. 2009, 24, 318–332. [Google Scholar] [CrossRef]
  153. Baker, K.; Rath, T.; Flak, M.B.; Arthur, J.C.; Chen, Z.; Glickman, J.N.; Zlobec, I.; Karamitopoulou, E.; Stachler, M.D.; Odze, R.D.; et al. Neonatal Fc receptor expression in dendritic cells mediates protective immunity against colorectal cancer. Immunity 2013, 39, 1095–1107. [Google Scholar] [CrossRef]
  154. Li, L.; Dong, M.; Wang, X.G. The Implication and Significance of Beta 2 Microglobulin: A Conservative Multifunctional Regulator. Chin. Med. J. 2016, 129, 448–455. [Google Scholar] [CrossRef]
  155. Akilesh, S.; Christianson, G.J.; Roopenian, D.O.C.; Shaw, A.S. Neonatal FcR expression in bone marrow-derived cells functions to protect serum IgG from catabolism. J. Immunol. 2007, 179, 4580–4588. [Google Scholar] [CrossRef]
  156. Sockolosky, J.T.; Szoka, F.C. The neonatal Fc receptor, FcRn, as a target for drug delivery and therapy. Adv. Drug Deliv. Rev. 2015, 91, 109–124. [Google Scholar] [CrossRef]
  157. Baldwin, W.M., III; Valujskikh, A.; Fairchild, R.L. The neonatal Fc receptor: Key to homeostasic control of IgG and IgG-related biopharmaceuticals. Am. J. Transpl. 2019, 19, 1881–1887. [Google Scholar] [CrossRef] [PubMed]
  158. Pyzik, M.; Rath, T.; Lencer, W.I.; Baker, K.; Blumberg, R.S. FcRn: The Architect Behind the Immune and Nonimmune Functions of IgG and Albumin. J. Immunol. 2015, 194, 4595–4603. [Google Scholar] [CrossRef] [PubMed]
  159. Simister, N.E.; Ahouse, J.C. The structure and evolution of FcRn. Res. Immunol. 1996, 147, 333–337; discussion 353. [Google Scholar] [CrossRef]
  160. Yang, R.; Zhang, W.; Shang, X.; Chen, H.; Mu, X.; Zhang, Y.; Zheng, Q.; Wang, X.; Liu, Y. Neutrophil-related genes predict prognosis and response to immune checkpoint inhibitors in bladder cancer. Front. Pharmacol. 2022, 13, 1013672. [Google Scholar] [CrossRef]
  161. Cadena Castaneda, D.; Brachet, G.; Goupille, C.; Ouldamer, L.; Gouilleux-Gruart, V. The neonatal Fc receptor in cancer FcRn in cancer. Cancer Med. 2020, 9, 4736–4742. [Google Scholar] [CrossRef]
  162. O’Shannessy, D.J.; Bendas, K.; Schweizer, C.; Wang, W.; Albone, E.; Somers, E.B.; Weil, S.; Meredith, R.K.; Wustner, J.; Grasso, L.; et al. Correlation of FCGRT genomic structure with serum immunoglobulin, albumin and farletuzumab pharmacokinetics in patients with first relapsed ovarian cancer. Genomics 2017, 109, 251–257. [Google Scholar] [CrossRef]
  163. Dalloneau, E.; Baroukh, N.; Mavridis, K.; Maillet, A.; Gueugnon, F.; Courty, Y.; Petit, A.; Kryza, T.; Del Rio, M.; Guyetant, S.; et al. Downregulation of the neonatal Fc receptor expression in non-small cell lung cancer tissue is associated with a poor prognosis. Oncotarget 2016, 7, 54415–54429. [Google Scholar] [CrossRef]
  164. Swiercz, R.; Mo, M.; Khare, P.; Schneider, Z.; Ober, R.J.; Ward, E.S. Loss of expression of the recycling receptor, FcRn, promotes tumor cell growth by increasing albumin consumption. Oncotarget 2017, 8, 3528–3541. [Google Scholar] [CrossRef]
  165. Castaneda, D.O.C.; Dhommee, C.; Baranek, T.; Dalloneau, E.; Lajoie, L.; Valayer, A.; Arnoult, C.; Demattei, M.V.; Fouquenet, D.; Parent, C.; et al. Lack of FcRn Impairs Natural Killer Cell Development and Functions in the Tumor Microenvironment. Front. Immunol. 2018, 9, 2259. [Google Scholar] [CrossRef]
  166. Challa, D.V.K.; Wang, X.; Montoyo, H.P.; Velmurugan, R.; Ober, R.J.; Ward, E.S. Neonatal Fc receptor expression in macrophages is indispensable for IgG homeostasis. MAbs 2019, 11, 848–860. [Google Scholar] [CrossRef]
  167. Gogesch, P.; Dudek, S.; van Zandbergen, G.; Waibler, Z.; Anzaghe, M. The Role of Fc Receptors on the Effectiveness of Therapeutic Monoclonal Antibodies. Int. J. Mol. Sci. 2021, 22, 8947. [Google Scholar] [CrossRef] [PubMed]
  168. Morosky, S.; Wells, A.V.I.; Lemon, K.; Evans, A.S.; Schamus, S.; Bakkenist, C.J.; Coyne, C.B. The neonatal Fc receptor is a pan-echovirus receptor. Proc. Natl. Acad. Sci. USA 2019, 116, 3758–3763. [Google Scholar] [CrossRef] [PubMed]
  169. Takeuchi, O.; Akira, S. Innate immunity to virus infection. Immunological. Rev. 2009, 227, 75–86. [Google Scholar] [CrossRef]
  170. Kawai, T.; Akira, S. Innate immune recognition of viral infection. Nat. Immunol. 2006, 7, 131–137. [Google Scholar] [CrossRef]
  171. Takeuchi, O.; Akira, S. Recognition of viruses by innate immunity. Immunol. Rev. 2007, 220, 214–224. [Google Scholar] [CrossRef]
  172. Carty, M.; Guy, C.; Bowie, A.G. Detection of Viral Infections by Innate Immunity. Biochem. Pharmacol. 2021, 183, 114316. [Google Scholar] [CrossRef]
  173. Barral, P.M.; Sarkar, D.; Su, Z.z.; Barber, G.N.; DeSalle, R.; Racaniello, V.R.; Fisher, P.B. Functions of the cytoplasmic RNA sensors RIG-I and MDA-5: Key regulators of innate immunity. Pharmacol. Ther. 2009, 124, 219–234. [Google Scholar] [CrossRef]
  174. Rehwinkel, J.; Gack, M.U. RIG-I-like receptors: Their regulation and roles in RNA sensing. Nat. Rev. Immunol. 2020, 20, 537–551. [Google Scholar] [CrossRef]
  175. Chen, K.R.; Ling, P. Interplays between Enterovirus A71 and the innate immune system. J. Biomed. Sci. 2019, 26, 95. [Google Scholar] [CrossRef]
  176. Li, D.; Lei, C.; Xu, Z.; Yang, F.; Liu, H.; Zhu, Z.; Li, S.; Liu, X.; Shu, H.; Zheng, H. Foot-and-mouth disease virus non-structural protein 3A inhibits the interferon-β signaling pathway. Sci. Rep. 2016, 6, 21888. [Google Scholar] [CrossRef]
  177. Elrick, M.J.; Pekosz, A.; Duggal, P. Enterovirus D68 molecular and cellular biology and pathogenesis. J. Biol. Chem. 2021, 296, 100317. [Google Scholar] [CrossRef]
  178. Kell, A.M.; Gale, M. RIG-I in RNA virus recognition. Virology 2015, 479, 110–121. [Google Scholar] [CrossRef] [PubMed]
  179. Nahavandi-Parizi, P.; Kariminik, A.; Montazeri, M. Retinoic acid-inducible gene 1 (RIG-1) and IFN-β promoter stimulator-1 (IPS-1) significantly down-regulated in the severe coronavirus disease 2019 (COVID-19). Mol. Biol. Rep. 2023, 50, 907–911. [Google Scholar] [CrossRef] [PubMed]
  180. Xu, L.-G.; Wang, Y.-Y.; Han, K.-J.; Li, L.-Y.; Zhai, Z.; Shu, H.-B. VISA is an adapter protein required for virus-triggered IFN-β signaling. Mol. Cell 2005, 19, 727–740. [Google Scholar] [CrossRef]
  181. Barral, P.M.; Sarkar, D.; Fisher, P.B.; Racaniello, V.R. RIG-I is cleaved during picornavirus infection. Virology 2009, 391, 171–176. [Google Scholar] [CrossRef]
  182. Feng, Q.; Langereis, M.A.; Lork, M.; Nguyen, M.; Hato, S.V.; Lanke, K.; Emdad, L.; Bhoopathi, P.; Fisher, P.B.; Lloyd, R.E.; et al. Enterovirus 2A pro Targets MDA5 and MAVS in Infected Cells. J. Virol. 2014, 88, 3369–3378. [Google Scholar] [CrossRef]
  183. Lei, X.; Liu, X.; Ma, Y.; Sun, Z.; Yang, Y.; Jin, Q.; He, B.; Wang, J. The 3C Protein of Enterovirus 71 Inhibits Retinoid Acid-Inducible Gene I-Mediated Interferon Regulatory Factor 3 Activation and Type I Interferon Responses. J. Virol. 2010, 84, 8051–8061. [Google Scholar] [CrossRef]
  184. Lei, X.; Xiao, X.; Xue, Q.; Jin, Q.; He, B.; Wang, J. Cleavage of Interferon Regulatory Factor 7 by Enterovirus 71 3C Suppresses Cellular Responses. J. Virol. 2013, 87, 1690–1698. [Google Scholar] [CrossRef]
  185. Feng, Q.; Langereis, M.A.; Olagnier, D.; Chiang, C.; van de Winkel, R.; van Essen, P.; Zoll, J.; Hiscott, J.; van Kuppeveld, F.J.M. Coxsackievirus cloverleaf RNA containing a 5′ triphosphate triggers an antiviral response via RIG-I activation. PLoS ONE 2014, 9, e95927. [Google Scholar] [CrossRef]
  186. Li, Y.; Li, M.; Jia, X.; Deng, H.; Wang, W.; Wu, F.; Wang, J.; Dang, S. Association of gene polymorphisms of pattern-recognition receptor signaling pathway with the risk and severity of hand, foot, and mouth disease caused by enterovirus 71 in Chinese Han population. J. Med. Virol. 2017, 90, 692–698. [Google Scholar] [CrossRef]
