Next Article in Journal
Pathological Features of Echovirus-11-Associated Brain Damage in Mice Based on RNA-Seq Analysis
Next Article in Special Issue
Zoonotic Origins of Human Metapneumovirus: A Journey from Birds to Humans
Previous Article in Journal
Prior Methamphetamine Use Disorder History Does Not Impair Interoceptive Processing of Soft Touch in HIV Infection
Previous Article in Special Issue
Comparable Infection Level and Tropism of Measles Virus and Canine Distemper Virus in Organotypic Brain Slice Cultures Obtained from Natural Host Species
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immunopathology of RSV: An Updated Review

by
Harrison C. Bergeron
and
Ralph A. Tripp
*
Department of Infectious Diseases, College of Veterinary Medicine, University of Georgia, Athens, GA 30602, USA
*
Author to whom correspondence should be addressed.
Submission received: 15 November 2021 / Revised: 6 December 2021 / Accepted: 8 December 2021 / Published: 10 December 2021
(This article belongs to the Special Issue Immunopathogenesis of Paramyxoviridae and Pneumoviridae)

Abstract

:
RSV is a leading cause of respiratory tract disease in infants and the elderly. RSV has limited therapeutic interventions and no FDA-approved vaccine. Gaps in our understanding of virus–host interactions and immunity contribute to the lack of biological countermeasures. This review updates the current understanding of RSV immunity and immunopathology with a focus on interferon responses, animal modeling, and correlates of protection.

1. Respiratory Syncytial Virus (RSV)

1.1. RSV Overview

RSV is a member of the Pneumoviridae genus and contains a single-stranded non-segmented negative-sense RNA genome approximately 15,200 nt in length [1]. Its genome contains 10 open reading frames (ORFs) which encode 11 proteins. From 3′ to 5′, these genes include two non-structural proteins (NS1 and NS2), two nucleocapsid proteins (N and P), one inner envelope membrane protein (M1), three surface proteins that coat the virion—small hydrophobic (SH), attachment (G), and fusion (F)—M2 which contains overlapping ORFs, resulting in the production of M2.1 and M2.2, and large (L) protein [2]. G and F proteins are the major antigenic proteins. RSV is pleomorphic, i.e., spherical, asymmetrical, and filamentous [3], and categorized into subgroups A and B based on the sequence of the G protein [4,5,6]. RSV subtypes co-circulate and often cause reinfection with the same strain [7].
RSV can be transmitted through respiratory droplets or fomites which infect the upper respiratory tract (URT) via nasopharyngeal or conjunctival mucosa [8]. From the URT, RSV spreads to the lower respiratory tract, primarily infecting polarized ciliated human airway epithelial cells (hAECs) [9], leading to lower respiratory tract infection (LRTI), bronchiolitis, and/or pneumonia [10,11,12]. Along with respiratory epithelial cells, RSV has also been reported to infect CX3CR1+ neonatal regulatory B lymphocytes [13], primary neurons [14], alveolar macrophages [15], dendritic cells [16], neutrophils [17], mast cells [18], and T cells [19].
Several host cell receptors for RSV are proposed. The attachment (G) protein is responsible for virus attachment and is a ~300 amino acid glycoprotein consisting of cytoplasmic (CP), transmembrane (TM), and extracellular (ecto) domains [20]. Due to an alternative translation site within the TM (Met48), RSV G protein is membrane bound (mG) and soluble (sG) [21]. sG can be detected as early as 12 h post-infection (pi) and is thought to act as an antigen decoy [22] and induce aberrant immune responses [21,23,24]. G protein contains two heavily glycosylated mucin-like domains that flank a highly conserved central conserved domain (CCD), a CX3C motif, and a heparin-binding domain (HBD) [25]. In immortalized cell lines, the HBD is responsible for substantial binding via cell surface glycosaminoglycans (GAGs) which can be blocked with the addition of exogenous heparin sulfate [20,26]. For hAECs lacking any detectable heparin sulfate or other proteoglycans, other modalities of attachment are required [27]. The conserved CX3C chemokine motif in the G protein binds to CX3CR1, a chemokine receptor found on some immune cells and respiratory epithelial cells [28,29]. CX3CR1 is the receptor for the chemokine CX3CL1 (fractalkine or FKN) [30,31,32,33,34]. Like G protein, FKN is membrane bound and soluble, is glycosylated, and contains HBDs [35]. Recent high-resolution crystal structures suggest conformational epitopes requiring proper folding of the cysteine noose located in the central conserved region, suggesting that G protein binds to CX3CR1 different from CX3CL1 despite functional mimicry [36]. Other studies have proposed that annexin II [37] and Toll-like receptor 4 (TLR-4) may also bind G protein [38].
Once G protein attaches to respiratory epithelial cells, the F protein mediates cell fusion, resulting in viral entry and infection [39,40]. This fusion event is catalyzed by the prefusion F protein binding to the host cell receptor causing a conformational change to a post-fusion conformation, fusing the virion with the host cell, and the formation of the fusion pore [41]. Receptors for this fusion event include nucleolin [42,43,44], and co-receptor candidates TLR-4 [45,46], EGFR [47], ICAM-1 [48], IGFR-1 [49], and C-type lectins [50]. Fusion releases vRNA into the host cytoplasm where the nucleoproteins (N, P, and L) initiate transcription in a 3′ to 5′ gradient fashion. The polymerase transcribes each gene, resulting in 5′ to 3′ subgenomic mRNAs, which are translated into viral proteins by the host cell machinery. To replicate the viral genome, polymerase converts the genome to antisense (5′ to 3′) to be used as the template for generating new negative-sense copies. Genomes assemble with viral proteins in the cytoplasm of infected cells to form new virions. Once formed, virion buds or syncytia are formed with neighboring cells mediated by F protein. Syncytia of epithelial cells leads to pathology through alteration of the airway integrity [51].

1.2. RSV Epidemiology

RSV infection causes a substantial disease burden in the infant, immunocompromised, and elderly populations with nearly all children infected with RSV by age two [52,53,54]. It is estimated to cause between 55,000 and 200,000 deaths in children under 5 years of age annually with the most serious disease in infants <1 year of age especially in low-income countries [55,56]. RSV is a leading cause of infant hospitalization and contributes substantially to medical intervention required for the elderly [57,58,59]. Moderate-to-severe RSV disease may also lead to the development of asthma and chronic wheeze later in life, even in children with no atopic predisposition [60,61,62,63]. This phenomenon is potentially mediated by the immune sensitization to RSV, lung development, neuronal development, or a combination of these factors and others [64]. Maternal antibodies may offer some protection to newborn infants; however, these antibodies wane within weeks after birth, and titers of antibodies vary between mothers [65,66]. Of note, during the COVID-19 pandemic, RSV infections (as well as other respiratory viruses included influenza A, influenza B, and adenovirus) fell possibly due to improved hygiene practices, social distancing, and school closures [67,68]. Children in Melbourne, Australia had between 68.8 and 100% reduction in RSV cases, which correlated with the strictness of the lockdown measures in place at that time [67].
Despite the known disease burden, there is no FDA-approved RSV vaccine available. As of July 2019, there were 121 clinical trials evaluating RSV vaccines [69]. In brief, vaccine platforms including virus particle based, nucleic acid, live attenuated, subunit and vector based are being pursued for infants, children, and the elderly as well as maternal vaccines to protect newborn infants [70,71,72]. Many vaccine candidates largely focus on generating pre-F antibodies that are potently neutralizing; live-attenuated vaccines may exclude certain proteins or epitopes while retaining the pre-F conformation [73]. Importantly, live-attenuated vaccines are likely exclusive to the infant population due to a lack of prior exposure while particle-based vaccines may be used for the various populations at risk for RSV disease [74].
The only specific prophylactic countermeasure available is palivizumab (Synagis®) which was licensed over 20 years ago [75]. Palivizumab is a humanized monoclonal Ab (mAb) targeting the F protein and is restricted for use in high-risk infants [76]. Palivizumab is administered monthly by injection and reduces hospitalization by ~50%; however, it is not recommended for treatment of wheeze or asthma by the American Academy of Pediatrics [77] and is only approved prophylactic [78]. Next-generation mAb candidates are in preclinical and clinical trials some demonstrating promise as a superior drug, e.g., nirsevimab [79,80]. mAbs including motavizumab and suptavumab (REGN2222) have recently failed to reach clinical trial endpoints [81]. Small- and large-molecule drugs in RSV therapeutic development have recently been reviewed [82].

2. The Immune Response to RSV

The respiratory system has a myriad of physical and biochemical features to protect it from agents such as viruses. Respiratory epithelial cells form tight junctions lining the respiratory tract and secrete mucins to help clear pathogens [83,84]. Surfactant-associated proteins (SPs) are members of collectins (collagenous C-type lectins) which reside on the apical surface of the epithelium and assist in the opsonization of potential pathogens [85]. Along with delivering a physical barrier, respiratory epithelial cells may also be phagocytic for pathogens such as bacterial and fungus as these cells also act as secondary phagocytic cells engulfing pathogens [86]. Respiratory epithelial cells are the primary target for RSV infection [87,88].

2.1. Innate Responses

Innate immunity uses germline-encoded receptors, i.e., pattern recognition receptors (PRRs) that respond to pathogen-associated molecular patterns (PAMPs) conserved among pathogens [89]. PRR-sensing leads to signaling cascades that induce transcription factors that upregulate antiviral and pro-inflammatory cytokines [90,91]. The antiviral cytokines include type I (IFNα, IFNβ) and type III (IFNλ) interferons (IFNs), whereas type II (IFNγ) IFN promotes immune cell activation [92,93]. Early IFN expression induces IFN-stimulated genes (ISGs) which modify the immune response [94]. Viral replication induces differential kinetics and magnitudes of the host response affecting the outcome of an innate and adaptive immune response. Leukocytes mediate the innate response, and the cytokines and chemokines produced to contribute to host protection from virus-induced pathology.

2.1.1. Neutrophils or PMNs

One of the first innate immune cells to respond to RSV infection are neutrophils or polymorphonuclear leukocytes (PMNs) that include neutrophils, eosinophils, basophils, and mast cells [95]. In one study sampling the bronchoalveolar lavage (BAL) of infants with severe RSV bronchiolitis, neutrophils accounted for a majority (76–93%) of innate immune cells [96]. RSV infection induces interleukin (IL)-8 secretion, a neutrophil chemokine [97,98]. It has been shown that RSV F protein can induce NETosis, a form of cell death characterized by the release of decondensed chromatin and granular contents from neutrophils to the extracellular space associated with binding of TLR-4, thus illustrating a potential mechanism for inflammation-induced by RSV F protein [99]. TLR-4 is canonically associated with lipopolysaccharides (LPS), an endotoxin found on Gram-negative bacteria that results in upregulation of IL-8 [100]. Thus, the initial mediation of TLR-4 by RSV F protein induces a positive neutrophil feedback loop recruiting more neutrophils to the lung and potential immunopathology.
Polymorphisms that result in increased IL-8 secretion are associated with greater RSV disease severity and wheeze [101,102], and neutrophils have been linked to epithelial cell damage [103] and mortality in young children with untreated RSV bronchiolitis [104]. Recent nasal immune profiling of infants with RSV bronchiolitis showed increased levels of IL-8 [105], and global mRNA expression showed increased neutrophil signatures in severe vs. mild RSV infection in infants [106]. Further, transient neutropenia due to young age was not a risk factor for immunocompetent infants to develop serious RSV disease [107]. TLR-2 is also involved in pulmonary neutrophil during early infection (24 hpi) linked to the expression of CCL2 [108] which may be in part driven by RSV G protein [109]. Along with NETosis, neutrophils also produce pro-inflammatory cytokines such as tumor necrosis factor-alpha (TNFα) which may contribute to immunopathology [110].
Early RSV vaccine trials investigated formalin-inactivated RSV (FI-RSV) vaccination of young children that unfortunately resulted in enhanced RSV disease including two infant deaths following natural RSV infection [111,112,113]. Initial reports suggested eosinophilia as driving immunopathology; however, recent analysis suggests that neutrophilia has a major role in enhanced respiratory disease (ERD) [12,114]. Contrary to these findings, one published report in mice suggests that neutrophils do not affect lung virus load or contribute to pro-inflammatory responses following RSV infection [115]. Comparing human responses to those in mice uses different metrics especially when trying to compare mild to severe disease given that mice are only semi-permissive to RSV, it is plausible that neutrophils impact RSV immunopathology.

2.1.2. Alveolar Macrophages

Alveolar macrophages (AMs) are present at the luminal surface of the alveolar lung space and are early responders to lung epithelial insult [116]. RSV infection mouse or human AMs induces TNFα-mediated necrosis mediated through the RIPK1/3/MLKL expression pathways [117]. RSV replication leads to upregulated macrophage migration inhibitory factor (MIF) expression leading to modified cytokine production by AMs [118]. Interestingly, this pathway is associated with increases in the pro-inflammatory TNFα as well as anti-inflammatory IL-10, suggesting that regulation of these cytokines is pivotal in balancing protection and pathology. One such regulator is myeloid PPAR-γ expression that in mice has been shown to reduce inflammatory markers such as TNFα and IL-1β [119]. AMs stimulated by TNFα and monocyte chemoattractant protein 1 (MCP-1) in the allergic airway murine model resulted in increased production of IFNγ and IL-27 [120]. Another regulator, i.e., transforming growth factor (TGF)-β1 disrupts antiviral host responses including type I IFNs during RSV infection in human and murine AMs [121]. AMs have been shown to also express IL-33, a driver of Th2-associated cytokine production, in a mitogen-activated protein kinase (MAPK)-dependent pathway leading to activation of nuclear factor kappa-light-chain enhancer of activated B cells (NF-κB) [122].
Relevant to FI-RSV-mediated ERD, challenging FI-RSV immune mice with RSV results in fewer AMs compared to non-primed or virus-like particle (VLP) F protein primed mice [123]. AMs expressing CD169+ are responsible for the capture of pathogens and are frequently the first cell type infected and thereby provide a confined source of antigen [124]. Interestingly, diphtheria toxin receptor (DTR) transgenic mice depleted of CD169+ AM cells had reduced pro-inflammatory BAL cytokines while chemokine levels were not affected, and there was an increase in lung inflammatory cells (monocytes, neutrophils, and eosinophils) following RSV challenge [125]. Analogously, RSV G and/or SH proteins have been shown to reduce chemokine expression (e.g., MCP-1 and MIPs) early after RSV infection with a reduction in non-tissue-resident macrophages occurring during RSV infection [126].

2.1.3. Eosinophils

Pulmonary eosinophilia is a hallmark of ERD in animal models of FI-RSV vaccination [127]. Some studies have linked ERD to RSV G protein sensitization. For example, sensitizing mice with recombinant vaccinia virus expressing RSV G protein (vvG), and not F protein or N protein, followed by RSV infection was shown to lead to substantially increased pulmonary eosinophils [128]. In contrast, mice intranasally infected with an RSV mutant virus lacking the G and SH genes develop pulmonary eosinophilia after RSV challenge indicating that the G and SH proteins are not solely responsible for ERD [129]. Other studies have shown that RSV G protein priming modulates eosinophil trafficking and function [130,131,132], and sublingual administration with the RSV G protein CCD primes mice for pulmonary eosinophilia and results in greater eosinophils compared to FI-RSV priming [133]. Sublingual priming was also shown to prime both pulmonary eosinophilia and neutrophilia in the lung tissue [134].
Eosinophilia and neutrophilia are linked to IL-17, as IL-17 depletion rescues this phenotype, and IL-17 is associated with other respiratory diseases such as asthma [135]. Additionally, the expression of IL-5 has been associated with pulmonary eosinophilia and airway hyperresponsiveness following RSV challenge or vaccinia virus G protein (vvG) priming [136,137]. Consistent with this finding, studies with overlapping G protein peptides showed that the G184–198 peptide encompassing the CX3C motif stimulates G protein primed murine splenocytes and peripheral blood mononuclear cells (PBMCs) to express IFNγ and IL-5 [138]. While this may indicate a level of immunopathology, one study showed a protective role for eosinophils expressing IL-5 in RSV clearance [139]. Another study examining IL-5/eotaxin knockout mice showed decreased mucus and airway inflammation following FI-RSV vaccination and RSV challenge with mice expressing lower IL-4 and IL-13 but increased IFNγ levels where the mice maintained RSV infection for a longer duration [140]. In summary, it is unclear if RSV G protein-induced pulmonary eosinophils are pathogenic or protective.
IL-13 expression, ERD, and pulmonary eosinophilia are mediated by vvG and FI-RSV priming [141,142,143]. sG protein has been shown to mediate pulmonary eosinophilia in mice and drive Th2 responses demonstrating its ability to induce aberrant immune responses [131]. These features may be advantageous to reduce RSV clearance in the host. Importantly, it has been shown that the form of G protein, i.e., sG or mG, and how it is delivered impacts the host response to RSV. For example, intranasal delivery of liposome-encapsulated G protein results in reduced pulmonary eosinophilia compared to G protein delivered without liposome encapsulation [144]. The significance of eosinophils in ERD is questioned by the findings from a study of eosinophil-deficient mice primed with vvG, as ERD as measured by weight loss, clinical scores, Penh levels, and pro-inflammatory cytokines (e.g., IFNs and TNFα) still developed despite the absence of eosinophils [145]. Given these different outcomes, it suggests that eosinophils are multifunctional, and their role remains under investigation particularly given the association of eosinophils and asthma [146].

2.1.4. Natural Killer (NK) Cells

NK cells are innate effector lymphocytes that typically control tumors and microbial infections, and are regulatory cells engaging in interactions with dendritic cells, macrophages, T cells, and endothelial cells [147]. NK cells can limit or exacerbate immune responses, promote and influence inflammatory responses, and NK cells can regulate RSV infection. It has been shown that severe RSV disease in infants correlates with single-nucleotide polymorphisms (SNPs) that increased leukocyte immunoglobulin-like receptor B1 (LILRB1+) NK cells [148]. LILRB1+ NK cells are noteworthy as the population of these cells in infants is generally low, and the receptor functions as an inhibitor of immune responses. Interestingly, in CD4 knockout mice, NK cells were shown to be recruited to the lung, and disease was reduced after FI-RSV priming and RSV challenge, suggesting that NK cells and disease are inversely correlated in the absence of CD4 T cells [149]. Mice infected with a recombinant RSV expressing IL-18 to enhance NK cell activation had reduced lung viral loads compared to wild-type infection and exhibited biphasic weight loss (days 2 and 6) not observed in wild-type mouse infection [150]. The role of NK cells was substantiated by depletion studies in mice infected with RSV expressing IL-18.
An in vitro study examining antibody-dependent enhancement (ADE) showed RSV co-incubated with suboptimal concentrations of neutralizing antibodies led to ADE and increased lung viral loads, enhanced numbers of NK cells, and increased IFNγ expression by NK cells [151]. The NK cells did not secrete increased perforin, suggesting that they were not directly cytotoxic. This observation is relevant when considering maternal antibodies that are induced by RSV vaccination as the findings imply that the maternal antibodies may prime NK cells; however, RSV vaccination may enhance or diminish the response [152].
The NK cell response has been linked to RSV pathogenesis as NK cell depletion reduced disease [153]. Interestingly, NK cell trafficking and function were shown to be diminished in TLR-4-deficient mice compared to C57BL/10Sn (TLR-4 expressing) mice after RSV infection [154]. This may be linked to RSV F and G proteins both of which are implicated in agonism and antagonism of TLR-4, respectively. Mice infected with wild-type RSV B1 have decreased NK cell infiltrates in the BAL compared to mice infected with RSV B1 lacking G and SH genes (CP52), suggesting a role for G and/or SH proteins in modifying NK cell responses [129]. The neuropeptide, substance P (SP), was shown to also influence NK cell immunity to RSV as anti-SP mAb treatment resulted in increased NK cells and IFNγ expression [155]. Notably, priming mice with vvG did not result in effect pulmonary NK cells while F and M2 priming did [156]. Taken together, the G protein is implicated in modifying the NK response.

