Open access peer-reviewed chapter

Respiratory Syncytial Virus

Written By

Sattya Narayan Talukdar and Masfique Mehedi

Submitted: 24 February 2022 Reviewed: 01 April 2022 Published: 28 July 2022

DOI: 10.5772/intechopen.104771

From the Edited Volume

RNA Viruses Infection

Edited by Yogendra Shah

Chapter metrics overview

146 Chapter Downloads

View Full Metrics

Abstract

Respiratory Syncytial Virus (RSV)-driven bronchiolitis is one of the most common causes of pediatric hospitalization. Every year, we face 33.1 million episodes of RSV-driven lower respiratory tract infection without any available vaccine or cost-effective therapeutics since the discovery of RSV eighty years before. RSV is an enveloped RNA virus belonging to the pneumoviridae family of viruses. This chapter aims to elucidate the structure and functions of the RSV genome and proteins and the mechanism of RSV infection in host cells from entry to budding, which will provide current insight into the RSV-host relationship. In addition, this book chapter summarizes the recent research outcomes regarding the structure of RSV and the functions of all viral proteins along with the RSV life cycle and cell-to-cell spread.

Keywords

  • RSV
  • RNA virus
  • RNA genome
  • replicative cycle
  • fusion protein
  • cell-to-cell spread
  • filopodia

1. Introduction

Human respiratory syncytial virus (RSV), despite being a human virus, was first isolated in 1955 from a chimpanzee with respiratory illness [1]. Since its first discovery, it did not take long to isolate RSV from infants with respiratory diseases. Indeed, serological studies verified the existence of RSV infection in infants and children [2, 3]. Now, RSV infection is a prominent cause of lower respiratory tract diseases (bronchiolitis and pneumonia) and hospitalization in children worldwide [4]. According to the most recent virus taxonomy, RSV now belongs to a new family Pneumoviridae of the order Mononegavirales [5].

Advertisement

2. RSV virion

RSV is an enveloped and cytoplasmic virus with non-segmented, negative-sense, single-stranded RNA genome [6]. RSV virions are known to bud out on the infected cell surface. The filamentous virion is up to 12 μm in length and 60 to 200 nm in diameter (Figure 1) [7, 8, 9]. RSV virions can be irregular-shaped spherical particles with a diameter ranging from 100 to 350 nm. Both filamentous and spherical virus particles mostly remain cell-associated [9].

Figure 1.

RSV virion. A photomicrograph of an RSV filamentous virion. The image was taken under electron microscope.

Advertisement

3. RSV strains

There are two RSV strains as RSV A and RSV B and are categorized on basis of genetic and antigenic differences [10]. However, mostly extensive antigenic and nucleotide sequence variation was observed between RSV A and RSV B, however, genetic as well as antigenic variability was also studied within the individual groups of RSV [11]. Multiple studies demonstrated the differences in viral replication between these two groups; specifically, RSV A replicated to higher titers than RSV B viruses in both cell culture and animal models [12, 13, 14, 15, 16]. In addition, RSV A infection is more virulent and severe than RSV B [17].

Advertisement

4. RSV RNA and proteins

The RSV genome is a single-stranded, negative-sense RNA whose length is ranging from 15,191 to 15,226 nucleotides [9]. The RSV genome contains ten genes in the order 3′-NS1-NS2-N-P-M-SH-G-F-M2-L-5′ that are transcribed sequentially into 10 independent messenger RNAs (mRNAs) (Figure 2). Each RSV mRNA encodes a single major protein except for M2, which encodes two separate open reading frames (ORF) for M2–1 and M2–2 proteins, respectively [9, 18, 19, 20]. The M2–1 and M2–2 ORF are located in the upstream and downstream parts of the mRNA, respectively [9]. Like many RNA viruses, RSV brings ribonucleoprotein (RNP) complex as a piece of transcriptional machinery for its genome transcription and replication inside the infected cell cytoplasm. The RNP complex consists of the viral genome, nucleoprotein (N), phosphoprotein (P), and RNA-dependent RNA polymerase [L) [21].

Figure 2.

Schematic of an RSV genome. RSV genome is a negative-sense non-segmented single-stranded RNA. The genome contained 10 genes oriented from 3′ end: NS1, NS2, N, P, M, SH, G, F, M2 (M2–1 and M2–2), and L.

4.1 Nucleoprotein (N)

The RNA genome is wrapped by N (391 amino acids) to create a nuclease-resistant, helical RNP complex called nucleocapsid (NC), and it functions as the template for both replication and transcription [22, 23]. RSV virus genome for replication does not follow the “rule of six” [24], which is common to most paramyxoviruses [25]. The three-dimensional (3D) crystal structure revealed a decameric, ribonucleoprotein complex of N protein and RNA with 3.3 A° resolution and suggested N protein can function as a helicase to separate temporary double-stranded RNA during RNA synthesis [23]. As a decameric structure, every N subunit has a core region comprising two domains, N-terminal and C-terminal, which are linked by a hinge region and the RNA genome turns inside a basic surface groove located at the interface of N-terminal/C-terminal; specifically, every N subunit interacts with seven ribonucleotides of RNA [23].

4.2 RNA-dependent RNA polymerase (L)

The L protein (2165 amino acids) has three enzymatic domains including RNA-dependent RNA polymerase (RdRp) domain, polyribonucleotidyl transferase domain which is essential for capping located in its N-terminal, and methyltransferase domain which is necessary for cap methylation located in C-terminal [26, 27, 28, 29]. Viral mRNA undergoes a post-transcriptional modification before translation and methyltransferase plays a significant role by catalyzing the methylation of cap structure at both N7- and 2′-O-positions because N7-methylation is vital for viral RNA translation and 2′-O-methylation is important for hiding viral RNA from the innate immunity system [30].

4.3 Phosphoprotein (P)

The P protein (241 amino acids) is a homotetrameric protein consisting of N-terminal domain, oligomerization domain, and C-terminal domain and it functions as a cofactor of RdRp and plays a significant role in transcription and replication by networking with other RSV proteins [31, 32, 33, 34, 35]. P protein functions as a multimodular adaptor for RNA synthesis by interacting with N-RNA, L, and M2–1 [36]. P can act as a chaperone for newly synthesized N (N0) protein by forming an N0-P complex that prevents the association of N0 with host RNA [37]. This protein is heavily phosphorylated by host kinase enzymes and it has 41 serine and threonine residues as potential phosphorylation sites; specifically, phosphorylation at residues T105, T188, T210, and S203 are essential for replication, and phosphorylation at residue S156 is vital for viral RNA synthesis [38].

4.4 RSV glycoproteins

As an enveloped virus, the RSV lipid envelope contains three transmembrane glycoproteins including a fusion (F) protein, an attachment glycoprotein (G), and a small hydrophobic (SH) protein; F and G proteins are essential for viral attachment and entry whereas SH protein is less likely involved in viral entry and budding [39, 40].

4.4.1 Fusion (F) protein

Fusion protein is a type 1 transmembrane protein (574 amino acids including a cytoplasmic tail domain of approximately 24 residues) involved in viral entry and assembly [39, 41]. Initially, F protein is synthesized as F0 protein and subsequently, F0 undergoes post-translational modification with multiple N-linked glycosylations depending on RSV strains [42]. To obtain fusion competence, precursor F0 protein (approximately 68–75 KDa) undergoes proteolytic cleavage by furin-like protease which cleaves two polybasic sites and removes a glycosylated peptide of 27 amino acids (Peptide 27 or Pep27) [43, 44]. This cleavage process occurs in the trans-Golgi network and then fusion protein transport to plasma membrane generating two subunits: one is amino-terminal F2 subunit (approximately 15–20 KDa) and another is carboxy-terminal F1 subunit (approximately 50–55 KDa) [45, 46]. A heterodimeric protomer is formed by F1 and F2 subunits covalently connected by disulfide bonds and three protomers combinedly form the matured trimeric form of F protein [47]. After trimerization, F protein exists as a prefusion conformation remaining approximately 12 nm above the membrane of the virus [48]. This prefusion conformation is not a stable form and undergoes a refolding process [6, 49]. This refolding process creates a more stable post-fusion conformation of F protein remaining approximately 17 nm above the viral membrane [50, 51]. The sequence difference of F ectodomains is almost 5% between RSV A and RSV B and therefore, F protein undergoes less antigenic drift and gets preference for suitable vaccine candidates [52].

4.4.2 Attachment glycoprotein (G)

In RSV-infected cells, G protein can exist in two forms; one is a membrane-bound form responsible for viral attachment and another is a secreted isoform responsible for immune evasion [53, 54]. The membrane-bound form (298 amino acids) is a type 2 integral membrane protein [55]. G protein has an amino-terminal cytoplasmic domain and a hydrophobic transmembrane domain; moreover, its ectodomain which undergoes post-translational modification with 4–5 N-linked glycans and 30–40 O-linked glycans, has two mucin-like regions and heparin-binding domains [55, 56, 57]. The translation of secreted G protein starts at an alternative AUG (Met48) located in the transmembrane domain allowing the ectodomain to secrete from the cell [58]. Both membrane-bound and secreted forms of G proteins are thought to be involved in RSV pathogenesis [59]. The higher variation of the mucin-like domain caused two subtypes of RSV: RSV A and RSV B [60].

