Next Article in Journal
An Investigation of Culicoides (Diptera: Ceratopogonidae) as Potential Vectors of Medically and Veterinary Important Arboviruses in South Africa
Next Article in Special Issue
SDAV, the Rat Coronavirus—How Much Do We Know about It in the Light of Potential Zoonoses
Previous Article in Journal
Best Molecular Tools to Investigate Coronavirus Diversity in Mammals: A Comparison
Previous Article in Special Issue
The Innate Immune DNA Sensing cGAS-STING Signaling Pathway Mediates Anti-PRRSV Function
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cell Entry of Animal Coronaviruses

1
Department of Pathology and Laboratory Medicine, Western University, 1151 Richmond Street, London, ON N6A 3K7, Canada
2
Institute of Infectious Disease, Beijing Ditan Hospital, Capital Medical University, Beijing 100015, China
3
Beijing Key Laboratory of Emerging Infectious Disease, Beijing 100015, China
4
Molecular Genetics Laboratory, London Health Sciences Centre, 800 Commissioners Road East, London, ON N6A 5W9, Canada
*
Authors to whom correspondence should be addressed.
Viruses 2021, 13(10), 1977; https://doi.org/10.3390/v13101977
Submission received: 10 September 2021 / Revised: 22 September 2021 / Accepted: 23 September 2021 / Published: 1 October 2021
(This article belongs to the Special Issue Enteric and Respiratory Viruses in Animals)

Abstract

:
Coronaviruses (CoVs) are a group of enveloped positive-sense RNA viruses and can cause deadly diseases in animals and humans. Cell entry is the first and essential step of successful virus infection and can be divided into two ongoing steps: cell binding and membrane fusion. Over the past two decades, stimulated by the global outbreak of SARS-CoV and pandemic of SARS-CoV-2, numerous efforts have been made in the CoV research. As a result, significant progress has been achieved in our understanding of the cell entry process. Here, we review the current knowledge of this essential process, including the viral and host components involved in cell binding and membrane fusion, molecular mechanisms of their interactions, and the sites of virus entry. We highlight the recent findings of host restriction factors that inhibit CoVs entry. This knowledge not only enhances our understanding of the cell entry process, pathogenesis, tissue tropism, host range, and interspecies-transmission of CoVs but also provides a theoretical basis to design effective preventive and therapeutic strategies to control CoVs infection.

1. Introduction

Coronaviruses (CoVs) are a group of enveloped, single-stranded positive-sense RNA viruses that cause diseases in the respiratory, gastrointestinal, hepatic, and central nervous systems [1]. The genome size ranges from 26.4 to 31.7 kb, making CoVs the largest RNA viruses known thus far. CoVs are associated with an extremely wide range of vertebrate hosts. So far, 45 CoV species have been identified from avian and mammalian species including human, domestic, wildlife, terrestrial, aerial, and marine animals. These CoVs fall into four genera, Alpha-, Beta-, Gamma-, and Deltacoronavirus, under the Orthocoronavirinae subfamily, Coronaviridae family, Cornidovirineae suborder, and Nidovirale order (https://talk.ictvonline.org/taxonomy/, accessed on 9 September 2021).
CoVs have long been known as economically important pathogens of livestock, poultry, and pet animals. For example, infectious bronchitis virus (IBV) was considered as the number one cause of infectious disease-related economic loss in the UK poultry industry [2]. Porcine transmissible gasteroenteritis virus (TGEV) causes nearly 100% fatality in piglets [3].
Cell entry is the first and essential step of virus infection (Figure 1). Although it is an ongoing process, conceptually, it can be divided into two steps: cell binding and membrane fusion. The former is involved in the binding of viral spike (S) protein to cell surface components, i.e., cellular receptors and/or attachment factors. Following cell binding, the S protein undergoes a conformational change, followed by membrane fusion between virus particles and host cells, and subsequent release of virus genome into the cytoplasm. Membrane fusion occurs either at the plasma membrane in a pH-independent manner, or in the endosome in a pH-dependent manner. Many factors affect virus entry, tissue tropism, persistence, virulence, and host range. These include the specific receptor usage, binding to attachment factors, susceptibility to protease cleavage, acid-induced conformational changes, and host restriction factors [4,5,6].
In this review, we primarily focus on the cell entry of coronaviruses infecting economically significant animals and pets, including porcine, chicken, bovine, canine, and feline coronaviruses, but will also cover the research advances in human coronaviruses (HCoVs), particularly in the discovery of host restriction factors for entry.

2. Overview of Coronaviral S Proteins

As the outermost component of the virion, the S protein is the major viral determinant for cell entry, host range, tissue, and cell tropisms. This is well exemplified by two reciprocal studies showing that murine hepatitis virus (MHV) recombined with feline infectious peritonitis virus (FIPV) S protein acquired the ability to infect feline cells and simultaneously lost the ability to infect mouse cells [7], while FIPV recombined with MHV S protein no longer infected feline cells but was able to infect mouse cells [8].
The S proteins of coronaviruses, varying from 1160 to 1452 amino acids in length, are a type I transmembrane protein that contains a signal peptide, a long N-terminal ectodomain, a transmembrane (TM) domain, and a short C-terminal cytoplasmic tail (Figure 2). The signal peptide directs the nascent S polypeptide to the endoplasmic reticulum (ER), where the signal peptide is cleaved and monomeric S precursors are synthesized. The precursors are heavily glycosylated to yield 150–200 kDa matured monomers, which are further oligomerized into homotrimers in the ER [9].
CoV S proteins are a class I fusion protein, similar to the S proteins of orthomyxoviruses, retroviruses, paramyxoviruses, and filoviruses. Class I fusion proteins have several common features: (i) they are type I transmembrane glycoproteins; (ii) the mature protein is a homotrimer; (iii) they can be cleaved into two noncovalently associated subunits, S1 and S2, by host proteases; (iv) the amino-terminal S1 subunit forms the globular head and is responsible for receptor binding, while the carboxyl-terminal S2 subunit forms a narrow stalk and mediates membrane fusion (Figure 2); and (v) the S2 domain harbors a fusion peptide and two heptad repeats (HRs) that forms a characteristic six-helix bundle structure during membrane fusion [6].
In addition to the role in receptor binding, the S1 domain is also the major determinant to elicit the production of neutralizing antibody. Some anti-S1 antibodies could block receptor binding and protect animals from infection [6,10]. Many coronaviruses have variations in the S1 domain that result in immune response escape. Minor variations as few as one amino acid could change receptor usage, tissue tropism, and virulence [11,12]. Compared to the S2 domain, which is considerably conserved, the S1 domains are diverse even between different strains, and this genetic diversity accounts for the different host range, receptor specificity, virus tropism, antigenicity, and virulence [4].

3. Cell Binding

3.1. Cellular Glycoprotein Receptor

Cellular receptors have been identified for several CoVs (Table 1). The MHV receptor, carcinoembryonic antigen-related cell adhesion molecule 1 (CEACAM1; also called CD66a), is the first entry receptor identified for CoVs [13,14,15]. Full-length CEACAM1 contains four Ig-like domains, but some isoforms have one or two domains missing and thus exhibit less S-binding affinity [13]. Glycosylation of CEACAM1 is required for receptor function [16]. The MHV S-binding domain and neutralizing antibody epitope are overlapping in the N-terminal 108aa of CEACAM1 [17]. Further characterization identified six critical S-binding residues (aa 38–43) on an exposed loop [18,19,20]. Crystal structure and mutagenesis analyses showed that a total of 17 residues in the domain D1 of CEACM1 interact with 14 residues in the N-terminal domain of MHV S1 subunit [21]. Among these, residues I41, R20, and N26 of CEACAM1 and residues I22, Y162, Y162, and Y162 of MHV S protein are particularly important. Although closely related to MHV, rat coronavirus (RCoV) and rat sialodacryoadenitis virus (SDAV) do not use CEACAM1 as an entry receptor [22].
Several alphacoronaviruses, including feline coronavirus (FCoV), canine coronavirus (CCoV), porcine CoVs (TGEV and porcine epidemic diarrhea virus (PEDV)), and HCoV-229E, use aminopeptidase N (APN; also known as CD13) of their own host species as receptors [57,58]. For example, human APN (hAPN) can only be used by HCoV-229E; porcine APN can only be used by TGEV and PEDV; and canine APN can only be used by CCoV. Intriguingly, feline APN can be used by HCoV-229E, FCoV, CCoV, and TGEV [59]. APN is expressed as heavily glycosylated dimers in a variety of cells including epithelial cells of the kidney, respiratory and enteric tracts, endothelial, macrophage, dendritic cell, and cells at synaptic junction site [60]. The high-level expression of APN on respiratory and enteric tracts may be important for virus infection, tissue tropism, and transmission.
Critical domains and residues for virus binding have been identified in the APNs. The introduction of a glycosylation site at residues 291–293 abolished the receptor activity for HCoV-229E [6]. For TGEV, a C-terminal domain, aa 717–813, of porcine APN is essential [61]. Additionally, a C-terminal domain (aa 643–841) of canine APN in the hAPN backbone can mediate entry of CCoV, TGEV, and FCoV [62]. In feline APN, the N-terminal aa 135–295 is required for HCoV-229E, while the C-terminal aa 670–840 is required for FCoV, TGEV, and CCoV [63]. Further characterization mapped the key determinant to aa 288–290 for HCoV-229E, aa 732–746 for TGEV, and both aa 732–746 and 764–788 for FCoV and CCoV. The importance of these determinants was verified by introducing into mouse APN that is not the receptor for these alphacoronaviruses. Introduction of aa 732–746 of feline APN into mouse APN conferred the ability to support TGEV entry, while the introduction of both aa 732–746 and 764–788 conferred receptor ability for FCoV and CCoV [64]. Therefore, the N-terminal domain of APNs is recognized by human CoV, while the C-terminal domain is recognized by animal CoVs. Tusell and Holmes (2007) also showed that residue T742 of feline APN plays a critical role in receptor function [64]. Substitution of threonine with valine at this residue abolished the receptor activity for FCoV, CCoV, and TGEV, whereas a serine substitution at this residue was tolerable, underscoring a role of the hydroxyl group of T742 in the interaction with S proteins. Similarly, a feline APN mutant with substitution of T742 by the hAPN counterpart (R741) no longer supported FCoV, CCoV, and TGEV entry. Conversely, an hAPN mutant with R741T substitution gained receptor activity for TGEV but not for FCoV and CCoV [64].
Surprisingly, HCoV-NL63, an alphacoronavirus that is closely related to HCoV-229E, does not use hAPN as a receptor. Rather, it shares the same receptor, human angiotensin-converting enzyme 2 (ACE2), with two deadly human betacoronaviruses, severe acute respiratory syndrome coronavirus (SARS-CoV) and severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) [65,66,67]. The receptor activity of animal ACE2 orthologs has also been examined. Both SARS-CoV and SARS-CoV-2 are able to use ACE2 from many other animal species (e.g., palm civet, raccoon dog, dog, cat, etc.) as a receptor [68,69,70,71,72,73]. Some SARS-like bat coronaviruses, e.g., WIV1, are also able to use bat ACE2 and multiple animal ACE2 orthologs as a receptor [74,75], highlighting the possibility that these bat CoVs may be able to jump to humans. Another deadly human betacoronavirus, middle east respiratory syndrome coronavirus (MERS-CoV), uses dipeptidyl peptidase 4 (DPP4, also known as CD26) as a functional receptor [76]. Interestingly, APN, ACE2, and DPP4 are all peptidases.

3.2. Receptor-Binding Domains and Residues in the Coronaviral S Proteins

Receptor binding is primarily mediated by an independently folded domain, called receptor-binding domain (RBD), within the S1 domain of S protein. RBDs have been mapped to aa 407–547 for HCoV-229E [32,33], aa 506–655 for TGEV [23], aa 1–330 for MHV [38], aa 318–510 for SARS [46,77], aa 358 to 588 for MERS-CoV [44], and aa 331–524 for SARS-CoV-2 [50] (Table 1). All these RBDs are located at the C-terminal domain of the S1 (S1-CTD), except for the MHV RBD that is located at the N-terminal domain of the S1 (S1-NTD).
Critical receptor-binding residues in the RBDs have been identified for several CoVs by mutagenesis analysis and crystal structure determination of receptor-RBD complex [21,58,78,79]. For example, residues 33, 62, 79, 82, 162, 183, 212, 214, and 216 in the MHV RBD are important for virus infection [21,80,81]. Unexpectedly, residue G29 was found to be critical for the conformational changes triggered by either receptor binding or high pH [82].
Based on the crystal structure of SARS-CoV RBD complexed with hACE2 [83], a total of 14 amino acids in the RBD interact with ACE2. Among these residues, N479 and T487 are critical for SARS-CoV to jump from the amplifying host civet to humans and contribute to the pathogenicity. Sequence analysis showed that all the severe human isolates have N479 and T487; all the mild human isolates have N479 and S487, while all the animal isolates have K479 and S487. Replacement of K479 and S487 in civet isolate with N479 and T487 significantly enhanced hACE2-binding affinity and virus entry [68].