2.1.5. Dendritic Cells

Dendritic cells (DCs) are professional antigen-presenting cells (APC) that function to bridge the innate and adaptive immune responses, and constantly scan the environment to present foreign antigens to adaptive immune cells. DCs have two subpopulations, i.e., plasmacytoid (pDC) or conventional (cDCs) [157]. pDCs and cDCs are similar in that they present antigen to T cells; however, pDCs secrete a substantial amount of type I IFNs [158,159]. pDCs produce higher levels of IFNβ via TLR7/MyD88 signaling during RSV infection compared to cDCs [160]. pDC-depletion in mice has been shown to impair the CTL response and result in increased lung viral loads. Human DCs upregulate CD38 expression via the type I IFN response at early time points post-infection [161]. In IFNβ/YFP reporter mice, MyD88 but not TLR-7 is required to induce IFNβ during RSV infection [162]. Taken together, pDCs secrete substantial type I IFNs during RSV infection via the MyD88 pathway.
cDCs have increased IL-4Rα expression that has been correlated with more severe immunopathology in mice and the development of a Th2 biasing response [163]. Since infants have greater IL-4Rα+ cDCs compared to older children this feature may correlate to increased disease severity in younger infants. Overexpression of IL-4Rα in murine cDCs leads to enhanced immunopathology similarly to that of neonatal mice following the RSV challenge [163]. During RSV infection, cDCs and pDCs are increased in the lung even weeks after the infection is cleared, and depleting pDCs enhances pathology and alters cytokine responses post-challenge [164,165]. Interestingly, it was shown that the IFNα produced by pDCs act to clear RSV but does not alter the adaptive immune response or subsequent pathology. In neonatal mice, RSV infection poorly induces IFNα production leading to an IL-4Rα-dependent Th2 response and a lack of viral clearance [159]. pDCs also express IL-33, a cytokine implicated in airway inflammation which leads to pathology during RSV infection [166]. Despite the presence of IFNα or IFNβ, neonatal cDCs are unable to upregulate T cell co-stimulatory molecules including CD80 and CD86 during antigen presentation, which reduces CTL cell priming [167]. Another study evaluating human cord blood cDCs showed limited maturation after sensitizing with RSV; however, influenza A virus sensitization resulted in the maturation of the cDCs demonstrating that while infant cDCs are capable of maturation, RSV is a poor inducer of maturation [168]. Interestingly, RSV infection induced a weaker IFNα response compared to human metapneumovirus (HMPV)-infected human blood monocyte differentiated DCs (moDCs) [159]. Another study showed lower IFNα expression in RSV-infected infants and young children compared to healthy adult pDCs, an effect that is potentially mediated by immature cytosolic RIG-I in younger patients [169].
Inhibiting fatty acid synthesis, and thus mitochondrial function, in DCs infected with RSV shifts a Th2/Th17 response to a Th1 response with reduced lung pathology [170]. The KDM6 gene, which codes for a demethylase, is upregulated during RSV infection and this gene activates transcription in DCs, resulting in the production of inflammatory cytokines, chemokines, resulting in pathology [171]. This is likely important in RSV vaccine development. In mice, RSV infection of pDCs that are co-cultured with T cells has reduced the transformation of Th17 cells to FoxP3+ Tregs [172]. Similarly in mice, RSV infection of DCs impairs T cell activation by affecting synapse assembly [173]. In humans, RSV infection of cord blood-derived DCs leads to altered cytokine profiles and reduced capacity to induce T cell proliferation and functional responses [174]. Taken together, these studies show that a protective DC response is related to T cell priming, as DC immaturity results in poor IFNα/β responses during RSV, resulting in Th2 cytokine responses.

2.2. Adaptive Immune Responses

Adaptive immunity follows innate immune responses. Adaptive immunity is characterized by immunological memory referring to the ability to quickly and robustly respond to previously encountered pathogens in an antigen-specific manner. Adaptive immunity is separated into humoral or cellular immunity. The humoral immune response encompasses B cells and their products, namely antibodies. The cellular immune response includes T cell subsets. Adaptive immunity is necessary for RSV clearance, and the establishment of long-term memory is needed to protect against future infections.

2.2.1. B Cells

B cells express antibodies that act as B cell receptors. When naïve or memory B cells are activated they proliferate and differentiate into antibody-secreting effector cells or plasmablasts [175]. B cells also present antigens and secrete cytokines. B cell activation occurs in the spleen and lymph nodes. Antigens that activate B cells with the help of T cells are known as T cell-dependent antigens, while antigens that activate B cells without T cell help are known as T cell-independent antigens [176]. RSV-infected infants have increased B cells in their peripheral blood mononuclear cells (PBMCs) [177,178], and B cell-activating factor (BAFF) and ‘a proliferation-inducing ligand’ or APRIL is present in the lung epithelium that correlates with IgA and mucosal antibody responses [179]. Similarly, a study of RSV-infected mice showed increased BAFF and B cell chemoattractant CXCL13 expression in the lung [180].
Concerning antibodies expressed by B cells, anti-F protein antibodies from the adenoids of young children have been shown to have high binding affinity and greater RSV neutralization compared to peripheral blood [181]. Anti-F protein antibodies from infants infected with RSV shows somatic hypermutations (SHM) and immunoglobulin (Ig) class switching that increases with age [182]. Notably, young infants have an immature SHM function that limits the antibody repertoire. Importantly, the most common V-gene antibodies pairs against F protein are VH3-21:VL1-40 or VH3-11:VL1-40 which targeted site III, are neutralizing, and do not require SHM. Another study showed post-fusion F protein vaccination induces pre- and post-F memory B cells and the antibodies were all neutralizing to some degree [183]. Importantly, RSV can infect neonatal B cells via CX3CR1 leading to increased pathology, secretion of IL-10, and an increase in the Th2 response [13]. It was recently discovered the B cell response to RSV is linked to type I IFN receptor expression response, and RSV is known to reduce robust IFN responses potentially translating to reduced B cell function in newborns [184]. ERD mediated by FI-RSV priming is also associated with low-affinity, non-neutralizing antibody responses further demonstrating importance of antibody function and epitope in the context of immunopathology [185].

2.2.2. T Cells

T cells express a T cell receptor (TCR) on their cell surface and belong to two major subpopulations, i.e., CD8+ cytotoxic T cells (CTLs) or CD4+ helper T cells. CTLs can directly kill virally infected cells, and T helper cells assist other lymphocytes including maturation of B cells into plasma cells and memory B cells, and activation of CTL and macrophages [186]. CD4+ T cells can secrete a myriad of cytokines that contribute to B cell stimulation, antibody proliferation and class switching, effective CTL responses, and innate cell activation [187]. CD4+ T cells can differentiate into one of several subtypes each of which has different roles. For example, Th1 cells are characterized by their expression of IFNγ and Tbet producing an inflammatory response [188]. Th2 cells are characterized by expression of IL-4 and drive differentiation and antibody production by B cells [189]. Th17cells express IL-17 and have a role in gut and mucosal defenses [190]. Th9 cells express IL-9 and defend against helminths [191], and Tfh express IL-21 and IL-4 providing B cell help [192,193]. Naive T cells can expand and differentiate into memory and effector T cells after they encounter their cognate antigen in the context of a major histocompatibility complex (MHC) [194]. There are several memory T cell subtypes including central memory T cells (CD45RO+, CCR7+, and CD62L), effector memory T cells (CD45RO+, CCR7-, CD62L-), and tissue-resident memory T cells (CD103+) [195,196].

2.2.3. CTLs

Studies depleting T cells in RSV-infected mice indicated the role of CTLs in controlling RSV and concomitant immunopathology [197]. In adults, it has been shown that resident memory CTL (Trm) proliferate extensively to RSV infection compared to the circulating T cells [198]. Interestingly, this study showed that the Trm were phenotypically lacking cytotoxic markers and had reduced production of pro-inflammatory cytokines but were correlated with protection from RSV disease. Consistent with the lack of cytotoxic function, perforin-depleted mice infected with RSV cleared RSV similarly as wild-type mice likely through Fas/FasL and pro-inflammatory cytokine expression, specifically TNFα [199]. TNFα may also contribute to lung pathology [200]. In mouse studies that examined CTL cell DNA vaccination with an RSV M2 gene, it was shown that DNA vaccination led to a non-protective, pathological response mediated by IFNγ and TNFα secreting T cells [201]. This study also evaluated the passive intranasal transfer of splenic CTLs from RSV-sensitized mice to naïve mice that resulted in protection, reduced RSV lung titers, and increased the IFNγ response post-RSV challenge.
The route of RSV vaccination can affect immunity and disease pathogenesis. For example, when murine cytomegalovirus (MCMV) expressing RSV M protein was administered intranasally (i.n.) as opposed to intraperitoneally (i.p.), disease pathology was ameliorated, and a robust lung-specific T cell response followed [202]. Interestingly, BALB/c mice intravenously (i.v.) vaccinated with an H-2kd restricted RSV M2 antigen generated greater protective polyfunctional CTLs compared to intraperitoneally by approximately 4-fold [203].
RSV G protein has a single MHC-I H-2Ld restricted epitope [204]; however, RSV G protein modifies CTL responses. For example, one study examining RSV G169–198 peptide nanoparticle vaccination showed increased IFNγ producing MHC class I H-2Kd restricted M2-specific CTLs following RSV challenge, an effect linked to CX3C–CX3CR1 interaction [205]. Interestingly, including a G protein CX3C motif in an influenza vaccine was shown to enhance anti-influenza CTL responses [206]. This may be because the G protein CX3C motif affects CX3CR1+ CTL trafficking to the lung mediating a Th2-type response [207].

2.2.4. CD4+ T Cells

Th cells function to help other immune cells. For example, they help activate B cells to produce antibodies, feedback to promote CTL activity and function and provide cytokine production which stimulates and activates immune cells [194]. The immunopathology of RSV infection is typically described in the context of Th1- or Th2-type responses. The canonical cytokines associated with Th1 cells are IFNγ and IL-2, while for Th2 cells often IL-4, IL-5, and IL-13 are used. Th1 cells defend the host against intracellular parasites including viruses; Th2 cells defend against extracellular parasites such as helminths. Th1 cells produce IFNγ that stimulates macrophage activation [208]. IL-4 results in IgG1 and IgE class switching [209], IL-5 is a potent eosinophil maturation molecule [210], and IL-13 affects mucus production and is also implicated in asthma [211]. The presence of Th1 cytokines downregulates Th2 cell activation and vice versa. The FI-RSV vaccine primes for Th2-type responses and is a comparator of RSV vaccine safety [212,213]. Further, infants with severe RSV disease have a higher ratio of Th2/Th1 cytokines compared to those having mild disease [213,214,215,216].
The RSV G protein influences Th2-type immune responses, and immune dysregulation is linked to CX3C–CX3CR1 interaction and sG expression [129,217,218,219]. The RSV G protein has been shown to prime for ERD in animal models [12,217]. This feature has hindered RSV vaccine development particularly for vaccine candidates involving the G protein. Accumulating evidence has shown the protective benefits of G protein-mediated antibodies specifically regarding disease pathogenesis, thus G protein-based vaccines are being reconsidered [220]. Studies in mice have suggested that modification of the CX3C motif [221,222] and/or CCD [223] could elicit a safe and protective response and shift Th2-type responses to Th1-type or balanced responses. A G protein DNA vaccine (pVAX1/3G148–198) was shown to induce a Th1-type biased response and protect mice from the disease [224]. Additionally, mice were inoculated with a temperature-sensitive (ts) live-attenuated influenza HA/G protein CCD vaccine and this provided protection against influenza and RSV [225]. Thus, the G protein offers an opportunity to induce protective immunity.
Various adjuvants have been tested to improve RSV vaccine candidates or to induce a Th1-type or more balanced Th responses. Vaccination studies in mice using alum and pre-fusion F protein led to ERD compared to a balanced Th1/Th2 response using the adjuvant Advax-SM, a plant-based polysaccharide adjuvant [226,227]. Using an adenovirus serotype 26 (Ad26) vaccine platform and the pre-fusion F protein, a Th1-type immune response was induced in vaccinated mice which lacked detectable immunopathology [228]. However, suboptimal vaccine dosing of pre- or post-fusion F protein caused ERD despite adjuvanting with a Th1-biasing TLR-4 agonist. ERD was prevented in mice optimally vaccinated using doses that induced a robust neutralizing antibody response demonstrating the importance of not only the adjuvant but proper dosing [229]. A multiplex vaccine containing both RSV F and G proteins, i.e., SBP-FG [230] and VLP-FG [231] resulted in a protective Th1-type response. Another multiplex vaccine called G1F/M2 incorporates a G protein neutralizing epitope, and an M2 CTL epitope, that is adjuvanted with CpG2006 (a Th1-biased TLR-9 agonist) was shown to induce a Th1-type response with reduced pulmonary inflammation and disease in RSV-infected mice [232]. Additionally, a recombinant baculovirus expressing a G protein fragment and M2 (Gcf A/Bac M2) was shown to induce an efficacious and protective Th1 response [233], and a multiplex vaccine encompassing the F, N, and M2-1 proteins on a chimpanzee adenovirus vector (PanAd3-RSV) offered protective, robust Th1-type responses [234]. Interestingly, vaccination of mice by microneedle patch using FI-RSV and TLR-4 adjuvant-induced a safe and effective response having reduced Th2-type and ERD responses post-RSV challenge [235]. Interestingly, a soluble F protein vaccine induced a Th2-type response characterized by high levels of IL-4, IL-5, and IL-13 in the BAL, while VLP-F induced a Th1-type response characterized by reduced Th2 cytokines and an increase in IFNγ [236]. Vaccination strategies that target the RSV F and G proteins may be beneficial as anti-bodies specific to this major cell surface viral proteins should protect by virus neutralization, reduce the chances of vaccine escape, and prevent disease pathogenesis, but vaccine platforms, adjuvants, dosing, and their delivery should be rationally developed.

2.2.5. Tregs Cells

Tregs cells are a subclass of CD4+ T cells expressing both the CD4, CD25, and the nuclear transcription factor Forkhead box P3 (FoxP3) which determines Treg development and function [237]. FoxP3 is crucial for maintaining the suppression of the immune system. Depleting Treg cells reduces functional B cells that may lead to increased disease in the mouse model [238]. Infants with RSV display reduced numbers of activated Treg cells in the periphery and the lower Treg cytokine (IL-17A, IL-1β, and IL-23) expression that is inversely correlated with disease severity [216,239]. In RSV-infected infants, the Treg responses in PBMCs are altered having decreased IL-2 and Foxp3 compared to healthy controls [236]. Another study showed NS1 suppressed Tregs while NS2 increased levels, suggesting that these proteins function independently to modulate the Treg response [234].
An FI-RSV study in mice showed Treg reduction, which correlated with severe disease was reversed by administering CCL-17 and CCL-22 to recruit pulmonary Tregs [240]. The RSV NS1 protein has been shown to antagonize immunity in part by reducing Treg cells while increasing Th2-type/Th17 cells [241,242]. RSV NS1 protein also increases Th2-type biasing of OX40L, and mAb treatment against OX40L increases Treg cells [243]. NS1 protein has also been shown to increase p-mTOR that reduces Foxp3 expression and Treg cell differentiation [241]. Treg cell depletion has been shown to delay RSV clearance and delay CTL trafficking to the lung [244]. Moreover, depletion increased the pro-inflammatory functions of the CTLs, suggesting that the Treg response is key for mediating the CTL response to clear RSV and reduce CTL-induced immunopathology. In a study that examined RSV-infected Treg-depleted mice depletion was shown to lead to improved lung virus clearance; however, depletion was associated with increased markers of enhanced disease including severe weight loss and BAL cellular influx [245]. Interestingly, this study also showed that pulmonary Tregs cells expressed granzyme B, suggesting a role in controlling RSV infection. In another study examining Treg cell-depleted mice, it was observed that depletion was associated with highly functional Th2-type pulmonary CD4 cells and enhanced pulmonary disease [246].

2.3. Interferons (IFNs) and Inteferon Stimulating Genes (ISGs)

IFNs impede viral replication and regulate RSV infection. IFNs are categorized into three types, i.e., type I, II, and type III IFNs. Type I IFNs include IFNα and IFNβ with at least 14 distinct isoforms of IFN-α [247]. pDCs are chief producers of type I IFNs that also include IFNs ε, κ, τ, μ, ζ, ω, and υ; however, these are not as well described. Type II IFN mediate T cell responses and activate macrophages [248]. IFNγ downregulates Th2-type cells and upregulates Th1-type cell responses. An infant cohort study showed RSV disease and wheeze associated with reduced IFNγ and a Th2-type cytokine propensity [249]. IFNγ is a canonical Th1-type cytokine. Type III IFNs include four isoforms of IFNλ (i.e., IFN-λ1, -λ2, -λ3, and -λ4) and are similar in function to type I IFNs [250].
IFNs bind their receptor and induce differential gene expression of interferon-stimulated genes (ISGs), resulting in an amplified antiviral response. Type I IFN binds to the ubiquitous heterodimeric membrane-bound receptor IFN-α/β-R1/2 (IFNAR) [251]. Binding results in activation of NF-κΒ, cAMP-response element-binding protein (CREB), and signal transducer and activator of transcription proteins (STAT). IFNγ binds to the ubiquitous receptor IFN-γR-I and -II. This binding results in the activation of the JAK-STAT pathway [252]. Type III IFNs bind to a complex of IL-10Rβ/2 and IFNLR1 that are found on subsets of epithelial cells and neutrophils, and this binding event cascades to activation of STAT and CREB [253]. These IFNs have distinct and overlapping roles in the context of RSV.