4.4.3 Small hydrophobic (SH) protein

SH glycoprotein is a small transmembrane protein (64 amino acids for RSV A and 65 amino acids for RSV B) attached by a hydrophobic signal-anchor sequence closer to the N-terminal with extracellular C-terminal orientation; in addition, this protein is considered as less immunogenic because of smaller size and lower abundance during RSV infection [61]. It can exist in several forms including full-length form or post-translational modified form by glycosylation and phosphorylation [62]. Although its function is not clearly understood like other glycoproteins, several studies suggested SH protein can play an auxiliary role during viral fusion along with F glycoprotein; however, SH protein is not crucial for viral entry and syncytium formation [63, 64, 65]. SH protein primarily amasses in the lipid raft membrane of the Golgi complex and endoplasmic reticulum; however, lower levels of SH protein are associated with the envelope of filamentous virus [40]. SH protein did not play an essential role during viral replication in cell culture but SH-deleted RSV infection caused 10-fold lower titers in animal models [39, 66]. It can induce membrane permeability and form pentameric ion channels suggesting its role as viroporins which are short (approximately 100 amino acids) membrane proteins forming oligomers to act as ion channels [67]. Moreover, SH protein is essential to activate the NLRP3 inflammasome [68, 69]. The role of SH protein on apoptosis is not clear because RSV infected A549 cells produced TNF-α and cells were not sensitive to TNF-α-induced death but cells demonstrated a higher level of apoptosis after SH-deleted RSV infection indicating that RSV SH protein may affect the TNF-α pathway resulting in apoptosis delay by an alternative mechanism [70].

4.5 RSV matrix proteins (M and M2)

RSV has two matrix proteins including M protein and M2 protein [58].

4.5.1 M protein

M protein (256 amino acids) is a non-glycosylated protein located in the innermost part of the viral envelope [71]. It is the main protein responsible for viral assembly and budding by interacting with the cell membrane, viral envelope, and viral nucleocapsid [72, 73]. M protein has a zinc finger domain, two clusters of basic amino acids indicating a nuclear localization signal and two nuclear export signals and its N-terminal has lower hydrophobicity; in contrast, C-terminal has higher hydrophobicity [74]. M protein contains multiple phosphorylation sites and undergoes phosphorylation during infection but it is unclear whether these phosphorylations control its function [75]. During the early phase of infection, M protein is present in the host nucleus and inhibits host cellular transcription [76]. During the late phase of infection, M protein is mostly cytoplasmic, interacts with nucleocapsid, and inhibits the activity of viral transcriptase [77]. M protein is located in the cytoplasmic part of the plasma membrane-associated with the lipid rafts along with G and N proteins implying that lipid rafts can function as a platform for the assembly and budding of RSV [73]. M protein is active in a dimer form and the conversion of M-M dimer to oligomer is essential for viral assembly because the interference of dimer formation reduces viral filament maturation and budding [21].

4.5.2 M2 (M2-1 and M2-2) protein

M2–1 and M2–2 are nucleocapsid associated proteins [78]. RSV M2 gene has two overlapping ORFs as M2–1 and M2–2 [79]. The recent crystal structure of the M2–1 (194 amino acids) protein has revealed its native tetrameric form with 2.5 Å resolution and each of its monomers contains three domains including zinc-binding, oligomerization, and core domains [80, 81]. M2–1 functions as a transcriptional anti-terminator and processivity factor [79, 82]. M2–1 did not affect genome and antigenome synthesis indicating that M2–1 is not involved in RNA replication [79, 83]. M2–2 protein (90 amino acids) acts as a regulatory factor switching from transcription to RNA replication because mRNA accumulation was intensely higher after 12–15 hours of infection and then flattened in case of wild-type virus infection but M2–2 knockout virus infection showed continued accumulation [80]. Another study showed M2–2 protein could negatively regulate transcription and positively modulate RNA replication because recombinant RSV infection without NS1 and M2–2 protein demonstrated ten times lower viral growth kinetics in the upper respiratory tract of infants [84].

4.6 RSV nonstructural (NS) proteins (NS1 and NS2)

RSV NS proteins including NS1 (139 amino acids) and NS2 (124 amino acids) play a crucial role in interfering with host innate immunity by forming a “Nonstructural degradosome complex” which can act as a proteasome-like complex that disintegrates a massive number of proteins involved in the innate immune system [85, 86]. Infection with NS1 and NS2 single- and double-gene-deleted RSV demonstrated that both proteins function individually and jointly to accomplish the complete inhibitory effect on type I and III IFNs whereas NS1 has a more individual function [87, 88]. Both NS1 and NS2 target retinoic acid-inducible gene I (RIG-I) like receptors (RLRs), which are considered as host pattern recognition receptors for RIG-I and melanoma differentiation-associated gene 5 (MDA5) [89]. Both NS1 and NS2 induce multiple chemokines and cytokines like RANTES, IL-8, TNFα during viral infection [90]. RIG-I activation by ubiquitination is vital for stimulating antiviral response and tripartite motif-containing protein 25 (TRIM25)-mediated K63-polyubiquitination is essential for RIG-I activation [91]. NS1 protein inhibits RIG-I ubiquitination by interacting with TRIM25 and eventually suppresses type-I interferon (IFN) signaling [92]. Cytosolic NS1 can go to the host nucleus and interacts with the gene regulatory domains of immune response genes, which can control gene transcription and eventually modulates host response against RSV infection [93]. NS1 localized to mitochondria inhibits type-I interferon (IFN) signaling by binding with mitochondrial antiviral signaling protein (MAVS) because the MAVS-RIG-1 complex is essential for type-I IFN activation [94]. NS1 also stimulates miR-29a expression, which affects mRNA coding for interferon alpha/beta receptor 1 (IFNAR1) [95]. NS1 enhances autophagy by the mTOR pathway, which is beneficial for RSV replication but inhibits apoptosis and multiple inflammatory cytokines and IFN-α [96]. Recombinant RSV (NS-deficient) infection showed that mostly NS1 (partially NS2) inhibits the maturation of Dendritic cells, which in turn activates B and T cell responses [97]. NS1 can also inhibit the anti-inflammatory effect of glucocorticoids [98]. The recent X-ray crystal structure of NS2 reveals that it has a unique fold that allows to target molecules different from NS1 and activates distinct IFN antagonism pathway compared to NS1 [99]. Recombinant RSV virus without NS2 showed lower viral growth indicating the role of NS2 in viral replication by evading host immunity [100]. The increased level of IFNβ was not as high when recombinant RSV without NS1 or NS1/NS2 were applied suggesting that both NS1 and NS2 work together for interferon signaling suppression [84]. NS2 also plays a significant role in NF-κB activation, which can initiate a cascade by binding transcription promoters of several proinflammatory cytokines along with IRF-3 and IFN-α/β [90]. In addition to innate immunity, NS2 interferes with adaptive immunity by suppressing CD8+ T-cell responses as a consequence of controlling type 1 IFN [101]. Mostly NS2 along with NS1 play a role in delaying apoptosis, which can enable prolonged RSV replication by activating 3-phophoinositide-dependent protein kinase (PDK)-RAC serine/threonine-protein kinase-glycogen synthase kinase (GSK) pathway [102]. In addition, NS2 plays a significant role in modulating cell morphology, which causes the shedding of infected cells and the spreading of RSV virions [103].

Advertisement

5. Replicative cycle of RSV

5.1 Entry

RSV infection mostly occurs in the apical side of ciliated cells and type 1 pneumocyte; however, several reports suggested the presence of RSV RNA in the extrapulmonary sites and fluids, but more investigations are required [104, 105, 106, 107]. RSV entry has two major phases; the first step is virion attachment to the host cell and the next step is the fusion of viral and host cell membranes in which host factors can involve in both or any individual phases [52]. Heparin-binding domain located between mucin-rich domains of G protein interacts with the unbranched disaccharide polymers specifically glycosaminoglycans (GAGs) connected to transmembrane proteins on the cell surface for the attachment observed in multiple cell culture studies [108, 109, 110]. Variation of G protein lacking heparin-binding domain showed viral attachment indicating the involvement of other regions of G protein during attachment [108]. Negatively charged regions of heparin sulfate contribute mostly and iduronic acid-containing GAG contributes minimally to the attachment [111, 112, 113]. Heparan sulfate proteoglycans (HSGP) act as the receptor for G protein in cell lines; however, recombinant RSV without G protein showed its infectivity; in contrast, HSGP does not express in ciliated epithelial cells, but G protein is still essential for infection in vivo [114, 115, 116]. However, the apical side of ciliated cells, which is the major site of RSV infection lack heparin sulfate indicating the involvement of other host factors, specifically, fractalkine receptor CX3C-chemokine receptor 1 (CX3CR1) bind to CX3C motif of G protein for the attachment [117, 118]. CX3CR1 expressed on the ciliated cells, acts as the receptor of G protein by interacting with its CX3C motif and mutations in the CX3C motif of G protein reduces RSV infection in vivo [117, 119, 120, 121]. F protein is involved in the viral attachment because RSV lacking G and SH proteins grows in cell culture studies and it interacts with heparin sulfate like G protein causing attachment and subsequent infection [63, 122, 123]. Almost 50% infection was observed after heparinase treatment and without GAG synthesis while RSV has F protein suggesting the interaction of F with other host factors; particularly, F protein facilitates entry by interacting with intercellular adhesion molecule 1, insulin-like growth factor 1, epidermal growth factor receptor, and nucleolin [124, 125, 126, 127]. Host and viral membrane then fuse after attachment so that viral particles can enter the cytoplasm and this fusion process is pH-independent and insensitive to lysosomal acidification [128, 129]. RSV infection induces an actin mediated rearrangement followed by plasma membrane blebbing and excess fluid uptake causing internalization of viral particles in a Rab5 positive macropinisome and this endocytic entry depends on the activation of F protein by a second proteolytic cleavage catalyzed by furin-like enzymes after endocytosis observed in A549 cell [130].