3.3. Attachment Factors

3.3.1. Sialic Acid (SA)

Sialic acid (SA) refers to a group of N- or O-substituted derivatives of neuraminic acid, a monosaccharide with a nine-carbon frame. They are normally conjugated with glycoproteins or glycolipids and are widely distributed in animal tissues. While some studies have claimed that SA functions as a receptor for HCoV-OC43 and BCoV [84], a more recent paper demonstrated that the binding of the S protein to SA did not trigger a conformational change in HCoV-OC43 [85]. Therefore, we still consider SA to be an attachment factor in CoVs that enhances virus binding to host cells and contributes to entry.
SA-binding activity of the S protein has been implicated in the enterotropism of TGEV [24]. TGEV is able to replicate in both respiratory and enteric epithelial cells and causes severe diarrhea in newborn piglets [86]. The S protein can induce erythrocyte agglutination, which is mediated by the S-SA interaction. TGEV mutant, with seven mutations in the S1-NTD, was found to lose the hemagglutinating (HA) activity and enteropathogenicity [12,87], suggesting a link between SA-binding ability and the enteropathogenicity. Interestingly, a naturally occurring TGEV variant, porcine respiratory coronavirus (PRCoV), with N-terminal 227 aa deletion in the S protein, also loses the HA activity and only causes moderate respiratory tract infection [24]. Since PRCoV and other HA-deficient mutants can still replicate in cultured cells, it seems that SA-binding activity is important only in vivo. Based on the finding that wt TGEV, but not HA-deficient mutant, could bind to a carbohydrate-rich mucin-like sialoglycoprotein (MGP), which is abundantly expressed in jejunal goblet cells [88,89], it was proposed that SA-binding activity may assist TGEV in access to its target cells. It should be noted that porcine APN-expressing enterocytes in the intestine are covered by thick layers of mucus and glycocalix that are rich in carbohydrates, including SA. Binding to MGP via SA may allow TGEV to stay longer in the intestine and more easily cross the layers to initiate infection in enterocytes [88,89].
Bovine coronavirus (BCoV) and HCoV-OC43 are closely related betacoronaviruses [39]. Both the S proteins exhibit HA activity and can use N-acetyl-9-O-sialic acid (Neu5,9Ac2) as an attachment factor [41,43]. Removal of SAs from the cell surface inhibits virus infection in cultured cells, while resialylation restores infection. Therefore, the role of SA-binding activity in infection is more important in BCoV/HCoV-OC43 infection than in TGEV infection. However, there is a subtle difference in recognizing SA by these two CoVs. BCoV is more efficient in recognizing α2,3-linked form, whereas HCoV-OC43 prefers the α2,6-linkage [90]. It is unclear if this moiety must be linked to a specific glycoprotein or glycolipid or if there is a specific protein receptor for these two viruses. It has been reported that HCoV-OC43 can bind to a major histocompatibility complex class I molecule [91], but its receptor function has never been established.
SA also plays an important role in IBV infection. IBV S protein also has HA activity and preferentially binds to α2,3-linked SA. Removal of SAs from the surface of Vero, hamster kidney, chicken embryonic kidney cells, and trachea epithelial cells with neuraminidase reduced their susceptibility to IBV infection. Therefore, SA might serve as the primary binding determinant for IBV to facilitate the binding to its authentic protein receptor, though tighter binding and subsequent membrane fusion may require interaction with a second receptor [54,92].
In contrast to most of the RBDs that are located at the C-terminal part of the S1 domain (S1-CTD), the N-terminal part of the S1 (S1-NTD) has been known to bind to SA, which is mediated by a sugar-binding structure similar to human galectin in the S1-NTD [21]. Interestingly, the S1-NTDs of IBV, BCoV, HCoV-OC43, TGEV, and MERS-CoV, but not those of MHV, HCoV-HKU1, and SARS-CoV, possess the sugar-binding ability [21,78,93]. This difference was found to be due to the different lengths of a loop in S1-NTDs among these CoVs [21].

3.3.2. Heparan Sulfate (HS)

Heparan sulfate (HS) is a common type of glycosaminoglycans (GAGs), a group of long unbranched polysaccharides that are composed of repeating disaccharide units. GAGs are attached to serine residues on proteoglycans during posttranslational modifications in the Golgi apparatus [94]. HS is found in the surfaces of most mammalian cell types and extracellular matrices.
In coronaviruses, using HS as an attachment factor is often a result of cell culture adaptation. The S protein of FCoV-UCD, a field isolate that does not grow in cell culture, is cleaved at a furin-recognition site, RRSRR, but cannot bind to heparin, an analog of cellular HS. On the contrary, a cell culture-adapted strain, UCD1, has an R-to-G substitution at the furin motif (RRSRG), which otherwise creates a heparin-binding motif (SRRSRG). Consistent with these genetic changes, the S protein is not cleaved but can bind to heparin. Meanwhile, the entry of UCD1 into FCWF cells became HS-dependent [35]. Similarly, persistent infection of MHV-A59 in murine cell culture resulted in an emergence of variant, MHV-BHK, which extended its host range by acquiring the ability to use HS for cell entry due to the acquisition of two HS-binding sites in its S protein [95,96]. Additionally, the S protein of this variant could no longer be cleaved, although the furin-cleavage motif is still there [95]. Together, these data suggest that acquisition of HS-binding ability is traded off for the loss of furin cleavage during cell culture adaptation. Unlike MHV-BHK, a neurotropic strain of MHV, JHMV, only binds to HS on the cell surface but does not use it as a receptor [97]. HS has also been shown to assist in SARS-CoV and SARS-CoV-2 entry by binding to the S protein directly, facilitating cell attachment and entry [47]. In gammacoronaviruses, HS has been considered to act as an attachment factor to mediate infection of IBV-Beaudette, an embryo-adapted strain that has extended the host range [55].

3.3.3. C-Type Lectins

DC-SIGN (dendritic cell-specific ICAM-3-grabbing nonintegrin) and its homolog L-SIGN (for liver/lymph node-specific; also called CD209L or DC-SIGNR) are type II C-type lectins. DC/L-SIGN mediate cell adhesion and pathogen recognition. They have been implicated in the entry, infection, and transmission of many different viruses [98,99,100,101]. On one hand, they enhance virus entry and facilitate infection in cells expressing the cognate receptor. This is called cis-infection. On the other hand, they capture and transfer virus particles to target cells. This is called trans-infection [100,102,103]. DC-SIGN has been found to reside within a well-defined cholesterol-rich microdomain, called lipid raft, on the plasma membrane, which may act as a docking site for viruses to invade host cells [104].
The role of DC/L-SIGN in SARS-CoV entry is controversial. It has been shown that DC-SIGN and/or L-SIGN could serve as an attachment factor to augment the entry of SARS-CoV [48,49], HCoV-NL63 [31], and FIPV [36]. DC/L-SIGN has also been confirmed to promote viral infection in IBV, TGEV, and the newly emerged SARS-CoV-2 [25,51,56]. However, some studies showed that DC/L-SIGN are alternative and independent albeit less efficient receptors for HCoV-229E [34] and SARS-CoV [105,106]. Interaction with DC/L-SIGN requires at least seven glycosylation sites (N109, N118, N119, N158, N227, N589, and N699) in the SARS-CoV S protein, which is outside of the ACE2-binding domain (aa 318-510) [105].
L-SECtin, another C-type lectin co-expressed with L-SIGN on sinusoidal endothelial cells in liver and lymph nodes also promotes SARS entry, but not through binding to high-mannose glycan as DC/L-SIGN do [107].

4. Membrane Fusion

4.1. Structure and Function of the S2 Domain

Membrane fusion is mediated by the S2 domain of coronaviral S proteins. The S2 domain contains several structurally conserved regions: fusion peptide (FP), two heptad repeats (HRs), TM domain, and cytoplasmic tail (Figure 2), all of which contribute to membrane fusion. The fusion peptide is highly hydrophobic and is able to cause leakage of liposomes [108]. Mutational analysis confirms that fusion peptide is important for membrane fusion [109,110,111].
The most striking feature of the S2 domain is two HRs. HR1 is located in the N-terminus, right downstream of the fusion peptide, and HR2 is located in the C-terminus, upstream of the TM domain. HR is a motif with seven-residue periodicity. The first and fourth residues are typically hydrophobic, whereas the others are hydrophilic. This motif is the basis of the α-helix coiled-coil structure and plays a central role in membrane fusion [112].
Based on the current model of membrane fusion of class I fusion proteins (Figure 3) [113], the S protein has three conformational states. Under native, pre-fusion state (state 1), the S protein is metastable, in which the S1 is associated with the S2. During this period, the fusion peptide is not exposed, and two HRs form trimeric helix coiled-coil individually. Following receptor binding, the S1 sheds from the S2. Subsequently, primed by the proteolytic cleavage at the S1/S2 boundary and/or at the S2′ site, the fusion peptide exposes and is inserted into the cell membrane. This state is known as the intermediate state (state 2). At state 3, HR2 folds back to HR1 and together forms a six-helix bundle, in which HR1 trimeric coiled-coil lies in the inner core that is wrapped by an outer layer of antiparallel HR2 coiled-coil [109,112,113,114,115]. The association of HR1 and HR2 represents the postfusion state. This six-helix bundle brings the fusion peptide, TM domain, and cell membrane into close proximity, destabilizing the lipid bilayers, forming a fusion pore, and releasing the viral genome into the cytoplasm [113,116,117]. HRs are one of the major targets for antiviral therapy. HR peptides or HR-specific antibodies can efficiently inhibit CoVs infection by blocking the formation of six-helix bundles [109,118,119,120].
Other regions in the S2 domain are also involved in membrane fusion. The region between two HRs is called the interhelica (IH) domain. It has been shown in MHV and SARS-CoV that a hydrophobic region of the IH domain is important for cell-to-cell fusion [110,111]. A short region immediately upstream of the TM domain, called juxtamembrane domain (JMD), is rich in aromatic residues (KWPWYVWL) and is highly conserved in CoVs. Alanine substitution of the aromatic residues results in reduced virus-to-cell fusion and cell-to-cell fusion [121].
The TM domain is highly conserved with up to 30% amino acid identity among different CoVs. Studies with SARS-CoV and MHV show that the TM domains contribute to the stability of S trimers and membrane fusion [122,123,124].
The cytoplasmic tail contains conserved sequences that are important for S protein trafficking through the cellular secretory pathway, virus particle assembly, and membrane fusion [125]. An ER-targeting signal is located at the very end of cytoplasmic tail. In IBV, the signal is a dilysine motif (KKSV); mutations of the residues KK cause faster trafficking of IBV S proteins to the cell surface, virus growth defect, and premature syncytia formation [126]. The S proteins of alphacoronaviruses and IBV also contain a tyrosine-based endocytic signal (YXXθ, where θ represents a bulky hydrophobic residue) upstream of the ER-targeting signal. This motif binds to the AP2 adaptor complex that, in turn, binds to clathrin to induce endocytosis [127]. Youn et al. (2005) demonstrated that this signal rapidly endocytosed IBV S proteins from the plasma membrane [126]. Deletion of such an ER-targeting signal and endocytic signal significantly enhanced cell entry of TGEV [128], HCoV-NL63 [129], SARS-CoV [130], and SARS-CoV-2 [73]. In the N-terminal and central regions of the cytoplasmic tail, there are four conserved cysteine clusters. These cysteines are important for membrane fusion. In particular, the clusters I and II cysteines near the TM domain play a major role [122,124,131]. These two cysteine clusters are also the major sites for palmitoylation in SARS-CoV S protein [131]. Palmitoylation in MHV S is critical for cell-to-cell fusion, assembly, and infectivity [132].