2.3.1. Type I IFN

The SOCS family of proteins are negative-feedback inhibitors of signaling induced by cytokines that act via the JAK/STAT pathway [254,255]. SOCS expression is induced by TLR-4. During RSV infection, early upregulation of SOCS3 results in downregulation of IFNα expression that is linked to RSV F and G protein induction of TLR-4 [256]. IFN signaling is tightly controlled by SOCS expression. Studies have shown that G protein expression modulates SOCS3, and reduces IFNβ levels likely through reduced expression of the ubiquitin-like protein ISG-15 [257]. Other studies haves shown a reduction in IFNβ is due to the blocking of membrane-bound TLR-3 and TLR-4 via sG antagonism [24]. Importantly, anti-G protein mAb that blocks CX3C–CX3CR1 interaction results in increased type I IFN expression, which correlated with reduced pathology [220]. Taken together, these studies show that RSV surface proteins can modulate type I IFN responses through membrane signaling.
NS1 and NS2 proteins disrupt innate immunity and type I IFN levels. Several studies have shown the RSV NS1/NS2 are involved in immune dysregulation during virus replication. It has been shown that IFN levels remain surprisingly low during RSV infection despite viral replication [258]. Type I IFN expression is substantially increased when cells are infected with an RSV deletion mutant lacking NS 1 and NS2 (RSVΔNS1/2) [259]. Several RSV proteins have been shown to suppress both the production and signaling of type I and III IFNs by counteracting host innate signaling proteins, and several ISGs are affected by NS proteins [258,260]. For these reasons, several attenuated vaccine candidates contain deleted NS genes. Infection of primary bronchial epithelial cells with RSVΔNS1/2 was shown to have reduced replication efficiency, suggesting a role for NS1/2 in replication [10]. Interestingly, neonatal mice treated with IFNα before the RSV challenge had increased B cells and IgA levels that correlated with an improved disease outcome [261], and infants with severe RSV had lower type I IFN gene expression compared to mild disease [262]. Taken together, type I IFNs may be protective during RSV infection; however, proteins including NS1, NS2, G, and F act to reduce type I responses. Intriguingly, NS1 has been shown to increase miR-29a levels that are known to decrease type I IFNAR expression, suggesting a means by which NS may act to quell the antiviral response [263].

2.3.2. Type III IFNs

Type I IFN receptors (IFNAR) are ubiquitous, whereas type III receptors are expressed principally on epithelial cells [264]. Type III IFN is thought to mediates a less efficient antiviral response compared to the type I IFN [265]. It is thought that the host reserves the more potent type I IFN response to be used if the local type III IFN response is insufficient [266]. Infants infected with RSV have increased IFNλ levels in nasal specimens that correlate with viral titers [267] and increased respiratory rates [268]. In primary nasal epithelial cells, RSV infection is associated with increases in IFNλ but not type I IFN which is facilitated through the RIG-I pathway [269]. Increasing IFNλ through blockade of prostaglandin D2 (PGD2): PGD2R2 has been shown to improve disease outcomes in neonatal mice and assist in lung viral clearance [270]. Recent studies suggest UV-inactivated RSV, and RSV F protein mediate an IFNλ blocking mechanism via epidermal growth factor receptor (EGFR) agonism [271]. F protein agonism of EGFR is linked to abundant mucin production [47]. Interestingly, RSV replication requires Rab5a involvement in intracellular membrane trafficking, and Rab5a downregulates the IFNλ response in the airway [272]. The RSV G protein modifies IFNλ responses, and infection of Calu-3 cells with a modified RSV having a G protein CX3C region ablation increased IFNλ expression indicating that CX3C–CX3CR1 interaction signals to modify the type III IFN response [273]. Additionally, the RSV G protein upregulates miRNA let-7f [274], and it is believed that this miRNA functions to reduce IFNλ translation. Thus, several RSV proteins may affect type III IFN responses to facilitate viral replication.

3. Animal Models

Animal models are used as a surrogate to better understand human responses to RSV infection, but few models are appropriate, as most are semi-permissive to RSV replication except chimpanzees [275,276,277]. Some studies have examined related pneumoviruses, e.g., pneumonia virus of mice (PVM) [278] and bovine RSV (BRSV) [279] as alternate virus models to RSV as these viruses have permissive animal hosts.

3.1. Mice

The most widely used species used in RSV studies is the inbred mouse; however, permissiveness to infection varies by strain by up to 100-fold [280]. Mouse strains have varying sensitivity and a vast majority of studies utilize the BALB/c mouse which are biased towards a Th2-type response and are H2d MHC restricted [281]. Interestingly, older mice have shown increased susceptibility to RSV infection [282]. Correlates of immunity and disease pathogenesis in mice can be assessed; however, weight loss is not typically observed in human disease, and does not always correlate with lung damage, lung function, or BAL infiltrates [283,284], thus, mice do not fully recapitulate human disease. Humanized mice have been investigated and in one study, human immune system (HIS) mice were generated from the NOD SCID gamma (NSG) mouse which possessed functional T cells and B cells and challenged with RSV line 19F [285]. Compared to NSG mice, HIS mice lost significant weight and had significantly more lung histopathology with increased BAL cell infiltrates, and increased IL-1β and MUC5ac, and reduced CCL3. Humanized mice appear to be a useful model but are laborious to develop and too costly. An alternative may be to challenge neonatal mice with the rA2-19F that closely mimics RSV disease symptoms in human infants [286]. rA2-19F is a chimeric RSV in which the F protein from the A2 strain was replaced with the F protein from the Line 19 strain (rA2-19F) and infection of adult mice with rA2-19F has been shown to mediate higher lung viral loads compared to either of its parent strains (A2 or line 19) [287]. This is likely a result of five mutations within the F protein which are involved with increased susceptibility to mouse lung, a reduction in type I IFN responses, and increased mucosal responses (including IL-13) [287]. rA2-19F infection has been shown to cause lung pathology, mucin secretion, airway hypersensitivity, and Th2 responses which are diminished via prophylaxis with anti-G protein mAb (131-2G) [287,288].

3.2. Cotton Rat

The cotton rat (CR) (Sigmodon hispidus) model has been used to justify the advancement of palivizumab in human clinical trials [289,290]. CRs are sensitive to RSV [291] and contain a functional Mx gene set which is important for mounting the innate antiviral response in humans which is absent in many inbred mouse strains [292]. However, the features of the immune response vary as CRs age [293,294]. CRs demonstrate a sex-dependent response to RSV with females showing increased airway hyperresponsiveness and pulmonary edema compared to male littermates [295]. Interestingly, the BAL fluid of CRs is primarily comprised of eosinophils, and challenge with RSV does not increase these cell numbers [296]. Moreover, CRs can be used in maternal antibody studies, as RSV antibodies transfer from dames to pups via the placenta and breast milk [297,298,299,300]. Further, CX3CR1 is a receptor used by RSV for infection of CRs [301]. While the cotton rat model has been used to evaluate drugs and vaccines, limitations of this model include cost, procurement (i.e., limited suppliers), and lack of robust tools to evaluate immune responses.

3.3. Non-Human Primates (NHPs)

NHPs are used as models because of their close evolutionary lineage to humans. RSV was first identified in chimpanzees and consequently named chimpanzee coryza agent (CCA) [302]. Apart from chimpanzees, other NHP species are only semi-permissive to RSV infection illustrating the specificity of RSV to humans. Nevertheless, NHPs have been used in several preclinical studies of drugs and vaccines. For example, the RSV fusion inhibitor, TMC353121, was tested in African green monkeys (AGMs) followed by RSV challenge which was shown to reduce lung RSV titers [303]. The RSV vaccine candidate, DS-Cav1-I53-50, is a stabilized pre-F protein nanoparticle vaccine that was evaluated in rhesus macaques and found to be highly immunogenic [304]. Another RSV nanoparticle vaccine candidate, i.e., pre-F-NP, a modified pre-F stabilized protein fused to ferritin nanoparticles also showed increased immunogenicity compared to the DS-Cav1 in Cynomolgus macaques [305]. Trivax, a candidate RSV vaccine that includes three RSV F protein neutralizing epitopes (i.e., 0, II, and IV) was evaluated in naïve and RSV-experienced AGMs [306]. Trivax was shown to be immunogenic in naive and AGMs sensitized to the RSV F protein. In addition, vectored virus vaccine candidates have been evaluated in Cynomolgus macaques, and importantly delivery via i.n. route resulted in broad and robust anti-F IgA responses [307]. A codon-deoptimized live-attenuated vaccine delivered to AGMs generated robust immunity and displayed minimal viral shedding, a point of interest in live-attenuated vaccines [308]. It is of note that many of the NHP studies evaluate responses against a common target, i.e., pre-F, yet differ in platform, adjuvant, and/or delivery.
The use of small animal models such as BALB/c mice aids immunological research and disease pathogenesis studies especially with regard to markers of ERD (i.e., Th2-type cytokines and chemokines), as mice have robust commercially available tools such as antibodies. However, mice are only semi-permissive to RSV and do not recapitulate human disease. CRs are useful for studies determining RSV transmission, including dam-to-pup transmission, as RSV is permissive in this model. However, similar to the mouse model, RSV disease is not recapitulated in the CR and moreover there are limited commercial tools available for studying CRs. NHPs, while requiring more resources than small animals, provide a good model for late preclinical vaccine and/or therapeutic evaluations, particularly when evaluating safety given their close lineage to humans. Thus, while these three animal models individually serve a niche in RSV research, no single animal model can fully serve as a surrogate for human RSV disease. The lack of an all-around animal model has hindered RSV vaccine and therapeutic development.

4. Correlates of Protection

A major hurdle in establishing countermeasures to RSV is limited correlates of protection (CoPs) needed to evaluate vaccines and therapeutics. CoPs are determined when there is a direct cause–effect relationship between a factor such as vaccination or treatment and an endpoint. Typically, virus-specific neutralizing antibody titers known to protect against RSV disease are used. Several studies have examined RSV-specific polyclonal or monoclonal antibodies responses associated with neutralizing activities as CoPs [80,309,310,311]. For example, antibody neutralization has been demonstrated by analyses of maternal sera [309,312], cord blood [313], breast milk [314], infant sera [309], and by passive antibody transfer studies [315]. CoPs for RSV remain under investigation and include specific immune responses, viral load, duration of infection, and microRNA responses [316]. One means of evaluating a CoP is to determine the cytokine/chemokine profiles of lung leukocytes and the Th cell subpopulations after sensitization [216]. In general, lower levels of Th2-type cytokines (e.g., IL-4) have been suggested to be associated with reduce disease [317], while MUC5AC expression has been correlated with more severe disease [318] and CXCL4 expression correlated with reduced disease severity [319]. Additionally, MIP-1α a potent eosinophil chemoattractant, has been positively correlated with severe disease in hospitalized infants, and interestingly RSV G protein may inhibit early MIP-1α expression in mice [126].
Along with Th1- or Th2-type polarization, there is macrophage polarization, where macrophages adopt different functional programs in response to the signals from their microenvironment. This ability is linked to their roles being effector cells in the innate immune response, and also removal of cellular debris, embryonic development and tissue repair [320]. Macrophages can be polarized into classic (M1) or alternatively activated (M2) groups based on in vitro studies, in which cultured macrophages were treated with molecules that stimulated their phenotype switching to a particular state [321]. M1 macrophages have been described as pro-inflammatory and important in directing host-defense against pathogens including phagocytosis and secretion of pro-inflammatory cytokines and microbicidal molecules. M2 macrophages have been described in the regulation of the resolution of inflammation and the repair of damaged tissues. M1 macrophages are induced by Th1-type cytokines and TLR-4 interaction, while M2 macrophages are induced by Th2-type cytokines [322]. The role of macrophage polarization and activation during RSV infection is significant. In the asthmatic murine model, TLR-4 signaling linked to by RSV induced M1 activation and reduction in Th2-type cytokines [323]. Other studies have shown an opposite function where M2 macrophages in the lung of RSV-infected mice contribution to Th2-type cytokine production mediated by STAT6 [324,325] contributing to a protective role in RSV infection [326]. For example, increasing M2 macrophage polarization by IL-4/anti-IL-4 immune complexes or PPARγ agonism has been shown to result in decreased lung pathology during RSV infection in mice [327]. Additionally, inhibition of cyclooxygenase-2 (COX-2) in RSV-infected CRs has been shown to increase M2 responses and result in decreased lung pathology [328]. Thus, while cytokines are relevant in RSV disease or protection, macrophage polarization may serve as relevant CoPs.
Another consideration when establishing CoPs is viral load and duration of infection. Traditionally, reduced viral load and a shorter duration of infection are indicative of moderated disease. Sterilizing immunity following RSV infection might likely prevent RSV disease; however, sterilizing immunity is not likely achievable given the many features of RSV to avoid immunity [12,329,330]. Reduced viral loads alone do not correlate with protection from disease as higher viral loads early after infection are correlated with protective innate immunity in infants [331] and elderly patients [332] hospitalized with RSV disease. In addition, antibody levels to RSV have been used to determine CoPs; however, the type of antibody induced and to what antigen (F or G protein) are important. In one study, infants with high levels of maternal antibodies in cord blood were protected from RSV during the first months of life [309], yet the type of antibodies elicited dictate the durability of protection as neutralizing maternal antibodies wane quickly with particular types of antibodies having varying half-lives [333]. Importantly, neutralizing antibody assays generally do not detect non-neutralizing, Fc-mediated responses that may also be protective. Mechanisms by which these responses are protective include antibody-dependent cytotoxicity (ADCC), antibody-dependent cellular phagocytosis (ADCP), and complement-dependent cytotoxicity [334]. However, the antigen–antibody complex associated by Fc and complement binding has been shown to result in ERD (and potential Th2-type biasing) in FI-RSV primed mice [335].
Original antigenic sin (OAS) is the tendency of the immune response to favor responses to an original antigen over a related or new antigen. OAS favors immunologic memory to the original antigen rather than that to new antigens leading to hierarchal responses. OAS can affect the memory response to the original infecting strain or vaccine antigen, accordingly OAS is important in vaccine development because vaccine efficacy could change upon the appearance of a new strain. Thus, the first RSV infection in the infant imprints immunity and may determine the outcome of disease particularly after reinfections. How OAS affects the original response to the infecting strain and subsequent infections needs to be better understood, particularly in RSV vaccine development. Importantly, OAS is an important caveat when evaluating vaccines and therapeutics in naïve animal models [336]. Thus, CoPs determined in the naïve animal models may not align with the markers of OAS-linked immune responses.
MicroRNAs (miRs) have been investigated as biomarkers that may serve as CoPs [337]. In one study, miR profiles expressed in mice vaccinated with different RSV vaccine candidates showed that miR regulation correlated with disease or protection [338]. The G protein CCD was shown to specifically downregulate let-7f and miR-24 [273]. Consistent with miR regulation, HEp-2 cells infected with RSV have downregulated levels of let-7f expression [339] and downregulated levels of miR-140-5p expression which was correlated with disease severity and linked to enhanced inflammation and decreased TLR-4 expression [340]. It has been shown that miR profiles may behave as biomarkers to determine disease severity as downregulation of miR-34b/c-5p has been associated with mucus secretion through reduced regulation of MUC5AC [341].
Importantly, RSV typically infects the very young and old but can infect others including the immune compromised [342] thus standardizing CoPs is complex as there is a wide range affecting outcomes. For example, infants have immunologic and anatomical immaturity, and may not be sensitized to RSV infection. Additionally, their lungs vary in architecture and disease susceptibility [343], and the infant immune system is biased towards a Th2-type response to prevent inflammatory damage to the lungs [344]. As described, Th2-type responses including increased IL-4Rα on cDCs is associated with increased disease in this population. Moreover, infants with genetic mutations for TLR-4, IL-4, IL-8, IL-13, and IL-4Rα are at greater risk for severe disease [102,345,346]. Hospitalization is associated with younger age so by age 5 into adulthood, RSV infections are generally mild, suggesting that the maturation of the immune system, memory T and B cells, and improved antiviral inflammatory responses (i.e., Th1-type) are considered protective and less pathogenic. In contrast, the elderly have a waning immunity and a reduced thymus size which can modify T cell responses, and often have comorbidities or lifestyle choices (i.e., cigarette smoking) that can complicate RSV disease [347]. Interestingly, while RSV disease presents differently in the elderly compared to the infant population [348], elderly immunopathology is similar to infants in that the immunosenescence results in increased Th2-type and Treg and reduced Th1-type and Tmem subset responses [349].
Despite the challenges with elucidating CoPs, Th2-type responses, IFNλ, and eosinophilia and/or neutrophilia are generally associated with immunopathogenesis while balanced and/or Th1-type responses, type-I IFNs and pDCs are considered protective.

5. Discussion and Conclusions

RSV causes substantial respiratory tract disease burden particularly in infants and the elderly. Unfortunately, there is no approved vaccine or post-exposure therapeutic available and prophylactic intervention is limited. RSV infection causes immune dysregulation and disease pathogenesis. B cells producing antibodies may neutralize RSV replication, and T cells such as CTL, Th, Treg, and Tmems are important for the resolution of infection. CTLs may also induce pathology, and CD4 Th1 or Th2 cells contribute to disease. Specifically, Th2-type cytokines are thought to be less protective and more pathogenic so a better understanding of the role of cytokines and chemokines, especially IFNs, is essential. While type I IFNs may protect from disease, RSV has various mechanisms to modify the type I response. RSV proteins can reduce antiviral IFNs, and IFNλ levels are key for controlling lung titers and disease severity. Finally, the lack of a reliable, reproducible, and cost-effective animal model has complicated RSV research and disease intervention strategies. Our knowledge of correlates of protection and biomarkers of immunity and disease are currently incomplete; however, new biomarker understanding should deepen our understanding.
Understanding host–pathogen interactions is vital to developing countermeasures to address unmet needs and disease burden. As viral load is not predictably correlated with disease, vaccine and therapeutic countermeasures should consider preventing both virus replication and the proinflammatory response to disease. Specifically, multivalent approaches that target both the G and F proteins could improve disease outcomes. Anti-F protein antibodies have shown clinical efficacy, e.g., palivizumab and nirsevimab, while G protein vaccine strategies have shown promise beyond proof of concept [350,351]. Vaccine candidates encompassing various epitopes such as live-attenuated and codon-deoptimized platforms may be advantageous [221,352]. Host-targeting drugs [353] or immunotherapies, e.g., IFNα [354] that interfere with the virus life cycle and/or improve the immune response may also potentially reduce disease burden with reduced risk of viral escape. Reverse vaccinology and rational structure-based antigen design may yield previously unobtainable responses by designing immunogens to induce potent and protective antibody responses.

Author Contributions

H.C.B. and R.A.T. wrote and edited the drafts. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Tripp laboratory, and supported by the Georgia Research Alliance (GRA).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

Authors have no conflict of interests to declare.