  187. Pang, L.; Gong, X.; Liu, N.; Xie, G.; Gao, W.; Kong, G.; Li, X.; Zhang, J.; Jin, Y.; Duan, Z. A polymorphism in melanoma differentiation-associated gene 5 may be a risk factor for enterovirus 71 infection. Clin. Microbiol. Infect. 2014, 20, O711–O717. [Google Scholar] [CrossRef] [PubMed]
  188. Oshiumi, H.; Okamoto, M.; Fujii, K.; Kawanishi, T.; Matsumoto, M.; Koike, S.; Seya, T. The TLR3/TICAM-1 pathway is mandatory for innate immune responses to poliovirus infection. J. Immunol. 2011, 187, 5320–5327. [Google Scholar] [CrossRef] [PubMed]
  189. Pathinayake, P.S.; Hsu, A.C.-Y.; Wark, P.A. Innate immunity and immune evasion by enterovirus 71. Viruses 2015, 7, 6613–6630. [Google Scholar] [CrossRef] [PubMed]
  190. Xu, C.; He, X.; Zheng, Z.; Zhang, Z.; Wei, C.; Guan, K.; Hou, L.; Zhang, B.; Zhu, L.; Cao, Y.; et al. Downregulation of MicroRNA miR-526a by Enterovirus Inhibits RIG-I-Dependent Innate Immune Response. J. Virol. 2014, 88, 11356–11368. [Google Scholar] [CrossRef]
  191. Li, J.; Xie, Y.; Li, L.; Li, X.; Shen, L.; Gong, J.; Zhang, R. MicroRNA-30a Modulates Type I Interferon Responses to Facilitate Coxsackievirus B3 Replication Via Targeting Tripartite Motif Protein 25. Front. Immunol. 2021, 11, 603437. [Google Scholar] [CrossRef]
  192. Martín-Vicente, M.; Medrano, L.M.; Resino, S.; García-Sastre, A.; Martínez, I. TRIM25 in the regulation of the antiviral innate immunity. Front. Immunol. 2017, 8, 1187. [Google Scholar] [CrossRef]
  193. Gack, M.U.; Shin, Y.C.; Joo, C.-H.; Urano, T.; Liang, C.; Sun, L.; Takeuchi, O.; Akira, S.; Chen, Z.; Inoue, S.; et al. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature 2007, 446, 916–920. [Google Scholar] [CrossRef]
  194. Seth, R.B.; Sun, L.; Ea, C.-K.; Chen, Z.J. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-κB and IRF3. Cell 2005, 122, 669–682. [Google Scholar] [CrossRef]
  195. Meylan, E.; Curran, J.; Hofmann, K.; Moradpour, D.; Binder, M.; Bartenschlager, R.; Tschopp, J. Cardif is an adaptor protein in the RIG-I antiviral pathway and is targeted by hepatitis C virus. Nature 2005, 437, 1167–1172. [Google Scholar] [CrossRef]
  196. Jacobs, J.L.; Coyne, C.B. Mechanisms of MAVS regulation at the mitochondrial membrane. J. Mol. Biol. 2013, 425, 5009–5019. [Google Scholar] [CrossRef]
  197. Honda, K.; Takaoka, A.; Taniguchi, T. Type I Inteferon Gene Induction by the Interferon Regulatory Factor Family of Transcription Factors. Immunity 2006, 25, 349–360. [Google Scholar] [CrossRef] [PubMed]
  198. Tenoever, B.R.; Ng, S.-L.; Chua, M.A.; McWhirter, S.M.; García-Sastre, A.; Maniatis, T. Multiple functions of the IKK-related kinase IKKε in interferon-mediated antiviral immunity. Science 2007, 315, 1274–1278. [Google Scholar] [CrossRef] [PubMed]
  199. Kawai, T.; Takahashi, K.; Sato, S.; Coban, C.; Kumar, H.; Kato, H.; Ishii, K.J.; Takeuchi, O.; Akira, S. IPS-1, an adaptor triggering RIG-I- and Mda5-mediated type I interferon induction. Nat. Immunol. 2005, 6, 981–988. [Google Scholar] [CrossRef] [PubMed]
  200. Fitzgerald, K.A.; McWhirter, S.M.; Faia, K.L.; Rowe, D.O.C.; Latz, E.; Golenbock, D.T.; Coyle, A.J.; Liao, S.-M.; Maniatis, T. IKKE and TBKI are essential components of the IRF3 signaling pathway. Nat. Immunol. 2003, 4, 491–496. [Google Scholar] [CrossRef] [PubMed]
  201. Huang, B.; Li, J.; Zhang, X.; Zhao, Q.; Lu, M.; Lv, Y. RIG-1 and MDA-5 signaling pathways contribute to IFN-β production and viral replication in porcine circovirus virus type 2-infected PK-15 cells in vitro. Vet. Microbiol. 2017, 211, 36–42. [Google Scholar] [CrossRef] [PubMed]
  202. Xue, B.; Li, H.; Guo, M.; Wang, J.; Xu, Y.; Zou, X.; Deng, R.; Li, G.; Zhu, H. TRIM21 Promotes Innate Immune Response to RNA Viral Infection through Lys27-Linked Polyubiquitination of MAVS. J. Virol. 2018, 92, 10–1128. [Google Scholar] [CrossRef]
  203. Liu, H.; Li, M.; Song, Y.; Xu, W. TRIM21 restricts coxsackievirus B3 replication, cardiac and pancreatic injury via interacting with MAVS and positively regulating IRF3-mediated Type-I interferon production. Front. Immunol. 2018, 9, 2479. [Google Scholar] [CrossRef]
  204. Lopacinski, A.B.; Sweatt, A.J.; Smolko, C.M.; Gray-Gaillard, E.; Borgman, C.A.; Shah, M.; Janes, K.A. Modeling the complete kinetics of coxsackievirus B3 reveals human determinants of host-cell feedback. Cell Syst. 2021, 12, 304–323.e13. [Google Scholar] [CrossRef]
  205. Zhang, Q.-M.; Song, W.-Q.; Li, Y.-J.; Qian, J.; Zhai, A.-X.; Wu, J.; Li, A.-M.; He, J.-M.; Zhao, J.-Y.; Yu, X.; et al. Over-expression of mitochondrial antiviral signaling protein inhibits coxsackievirus B3 infection by enhancing type-I interferons production. Virol. J. 2012, 9, 312. [Google Scholar] [CrossRef]
  206. Shi, Y.; Liu, Y.; Zheng, Y.; Tang, Y.; Zhu, G.; Qiu, W.; Huang, L.; Han, S.; Yin, J.; Peng, B.; et al. Autophagy triggered by MAVS inhibits Coxsackievirus A16 replication. Acta Virol. 2019, 63, 392–402. [Google Scholar] [CrossRef]
  207. Li, X.-D.; Sun, L.; Seth, R.B.; Pineda, G.; Chen, Z.J. Hepatitis C virus protease NS3/4A cleaves mitochondrial antiviral signaling protein off the mitochondria to evade innate immunity. Proc. Natl. Acad. Sci. USA 2005, 102, 17717–17722. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, Y.; Ma, L.; Stipkovits, L.; Szathmary, S.; Li, X.; Liu, Y. The Strategy of Picornavirus Evading Host Antiviral Responses: Non-structural Proteins Suppress the Production of IFNs. Front. Microbiol. 2018, 9, 2943. [Google Scholar] [CrossRef] [PubMed]
  209. Li, X.; Yang, E.; Li, X.; Fan, T.; Guo, S.; Yang, H.; Wu, B.; Wang, H. MAVS-Based Reporter Systems for Real-Time Imaging of EV71 Infection and Antiviral Testing. Viruses 2023, 15, 1064. [Google Scholar] [CrossRef]
  210. Varki, A. Sialic acids in human health and disease. Trends Mol. Med. 2008, 14, 351–360. [Google Scholar] [CrossRef]
  211. Ghosh, S. Sialic Acids and Sialoglycoconjugates in the Biology of Life, Health and Disease; Academic Press: Cambridge, MA, USA, 2020; ISBN 978-0-12-816126-5. [Google Scholar] [CrossRef]
  212. Varki, A.; Gagneux, P. Multifarious roles of sialic acids in immunity. Ann. N.Y. Acad. Sci. 2012, 1253, 16–36. [Google Scholar] [CrossRef]
  213. Peters, C.E.; Carette, J.E. Return of the neurotropic enteroviruses: Co-opting cellular pathways for infection. Viruses 2021, 13, 166. [Google Scholar] [CrossRef]
  214. Rosenfeld, A.B.; Warren, A.V.L.; Racaniello, V.R. Neurotropism of enterovirus D68 isolates is independent of sialic acid and is not a recently acquired phenotype. mBio 2019, 10, e02370-e19. [Google Scholar] [CrossRef]
  215. Sun, J.; Hu, X.-Y.; Yu, X.-F. Current understanding of human enterovirus D68. Viruses 2019, 11, 490. [Google Scholar] [CrossRef]
  216. Heida, R.; Bhide, Y.C.; Gasbarri, M.; Kocabiyik, Ö.; Stellacci, F.; Huckriede, A.V.L.; Hinrichs, W.L.; Frijlink, H.W. Advances in the development of entry inhibitors for sialic-acid-targeting viruses. Drug Discov. Today 2020, 26, 122–137. [Google Scholar] [CrossRef]
  217. Mistry, N.; Inoue, H.; Jamshidi, F.; Storm, R.J.; Oberste, M.S.; Arnberg, N. Coxsackievirus A24 Variant Uses Sialic Acid-Containing O -Linked Glycoconjugates as Cellular Receptors on Human Ocular Cells. J. Virol. 2011, 85, 11283–11290. [Google Scholar] [CrossRef]
  218. Alexander, D.A.; Dimock, K. Sialic Acid Functions in Enterovirus 70 Binding and Infection. J. Virol. 2002, 76, 11265–11272. [Google Scholar] [CrossRef] [PubMed]