4.2. Cleavage of S Proteins by Host Proteases

Proteolytic cleavage of the S proteins is a critical priming step for membrane fusion mediated by class I fusion proteins. Coronaviral S proteins can be cleaved at different sites. Two main types of protease cleavage sites, S1/S2 and S2′, have been identified. The former is located at the S1/S2 boundary; the latter is located right upstream of the fusion peptide (Figure 2) [5]. Depending on the cell type, S proteins are cleaved at different stages of the virus life cycle and at different cellular sites, e.g., during the S protein biogenesis at the ER/Golgi and trans-Golgi compartment, on the plasm membrane or in endosome/lysosome during cell entry, and in the extracellular space during infection. This spatially and temporally controlled and cell-type-dependent cleavage modulates the pathogenicity, cell and tissue tropism, and host range of coronaviruses.
A number of host proteases have been identified to activate membrane fusion of CoVs. These include trypsin, furin, cathepsins, TMPRSS, elastase, etc. [5]. Among these, furin, trypsin, cathepsin L and TMPRSS could cleave at the S1/S2 site, while trypsin and elastase could cleave at the S2ʹ site. In general, furin cleavage occurs during biosynthesis; trypsin and TMPRSS cleavage occur in the extracellular space and plasm membrane; cathepsin cleavage occurs in endosome/lysosome (Figure 1).
For some CoVs, e.g., IBV, MHV, CCoV, and FCoV-I, the S proteins have a multibasic furin cleavage motif (often RRXRR) at the S1/S2 boundary. Thus, these S proteins are cleaved during biosynthesis by furin-like protease into S1 and S2 subunits prior to assembly into virion [35,133]. Furin cleavage is not essential for virus–cell-membrane fusion, but it enhances cell-to-cell fusion [122,134,135,136]. Persistent infection of cell culture often leads to loss of furin cleavage site, and such cell-adapted variants are generally attenuated in vivo [134]. For MHV-A59 strain, the S protein is cleaved a second time at the S2ʹ site during virus entry [137]. The S protein of IBV is also cleaved at the S2ʹ site during protein biosynthesis and during infection, and this cleavage has been shown to be important for virus entry [138].
For those CoVs that do not have furin-cleavage sites, e.g., PEDV and FCoV-II, other proteases, including TMPRSS2, trypsin, or cathepsin, are used. For example, PEDV S protein was found to be cleaved at the S2ʹ site by trypsin after receptor binding [139], and the virus entry was shown to be dependent on a low pH and endosomal cathepsin [140]. Cleavage at the S2ʹ site by cathepsin was also found to be critical for FCoV-II entry [141]. Proteolytic cleavage by TMPRSS2 has been found to promote entry at the plasm membrane for SARS-CoV [142], MERS-CoV [143], and SARS-CoV-2 [144]. In addition to TMPRSS2, SARS-CoV also uses other proteases, e.g., cathepsin L, trypsin, thermolysin, elastase, and factor Xa, to cleave the S protein and promote entry [145,146,147].
Gain or loss of proteolytic cleavage site is often associated with changes in the entry route, dependence on pH, and cell-to-cell fusion ability. HCoV-229E S protein does not have a furin cleavage site but is cleaved by cathepsin L and maybe other proteases. This cleavage is required for membrane fusion in acidic endosomes. However, if the virus is treated with trypsin before infection, the S protein will be cleaved, coupled with cell surface entry under neutral pH and enhanced syncitia formation [148]. This phenomenon is also observed in SARS-CoV infection with trypsin treatment; furthermore, the introduction of furin cleavage site into SARS-CoV S results in a cleaved S that mediates entry at the cell surface in a pH-independent fashion [146,149]. Similarly, MHV-A59 expresses a cleavable S protein that mediates entry at the cell surface and cell-to-cell fusion under neutral pH. In contrast, MHV-2 expresses an uncleaved S protein that mediates cathepsin-dependent entry in the acidic endosome and cannot mediate cell-to-cell fusion. Recombinant virus bearing a cleaved MHV-2 S protein has an A59-like phenotype in entry [150]. These findings suggest that, during in vivo infection, HCoV-229E, SARS-CoV, and MHV-2 may bypass the endocytosis pathway and directly enter target cells from the cell surface, since their S proteins may be cleaved by various proteases secreted extracellularly.

5. Host Restriction Factors for Coronavirus Entry

5.1. IFITM

Interferon-induced transmembrane proteins (IFITMs), more specifically, IFITM1, -2, or -3, have been shown to inhibit cell entry of many different enveloped RNA viruses [151,152,153]. Several HCoVs, including SARS-CoV [154], SARS-CoV-2 [155], 229E [156], and MERS-CoV [157], have been found to be sensitive to IFITMs restriction.
How IFITMs inhibit virus entry is not completely understood. The current model proposes that IFITMs inhibit virus-cell fusion at the hemifusion or pore formation stage by modifying the rigidity and/or curvature of the membranes in which they reside [158]. IFITM1 mostly resides at the plasma membrane, while IFITM2 and -3 are more found in endosome/lysosomal. A compelling model proposes that IFITM2/3 block virus entry at endosome/lysosome by interrupting the membrane fusion [151]. The following evidence supports this model: (i) IFITMs-insensitive viruses fuse at the plasma membrane; (ii) mutant IFITM3 that could not be internalized into endosome/lysosome had impaired restriction activity against influenza virus; (iii) restriction of SARS-CoV, SARS-CoV-2, and bat CoV WIV1 by IFITMs could be circumvented by pre-treatment with trypsin or TMPRSS2 that bypasses the dependence on lysosomal cathepsin L [75,154,159]; and (iv) removing the furin cleavage site in the S protein of SARS-CoV-2 made the S proteins remain noncleaved during biosynthesis and the entry more dependent on low pH and endosome/lysosome, and as a result, the entry is more sensitive to IFITM2 restriction [155].
IFITMs do not always inhibit viral entry. HCoV-OC43 could use IFITM2 or IFITM3 to promote its cell entry and infection, presumably by modulating membrane fusion [153,160]. It remains largely unknown how IFITMs promote CoV entry. Recent studies have indicated that specific mutations can switch the biological activity of IFITMs from inhibiting to enhancing the infection of selected human coronaviruses [161]. Three distinct mutations in IFITM1 or IFITM3 were found to allow the host restriction factors to promote viral entry in SARS-CoV and/or MERS-CoV. For example, an alanine or aspartic acid replacement of IFITM3 tyrosine 20 to mimic unphosphorylated or phosphorylated IFITM3 reduced its inhibition activity on HCoV-NL63 and -229E entry but enhanced the entry of SARS-CoV and MERS-CoV. These findings suggest that the residues and structural motifs of IFITMs are vital for modulating HCoV entries, likely through viral/host cellular components’ interactions at the site of entry to modulate membrane fusion [161].

5.2. LY6E

Lymphocyte antigen 6 family member E (LY6E) potently restricts infection by multiple CoVs, including MHV and several HCoVs (229E, OC43, MERS-CoV, SARS-CoV, SARS-CoV-2) [160,162]. Mechanistic studies revealed that LY6E inhibits coronavirus entry into cells by interfering with S protein-mediated membrane fusion [162]. Mice lacking LY6E in immune cells were highly susceptible to mouse hepatitis virus (MHV). In LY6E knockout mice, an aggravated viral pathogenesis was accompanied by loss of hepatic immune cells, higher splenic viral burden, and reduction in global antiviral gene pathways. Furthermore, constitutive LY6E directly protects primary B cells from murine coronavirus infection [162]. While overexpression of amphotericin treatment or TMPRSS2 neutralized the restriction of IFITM3 on virus entry, LY6E restriction on human CoV entry was not affected [160]. These findings suggest that LY6E controls CoV infection via a mechanism distinct from IFITMs.

5.3. CD74

CD74 has been recently identified as a direct broad-range antiviral effector protein [163]. Among the four isoforms of CD74, only p41 isoform is able to block the endosomal entry pathway of EBOV and coronaviruses, including SARS-CoV, SARS-CoV-2, and Bat coronavirus WIV1 [163]. p41 is up-regulated by the major histocompatibility complex (MHC) class II transactivator (CIITA) but inhibits virus entry independently of CIITA expression. The p41 isoform contains the thyroglobulin domain, lacks the ER retention signal, and normally accumulates in endosomes. The CD74 thyroglobulin domain inhibits cathepsins, suggesting that this may be the mechanism for antiviral activity. Disruption of the p41 CTSL binding site by mutation completely inhibited antiviral activity.

5.4. Ezrin

Ezrin, a membrane actin-linker, was discovered as a CoV entry restrictor by using the cytoplasmic tail of SARS-CoV S protein as bait in a yeast two-hybrid screen [164]. Further characterization showed that the F1 lobe of the FERM domain of Ezrin interacts with both the eight C-terminal residues and the conserved cysteine cluster near the TM domain. Mechanistic study implies that Ezrin inhibits SARS-CoV entry possibly by restricting the fusion pore opening and trapping incoming particles within the intracellular filamentous actin network [164].

5.5. GILT

Recently, we found that gamma-interferon-inducible lysosomal thiol reductase (GILT) could inhibit pseudotyped virus entry of multiple enveloped RNA viruses, including SARS-CoV [165]. GILT is constitutively expressed in lung epithelial cells and fibroblasts. Its expression could be further induced by type II interferon. We also found that GILT expression reduced the level and activity of endosome/lysosome-associated cathepsin L, while trypsin treatment could abrogate the entry restriction by GILT. These findings suggest that GILT exerts its restriction in the endosome/lysosome.

5.6. CH25H

Cholesterol 25-hydroxylase (CH25H) is also an interferon (IFN)-stimulated gene. The enzyme converts cholesterol to 25-hydrocholesterol (25HC) that has been known to possess broad antiviral activities against a wide range of enveloped viruses by blocking membrane fusion. Most recently, 25HC was shown to accumulate in the late endosomes and inhibit SARS-CoV-2 infection [166,167]. Mechanistically, this inhibitor may be achieved by activating the ER-localized acyl-CoA:cholesterol acyltransferase (ACAT) that depletes cholesterol from the plasma membrane, which leads to interruption of membrane fusion.

6. Perspective of Future Study

Coronavirus research has been significantly stimulated twice over the past two decades. The first was triggered by the outbreak of SARS-CoV during 2002-2003 that caused 774 deaths in 8096 infected people in the world (http://www.who.int/csr/sars/country/table2004_04_21/en/index.html, accessed on 9 September 2021). The second was initiated by the global pandemic of SARS-CoV-2 since December 2019. As of 30 August 2021, this virus has caused ~220 million confirmed cases and nearly 4.5 million deaths (https://COVID19.who.int/, accessed on 9 September 2021). As a result, numerous efforts have been made, and accordingly, significant progress has been made in every direction of CoV research. The most striking finding in the field of CoV entry over the past ten years is probably the identification of a number of host restriction factors. These factors share several common features. First, they are all interferon-stimulated genes (ISGs). Second, they are against a broad range of RNA viruses. Third, they work by inhibiting the membrane fusion during virus entry. Despite this progress, we still have very limited information as to how these restriction factors exert their inhibitor effect mechanistically. Additionally, there are hundreds of ISGs. It was estimated that nearly 10% of human genes are able to be stimulated by interferons [168]. We speculate that there will be more ISGs that are able to inhibit CoVs entry. This field warrants more studies.
Another important field to pursue in the future is the identification of host factors, including receptors, attachment factors, and other host proteins, that promote cell entry of CoVs. So far, 45 CoV species have been recognized, but only a handful of them have receptors identified (Table 1). Receptors of several important animal CoVs, e.g., IBV, BCoV, and swine acute diarrhea syndrome coronavirus (SADS-CoV), remain unknown. SADS-CoV is a novel emerging virus that originated from a bat coronavirus by cross-species transmission. It caused a large-scale outbreak of fatal disease in pigs, resulting in 24,693 piglet deaths across four farms in China [169]. Initial studies showed that several known receptors for CoVs, including ACE2, APN, and DPP4, are not the receptor for SADS-CoV [169,170].
Significant progress has also been made in understanding the S-receptor molecular interaction by determining the crystal structures of their complex. However, such information is still lacking for many important CoVs and thus further investigation is needed.

Author Contributions

H.L. and X.Z. conceptualized the structure of and revised this manuscript. Y.-R.C. prepared the original draft. X.L. prepared the figures. All authors have read and agreed to the published version of the manuscript.