5.2 Transcription and replication

RSV replication and transcription are dependent on viral components including viral RNA, N, P, L, and M2–1 [131]. RSV utilizes its own machinery (RNP complex) to replicate in the host cytoplasm [132]. Inclusion body formation is a hallmark of RSV infection produced by multiple viral proteins including N, P, L, and M2–1 and this cytoplasmic structure is increased with RSV infection in epithelial cells [72, 133, 134]. Specifically, N and P proteins are important for inclusion body formation because the expression of these proteins with or without RSV infection showed inclusion body formation [135]. P protein can hijack host cell machinery by forming a complex with host phosphatase (PP1) and this P-PP1 complex dephosphorylates M2–1, as a result, P protein can recruit M2–1 protein in the inclusion body to facilitate viral RNA synthesis [136]. M protein is also reported to localize in inclusion bodies mediated by M2–1 protein [137]. The inclusion body is thought to be the first place where M protein interacts with the ribonucleoprotein complex and M protein is involved in the release of RNP from inclusion bodies towards budding [138]. Host actin cytoskeleton and Hsp70 proteins are also observed in inclusion bodies, but their role is not clear yet and they perhaps facilitate viral machinery [139]. RSV infection causes vigorous stress on the host cell resulting formation of cytoplasmic stress granules, which are different from cytoplasmic inclusion bodies and these stress granule formations facilitates viral replication [140].

Both viral RNA replication and mRNA transcription start from the same single promoter in leader (le) region (44-nucleotide long) at the 3′ end of RSV genome and it produces methyl-guanosine capped and polyadenylated mRNA during transcription and antigenome during replication [20, 141, 142, 143]. Each RSV gene has two conserved cis-acting elements including a gene start (gs) signal at the beginning and a gene end (ge) signal at the end [144]. The promoter of leader (Le) region at the 3′ end of RSV genome has two initiation sites, one is at position +1 or 1 U required for replication and another one is at position +3 or 3C required for transcription [145]. 9 out of 10 gs signaling sequences are highly conserved whereas the tenth one has minimal sequence difference in RSV genome [19]. During transcription, both gs and ge signaling sequences play significant role, specifically, gs signal provides direction to RNA-dependent RNA-polymerase (RdRp) for initiating RNA synthesis and ge signal provides direction to RdRp to polyadenylate and release the mRNA [146, 147]. Then RdRp connected to the template can initiate transcription again at the next gs signal and this process persists along RSV genome [144]. During replication, RdRp attaches a similar promoter sequence in le region, but it ignores ge signal and continues to proceed throughout the genome to produce an antigenome, which is a full-length positive-sense complement of RSV genome [145]. Viral genome and antigenome RNA are encapsidated in RSV nucleoprotein whereas viral mRNAs are not encapsidated [145]. Every nucleoprotein monomer interacts with 7 nucleotides of viral RNA and this complex forms a helical nucleocapsid acting as a template for the next RNA synthesis. This encapsidation is thought to increase RdRp activities to override ge signal during replication, therefore, encapsidation is the distinguishing factor between replication and transcription [23, 148, 149]. The trailer (tr) region (155-neucleotide long) at the 3′ end of RSV antigenome has a promoter, which allows RdRp towards RSV genome synthesis [142, 143, 150]. The first 12 nucleotides of tr promoter are like those of the le promoter and the signal starts from position +1 and + 3 undergoes replication and transcription, respectively, but tr promoter cannot produce capped and polyadenylated mRNA because of lacking ge signal sequence adjacent to tr promoter [151, 152]. However, it is reported that tr promoter can initiate transcription of short RNA, which can inhibit cellular stress granules [153]. The concentration of ATP or GTP can determine the fate of replication and transcription at positions +1 (1 U) or position +3 (3C) observed at in vitro model, specifically, higher ATP concentration stimulates initiation from 1 U and evades initiation at 3C, in contrast, higher GTP concentration displays opposite effect [154]. Overall, L and P proteins form the core RdRp and L-P complex then form L-P-N and L-P-M2–1 complex to initiate replication and transcription, respectively [79, 155].

5.3 Virion assembly and budding

Both assembly and budding of RSV occur at the apical side of ciliated cells [156]. RSV assembly is associated with lipid microdomain or lipid raft rich in cholesterol and sphingolipids; specifically, RSV filament formation observed in caveolin-1 and lipid-raft ganglioside GM1 rich regions of host cell surface membrane [157, 158, 159]. RSV assembly into viral filament occurs at the cell surface requiring the activity of F protein cytoplasmic tail and M protein and this process are not dependent on actin polymerization [160]. However, Mehedi et al., showed the depletion of ARP2 resulted in perturbation of RSV progeny virion on the infected cell surface, consequently reducing viral shedding [8]. Viral assembly requires the activity of F protein cytoplasmic tail and M protein because both proteins accumulate in inclusion bodies cytoplasmic tail of F protein enables the release of the complex of matrix and RNP from inclusion bodies [161]. Although previous studies showed that three proteins including M, P, and F proteins are enough to create virus-like particles, a recent nuclear magnetic resonance study suggests that three novel interaction sites of M on P including site I in αN2 region, site II in 115 to 125 region and oligomerization domain where oligomerization domain is necessary for virus-like structure formation and virus release [137]. The incorporation of RSV proteins into lipid microdomains during virus assembly can cause the interaction of F protein with host factors including caveolin-1, CD44, RhoA, causing microvillus-like projections essential for virus filament and syncytium formation [162, 163]. Actin cytoskeleton and actin-associated pathways linked with PI3K and Rac GTPase are involved in RSV assembly [164]. M protein can bind DNA as well as RNA and it localizes into the nucleus mediated by importin-β1 nuclear import receptor, which forms a complex with guanine nucleotide-binding protein Ran and binds M protein amino acid 100–183 [165, 166]. During the early phase of infection, nuclear accumulation of M protein was observed when M protein interacts with nuclear components mediated by its zinc finger domain resulting in the inhibition of host cell transcription [165]. During the later phase of infection, M protein undergoes phosphorylation inducing nuclear export mediated by Crm1 by unmasking the nuclear export signal [78]. Therefore, M protein is thought to play a regulatory role as a transcription inhibitory factor by inhibiting viral transcriptase to facilitate RSV assembly and budding [77, 167]. RSV glycoprotein and RNP vesicles combined together prior to the filamentous virus formation and G protein recycling has been observed via clathrin-mediated endocytosis, which might be connected with filamentous RSV formation [168]. RSV budding preferentially appears at the apical membrane of epithelial cells by an apical recycling endosome (ARE)-mediated apical protein sorting pathway [169]. RSV budding is independent of the endosomal sorting complex necessary for transport (ESCRT) mechanism controlled by ARE-associated protein, Rab11 family interacting protein 2 (FIP2) [170]. Recently, ARP2 is identified as a novel host factor of RSV budding and cell-to-cell spread [8].

Advertisement

6. RSV cell-to-cell spread

Although RSV progeny virions mostly remain cell-associated, virus shedding occurs from the infected cell’s surface and through cellular protrusions namely filopodia [8, 9]. RSV-induced syncytium (multinucleated cell) formation is a common feature of RSV infection in vitro. The syncytium involves the merging of infected cells with the adjacent uninfected cells, which allows the transfer of viral components from infected cells to the adjacent uninfected cells [171] (Figure 3). Mehedi et al., first showed that RSV uses a novel filopodia-driven cell-to-cell spread mechanism in the lung epithelial cells in vitro (Figure 4). It appears that RSV infection modulates cellular actin dynamics; particularly, actin-related protein 2/3 (ARP2/3) complex-driven actin polymerization contributes to lamellipodium and filopodium formation of cell motility. They showed the depletion of ARP2, a major constituent of the ARP2/3 complex resulted in a substantial reduction of RSV budding and filopodia-driven cell-to-cell spread [8, 172, 173, 174].

Figure 3.