  219. Su, P.-Y.; Liu, Y.-T.; Chang, H.-Y.; Huang, S.-W.; Wang, Y.-F.; Yu, C.-K.; Wang, J.-R.; Chang, C.-F. Cell surface sialylation affects binding of enterovirus 71 to rhabdomyosarcoma and neuroblastoma cells. BMC Microbiol. 2012, 12, 162. [Google Scholar] [CrossRef] [PubMed]
  220. Matrosovich, M.; Herrler, G.; Klenk, H.D. Sialic acid receptors of viruses. Top. Curr. Chem. 2015, 367, 1–28. [Google Scholar] [CrossRef]
  221. Salomon, R.; Webster, R.G. The Influenza Virus Enigma. Cell 2009, 136, 402–410. [Google Scholar] [CrossRef]
  222. Kamiki, H.; Murakami, S.; Nishikaze, T.; Hiono, T.; Igarashi, M.; Furuse, Y.; Matsugo, H.; Ishida, H.; Katayama, M.; Sekine, W.; et al. Influenza A Virus Agnostic Receptor Tropism Revealed Using a Novel Biological System with Terminal Sialic Acid Knockout Cells. J. Virol. 2022, 96, e0041622. [Google Scholar] [CrossRef]
  223. Turnaev, I.I.; Bocharnikova, M.E.; Afonnikov, D.A. Human phospholipases A2: A functional and evolutionary analysis. Vavilovskii Zhurnal Genet. I Sel. 2022, 26, 787–797. [Google Scholar] [CrossRef]
  224. Wilton, D.O.C. Phospholipases A2: Structure and function. Eur. J. Lipid Sci. Technol. 2005, 107, 193–205. [Google Scholar] [CrossRef]
  225. Baggen, J.; Liu, Y.; Lyoo, H.; van Vliet, A.V.L.W.; Wahedi, M.; de Bruin, J.W.; Roberts, R.W.; Overduin, P.; Meijer, A.; Rossmann, M.G.; et al. Bypassing pan-enterovirus host factor PLA2G16. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef]
  226. Staring, J.; von Castelmur, E.; Blomen, V.A.; van den Hengel, L.G.; Brockmann, M.; Baggen, J.; Thibaut, H.J.; Nieuwenhuis, J.; Janssen, H.; van Kuppeveld, F.J.; et al. PLA2G16 represents a switch between entry and clearance of Picornaviridae. Nature 2017, 541, 412–416. [Google Scholar] [CrossRef]
  227. Elling, U.; Wimmer, R.A.; Leibbrandt, A.; Burkard, T.; Michlits, G.; Leopoldi, A.; Micheler, T.; Abdeen, D.; Zhuk, S.; Aspalter, I.M.; et al. A reversible haploid mouse embryonic stem cell biobank resource for functional genomics. Nature 2017, 550, 114–118. [Google Scholar] [CrossRef]
  228. Xiong, S.; Tu, H.; Kollareddy, M.; Pant, V.; Li, Q.; Zhang, Y.; Jackson, J.G.; Suh, Y.-A.; Elizondo-Fraire, A.C.; Yang, P.; et al. Pla2g16 phospholipase mediates gain-of-function activities of mutant p53. Proc. Natl. Acad. Sci. USA 2014, 111, 11145–11150. [Google Scholar] [CrossRef] [PubMed]
  229. The Human Protein Atlas. Available online: https://www.proteinatlas.org/ (accessed on 19 August 2024).
  230. Uhlen, M.; Fagerberg, L.; Hallstrom, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, A.; Kampf, C.; Sjostedt, E.; Asplund, A.; et al. Proteomics. Tissue-based map of the human proteome. Science 2015, 347, 1260419. [Google Scholar] [CrossRef] [PubMed]
  231. Masson, D.; Jarry, A.; Baury, B.; Blanchardie, P.; Laboisse, C.; Lustenberger, P.; Denis, M.G. Overexpression of the CD155 gene in human colorectal carcinoma. Gut 2001, 49, 236–240. [Google Scholar] [CrossRef]
  232. Murakami, D.; Matsuda, K.; Iwamoto, H.; Mitani, Y.; Mizumoto, Y.; Nakamura, Y.; Matsuzaki, I.; Iwamoto, R.; Takahashi, Y.; Kojima, F.; et al. Prognostic value of CD155/TIGIT expression in patients with colorectal cancer. PLoS ONE 2022, 17, e0265908. [Google Scholar] [CrossRef]
  233. Sloan, K.E.; Eustace, B.K.; Stewart, J.K.; Zehetmeier, C.; Torella, C.; Simeone, M.; Roy, J.E.; Unger, C.; Louis, D.N.; Ilag, L.L.; et al. CD155/PVR plays a key role in cell motility during tumor cell invasion and migration. BMC Cancer 2004, 4, 73. [Google Scholar] [CrossRef]
  234. Ma, H.; Chen, X.; Mo, S.; Mao, X.; Chen, J.; Liu, Y.; Lu, Z.; Yu, S.; Chen, J. The spatial coexistence of TIGIT/CD155 defines poorer survival and resistance to adjuvant chemotherapy in pancreatic ductal adenocarcinoma. Theranostics 2023, 13, 4601–4614. [Google Scholar] [CrossRef]
  235. Oyama, R.; Kanayama, M.; Mori, M.; Matsumiya, H.; Taira, A.; Shinohara, S.; Takenaka, M.; Yoneda, K.; Kuroda, K.; Tanaka, F. CD155 expression and its clinical significance in non-small cell lung cancer. Oncol. Lett. 2022, 23, 166. [Google Scholar] [CrossRef]
  236. Liu, X.; Xu, C.; Guo, T.; Zhan, S.; Quan, Q.; Li, M.; Wang, Z.; Zhang, X.; Guo, L.; Cao, L. Clinical significance of CD155 expression and correlation with cellular components of tumor microenvironment in gastric adenocarcinoma. Front. Immunol. 2023, 14, 1173524. [Google Scholar] [CrossRef]
  237. Inozume, T.; Yaguchi, T.; Furuta, J.; Harada, K.; Kawakami, Y.; Shimada, S. Melanoma Cells Control Antimelanoma CTL Responses via Interaction between TIGIT and CD155 in the Effector Phase. J. Investig. Dermatol. 2016, 136, 255–263. [Google Scholar] [CrossRef]
  238. Luo, C.; Ye, W.; Hu, J.; Othmane, B.; Li, H.; Chen, J.; Zu, X. A Poliovirus Receptor (CD155)-Related Risk Signature Predicts the Prognosis of Bladder Cancer. Front. Oncol. 2021, 11, 660273. [Google Scholar] [CrossRef]
  239. Huang, D.W.; Huang, M.; Lin, X.S.; Huang, Q. CD155 expression and its correlation with clinicopathologic characteristics, angiogenesis, and prognosis in human cholangiocarcinoma. Onco. Targets Ther. 2017, 10, 3817–3825. [Google Scholar] [CrossRef] [PubMed]
  240. Enloe, B.M.; Jay, D.G. Inhibition of Necl-5 (CD155/PVR) reduces glioblastoma dispersal and decreases MMP-2 expression and activity. J. Neurooncol. 2011, 102, 225–235. [Google Scholar] [CrossRef] [PubMed]
  241. Mori, K.; Matsumoto, K.; Amano, N.; Koguchi, D.; Shimura, S.; Hagiwara, M.; Shimizu, Y.; Ikeda, M.; Sato, Y.; Iwamura, M. Expression of Membranous CD155 Is Associated with Aggressive Phenotypes and a Poor Prognosis in Patients with Bladder Cancer. Cancers 2022, 14, 1576. [Google Scholar] [CrossRef]
  242. Zhang, J.; Zhu, Y.; Wang, Q.; Kong, Y.; Sheng, H.; Guo, J.; Xu, J.; Dai, B. Poliovirus receptor CD155 is up-regulated in muscle-invasive bladder cancer and predicts poor prognosis. Urol. Oncol. 2020, 38, 41.e11–41.e18. [Google Scholar] [CrossRef]
  243. Yong, H.; Cheng, R.; Li, X.; Gao, G.; Jiang, X.; Cheng, H.; Zhou, X.; Zhao, W. CD155 expression and its prognostic value in postoperative patients with breast cancer. Biomed. Pharmacother. 2019, 115, 108884. [Google Scholar] [CrossRef]
  244. Iguchi-Manaka, A.; Okumura, G.; Ichioka, E.; Kiyomatsu, H.; Ikeda, T.; Bando, H.; Shibuya, A.; Shibuya, K. High expression of soluble CD155 in estrogen receptor-negative breast cancer. Breast Cancer 2020, 27, 92–99. [Google Scholar] [CrossRef]
  245. Solecki, D.J.; Gromeier, M.; Mueller, S.; Bernhardt, G.; Wimmer, E. Expression of the human poliovirus receptor/CD155 gene is activated by sonic hedgehog. J. Biol. Chem. 2002, 277, 25697–25702. [Google Scholar] [CrossRef]
  246. Thompson, E.M.; Brown, M.; Dobrikova, E.; Ramaswamy, V.; Taylor, M.D.; McLendon, R.; Sanks, J.; Chandramohan, V.; Bigner, D.; Gromeier, M. Poliovirus Receptor (CD155) Expression in Pediatric Brain Tumors Mediates Oncolysis of Medulloblastoma and Pleomorphic Xanthoastrocytoma. J. Neuropathol. Exp. Neurol. 2018, 77, 696–702. [Google Scholar] [CrossRef]
  247. Li, S.; McLendon, R.; Sankey, E.; Kornahrens, R.; Lyne, A.M.; Cavalli, F.M.G.; McKay, Z.; Herndon, J.E., II; Remke, M.; Picard, D.; et al. CD155 is a putative therapeutic target in medulloblastoma. Clin. Transl. Oncol. 2023, 25, 696–705. [Google Scholar] [CrossRef]
  248. Matsudo, K.; Takada, K.; Kinoshita, F.; Hashinokuchi, A.; Nagano, T.; Akamine, T.; Kohno, M.; Takenaka, T.; Shimokawa, M.; Oda, Y.; et al. CD155 Expression in Early-Stage Lung Adenocarcinoma. Ann. Thorac. Surg. 2024, in press. [Google Scholar] [CrossRef]
  249. Iguchi-Manaka, A.; Okumura, G.; Kojima, H.; Cho, Y.; Hirochika, R.; Bando, H.; Sato, T.; Yoshikawa, H.; Hara, H.; Shibuya, A.; et al. Increased Soluble CD155 in the Serum of Cancer Patients. PLoS ONE 2016, 11, e0152982. [Google Scholar] [CrossRef] [PubMed]