Funding

X.Z. was funded by the National Science Foundation of China (81772173 and 81971916) and the National Science and Technology Mega-Project of China (2018ZX10301-408-002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lai, M.M.; Holmes, K.V. Coronaviridae: The Viruses and Their Replication. In Fields Virology; Knipe, D.M., Howley, P.M., Eds.; Lippincott-Raven: Philadelphia, PA, USA, 2001; pp. 1163–1186. [Google Scholar]
  2. Cavanagh, D. Coronavirus Avian Infectious Bronchitis Virus. Vet. Res. 2007, 38, 281–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Laude, H.; Rasschaert, D.; Delmas, B.; Godet, M.; Gelfi, J.; Charley, B. Molecular Biology of Transmissible Gastroenteritis Virus. Vet. Microbiol. 1990, 23, 147–154. [Google Scholar] [CrossRef]
  4. Gallagher, T.M.; Buchmeier, M.J. Coronavirus Spike Proteins in Viral Entry and Pathogenesis. Virology 2001, 279, 371–374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Millet, J.K.; Whittaker, G.R. Host Cell Proteases: Critical Determinants of Coronavirus Tropism and Pathogenesis. Virus Res. 2015, 202, 120–134. [Google Scholar] [CrossRef] [PubMed]
  6. Wentworth, D.E.; Holmes, K.V. Coronavirus Binding and Entry. In Coronaviruses: Molecular and Cellular Biology; Thiel, V., Ed.; Caister Academic Press: Norfolk, UK, 2007; pp. 3–30. [Google Scholar]
  7. Kuo, L.; Godeke, G.J.; Raamsman, M.J.; Masters, P.S.; Rottier, P.J. Retargeting of Coronavirus by Substitution of the Spike Glycoprotein Ectodomain: Crossing the Host Cell Species Barrier. J. Virol. 2000, 74, 1393–1406. [Google Scholar] [CrossRef] [Green Version]
  8. Haijema, B.J.; Volders, H.; Rottier, P.J.M. Switching Species Tropism: An Effective Way to Manipulate the Feline Coronavirus Genome. J. Virol. 2003, 77, 4528–4538. [Google Scholar] [CrossRef] [Green Version]
  9. Masters, P.S. The Molecular Biology of Coronaviruses. Adv. Virus Res. 2006, 66, 193–292. [Google Scholar] [CrossRef]
  10. Du, L.; He, Y.; Zhou, Y.; Liu, S.; Zheng, B.-J.; Jiang, S. The Spike Protein of SARS-CoV-a Target for Vaccine and Therapeutic Development. Nat. Rev. Microbiol. 2009, 7, 226–236. [Google Scholar] [CrossRef]
  11. Ballesteros, M.L.; Sánchez, C.M.; Enjuanes, L. Two Amino Acid Changes at the N-Terminus of Transmissible Gastroenteritis Coronavirus Spike Protein Result in the Loss of Enteric Tropism. Virology 1997, 227, 378–388. [Google Scholar] [CrossRef] [Green Version]
  12. Krempl, C.; Schultze, B.; Laude, H.; Herrler, G. Point Mutations in the S Protein Connect the Sialic Acid Binding Activity with the Enteropathogenicity of Transmissible Gastroenteritis Coronavirus. J. Virol. 1997, 71, 3285–3287. [Google Scholar] [CrossRef] [Green Version]
  13. Dveksler, G.S.; Dieffenbach, C.W.; Cardellichio, C.B.; McCuaig, K.; Pensiero, M.N.; Jiang, G.S.; Beauchemin, N.; Holmes, K.V. Several Members of the Mouse Carcinoembryonic Antigen-Related Glycoprotein Family Are Functional Receptors for the Coronavirus Mouse Hepatitis Virus-A59. J. Virol. 1993, 67, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Dveksler, G.S.; Pensiero, M.N.; Cardellichio, C.B.; Williams, R.K.; Jiang, G.S.; Holmes, K.V.; Dieffenbach, C.W. Cloning of the Mouse Hepatitis Virus (MHV) Receptor: Expression in Human and Hamster Cell Lines Confers Susceptibility to MHV. J. Virol. 1991, 65, 6881–6891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Williams, R.K.; Jiang, G.S.; Holmes, K.V. Receptor for Mouse Hepatitis Virus Is a Member of the Carcinoembryonic Antigen Family of Glycoproteins. Proc. Natl. Acad. Sci. USA 1991, 88, 5533–5536. [Google Scholar] [CrossRef] [Green Version]
  16. Pensiero, M.N.; Dveksler, G.S.; Cardellichio, C.B.; Jiang, G.S.; Elia, P.E.; Dieffenbach, C.W.; Holmes, K.V. Binding of the Coronavirus Mouse Hepatitis Virus A59 to Its Receptor Expressed from a Recombinant Vaccinia Virus Depends on Posttranslational Processing of the Receptor Glycoprotein. J. Virol. 1992, 66, 4028–4039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Dveksler, G.S.; Pensiero, M.N.; Dieffenbach, C.W.; Cardellichio, C.B.; Basile, A.A.; Elia, P.E.; Holmes, K.V. Mouse Hepatitis Virus Strain A59 and Blocking Antireceptor Monoclonal Antibody Bind to the N-Terminal Domain of Cellular Receptor. Proc. Natl. Acad. Sci. USA 1993, 90, 1716. [Google Scholar] [CrossRef] [Green Version]
  18. Rao, P.V.; Kumari, S.; Gallagher, T.M. Identification of a Contiguous 6-Residue Determinant in the MHV Receptor That Controls the Level of Virion Binding to Cells. Virology 1997, 229, 336–348. [Google Scholar] [CrossRef] [Green Version]
  19. Tan, K.; Zelus, B.D.; Meijers, R.; Liu, J.; Bergelson, J.M.; Duke, N.; Zhang, R.; Joachimiak, A.; Holmes, K.V.; Wang, J. Crystal Structure of Murine SCEACAM1a [1,4]: A Coronavirus Receptor in the CEA Family. EMBO J. 2002, 21, 2076–2086. [Google Scholar] [CrossRef] [Green Version]
  20. Wessner, D.R.; Shick, P.C.; Lu, J.H.; Cardellichio, C.B.; Gagneten, S.E.; Beauchemin, N.; Holmes, K.V.; Dveksler, G.S. Mutational Analysis of the Virus and Monoclonal Antibody Binding Sites in MHVR, the Cellular Receptor of the Murine Coronavirus Mouse Hepatitis Virus Strain A59. J. Virol. 1998, 72, 1941–1948. [Google Scholar] [CrossRef] [Green Version]
  21. Peng, G.; Sun, D.; Rajashankar, K.R.; Qian, Z.; Holmes, K.V.; Li, F. Crystal Structure of Mouse Coronavirus Receptor-Binding Domain Complexed with Its Murine Receptor. Proc. Natl. Acad. Sci. USA 2011, 108, 10696–10701. [Google Scholar] [CrossRef] [Green Version]
  22. Gagneten, S.; Scanga, C.A.; Dveksler, G.S.; Beauchemin, N.; Percy, D.; Holmes, K.V. Attachment Glycoproteins and Receptor Specificity of Rat Coronaviruses. Lab. Anim. Sci. 1996, 46, 159–166. [Google Scholar]
  23. Godet, M.; Grosclaude, J.; Delmas, B.; Laude, H. Major Receptor-Binding and Neutralization Determinants Are Located within the Same Domain of the Transmissible Gastroenteritis Virus (Coronavirus) Spike Protein. J. Virol. 1994, 68, 8008–8016. [Google Scholar] [CrossRef] [Green Version]
  24. Schultze, B.; Krempl, C.; Ballesteros, M.L.; Shaw, L.; Schauer, R.; Enjuanes, L.; Herrler, G. Transmissible Gastroenteritis Coronavirus, but Not the Related Porcine Respiratory Coronavirus, Has a Sialic Acid (N-Glycolylneuraminic Acid) Binding Activity. J. Virol. 1996, 70, 5634–5637. [Google Scholar] [CrossRef] [Green Version]
  25. Zhao, P.; Xu, L.-D.; Zhang, Y.; Cao, H.; Chen, R.; Wang, B.; Huang, Y.-W. Expression of the Human or Porcine C-Type Lectins DC-SIGN/L-SIGN Confers Susceptibility to Porcine Epidemic Diarrhea Virus Entry and Infection in Otherwise Refractory Cell Lines. Microb. Pathog. 2021, 157, 104956. [Google Scholar] [CrossRef] [PubMed]
  26. Deng, F.; Ye, G.; Liu, Q.; Navid, M.T.; Zhong, X.; Li, Y.; Wan, C.; Xiao, S.; He, Q.; Fu, Z.F.; et al. Identification and Comparison of Receptor Binding Characteristics of the Spike Protein of Two Porcine Epidemic Diarrhea Virus Strains. Viruses 2016, 8, 55. [Google Scholar] [CrossRef] [Green Version]
  27. Sun, X.-L. The Role of Cell Surface Sialic Acids for SARS-CoV-2 Infection. Glycobiology 2021. [Google Scholar] [CrossRef]
  28. Huan, C.; Wang, Y.; Ni, B.; Wang, R.; Huang, L.; Ren, X.; Tong, G.; Ding, C.; Fan, H.; Mao, X. Porcine Epidemic Diarrhea Virus Uses Cell-Surface Heparan Sulfate as an Attachment Factor. Arch. Virol. 2015, 160, 1621–1628. [Google Scholar] [CrossRef] [PubMed]
  29. Lin, H.-X.; Feng, Y.; Wong, G.; Wang, L.; Li, B.; Zhao, X.; Li, Y.; Smaill, F.; Zhang, C. 2008 Identification of Residues in the Receptor-Binding Domain (RBD) of the Spike Protein of Human Coronavirus NL63 That Are Critical for the RBD–ACE2 Receptor Interaction. J. Gen. Virol. 2008, 89, 1015–1024. [Google Scholar] [CrossRef]
  30. Milewska, A.; Zarebski, M.; Nowak, P.; Stozek, K.; Potempa, J.; Pyrc, K. Human Coronavirus NL63 Utilizes Heparan Sulfate Proteoglycans for Attachment to Target Cells. J. Virol. 2014, 88, 13221–13230. [Google Scholar] [CrossRef] [Green Version]
  31. Hofmann, H.; Simmons, G.; Rennekamp, A.J.; Chaipan, C.; Gramberg, T.; Heck, E.; Geier, M.; Wegele, A.; Marzi, A.; Bates, P.; et al. Highly Conserved Regions within the Spike Proteins of Human Coronaviruses 229E and NL63 Determine Recognition of Their Respective Cellular Receptors. J. Virol. 2006, 80, 8639–8652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Bonavia, A.; Zelus, B.D.; Wentworth, D.E.; Talbot, P.J.; Holmes, K.V. Identification of a Receptor-Binding Domain of the Spike Glycoprotein of Human Coronavirus HCoV-229E. J. Virol. 2003, 77, 2530–2538. [Google Scholar] [CrossRef] [Green Version]
  33. Breslin, J.J.; Mørk, I.; Smith, M.K.; Vogel, L.K.; Hemmila, E.M.; Bonavia, A.; Talbot, P.J.; Sjöström, H.; Norén, O.; Holmes, K.V. Human Coronavirus 229E: Receptor Binding Domain and Neutralization by Soluble Receptor at 37 Degrees C. J. Virol. 2003, 77, 4435–4438. [Google Scholar] [CrossRef] [Green Version]
  34. Jeffers, S.A.; Hemmila, E.M.; Holmes, K.V. Human Coronavirus 229E Can Use CD209L (L-SIGN) to Enter Cells. Adv. Exp. Med. Biol. 2006, 581, 265–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. De Haan, C.; Haijema, B.; Schellen, P.; Wichgers Schreur, P.; Lintelo, E.; Vennema, H.; Rottier, P. Cleavage of Group 1 Coronavirus Spike Proteins: How Furin Cleavage Is Traded Off against Heparan Sulfate Binding upon Cell Culture Adaptation. J. Virol. 2008, 82, 6078–6083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Regan, A.D.; Whittaker, G.R. Utilization of DC-SIGN for Entry of Feline Coronaviruses into Host Cells. J. Virol. 2008, 82, 11992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Regan, A.D.; Millet, J.K.; Tse, L.P.V.; Chillag, Z.; Rinaldi, V.D.; Licitra, B.N.; Dubovi, E.J.; Town, C.D.; Whittaker, G.R. Characterization of a Recombinant Canine Coronavirus with a Distinct Receptor-Binding (S1) Domain. Virology 2012, 430, 90–99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Kubo, H.; Yamada, Y.K.; Taguchi, F. Localization of Neutralizing Epitopes and the Receptor-Binding Site within the Amino-Terminal 330 Amino Acids of the Murine Coronavirus Spike Protein. J. Virol. 1994, 68, 5403–5410. [Google Scholar] [CrossRef] [Green Version]
  39. Szczepanski, A.; Owczarek, K.; Bzowska, M.; Gula, K.; Drebot, I.; Ochman, M.; Maksym, B.; Rajfur, Z.; Mitchell, J.A.; Pyrc, K. Canine Respiratory Coronavirus, Bovine Coronavirus, and Human Coronavirus OC43: Receptors and Attachment Factors. Viruses 2019, 11, 328. [Google Scholar] [CrossRef] [Green Version]