References

  1. Collins, P.L.; Fearns, R.; Graham, B.S. Respiratory Syncytial Virus: Virology, Reverse Genetics, and Pathogenesis of Disease. Curr. Top. Microbiol. Immunol. 2013, 372, 3–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Fearns, R.; Deval, J. New antiviral approaches for respiratory syncytial virus and other mononegaviruses: Inhibiting the RNA polymerase. Antivir. Res. 2016, 134, 63–76. [Google Scholar] [CrossRef] [Green Version]
  3. Ke, Z.; Dillard, R.S.; Chirkova, T.; Leon, F.; Stobart, C.C.; Hampton, C.M.; Strauss, J.D.; Rajan, D.; Rostad, C.A.; Taylor, J.V.; et al. The Morphology and Assembly of Respiratory Syncytial Virus Revealed by Cryo-Electron Tomography. Viruses 2018, 10, 446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Anderson, L.J.; Hierholzer, J.C.; Tsou, C.; Hendry, R.M.; Fernie, B.F.; Stone, Y.; McIntosh, K. Antigenic Characterization of Respiratory Syncytial Virus Strains with Monoclonal Antibodies. J. Infect. Dis. 1985, 151, 626–633. [Google Scholar] [CrossRef] [PubMed]
  5. Choi, Y.; Mason, C.S.; Jones, L.P.; Crabtree, J.; Jorquera, P.A.; Tripp, R.A. Antibodies to the Central Conserved Region of Respiratory Syncytial Virus (RSV) G Protein Block RSV G Protein CX3C-CX3CR1 Binding and Cross-Neutralize RSV A and B Strains. Viral Immunol. 2012, 25, 193–203. [Google Scholar] [CrossRef] [Green Version]
  6. Johnson, P.R.; Spriggs, M.K.; Olmsted, R.A.; Collins, P.L. The G glycoprotein of human respiratory syncytial viruses of subgroups A and B: Extensive sequence divergence between antigenically related proteins. Proc. Natl. Acad. Sci. USA 1987, 84, 5625–5629. [Google Scholar] [CrossRef] [Green Version]
  7. Borchers, A.T.; Chang, C.; Gershwin, M.E.; Gershwin, L.J. Respiratory Syncytial Virus—A Comprehensive Review. Clin. Rev. Allergy Immunol. 2013, 45, 331–379. [Google Scholar] [CrossRef]
  8. Piedimonte, G.; Perez, M.K. Respiratory Syncytial Virus Infection and Bronchiolitis. Pediatr. Rev. 2014, 35, 519–530. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, L.; Peeples, M.E.; Boucher, R.C.; Collins, P.L.; Pickles, R.J. Respiratory Syncytial Virus Infection of Human Airway Epithelial Cells Is Polarized, Specific to Ciliated Cells, and without Obvious Cytopathology. J. Virol. 2002, 76, 5654–5666. [Google Scholar] [CrossRef] [Green Version]
  10. Villenave, R.; Broadbent, L.; Douglas, I.; Lyons, J.D.; Coyle, P.V.; Teng, M.N.; Tripp, R.A.; Heaney, L.G.; Shields, M.D.; Power, U.F. Induction and Antagonism of Antiviral Responses in Respiratory Syncytial Virus-Infected Pediatric Airway Epithelium. J. Virol. 2015, 89, 12309–12318. [Google Scholar] [CrossRef] [Green Version]
  11. van Benten, I.J.; van Drunen, C.M.; Koopman, L.P.; KleinJan, A.; van Middelkoop, B.C.; de Waal, L.; Osterhaus, A.; Neijens, H.J.; Fokkens, W.J. RSV-induced bronchiolitis but not upper respiratory tract infection is accompanied by an increased nasal IL-18 response. J. Med. Virol. 2003, 71, 290–297. [Google Scholar] [CrossRef]
  12. Jorquera, P.A.; Anderson, L.; Tripp, R.A. Understanding respiratory syncytial virus (RSV) vaccine development and aspects of disease pathogenesis. Expert Rev. Vaccines 2015, 15, 173–187. [Google Scholar] [CrossRef]
  13. Zhivaki, D.; Lemoine, S.; Lim, A.; Morva, A.; Vidalain, P.-O.; Schandene, L.; Casartelli, N.; Rameix-Welti, M.-A.; Hervé, P.-L.; Dériaud, E.; et al. Respiratory Syncytial Virus Infects Regulatory B Cells in Human Neonates via Chemokine Receptor CX3CR1 and Promotes Lung Disease Severity. Immunity 2017, 46, 301–314. [Google Scholar] [CrossRef] [Green Version]
  14. Li, X.-Q.; Fu, Z.F.; Alvarez, R.; Henderson, C.; Tripp, R.A. Respiratory Syncytial Virus (RSV) Infects Neuronal Cells and Processes That Innervate the Lung by a Process Involving RSV G Protein. J. Virol. 2006, 80, 537–540. [Google Scholar] [CrossRef] [Green Version]
  15. Panuska, J.R.; Cirino, N.M.; Midulla, F.; Despot, J.E.; McFadden, E.R.; Huang, Y.T. Productive infection of isolated human alveolar macrophages by respiratory syncytial virus. J. Clin. Investig. 1990, 86, 113–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Johnson, T.R.; Johnson, C.N.; Corbett, K.S.; Edwards, G.C.; Graham, B.S. Primary Human mDC1, mDC2, and pDC Dendritic Cells Are Differentially Infected and Activated by Respiratory Syncytial Virus. PLoS ONE 2011, 6, e16458. [Google Scholar] [CrossRef] [Green Version]
  17. Halfhide, C.P.; Flanagan, B.F.; Brearey, S.P.; Hunt, J.A.; Fonceca, A.M.; McNamara, P.S.; Howarth, D.; Edwards, S.; Smyth, R.L. Respiratory Syncytial Virus Binds and Undergoes Transcription in Neutrophils from the Blood and Airways of Infants with Severe Bronchiolitis. J. Infect. Dis. 2011, 204, 451–458. [Google Scholar] [CrossRef] [PubMed]
  18. Al-Afif, A.; Alyazidi, R.; Oldford, S.A.; Huang, Y.Y.; King, C.A.; Marr, N.; Haidl, I.D.; Anderson, R.; Marshall, J.S. Respiratory syncytial virus infection of primary human mast cells induces the selective production of type I interferons, CXCL10, and CCL4. J. Allergy Clin. Immunol. 2015, 136, 1346–1354.e1. [Google Scholar] [CrossRef] [Green Version]
  19. Raiden, S.; Sananez, I.; Remes-Lenicov, F.; Pandolfi, J.; Romero, C.; De Lillo, L.; Ceballos, A.; Geffner, J.; Arruvito, L. Respiratory Syncytial Virus (RSV) Infects CD4+ T Cells: Frequency of Circulating CD4+ RSV+ T Cells as a Marker of Disease Severity in Young Children. J. Infect. Dis. 2017, 215, 1049–1058. [Google Scholar] [CrossRef] [Green Version]
  20. Tripp, R.A.; Jones, L.P.; Haynes, L.M.; Zheng, H.; Murphy, P.M.; Anderson, L.J. CX3C chemokine mimicry by respiratory syncytial virus G glycoprotein. Nat. Immunol. 2001, 2, 732–738. [Google Scholar] [CrossRef] [PubMed]
  21. Teng, M.N.; Whitehead, S.S.; Collins, P.L. Contribution of the Respiratory Syncytial Virus G Glycoprotein and Its Secreted and Membrane-Bound Forms to Virus Replication in Vitro and in Vivo. Virology 2001, 289, 283–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Bukreyev, A.; Yang, L.; Fricke, J.; Cheng, L.; Ward, J.M.; Murphy, B.R.; Collins, P.L. The Secreted Form of Respiratory Syncytial Virus G Glycoprotein Helps the Virus Evade Antibody-Mediated Restriction of Replication by Acting as an Antigen Decoy and through Effects on Fc Receptor-Bearing Leukocytes. J. Virol. 2008, 82, 12191–12204. [Google Scholar] [CrossRef] [Green Version]
  23. Bukreyev, A.; Yang, L.; Collins, P.L. The Secreted G Protein of Human Respiratory Syncytial Virus Antagonizes Antibody-Mediated Restriction of Replication Involving Macrophages and Complement. J. Virol. 2012, 86, 10880–10884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Shingai, M.; Azuma, M.; Ebihara, T.; Sasai, M.; Funami, K.; Ayata, M.; Ogura, H.; Tsutsumi, H.; Matsumoto, M.; Seya, T. Soluble G protein of respiratory syncytial virus inhibits Toll-like receptor 3/4-mediated IFN-beta induction. Int. Immunol. 2008, 20, 1169–1180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Feldman, S.A.; Hendry, R.M.; Beeler, J.A. Identification of a Linear Heparin Binding Domain for Human Respiratory Syncytial Virus Attachment Glycoprotein G. J. Virol. 1999, 73, 6610–6617. [Google Scholar] [CrossRef] [Green Version]
  26. Hallak, L.K.; Spillmann, D.; Collins, P.L.; Peeples, M.E. Glycosaminoglycan Sulfation Requirements for Respiratory Syncytial Virus Infection. J. Virol. 2000, 74, 10508–10513. [Google Scholar] [CrossRef] [Green Version]
  27. Shields, B.; Mills, J.; Ghildyal, R.; Gooley, P.; Meanger, J. Multiple heparin binding domains of respiratory syncytial virus G mediate binding to mammalian cells. Arch. Virol. 2003, 148, 1987–2003. [Google Scholar] [CrossRef]
  28. Chirkova, T.; Lin, S.; Oomens, A.G.P.; Gaston, K.A.; Boyoglu-Barnum, S.; Meng, J.; Stobart, C.; Cotton, C.U.; Hartert, T.V.; Moore, M.L.; et al. CX3CR1 is an important surface molecule for respiratory syncytial virus infection in human airway epithelial cells. J. Gen. Virol. 2015, 96, 2543–2556. [Google Scholar] [CrossRef]
  29. Johnson, S.M.; McNally, B.A.; Ioannidis, I.; Flano, E.; Teng, M.N.; Oomens, A.G.; Walsh, E.E.; Peeples, M.E. Respiratory Syncytial Virus Uses CX3CR1 as a Receptor on Primary Human Airway Epithelial Cultures. PLoS Pathog. 2015, 11, e1005318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Panek, C.A.; Ramos, M.V.; Mejias, M.P.; Abrey-Recalde, M.J.; Fernandez-Brando, R.J.; Gori, M.S.; Salamone, G.V.; Palermo, M.S. Differential expression of the fractalkine chemokine receptor (CX3CR1) in human monocytes during differentiation. Cell. Mol. Immunol. 2014, 12, 669–680. [Google Scholar] [CrossRef] [Green Version]
  31. Nishimura, M.; Umehara, H.; Nakayama, T.; Yoneda, O.; Hieshima, K.; Kakizaki, M.; Dohmae, N.; Yoshie, O.; Imai, T. Dual Functions of Fractalkine/CX3C Ligand 1 in Trafficking of Perforin+/Granzyme B+Cytotoxic Effector Lymphocytes That Are Defined by CX3CR1 Expression. J. Immunol. 2002, 168, 6173–6180. [Google Scholar] [CrossRef] [Green Version]
  32. Umehara, H.; Goda, S.; Imai, T.; Nagano, Y.; Minami, Y.; Tanaka, Y.; Okazaki, T.; Bloom, E.T.; Domae, N. Fractalkine, a CX 3 C-chemokine, functions predominantly as an adhesion molecule in monocytic cell line THP-1. Immunol. Cell Biol. 2001, 79, 298–302. [Google Scholar] [CrossRef]
  33. Combadiere, C.; Salzwedel, K.; Smith, E.D.; Tiffany, H.L.; Berger, E.A.; Murphy, P.M. Identification of CX3CR1. A chemotactic receptor for the human CX3C chemokine fractalkine and a fusion coreceptor for HIV-1. J. Biol. Chem. 1998, 273, 23799–23804. [Google Scholar] [CrossRef] [Green Version]
  34. Bazan, J.F.; Bacon, K.B.; Hardiman, G.; Wang, W.; Soo, K.; Rossi, D.; Greaves, D.R.; Zlotnik, A.; Schall, T.J. A new class of membrane-bound chemokine with a CX3C motif. Nature 1997, 385, 640–644. [Google Scholar] [CrossRef]
  35. Kim, K.-W.; Vallon-Eberhard, A.; Zigmond, E.; Farache, J.; Shezen, E.; Shakhar, G.; Ludwig, A.; Lira, S.A.; Jung, S. In vivo structure/function and expression analysis of the CX3C chemokine fractalkine. Blood 2011, 118, e156–e167. [Google Scholar] [CrossRef] [Green Version]
  36. Jones, H.G.; Ritschel, T.; Pascual, G.; Brakenhoff, J.P.J.; Keogh, E.; Furmanova-Hollenstein, P.; Lanckacker, E.; Wadia, J.S.; Gilman, M.S.A.; Williamson, R.A.; et al. Structural basis for recognition of the central conserved region of RSV G by neutralizing human antibodies. PLoS Pathog. 2018, 14, e1006935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Malhotra, R.; Ward, M.; Bright, H.; Priest, R.; Foster, M.R.; Hurle, M.; Blair, E.; Bird, M. Isolation and characterisation of potential respiratory syncytial virus receptor(s) on epithelial cells. Microbes Infect. 2003, 5, 123–133. [Google Scholar] [CrossRef]
  38. Griffiths, C.; Drews, S.J.; Marchant, D.J. Respiratory Syncytial Virus: Infection, Detection, and New Options for Prevention and Treatment. Clin. Microbiol. Rev. 2017, 30, 277–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Jones, H.G.; Battles, M.B.; Lin, C.-C.; Bianchi, S.; Corti, D.; McLellan, J.S. Alternative conformations of a major antigenic site on RSV F. PLoS Pathog. 2019, 15, e1007944. [Google Scholar] [CrossRef] [Green Version]
  40. McLellan, J.; Ray, W.C.; Peeples, M.E. Structure and Function of Respiratory Syncytial Virus Surface Glycoproteins. Curr. Top. Microbiol. Immunol. 2013, 372, 83–104. [Google Scholar] [CrossRef] [Green Version]
  41. Murineddu, G.; Murruzzu, C.; Pinna, G. An Overview on Different Classes of Viral Entry and Respiratory Syncitial Virus (RSV) Fusion Inhibitors. Curr. Med. Chem. 2010, 17, 1067–1091. [Google Scholar] [CrossRef]
  42. Yuan, X.; Hu, T.; He, H.; Qiu, H.; Wu, X.; Chen, J.; Wang, M.; Chen, C.; Huang, S. Respiratory syncytial virus prolifically infects N2a neuronal cells, leading to TLR4 and nucleolin protein modulations and RSV F protein co-localization with TLR4 and nucleolin. J. Biomed. Sci. 2018, 25, 13. [Google Scholar] [CrossRef] [Green Version]
  43. Mastrangelo, P.; Chin, A.; Tan, S.; Jeon, A.; Ackerley, C.; Siu, K.; Lee, J.; Hegele, R. Identification of RSV Fusion Protein Interaction Domains on the Virus Receptor, Nucleolin. Viruses 2021, 13, 261. [Google Scholar] [CrossRef] [PubMed]
  44. Anderson, C.S.; Chirkova, T.; Slaunwhite, C.G.; Qiu, X.; Walsh, E.E.; Anderson, L.J.; Mariani, T.J. CX3CR1 Engagement by Respiratory Syncytial Virus Leads to Induction of Nucleolin and Dysregulation of Cilia-related Genes. J. Virol. 2021, 95, e00095-21. [Google Scholar] [CrossRef] [PubMed]
  45. Marr, N.; Turvey, S.E. Role of human TLR4 in respiratory syncytial virus-induced NF-κB activation, viral entry and replication. Innate Immun. 2012, 18, 856–865. [Google Scholar] [CrossRef] [PubMed]
  46. Kurt-Jones, E.A.; Popova, L.; Kwinn, L.A.; Haynes, L.M.; Jones, L.P.; Tripp, R.; Walsh, E.E.; Freeman, M.W.; Golenbock, D.T.; Anderson, L.J.; et al. Pattern recognition receptors TLR4 and CD14 mediate response to respiratory syncytial virus. Nat. Immunol. 2000, 1, 398–401. [Google Scholar] [CrossRef]
  47. Currier, M.G.; Lee, S.; Stobart, C.; Hotard, A.L.; Villenave, R.; Meng, J.; Pretto-Kernahan, C.; Shields, M.D.; Nguyen, M.T.; Todd, S.O.; et al. EGFR Interacts with the Fusion Protein of Respiratory Syncytial Virus Strain 2-20 and Mediates Infection and Mucin Expression. PLoS Pathog. 2016, 12, e1005622. [Google Scholar] [CrossRef] [Green Version]
  48. Behera, A.K.; Matsuse, H.; Kumar, M.; Kong, X.; Lockey, R.F.; Mohapatra, S.S. Blocking Intercellular Adhesion Molecule-1 on Human Epithelial Cells Decreases Respiratory Syncytial Virus Infection. Biochem. Biophys. Res. Commun. 2001, 280, 188–195. [Google Scholar] [CrossRef] [PubMed]
  49. Griffiths, C.D.; Bilawchuk, L.M.; McDonough, J.E.; Jamieson, K.C.; Elawar, F.; Cen, Y.; Duan, W.; Lin, C.; Song, H.; Casanova, J.-L.; et al. IGF1R is an entry receptor for respiratory syncytial virus. Nature 2020, 583, 615–619. [Google Scholar] [CrossRef] [PubMed]
  50. Ghildyal, R.; Hartley, C.; Varrasso, A.; Meanger, J.; Voelker, D.R.; Anders, E.M.; Mills, J. Surfactant Protein A Binds to the Fusion Glycoprotein of Respiratory Syncytial Virus and Neutralizes Virion Infectivity. J. Infect. Dis. 1999, 180, 2009–2013. [Google Scholar] [CrossRef] [PubMed]
  51. McNamara, P.; Smyth, R.L. The pathogenesis of respiratory syncytial virus disease in childhood. Br. Med. Bull. 2002, 61, 13–28. [Google Scholar] [CrossRef]
  52. Glezen, W.P.; Taber, L.H.; Frank, A.L.; Kasel, J.A. Risk of Primary Infection and Reinfection with Respiratory Syncytial Virus. Arch. Pediatr. Adolesc. Med. 1986, 140, 543–546. [Google Scholar] [CrossRef] [PubMed]
  53. Baker, K.A.; Ryan, M.E. RSV infection in infants and young children. What’s new in diagnosis, treatment, and prevention? Postgrad. Med. 1999, 106, 97–111. [Google Scholar] [CrossRef]
  54. Anderson, N.W.; Binnicker, M.J.; Harris, D.M.; Chirila, R.M.; Brumble, L.; Mandrekar, J.; Hata, D.J. Morbidity and mortality among patients with respiratory syncytial virus infection: A 2-year retrospective review. Diagn. Microbiol. Infect. Dis. 2016, 85, 367–371. [Google Scholar] [CrossRef] [PubMed]