  250. Atsumi, S.; Matsumine, A.; Toyoda, H.; Niimi, R.; Iino, T.; Sudo, A. Prognostic significance of CD155 mRNA expression in soft tissue sarcomas. Oncol. Lett. 2013, 5, 1771–1776. [Google Scholar] [CrossRef] [PubMed]
  251. Sloan, K.E.; Stewart, J.K.; Treloar, A.F.; Matthews, R.T.; Jay, D.G. CD155/PVR enhances glioma cell dispersal by regulating adhesion signaling and focal adhesion dynamics. Cancer Res. 2005, 65, 10930–10937. [Google Scholar] [CrossRef]
  252. Nishiwada, S.; Sho, M.; Yasuda, S.; Shimada, K.; Yamato, I.; Akahori, T.; Kinoshita, S.; Nagai, M.; Konishi, N.; Nakajima, Y. Clinical significance of CD155 expression in human pancreatic cancer. Anticancer Res. 2015, 35, 2287–2297. [Google Scholar]
  253. Zhao, K.; Ma, L.; Feng, L.; Huang, Z.; Meng, X.; Yu, J. CD155 Overexpression Correlates With Poor Prognosis in Primary Small Cell Carcinoma of the Esophagus. Front. Mol. Biosci. 2020, 7, 608404. [Google Scholar] [CrossRef]
  254. Lee, B.H.; Kim, J.H.; Kang, K.W.; Lee, S.R.; Park, Y.; Sung, H.J.; Kim, B.S. PVR (CD155) Expression as a Potential Prognostic Marker in Multiple Myeloma. Biomedicines 2022, 10, 1099. [Google Scholar] [CrossRef]
  255. Zheng, Q.; Gao, J.; Yin, P.; Wang, W.; Wang, B.; Li, Y.; Zhao, C. CD155 contributes to the mesenchymal phenotype of triple-negative breast cancer. Cancer Sci. 2020, 111, 383–394. [Google Scholar] [CrossRef]
  256. Carlsten, M.; Norell, H.; Bryceson, Y.T.; Poschke, I.; Schedvins, K.; Ljunggren, H.G.; Kiessling, R.; Malmberg, K.J. Primary human tumor cells expressing CD155 impair tumor targeting by down-regulating DNAM-1 on NK cells. J. Immunol. 2009, 183, 4921–4930. [Google Scholar] [CrossRef]
  257. Lepletier, A.; Madore, J.; O’Donnell, J.S.; Johnston, R.L.; Li, X.Y.; McDonald, E.; Ahern, E.; Kuchel, A.; Eastgate, M.; Pearson, S.A.; et al. Tumor CD155 Expression Is Associated with Resistance to Anti-PD1 Immunotherapy in Metastatic Melanoma. Clin. Cancer Res. 2020, 26, 3671–3681. [Google Scholar] [CrossRef]
  258. Kernek, C.B. Avascular nonunion of a subtrochanteric femur fracture with formation of a heterotopic bone strut. Orthopedics 1988, 11, 36–52. [Google Scholar]
  259. Regev, O.; Kizner, M.; Roncato, F.; Dadiani, M.; Saini, M.; Castro-Giner, F.; Yajuk, O.; Kozlovski, S.; Levi, N.; Addadi, Y.; et al. ICAM-1 on Breast Cancer Cells Suppresses Lung Metastasis but Is Dispensable for Tumor Growth and Killing by Cytotoxic T Cells. Front. Immunol. 2022, 13, 849701. [Google Scholar] [CrossRef] [PubMed]
  260. Guo, P.; Huang, J.; Wang, L.; Jia, D.; Yang, J.; Dillon, D.A.; Zurakowski, D.; Mao, H.; Moses, M.A.; Auguste, D.T. ICAM-1 as a molecular target for triple negative breast cancer. Proc. Natl. Acad. Sci. USA 2014, 111, 14710–14715. [Google Scholar] [CrossRef] [PubMed]
  261. Schroder, C.; Witzel, I.; Muller, V.; Krenkel, S.; Wirtz, R.M.; Janicke, F.; Schumacher, U.; Milde-Langosch, K. Prognostic value of intercellular adhesion molecule (ICAM)-1 expression in breast cancer. J. Cancer Res. Clin. Oncol. 2011, 137, 1193–1201. [Google Scholar] [CrossRef]
  262. Taftaf, R.; Liu, X.; Singh, S.; Jia, Y.; Dashzeveg, N.K.; Hoffmann, A.D.; El-Shennawy, L.; Ramos, E.K.; Adorno-Cruz, V.; Schuster, E.J.; et al. ICAM1 initiates CTC cluster formation and trans-endothelial migration in lung metastasis of breast cancer. Nat. Commun. 2021, 12, 4867. [Google Scholar] [CrossRef] [PubMed]
  263. Luo, L.; Xia, L.; Zha, B.; Zuo, C.; Deng, D.; Chen, M.; Hu, L.; He, Y.; Dai, F.; Wu, J.; et al. miR-335-5p targeting ICAM-1 inhibits invasion and metastasis of thyroid cancer cells. Biomed. Pharmacother. 2018, 106, 983–990. [Google Scholar] [CrossRef] [PubMed]
  264. Vedvyas, Y.; McCloskey, J.E.; Yang, Y.; Min, I.M.; Fahey, T.J.; Zarnegar, R.; Hsu, Y.S.; Hsu, J.M.; Van Besien, K.; Gaudet, I.; et al. Manufacturing and preclinical validation of CAR T cells targeting ICAM-1 for advanced thyroid cancer therapy. Sci. Rep. 2019, 9, 10634. [Google Scholar] [CrossRef]
  265. Buitrago, D.; Keutgen, X.M.; Crowley, M.; Filicori, F.; Aldailami, H.; Hoda, R.; Liu, Y.F.; Hoda, R.S.; Scognamiglio, T.; Jin, M.; et al. Intercellular adhesion molecule-1 (ICAM-1) is upregulated in aggressive papillary thyroid carcinoma. Ann. Surg. Oncol. 2012, 19, 973–980. [Google Scholar] [CrossRef]
  266. Song, P.; Xu, Y.; Ye, G. B7-H3 and ICAM-1 are potentially therapeutic targets for thyroid carcinoma. Diagn. Pathol. 2024, 19, 77. [Google Scholar] [CrossRef]
  267. Shi, X.; Jiang, J.; Ye, X.; Liu, Y.; Wu, Q.; Wang, L. Prognostic prediction and diagnostic role of intercellular adhesion molecule-1 (ICAM1) expression in clear cell renal cell carcinoma. J. Mol. Histol. 2014, 45, 427–434. [Google Scholar] [CrossRef]
  268. Maruo, Y.; Gochi, A.; Kaihara, A.; Shimamura, H.; Yamada, T.; Tanaka, N.; Orita, K. ICAM-1 expression and the soluble ICAM-1 level for evaluating the metastatic potential of gastric cancer. Int. J. Cancer 2002, 100, 486–490. [Google Scholar] [CrossRef]
  269. Maeda, K.; Kang, S.M.; Sawada, T.; Nishiguchi, Y.; Yashiro, M.; Ogawa, Y.; Ohira, M.; Ishikawa, T.; Hirakawa, Y.S.C.K. Expression of intercellular adhesion molecule-1 and prognosis in colorectal cancer. Oncol. Rep. 2002, 9, 511–514. [Google Scholar] [CrossRef] [PubMed]
  270. Zarzycka, M.; Kotula-Balak, M.; Gil, D. The mechanism of the contribution of ICAM-1 to epithelial-mesenchymal transition (EMT) in bladder cancer. Hum. Cell 2024, 37, 801–816. [Google Scholar] [CrossRef] [PubMed]
  271. Hamai, A.; Meslin, F.; Benlalam, H.; Jalil, A.; Mehrpour, M.; Faure, F.; Lecluse, Y.; Vielh, P.; Avril, M.F.; Robert, C.; et al. ICAM-1 has a critical role in the regulation of metastatic melanoma tumor susceptibility to CTL lysis by interfering with PI3K/AKT pathway. Cancer Res. 2008, 68, 9854–9864. [Google Scholar] [CrossRef] [PubMed]
  272. Howell, W.M.; Rose-Zerilli, M.J.; Theaker, J.M.; Bateman, A.C. ICAM-1 polymorphisms and development of cutaneous malignant melanoma. Int. J. Immunogenet. 2005, 32, 367–373. [Google Scholar] [CrossRef]
  273. Usami, Y.; Ishida, K.; Sato, S.; Kishino, M.; Kiryu, M.; Ogawa, Y.; Okura, M.; Fukuda, Y.; Toyosawa, S. Intercellular adhesion molecule-1 (ICAM-1) expression correlates with oral cancer progression and induces macrophage/cancer cell adhesion. Int. J. Cancer 2013, 133, 568–578. [Google Scholar] [CrossRef]
  274. Kotteas, E.A.; Boulas, P.; Gkiozos, I.; Tsagkouli, S.; Tsoukalas, G.; Syrigos, K.N. The intercellular cell adhesion molecule-1 (icam-1) in lung cancer: Implications for disease progression and prognosis. Anticancer Res. 2014, 34, 4665–4672. [Google Scholar]
  275. Relph, K.; Arif, M.; Pandha, H.; Annels, N.; Simpson, G.R. Analysis of ICAM-1 Expression on Bladder Carcinoma Cell Lines and Infectivity and Oncolysis by Coxsackie Virus A21. Methods Mol. Biol. 2023, 2684, 319–327. [Google Scholar] [CrossRef]
  276. Madigan, M.C.; Penfold, P.L.; King, N.J.; Billson, F.A.; Conway, R.M. Immunoglobulin superfamily expression in primary retinoblastoma and retinoblastoma cell lines. Oncol. Res. 2002, 13, 103–111. [Google Scholar]
  277. Arnold, J.M.; Cummings, M.; Purdie, D.; Chenevix-Trench, G. Reduced expression of intercellular adhesion molecule-1 in ovarian adenocarcinomas. Br. J. Cancer 2001, 85, 1351–1358. [Google Scholar] [CrossRef]
  278. Allende-Vega, N.; Marco Brualla, J.; Falvo, P.; Alexia, C.; Constantinides, M.; Fayd’herbe de Maudave, A.; Coenon, L.; Gitenay, D.; Mitola, G.; Massa, P.; et al. Metformin sensitizes leukemic cells to cytotoxic lymphocytes by increasing expression of intercellular adhesion molecule-1 (ICAM-1). Sci. Rep. 2022, 12, 1341. [Google Scholar] [CrossRef]