  40. Borucki, M.K.; Allen, J.E.; Chen-Harris, H.; Zemla, A.; Vanier, G.; Mabery, S.; Torres, C.; Hullinger, P.; Slezak, T. The Role of Viral Population Diversity in Adaptation of Bovine Coronavirus to New Host Environments. PLoS ONE 2013, 8, e52752. [Google Scholar] [CrossRef]
  41. Schultze, B.; Herrler, G. Bovine Coronavirus Uses N-Acetyl-9-O-Acetylneuraminic Acid as a Receptor Determinant to Initiate the Infection of Cultured Cells. J. Gen. Virol. 1992, 73 Pt 4, 901–906. [Google Scholar] [CrossRef]
  42. Lau, S.K.P.; Lee, P.; Tsang, A.K.L.; Yip, C.C.Y.; Tse, H.; Lee, R.A.; So, L.-Y.; Lau, Y.-L.; Chan, K.-H.; Woo, P.C.Y.; et al. Molecular Epidemiology of Human Coronavirus OC43 Reveals Evolution of Different Genotypes over Time and Recent Emergence of a Novel Genotype Due to Natural Recombination. J. Virol. 2011, 85, 11325–11337. [Google Scholar] [CrossRef] [Green Version]
  43. Vlasak, R.; Luytjes, W.; Spaan, W.; Palese, P. Human and Bovine Coronaviruses Recognize Sialic Acid-Containing Receptors Similar to Those of Influenza C Viruses. Proc. Natl. Acad. Sci. USA 1988, 85, 4526–4529. [Google Scholar] [CrossRef] [Green Version]
  44. Mou, H.; Raj, V.S.; van Kuppeveld, F.J.M.; Rottier, P.J.M.; Haagmans, B.L.; Bosch, B.J. The Receptor Binding Domain of the New Middle East Respiratory Syndrome Coronavirus Maps to a 231-Residue Region in the Spike Protein That Efficiently Elicits Neutralizing Antibodies. J. Virol. 2013, 87, 9379–9383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yan, L.; Song, Y.; Xia, K.; He, P.; Zhang, F.; Chen, S.; Pouliot, R.; Weiss, D.J.; Tandon, R.; Bates, J.T.; et al. Heparan Sulfates from Bat and Human Lung and Their Binding to the Spike Protein of SARS-CoV-2 Virus. Carbohydr. Polym. 2021, 260, 117797. [Google Scholar] [CrossRef] [PubMed]
  46. Wong, S.K.; Li, W.; Moore, M.J.; Choe, H.; Farzan, M. A 193-Amino Acid Fragment of the SARS Coronavirus S Protein Efficiently Binds Angiotensin-Converting Enzyme 2. J. Biol. Chem. 2004, 279, 3197–3201. [Google Scholar] [CrossRef] [Green Version]
  47. Zhang, Q.; Chen, C.Z.; Swaroop, M.; Xu, M.; Wang, L.; Lee, J.; Wang, A.Q.; Pradhan, M.; Hagen, N.; Chen, L.; et al. Heparan Sulfate Assists SARS-CoV-2 in Cell Entry and Can Be Targeted by Approved Drugs in Vitro. Cell Discov. 2020, 6, 1–14. [Google Scholar] [CrossRef] [PubMed]
  48. Marzi, A.; Gramberg, T.; Simmons, G.; Möller, P.; Rennekamp, A.; Krumbiegel, M.; Geier, M.; Eisemann, J.; Turza, N.; Saunier, B.; et al. DC-SIGN and DC-SIGNR Interact with the Glycoprotein of Marburg Virus and the S Protein of Severe Acute Respiratory Syndrome Coronavirus. J. Virol. 2004, 78, 12090–12095. [Google Scholar] [CrossRef] [Green Version]
  49. Yang, Z.-Y.; Huang, Y.; Ganesh, L.; Leung, K.; Kong, W.-P.; Schwartz, O.; Subbarao, K.; Nabel, G.J. PH-Dependent Entry of Severe Acute Respiratory Syndrome Coronavirus Is Mediated by the Spike Glycoprotein and Enhanced by Dendritic Cell Transfer through DC-SIGN. J. Virol. 2004, 78, 5642–5650. [Google Scholar] [CrossRef] [Green Version]
  50. Tai, W.; He, L.; Zhang, X.; Pu, J.; Voronin, D.; Jiang, S.; Zhou, Y.; Du, L. Characterization of the Receptor-Binding Domain (RBD) of 2019 Novel Coronavirus: Implication for Development of RBD Protein as a Viral Attachment Inhibitor and Vaccine. Cell. Mol. Immunol. 2020, 17, 613–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Thépaut, M.; Luczkowiak, J.; Vivès, C.; Labiod, N.; Bally, I.; Lasala, F.; Grimoire, Y.; Fenel, D.; Sattin, S.; Thielens, N.; et al. DC/L-SIGN Recognition of Spike Glycoprotein Promotes SARS-CoV-2 Trans-Infection and Can Be Inhibited by a Glycomimetic Antagonist. PLoS Pathog. 2021, 17, e1009576. [Google Scholar] [CrossRef]
  52. Huang, X.; Dong, W.; Milewska, A.; Golda, A.; Qi, Y.; Zhu, Q.K.; Marasco, W.A.; Baric, R.S.; Sims, A.C.; Pyrc, K.; et al. Human Coronavirus HKU1 Spike Protein Uses O-Acetylated Sialic Acid as an Attachment Receptor Determinant and Employs Hemagglutinin-Esterase Protein as a Receptor-Destroying Enzyme. J. Virol. 2015, 89, 7202–7213. [Google Scholar] [CrossRef] [Green Version]
  53. Promkuntod, N.; van Eijndhoven, R.E.W.; de Vrieze, G.; Gröne, A.; Verheije, M.H. Mapping of the Receptor-Binding Domain and Amino Acids Critical for Attachment in the Spike Protein of Avian Coronavirus Infectious Bronchitis Virus. Virology 2014, 448, 26–32. [Google Scholar] [CrossRef] [Green Version]
  54. Winter, C.; Schwegmann-Weßels, C.; Cavanagh, D.; Neumann, U.; Herrler, G. Sialic Acid Is a Receptor Determinant for Infection of Cells by Avian Infectious Bronchitis Virus. J. Gen. Virol. 2006, 87, 1209–1216. [Google Scholar] [CrossRef]
  55. Madu, I.G.; Chu, V.C.; Lee, H.; Regan, A.D.; Bauman, B.E.; Whittaker, G.R. Heparan Sulfate Is a Selective Attachment Factor for the Avian Coronavirus Infectious Bronchitis Virus Beaudette. Avian. Dis. 2007, 51, 45–51. [Google Scholar] [CrossRef] [Green Version]
  56. Zhang, Y.; Buckles, E.; Whittaker, G.R. Expression of the C-Type Lectins DC-SIGN or L-SIGN Alters Host Cell Susceptibility for the Avian Coronavirus, Infectious Bronchitis Virus. Vet. Microbiol. 2012, 157, 285–293. [Google Scholar] [CrossRef]
  57. Yuan, P.; Yang, Z.; Song, H.; Wang, K.; Yang, Y.; Xie, L.; Huang, S.; Liu, J.; Ran, L.; Song, Z. Three Main Inducers of Alphacoronavirus Infection of Enterocytes: Sialic Acid, Proteases, and Low PH. INT 2018, 61, 53–63. [Google Scholar] [CrossRef]
  58. Reguera, J.; Santiago, C.; Mudgal, G.; Ordoño, D.; Enjuanes, L.; Casasnovas, J.M. Structural Bases of Coronavirus Attachment to Host Aminopeptidase N and Its Inhibition by Neutralizing Antibodies. PLoS Pathog. 2012, 8, e1002859. [Google Scholar] [CrossRef] [Green Version]
  59. Tresnan, D.B.; Levis, R.; Holmes, K.V. Feline Aminopeptidase N Serves as a Receptor for Feline, Canine, Porcine, and Human Coronaviruses in Serogroup I. J. Virol. 1996, 70, 8669–8674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Luan, Y.; Xu, W. The Structure and Main Functions of Aminopeptidase N. Curr. Med. Chem. 2007, 14, 639–647. [Google Scholar] [CrossRef] [PubMed]
  61. Delmas, B.; Gelfi, J.; Kut, E.; Sjöström, H.; Noren, O.; Laude, H. Determinants Essential for the Transmissible Gastroenteritis Virus-Receptor Interaction Reside within a Domain of Aminopeptidase-N That Is Distinct from the Enzymatic Site. J. Virol. 1994, 68, 5216–5224. [Google Scholar] [CrossRef] [Green Version]
  62. Benbacer, L.; Kut, E.; Besnardeau, L.; Laude, H.; Delmas, B. Interspecies Aminopeptidase-N Chimeras Reveal Species-Specific Receptor Recognition by Canine Coronavirus, Feline Infectious Peritonitis Virus, and Transmissible Gastroenteritis Virus. J. Virol. 1997, 71, 734–737. [Google Scholar] [CrossRef] [Green Version]
  63. Hegyi, A.; Kolb, A.F. Characterization of Determinants Involved in the Feline Infectious Peritonitis Virus Receptor Function of Feline Aminopeptidase N. J. Gen. Virol. 1998, 79 Pt 6, 1387–1391. [Google Scholar] [CrossRef] [Green Version]
  64. Tusell, S.M.; Schittone, S.A.; Holmes, K.V. Mutational Analysis of Aminopeptidase N, a Receptor for Several Group 1 Coronaviruses, Identifies Key Determinants of Viral Host Range. J. Virol. 2007, 81, 1261–1273. [Google Scholar] [CrossRef] [Green Version]
  65. Hofmann, H.; Pyrc, K.; van der Hoek, L.; Geier, M.; Berkhout, B.; Pöhlmann, S. Human Coronavirus NL63 Employs the Severe Acute Respiratory Syndrome Coronavirus Receptor for Cellular Entry. Proc. Natl. Acad. Sci. USA 2005, 102, 7988–7993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Li, W.; Moore, M.J.; Vasilieva, N.; Sui, J.; Wong, S.K.; Berne, M.A.; Somasundaran, M.; Sullivan, J.L.; Luzuriaga, K.; Greenough, T.C.; et al. Angiotensin-Converting Enzyme 2 Is a Functional Receptor for the SARS Coronavirus. Nature 2003, 426, 450–454. [Google Scholar] [CrossRef] [Green Version]
  67. Zhou, P.; Yang, X.-L.; Wang, X.-G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.-R.; Zhu, Y.; Li, B.; Huang, C.-L.; et al. A Pneumonia Outbreak Associated with a New Coronavirus of Probable Bat Origin. Nature 2020, 579, 270–273. [Google Scholar] [CrossRef] [Green Version]
  68. Li, W.; Zhang, C.; Sui, J.; Kuhn, J.H.; Moore, M.J.; Luo, S.; Wong, S.-K.; Huang, I.-C.; Xu, K.; Vasilieva, N.; et al. Receptor and Viral Determinants of SARS-Coronavirus Adaptation to Human ACE2. EMBO J. 2005, 24, 1634–1643. [Google Scholar] [CrossRef] [Green Version]
  69. Chen, Y.; Liu, L.; Wei, Q.; Zhu, H.; Jiang, H.; Tu, X.; Qin, C.; Chen, Z. Rhesus Angiotensin Converting Enzyme 2 Supports Entry of Severe Acute Respiratory Syndrome Coronavirus in Chinese Macaques. Virology 2008, 381, 89–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Xu, L.; Zhang, Y.; Liu, Y.; Chen, Z.; Deng, H.; Ma, Z.; Wang, H.; Hu, Z.; Deng, F. Angiotensin-Converting Enzyme 2 (ACE2) from Raccoon Dog Can Serve as an Efficient Receptor for the Spike Protein of Severe Acute Respiratory Syndrome Coronavirus. J. Gen. Virol. 2009, 90, 2695–2703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Zamoto, A.; Taguchi, F.; Fukushi, S.; Morikawa, S.; Yamada, Y.K. Identification of Ferret ACE2 and Its Receptor Function for SARS-Coronavirus. Adv. Exp. Med. Biol. 2006, 581, 519–522. [Google Scholar] [CrossRef] [Green Version]
  72. Heller, L.K.; Gillim-Ross, L.; Olivieri, E.R.; Wentworth, D.E. Mustela Vison ACE2 Functions as a Receptor for Sars-Coronavirus. Nidoviruses 2006, 581, 507–510. [Google Scholar] [CrossRef] [Green Version]
  73. Zhao, X.; Chen, D.; Szabla, R.; Zheng, M.; Li, G.; Du, P.; Zheng, S.; Li, X.; Song, C.; Li, R.; et al. Broad and Differential Animal Angiotensin-Converting Enzyme 2 Receptor Usage by SARS-CoV-2. J. Virol. 2020, 94, e00940-20. [Google Scholar] [CrossRef]
  74. Ge, X.-Y.; Li, J.-L.; Yang, X.-L.; Chmura, A.A.; Zhu, G.; Epstein, J.H.; Mazet, J.K.; Hu, B.; Zhang, W.; Peng, C.; et al. Isolation and Characterization of a Bat SARS-like Coronavirus That Uses the ACE2 Receptor. Nature 2013, 503, 535–538. [Google Scholar] [CrossRef] [PubMed]
  75. Zheng, M.; Zhao, X.; Zheng, S.; Chen, D.; Du, P.; Li, X.; Jiang, D.; Guo, J.-T.; Zeng, H.; Lin, H. Bat SARS-Like WIV1 Coronavirus Uses the ACE2 of Multiple Animal Species as Receptor and Evades IFITM3 Restriction via TMPRSS2 Activation of Membrane Fusion. Emerg. Microbes Infect. 2020, 9, 1567–1579. [Google Scholar] [CrossRef]