RSV-induced syncytium (multinucleated cell) formation. A549 cells were infected with GFP-expressing RSV (RSV-GFP) at a multiplicity of infection of 1. At 48-hour post-infection, cells were fixed and imaged under an epifluorescence.

Figure 4.

Filopodia-driven RSV cell-to-cell spread. A549 cells were infected with RSV-WT (strain A) at a multiplicity of infection of 1. At 24-hour post-infection, cells were fixed, permeabilized, and stained for RSV F protein by using F-specific immunofluorescence (IFA) (green). F-actin was detected by rhodamine phalloidin staining (red). The image was taken under a stimulated emission depletion (STED) microscope.

References

  1. 1. Morris J, Blount R Jr, Savage R. Recovery of cytopathogenic agent from chimpanzees with goryza. Proceedings of the Society for Experimental Biology and Medicine. 1956;92(3):544-549
  2. 2. Moris JABR, Savage RE. Recovery of cytopathic agent from chimpanzees with coryza. Proc Sci Exp Biol Med. 1956;92:544-550
  3. 3. Chanock R, Finberg L. Recovery from infants with respiratory illness of a virus related to chimpanzee coryza agent (CCA). II. Epidemiologic aspects of infection in infants and young children. American Journal of Hygiene. 1957;66(3):291-300
  4. 4. Verwey C, Nunes MC. RSV lower respiratory tract infection and lung health in the first 2 years of life. The Lancet Global Health. 2020;8(10):e1247-e12e8
  5. 5. Rima B, Collins P, Easton A, Fouchier R, Kurath G, Lamb RA, et al. ICTV virus taxonomy profile: Pneumoviridae. Journal of General Virology. 2017;98(12):2912-2913
  6. 6. Liljeroos L, Krzyzaniak MA, Helenius A, Butcher SJ. Architecture of respiratory syncytial virus revealed by electron cryotomography. Proceedings of the National Academy of Sciences. 2013;110(27):11133-11138
  7. 7. Ke Z, Dillard RS, Chirkova T, Leon F, Stobart CC, Hampton CM, et al. The morphology and assembly of respiratory syncytial virus revealed by cryo-electron tomography. Viruses. 2018;10(8):446
  8. 8. Mehedi M, McCarty T, Martin SE, Le Nouen C, Buehler E, Chen YC, et al. Actin-related protein 2 (ARP2) and virus-induced Filopodia facilitate human respiratory syncytial virus spread. PLoS Pathogens. 2016;12(12):e1006062
  9. 9. Collins PL, Karron RA. Respiratory syncytial virus and Metapneumovirus. In: Knipe DMH, PM editor. Fields Virology. Sixth ed. Philadelphia: Lippincott Williums & Wilkins; 2013
  10. 10. Mufson MA, Örvell C, Rafnar B, Norrby E. Two distinct subtypes of human respiratory syncytial virus. Journal of General Virology. 1985;66(10):2111-2124
  11. 11. Storch GA, Park CS. Monoclonal antibodies demonstrate heterogeneity in the G glycoprotein of prototype strains and clinical isolates of respiratory syncytial virus. Journal of Medical Virology. 1987;22(4):345-356
  12. 12. Crowe JE Jr, Bui PT, Firestone C-Y, Connors M, Elkins WR, Chanock RM, et al. Live subgroup B respiratory syncytial virus vaccines that are attenuated, genetically stable, and immunogenic in rodents and nonhuman primates. Journal of Infectious Diseases. 1996;173(4):829-839
  13. 13. Hierholzer J, Tannock G, Hierholzer CM, Coombs R, Kennett ML, Phillips P, et al. Subgrouping of respiratory syncytial virus strains from Australia and Papua New Guinea by biological and antigenic characteristics. Archives of Virology. 1994;136(1-2):133-147
  14. 14. Morgan L, Routledge E, Willcocks M, Samson A, Scott R, Toms G. Strain variation of respiratory syncytial virus. Journal of General Virology. 1987;68(11):2781-2788
  15. 15. Stott E, Taylor G, Ball L, Anderson K, Young K, King A, et al. Immune and histopathological responses in animals vaccinated with recombinant vaccinia viruses that express individual genes of human respiratory syncytial virus. Journal of Virology. 1987;61(12):3855-3861
  16. 16. Sullender WM, Anderson K, Wertz GW. The respiratory syncytial virus subgroup B attachment glycoprotein: Analysis of sequence, expression from a recombinant vector, and evaluation as an immunogen against homologous and heterologous subgroup virus challenge. Virology. 1990;178(1):195-203
  17. 17. Walsh EE, McConnochie KM, Long CE, Hall CB. Severity of respiratory syncytial virus infection is related to virus strain. Journal of Infectious Diseases. 1997;175(4):814-820
  18. 18. Collins PL, Wertz GW. cDNA cloning and transcriptional mapping of nine polyadenylylated RNAs encoded by the genome of human respiratory syncytial virus. Proceedings of the National Academy of Sciences. 1983;80(11):3208-3212
  19. 19. Collins PL, Dickens LE, Buckler-White A, Olmsted RA, Spriggs MK, Camargo E, et al. Nucleotide sequences for the gene junctions of human respiratory syncytial virus reveal distinctive features of intergenic structure and gene order. Proceedings of the National Academy of Sciences. 1986;83(13):4594-4598
  20. 20. Dickens L, Collins P, Wertz G. Transcriptional mapping of human respiratory syncytial virus. Journal of Virology. 1984;52(2):364-369
  21. 21. Förster A, Maertens GN, Farrell PJ, Bajorek M. Dimerization of matrix protein is required for budding of respiratory syncytial virus. Journal of Virology. 2015;89(8):4624-4635
  22. 22. Bhella D, Ralph A, Yeo RP. Conformational flexibility in recombinant measles virus nucleocapsids visualised by cryo-negative stain electron microscopy and real-space helical reconstruction. Journal of Molecular Biology. 2004;340(2):319-331
  23. 23. Tawar RG, Duquerroy S, Vonrhein C, Varela PF, Damier-Piolle L, Castagné N, et al. Crystal structure of a nucleocapsid-like nucleoprotein-RNA complex of respiratory syncytial virus. Science. 2009;326(5957):1279-1283
  24. 24. Samal SK, Collins PL. RNA replication by a respiratory syncytial virus RNA analog does not obey the rule of six and retains a nonviral trinucleotide extension at the leader end. Journal of Virology. 1996;70(8):5075-5082
  25. 25. Kolakofsky D, Pelet T, Garcin D, Hausmann S, Curran J, Roux L. Paramyxovirus RNA synthesis and the requirement for hexamer genome length: The rule of six revisited. Journal of Virology. 1998;72(2):891-899
  26. 26. Clarke MO, Mackman R, Byun D, Hui H, Barauskas O, Birkus G, et al. Discovery of β-d-2′-deoxy-2′-α-fluoro-4′-α-cyano-5-aza-7, 9-dideaza adenosine as a potent nucleoside inhibitor of respiratory syncytial virus with excellent selectivity over mitochondrial RNA and DNA polymerases. Bioorganic & Medicinal Chemistry Letters. 2015;25(12):2484-2487
  27. 27. Deval J, Hong J, Wang G, Taylor J, Smith LK, Fung A, et al. Molecular basis for the selective inhibition of respiratory syncytial virus RNA polymerase by 2′-fluoro-4′-chloromethyl-cytidine triphosphate. PLoS Pathogens. 2015;11(6):e1004995
  28. 28. Wang G, Deval J, Hong J, Dyatkina N, Prhavc M, Taylor J, et al. Discovery of 4′-chloromethyl-2′-deoxy-3′, 5′-di-O-isobutyryl-2′-fluorocytidine (ALS-8176), a first-in-class RSV polymerase inhibitor for treatment of human respiratory syncytial virus infection. Journal of Medicinal Chemistry. 2015;58(4):1862-1878
  29. 29. Fix J, Galloux M, Blondot M-L, Eléouët J-F. The insertion of fluorescent proteins in a variable region of respiratory syncytial virus L polymerase results in fluorescent and functional enzymes but with reduced activities. The open virology journal. 2011;5:103
  30. 30. Sutto-Ortiz P, Tcherniuk S, Ysebaert N, Abeywickrema P, Noël M, Decombe A, et al. The methyltransferase domain of the respiratory syncytial virus L protein catalyzes cap N7 and 2’-O-methylation. PLoS Pathogens. 2021;17(5):e1009562
  31. 31. SnA E, Paris G, de Prat-Gay G. Modular unfolding and dissociation of the human respiratory syncytial virus phosphoprotein P and its interaction with the M2-1 antiterminator: A singular tetramer–tetramer interface arrangement. Biochemistry. 2012;51(41):8100-8110
  32. 32. Mazumder B, Barik S. Requirement of casein kinase II-mediated phosphorylation for the transcriptional activity of human respiratory syncytial viral phosphoprotein P: Transdominant negative phenotype of phosphorylation-defective P mutants. Virology. 1994;205(1):104-111
  33. 33. Asenjo A, Villanueva N. Regulated but not constitutive human respiratory syncytial virus (HRSV) P protein phosphorylation is essential for oligomerization. FEBS Letters. 2000;467(2-3):279-284