  279. Lin, Y.M.; Chang, Z.L.; Liao, Y.Y.; Chou, M.C.; Tang, C.H. IL-6 promotes ICAM-1 expression and cell motility in human osteosarcoma. Cancer Lett. 2013, 328, 135–143. [Google Scholar] [CrossRef] [PubMed]
  280. Vega, M.A.; Segui-Real, B.; Garcia, J.A.; Cales, C.; Rodriguez, F.; Vanderkerckhove, J.; Sandoval, I.V. Cloning, sequencing, and expression of a cDNA encoding rat LIMP II, a novel 74-kDa lysosomal membrane protein related to the surface adhesion protein CD36. J. Biol. Chem. 1991, 266, 16818–16824. [Google Scholar] [CrossRef] [PubMed]
  281. Calvo, D.; Vega, M.A. Identification, primary structure, and distribution of CLA-1, a novel member of the CD36/LIMPII gene family. J. Biol. Chem. 1993, 268, 18929–18935. [Google Scholar] [CrossRef]
  282. Mulcahy, J.V.; Riddell, D.R.; Owen, J.S. Human scavenger receptor class B type II (SR-BII) and cellular cholesterol efflux. Biochem. J. 2004, 377, 741–747. [Google Scholar] [CrossRef] [PubMed]
  283. Ritsch, A.; Tancevski, I.; Schgoer, W.; Pfeifhofer, C.; Gander, R.; Eller, P.; Foeger, B.; Stanzl, U.; Patsch, J.R. Molecular characterization of rabbit scavenger receptor class B types I and II: Portal to central vein gradient of expression in the liver. J. Lipid Res. 2004, 45, 214–222. [Google Scholar] [CrossRef]
  284. Grove, J.; Huby, T.; Stamataki, Z.; Vanwolleghem, T.; Meuleman, P.; Farquhar, M.; Schwarz, A.; Moreau, M.; Owen, J.S.; Leroux-Roels, G.; et al. Scavenger receptor BI and BII expression levels modulate hepatitis C virus infectivity. J. Virol. 2007, 81, 3162–3169. [Google Scholar] [CrossRef]
  285. Zhang, X.; Wang, H.; Sun, Y.; Qi, M.; Li, W.; Zhang, Z.; Zhang, X.E.; Cui, Z. Enterovirus A71 Oncolysis of Malignant Gliomas. Mol. Ther. 2020, 28, 1533–1546. [Google Scholar] [CrossRef]
  286. Zhang, D.; Fang, J.; Shan, J.; Xu, L.; Wu, Y.; Lu, B.; Zhang, X.; Wang, C.; Sun, P.; Wang, Q. SCARB2 associates with tumor-infiltrating neutrophils and predicts poor prognosis in breast cancer. Breast Cancer Res. Treat. 2024, 207, 15–24. [Google Scholar] [CrossRef]
  287. Montastruc, M.; Reiffers, J.; Stoppa, A.M.; Sotto, J.J.; Corront, B.; Marit, G.; Maraninchi, D.; Michallet, M.; Gastaut, J.A.; Broustet, A.; et al. Treatment of acute myeloid leukemia in elderly patients: The influence of maintenance therapy (BGM 84 protocol). Nouv. Rev. Fr. Hematol. 1990, 32, 147–152. [Google Scholar]
  288. Reeh, M.; Bockhorn, M.; Gorgens, D.; Vieth, M.; Hoffmann, T.; Simon, R.; Izbicki, J.R.; Sauter, G.; Schumacher, U.; Anders, M. Presence of the coxsackievirus and adenovirus receptor (CAR) in human neoplasms: A multitumour array analysis. Br. J. Cancer 2013, 109, 1848–1858. [Google Scholar] [CrossRef]
  289. Zeimet, A.G.; Muller-Holzner, E.; Schuler, A.; Hartung, G.; Berger, J.; Hermann, M.; Widschwendter, M.; Bergelson, J.M.; Marth, C. Determination of molecules regulating gene delivery using adenoviral vectors in ovarian carcinomas. Gene Ther. 2002, 9, 1093–1100. [Google Scholar] [CrossRef] [PubMed]
  290. Hasenburg, A.; Fischer, D.O.C.; Tong, X.W.; Rojas-Martinez, A.; Kaufman, R.H.; Ramzy, I.; Kohlberger, P.; Orlowska-Volk, M.; Aguilar-Cordova, E.; Kieback, D.G. Adenovirus-mediated thymidine kinase gene therapy for recurrent ovarian cancer: Expression of coxsackie-adenovirus receptor and integrins alphavbeta3 and alphavbeta5. J. Soc. Gynecol. Investig. 2002, 9, 174–180. [Google Scholar] [CrossRef] [PubMed]
  291. Reimer, D.; Steppan, I.; Wiedemair, A.; Concin, N.; Hofstetter, G.; Marth, C.; Muller-Holzner, E.; Zeimet, A.G. Soluble isoforms but not the transmembrane form of coxsackie-adenovirus receptor are of clinical relevance in epithelial ovarian cancer. Int. J. Cancer 2007, 120, 2568–2575. [Google Scholar] [CrossRef]
  292. Qin, M.; Escuadro, B.; Dohadwala, M.; Sharma, S.; Batra, R.K. A novel role for the coxsackie adenovirus receptor in mediating tumor formation by lung cancer cells. Cancer Res. 2004, 64, 6377–6380. [Google Scholar] [CrossRef]
  293. Wang, Y.; Wang, S.; Bao, Y.; Ni, C.; Guan, N.; Zhao, J.; Salford, L.G.; Widegren, B.; Fan, X. Coxsackievirus and adenovirus receptor expression in non-malignant lung tissues and clinical lung cancers. J. Mol. Histol. 2006, 37, 153–160. [Google Scholar] [CrossRef]
  294. Zhang, X.; Fang, B.; Mohan, R.; Chang, J.Y. Coxsackie-adenovirus receptor as a novel marker of stem cells in treatment-resistant non-small cell lung cancer. Radiother. Oncol. 2012, 105, 250–257. [Google Scholar] [CrossRef]
  295. Wunder, T.; Schmid, K.; Wicklein, D.; Groitl, P.; Dobner, T.; Lange, T.; Anders, M.; Schumacher, U. Expression of the coxsackie adenovirus receptor in neuroendocrine lung cancers and its implications for oncolytic adenoviral infection. Cancer Gene Ther. 2013, 20, 25–32. [Google Scholar] [CrossRef]
  296. Uotani, K.; Tazawa, H.; Hasei, J.; Fujiwara, T.; Yoshida, A.; Yamakawa, Y.; Omori, T.; Sugiu, K.; Komatsubara, T.; Kondo, H.; et al. Fluorescence-guided assessment of bone and soft-tissue sarcomas for predicting the efficacy of telomerase-specific oncolytic adenovirus. PLoS ONE 2024, 19, e0298292. [Google Scholar] [CrossRef]
  297. Giaginis, C.; Zarros, A.; Alexandrou, P.; Klijanienko, J.; Delladetsima, I.; Theocharis, S. Evaluation of coxsackievirus and adenovirus receptor expression in human benign and malignant thyroid lesions. APMIS 2010, 118, 210–221. [Google Scholar] [CrossRef]
  298. Rauen, K.A.; Sudilovsky, D.; Le, J.L.; Chew, K.L.; Hann, B.; Weinberg, V.; Schmitt, L.D.; McCormick, F. Expression of the coxsackie adenovirus receptor in normal prostate and in primary and metastatic prostate carcinoma: Potential relevance to gene therapy. Cancer Res. 2002, 62, 3812–3818. [Google Scholar]
  299. Pandha, H.S.; Stockwin, L.T.H.; Eaton, J.; Clarke, I.A.; Dalgleish, A.G.; Todryk, S.M.; Blair, G.E. Coxsackie B and adenovirus receptor, integrin and major histocompatibility complex class I expression in human prostate cancer cell lines: Implications for gene therapy strategies. Prostate Cancer Prostatic. Dis. 2003, 6, 6–11. [Google Scholar] [CrossRef] [PubMed]
  300. Martin, T.A.; Watkins, G.; Jiang, W.G. The Coxsackie-adenovirus receptor has elevated expression in human breast cancer. Clin. Exp. Med. 2005, 5, 122–128. [Google Scholar] [CrossRef] [PubMed]
  301. Auer, D.; Reimer, D.; Porto, V.; Fleischer, M.; Roessler, J.; Wiedemair, A.; Marth, C.; Muller-Holzner, E.; Daxenbichler, G.; Zeimet, A.G. Expression of coxsackie-adenovirus receptor is related to estrogen sensitivity in breast cancer. Breast Cancer Res. Treat. 2009, 116, 103–111. [Google Scholar] [CrossRef]
  302. Vindrieux, D.; Le Corre, L.; Hsieh, J.T.; Metivier, R.; Escobar, P.; Caicedo, A.; Brigitte, M.; Lazennec, G. Coxsackie and adenovirus receptor is a target and a mediator of estrogen action in breast cancer. Endocr. Relat. Cancer 2011, 18, 311–321. [Google Scholar] [CrossRef]
  303. Gu, W.; Ogose, A.; Kawashima, H.; Ito, M.; Ito, T.; Matsuba, A.; Kitahara, H.; Hotta, T.; Tokunaga, K.; Hatano, H.; et al. High-level expression of the coxsackievirus and adenovirus receptor messenger RNA in osteosarcoma, Ewing’s sarcoma, and benign neurogenic tumors among musculoskeletal tumors. Clin. Cancer Res. 2004, 10, 3831–3838. [Google Scholar] [CrossRef]
  304. Kawashima, H.; Ogose, A.; Yoshizawa, T.; Kuwano, R.; Hotta, Y.; Hotta, T.; Hatano, H.; Kawashima, H.; Endo, N. Expression of the coxsackievirus and adenovirus receptor in musculoskeletal tumors and mesenchymal tissues: Efficacy of adenoviral gene therapy for osteosarcoma. Cancer Sci. 2003, 94, 70–75. [Google Scholar] [CrossRef]