  76. Raj, V.S.; Mou, H.; Smits, S.L.; Dekkers, D.H.W.; Müller, M.A.; Dijkman, R.; Muth, D.; Demmers, J.A.A.; Zaki, A.; Fouchier, R.A.M.; et al. Dipeptidyl Peptidase 4 Is a Functional Receptor for the Emerging Human Coronavirus-EMC. Nature 2013, 495, 251–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Babcock, G.J.; Esshaki, D.J.; Thomas, W.D.; Ambrosino, D.M. Amino Acids 270 to 510 of the Severe Acute Respiratory Syndrome Coronavirus Spike Protein Are Required for Interaction with Receptor. J. Virol. 2004, 78, 4552–4560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Li, F. Receptor Recognition Mechanisms of Coronaviruses: A Decade of Structural Studies. J. Virol. 2014, 89, 1954–1964. [Google Scholar] [CrossRef] [Green Version]
  79. Peng, G.; Xu, L.; Lin, Y.-L.; Chen, L.; Pasquarella, J.R.; Holmes, K.V.; Li, F. Crystal Structure of Bovine Coronavirus Spike Protein Lectin Domain. J. Biol. Chem. 2012, 287, 41931–41938. [Google Scholar] [CrossRef] [Green Version]
  80. Suzuki, H.; Taguchi, F. Analysis of the Receptor-Binding Site of Murine Coronavirus Spike Protein. J. Virol. 1996, 70, 2632–2636. [Google Scholar] [CrossRef] [Green Version]
  81. Thackray, L.B.; Turner, B.C.; Holmes, K.V. Substitutions of Conserved Amino Acids in the Receptor-Binding Domain of the Spike Glycoprotein Affect Utilization of Murine CEACAM1a by the Murine Coronavirus MHV-A59. Virology 2005, 334, 98–110. [Google Scholar] [CrossRef] [Green Version]
  82. Mi, D.; Ou, X.; Li, P.; Peng, G.; Liu, Y.; Guo, R.; Mu, Z.; Li, F.; Holmes, K.; Qian, Z. Glycine 29 Is Critical for Conformational Changes of the Spike Glycoprotein of Mouse Hepatitis Virus A59 Triggered by Either Receptor Binding or High PH. J. Virol. 2019, 93, e01046-19. [Google Scholar] [CrossRef] [Green Version]
  83. Li, F.; Li, W.; Farzan, M.; Harrison, S.C. Structure of SARS Coronavirus Spike Receptor-Binding Domain Complexed with Receptor. Science 2005, 309, 1864–1868. [Google Scholar] [CrossRef] [PubMed]
  84. Hulswit, R.J.G.; Lang, Y.; Bakkers, M.J.G.; Li, W.; Li, Z.; Schouten, A.; Ophorst, B.; Kuppeveld, F.J.M.; van Boons, G.-J.; Bosch, B.-J.; et al. Human Coronaviruses OC43 and HKU1 Bind to 9-O-Acetylated Sialic Acids via a Conserved Receptor-Binding Site in Spike Protein Domain A. Proc. Natl. Acad. Sci. USA 2019, 116, 2681–2690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Tortorici, M.A.; Walls, A.C.; Lang, Y.; Wang, C.; Li, Z.; Koerhuis, D.; Boons, G.-J.; Bosch, B.-J.; Rey, F.A.; de Groot, R.J.; et al. Structural Basis for Human Coronavirus Attachment to Sialic Acid Receptors. Nat. Struct. Mol. Biol. 2019, 26, 481–489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Sánchez, C.M.; Gebauer, F.; Suñé, C.; Mendez, A.; Dopazo, J.; Enjuanes, L. Genetic Evolution and Tropism of Transmissible Gastroenteritis Coronaviruses. Virology 1992, 190, 92–105. [Google Scholar] [CrossRef]
  87. Krempl, C.; Ballesteros, M.L.; Zimmer, G.; Enjuanes, L.; Klenk, H.D.; Herrler, G. Characterization of the Sialic Acid Binding Activity of Transmissible Gastroenteritis Coronavirus by Analysis of Haemagglutination-Deficient Mutants. J. Gen. Virol. 2000, 81, 489–496. [Google Scholar] [CrossRef] [PubMed]
  88. Schwegmann-Weßels, C.; Zimmer, G.; Laude, H.; Enjuanes, L.; Herrler, G. Binding of Transmissible Gastroenteritis Coronavirus to Cell Surface Sialoglycoproteins. J. Virol. 2002, 76, 6037–6043. [Google Scholar] [CrossRef] [Green Version]
  89. Schwegmann-Wessels, C.; Zimmer, G.; Schröder, B.; Breves, G.; Herrler, G. Binding of Transmissible Gastroenteritis Coronavirus to Brush Border Membrane Sialoglycoproteins. J. Virol. 2003, 77, 11846–11848. [Google Scholar] [CrossRef] [Green Version]
  90. Künkel, F.; Herrler, G. Structural and Functional Analysis of the Surface Protein of Human Coronavirus OC43. Virology 1993, 195, 195–202. [Google Scholar] [CrossRef] [PubMed]
  91. Collins, A.R. Human Coronavirus OC43 Interacts with Major Histocompatibility Complex Class I Molecules at the Cell Surface to Establish Infection. Immunol. Investig. 1994, 23, 313–321. [Google Scholar] [CrossRef] [Green Version]
  92. Winter, C.; Herrler, G.; Neumann, U. Infection of the Tracheal Epithelium by Infectious Bronchitis Virus Is Sialic Acid Dependent. Microbes Infect. 2008, 10, 367. [Google Scholar] [CrossRef]
  93. Li, W.; Hulswit, R.J.G.; Widjaja, I.; Raj, V.S.; McBride, R.; Peng, W.; Widagdo, W.; Tortorici, M.A.; van Dieren, B.; Lang, Y.; et al. Identification of Sialic Acid-Binding Function for the Middle East Respiratory Syndrome Coronavirus Spike Glycoprotein. Proc. Natl. Acad. Sci. USA 2017, 114, E8508–E8517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Liu, J.; Thorp, S.C. Cell Surface Heparan Sulfate and Its Roles in Assisting Viral Infections. Med. Res. Rev. 2002, 22, 1–25. [Google Scholar] [CrossRef]
  95. De Haan, C.A.M.; Li, Z.; Lintelo, E.; te Bosch, B.J.; Haijema, B.J.; Rottier, P.J.M. Murine Coronavirus with an Extended Host Range Uses Heparan Sulfate as an Entry Receptor. J. Virol. 2005, 79, 14451. [Google Scholar] [CrossRef] [Green Version]
  96. Schickli, J.H.; Zelus, B.D.; Wentworth, D.E.; Sawicki, S.G.; Holmes, K.V. The Murine Coronavirus Mouse Hepatitis Virus Strain A59 from Persistently Infected Murine Cells Exhibits an Extended Host Range. J. Virol. 1997, 71. [Google Scholar] [CrossRef] [Green Version]
  97. Watanabe, R.; Sawicki, S.G.; Taguchi, F. Heparan Sulfate Is a Binding Molecule but Not a Receptor for CEACAM1-Independent Infection of Murine Coronavirus. Virology 2007, 366, 16–22. [Google Scholar] [CrossRef]
  98. Baribaud, F.; Pöhlmann, S.; Doms, R.W. The Role of DC-SIGN and DC-SIGNR in HIV and SIV Attachment, Infection, and Transmission. Virology 2001, 286, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Jain, P.; Manuel, S.L.; Khan, Z.K.; Ahuja, J.; Quann, K.; Wigdahl, B. DC-SIGN Mediates Cell-Free Infection and Transmission of Human T-Cell Lymphotropic Virus Type 1 by Dendritic Cells. J. Virol. 2009, 83, 10908–10921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Lozach, P.-Y.; Burleigh, L.; Staropoli, I.; Amara, A. The C Type Lectins DC-SIGN and L-SIGN. Glycovirology Protoc. 2007, 379, 51–68. [Google Scholar] [CrossRef]
  101. Schneider-Schaulies, J.; Schneider-Schaulies, S. Receptor Interactions, Tropism, and Mechanisms Involved in Morbillivirus-Induced Immunomodulation. Adv. Virus Res. 2008, 71, 173–205. [Google Scholar] [CrossRef] [PubMed]
  102. Cambi, A.; Koopman, M.; Figdor, C.G. How C-Type Lectins Detect Pathogens. Cell. Microbiol. 2005, 7, 481–488. [Google Scholar] [CrossRef]
  103. Geijtenbeek, T.B.H.; Gringhuis, S.I. Signalling through C-Type Lectin Receptors: Shaping Immune Responses. Nat. Rev. Immunol. 2009, 9, 465–479. [Google Scholar] [CrossRef]
  104. Cambi, A.; de Lange, F.; van Maarseveen, N.M.; Nijhuis, M.; Joosten, B.; van Dijk, E.M.H.P.; de Bakker, B.I.; Fransen, J.A.M.; Bovee-Geurts, P.H.M.; van Leeuwen, F.N.; et al. Microdomains of the C-Type Lectin DC-SIGN Are Portals for Virus Entry into Dendritic Cells. J. Cell Biol. 2004, 164, 145–155. [Google Scholar] [CrossRef]
  105. Han, D.P.; Lohani, M.; Cho, M.W. Specific Asparagine-Linked Glycosylation Sites Are Critical for DC-SIGN- and L-SIGN-Mediated Severe Acute Respiratory Syndrome Coronavirus Entry. J. Virol. 2007, 81, 12029–12039. [Google Scholar] [CrossRef] [Green Version]
  106. Jeffers, S.A.; Tusell, S.M.; Gillim-Ross, L.; Hemmila, E.M.; Achenbach, J.E.; Babcock, G.J.; Thomas, W.D.; Thackray, L.B.; Young, M.D.; Mason, R.J.; et al. CD209L (L-SIGN) Is a Receptor for Severe Acute Respiratory Syndrome Coronavirus. Proc. Natl. Acad. Sci. USA 2004, 101, 15748–15753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Gramberg, T.; Hofmann, H.; Möller, P.; Lalor, P.F.; Marzi, A.; Geier, M.; Krumbiegel, M.; Winkler, T.; Kirchhoff, F.; Adams, D.H.; et al. LSECtin Interacts with Filovirus Glycoproteins and the Spike Protein of SARS Coronavirus. Virology 2005, 340, 224–236. [Google Scholar] [CrossRef] [Green Version]
  108. Guillén, J.; Pérez-Berná, A.J.; Moreno, M.R.; Villalaín, J. Identification of the Membrane-Active Regions of the Severe Acute Respiratory Syndrome Coronavirus Spike Membrane Glycoprotein Using a 16/18-Mer Peptide Scan: Implications for the Viral Fusion Mechanism. J. Virol. 2005, 79, 1743–1752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Bosch, B.J.; van der Zee, R.; de Haan, C.A.M.; Rottier, P.J.M. The Coronavirus Spike Protein Is a Class I Virus Fusion Protein: Structural and Functional Characterization of the Fusion Core Complex. J. Virol. 2003, 77, 8801–8811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Luo, Z.; Weiss, S.R. Roles in Cell-to-Cell Fusion of Two Conserved Hydrophobic Regions in the Murine Coronavirus Spike Protein. Virology 1998, 244, 483–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Petit, C.M.; Melancon, J.M.; Chouljenko, V.N.; Colgrove, R.; Farzan, M.; Knipe, D.M.; Kousoulas, K.G. Genetic Analysis of the SARS-Coronavirus Spike Glycoprotein Functional Domains Involved in Cell-Surface Expression and Cell-to-Cell Fusion. Virology 2005, 341, 215–230. [Google Scholar] [CrossRef] [Green Version]
  112. De Groot, R.J.; Luytjes, W.; Horzinek, M.C.; van der Zeijst, B.A.; Spaan, W.J.; Lenstra, J.A. Evidence for a Coiled-Coil Structure in the Spike Proteins of Coronaviruses. J. Mol. Biol. 1987, 196, 963–966. [Google Scholar] [CrossRef]
  113. Tripet, B.; Howard, M.W.; Jobling, M.; Holmes, R.K.; Holmes, K.V.; Hodges, R.S. Structural Characterization of the SARS-Coronavirus Spike S Fusion Protein Core. J. Biol. Chem. 2004, 279, 20836–20849. [Google Scholar] [CrossRef] [Green Version]
  114. Xu, Y.; Liu, Y.; Lou, Z.; Qin, L.; Li, X.; Bai, Z.; Pang, H.; Tien, P.; Gao, G.F.; Rao, Z. Structural Basis for Coronavirus-Mediated Membrane Fusion: Crystal structure of mouse hepatitis virus spike protein fusion core. J. Biol. Chem. 2004, 279, 30514–30522. [Google Scholar] [CrossRef] [Green Version]
  115. Zheng, Q.; Deng, Y.; Liu, J.; van der Hoek, L.; Berkhout, B.; Lu, M. Core Structure of S2 from the Human Coronavirus NL63 Spike Glycoprotein. Biochemistry 2006, 45, 15205–15215. [Google Scholar] [CrossRef] [PubMed]
  116. Eckert, D.M.; Kim, P.S. Mechanisms of Viral Membrane Fusion and Its Inhibition. Annu. Rev. Biochem. 2001, 70, 777–810. [Google Scholar] [CrossRef] [Green Version]
  117. Epand, R.M. Fusion Peptides and the Mechanism of Viral Fusion. Biochim. Biophys. Acta 2003, 1614, 116–121. [Google Scholar] [CrossRef] [Green Version]
  118. Liu, S.; Xiao, G.; Chen, Y.; He, Y.; Niu, J.; Escalante, C.R.; Xiong, H.; Farmar, J.; Debnath, A.K.; Tien, P.; et al. Interaction between Heptad Repeat 1 and 2 Regions in Spike Protein of SARS-Associated Coronavirus: Implications for Virus Fusogenic Mechanism and Identification of Fusion Inhibitors. Lancet 2004, 363, 938–947. [Google Scholar] [CrossRef] [Green Version]