  34. 34. Llorente MT, Taylor IA, López-Viñas E, Gomez-Puertas P, Calder LJ, García-Barreno B, et al. Structural properties of the human respiratory syncytial virus P protein: Evidence for an elongated homotetrameric molecule that is the smallest orthologue within the family of paramyxovirus polymerase cofactors. Proteins: Structure, Function, and Bioinformatics. 2008;72(3):946-958
  35. 35. Llorente MT, García-Barreno B, Calero M, Camafeita E, Lopez JA, Longhi S, et al. Structural analysis of the human respiratory syncytial virus phosphoprotein: Characterization of an α-helical domain involved in oligomerization. Journal of General Virology. 2006;87(1):159-169
  36. 36. Blondot M-L, Dubosclard V, Fix J, Lassoued S, Aumont-Nicaise M, Bontems F, et al. Structure and functional analysis of the RNA-and viral phosphoprotein-binding domain of respiratory syncytial virus M2-1 protein. PLoS Pathogens. 2012;8(5):e1002734
  37. 37. Galloux M, Gabiane G, Sourimant J, Richard C-A, England P, Moudjou M, et al. Identification and characterization of the binding site of the respiratory syncytial virus phosphoprotein to RNA-free nucleoprotein. Journal of Virology. 2015;89(7):3484-3496
  38. 38. Beavis AC, Tran KC, Barrozo ER, Phan SI, Teng MN, He B. Respiratory syncytial virus phosphoprotein residue S156 plays a role in regulating genome transcription and replication. Journal of Virology. 2021;95(24):e01206-e01221
  39. 39. Bukreyev A, Whitehead SS, Murphy BR, Collins PL. Recombinant respiratory syncytial virus from which the entire SH gene has been deleted grows efficiently in cell culture and exhibits site-specific attenuation in the respiratory tract of the mouse. Journal of Virology. 1997;71(12):8973-8982
  40. 40. Rixon HWM, Brown G, Aitken J, McDonald T, Graham S, Sugrue RJ. The small hydrophobic (SH) protein accumulates within lipid-raft structures of the Golgi complex during respiratory syncytial virus infection. Journal of General Virology. 2004;85(5):1153-1165
  41. 41. Oomens AG, Bevis KP, Wertz GW. The cytoplasmic tail of the human respiratory syncytial virus F protein plays critical roles in cellular localization of the F protein and infectious progeny production. Journal of Virology. 2006;80(21):10465-10477
  42. 42. Collins PL, Huang YT, Wertz GW. Nucleotide sequence of the gene encoding the fusion (F) glycoprotein of human respiratory syncytial virus. Proceedings of the National Academy of Sciences. 1984;81(24):7683-7687
  43. 43. Zimmer G, Budz L, Herrler G. Proteolytic activation of respiratory syncytial virus fusion protein: Cleavage at two furin consensus sequences. Journal of Biological Chemistry. 2001;276(34):31642-31650
  44. 44. González-Reyes L, Ruiz-Argüello MB, García-Barreno B, Calder L, López JA, Albar JP, et al. Cleavage of the human respiratory syncytial virus fusion protein at two distinct sites is required for activation of membrane fusion. Proceedings of the National Academy of Sciences. 2001;98(17):9859-9864
  45. 45. Ghildyal R, Jans A, D, G Bardin P, Mills J. Protein-protein interactions in RSV assembly: Potential targets for attenuating RSV strains. Infectious Disorders-Drug Targets (Formerly Current Drug Targets-Infectious Disorders). 2012;12(2):103-109
  46. 46. Bolt G, Pedersen LØ, Birkeslund HH. Cleavage of the respiratory syncytial virus fusion protein is required for its surface expression: Role of furin. Virus Research. 2000;68(1):25-33
  47. 47. Day ND, Branigan PJ, Liu C, Gutshall LL, Luo J, Melero JA, et al. Contribution of cysteine residues in the extracellular domain of the F protein of human respiratory syncytial virus to its function. Virology Journal. 2006;3(1):1-11
  48. 48. McLellan JS, Chen M, Leung S, Graepel KW, Du X, Yang Y, et al. Structure of RSV fusion glycoprotein trimer bound to a prefusion-specific neutralizing antibody. Science. 2013;340(6136):1113-1117
  49. 49. Killikelly AM, Kanekiyo M, Graham BS. Pre-fusion F is absent on the surface of formalin-inactivated respiratory syncytial virus. Scientific Reports. 2016;6(1):1-7
  50. 50. McLellan JS, Yang Y, Graham BS, Kwong PD. Structure of respiratory syncytial virus fusion glycoprotein in the postfusion conformation reveals preservation of neutralizing epitopes. Journal of Virology. 2011;85(15):7788-7796
  51. 51. Swanson KA, Settembre EC, Shaw CA, Dey AK, Rappuoli R, Mandl CW, et al. Structural basis for immunization with postfusion respiratory syncytial virus fusion F glycoprotein (RSV F) to elicit high neutralizing antibody titers. Proceedings of the National Academy of Sciences. 2011;108(23):9619-9624
  52. 52. Battles MB, McLellan JS. Respiratory syncytial virus entry and how to block it. Nature Reviews Microbiology. 2019;17(4):233-245
  53. 53. Hendricks DA, Baradaran K, McIntosh K, Patterson JL. Appearance of a soluble form of the G protein of respiratory syncytial virus in fluids of infected cells. Journal of General Virology. 1987;68(6):1705-1714
  54. 54. Levine S, Klaiber-Franco R, Paradiso P. Demonstration that glycoprotein G is the attachment protein of respiratory syncytial virus. Journal of General Virology. 1987;68(9):2521-2524
  55. 55. Wertz GW, Collins PL, Huang Y, Gruber C, Levine S, Ball LA. Nucleotide sequence of the G protein gene of human respiratory syncytial virus reveals an unusual type of viral membrane protein. Proceedings of the National Academy of Sciences. 1985;82(12):4075-4079
  56. 56. Collins PL, Mottet G. Oligomerization and post-translational processing of glycoprotein G of human respiratory syncytial virus: Altered O-glycosylation in the presence of brefeldin A. Journal of General Virology. 1992;73(4):849-863
  57. 57. Satake M, Coligan JE, Elango N, Norrby E, Venkatesan S. Respiratory syncytial virus envelope glycoprotein (G) has a novel structure. Nucleic Acids Research. 1985;13(21):7795-7812
  58. 58. Sullender WM. Respiratory syncytial virus genetic and antigenic diversity. Clinical Microbiology Reviews. 2000;13(1):1-15
  59. 59. Johnson TR, Johnson JE, Roberts SR, Wertz GW, Parker RA, Graham BS. Priming with secreted glycoprotein G of respiratory syncytial virus (RSV) augments interleukin-5 production and tissue eosinophilia after RSV challenge. Journal of Virology. 1998;72(4):2871-2880
  60. 60. Kapikian AZ, Mitchell RH, Chanock RM, Shvedoff RA, Stewart CE. An epidemiologic study of altered clinical reactivity to respiratory syncytial (RS) virus infection in children previously vaccinated with an inactivated RS virus vaccine. American Journal of Epidemiology. 1969;89(4):405-421
  61. 61. Åkerlind-Stopner B, Hu A, Mufson M, Utter G, Norrby E. Antibody responses of children to the C-terminal peptide of the SH protein of respiratory syncytial virus and the immunological characterization of this protein. Journal of Medical Virology. 1993;40(2):112-120
  62. 62. Olmsted RA, Collins PL. The 1A protein of respiratory syncytial virus is an integral membrane protein present as multiple, structurally distinct species. Journal of Virology. 1989;63(5):2019-2029
  63. 63. Techaarpornkul S, Barretto N, Peeples ME. Functional analysis of recombinant respiratory syncytial virus deletion mutants lacking the small hydrophobic and/or attachment glycoprotein gene. Journal of Virology. 2001;75(15):6825-6834
  64. 64. Heminway B, Yu Y, Tanaka Y, Perrine K, Gustafson E, Bernstein J, et al. Analysis of respiratory syncytial virus F, G, and SH proteins in cell fusion. Virology. 1994;200(2):801-805
  65. 65. Feldman S, Crim R, Audet S, Beeler J. Human respiratory syncytial virus surface glycoproteins F, G and SH form an oligomeric complex. Archives of Virology. 2001;146(12):2369-2383
  66. 66. Jin H, Zhou H, Cheng X, Tang R, Munoz M, Nguyen N. Recombinant respiratory syncytial viruses with deletions in the NS1, NS2, SH, and M2-2 genes are attenuated in vitro and in vivo. Virology. 2000;273(1):210-218
  67. 67. Gan S-W, Tan E, Lin X, Yu D, Wang J, Tan GM-Y, et al. The small hydrophobic protein of the human respiratory syncytial virus forms pentameric ion channels. Journal of Biological Chemistry. 2012;287(29):24671-24689