  55. Histoshi, T.; McAllister, D.A.; O’Brien, K.L.; Simoes, E.A.F.; Madhi, S.A.; Gessner, B.D.; Polack, F.P.; Balsells, E.; Acacio, S.; Aguayo, C.; et al. Global, regional, and national disease burden estimates of acute lower respiratory infections due to respiratory syncytial virus in young children in 2015: A systematic review and modelling study. Lancet 2017, 390, 946–958. [Google Scholar] [CrossRef] [Green Version]
  56. Scheltema, N.M.; Gentile, A.; Lucion, F.; Nokes, D.J.; Munywoki, P.K.; Madhi, S.; Groome, M.; Cohen, C.; Moyes, J.; Thorburn, K.; et al. Global respiratory syncytial virus-associated mortality in young children (RSV GOLD): A retrospective case series. Lancet Glob. Health 2017, 5, e984–e991. [Google Scholar] [CrossRef] [Green Version]
  57. Falsey, A.R.; Hennessey, P.A.; Formica, M.A.; Cox, C.; Walsh, E.E. Respiratory Syncytial Virus Infection in Elderly and High-Risk Adults. N. Engl. J. Med. 2005, 352, 1749–1759. [Google Scholar] [CrossRef] [PubMed]
  58. Bosco, E.; van Aalst, R.; McConeghy, K.W.; Silva, J.; Moyo, P.; Eliot, M.N.; Chit, A.; Gravenstein, S.; Zullo, A.R. Estimated Cardiorespiratory Hospitalizations Attributable to Influenza and Respiratory Syncytial Virus Among Long-term Care Facility Residents. JAMA Netw. Open 2021, 4, e2111806. [Google Scholar] [CrossRef]
  59. Nokes, D.J.; Okiro, E.; Ngama, M.; Ochola, R.; White, L.; Scott, P.D.; English, M.; Cane, P.A.; Medley, G. Respiratory Syncytial Virus Infection and Disease in Infants and Young Children Observed from Birth in Kilifi District, Kenya. Clin. Infect. Dis. 2008, 46, 50–57. [Google Scholar] [CrossRef]
  60. Zhou, Y.; Tong, L.; Li, M.; Wang, Y.; Li, L.; Yang, D.; Zhang, Y.; Chen, Z. Recurrent Wheezing and Asthma After Respiratory Syncytial Virus Bronchiolitis. Front. Pediatr. 2021, 9, 649003. [Google Scholar] [CrossRef]
  61. Henderson, J.; Hilliard, T.N.; Sherriff, A.; Stalker, D.; Shammari, N.A.; Thomas, H.M. Hospitalization for RSV bronchiolitis before 12 months of age and subsequent asthma, atopy and wheeze: A longitudinal birth cohort study. Pediatr. Allergy Immunol. 2005, 16, 386–392. [Google Scholar] [CrossRef] [PubMed]
  62. Korsten, K.; Blanken, M.O.; Buiteman, B.J.M.; Nibbelke, E.E.; Naaktgeboren, C.A.; Bont, L.J.; Wildenbeest, J.G. RSV hospitalization in infancy increases the risk of current wheeze at age 6 in late preterm born children without atopic predisposition. Eur. J. Nucl. Med. Mol. Imaging 2019, 178, 455–462. [Google Scholar] [CrossRef] [PubMed]
  63. Jartti, T.; Gern, J.E. Role of viral infections in the development and exacerbation of asthma in children. J. Allergy Clin. Immunol. 2017, 140, 895–906. [Google Scholar] [CrossRef] [Green Version]
  64. Wu, P.; Hartert, T.V. Evidence for a causal relationship between respiratory syncytial virus infection and asthma. Expert Rev. Anti-Infect. Ther. 2011, 9, 731–745. [Google Scholar] [CrossRef] [Green Version]
  65. Brandenburg, A.H.; Groen, J.; van Steensel-Moll, H.A.; Claas, E.C.; Rothbarth, P.H.; Neijens, H.J.; Osterhaus, A.D. Respiratory syncytial virus specific serum antibodies in infants under six months of age: Limited serological response upon infection. J. Med. Virol. 1997, 52, 97–104. [Google Scholar] [CrossRef] [Green Version]
  66. Jans, J.; Wicht, O.; Widjaja, I.; Ahout, I.M.L.; De Groot, R.; Guichelaar, T.; Luytjes, W.; De Jonge, M.I.; de Haan, C.; Ferwerda, G. Characteristics of RSV-Specific Maternal Antibodies in Plasma of Hospitalized, Acute RSV Patients under Three Months of Age. PLoS ONE 2017, 12, e0170877. [Google Scholar] [CrossRef] [PubMed]
  67. Abo, Y.; Clifford, V.; Lee, L.; Costa, A.; Crawford, N.; Wurzel, D.; Daley, A.J. COVID-19 public health measures and respiratory viruses in children in Melbourne. J. Paediatr. Child Health 2021. [Google Scholar] [CrossRef]
  68. Zhu, Y.; Li, W.; Yang, B.; Qian, R.; Wu, F.; He, X.; Zhu, Q.; Liu, J.; Ni, Y.; Wang, J.; et al. Epidemiological and virological characteristics of respiratory tract infections in children during COVID-19 outbreak. BMC Pediatr. 2021, 21, 1–8. [Google Scholar] [CrossRef]
  69. Killikelly, A.; Tunis, M.; House, A.; Quach, C.; Vaudry, W.; Moore, D. Overview of the respiratory syncytial virus vaccine candidate pipeline in Canada. Can. Commun. Dis. Rep. 2020, 46, 56–61. [Google Scholar] [CrossRef]
  70. Mazur, N.I.; Higgins, D.; Nunes, M.C.; Melero, J.A.; Langedijk, A.C.; Horsley, N.; Buchholz, U.J.; Openshaw, P.J.; McLellan, J.; Englund, J.A.; et al. The respiratory syncytial virus vaccine landscape: Lessons from the graveyard and promising candidates. Lancet Infect. Dis. 2018, 18, e295–e311. [Google Scholar] [CrossRef] [Green Version]
  71. Graham, B.S. Vaccine development for respiratory syncytial virus. Curr. Opin. Virol. 2017, 23, 107–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Al-Halifa, S.; Gauthier, L.; Arpin, D.; Bourgault, S.; Archambault, D. Nanoparticle-Based Vaccines Against Respiratory Viruses. Front. Immunol. 2019, 10, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Glowinski, R.; Mejias, A.; Ramilo, O. New preventive strategies for respiratory syncytial virus infection in children. Curr. Opin. Virol. 2021, 51, 216–223. [Google Scholar] [CrossRef] [PubMed]
  74. Domachowske, J.B.; Anderson, E.J.; Goldstein, M. The Future of Respiratory Syncytial Virus Disease Prevention and Treatment. Infect. Dis. Ther. 2021, 10, 47–60. [Google Scholar] [CrossRef] [PubMed]
  75. The IMpact-RSV Study Group. Palivizumab, a humanized respiratory syncytial virus monoclonal antibody, reduces hospitalization from respiratory syncytial virus infection in high-risk infants. Pediatrics 1998, 102, 531–537. [Google Scholar] [CrossRef]
  76. Resch, B. Product review on the monoclonal antibody palivizumab for prevention of respiratory syncytial virus infection. Hum. Vaccines Immunother. 2017, 13, 2138–2149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Committee on Infectious Diseases. Updated guidance for palivizumab prophylaxis among infants and young children at increased risk of hospitalization for respiratory syncytial virus infection. Pediatrics 2014, 134, 415–420. [Google Scholar] [CrossRef] [Green Version]
  78. Smith, D.K.; Seales, S.; Budzik, C. Respiratory Syncytial Virus Bronchiolitis in Children. Am. Fam. Physician 2017, 95, 94–99. [Google Scholar]
  79. Griffin, M.P.; Yuan, Y.; Takas, T.; Domachowske, J.B.; Madhi, S.A.; Manzoni, P.; Simões, E.A.F.; Esser, M.T.; Khan, A.A.; Dubovsky, F.; et al. Single-Dose Nirsevimab for Prevention of RSV in Preterm Infants. N. Engl. J. Med. 2020, 383, 415–425. [Google Scholar] [CrossRef]
  80. Zhu, Q.; McLellan, J.S.; Kallewaard, N.L.; Ulbrandt, N.D.; Palaszynski, S.; Zhang, J.; Moldt, B.; Khan, A.; Svabek, C.; McAuliffe, J.M.; et al. A highly potent extended half-life antibody as a potential RSV vaccine surrogate for all infants. Sci. Transl. Med. 2017, 9, eaaj1928. [Google Scholar] [CrossRef]
  81. Lai, S.K.; McSweeney, M.D.; Pickles, R.J. Learning from past failures: Challenges with monoclonal antibody therapies for COVID-19. J. Control. Release 2020, 329, 87–95. [Google Scholar] [CrossRef]
  82. Bergeron, H.C.; Tripp, R.A. Emerging small and large molecule therapeutics for respiratory syncytial virus. Expert Opin. Investig. Drugs 2020, 29, 285–294. [Google Scholar] [CrossRef] [PubMed]
  83. Whitsett, J.A.; Alenghat, T. Respiratory epithelial cells orchestrate pulmonary innate immunity. Nat. Immunol. 2014, 16, 27–35. [Google Scholar] [CrossRef] [Green Version]
  84. Whitsett, J.A. Airway Epithelial Differentiation and Mucociliary Clearance. Ann. Am. Thorac. Soc. 2018, 15, S143–S148. [Google Scholar] [CrossRef]
  85. Voelker, D.R.; Numata, M. Phospholipid regulation of innate immunity and respiratory viral infection. J. Biol. Chem. 2019, 294, 4282–4289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Bertuzzi, M.; Hayes, G.E.; Bignell, E.M. Microbial uptake by the respiratory epithelium: Outcomes for host and pathogen. FEMS Microbiol. Rev. 2019, 43, 145–161. [Google Scholar] [CrossRef]
  87. Ioannidis, I.; McNally, B.; Willette, M.; Peeples, M.E.; Chaussabel, D.; Durbin, J.E.; Ramilo, O.; Mejias, A.; Flano, E. Plasticity and Virus Specificity of the Airway Epithelial Cell Immune Response during Respiratory Virus Infection. J. Virol. 2012, 86, 5422–5436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Villenave, R.; Shields, M.D.; Power, U.F. Respiratory syncytial virus interaction with human airway epithelium. Trends Microbiol. 2013, 21, 238–244. [Google Scholar] [CrossRef] [PubMed]
  89. Carty, M.; Guy, C.; Bowie, A.G. Detection of Viral Infections by Innate Immunity. Biochem. Pharmacol. 2020, 183, 114316. [Google Scholar] [CrossRef] [PubMed]
  90. Thompson, M.R.; Kaminski, J.J.; Kurt-Jones, E.A.; Fitzgerald, K.A. Pattern Recognition Receptors and the Innate Immune Response to Viral Infection. Viruses 2011, 3, 920–940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Cavlar, T.; Ablasser, A.; Hornung, V. Induction of type I IFNs by intracellular DNA-sensing pathways. Immunol. Cell Biol. 2012, 90, 474–482. [Google Scholar] [CrossRef] [PubMed]
  92. Rojas, J.M.; Alejo, A.; Martín, V.; Sevilla, N. Viral pathogen-induced mechanisms to antagonize mammalian interferon (IFN) signaling pathway. Cell. Mol. Life Sci. 2020, 78, 1423–1444. [Google Scholar] [CrossRef]
  93. Mezouar, S.; Mege, J. Changing the paradigm of IFN-γ at the interface between innate and adaptive immunity: Macrophage-derived IFN-γ. J. Leukoc. Biol. 2020, 108, 419–426. [Google Scholar] [CrossRef] [PubMed]
  94. Sobah, M.L.; Liongue, C.; Ward, A.C. SOCS Proteins in Immunity, Inflammatory Diseases, and Immune-Related Cancer. Front. Med. 2021, 8, 727987. [Google Scholar] [CrossRef] [PubMed]
  95. Smith, P.K.; Wang, S.Z.; Dowling, K.D.; Forsyth, K.D. Leucocyte populations in respiratory syncytial virus-induced bronchiolitis. J. Paediatr. Child Health 2001, 37, 146–151. [Google Scholar] [CrossRef]
  96. Everard, M.L.; Swarbrick, A.; Wrightham, M.; McIntyre, J.; Dunkley, C.; James, P.D.; Sewell, H.F.; Milner, A.D. Analysis of cells obtained by bronchial lavage of infants with respiratory syncytial virus infection. Arch. Dis. Child. 1994, 71, 428–432. [Google Scholar] [CrossRef] [Green Version]
  97. Qin, L.; Feng, J.; Hu, C.; Li, Y.; Niu, R. Th17/Treg imbalance mediated by IL-8 in RSV-infected bronchial epithelial cells. Zhong Nan Da Xue Xue Bao Yi Xue Ban 2016, 41, 337–344. [Google Scholar] [CrossRef]
  98. Oshansky, C.M.; Barber, J.P.; Crabtree, J.; Tripp, R.A. Respiratory Syncytial Virus F and G Proteins Induce Interleukin 1α, CC, and CXC Chemokine Responses by Normal Human Bronchoepithelial Cells. J. Infect. Dis. 2010, 201, 1201–1207. [Google Scholar] [CrossRef] [Green Version]
  99. Funchal, G.A.; Jaeger, N.; Czepielewski, R.S.; Machado, M.S.; Muraro, S.; Stein, R.; Bonorino, C.B.C.; Porto, B.N. Respiratory Syncytial Virus Fusion Protein Promotes TLR-4—Dependent Neutrophil Extracellular Trap Formation by Human Neutrophils. PLoS ONE 2015, 10, e0124082. [Google Scholar] [CrossRef]
  100. He, W.; Qu, T.; Yu, Q.; Wang, Z.; Lv, H.; Zhang, J.; Zhao, X.; Wang, P. LPS induces IL-8 expression through TLR4, MyD88, NF-kappaB and MAPK pathways in human dental pulp stem cells. Int. Endod. J. 2012, 46, 128–136. [Google Scholar] [CrossRef]
  101. Goetghebuer, T.; Isles, K.; Moore, C.; Thomson, A.; Kwiatkowski, D.; Hull, J. Genetic predisposition to wheeze following respiratory syncytial virus bronchiolitis. Clin. Exp. Allergy 2004, 34, 801–803. [Google Scholar] [CrossRef]
  102. Hull, J.; Thomson, A.; Kwiatkowski, D. Association of respiratory syncytial virus bronchiolitis with the interleukin 8 gene region in UK families. Thorax 2000, 55, 1023–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Deng, Y.; Herbert, J.; Robinson, E.; Ren, L.; Smyth, R.L.; Smith, C.M. Neutrophil-Airway Epithelial Interactions Result in Increased Epithelial Damage and Viral Clearance during Respiratory Syncytial Virus Infection. J. Virol. 2020, 94, e02161-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Johnson, J.E.; Gonzales, R.A.; Olson, S.J.; Wright, P.F.; Graham, B.S. The histopathology of fatal untreated human respiratory syncytial virus infection. Mod. Pathol. 2006, 20, 108–119. [Google Scholar] [CrossRef]
  105. Alba, C.; Aparicio, M.; González-Martínez, F.; González-Sánchez, M.I.; Pérez-Moreno, J.; del Castillo, B.T.; Rodríguez, J.M.; Rodríguez-Fernández, R.; Fernández, L. Nasal and Fecal Microbiota and Immunoprofiling of Infants with and without RSV Bronchiolitis. Front. Microbiol. 2021, 12, 667832. [Google Scholar] [CrossRef] [PubMed]
  106. Dapat, C.; Kumaki, S.; Sakurai, H.; Nishimura, H.; Labayo, H.K.M.; Okamoto, M.; Saito, M.; Oshitani, H. Gene signature of children with severe respiratory syncytial virus infection. Pediatr. Res. 2021, 89, 1664–1672. [Google Scholar] [CrossRef]
  107. Korematsu, T.; Koga, H. Transient Neutropenia in Immunocompetent Infants with Respiratory Syncytial Virus Infection. Viruses 2021, 13, 301. [Google Scholar] [CrossRef]
  108. Murawski, M.R.; Bowen, G.N.; Cerny, A.M.; Anderson, L.J.; Haynes, L.M.; Tripp, R.; Kurt-Jones, E.A.; Finberg, R.W. Respiratory Syncytial Virus Activates Innate Immunity through Toll-Like Receptor 2. J. Virol. 2009, 83, 1492–1500. [Google Scholar] [CrossRef] [Green Version]
  109. Culley, F.J.; Pennycook, A.M.J.; Tregoning, J.S.; Hussell, T.; Openshaw, P.J.M. Differential Chemokine Expression following Respiratory Virus Infection Reflects Th1- or Th2-Biased Immunopathology. J. Virol. 2006, 80, 4521–4527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Ullah, A.; Rittchen, S.; Li, J.; Hasnain, S.Z.; Phipps, S. DP1 prostanoid receptor activation increases the severity of an acute lower respiratory viral infection in mice via TNF-α-induced immunopathology. Mucosal Immunol. 2021, 14, 963–972. [Google Scholar] [CrossRef]
  111. Weibel, R.E.; Stokes, J.; Leagus, M.B.; Mascoli, C.C.; Tytell, A.A.; Woodhour, A.F.; Vella, P.P.; Hilleman, M.R. Respiratory virus vaccines. VII. Field evaluation of respiratory syncytial, parainfluenza 1, 2, 3, and Mycoplasma pneumoniae vaccines, 1965 to 1966. Am. Rev. Respir. Dis. 1967, 96, 724–739. [Google Scholar] [CrossRef] [PubMed]
  112. Chin, J.; Magoffin, R.L.; Shearer, L.A.; Schieble, J.H.; Lennette, E.H. Field evaluation of a respiratory syncytial virus vaccine and a trivalent parainfluenza virus vaccine in a pediatric population. Am. J. Epidemiol. 1969, 89, 449–463. [Google Scholar] [CrossRef] [PubMed]
  113. Fulginiti, V.A.; Eller, J.J.; Sieber, O.F.; Joyner, J.W.; Minamitani, M.; Meiklejohn, G. Respiratory virus immunization. I. A field trial of two inactivated respiratory virus vaccines; an aqueous trivalent parainfluenza virus vaccine and an alum-precipitated respiratory syncytial virus vaccine. Am. J. Epidemiol. 1969, 89, 435–448. [Google Scholar] [CrossRef] [PubMed]
  114. Acosta, P.; Caballero, M.T.; Polack, F.P. Brief History and Characterization of Enhanced Respiratory Syncytial Virus Disease. Clin. Vaccine Immunol. 2016, 23, 189–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Kirsebom, F.; Michalaki, C.; Oyarzabal, M.A.; Johansson, C. Neutrophils do not impact viral load or the peak of disease severity during RSV infection. Sci. Rep. 2020, 10, 1–12. [Google Scholar] [CrossRef] [Green Version]