  119. Sainz, B.; Mossel, E.C.; Gallaher, W.R.; Wimley, W.C.; Peters, C.J.; Wilson, R.B.; Garry, R.F. Inhibition of Severe Acute Respiratory Syndrome-Associated Coronavirus (SARS-CoV) Infectivity by Peptides Analogous to the Viral Spike Protein. Virus Res. 2006, 120, 146–155. [Google Scholar] [CrossRef] [PubMed]
  120. Tripet, B.; Kao, D.; Jeffers, S.; Holmes, K.; Hodges, R. Template-Based Coiled-Coil Antigens Elicit Neutralizing Antibodies to the SARS-Coronavirus. J. Struct. Biol. 2006, 155, 176–194. [Google Scholar] [CrossRef] [PubMed]
  121. Howard, M.W.; Travanty, E.A.; Jeffers, S.A.; Smith, M.K.; Wennier, S.T.; Thackray, L.B.; Holmes, K.V. Aromatic Amino Acids in the Juxtamembrane Domain of Severe Acute Respiratory Syndrome Coronavirus Spike Glycoprotein Are Important for Receptor-Dependent Virus Entry and Cell-Cell Fusion. J. Virol. 2008, 82, 2883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Bos, E.C.; Heijnen, L.; Luytjes, W.; Spaan, W.J. Mutational Analysis of the Murine Coronavirus Spike Protein: Effect on Cell-to-Cell Fusion. Virology 1995, 214, 453–463. [Google Scholar] [CrossRef] [Green Version]
  123. Broer, R.; Boson, B.; Spaan, W.; Cosset, F.-L.; Corver, J. Important Role for the Transmembrane Domain of Severe Acute Respiratory Syndrome Coronavirus Spike Protein during Entry. J. Virol. 2006, 80, 1302–1310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Chang, K.W.; Sheng, Y.; Gombold, J.L. Coronavirus-Induced Membrane Fusion Requires the Cysteine-Rich Domain in the Spike Protein. Virology 2000, 269, 212–224. [Google Scholar] [CrossRef] [PubMed]
  125. Hogue, B.G.; Machamer, C.E. Coronavirus Structural Proteins and Virus Assembly. In Nidoviruses; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2007; pp. 179–200. [Google Scholar]
  126. Youn, S.; Collisson, E.W.; Machamer, C.E. Contribution of Trafficking Signals in the Cytoplasmic Tail of the Infectious Bronchitis Virus Spike Protein to Virus Infection. J. Virol. 2005, 79, 13209–13217. [Google Scholar] [CrossRef] [Green Version]
  127. Bonifacino, J.S.; Jackson, C.L. Endosome-Specific Localization and Function of the ARF Activator GNOM. Cell 2003, 112, 141–142. [Google Scholar] [CrossRef] [Green Version]
  128. Schwegmann-Wessels, C.; Glende, J.; Ren, X.; Qu, X.; Deng, H.; Enjuanes, L.; Herrler, G. Comparison of Vesicular Stomatitis Virus Pseudotyped with the S Proteins from a Porcine and a Human Coronavirus. J. Gen. Virol. 2009, 90, 1724–1729. [Google Scholar] [CrossRef]
  129. Lin, H.-X.; Feng, Y.; Tu, X.; Zhao, X.; Hsieh, C.-H.; Griffin, L.; Junop, M.; Zhang, C. Characterization of the Spike Protein of Human Coronavirus NL63 in Receptor Binding and Pseudotype Virus Entry. Virus Res. 2011, 160, 283–293. [Google Scholar] [CrossRef]
  130. Giroglou, T.; Cinatl, J.; Rabenau, H.; Drosten, C.; Schwalbe, H.; Doerr, H.W.; von Laer, D. Retroviral Vectors Pseudotyped with Severe Acute Respiratory Syndrome Coronavirus S Protein. J. Virol. 2004, 78, 9007–9015. [Google Scholar] [CrossRef] [Green Version]
  131. Petit, C.M.; Chouljenko, V.N.; Iyer, A.; Colgrove, R.; Farzan, M.; Knipe, D.M.; Kousoulas, K.G. Palmitoylation of the Cysteine-Rich Endodomain of the SARS-Coronavirus Spike Glycoprotein Is Important for Spike-Mediated Cell Fusion. Virology 2007, 360, 264–274. [Google Scholar] [CrossRef] [Green Version]
  132. Thorp, E.B.; Boscarino, J.A.; Logan, H.L.; Goletz, J.T.; Gallagher, T.M. Palmitoylations on Murine Coronavirus Spike Proteins Are Essential for Virion Assembly and Infectivity. J. Virol. 2006, 80, 1280–1289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. De Haan, C.A.M.; Rottier, P.J.M. Molecular Interactions in the Assembly of Coronaviruses. Adv. Virus Res. 2005, 64, 165–230. [Google Scholar] [CrossRef]
  134. De Haan, C.A.M.; Stadler, K.; Godeke, G.-J.; Bosch, B.J.; Rottier, P.J.M. Cleavage Inhibition of the Murine Coronavirus Spike Protein by a Furin-like Enzyme Affects Cell-Cell but Not Virus-Cell Fusion. J. Virol. 2004, 78, 6048–6054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Stauber, R.; Pfleiderera, M.; Siddell, S. Proteolytic Cleavage of the Murine Coronavirus Surface Glycoprotein Is Not Required for Fusion Activity. J. Gen. Virol. 1993, 74 Pt 2, 183–191. [Google Scholar] [CrossRef] [PubMed]
  136. Taguchi, F. Fusion Formation by the Uncleaved Spike Protein of Murine Coronavirus JHMV Variant Cl-2. J. Virol. 1993, 67, 1195. [Google Scholar] [CrossRef] [Green Version]
  137. Wicht, O.; Burkard, C.; de Haan, C.A.M.; van Kuppeveld, F.J.M.; Rottier, P.J.M.; Bosch, B.J. Identification and Characterization of a Proteolytically Primed Form of the Murine Coronavirus Spike Proteins after Fusion with the Target Cell. J. Virol. 2014, 88, 4943–4952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Yamada, Y.; Liu, D.X. Proteolytic Activation of the Spike Protein at a Novel RRRR/S Motif Is Implicated in Furin-Dependent Entry, Syncytium Formation, and Infectivity of Coronavirus Infectious Bronchitis Virus in Cultured Cells. J. Virol. 2009, 83, 8744–8758. [Google Scholar] [CrossRef] [Green Version]
  139. Wicht, O.; Li, W.; Willems, L.; Meuleman, T.J.; Wubbolts, R.W.; van Kuppeveld, F.J.M.; Rottier, P.J.M.; Bosch, B.J. Proteolytic Activation of the Porcine Epidemic Diarrhea Coronavirus Spike Fusion Protein by Trypsin in Cell Culture. J. Virol. 2014, 88, 7952–7961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Park, J.-E.; Cruz, D.J.M.; Shin, H.-J. Clathrin- and Serine Proteases-Dependent Uptake of Porcine Epidemic Diarrhea Virus into Vero Cells. Virus Res. 2014, 191, 21–29. [Google Scholar] [CrossRef]
  141. Regan, A.D.; Shraybman, R.; Cohen, R.D.; Whittaker, G.R. Differential Role for Low PH and Cathepsin-Mediated Cleavage of the Viral Spike Protein during Entry of Serotype II Feline Coronaviruses. Vet. Microbiol. 2008, 132, 235–248. [Google Scholar] [CrossRef] [PubMed]
  142. Glowacka, I.; Bertram, S.; Müller, M.A.; Allen, P.; Soilleux, E.; Pfefferle, S.; Steffen, I.; Tsegaye, T.S.; He, Y.; Gnirss, K.; et al. Evidence That TMPRSS2 Activates the Severe Acute Respiratory Syndrome Coronavirus Spike Protein for Membrane Fusion and Reduces Viral Control by the Humoral Immune Response. J. Virol. 2011, 85, 4122–4134. [Google Scholar] [CrossRef] [Green Version]
  143. Shirato, K.; Kawase, M.; Matsuyama, S. Middle East Respiratory Syndrome Coronavirus Infection Mediated by the Transmembrane Serine Protease TMPRSS2. J. Virol. 2013, 87, 12552–12561. [Google Scholar] [CrossRef] [Green Version]
  144. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Krüger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.-H.; Nitsche, A.; et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell 2020, 181, 271–280. [Google Scholar] [CrossRef] [PubMed]
  145. Du, L.; Kao, R.Y.; Zhou, Y.; He, Y.; Zhao, G.; Wong, C.; Jiang, S.; Yuen, K.-Y.; Jin, D.-Y.; Zheng, B.-J. Cleavage of Spike Protein of SARS Coronavirus by Protease Factor Xa Is Associated with Viral Infectivity. Biochem. Biophys. Res. Commun. 2007, 359, 174–179. [Google Scholar] [CrossRef] [PubMed]
  146. Matsuyama, S.; Ujike, M.; Morikawa, S.; Tashiro, M.; Taguchi, F. Protease-Mediated Enhancement of Severe Acute Respiratory Syndrome Coronavirus Infection. Proc. Natl. Acad. Sci. USA 2005, 102, 12543–12547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Simmons, G.; Gosalia, D.N.; Rennekamp, A.J.; Reeves, J.D.; Diamond, S.L.; Bates, P. Inhibitors of Cathepsin L Prevent Severe Acute Respiratory Syndrome Coronavirus Entry. Proc. Natl. Acad. Sci. USA 2005, 102, 11876–11881. [Google Scholar] [CrossRef] [Green Version]
  148. Kawase, M.; Shirato, K.; Matsuyama, S.; Taguchi, F. Protease-Mediated Entry via the Endosome of Human Coronavirus 229E. J. Virol. 2009, 83, 712–721. [Google Scholar] [CrossRef] [Green Version]
  149. Watanabe, R.; Matsuyama, S.; Shirato, K.; Maejima, M.; Fukushi, S.; Morikawa, S.; Taguchi, F. Entry from the Cell Surface of Severe Acute Respiratory Syndrome Coronavirus with Cleaved S Protein as Revealed by Pseudotype Virus Bearing Cleaved S Protein. J. Virol. 2008, 82, 11985–11991. [Google Scholar] [CrossRef] [Green Version]
  150. Qiu, Z.; Hingley, S.; Simmons, G.; Yu, C.; Sarma, J.; Bates, P.; Weiss, S. Endosomal Proteolysis by Cathepsins Is Necessary for Murine Coronavirus Mouse Hepatitis Virus Type 2 Spike-Mediated Entry. J. Virol. 2006, 80, 5768–5776. [Google Scholar] [CrossRef] [Green Version]
  151. Bailey, C.C.; Zhong, G.; Huang, I.-C.; Farzan, M. IFITM-Family Proteins: The Cell’s First Line of Antiviral Defense. Annu. Rev. Virol. 2014, 1, 261–283. [Google Scholar] [CrossRef] [Green Version]
  152. Perreira, J.M.; Chin, C.R.; Feeley, E.M.; Brass, A.L. IFITMs Restrict the Replication of Multiple Pathogenic Viruses. J. Mol. Biol. 2013, 425, 4937–4955. [Google Scholar] [CrossRef]
  153. Zhao, X.; Li, J.; Winkler, C.A.; An, P.; Guo, J.-T. IFITM Genes, Variants, and Their Roles in the Control and Pathogenesis of Viral Infections. Front. Microbiol. 2019, 9, 3228. [Google Scholar] [CrossRef] [Green Version]
  154. Huang, I.-C.; Bailey, C.C.; Weyer, J.L.; Radoshitzky, S.R.; Becker, M.M.; Chiang, J.J.; Brass, A.L.; Ahmed, A.A.; Chi, X.; Dong, L.; et al. Distinct Patterns of IFITM-Mediated Restriction of Filoviruses, SARS Coronavirus, and Influenza A Virus. PLoS Pathog 2011, 7, e1001258. [Google Scholar] [CrossRef]
  155. Winstone, H.; Lista, M.J.; Reid, A.C.; Bouton, C.; Pickering, S.; Galao, R.P.; Kerridge, C.; Doores, K.J.; Swanson, C.M.; Neil, S.J.D. The Polybasic Cleavage Site in SARS-CoV-2 Spike Modulates Viral Sensitivity to Type I Interferon and IFITM2. J. Virol. 2021, 95, e02422-20. [Google Scholar] [CrossRef]
  156. Bertram, S.; Dijkman, R.; Habjan, M.; Heurich, A.; Gierer, S.; Glowacka, I.; Welsch, K.; Winkler, M.; Schneider, H.; Hofmann-Winkler, H.; et al. TMPRSS2 Activates the Human Coronavirus 229E for Cathepsin-Independent Host Cell Entry and Is Expressed in Viral Target Cells in the Respiratory Epithelium. J. Virol. 2013, 87, 6150–6160. [Google Scholar] [CrossRef] [Green Version]
  157. Wrensch, F.; Winkler, M.; Pöhlmann, S. IFITM Proteins Inhibit Entry Driven by the MERS-Coronavirus Spike Protein: Evidence for Cholesterol-Independent Mechanisms. Viruses 2014, 6, 3683–3698. [Google Scholar] [CrossRef] [Green Version]
  158. Shi, G.; Kenney, A.D.; Kudryashova, E.; Zani, A.; Zhang, L.; Lai, K.K.; Hall-Stoodley, L.; Robinson, R.T.; Kudryashov, D.S.; Compton, A.A.; et al. Opposing Activities of IFITM Proteins in SARS-CoV-2 Infection. EMBO J. 2021, 40, e106501. [Google Scholar] [CrossRef] [PubMed]