  68. 68. Triantafilou K, Kar S, Vakakis E, Kotecha S, Triantafilou M. Human respiratory syncytial virus viroporin SH: A viral recognition pathway used by the host to signal inflammasome activation. Thorax. 2013;68(1):66-75
  69. 69. Shen C, Zhang Z, Xie T, Ji J, Xu J, Lin L, et al. Rhein suppresses lung inflammatory injury induced by human respiratory syncytial virus through inhibiting NLRP3 inflammasome activation via NF-κB pathway in mice. Frontiers in Pharmacology. 2020;10:1600
  70. 70. Fuentes S, Tran KC, Luthra P, Teng MN, He B. Function of the respiratory syncytial virus small hydrophobic protein. Journal of Virology. 2007;81(15):8361-8366
  71. 71. Latiff K, Meanger J, Mills J, Ghildyal R. Sequence and structure relatedness of matrix protein of human respiratory syncytial virus with matrix proteins of other negative-sense RNA viruses. Clinical Microbiology and Infection. 2004;10(10):945-948
  72. 72. Li D, Jans DA, Bardin PG, Meanger J, Mills J, Ghildyal R. Association of respiratory syncytial virus M protein with viral nucleocapsids is mediated by the M2-1 protein. Journal of Virology. 2008;82(17):8863-8870
  73. 73. Marty A, Meanger J, Mills J, Shields B, Ghildyal R. Association of matrix protein of respiratory syncytial virus with the host cell membrane of infected cells. Archives of Virology. 2003;149(1):199-210
  74. 74. Ghildyal R, Ho A, Jans DA. Central role of the respiratory syncytial virus matrix protein in infection. FEMS Microbiology Reviews. 2006;30(5):692-705
  75. 75. Peeples ME. Paramyxovirus M proteins. The paramyxoviruses: Springer; 1991. pp. 427-456
  76. 76. Ghildyal R, Baulch-Brown C, Mills J, Meanger J. The matrix protein of human respiratory syncytial virus localises to the nucleus of infected cells and inhibits transcription. Archives of Virology. 2003;148(7):1419-1429
  77. 77. Ghildyal R, Mills J, Murray M, Vardaxis N, Meanger J. Respiratory syncytial virus matrix protein associates with nucleocapsids in infected cells. Journal of General Virology. 2002;83(4):753-757
  78. 78. Ghildyal R, Ho A, Dias M, Soegiyono L, Bardin PG, Tran KC, et al. The respiratory syncytial virus matrix protein possesses a Crm1-mediated nuclear export mechanism. Journal of Virology. 2009;83(11):5353-5362
  79. 79. Collins PL, Hill MG, Cristina J, Grosfeld H. Transcription elongation factor of respiratory syncytial virus, a nonsegmented negative-strand RNA virus. Proceedings of the National Academy of Sciences. 1996;93(1):81-85
  80. 80. Bermingham A, Collins PL. The M2-2 protein of human respiratory syncytial virus is a regulatory factor involved in the balance between RNA replication and transcription. Proceedings of the National Academy of Sciences. 1999;96(20):11259-11264
  81. 81. Tanner SJ, Ariza A, Richard C-A, Kyle HF, Dods RL, Blondot M-L, et al. Crystal structure of the essential transcription antiterminator M2-1 protein of human respiratory syncytial virus and implications of its phosphorylation. Proceedings of the National Academy of Sciences. 2014;111(4):1580-1585
  82. 82. Hardy RW, Wertz GW. The product of the respiratory syncytial virus M2 gene ORF1 enhances readthrough of intergenic junctions during viral transcription. Journal of Virology. 1998;72(1):520-526
  83. 83. Fearns R, Collins PL. Role of the M2-1 transcription antitermination protein of respiratory syncytial virus in sequential transcription. Journal of Virology. 1999;73(7):5852-5864
  84. 84. Teng MN, Whitehead SS, Bermingham A, St. Claire M, Elkins WR, Murphy BR, et al. Recombinant respiratory syncytial virus that does not express the NS1 or M2-2 protein is highly attenuated and immunogenic in chimpanzees. Journal of Virology 2000;74(19):9317-9321.
  85. 85. Boyoglu-Barnum S, Chirkova T, Anderson LJ. Biology of infection and disease pathogenesis to guide RSV vaccine development. Frontiers in Immunology. 2019;10:1675
  86. 86. Sedeyn K, Schepens B, Saelens X. Respiratory syncytial virus nonstructural proteins 1 and 2: Exceptional disrupters of innate immune responses. PLoS Pathogens. 2019;15(10):e1007984
  87. 87. Spann KM, Tran K-C, Chi B, Rabin RL, Collins PL. Suppression of the induction of alpha, beta, and gamma interferons by the NS1 and NS2 proteins of human respiratory syncytial virus in human epithelial cells and macrophages. Journal of Virology. 2004;78(8):4363-4369
  88. 88. Ren J, Liu T, Pang L, Li K, Garofalo RP, Casola A, et al. A novel mechanism for the inhibition of interferon regulatory factor-3-dependent gene expression by human respiratory syncytial virus NS1 protein. The Journal of General Virology. 2011;92(Pt 9):2153
  89. 89. Brisse M, Ly H. Comparative structure and function analysis of the RIG-I-like receptors: RIG-I and MDA5. Frontiers in Immunology. 2019;10:1586
  90. 90. Spann KM, Tran KC, Collins PL. Effects of nonstructural proteins NS1 and NS2 of human respiratory syncytial virus on interferon regulatory factor 3, NF-κB, and proinflammatory cytokines. Journal of Virology. 2005;79(9):5353-5362
  91. 91. Martín-Vicente M, Medrano LM, Resino S, García-Sastre A, Martínez I. TRIM25 in the regulation of the antiviral innate immunity. Frontiers in Immunology. 2017;8:1187
  92. 92. Ban J, Lee N-R, Lee N-J, Lee JK, Quan F-S, Inn K-S. Human respiratory syncytial virus NS 1 targets TRIM25 to suppress RIG-I ubiquitination and subsequent RIG-I-mediated antiviral signaling. Viruses. 2018;10(12):716
  93. 93. Pei J, Beri NR, Zou AJ, Hubel P, Dorando HK, Bergant V, et al. Nuclear-localized human respiratory syncytial virus NS1 protein modulates host gene transcription. Cell Reports. 2021;37(2):109803
  94. 94. Boyapalle S, Wong T, Garay J, Teng M, San Juan-Vergara H, Mohapatra S, et al. Respiratory syncytial virus NS1 protein colocalizes with mitochondrial antiviral signaling protein MAVS following infection. PLoS One. 2012;7(2):e29386
  95. 95. Zhang Y, Yang L, Wang H, Zhang G, Sun X. Respiratory syncytial virus non-structural protein 1 facilitates virus replication through miR-29a-mediated inhibition of interferon-α receptor. Biochemical and Biophysical Research Communications. 2016;478(3):1436-1441
  96. 96. Han B, Wang Y, Zheng M. Inhibition of autophagy promotes human RSV NS1-induced inflammation and apoptosis in vitro. Experimental and Therapeutic Medicine. 2021;22(4):1-9
  97. 97. Munir S, Le Nouen C, Luongo C, Buchholz UJ, Collins PL, Bukreyev A. Nonstructural proteins 1 and 2 of respiratory syncytial virus suppress maturation of human dendritic cells. Journal of Virology. 2008;82(17):8780-8796
  98. 98. Xie J, Long X, Gao L, Chen S, Zhao K, Li W, et al. Respiratory syncytial virus nonstructural protein 1 blocks glucocorticoid receptor nuclear translocation by targeting IPO13 and may account for glucocorticoid insensitivity. The Journal of Infectious Diseases. 2018;217(1):35-46
  99. 99. Pei J, Wagner ND, Zou AJ, Chatterjee S, Borek D, Cole AR, et al. Structural basis for IFN antagonism by human respiratory syncytial virus nonstructural protein 2. Proceedings of the National Academy of Sciences. 2021;118(10):1-12
  100. 100. Teng MN, Collins PL. Identification of the respiratory syncytial virus proteins required for formation and passage of helper-dependent infectious particles. Journal of Virology. 1998;72(7):5707-5716
  101. 101. Kotelkin A, Belyakov IM, Yang L, Berzofsky JA, Collins PL, Bukreyev A. The NS2 protein of human respiratory syncytial virus suppresses the cytotoxic T-cell response as a consequence of suppressing the type I interferon response. Journal of Virology. 2006;80(12):5958-5967
  102. 102. Bitko V, Shulyayeva O, Mazumder B, Musiyenko A, Ramaswamy M, Look DC, et al. Nonstructural proteins of respiratory syncytial virus suppress premature apoptosis by an NF-κB-dependent, interferon-independent mechanism and facilitate virus growth. Journal of Virology. 2007;81(4):1786-1795
  103. 103. Liesman RM, Buchholz UJ, Luongo CL, Yang L, Proia AD, DeVincenzo JP, et al. RSV-encoded NS2 promotes epithelial cell shedding and distal airway obstruction. The Journal of Clinical Investigation. 2014;124(5):2219-2233