  305. Rice, A.M.; Currier, M.A.; Adams, L.C.; Bharatan, N.S.; Collins, M.H.; Snyder, J.D.; Khan, J.; Cripe, T.P. Ewing sarcoma family of tumors express adenovirus receptors and are susceptible to adenovirus-mediated oncolysis. J. Pediatr. Hematol. Oncol. 2002, 24, 527–533. [Google Scholar] [CrossRef]
  306. Witlox, M.A.; Van Beusechem, V.W.; Grill, J.; Haisma, H.J.; Schaap, G.; Bras, J.; Van Diest, P.; De Gast, A.; Curiel, D.T.; Pinedo, H.M.; et al. Epidermal growth factor receptor targeting enhances adenoviral vector based suicide gene therapy of osteosarcoma. J. Gene Med. 2002, 4, 510–516. [Google Scholar] [CrossRef]
  307. Galetke, W.; Randerath, W.; Feldmeyer, F.; David, M.; Trappe, A.; Ingenabel, F. Importance of routine examinations in patients with obstructive sleep apnea syndrome. Pneumologie 2002, 56, 432–437. [Google Scholar] [CrossRef]
  308. Giaginis, C.T.; Zarros, A.C.; Papaefthymiou, M.A.; Papadopouli, A.E.; Sfiniadakis, I.K.; Theocharis, S.E. Coxsackievirus and adenovirus receptor expression in human endometrial adenocarcinoma: Possible clinical implications. World J. Surg. Oncol. 2008, 6, 59. [Google Scholar] [CrossRef]
  309. Jee, Y.S.; Lee, S.G.; Lee, J.C.; Kim, M.J.; Lee, J.J.; Kim, D.Y.; Park, S.W.; Sung, M.W.; Heo, D.S. Reduced expression of coxsackievirus and adenovirus receptor (CAR) in tumor tissue compared to normal epithelium in head and neck squamous cell carcinoma patients. Anticancer Res. 2002, 22, 2629–2634. [Google Scholar] [PubMed]
  310. Wunder, T.; Schumacher, U.; Friedrich, R.E. Coxsackie adenovirus receptor expression in carcinomas of the head and neck. Anticancer Res. 2012, 32, 1057–1062. [Google Scholar] [PubMed]
  311. Bernal, R.M.; Sharma, S.; Gardner, B.K.; Douglas, J.T.; Bergelson, J.M.; Dubinett, S.M.; Batra, R.K. Soluble coxsackievirus adenovirus receptor is a putative inhibitor of adenoviral gene transfer in the tumor milieu. Clin. Cancer Res. 2002, 8, 1915–1923. [Google Scholar]
  312. Sachs, M.D.; Rauen, K.A.; Ramamurthy, M.; Dodson, J.L.; De Marzo, A.M.; Putzi, M.J.; Schoenberg, M.P.; Rodriguez, R. Integrin alpha(v) and coxsackie adenovirus receptor expression in clinical bladder cancer. Urology 2002, 60, 531–536. [Google Scholar] [CrossRef]
  313. Buscarini, M.; Quek, M.L.; Gilliam-Hegarich, S.; Kasahara, N.; Bochner, B. Adenoviral receptor expression of normal bladder and transitional cell carcinoma of the bladder. Urol. Int. 2007, 78, 160–166. [Google Scholar] [CrossRef]
  314. Korn, W.M.; Macal, M.; Christian, C.; Lacher, M.D.; McMillan, A.; Rauen, K.A.; Warren, R.S.; Ferrell, L. Expression of the coxsackievirus- and adenovirus receptor in gastrointestinal cancer correlates with tumor differentiation. Cancer Gene Ther. 2006, 13, 792–797. [Google Scholar] [CrossRef]
  315. Hoshino, I.; Matsubara, H.; Akutsu, Y.; Nishimori, T.; Yoneyama, Y.; Murakami, K.; Sakata, H.; Matsushita, K.; Komatsu, A.; Brooks, R.; et al. Role of histone deacetylase inhibitor in adenovirus-mediated p53 gene therapy in esophageal cancer. Anticancer Res. 2008, 28, 665–671. [Google Scholar]
  316. Anders, M.; Vieth, M.; Rocken, C.; Ebert, M.; Pross, M.; Gretschel, S.; Schlag, P.M.; Wiedenmann, B.; Kemmner, W.; Hocker, M. Loss of the coxsackie and adenovirus receptor contributes to gastric cancer progression. Br. J. Cancer 2009, 100, 352–359. [Google Scholar] [CrossRef]
  317. Yang, X.; Li, S.; Wang, H.; Chen, W.; Mou, X.; Wang, S. Expression of coxsackie and adenovirus receptor is correlated with inferior prognosis in liver cancer patients. Oncol. Lett. 2019, 17, 2485–2490. [Google Scholar] [CrossRef]
  318. Ma, Y.Y.; Wang, X.J.; Han, Y.; Li, G.; Wang, H.J.; Wang, S.B.; Chen, X.Y.; Liu, F.L.; He, X.L.; Tong, X.M.; et al. Loss of coxsackie and adenovirus receptor expression in human colorectal cancer: A potential impact on the efficacy of adenovirus-mediated gene therapy in Chinese Han population. Mol. Med. Rep. 2016, 14, 2541–2547. [Google Scholar] [CrossRef]
  319. Stecker, K.; Vieth, M.; Koschel, A.; Wiedenmann, B.; Rocken, C.; Anders, M. Impact of the coxsackievirus and adenovirus receptor on the adenoma-carcinoma sequence of colon cancer. Br. J. Cancer 2011, 104, 1426–1433. [Google Scholar] [CrossRef] [PubMed]
  320. Anders, M.; Rosch, T.; Kuster, K.; Becker, I.; Hofler, H.; Stein, H.J.; Meining, A.; Wiedenmann, B.; Sarbia, M. Expression and function of the coxsackie and adenovirus receptor in Barrett’s esophagus and associated neoplasia. Cancer Gene Ther. 2009, 16, 508–515. [Google Scholar] [CrossRef] [PubMed]
  321. Fuxe, J.; Liu, L.; Malin, S.; Philipson, L.; Collins, V.P.; Pettersson, R.F. Expression of the coxsackie and adenovirus receptor in human astrocytic tumors and xenografts. Int. J. Cancer 2003, 103, 723–729. [Google Scholar] [CrossRef] [PubMed]
  322. Li, Y.; Pong, R.C.; Bergelson, J.M.; Hall, M.C.; Sagalowsky, A.V.I.; Tseng, C.P.; Wang, Z.; Hsieh, J.T. Loss of adenoviral receptor expression in human bladder cancer cells: A potential impact on the efficacy of gene therapy. Cancer Res. 1999, 59, 325–330. [Google Scholar]
  323. Matsumoto, K.; Shariat, S.F.; Ayala, G.E.; Rauen, K.A.; Lerner, S.P. Loss of coxsackie and adenovirus receptor expression is associated with features of aggressive bladder cancer. Urology 2005, 66, 441–446. [Google Scholar] [CrossRef]
  324. Huang, K.C.; Altinoz, M.; Wosik, K.; Larochelle, N.; Koty, Z.; Zhu, L.; Holland, P.M.C.; Nalbantoglu, J. Impact of the coxsackie and adenovirus receptor (CAR) on glioma cell growth and invasion: Requirement for the C-terminal domain. Int. J. Cancer 2005, 113, 738–745. [Google Scholar] [CrossRef]
  325. Asaoka, K.; Tada, M.; Sawamura, Y.; Ikeda, J.; Abe, H. Dependence of efficient adenoviral gene delivery in malignant glioma cells on the expression levels of the Coxsackievirus and adenovirus receptor. J. Neurosurg. 2000, 92, 1002–1008. [Google Scholar] [CrossRef]
  326. Yamashita, M.; Ino, A.; Kawabata, K.; Sakurai, F.; Mizuguchi, H. Expression of coxsackie and adenovirus receptor reduces the lung metastatic potential of murine tumor cells. Int. J. Cancer 2007, 121, 1690–1696. [Google Scholar] [CrossRef]
  327. Kuster, K.; Grotzinger, C.; Koschel, A.; Fischer, A.; Wiedenmann, B.; Anders, M. Sodium butyrate increases expression of the coxsackie and adenovirus receptor in colon cancer cells. Cancer Investig. 2010, 28, 268–274. [Google Scholar] [CrossRef]
  328. Kuster, K.; Koschel, A.; Rohwer, N.; Fischer, A.; Wiedenmann, B.; Anders, M. Downregulation of the coxsackie and adenovirus receptor in cancer cells by hypoxia depends on HIF-1alpha. Cancer Gene Ther. 2010, 17, 141–146. [Google Scholar] [CrossRef]
  329. Bruning, A.; Runnebaum, I.B. CAR is a cell-cell adhesion protein in human cancer cells and is expressionally modulated by dexamethasone, TNFalpha, and TGFbeta. Gene Ther. 2003, 10, 198–205. [Google Scholar] [CrossRef] [PubMed]
  330. Thorsteinsson, L.; O’Dowd, G.M.; Harrington, P.M.; Johnson, P.M. The complement regulatory proteins CD46 and CD59, but not CD55, are highly expressed by glandular epithelium of human breast and colorectal tumour tissues. APMIS 1998, 106, 869–878. [Google Scholar] [CrossRef] [PubMed]
  331. Spiller, O.B.; Moretto, G.; Kim, S.U.; Morgan, B.P.; Devine, D.V. Complement expression on astrocytes and astrocytoma cell lines: Failure of complement regulation at the C3 level correlates with very low CD55 expression. J. Neuroimmunol. 1996, 71, 97–106. [Google Scholar] [CrossRef]