  159. Buchrieser, J.; Dufloo, J.; Hubert, M.; Monel, B.; Planas, D.; Rajah, M.M.; Planchais, C.; Porrot, F.; Guivel-Benhassine, F.; van der Werf, S.; et al. Syncytia Formation by SARS-CoV-2-Infected Cells. EMBO J. 2020, 39, e106267. [Google Scholar] [CrossRef]
  160. Zhao, X.; Zheng, S.; Chen, D.; Zheng, M.; Li, X.; Li, G.; Lin, H.; Chang, J.; Zeng, H.; Guo, J.-T. LY6E Restricts Entry of Human Coronaviruses, Including Currently Pandemic SARS-CoV-2. J. Virol. 2020, 94, e00562-20. [Google Scholar] [CrossRef] [PubMed]
  161. Zhao, X.; Sehgal, M.; Hou, Z.; Cheng, J.; Shu, S.; Wu, S.; Guo, F.; Le Marchand, S.J.; Lin, H.; Chang, J.; et al. Identification of Residues Controlling Restriction versus Enhancing Activities of IFITM Proteins on Entry of Human Coronaviruses. J. Virol. 2018, 92, e01535-17. [Google Scholar] [CrossRef] [Green Version]
  162. Pfaender, S.; Mar, K.B.; Michailidis, E.; Kratzel, A.; Boys, I.N.; V’kovski, P.; Fan, W.; Kelly, J.N.; Hirt, D.; Ebert, N.; et al. LY6E Impairs Coronavirus Fusion and Confers Immune Control of Viral Disease. Nat. Microbiol. 2020, 5, 1330–1339. [Google Scholar] [CrossRef]
  163. Bruchez, A.; Sha, K.; Johnson, J.; Chen, L.; Stefani, C.; McConnell, H.; Gaucherand, L.; Prins, R.; Matreyek, K.A.; Hume, A.J.; et al. MHC Class II Transactivator CIITA Induces Cell Resistance to Ebola Virus and SARS-like Coronaviruses. Science 2020, 370, 241–247. [Google Scholar] [CrossRef] [PubMed]
  164. Millet, J.K.; Kien, F.; Cheung, C.-Y.; Siu, Y.-L.; Chan, W.-L.; Li, H.; Leung, H.-L.; Jaume, M.; Bruzzone, R.; Peiris, J.S.M.; et al. Ezrin Interacts with the SARS Coronavirus Spike Protein and Restrains Infection at the Entry Stage. PLoS ONE 2012, 7, e49566. [Google Scholar] [CrossRef]
  165. Chen, D.; Hou, Z.; Jiang, D.; Zheng, M.; Li, G.; Zhang, Y.; Li, R.; Lin, H.; Chang, J.; Zeng, H.; et al. GILT Restricts the Cellular Entry Mediated by the Envelope Glycoproteins of SARS-CoV, Ebola Virus and Lassa Fever Virus. Emerg. Microbes Infect. 2019, 8, 1511–1523. [Google Scholar] [CrossRef] [Green Version]
  166. Zang, R.; Case, J.B.; Yutuc, E.; Ma, X.; Shen, S.; Gomez Castro, M.F.; Liu, Z.; Zeng, Q.; Zhao, H.; Son, J.; et al. Cholesterol 25-Hydroxylase Suppresses SARS-CoV-2 Replication by Blocking Membrane Fusion. Proc. Natl. Acad. Sci. USA 2020, 117, 32105–32113. [Google Scholar] [CrossRef] [PubMed]
  167. Wang, S.; Li, W.; Hui, H.; Tiwari, S.K.; Zhang, Q.; Croker, B.A.; Rawlings, S.; Smith, D.; Carlin, A.F.; Rana, T.M. Cholesterol 25-Hydroxylase Inhibits SARS-CoV-2 and Other Coronaviruses by Depleting Membrane Cholesterol. EMBO J. 2020, 39, e106057. [Google Scholar] [CrossRef] [PubMed]
  168. Schoggins, J. Interferon-Stimulated Genes: What Do They All Do? Annu. Rev. Virol. 2019, 6, 567–584. [Google Scholar] [CrossRef]
  169. Zhou, P.; Fan, H.; Lan, T.; Yang, X.-L.; Shi, W.-F.; Zhang, W.; Zhu, Y.; Zhang, Y.-W.; Xie, Q.-M.; Mani, S.; et al. Fatal Swine Acute Diarrhoea Syndrome Caused by an HKU2-Related Coronavirus of Bat Origin. Nature 2018, 556, 255–258. [Google Scholar] [CrossRef] [PubMed]
  170. Edwards, C.E.; Yount, B.L.; Graham, R.L.; Leist, S.R.; Hou, Y.J.; Dinnon, K.H.; Sims, A.C.; Swanstrom, J.; Gully, K.; Scobey, T.D.; et al. Swine Acute Diarrhea Syndrome Coronavirus Replication in Primary Human Cells Reveals Potential Susceptibility to Infection. Proc. Natl. Acad. Sci. USA 2020, 117, 26915–26925. [Google Scholar] [CrossRef]
Figure 1. Schematic model for cell entry process of coronaviruses. The virus entry process is initiated by cell binding, which is mediated by the interaction between the viral spike (S) protein and cellular receptor and/or attachment factor, followed by membrane fusion and release of the viral genome into the cytoplasm for replication. Membrane fusion occurs at the plasma membrane (early entry) in a pH-independent manner or in endosome/lysosome compartments (late entry), in a pH-dependent manner, or both. Proteolytic cleavage of the S protein is a critical priming step for membrane fusion. Cleavage can occur at the plasma membrane by TMPRSS2 or trypsin or in the endosome/lysosome by cathepsin. Membrane fusion occurring at the cell surface could be inhibited by interferon-inducible host proteins LY6E and IFITM1. Membrane fusion in the endosome/lysosome can be restricted by several other interferon-inducible proteins, including LY6E, IFITM2, IFITM3, CD74, CH25H, and GILT.
Figure 1. Schematic model for cell entry process of coronaviruses. The virus entry process is initiated by cell binding, which is mediated by the interaction between the viral spike (S) protein and cellular receptor and/or attachment factor, followed by membrane fusion and release of the viral genome into the cytoplasm for replication. Membrane fusion occurs at the plasma membrane (early entry) in a pH-independent manner or in endosome/lysosome compartments (late entry), in a pH-dependent manner, or both. Proteolytic cleavage of the S protein is a critical priming step for membrane fusion. Cleavage can occur at the plasma membrane by TMPRSS2 or trypsin or in the endosome/lysosome by cathepsin. Membrane fusion occurring at the cell surface could be inhibited by interferon-inducible host proteins LY6E and IFITM1. Membrane fusion in the endosome/lysosome can be restricted by several other interferon-inducible proteins, including LY6E, IFITM2, IFITM3, CD74, CH25H, and GILT.
Viruses 13 01977 g001
Figure 2. Schematic diagram of the domain architecture of coronaviral spike (S) proteins. Upper panel: domain architecture. Lower panel: a model of S trimer where the S1 forms a globular head and the S2 forms a narrow stalk. FP: fusion peptide. HR1: heptad repeat 1. HR2: heptad repeat 2. TM: transmembrane domain. S1/S2 and S2′ are two types of proteolytic cleavage sites. The S1 domain is required for receptor binding, while the S2 domain mediates membrane fusion.
Figure 2. Schematic diagram of the domain architecture of coronaviral spike (S) proteins. Upper panel: domain architecture. Lower panel: a model of S trimer where the S1 forms a globular head and the S2 forms a narrow stalk. FP: fusion peptide. HR1: heptad repeat 1. HR2: heptad repeat 2. TM: transmembrane domain. S1/S2 and S2′ are two types of proteolytic cleavage sites. The S1 domain is required for receptor binding, while the S2 domain mediates membrane fusion.
Viruses 13 01977 g002
Figure 3. Schematic model for membrane fusion between coronavirus particles and host cells. State 1 (native): the S1 domain is associated with the S2 domain. The fusion peptide (FP) is buried inside the structure and both of the heptad repeats (HRs) exist as trimers individually. State 2 (intermediate state): the S1 domain is dissociated from the S2 domain after receptor binding. Proteolytic cleavage at the S1/S2 boundary and/or at the S2′ site helps the FP expose and then insert into the cell membrane. State 3 (fusion active state): HR2 folds back to HR1 and forms a stable six-helix bundle, which brings viral and cellular membranes into close proximity, forming a fusion pore, which is followed by releasing viral genome into the cytoplasm of host cells. FP: fusion peptide. HR1: heptad repeat 1. HR2: heptad repeat 2. 6-HB: 6-helix bundle.
Figure 3. Schematic model for membrane fusion between coronavirus particles and host cells. State 1 (native): the S1 domain is associated with the S2 domain. The fusion peptide (FP) is buried inside the structure and both of the heptad repeats (HRs) exist as trimers individually. State 2 (intermediate state): the S1 domain is dissociated from the S2 domain after receptor binding. Proteolytic cleavage at the S1/S2 boundary and/or at the S2′ site helps the FP expose and then insert into the cell membrane. State 3 (fusion active state): HR2 folds back to HR1 and forms a stable six-helix bundle, which brings viral and cellular membranes into close proximity, forming a fusion pore, which is followed by releasing viral genome into the cytoplasm of host cells. FP: fusion peptide. HR1: heptad repeat 1. HR2: heptad repeat 2. 6-HB: 6-helix bundle.
Viruses 13 01977 g003
Table 1. Host receptor usage, receptor binding domains, and attachment factors for CoVs.
Table 1. Host receptor usage, receptor binding domains, and attachment factors for CoVs.
GeneraSpecies *ReceptorReceptor Binding Domain (RBD)Attachment Factor
AlphacoronavirusesTGEVPorcine APNC-terminal aa 506–655 [23]Sialic acid [24], DC/L-SIGN [25]
PEDVPorcine APNC-terminal [26]Sialic acid [27], heparan sulfate [28], DC/L-SIGN [25]
HCoV-NL63ACE2C-terminal aa 476–616 [29]Heparan sulfate [30], DC/L-SIGN [31]
HCoV-229EHuman APN (also called CD13)C-terminal aa 407–547 [32,33]DC/L-SIGN [34]
FCoVFeline APNUnknownHeparan sulfate [35], DC/L-SIGN [36]
CCoVCanine APNC-terminal aa 526–676 [37]
BetacoronavirusMHVCEACAM1 (also called CD66a)N-terminal aa 1–330 [38]Sialic acid [27], Heparan sulfate [39]
BCoVUnknownNot yet determined but falls within the N-terminal domain [40]Sialic acid [41], heparan sulfate [39]
HCoV-OC43Unknownaa 339–549 [42]Sialic acid [43], heparan sulfate [35]
MERS-CoVDPP4 (also called CD26)aa 358–588 [44]Sialic acid [27], heparan sulfate [45]
SARS-CoVACE2aa 318–510 [46]Heparan sulfate [47], DC/L-SIGN [48,49]
SARS-CoV-2ACE2aa 331–524 [50]Heparan sulfate [47], DC/L-SIGN [51]
HCoV-HKU1UnknownN/AO-Acetylated Sialic Acid [52]
GammacoronavirusIBVUnknownN-terminal residues 19–272 [53]Sialic acid [54], Heparan sulfate [55], DC/L-SIGN [56]
* TGEV: porcine transmissible gastroenteritis virus; PEDV: porcine epidemic diarrhea virus; HCoV-NL63: human coronavirus NL63; HCoV-229E: human coronavirus 229E; FCoV: feline coronavirus; CCoV: canine coronavirus; MHV: mouse hepatitis virus; BCoV: bovine coronavirus; HCoV-OC43: human coronavirus OC43; MERS-CoV: middle east respiratory syndrome coronavirus; SARS-CoV: severe acute respiratory syndrome coronavirus; SARS-CoV-2: severe acute respiratory syndrome coronavirus 2; HCoV-HKU1: human coronavirus HKU1; IBV: infectious bronchitis virus. The references are provided in the bracket for the RBDs and attachment factors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cheng, Y.-R.; Li, X.; Zhao, X.; Lin, H. Cell Entry of Animal Coronaviruses. Viruses 2021, 13, 1977. https://doi.org/10.3390/v13101977

AMA Style

Cheng Y-R, Li X, Zhao X, Lin H. Cell Entry of Animal Coronaviruses. Viruses. 2021; 13(10):1977. https://doi.org/10.3390/v13101977

Chicago/Turabian Style

Cheng, Yang-Ran, Xinglin Li, Xuesen Zhao, and Hanxin Lin. 2021. "Cell Entry of Animal Coronaviruses" Viruses 13, no. 10: 1977. https://doi.org/10.3390/v13101977

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