  116. Joshi, N.; Walter, J.M.; Misharin, A.V. Alveolar Macrophages. Cell. Immunol. 2018, 330, 86–90. [Google Scholar] [CrossRef]
  117. Santos, L.D.; Antunes, K.H.; Muraro, S.P.; de Souza, G.F.; da Silva, A.G.; Felipe, J.D.S.; Zanetti, L.C.; Czepielewski, R.S.; Magnus, K.; Scotta, M.; et al. TNF-mediated alveolar macrophage necroptosis drives disease pathogenesis during respiratory syncytial virus infection. Eur. Respir. J. 2020, 57, 2003764. [Google Scholar] [CrossRef]
  118. de Souza, G.F.; Muraro, S.; Santos, L.D.; Monteiro, A.P.T.; Da Silva, A.G.; Souza, A.P.; Stein, R.; Bozza, P.; Porto, B.N. Macrophage migration inhibitory factor (MIF) controls cytokine release during respiratory syncytial virus infection in macrophages. Inflamm. Res. 2019, 68, 481–491. [Google Scholar] [CrossRef] [PubMed]
  119. Huang, S.; Zhu, B.; Cheon, I.S.; Goplen, N.P.; Jiang, L.; Zhang, R.; Peebles, R.S.; Mack, M.; Kaplan, M.H.; Limper, A.H.; et al. PPAR-γ in Macrophages Limits Pulmonary Inflammation and Promotes Host Recovery following Respiratory Viral Infection. J. Virol. 2019, 93, e00030-19. [Google Scholar] [CrossRef] [Green Version]
  120. Nguyen, T.H.; Maltby, S.; Tay, H.L.; Eyers, F.; Foster, P.S.; Yang, M. Identification of IFN-γ and IL-27 as Critical Regulators of Respiratory Syncytial Virus–Induced Exacerbation of Allergic Airways Disease in a Mouse Model. J. Immunol. 2017, 200, 237–247. [Google Scholar] [CrossRef]
  121. Grunwell, J.R.; Yeligar, S.M.; Stephenson, S.; Du Ping, X.; Gauthier, T.W.; Fitzpatrick, A.M.; Brown, L.A.S. TGF-β1 Suppresses the Type I IFN Response and Induces Mitochondrial Dysfunction in Alveolar Macrophages. J. Immunol. 2018, 200, 2115–2128. [Google Scholar] [CrossRef] [Green Version]
  122. Qi, F.; Bai, S.; Wang, D.; Xu, L.; Hu, H.; Zeng, S.; Chai, R.; Liu, B. Macrophages produce IL-33 by activating MAPK signaling pathway during RSV infection. Mol. Immunol. 2017, 87, 284–292. [Google Scholar] [CrossRef] [PubMed]
  123. Lee, Y.-T.; Ko, E.-J.; Kim, K.-H.; Hwang, H.S.; Lee, Y.; Kwon, Y.-M.; Kim, M.-C.; Lee, Y.-N.; Jung, Y.-J.; Kang, S.-M. Cellular Immune Correlates Preventing Disease Against Respiratory Syncytial Virus by Vaccination with Virus-Like Nanoparticles Carrying Fusion Proteins. J. Biomed. Nanotechnol. 2017, 13, 84–98. [Google Scholar] [CrossRef] [Green Version]
  124. Ducreux, J.; Crocker, P.; Vanbever, R. Analysis of sialoadhesin expression on mouse alveolar macrophages. Immunol. Lett. 2009, 124, 77–80. [Google Scholar] [CrossRef]
  125. Oh, D.S.; Oh, J.E.; Jung, H.E.; Lee, H.K. Transient Depletion of CD169+ Cells Contributes to Impaired Early Protection and Effector CD8+ T Cell Recruitment against Mucosal Respiratory Syncytial Virus Infection. Front. Immunol. 2017, 8, 819. [Google Scholar] [CrossRef]
  126. Tripp, R.A.; Jones, L.; Anderson, L.J. Respiratory Syncytial Virus G and/or SH Glycoproteins Modify CC and CXC Chemokine mRNA Expression in the BALB/c Mouse. J. Virol. 2000, 74, 6227–6229. [Google Scholar] [CrossRef] [Green Version]
  127. Castilow, E.M.; Olson, M.R.; Varga, S.M. Understanding respiratory syncytial virus (RSV) vaccine-enhanced disease. Immunol. Res. 2007, 39, 225–239. [Google Scholar] [CrossRef] [PubMed]
  128. Openshaw, P.J.M.; Clarke, S.L.; Record, F.M. Pulmonary eosinophilic response to respiratory syncytial virus infection in mice sensitized to the major surface glycoprotein G. Int. Immunol. 1992, 4, 493–500. [Google Scholar] [CrossRef]
  129. Tripp, R.; Moore, D.; Jones, L.; Sullender, W.; Winter, J.; Anderson, L.J. Respiratory Syncytial Virus G and/or SH Protein Alters Th1 Cytokines, Natural Killer Cells, and Neutrophils Responding to Pulmonary Infection in BALB/c Mice. J. Virol. 1999, 73, 7099–7107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Hancock, G.E.; Speelman, D.J.; Heers, K.; Bortell, E.; Smith, J.; Cosco, C. Generation of atypical pulmonary inflammatory responses in BALB/c mice after immunization with the native attachment (G) glycoprotein of respiratory syncytial virus. J. Virol. 1996, 70, 7783–7791. [Google Scholar] [CrossRef] [Green Version]
  131. Johnson, T.R.; Johnson, J.E.; Roberts, S.R.; Wertz, G.W.; Parker, R.A.; Graham, B.S. Priming with Secreted Glycoprotein G of Respiratory Syncytial Virus (RSV) Augments Interleukin-5 Production and Tissue Eosinophilia after RSV Challenge. J. Virol. 1998, 72, 2871–2880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Bembridge, G.P.; Garcia-Beato, R.; Lopez, J.A.; Melero, J.A.; Taylor, G. Subcellular site of expression and route of vaccination influence pulmonary eosinophilia following respiratory syncytial virus challenge in BALB/c mice sensitized to the attachment G protein. J. Immunol. 1998, 161, 2473–2480. [Google Scholar] [PubMed]
  133. Johnson, T.R.; Teng, M.N.; Collins, P.L.; Graham, B.S. Respiratory Syncytial Virus (RSV) G Glycoprotein Is Not Necessary for Vaccine-Enhanced Disease Induced by Immunization with Formalin-Inactivated RSV. J. Virol. 2004, 78, 6024–6032. [Google Scholar] [CrossRef] [Green Version]
  134. Cheon, I.S.; Kim, J.; Choi, Y.; Shim, B.-S.; Choi, J.-A.; Jung, D.-I.; Kim, J.-O.; Braciale, T.J.; Youn, H.; Song, M.K.; et al. Sublingual Immunization With an RSV G Glycoprotein Fragment Primes IL-17-Mediated Immunopathology Upon Respiratory Syncytial Virus Infection. Front. Immunol. 2019, 10, 567. [Google Scholar] [CrossRef] [PubMed]
  135. Ramakrishnan, R.K.; Al Heialy, S.; Hamid, Q. Role of IL-17 in asthma pathogenesis and its implications for the clinic. Expert Rev. Respir. Med. 2019, 13, 1057–1068. [Google Scholar] [CrossRef]
  136. Johnson, T.R.; Rothenberg, M.E.; Graham, B.S. Pulmonary eosinophilia requires interleukin-5, eotaxin-1, and CD4+ T cells in mice immunized with respiratory syncytial virus G glycoprotein. J. Leukoc. Biol. 2008, 84, 748–759. [Google Scholar] [CrossRef]
  137. Schwarze, J.; Cieslewicz, G.; Hamelmann, E.; Joetham, A.; Shultz, L.D.; Lamers, M.C.; Gelfand, E.W. IL-5 and eosinophils are essential for the development of airway hyperresponsiveness following acute respiratory syncytial virus infection. J. Immunol. 1999, 162, 2997–3004. [Google Scholar] [PubMed]
  138. Tebbey, P.W.; Hagen, M.; Hancock, G.E. Atypical Pulmonary Eosinophilia Is Mediated by a Specific Amino Acid Sequence of the Attachment (G) Protein of Respiratory Syncytial Virus. J. Exp. Med. 1998, 188, 1967–1972. [Google Scholar] [CrossRef] [Green Version]
  139. Phipps, S.; Lam, C.E.; Mahalingam, S.; Newhouse, M.; Ramirez, R.; Rosenberg, H.F.; Foster, P.S.; Matthaei, K.I. Eosinophils contribute to innate antiviral immunity and promote clearance of respiratory syncytial virus. Blood 2007, 110, 1578–1586. [Google Scholar] [CrossRef] [Green Version]
  140. Suresh, M.; Townsend, D.; Herrero, L.; Zaid, A.; Rolph, M.S.; Gahan, M.; Nelson, M.A.; Rudd, P.A.; Matthaei, K.I.; Foster, P.S.; et al. Dual Proinflammatory and Antiviral Properties of Pulmonary Eosinophils in Respiratory Syncytial Virus Vaccine-Enhanced Disease. J. Virol. 2014, 89, 1564–1578. [Google Scholar] [CrossRef] [Green Version]
  141. Castilow, E.M.; Meyerholz, D.; Varga, S. IL-13 Is Required for Eosinophil Entry into the Lung during Respiratory Syncytial Virus Vaccine-Enhanced Disease. J. Immunol. 2008, 180, 2376–2384. [Google Scholar] [CrossRef] [Green Version]
  142. Johnson, T.R.; Parker, R.A.; Johnson, J.E.; Graham, B.S. IL-13 Is Sufficient for Respiratory Syncytial Virus G Glycoprotein-Induced Eosinophilia After Respiratory Syncytial Virus Challenge. J. Immunol. 2003, 170, 2037–2045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Rosenberg, H.F.; Dyer, K.D.; Domachowske, J. Respiratory viruses and eosinophils: Exploring the connections. Antivir. Res. 2009, 83, 1–9. [Google Scholar] [CrossRef]
  144. Mader, D.; Huang, Y.; Wang, C.; Fraser, R.; Issekutz, A.C.; Stadnyk, A.W.; Anderson, R. Liposome encapsulation of a soluble recombinant fragment of the respiratory syncytial virus (RSV) G protein enhances immune protection and reduces lung eosinophilia associated with virus challenge. Vaccine 2000, 18, 1110–1117. [Google Scholar] [CrossRef]
  145. Castilow, E.M.; Legge, K.L.; Varga, S.M. Eosinophils do not contribute to respiratory syncytial virus vaccine-enhanced disease1. J. Immunol. 2008, 181, 6692–6696. [Google Scholar] [CrossRef] [Green Version]
  146. Van Hulst, G.; Batugedara, H.M.; Jorssen, J.; Louis, R.; Bureau, F.; Desmet, C.J. Eosinophil diversity in asthma. Biochem. Pharmacol. 2020, 179, 113963. [Google Scholar] [CrossRef] [PubMed]
  147. Bhat, R.; Farrag, M.; Almajhdi, F.N. Double-edged role of natural killer cells during RSV infection. Int. Rev. Immunol. 2020, 39, 233–244. [Google Scholar] [CrossRef]
  148. Noyola, D.E.; Juárez-Vega, G.; Monjarás-Ávila, C.; Escalante-Padrón, F.; Rangel-Ramírez, V.; Cadena-Mota, S.; Monsiváis-Urenda, A.; García-Sepúlveda, C.A.; Gonzalez-Amaro, R. NK cell immunophenotypic and genotypic analysis of infants with severe respiratory syncytial virus infection. Microbiol. Immunol. 2015, 59, 389–397. [Google Scholar] [CrossRef] [Green Version]
  149. Ko, E.-J.; Lee, Y.; Lee, Y.-T.; Hwang, H.S.; Park, Y.; Kim, K.-H.; Kang, S.-M. Natural Killer and CD8 T Cells Contribute to Protection by Formalin Inactivated Respiratory Syncytial Virus Vaccination under a CD4-Deficient Condition. Immune Netw. 2020, 20, e51. [Google Scholar] [CrossRef]
  150. Harker, J.A.; Godlee, A.; Wahlsten, J.L.; Lee, D.C.P.; Thorne, L.G.; Sawant, D.; Tregoning, J.S.; Caspi, R.R.; Bukreyev, A.; Collins, P.L.; et al. Interleukin 18 Coexpression during Respiratory Syncytial Virus Infection Results in Enhanced Disease Mediated by Natural Killer Cells. J. Virol. 2010, 84, 4073–4082. [Google Scholar] [CrossRef] [Green Version]
  151. Van Erp, E.A.; Feyaerts, D.; Duijst, M.; Mulder, H.L.; Wicht, O.; Luytjes, W.; Ferwerda, G.; Van Kasteren, P.B. Respiratory Syncytial Virus Infects Primary Neonatal and Adult Natural Killer Cells and Affects Their Antiviral Effector Function. J. Infect. Dis. 2018, 219, 723–733. [Google Scholar] [CrossRef]
  152. Erp, E.A.; Lakerveld, A.J.; de Graaf, E.; Larsen, M.D.; Schepp, R.M.; Ederveen, A.L.H.; Ahout, I.M.; de Haan, C.A.; Wuhrer, M.; Luytjes, W.; et al. Natural killer cell activation by respiratory syncytial virus-specific antibodies is decreased in infants with severe respiratory infections and correlates with Fc-glycosylation. Clin. Transl. Immunol. 2020, 9, e1112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Li, F.; Zhu, H.; Sun, R.; Wei, H.; Tian, Z. Natural Killer Cells Are Involved in Acute Lung Immune Injury Caused by Respiratory Syncytial Virus Infection. J. Virol. 2011, 86, 2251–2258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Haynes, L.M.; Moore, D.D.; Kurt-Jones, E.A.; Finberg, R.W.; Anderson, L.J.; Tripp, R.A. Involvement of Toll-Like Receptor 4 in Innate Immunity to Respiratory Syncytial Virus. J. Virol. 2001, 75, 10730–10737. [Google Scholar] [CrossRef] [Green Version]
  155. Tripp, R.A.; Moore, D.; Winter, J.; Anderson, L.J. Respiratory Syncytial Virus Infection and G and/or SH Protein Expression Contribute to Substance P, Which Mediates Inflammation and Enhanced Pulmonary Disease in BALB/c Mice. J. Virol. 2000, 74, 1614–1622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Hussell, T.; Openshaw, P.J. Intracellular IFN-gamma expression in natural killer cells precedes lung CD8+ T cell recruitment during respiratory syncytial virus infection. J. Gen. Virol. 1998, 79, 2593–2601. [Google Scholar] [CrossRef]
  157. Segura, E. Review of Mouse and Human Dendritic Cell Subsets. Methods Mol. Biol. 2016, 1423, 3–15. [Google Scholar] [CrossRef]
  158. Jung, H.E.; Kim, T.H.; Lee, H.K. Contribution of Dendritic Cells in Protective Immunity against Respiratory Syncytial Virus Infection. Viruses 2020, 12, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Guerrero-Plata, A.; Casola, A.; Suarez, G.; Yu, X.; Spetch, L.; Peeples, M.E.; Garofalo, R.P. Differential Response of Dendritic Cells to Human Metapneumovirus and Respiratory Syncytial Virus. Am. J. Respir. Cell Mol. Biol. 2006, 34, 320–329. [Google Scholar] [CrossRef] [Green Version]
  160. Kim, T.H.; Oh, D.S.; Jung, H.E.; Chang, J.; Lee, H.K. Plasmacytoid Dendritic Cells Contribute to the Production of IFN-β via TLR7-MyD88-Dependent Pathway and CTL Priming during Respiratory Syncytial Virus Infection. Viruses 2019, 11, 730. [Google Scholar] [CrossRef] [Green Version]
  161. Schiavoni, I.; Scagnolari, C.; Horenstein, A.L.; Leone, P.; Pierangeli, A.; Malavasi, F.; Ausiello, C.M.; Fedele, G. CD38 modulates respiratory syncytial virus-driven proinflammatory processes in human monocyte-derived dendritic cells. Immunology 2017, 154, 122–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Oh, D.S.; Kim, T.H.; Lee, H.K. Differential Role of Anti-Viral Sensing Pathway for the Production of Type I Interferon β in Dendritic Cells and Macrophages Against Respiratory Syncytial Virus A2 Strain Infection. Viruses 2019, 11, 62. [Google Scholar] [CrossRef] [Green Version]
  163. Shrestha, B.; You, D.; Saravia, J.; Siefker, D.T.; Jaligama, S.; Lee, G.I.; Sallam, A.A.; Harding, J.N.; Cormier, S.A. IL-4Rα on dendritic cells in neonates and Th2 immunopathology in respiratory syncytial virus infection. J. Leukoc. Biol. 2017, 102, 153–161. [Google Scholar] [CrossRef]
  164. Smit, J.; Rudd, B.D.; Lukacs, N.W. Plasmacytoid dendritic cells inhibit pulmonary immunopathology and promote clearance of respiratory syncytial virus. J. Exp. Med. 2006, 203, 1153–1159. [Google Scholar] [CrossRef] [PubMed]
  165. Beyer, M.; Bartz, H.; Hörner, K.; Doths, S.; Koerner-Rettberg, C.; Schwarze, J. Sustained increases in numbers of pulmonary dendritic cells after respiratory syncytial virus infection. J. Allergy Clin. Immunol. 2004, 113, 127–133. [Google Scholar] [CrossRef] [PubMed]
  166. Qi, F.; Wang, D.; Liu, J.; Zeng, S.; Xu, L.; Hu, H.; Liu, B. Respiratory macrophages and dendritic cells mediate respiratory syncytial virus-induced IL-33 production in TLR3- or TLR7-dependent manner. Int. Immunopharmacol. 2015, 29, 408–415. [Google Scholar] [CrossRef] [PubMed]
  167. Lau-Kilby, A.; Turfkruyer, M.; Kehl, M.; Yang, L.; Buchholz, U.J.; Hickey, K.; Malloy, A.M.W. Type I IFN ineffectively activates neonatal dendritic cells limiting respiratory antiviral T-cell responses. Mucosal Immunol. 2019, 13, 371–380. [Google Scholar] [CrossRef] [PubMed]
  168. Le Nouën, C.; Hillyer, P.; Levenson, E.; Martens, C.; Rabin, R.L.; Collins, P.L.; Buchholz, U.J. Lack of Activation Marker Induction and Chemokine Receptor Switch in Human Neonatal Myeloid Dendritic Cells in Response to Human Respiratory Syncytial Virus. J. Virol. 2019, 93, e01216-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Marr, N.; Wang, T.-I.; Kam, S.H.Y.; Hu, Y.S.; Sharma, A.A.; Lam, A.; Markowski, J.; Solimano, A.; Lavoie, P.; Turvey, S. Attenuation of Respiratory Syncytial Virus–Induced and RIG-I–Dependent Type I IFN Responses in Human Neonates and Very Young Children. J. Immunol. 2014, 192, 948–957. [Google Scholar] [CrossRef]
  170. Elesela, S.; Morris, S.B.; Narayanan, S.; Kumar, S.; Lombard, D.B.; Lukacs, N.W. Sirtuin 1 regulates mitochondrial function and immune homeostasis in respiratory syncytial virus infected dendritic cells. PLoS Pathog. 2020, 16, e1008319. [Google Scholar] [CrossRef]