  332. Shinoura, N.; Heffelfinger, S.C.; Miller, M.; Shamraj, O.I.; Miura, N.H.; Larson, J.J.; DeTribolet, N.; Warnick, R.E.; Tew, J.J.; Menon, A.G. RNA expression of complement regulatory proteins in human brain tumors. Cancer Lett. 1994, 86, 143–149. [Google Scholar] [CrossRef]
  333. Gasque, P.; Thomas, A.; Fontaine, M.; Morgan, B.P. Complement activation on human neuroblastoma cell lines in vitro: Route of activation and expression of functional complement regulatory proteins. J. Neuroimmunol. 1996, 66, 29–40. [Google Scholar] [CrossRef]
  334. Maenpaa, A.; Junnikkala, S.; Hakulinen, J.; Timonen, T.; Meri, S. Expression of complement membrane regulators membrane cofactor protein (CD46), decay accelerating factor (CD55), and protectin (CD59) in human malignant gliomas. Am. J. Pathol. 1996, 148, 1139–1152. [Google Scholar]
  335. Fukuda, H.; Seya, T.; Hara, T.; Matsumoto, M.; Kinoshita, T.; Masaoka, T. Deficiency of complement decay-accelerating factor (DAF, CD55) in non-Hodgkin’s lymphoma. Immunol. Lett. 1991, 29, 205–209. [Google Scholar] [CrossRef]
  336. Hara, T.; Kojima, A.; Fukuda, H.; Masaoka, T.; Fukumori, Y.; Matsumoto, M.; Seya, T. Levels of complement regulatory proteins, CD35 (CR1), CD46 (MCP) and CD55 (DAF) in human haematological malignancies. Br. J. Haematol. 1992, 82, 368–373. [Google Scholar] [CrossRef]
  337. Kuraya, M.; Yefenof, E.; Klein, G.; Klein, E. Expression of the complement regulatory proteins CD21, CD55 and CD59 on Burkitt lymphoma lines: Their role in sensitivity to human serum-mediated lysis. Eur. J. Immunol. 1992, 22, 1871–1876. [Google Scholar] [CrossRef]
  338. Guc, D.; Canpinar, H.; Kucukaksu, C.; Kansu, E. Expression of complement regulatory proteins CR1, DAF, MCP and CD59 in haematological malignancies. Eur. J. Haematol. 2000, 64, 3–9. [Google Scholar] [CrossRef]
  339. Bomstein, Y.; Fishelson, Z. Enhanced sensitivity of P-glycoprotein-positive multidrug resistant tumor cells to complement-mediated lysis. Eur. J. Haematol. 1997, 27, 2204–2211. [Google Scholar] [CrossRef]
  340. Hiraoka, S.; Mizuno, M.; Nasu, J.; Okazaki, H.; Makidono, C.; Okada, H.; Terada, R.; Yamamoto, K.; Fujita, T.; Shiratori, Y. Enhanced expression of decay-accelerating factor, a complement-regulatory protein, in the specialized intestinal metaplasia of Barrett’s esophagus. J. Lab. Clin. Med. 2004, 143, 201–206. [Google Scholar] [CrossRef] [PubMed]
  341. Murao, T.; Shiotani, A.; Fujita, Y.; Yamanaka, Y.; Kamada, T.; Manabe, N.; Hata, J.; Nishio, K.; Haruma, K. Overexpression of CD55 from Barrett’s esophagus is associated with esophageal adenocarcinoma risk. J. Gastroenterol. Hepatol. 2016, 31, 99–106. [Google Scholar] [CrossRef]
  342. Juhl, H.; Helmig, F.; Baltzer, K.; Kalthoff, H.; Henne-Bruns, D.; Kremer, B. Frequent expression of complement resistance factors CD46, CD55, and CD59 on gastrointestinal cancer cells limits the therapeutic potential of monoclonal antibody 17-1A. J. Surg. Oncol. 1997, 64, 222–230. [Google Scholar] [CrossRef]
  343. He, Z.; Wu, H.; Jiao, Y.; Zheng, J. Expression and prognostic value of CD97 and its ligand CD55 in pancreatic cancer. Oncol. Lett. 2015, 9, 793–797. [Google Scholar] [CrossRef]
  344. Liu, Y.; Chen, L.; Peng, S.; Chen, Z.; Gimm, O.; Finke, R.; Hoang-Vu, C. The expression of CD97EGF and its ligand CD55 on marginal epithelium is related to higher stage and depth of tumor invasion of gastric carcinomas. Oncol. Rep. 2005, 14, 1413–1420. [Google Scholar] [CrossRef]
  345. Hosch, S.B.; Scheunemann, P.; Luth, M.; Inndorf, S.; Stoecklein, N.H.; Erbersdobler, A.; Rehders, A.; Gundlach, M.; Knoefel, W.T.; Izbicki, J.R. Expression of 17-1A antigen and complement resistance factors CD55 and CD59 on liver metastasis in colorectal cancer. J. Gastrointest. Surg. 2001, 5, 673–679. [Google Scholar] [CrossRef]
  346. Shang, Y.; Chai, N.; Gu, Y.; Ding, L.; Yang, Y.; Zhou, J.; Ren, G.; Hao, X.; Fan, D.; Wu, K.; et al. Systematic immunohistochemical analysis of the expression of CD46, CD55, and CD59 in colon cancer. Arch. Pathol. Lab. Med. 2014, 138, 910–919. [Google Scholar] [CrossRef]
  347. Li, L.; Spendlove, I.; Morgan, J.; Durrant, L.G. CD55 is over-expressed in the tumour environment. Br. J. Cancer 2001, 84, 80–86. [Google Scholar] [CrossRef]
  348. Meng, Z.W.; Liu, M.C.; Hong, H.J.; Du, Q.; Chen, Y.L. Expression and prognostic value of soluble CD97 and its ligand CD55 in intrahepatic cholangiocarcinoma. Tumor Biol. 2017, 39, 1010428317694319. [Google Scholar] [CrossRef]
  349. Wu, J.; Lei, L.; Wang, S.; Gu, D.; Zhang, J. Immunohistochemical expression and prognostic value of CD97 and its ligand CD55 in primary gallbladder carcinoma. J. Biomed. Biotechnol. 2012, 2012, 587672. [Google Scholar] [CrossRef] [PubMed]
  350. Mustafa, T.; Eckert, A.; Klonisch, T.; Kehlen, A.; Maurer, P.; Klintschar, M.; Erhuma, M.; Zschoyan, R.; Gimm, O.; Dralle, H.; et al. Expression of the epidermal growth factor seven-transmembrane member CD97 correlates with grading and staging in human oral squamous cell carcinomas. Cancer Epidemiol. Biomark. Prev. 2005, 14, 108–119. [Google Scholar] [CrossRef]
  351. Sukri, A.; Hanafiah, A.; Kosai, N.R.; Mohammed Taher, M.; Mohamed, R. New insight on the role of Helicobacter pylori cagA in the expression of cell surface antigens with important biological functions in gastric carcinogenesis. Helicobacter 2022, 27, e12913. [Google Scholar] [CrossRef] [PubMed]
  352. Kaneko, K.; Zaitoun, A.M.; Letley, D.P.; Rhead, J.L.; Torres, J.; Spendlove, I.; Atherton, J.C.; Robinson, K. The active form of Helicobacter pylori vacuolating cytotoxin induces decay-accelerating factor CD55 in association with intestinal metaplasia in the human gastric mucosa. J. Pathol. 2022, 258, 199–209. [Google Scholar] [CrossRef]
  353. Yamakawa, M.; Yamada, K.; Tsuge, T.; Ohrui, H.; Ogata, T.; Dobashi, M.; Imai, Y. Protection of thyroid cancer cells by complement-regulatory factors. Cancer 1994, 73, 2808–2817. [Google Scholar] [CrossRef]
  354. Steck, T.; Westphal, E.; Wurfel, W. Maternal immunization by husband’s leukocytes for repeated fetal death associated with mild pre-eclampsia—Case report with successful outcome. Arch. Gynecol. Obstet. 1992, 252, 103–107. [Google Scholar] [CrossRef]
  355. Rushmere, N.K.; Knowlden, J.M.; Gee, J.M.; Harper, M.E.; Robertson, J.F.; Morgan, B.P.; Nicholson, R.I. Analysis of the level of mRNA expression of the membrane regulators of complement, CD59, CD55 and CD46, in breast cancer. Int. J. Cancer 2004, 108, 930–936. [Google Scholar] [CrossRef]
  356. Gorter, A.; Blok, V.T.; Haasnoot, W.H.; Ensink, N.G.; Daha, M.R.; Fleuren, G.J. Expression of CD46, CD55, and CD59 on renal tumor cell lines and their role in preventing complement-mediated tumor cell lysis. Lab. Investig. 1996, 74, 1039–1049. [Google Scholar]
  357. Ravindranath, N.M.; Shuler, C. Expression of complement restriction factors (CD46, CD55 & CD59) in head and neck squamous cell carcinomas. J. Oral Pathol. Med. 2006, 35, 560–567. [Google Scholar] [CrossRef]
  358. Kesselring, R.; Thiel, A.; Pries, R.; Fichtner-Feigl, S.; Brunner, S.; Seidel, P.; Bruchhage, K.L.; Wollenberg, B. The complement receptors CD46, CD55 and CD59 are regulated by the tumour microenvironment of head and neck cancer to facilitate escape of complement attack. Eur. J. Cancer 2014, 50, 2152–2161. [Google Scholar] [CrossRef]
  359. Shen, Y.; Yin, R.; Deng, X.; Shen, H. Increased expression of CD55 correlates with tumor progression and poor prognosis in nasopharyngeal carcinoma. Clin. Investig. Med. 2012, 35, E34–E39. [Google Scholar] [CrossRef] [PubMed]
  360. He, Y.; Wang, W.; Xu, L.; Li, L.; Liu, J.; Feng, M.; Bu, H. Immunohistochemical Expression and Prognostic Significance of CD97 and its Ligand DAF in Human Cervical Squamous Cell Carcinoma. Int. J. Gynecol. Pathol. 2015, 34, 473–479. [Google Scholar] [CrossRef] [PubMed]