  104. 104. Johnson JE, Gonzales RA, Olson SJ, Wright PF, Graham BS. The histopathology of fatal untreated human respiratory syncytial virus infection. Modern Pathology. 2007;20(1):108-119
  105. 105. Xu L, Gao H, Zeng J, Liu J, Lu C, Guan X, et al. A fatal case associated with respiratory syncytial virus infection in a young child. BMC Infectious Diseases. 2018;18(1):1-7
  106. 106. Pitkäranta A, Virolainen A, Jero J, Arruda E, Hayden FG. Detection of rhinovirus, respiratory syncytial virus, and coronavirus infections in acute otitis media by reverse transcriptase polymerase chain reaction. Pediatrics. 1998;102(2):291-295
  107. 107. Rohwedder A, Keminer O, Forster J, Schneider K, Schneider E, Werchau H. Detection of respiratory syncytial virus RNA in blood of neonates by polymerase chain reaction. Journal of Medical Virology. 1998;54(4):320-327
  108. 108. Escribano-Romero E, Rawling J, García-Barreno B, Melero JA. The soluble form of human respiratory syncytial virus attachment protein differs from the membrane-bound form in its oligomeric state but is still capable of binding to cell surface proteoglycans. Journal of Virology. 2004;78(7):3524-3532
  109. 109. Krusat T, Streckert H-J. Heparin-dependent attachment ofrespiratory syncytial virus (RSV) to host cells. Archives of Virology. 1997;142(6):1247-1254
  110. 110. Feldman SA, Hendry RM, Beeler JA. Identification of a linear heparin binding domain for human respiratory syncytial virus attachment glycoprotein G. Journal of Virology. 1999;73(8):6610-6617
  111. 111. Hallak LK, Spillmann D, Collins PL, Peeples ME. Glycosaminoglycan sulfation requirements for respiratory syncytial virus infection. Journal of Virology. 2000;74(22):10508-10513
  112. 112. Martínez I, Melero JA. Binding of human respiratory syncytial virus to cells: Implication of sulfated cell surface proteoglycans. Journal of General Virology. 2000;81(11):2715-2722
  113. 113. Hallak LK, Collins PL, Knudson W, Peeples ME. Iduronic acid-containing glycosaminoglycans on target cells are required for efficient respiratory syncytial virus infection. Virology. 2000;271(2):264-275
  114. 114. Teng MN, Collins PL. The central conserved cystine noose of the attachment G protein of human respiratory syncytial virus is not required for efficient viral infection in vitro or in vivo. Journal of Virology. 2002;76(12):6164-6171
  115. 115. Widjojoatmodjo MN, Boes J, van Bers M, van Remmerden Y, Roholl PJ, Luytjes W. A highly attenuated recombinant human respiratory syncytial virus lacking the G protein induces long-lasting protection in cotton rats. Virology Journal. 2010;7(1):1-10
  116. 116. Zhang L, Bukreyev A, Thompson CI, Watson B, Peeples ME, Collins PL, et al. Infection of ciliated cells by human parainfluenza virus type 3 in an in vitro model of human airway epithelium. Journal of Virology. 2005;79(2):1113-1124
  117. 117. Johnson SM, McNally BA, Ioannidis I, Flano E, Teng MN, Oomens AG, et al. Respiratory syncytial virus uses CX3CR1 as a receptor on primary human airway epithelial cultures. PLoS Pathogens. 2015;11(12):e1005318
  118. 118. Tripp RA, Jones LP, Haynes LM, Zheng H, Murphy PM, Anderson LJ. CX3C chemokine mimicry by respiratory syncytial virus G glycoprotein. Nature Immunology. 2001;2(8):732-738
  119. 119. Ha B, Chirkova T, Boukhvalova MS, Sun HY, Walsh EE, Anderson CS, et al. Mutation of respiratory syncytial virus G protein’s CX3C motif attenuates infection in cotton rats and primary human airway epithelial cells. Vaccine. 2019;7(3):69
  120. 120. Lee H-J, Lee J-Y, Park M-H, Kim J-Y, Chang J. Monoclonal antibody against G glycoprotein increases respiratory syncytial virus clearance in vivo and prevents vaccine-enhanced diseases. PLoS One. 2017;12(1):e0169139
  121. 121. Jeong K-I, Piepenhagen PA, Kishko M, DiNapoli JM, Groppo RP, Zhang L, et al. CX3CR1 is expressed in differentiated human ciliated airway cells and co-localizes with respiratory syncytial virus on cilia in a G protein-dependent manner. PLoS One. 2015;10(6):e0130517
  122. 122. Karron RA, Buonagurio DA, Georgiu AF, Whitehead SS, Adamus JE, Clements-Mann ML, et al. Respiratory syncytial virus (RSV) SH and G proteins are not essential for viral replication in vitro: Clinical evaluation and molecular characterization of a cold-passaged, attenuated RSV subgroup B mutant. Proceedings of the National Academy of Sciences. 1997;94(25):13961-13966
  123. 123. Feldman SA, Audet S, Beeler JA. The fusion glycoprotein of human respiratory syncytial virus facilitates virus attachment and infectivity via an interaction with cellular heparan sulfate. Journal of Virology. 2000;74(14):6442-6447
  124. 124. Behera AK, Matsuse H, Kumar M, Kong X, Lockey RF, Mohapatra SS. Blocking intercellular adhesion molecule-1 on human epithelial cells decreases respiratory syncytial virus infection. Biochemical and Biophysical Research Communications. 2001;280(1):188-195
  125. 125. Currier MG, Lee S, Stobart CC, Hotard AL, Villenave R, Meng J, et al. EGFR interacts with the fusion protein of respiratory syncytial virus strain 2-20 and mediates infection and mucin expression. PLoS Pathogens. 2016;12(5):e1005622
  126. 126. Tayyari F, Marchant D, Moraes TJ, Duan W, Mastrangelo P, Hegele RG. Identification of nucleolin as a cellular receptor for human respiratory syncytial virus. Nature Medicine. 2011;17(9):1132-1135
  127. 127. Griffiths CD, Bilawchuk LM, McDonough JE, Jamieson KC, Elawar F, Cen Y, et al. IGF1R is an entry receptor for respiratory syncytial virus. Nature. 2020;583(7817):615-619
  128. 128. Srinivasakumar N, Ogra P, Flanagan T. Characteristics of fusion of respiratory syncytial virus with HEp-2 cells as measured by R18 fluorescence dequenching assay. Journal of Virology. 1991;65(8):4063-4069
  129. 129. Kahn JS, Schnell MJ, Buonocore L, Rose JK. Recombinant vesicular stomatitis virus expressing respiratory syncytial virus (RSV) glycoproteins: RSV fusion protein can mediate infection and cell fusion. Virology. 1999;254(1):81-91
  130. 130. Krzyzaniak MA, Zumstein MT, Gerez JA, Picotti P, Helenius A. Host cell entry of respiratory syncytial virus involves macropinocytosis followed by proteolytic activation of the F protein. PLoS Pathogens. 2013;9(4):e1003309
  131. 131. Carromeu C, Simabuco F, Tamura R, Arcieri LF, Ventura A. Intracellular localization of human respiratory syncytial virus L protein. Archives of Virology. 2007;152(12):2259-2263
  132. 132. Fearns R, Deval J. New antiviral approaches for respiratory syncytial virus and other mononegaviruses: Inhibiting the RNA polymerase. Antiviral Research. 2016;134:63-76
  133. 133. Garcı́a J, Garcı́a-Barreno B, Vivo A, Melero JA. Cytoplasmic inclusions of respiratory syncytial virus-infected cells: Formation of inclusion bodies in transfected cells that coexpress the nucleoprotein, the phosphoprotein, and the 22K protein. Virology. 1993;195(1):243-247
  134. 134. Norrby E, Marusyk H, Örvell C. Morphogenesis of respiratory syncytial virus in a green monkey kidney cell line (Vero). Journal of Virology. 1970;6(2):237-242
  135. 135. García-Barreno B, Delgado T, Melero JA. Identification of protein regions involved in the interaction of human respiratory syncytial virus phosphoprotein and nucleoprotein: Significance for nucleocapsid assembly and formation of cytoplasmic inclusions. Journal of Virology. 1996;70(2):801-808
  136. 136. Richard C-A, Rincheval V, Lassoued S, Fix J, Cardone C, Esneau C, et al. RSV hijacks cellular protein phosphatase 1 to regulate M2-1 phosphorylation and viral transcription. PLoS Pathogens. 2018;14(3):e1006920
  137. 137. Bajorek M, Galloux M, Richard C-A, Szekely O, Rosenzweig R, Sizun C, et al. Tetramerization of phosphoprotein is essential for respiratory syncytial virus budding while its N-terminal region mediates direct interactions with the matrix protein. Journal of Virology. 2021;95(7):e02217-e02220
  138. 138. Mitra R, Baviskar P, Duncan-Decocq RR, Patel D, Oomens AG. The human respiratory syncytial virus matrix protein is required for maturation of viral filaments. Journal of Virology. 2012;86(8):4432-4443