  171. Malinczak, C.-A.; Rasky, A.J.; Fonseca, W.; Schaller, M.A.; Allen, R.M.; Ptaschinski, C.; Morris, S.; Lukacs, N.W. Upregulation of H3K27 Demethylase KDM6 During Respiratory Syncytial Virus Infection Enhances Proinflammatory Responses and Immunopathology. J. Immunol. 2019, 204, 159–168. [Google Scholar] [CrossRef] [PubMed]
  172. Jin, L.; Hu, Q.; Hu, Y.; Chen, Z.; Liao, W. Respiratory Syncytial Virus Infection Reduces Kynurenic Acid Production and Reverses Th17/Treg Balance by Modulating Indoleamine 2,3-Dioxygenase (IDO) Molecules in Plasmacytoid Dendritic Cells. Med. Sci. Monit. 2020, 26, e926763. [Google Scholar] [CrossRef]
  173. Gonzalez, P.A.; Prado, C.E.; Leiva, E.; Carreno, L.J.; Bueno, S.M.; Riedel, C.; Kalergis, A.M. Respiratory syncytial virus impairs T cell activation by preventing synapse assembly with dendritic cells. Proc. Natl. Acad. Sci. USA 2008, 105, 14999–15004. [Google Scholar] [CrossRef] [Green Version]
  174. Rothoeft, T.; Fischer, K.; Zawatzki, S.; Schulz, V.; Schauer, U.; Rettberg, C.K. Differential response of human naive and memory/effector T cells to dendritic cells infected by respiratory syncytial virus. Clin. Exp. Immunol. 2007, 150, 263–273. [Google Scholar] [CrossRef]
  175. Ryan, D.; Tang, J. Regulation of Human B Cell Lymphopoiesis by Adhesion Molecules and Cytokines. Leuk. Lymphoma 1995, 17, 375–389. [Google Scholar] [CrossRef] [PubMed]
  176. Vinuesa, C.G.; Linterman, M.; Goodnow, C.; Randall, K.L. T cells and follicular dendritic cells in germinal center B-cell formation and selection. Immunol. Rev. 2010, 237, 72–89. [Google Scholar] [CrossRef] [PubMed]
  177. Raes, M.; Peeters, V.; Alliet, P.; Gillis, P.; Kortleven, J.; Magerman, K.; Rummens, J.L. Peripheral blood T and B lymphocyte subpopulations in infants with acute respiratory syncytial virus brochiolitis. Pediatr. Allergy Immunol. 1997, 8, 97–102. [Google Scholar] [CrossRef] [PubMed]
  178. Román, M.; Calhoun, W.J.; Hinton, K.L.; Avendaño, L.F.; Simon, V.; Escobar, A.M.; Gaggero, A.; Díaz, P.V. Respiratory Syncytial Virus Infection in Infants Is Associated with Predominant Th-2-like Response. Am. J. Respir. Crit. Care Med. 1997, 156, 190–195. [Google Scholar] [CrossRef]
  179. Reed, J.L.; Welliver, T.P.; Sims, G.P.; McKinney, L.; Velozo, L.; Avendano, L.; Hintz, K.; Luma, J.; Coyle, A.J.; Welliver, S.R.C. Innate Immune Signals Modulate Antiviral and Polyreactive Antibody Responses during Severe Respiratory Syncytial Virus Infection. J. Infect. Dis. 2009, 199, 1128–1138. [Google Scholar] [CrossRef] [PubMed]
  180. Alturaiki, W.; McFarlane, A.J.; Rose, K.; Corkhill, R.; McNamara, P.S.; Schwarze, J.; Flanagan, B.F. Expression of the B cell differentiation factor BAFF and chemokine CXCL13 in a murine model of Respiratory Syncytial Virus infection. Cytokine 2018, 110, 267–271. [Google Scholar] [CrossRef] [PubMed]
  181. Shehata, L.; Wieland-Alter, W.F.; Maurer, D.; Chen, E.; Connor, R.I.; Wright, P.F.; Walker, L.M. Systematic comparison of respiratory syncytial virus-induced memory B cell responses in two anatomical compartments. Nat. Commun. 2019, 10, 1–9. [Google Scholar] [CrossRef] [Green Version]
  182. Goodwin, E.; Gilman, M.S.; Wrapp, D.; Chen, M.; Ngwuta, J.O.; Moin, S.; Bai, P.; Sivasubramanian, A.; Connor, R.I.; Wright, P.F.; et al. Infants Infected with Respiratory Syncytial Virus Generate Potent Neutralizing Antibodies that Lack Somatic Hypermutation. Immunity 2018, 48, 339–349.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Schneikart, G.; Tavarini, S.; Sammicheli, C.; Torricelli, G.; Guidotti, S.; Andreano, E.; Buricchi, F.; D’Oro, U.; Finco, O.; Bardelli, M. The respiratory syncytial virus fusion protein-specific B cell receptor repertoire reshaped by post-fusion subunit vaccination. Vaccine 2020, 38, 7916–7927. [Google Scholar] [CrossRef]
  184. Jans, J.; Pettengill, M.; Kim, D.; van der Made, C.; de Groot, R.; Henriet, S.; de Jonge, M.I.; Ferwerda, G.; Levy, O. Human newborn B cells mount an interferon-α/β receptor-dependent humoral response to respiratory syncytial virus. J. Allergy Clin. Immunol. 2016, 139, 1997–2000.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Murphy, B.R.; Walsh, E.E. Formalin-inactivated respiratory syncytial virus vaccine induces antibodies to the fusion glycoprotein that are deficient in fusion-inhibiting activity. J. Clin. Microbiol. 1988, 26, 1595–1597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Son, Y.M.; Sun, J. Co-Ordination of Mucosal B Cell and CD8 T Cell Memory by Tissue-Resident CD4 Helper T Cells. Cells 2021, 10, 2355. [Google Scholar] [CrossRef]
  187. Swain, S.L.; McKinstry, K.K.; Strutt, T.M. Expanding roles for CD4+ T cells in immunity to viruses. Nat. Rev. Immunol. 2012, 12, 136–148. [Google Scholar] [CrossRef]
  188. Gagliani, N.; Huber, S. Basic Aspects of T Helper Cell Differentiation. Methods Mol. Biol. 2016, 1514, 19–30. [Google Scholar] [CrossRef]
  189. Chapoval, S.; Dasgupta, P.; Dorsey, N.J.; Keegan, A.D. Regulation of the T helper cell type 2 (Th2)/T regulatory cell (Treg) balance by IL-4 and STAT6. J. Leukoc. Biol. 2010, 87, 1011–1018. [Google Scholar] [CrossRef] [Green Version]
  190. Iwanaga, N.; Kolls, J.K. Updates on T helper type 17 immunity in respiratory disease. Immunology 2018, 156, 3–8. [Google Scholar] [CrossRef] [Green Version]
  191. Licona-Limón, P.; Arias-Rojas, A.; Olguín-Martínez, E. IL-9 and Th9 in parasite immunity. Semin. Immunopathol. 2016, 39, 29–38. [Google Scholar] [CrossRef]
  192. Eto, D.; Lao, C.; DiToro, D.; Barnett, B.; Escobar, T.C.; Kageyama, R.; Yusuf, I.; Crotty, S. IL-21 and IL-6 Are Critical for Different Aspects of B Cell Immunity and Redundantly Induce Optimal Follicular Helper CD4 T Cell (Tfh) Differentiation. PLoS ONE 2011, 6, e17739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Glatman Zaretsky, A.; Taylor, J.J.; King, I.L.; Marshall, F.A.; Mohrs, M.; Pearce, E.J. T follicular helper cells differentiate from Th2 cells in response to helminth antigens. J. Exp. Med. 2009, 206, 991–999. [Google Scholar] [CrossRef] [Green Version]
  194. Zhu, J. T Helper Cell Differentiation, Heterogeneity, and Plasticity. Cold Spring Harb. Perspect. Biol. 2017, 10, a030338. [Google Scholar] [CrossRef] [Green Version]
  195. Sallusto, F.; Geginat, J.; Lanzavecchia, A. Central Memory and Effector Memory T Cell Subsets: Function, Generation, and Maintenance. Annu. Rev. Immunol. 2004, 22, 745–763. [Google Scholar] [CrossRef] [PubMed]
  196. Mahnke, Y.; Brodie, T.M.; Sallusto, F.; Roederer, M.; Lugli, E. The who’s who of T-cell differentiation: Human memory T-cell subsets. Eur. J. Immunol. 2013, 43, 2797–2809. [Google Scholar] [CrossRef] [PubMed]
  197. Graham, B.S.; Bunton, L.A.; Wright, P.F.; Karzon, D.T. Role of T lymphocyte subsets in the pathogenesis of primary infection and rechallenge with respiratory syncytial virus in mice. J. Clin. Investig. 1991, 88, 1026–1033. [Google Scholar] [CrossRef]
  198. Jozwik, A.; Habibi, M.S.; Paras, A.; Zhu, J.; Guvenel, A.; Dhariwal, J.; Almond, M.; Wong, E.H.C.; Sykes, A.; Maybeno, M.; et al. RSV-specific airway resident memory CD8+ T cells and differential disease severity after experimental human infection. Nat. Commun. 2015, 6, 10224. [Google Scholar] [CrossRef] [PubMed]
  199. Rutigliano, J.A.; Graham, B.S. Prolonged Production of TNF-α Exacerbates Illness during Respiratory Syncytial Virus Infection. J. Immunol. 2004, 173, 3408–3417. [Google Scholar] [CrossRef] [Green Version]
  200. Aung, S.; Rutigliano, J.A.; Graham, B.S. Alternative Mechanisms of Respiratory Syncytial Virus Clearance in Perforin Knockout Mice Lead to Enhanced Disease. J. Virol. 2001, 75, 9918–9924. [Google Scholar] [CrossRef] [Green Version]
  201. Kinnear, E.; Lambert, L.; McDonald, J.U.; Cheeseman, H.M.; Caproni, L.J.; Tregoning, J.S. Airway T cells protect against RSV infection in the absence of antibody. Mucosal Immunol. 2017, 11, 249–256. [Google Scholar] [CrossRef] [Green Version]
  202. Morabito, K.; Ruckwardt, T.R.; Redwood, A.; Moin, S.; Price, D.; Graham, B.S. Intranasal administration of RSV antigen-expressing MCMV elicits robust tissue-resident effector and effector memory CD8+ T cells in the lung. Mucosal Immunol. 2016, 10, 545–554. [Google Scholar] [CrossRef] [Green Version]
  203. Lee, S.; Stokes, K.L.; Currier, M.G.; Sakamoto, K.; Lukacs, N.W.; Celis, E.; Moore, M.L. Vaccine-Elicited CD8+ T Cells Protect against Respiratory Syncytial Virus Strain A2-Line19F-Induced Pathogenesis in BALB/c Mice. J. Virol. 2012, 86, 13016–13024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. McDermott, D.S.; Knudson, C.J.; Varga, S.M. Determining the Breadth of the Respiratory Syncytial Virus-Specific T Cell Response. J. Virol. 2014, 88, 3135–3143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Jorquera, P.A.; Choi, Y.; Oakley, K.E.; Powell, T.J.; Boyd, J.G.; Palath, N.; Haynes, L.M.; Anderson, L.J.; Tripp, R.A. Nanoparticle Vaccines Encompassing the Respiratory Syncytial Virus (RSV) G Protein CX3C Chemokine Motif Induce Robust Immunity Protecting from Challenge and Disease. PLoS ONE 2013, 8, e74905. [Google Scholar] [CrossRef] [Green Version]
  206. Melendi, G.A.; Bridget, D.; Monsalvo, A.C.; Laham, F.F.; Acosta, P.; Delgado, M.F.; Polack, F.P.; Irusta, P.M. Conserved cysteine residues within the attachment G glycoprotein of respiratory syncytial virus play a critical role in the enhancement of cytotoxic T-lymphocyte responses. Virus Genes 2010, 42, 46–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Harcourt, J.; Alvarez, R.; Jones, L.P.; Henderson, C.; Anderson, L.J.; Tripp, R. Respiratory Syncytial Virus G Protein and G Protein CX3C Motif Adversely Affect CX3CR1+ T Cell Responses. J. Immunol. 2006, 176, 1600–1608. [Google Scholar] [CrossRef] [Green Version]
  208. Wu, C.; Xue, Y.; Wang, P.; Lin, L.; Liu, Q.; Li, N.; Xu, J.; Cao, X. IFN-γ Primes Macrophage Activation by Increasing Phosphatase and Tensin Homolog via Downregulation of miR-3473b. J. Immunol. 2014, 193, 3036–3044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Kashiwada, M.; Levy, D.M.; McKeag, L.; Murray, K.; Schröder, A.; Canfield, S.M.; Traver, G.; Rothman, P.B. IL-4-induced transcription factor NFIL3/E4BP4 controls IgE class switching. Proc. Natl. Acad. Sci. USA 2009, 107, 821–826. [Google Scholar] [CrossRef] [Green Version]
  210. Kouro, T.; Takatsu, K. IL-5- and eosinophil-mediated inflammation: From discovery to therapy. Int. Immunol. 2009, 21, 1303–1309. [Google Scholar] [CrossRef] [Green Version]
  211. Patel, S.S.; Casale, T.B.; Cardet, J.C. Biological therapies for eosinophilic asthma. Expert Opin. Biol. Ther. 2018, 18, 747–754. [Google Scholar] [CrossRef] [PubMed]
  212. Becker, Y. Respiratory syncytial virus (RSV) evades the human adaptive immune system by skewing the Th1/Th2 cytokine balance toward increased levels of Th2 cytokines and IgE, markers of allergy—A review. Virus Genes 2006, 33, 235–252. [Google Scholar] [CrossRef]
  213. Christiaansen, A.; Knudson, C.J.; Weiss, K.A.; Varga, S.M. The CD4 T cell response to respiratory syncytial virus infection. Immunol. Res. 2014, 59, 109–117. [Google Scholar] [CrossRef]
  214. Nenna, R.; Fedele, G.; Frassanito, A.; Petrarca, L.; Di Mattia, G.; Pierangeli, A.; Scagnolari, C.; Papoff, P.; Schiavoni, I.; Leone, P.; et al. Increased T-helper Cell 2 Response in Infants with Respiratory Syncytial Virus Bronchiolitis Hospitalized Outside Epidemic Peak. Pediatr. Infect. Dis. J. 2020, 39, 61–67. [Google Scholar] [CrossRef] [PubMed]
  215. Pinto, R.A.; Arredondo, S.M.; Bono, M.R.; Gaggero, A.A.; Diaz, P.V. T Helper 1/T Helper 2 Cytokine Imbalance in Respiratory Syncytial Virus Infection Is Associated with Increased Endogenous Plasma Cortisol. Pediatrics 2006, 117, e878–e886. [Google Scholar] [CrossRef] [Green Version]
  216. Siefker, D.T.; Vu, L.; You, D.; McBride, A.; Taylor, R.; Jones, T.L.; DeVincenzo, J.; Cormier, S.A. Respiratory Syncytial Virus Disease Severity Is Associated with Distinct CD8+ T-Cell Profiles. Am. J. Respir. Crit. Care Med. 2020, 201, 325–334. [Google Scholar] [CrossRef] [PubMed]
  217. Hussell, T.; Baldwin, C.J.; O’Garra, A.; Openshaw, P.J.M. CD8+ T cells control Th2-driven pathology during pulmonary respiratory syncytial virus infection. Eur. J. Immunol. 1997, 27, 3341–3349. [Google Scholar] [CrossRef]
  218. Liang, B.; Kabatova, B.; Kabat, J.; Dorward, D.W.; Liu, X.; Surman, S.; Liu, X.; Moseman, A.P.; Buchholz, U.J.; Collins, P.L.; et al. Effects of Alterations to the CX3C Motif and Secreted Form of Human Respiratory Syncytial Virus (RSV) G Protein on Immune Responses to a Parainfluenza Virus Vector Expressing the RSV G Protein. J. Virol. 2019, 93, e02043-18. [Google Scholar] [CrossRef] [Green Version]
  219. Chirkova, T.; Boyoglu-Barnum, S.; Gaston, K.A.; Malik, F.M.; Trau, S.; Oomens, A.G.P.; Anderson, L.J. Respiratory Syncytial Virus G Protein CX3C Motif Impairs Human Airway Epithelial and Immune Cell Responses. J. Virol. 2013, 87, 13466–13479. [Google Scholar] [CrossRef] [Green Version]
  220. Tripp, R.A.; Power, U.F.; Openshaw, P.J.M.; Kauvar, L.M. Respiratory Syncytial Virus: Targeting the G Protein Provides a New Approach for an Old Problem. J. Virol. 2018, 92, e01302-17. [Google Scholar] [CrossRef] [Green Version]
  221. Boyoglu-Barnum, S.; Todd, S.O.; Meng, J.; Barnum, T.R.; Chirkova, T.; Haynes, L.M.; Jadhao, S.J.; Tripp, R.A.; Oomens, A.G.; Moore, M.L.; et al. Mutating the CX3C Motif in the G Protein Should Make a Live Respiratory Syncytial Virus Vaccine Safer and More Effective. J. Virol. 2017, 91, e02059-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Ha, B.; Chirkova, T.; Boukhvalova, M.S.; Sun, H.Y.; Walsh, E.E.; Anderson, C.S.; Mariani, T.J.; Anderson, L.J. Mutation of Respiratory Syncytial Virus G Protein’s CX3C Motif Attenuates Infection in Cotton Rats and Primary Human Airway Epithelial Cells. Vaccines 2019, 7, 69. [Google Scholar] [CrossRef] [Green Version]
  223. Bergeron, H.; Murray, J.; Castrejon, A.N.; DuBois, R.; Tripp, R. Respiratory Syncytial Virus (RSV) G Protein Vaccines with Central Conserved Domain Mutations Induce CX3C-CX3CR1 Blocking Antibodies. Viruses 2021, 13, 352. [Google Scholar] [CrossRef]
  224. Hua, Y.; Jiao, Y.-Y.; Ma, Y.; Peng, X.-L.; Fu, Y.-H.; Zheng, Y.-P.; Hong, T.; He, J.-S. DNA vaccine encoding central conserved region of G protein induces Th1 predominant immune response and protection from RSV infection in mice. Immunol. Lett. 2016, 179, 95–101. [Google Scholar] [CrossRef]
  225. Jung, Y.-J.; Lee, Y.-N.; Kim, K.-H.; Lee, Y.; Jeeva, S.; Park, B.R.; Kang, S.-M. Recombinant Live Attenuated Influenza Virus Expressing Conserved G-Protein Domain in a Chimeric Hemagglutinin Molecule Induces G-Specific Antibodies and Confers Protection against Respiratory Syncytial Virus. Vaccines 2020, 8, 716. [Google Scholar] [CrossRef]
  226. Eichinger, K.M.; Kosanovich, J.L.; Gidwani, S.V.; Zomback, A.; Lipp, M.A.; Perkins, T.N.; Oury, T.D.; Petrovsky, N.; Marshall, C.P.; Yondola, M.A.; et al. Prefusion RSV F Immunization Elicits Th2-Mediated Lung Pathology in Mice When Formulated with a Th2 (but Not a Th1/Th2-Balanced) Adjuvant Despite Complete Viral Protection. Front. Immunol. 2020, 11, 1673. [Google Scholar] [CrossRef] [PubMed]