  361. Kapka-Skrzypczak, L.; Wolinska, E.; Szparecki, G.; Wilczynski, G.M.; Czajka, M.; Skrzypczak, M. CD55, CD59, factor H and factor H-like 1 gene expression analysis in tumors of the ovary and corpus uteri origin. Immunol. Lett. 2015, 167, 67–71. [Google Scholar] [CrossRef] [PubMed]
  362. Nowicki, S.; Nowicki, B.; Pham, T.; Hasan, R.; Nagamani, M. Expression of decay accelerating factor in endometrial adenocarcinoma is inversely related to the stage of tumor. Am. J. Reprod. Immunol. 2001, 46, 144–148. [Google Scholar] [CrossRef]
  363. Murray, K.P.; Mathure, S.; Kaul, R.; Khan, S.; Carson, L.F.; Twiggs, L.B.; Martens, M.G.; Kaul, A. Expression of complement regulatory proteins-CD 35, CD 46, CD 55, and CD 59-in benign and malignant endometrial tissue. Gynecol. Oncol. 2000, 76, 176–182. [Google Scholar] [CrossRef]
  364. Bellone, S.; Roque, D.; Cocco, E.; Gasparrini, S.; Bortolomai, I.; Buza, N.; Abu-Khalaf, M.; Silasi, D.A.; Ratner, E.; Azodi, M.; et al. Downregulation of membrane complement inhibitors CD55 and CD59 by siRNA sensitises uterine serous carcinoma overexpressing Her2/neu to complement and antibody-dependent cell cytotoxicity in vitro: Implications for trastuzumab-based immunotherapy. Br. J. Cancer 2012, 106, 1543–1550. [Google Scholar] [CrossRef]
  365. Varsano, S.; Rashkovsky, L.; Shapiro, H.; Ophir, D.; Mark-Bentankur, T. Human lung cancer cell lines express cell membrane complement inhibitory proteins and are extremely resistant to complement-mediated lysis; a comparison with normal human respiratory epithelium in vitro, and an insight into mechanism(s) of resistance. Clin. Exp. Immunol. 1998, 113, 173–182. [Google Scholar] [CrossRef]
  366. Durrant, L.G.; Chapman, M.A.; Buckley, D.J.; Spendlove, I.; Robins, R.A.; Armitage, N.C. Enhanced expression of the complement regulatory protein CD55 predicts a poor prognosis in colorectal cancer patients. Cancer Immunol. Immunother. 2003, 52, 638–642. [Google Scholar] [CrossRef]
  367. Cimmino, F.; Avitabile, M.; Pezone, L.; Scalia, G.; Montanaro, D.; Andreozzi, M.; Terracciano, L.; Iolascon, A.; Capasso, M. CD55 is a HIF-2alpha marker with anti-adhesive and pro-invading properties in neuroblastoma. Oncogenesis 2016, 5, e212. [Google Scholar] [CrossRef]
  368. Kapka-Skrzypczak, L.; Wolinska, E.; Szparecki, G.; Czajka, M.; Skrzypczak, M. The immunohistochemical analysis of membrane-bound CD55, CD59 and fluid-phase FH and FH-like complement inhibitors in cancers of ovary and corpus uteri origin. Cent. Eur. J. Immunol. 2015, 40, 349–353. [Google Scholar] [CrossRef]
  369. Simpson, K.L.; Jones, A.; Norman, S.; Holmes, C.H. Expression of the complement regulatory proteins decay accelerating factor (DAF, CD55), membrane cofactor protein (MCP, CD46) and CD59 in the normal human uterine cervix and in premalignant and malignant cervical disease. Am. J. Pathol. 1997, 151, 1455–1467. [Google Scholar] [PubMed]
  370. Li, C.; Chen, H.; Hu, L.; Xing, Y.; Sasaki, T.; Villosis, M.F.; Li, J.; Nishita, M.; Minami, Y.; Minoo, P. Ror2 modulates the canonical Wnt signaling in lung epithelial cells through cooperation with Fzd2. BMC Mol. Biol. 2008, 9, 11. [Google Scholar] [CrossRef] [PubMed]
  371. Clines, K.L.; Clines, G.A. DKK1 and Kremen Expression Predicts the Osteoblastic Response to Bone Metastasis. Transl. Oncol. 2018, 11, 873–882. [Google Scholar] [CrossRef]
  372. Sumia, I.; Pierani, A.; Causeret, F. Kremen1-induced cell death is regulated by homo- and heterodimerization. Cell Death Discov. 2019, 5, 91. [Google Scholar] [CrossRef]
  373. Dun, X.; Jiang, H.; Zou, J.; Shi, J.; Zhou, L.; Zhu, R.; Hou, J. Differential expression of DKK-1 binding receptors on stromal cells and myeloma cells results in their distinct response to secreted DKK-1 in myeloma. Mol. Cancer 2010, 9, 247. [Google Scholar] [CrossRef]
  374. Larsen, M.T.; Mandrup, O.N.A.; Schelde, K.K.; Luo, Y.; Sorensen, K.D.; Dagnaes-Hansen, F.; Cameron, J.; Stougaard, M.; Steiniche, T.; Howard, K.A. FcRn overexpression in human cancer drives albumin recycling and cell growth; a mechanistic basis for exploitation in targeted albumin-drug designs. J. Control Release 2020, 322, 53–63. [Google Scholar] [CrossRef]
  375. Kim, M.H.; Lee, J.H.; Lee, J.S.; Kim, D.O.C.; Yang, J.W.; An, H.J.; Na, J.M.; Shin, M.C.; Song, D.H. Fc Receptor Expression as a Prognostic Factor in Patients with Non-small-cell Lung Cancer. Vivo 2022, 36, 2708–2713. [Google Scholar] [CrossRef]
  376. Huang, H.; Hara, A.; Homma, T.; Yonekawa, Y.; Ohgaki, H. Altered expression of immune defense genes in pilocytic astrocytomas. J. Neuropathol. Exp. Neurol. 2005, 64, 891–901. [Google Scholar] [CrossRef]
  377. Cejas, R.B.; Ferguson, D.O.C.; Quinones-Lombrana, A.; Bard, J.E.; Blanco, J.G. Contribution of DNA methylation to the expression of FCGRT in human liver and myocardium. Sci. Rep. 2019, 9, 8674. [Google Scholar] [CrossRef]
  378. Zhu, X.; Meng, G.; Dickinson, B.L.; Li, X.; Mizoguchi, E.; Miao, L.; Wang, Y.; Robert, C.; Wu, B.; Smith, P.D.; et al. MHC class I-related neonatal Fc receptor for IgG is functionally expressed in monocytes, intestinal macrophages, and dendritic cells. J. Immunol. 2001, 166, 3266–3276. [Google Scholar] [CrossRef]
  379. Chapman, N.M. Persistent Enterovirus Infection: Little Deletions, Long Infections. Vaccines 2022, 10, 770. [Google Scholar] [CrossRef] [PubMed]
  380. Saccon, E.; Chen, X.; Mikaeloff, F.; Rodriguez, J.E.; Szekely, L.; Vinhas, B.S.; Krishnan, S.; Byrareddy, S.N.; Frisan, T.; Vegvari, A.; et al. Cell-type-resolved quantitative proteomics map of interferon response against SARS-CoV-2. iScience 2021, 24, 102420. [Google Scholar] [CrossRef] [PubMed]
  381. Smirnova, O.N.A.; Ivanova, O.N.; Fedyakina, I.T.; Yusubalieva, G.M.; Baklaushev, V.P.; Yanvarev, D.V.; Kechko, O.I.; Mitkevich, V.A.; Vorobyev, P.O.; Fedorov, V.S.; et al. SARS-CoV-2 Establishes a Productive Infection in Hepatoma and Glioblastoma Multiforme Cell Lines. Cancers 2023, 15, 632. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Receptors used by enteroviruses for cell entry.
Figure 1. Receptors used by enteroviruses for cell entry.
Cancers 16 03139 g001
Figure 2. A schematic representation of interferon induction system activation during enteroviral infection.
Figure 2. A schematic representation of interferon induction system activation during enteroviral infection.
Cancers 16 03139 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alekseeva, O.N.; Hoa, L.T.; Vorobyev, P.O.; Kochetkov, D.V.; Gumennaya, Y.D.; Naberezhnaya, E.R.; Chuvashov, D.O.; Ivanov, A.V.; Chumakov, P.M.; Lipatova, A.V. Receptors and Host Factors for Enterovirus Infection: Implications for Cancer Therapy. Cancers 2024, 16, 3139. https://doi.org/10.3390/cancers16183139

AMA Style

Alekseeva ON, Hoa LT, Vorobyev PO, Kochetkov DV, Gumennaya YD, Naberezhnaya ER, Chuvashov DO, Ivanov AV, Chumakov PM, Lipatova AV. Receptors and Host Factors for Enterovirus Infection: Implications for Cancer Therapy. Cancers. 2024; 16(18):3139. https://doi.org/10.3390/cancers16183139

Chicago/Turabian Style

Alekseeva, Olga N., Le T. Hoa, Pavel O. Vorobyev, Dmitriy V. Kochetkov, Yana D. Gumennaya, Elizaveta R. Naberezhnaya, Denis O. Chuvashov, Alexander V. Ivanov, Peter M. Chumakov, and Anastasia V. Lipatova. 2024. "Receptors and Host Factors for Enterovirus Infection: Implications for Cancer Therapy" Cancers 16, no. 18: 3139. https://doi.org/10.3390/cancers16183139

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

Article Metrics

Back to TopTop