  139. 139. Santangelo P, Nitin N, LaConte L, Woolums A, Bao G. Live-cell characterization and analysis of a clinical isolate of bovine respiratory syncytial virus, using molecular beacons. Journal of Virology. 2006;80(2):682-688
  140. 140. Lindquist ME, Lifland AW, Utley TJ, Santangelo PJ, Crowe JE Jr. Respiratory syncytial virus induces host RNA stress granules to facilitate viral replication. Journal of Virology. 2010;84(23):12274-12284
  141. 141. Fearns R, Peeples ME, Collins PL. Mapping the transcription and replication promoters of respiratory syncytial virus. Journal of Virology. 2002;76(4):1663-1672
  142. 142. Collins PL, Mink MA, Stec DS. Rescue of synthetic analogs of respiratory syncytial virus genomic RNA and effect of truncations and mutations on the expression of a foreign reporter gene. Proceedings of the National Academy of Sciences. 1991;88(21):9663-9667
  143. 143. Mink MA, Stec DS, Collins PL. Nucleotide sequences of the 3′ leader and 5′ trailer regions of human respiratory syncytial virus genomic RNA. Virology. 1991;185(2):615-624
  144. 144. Kuo L, Grosfeld H, Cristina J, Hill MG, Collins PL. Effects of mutations in the gene-start and gene-end sequence motifs on transcription of monocistronic and dicistronic minigenomes of respiratory syncytial virus. Journal of Virology. 1996;70(10):6892-6901
  145. 145. Tremaglio CZ, Noton SL, Deflubé LR, Fearns R. Respiratory syncytial virus polymerase can initiate transcription from position 3 of the leader promoter. Journal of Virology. 2013;87(6):3196-3207
  146. 146. Kuo L, Fearns R, Collins PL. The structurally diverse intergenic regions of respiratory syncytial virus do not modulate sequential transcription by a dicistronic minigenome. Journal of Virology. 1996;70(9):6143-6150
  147. 147. Liuzzi M, Mason SW, Cartier M, Lawetz C, McCollum RS, Dansereau N, et al. Inhibitors of respiratory syncytial virus replication target cotranscriptional mRNA guanylylation by viral RNA-dependent RNA polymerase. Journal of Virology. 2005;79(20):13105-13115
  148. 148. Gubbay O, Curran J, Kolakofsky D. Sendai virus genome synthesis and assembly are coupled: A possible mechanism to promote viral RNA polymerase processivity. Journal of General Virology. 2001;82(12):2895-2903
  149. 149. McGivern DR, Collins PL, Fearns R. Identification of internal sequences in the 3′ leader region of human respiratory syncytial virus that enhance transcription and confer replication processivity. Journal of Virology. 2005;79(4):2449-2460
  150. 150. Peeples ME, Collins PL. Mutations in the 5′ trailer region of a respiratory syncytial virus minigenome which limit RNA replication to one step. Journal of Virology. 2000;74(1):146-155
  151. 151. Noton SL, Deflubé LR, Tremaglio CZ, Fearns R. The respiratory syncytial virus polymerase has multiple RNA synthesis activities at the promoter. 2012 PLoS Pathogens. 2012;8(10): e1002980
  152. 152. Noton SL, Cowton VM, Zack CR, McGivern DR, Fearns R. Evidence that the polymerase of respiratory syncytial virus initiates RNA replication in a nontemplated fashion. Proceedings of the National Academy of Sciences. 2010;107(22):10226-10231
  153. 153. Hanley LL, McGivern DR, Teng MN, Djang R, Collins PL, Fearns R. Roles of the respiratory syncytial virus trailer region: Effects of mutations on genome production and stress granule formation. Virology. 2010;406(2):241-252
  154. 154. Cressey TN, Noton SL, Nagendra K, Braun MR, Fearns R. Mechanism for de novo initiation at two sites in the respiratory syncytial virus promoter. Nucleic Acids Research. 2018;46(13):6785-6796
  155. 155. Grosfeld H, Hill MG, Collins PL. RNA replication by respiratory syncytial virus (RSV) is directed by the N, P, and L proteins; transcription also occurs under these conditions but requires RSV superinfection for efficient synthesis of full-length mRNA. Journal of Virology. 1995;69(9):5677-5686
  156. 156. Roberts SR, Compans RW, Wertz GW. Respiratory syncytial virus matures at the apical surfaces of polarized epithelial cells. Journal of Virology. 1995;69(4):2667-2673
  157. 157. Brown G, Aitken J, Rixon HWM, Sugrue RJ. Caveolin-1 is incorporated into mature respiratory syncytial virus particles during virus assembly on the surface of virus-infected cells. Journal of General Virology. 2002;83(3):611-621
  158. 158. Brown G, Rixon HWM, Sugrue RJ. Respiratory syncytial virus assembly occurs in GM1-rich regions of the host-cell membrane and alters the cellular distribution of tyrosine phosphorylated caveolin-1. Journal of General Virology. 2002;83(8):1841-1850
  159. 159. Simons K. Ikonen E. Functional rafts in cell membranes. nature. 1997;387(6633):569-572
  160. 160. Shaikh FY, Utley TJ, Craven RE, Rogers MC, Lapierre LA, Goldenring JR, et al. Respiratory syncytial virus assembles into structured filamentous virion particles independently of host cytoskeleton and related proteins. PLoS One. 2012;7(7):e40826
  161. 161. Baviskar PS, Hotard AL, Moore ML, Oomens AG. The respiratory syncytial virus fusion protein targets to the perimeter of inclusion bodies and facilitates filament formation by a cytoplasmic tail-dependent mechanism. Journal of Virology. 2013;87(19):10730-10741
  162. 162. McCurdy LH, Graham BS. Role of plasma membrane lipid microdomains in respiratory syncytial virus filament formation. Journal of Virology. 2003;77(3):1747-1756
  163. 163. Gower TL, Pastey MK, Peeples ME, Collins PL, McCurdy LH, Hart TK, et al. RhoA signaling is required for respiratory syncytial virus-induced syncytium formation and filamentous virion morphology. Journal of Virology. 2005;79(9):5326-5336
  164. 164. Jeffree CE, Brown G, Aitken J, Su-Yin DY, Tan B-H, Sugrue RJ. Ultrastructural analysis of the interaction between F-actin and respiratory syncytial virus during virus assembly. Virology. 2007;369(2):309-323
  165. 165. Ghildyal R, Ho A, Wagstaff KM, Dias MM, Barton CL, Jans P, et al. Nuclear import of the respiratory syncytial virus matrix protein is mediated by importin β1 independent of importin α. Biochemistry. 2005;44(38):12887-12895
  166. 166. Rodríguez L, Cuesta I, Asenjo A, Villanueva N. Human respiratory syncytial virus matrix protein is an RNA-binding protein: Binding properties, location and identity of the RNA contact residues. Journal of General Virology. 2004;85(3):709-719
  167. 167. Ghildyal R, Li D, Peroulis I, Shields B, Bardin PG, Teng MN, et al. Interaction between the respiratory syncytial virus G glycoprotein cytoplasmic domain and the matrix protein. Journal of General Virology. 2005;86(7):1879-1884
  168. 168. Vanover D, Smith DV, Blanchard EL, Alonas E, Kirschman JL, Lifland AW, et al. RSV glycoprotein and genomic RNA dynamics reveal filament assembly prior to the plasma membrane. Nature Communications. 2017;8(1):1-15
  169. 169. Brock SC, Goldenring JR, Crowe JE. Apical recycling systems regulate directional budding of respiratory syncytial virus from polarized epithelial cells. Proceedings of the National Academy of Sciences. 2003;100(25):15143-15148
  170. 170. Utley TJ, Ducharme NA, Varthakavi V, Shepherd BE, Santangelo PJ, Lindquist ME, et al. Respiratory syncytial virus uses a Vps4-independent budding mechanism controlled by Rab11-FIP2. Proceedings of the National Academy of Sciences. 2008;105(29):10209-10214
  171. 171. Cifuentes-Munoz N, Dutch RE, Cattaneo R. Direct cell-to-cell transmission of respiratory viruses: The fast lanes. PLoS Pathogens. 2018;14(6):e1007015
  172. 172. Mehedi M, Collins PL, Buchholz UJ. A novel host factor for human respiratory syncytial virus. Communicative & Integrative Biology. 2017;10(3):e1319025
  173. 173. Mehedi M, Smelkinson M, Kabat J, Ganesan S, Collins PL, Buchholz UJ. Multicolor stimulated emission depletion (STED) microscopy to generate high-resolution images of respiratory syncytial virus particles and infected cells. Bio-Protocol. 2017;7(17):1-10
  174. 174. Paluck A, Osan J, Hollingsworth L, Talukdar SN, Saegh AA, Mehedi M. Role of ARP2/3 complex-driven actin polymerization in RSV infection. Pathogens. 2021;11(1):1-12

Written By

Sattya Narayan Talukdar and Masfique Mehedi

Submitted: 24 February 2022 Reviewed: 01 April 2022 Published: 28 July 2022