  227. Kosanovich, J.L.; Eichinger, K.M.; Lipp, M.A.; Yondola, M.A.; Perkins, T.N.; Empey, K.M. Formulation of the prefusion RSV F protein with a Th1/Th2-balanced adjuvant provides complete protection without Th2-skewed immunity in RSV-experienced young mice. Vaccine 2020, 38, 6357–6362. [Google Scholar] [CrossRef] [PubMed]
  228. Van Der Fits, L.; Bolder, R.; Der Meer, M.H.-V.; Drijver, J.; Van Polanen, Y.; Serroyen, J.; Langedijk, J.P.M.; Schuitemaker, H.; Saeland, E.; Zahn, R. Adenovector 26 encoded prefusion conformation stabilized RSV-F protein induces long-lasting Th1-biased immunity in neonatal mice. npj Vaccines 2020, 5, 49. [Google Scholar] [CrossRef]
  229. Schneider-Ohrum, K.; Cayatte, C.; Bennett, A.S.; Rajani, G.M.; McTamney, P.; Nacel, K.; Hostetler, L.; Cheng, L.; Ren, K.; O’Day, T.; et al. Immunization with Low Doses of Recombinant Postfusion or Prefusion Respiratory Syncytial Virus F Primes for Vaccine-Enhanced Disease in the Cotton Rat Model Independently of the Presence of a Th1-Biasing (GLA-SE) or Th2-Biasing (Alum) Adjuvant. J. Virol. 2017, 91, e02180-16. [Google Scholar] [CrossRef] [Green Version]
  230. Khan, I.U.; Ahmad, F.; Zhang, S.; Lu, P.; Wang, J.; Xie, J.; Zhu, N. Respiratory syncytial virus F and G protein core fragments fused to HBsAg-binding protein (SBP) induce a Th1-dominant immune response without vaccine-enhanced disease. Int. Immunol. 2018, 31, 199–209. [Google Scholar] [CrossRef]
  231. Hwang, H.S.; Kim, K.-H.; Lee, Y.; Lee, Y.-T.; Ko, E.-J.; Park, S.; Lee, J.S.; Lee, B.-C.; Kwon, Y.-M.; Moore, M.L.; et al. Virus-like particle vaccines containing F or F and G proteins confer protection against respiratory syncytial virus without pulmonary inflammation in cotton rats. Hum. Vaccines Immunother. 2017, 13, 1031–1039. [Google Scholar] [CrossRef]
  232. Li, N.; Zhang, L.; Zheng, B.; Li, W.; Liu, J.; Zhang, H.; Zeng, R. RSV recombinant candidate vaccine G1F/M2 with CpG as an adjuvant prevents vaccine-associated lung inflammation, which may be associated with the appropriate types of immune memory in spleens and lungs. Hum. Vaccines Immunother. 2019, 15, 2684–2694. [Google Scholar] [CrossRef] [PubMed]
  233. Jo, Y.-M.; Kim, J.; Chang, J. Vaccine containing G protein fragment and recombinant baculovirus expressing M2 protein induces protective immunity to respiratory syncytial virus. Clin. Exp. Vaccine Res. 2019, 8, 43–53. [Google Scholar] [CrossRef]
  234. Green, C.A.; Scarselli, E.; Sande, C.J.; Thompson, A.J.; De Lara, C.M.; Taylor, K.S.; Haworth, K.; Del Sorbo, M.; Angus, B.; Siani, L.; et al. Chimpanzee adenovirus- and MVA-vectored respiratory syncytial virus vaccine is safe and immunogenic in adults. Sci. Transl. Med. 2015, 7, 300ra126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Park, S.; Lee, Y.; Kwon, Y.-M.; Lee, Y.-T.; Kim, K.-H.; Ko, E.-J.; Jung, J.H.; Song, M.; Graham, B.; Prausnitz, M.; et al. Vaccination by microneedle patch with inactivated respiratory syncytial virus and monophosphoryl lipid A enhances the protective efficacy and diminishes inflammatory disease after challenge. PLoS ONE 2018, 13, e0205071. [Google Scholar] [CrossRef]
  236. Lee, Y.; Lee, Y.-T.; Ko, E.-J.; Kim, K.-H.; Hwang, H.S.; Park, S.; Kwon, Y.-M.; Kang, S.M. Soluble F proteins exacerbate pulmonary histopathology after vaccination upon respiratory syncytial virus challenge but not when presented on virus-like particles. Hum. Vaccines Immunother. 2017, 13, 2594–2605. [Google Scholar] [CrossRef] [PubMed]
  237. Sjaastad, L.E.; Owen, D.L.; Tracy, S.I.; Farrar, M.A. Phenotypic and Functional Diversity in Regulatory T Cells. Front. Cell Dev. Biol. 2021, 9, 715901. [Google Scholar] [CrossRef]
  238. Shao, H.-Y.; Huang, J.-Y.; Lin, Y.-W.; Yu, S.-L.; Chitra, E.; Chang, C.-K.; Sung, W.-C.; Chong, P.; Chow, Y.-H. Depletion of regulatory T-cells leads to moderate B-cell antigenicity in respiratory syncytial virus infection. Int. J. Infect. Dis. 2015, 41, 56–64. [Google Scholar] [CrossRef] [Green Version]
  239. Christiaansen, A.; Syed, M.A.; Eyck, P.P.T.; Hartwig, S.M.; Durairaj, L.; Kamath, S.S.; Varga, S.M. Altered Treg and cytokine responses in RSV-infected infants. Pediatr. Res. 2016, 80, 702–709. [Google Scholar] [CrossRef]
  240. Loebbermann, J.; Durant, L.; Thornton, H.; Johansson, C.; Openshaw, P.J. Defective immunoregulation in RSV vaccine-augmented viral lung disease restored by selective chemoattraction of regulatory T cells. Proc. Natl. Acad. Sci. USA 2013, 110, 2987–2992. [Google Scholar] [CrossRef] [Green Version]
  241. Fan, P.; Liu, Z.; Zheng, M.; Chen, M.; Xu, Y.; Zhao, D. Respiratory syncytial virus nonstructural protein 1 breaks immune tolerance in mice by downregulating Tregs through TSLP-OX40/OX40L-mTOR axis. Mol. Immunol. 2021, 138, 20–30. [Google Scholar] [CrossRef] [PubMed]
  242. Yang, P.; Zheng, J.; Wang, S.; Liu, P.; Xie, M.; Zhao, D. Respiratory syncytial virus nonstructural proteins 1 and 2 are crucial pathogenic factors that modulate interferon signaling and Treg cell distribution in mice. Virology 2015, 485, 223–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Han, J.; Dakhama, A.; Jia, Y.; Wang, M.; Zeng, W.; Takeda, K.; Shiraishi, Y.; Okamoto, M.; Ziegler, S.F.; Gelfand, E.W. Responsiveness to respiratory syncytial virus in neonates is mediated through thymic stromal lymphopoietin and OX40 ligand. J. Allergy Clin. Immunol. 2012, 130, 1175–1186.e9. [Google Scholar] [CrossRef] [Green Version]
  244. Fulton, R.B.; Meyerholz, D.; Varga, S.M. Foxp3+ CD4 Regulatory T Cells Limit Pulmonary Immunopathology by Modulating the CD8 T Cell Response during Respiratory Syncytial Virus Infection. J. Immunol. 2010, 185, 2382–2392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Loebbermann, J.; Thornton, H.; Durant, L.; Sparwasser, T.; Webster, K.E.; Sprent, J.; Culley, F.; Johansson, C.; Openshaw, P.J. Regulatory T cells expressing granzyme B play a critical role in controlling lung inflammation during acute viral infection. Mucosal Immunol. 2012, 5, 161–172. [Google Scholar] [CrossRef] [Green Version]
  246. Durant, L.; Makris, S.; Voorburg, C.M.; Loebbermann, J.; Johansson, C.; Openshaw, P.J.M. Regulatory T Cells Prevent Th2 Immune Responses and Pulmonary Eosinophilia during Respiratory Syncytial Virus Infection in Mice. J. Virol. 2013, 87, 10946–10954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Pestka, S.; Krause, C.D.; Walter, M.R. Interferons, interferon-like cytokines, and their receptors. Immunol. Rev. 2004, 202, 8–32. [Google Scholar] [CrossRef]
  248. Schroder, K.; Hertzog, P.J.; Ravasi, T.; Hume, D.A. Interferon-γ: An overview of signals, mechanisms and functions. J. Leukoc. Biol. 2004, 75, 163–189. [Google Scholar] [CrossRef]
  249. Turi, K.N.; Shankar, J.; Anderson, L.J.; Rajan, D.; Gaston, K.; Gebretsadik, T.; Das, S.R.; Stone, C.; Larkin, E.K.; Rosas-Salazar, C.; et al. Infant Viral Respiratory Infection Nasal Immune-Response Patterns and Their Association with Subsequent Childhood Recurrent Wheeze. Am. J. Respir. Crit. Care Med. 2018, 198, 1064–1073. [Google Scholar] [CrossRef] [PubMed]
  250. Zanoni, I.; Granucci, F.; Broggi, A. Interferon (IFN)-λ Takes the Helm: Immunomodulatory Roles of Type III IFNs. Front. Immunol. 2017, 8, 1661. [Google Scholar] [CrossRef]
  251. De Weerd, N.A.; Vivian, J.P.; Lim, S.S.; Huang, S.U.; Hertzog, P.J. Structural integrity with functional plasticity: What type I IFN receptor polymorphisms reveal. J. Leukoc. Biol. 2020, 108, 909–924. [Google Scholar] [CrossRef] [PubMed]
  252. Platanias, L.C. Signaling pathways activated by interferons. Exp. Hematol. 1999, 27, 1583–1592. [Google Scholar] [CrossRef]
  253. Tovey, M.G.; Lallemand, C. Safety, Tolerability, and Immunogenicity of Interferons. Pharmaceuticals 2010, 3, 1162–1186. [Google Scholar] [CrossRef] [Green Version]
  254. Starr, R.; Willson, T.A.; Viney, E.M.; Murray, L.J.L.; Rayner, J.R.; Jenkins, B.; Gonda, T.J.; Alexander, W.S.; Metcalf, D.; Nicola, N.; et al. A family of cytokine-inducible inhibitors of signalling. Nature 1997, 387, 917–921. [Google Scholar] [CrossRef] [PubMed]
  255. Liau, N.P.D.; Laktyushin, A.; Lucet, I.S.; Murphy, J.M.; Yao, S.; Whitlock, E.; Callaghan, K.; Nicola, N.A.; Kershaw, N.J.; Babon, J.J. The molecular basis of JAK/STAT inhibition by SOCS1. Nat. Commun. 2018, 9, 1–14. [Google Scholar] [CrossRef]
  256. Oshansky, C.M.; Krunkosky, T.M.; Barber, J.; Jones, L.P.; Tripp, R.A. Respiratory Syncytial Virus Proteins Modulate Suppressors of Cytokine Signaling 1 and 3 and the Type I Interferon Response to Infection by a Toll-Like Receptor Pathway. Viral Immunol. 2009, 22, 147–161. [Google Scholar] [CrossRef] [PubMed]
  257. Moore, E.C.; Barber, J.; Tripp, R.A. Respiratory syncytial virus (RSV) attachment and nonstructural proteins modify the type I interferon response associated with suppressor of cytokine signaling (SOCS) proteins and IFN-stimulated gene-15 (ISG15). Virol. J. 2008, 5, 116. [Google Scholar] [CrossRef] [Green Version]
  258. Sedeyn, K.; Schepens, B.; Saelens, X. Respiratory syncytial virus nonstructural proteins 1 and 2: Exceptional disrupters of innate immune responses. PLoS Pathog. 2019, 15, e1007984. [Google Scholar] [CrossRef] [PubMed]
  259. Spann, K.M.; Tran, K.-C.; Chi, B.; Rabin, R.L.; Collins, P.L. 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. J. Virol. 2004, 78, 4363–4369. [Google Scholar] [CrossRef] [Green Version]
  260. Thornhill, E.M.; Verhoeven, D. Respiratory Syncytial Virus’s Non-structural Proteins: Masters of Interference. Front. Cell. Infect. Microbiol. 2020, 10, 225. [Google Scholar] [CrossRef]
  261. Hijano, D.R.; Siefker, D.T.; Shrestha, B.; Jaligama, S.; Vu, L.D.; Tillman, H.; Finkelstein, D.; Saravia, J.; You, D.; Cormier, S.A. Type I Interferon Potentiates IgA Immunity to Respiratory Syncytial Virus Infection During Infancy. Sci. Rep. 2018, 8, 11034. [Google Scholar] [CrossRef]
  262. Thwaites, R.; Coates, M.; Ito, K.; Ghazaly, M.; Feather, C.; Abdulla, F.; Tunstall, T.; Jain, P.; Cass, L.; Rapeport, G.; et al. Reduced Nasal Viral Load and IFN Responses in Infants with Respiratory Syncytial Virus Bronchiolitis and Respiratory Failure. Am. J. Respir. Crit. Care Med. 2018, 198, 1074–1084. [Google Scholar] [CrossRef]
  263. 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. Biochem. Biophys. Res. Commun. 2016, 478, 1436–1441. [Google Scholar] [CrossRef]
  264. Stanifer, M.L.; Pervolaraki, K.; Boulant, S. Differential Regulation of Type I and Type III Interferon Signaling. Int. J. Mol. Sci. 2019, 20, 1445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Lozhkov, A.A.; Klotchenko, S.A.; Ramsay, E.S.; Moshkoff, H.D.; Moshkoff, D.A.; Vasin, A.V.; Salvato, M.S. The Key Roles of Interferon Lambda in Human Molecular Defense against Respiratory Viral Infections. Pathogens 2020, 9, 989. [Google Scholar] [CrossRef]
  266. Lazear, H.M.; Schoggins, J.W.; Diamond, M.S. Shared and Distinct Functions of Type I and Type III Interferons. Immunity 2019, 50, 907–923. [Google Scholar] [CrossRef] [PubMed]
  267. Salka, K.; Arroyo, M.; Chorvinsky, E.; Abutaleb, K.; Perez, G.F.; Wolf, S.; Xuchen, X.; Weinstock, J.; Gutierrez, M.J.; Pérez-Losada, M.; et al. Innate IFN-lambda responses to dsRNA in the human infant airway epithelium and clinical regulatory factors during viral respiratory infections in early life. Clin. Exp. Allergy 2020, 50, 1044–1054. [Google Scholar] [CrossRef]
  268. Selvaggi, C.; Pierangeli, A.; Fabiani, M.; Spano, L.; Nicolai, A.; Papoff, P.; Moretti, C.; Midulla, F.; Antonelli, G.; Scagnolari, C. Interferon lambda 1–3 expression in infants hospitalized for RSV or HRV associated bronchiolitis. J. Infect. 2014, 68, 467–477. [Google Scholar] [CrossRef] [PubMed]
  269. Okabayashi, T.; Kojima, T.; Masaki, T.; Yokota, S.-I.; Imaizumi, T.; Tsutsumi, H.; Himi, T.; Fujii, N.; Sawada, N. Type-III interferon, not type-I, is the predominant interferon induced by respiratory viruses in nasal epithelial cells. Virus Res. 2011, 160, 360–366. [Google Scholar] [CrossRef]
  270. Werder, R.B.; Lynch, J.P.; Simpson, J.C.; Zhang, V.; Hodge, N.H.; Poh, M.; Forbes-Blom, E.; Kulis, C.; Smythe, M.L.; Upham, J.W.; et al. PGD2/DP2 receptor activation promotes severe viral bronchiolitis by suppressing IFN-λ production. Sci. Transl. Med. 2018, 10, eaao0052. [Google Scholar] [CrossRef] [Green Version]
  271. Kalinowski, A.; Galen, B.T.; Ueki, I.F.; Sun, Y.; Mulenos, A.; Osafo-Addo, A.; Clark, B.; Joerns, J.; Liu, W.; Nadel, J.A.; et al. Respiratory syncytial virus activates epidermal growth factor receptor to suppress interferon regulatory factor 1-dependent interferon-lambda and antiviral defense in airway epithelium. Mucosal Immunol. 2018, 11, 958–967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Mo, S.; Tang, W.; Xie, J.; Chen, S.; Ren, L.; Zang, N.; Xie, X.; Deng, Y.; Gao, L.; Liu, E. Respiratory syncytial virus activates Rab5a to suppress IRF1-dependent IFN-lambda production, subverting the antiviral defense of airway epithelial cells. J. Virol. 2021, 95, e02333-20. [Google Scholar] [CrossRef]
  273. Bakre, A.A.; Harcourt, J.L.; Haynes, L.M.; Anderson, L.J.; Tripp, R.A. The Central Conserved Region (CCR) of Respiratory Syncytial Virus (RSV) G Protein Modulates Host miRNA Expression and Alters the Cellular Response to Infection. Vaccines 2017, 5, 16. [Google Scholar] [CrossRef] [PubMed]
  274. Bakre, A.; Mitchell, P.; Coleman, J.K.; Jones, L.P.; Saavedra, G.; Teng, M.; Tompkins, S.; Tripp, R.A. Respiratory syncytial virus modifies microRNAs regulating host genes that affect virus replication. J. Gen. Virol. 2012, 93, 2346–2356. [Google Scholar] [CrossRef]
  275. Taylor, G. Animal models of respiratory syncytial virus infection. Vaccine 2016, 35, 469–480. [Google Scholar] [CrossRef] [Green Version]
  276. Sitthicharoenchai, P.; Alnajjar, S.; Ackermann, M.R. A model of respiratory syncytial virus (RSV) infection of infants in newborn lambs. Cell Tissue Res. 2020, 380, 313–324. [Google Scholar] [CrossRef] [PubMed]
  277. Guerra-Maupome, M.; Palmer, M.V.; McGill, J.L.; Sacco, R.E. Utility of the Neonatal Calf Model for Testing Vaccines and Intervention Strategies for Use against Human RSV Infection. Vaccines 2019, 7, 7. [Google Scholar] [CrossRef] [Green Version]
  278. Dyer, K.D.; Garcia-Crespo, K.E.; Glineur, S.; Domachowske, J.B.; Rosenberg, H.F. The Pneumonia Virus of Mice (PVM) Model of Acute Respiratory Infection. Viruses 2012, 4, 3494–3510. [Google Scholar] [CrossRef]
  279. Larsen, L.E. Bovine Respiratory Syncytial Virus (BRSV): A review. Acta Vet. Scand. 2000, 41, 1–24. [Google Scholar] [CrossRef]
  280. Prince, G.A.; Horswood, R.L.; Berndt, J.; Suffin, S.C.; Chanock, R.M. Respiratory syncytial virus infection in inbred mice. Infect. Immun. 1979, 26, 764–766. [Google Scholar] [CrossRef] [Green Version]
  281. Sellers, R.S.; Clifford, C.B.; Treuting, P.M.; Brayton, C. Immunological Variation Between Inbred Laboratory Mouse Strains: Points to consider in phenotyping genetically immunomodified mice. Vet. Pathol. 2011, 49, 32–43. [Google Scholar] [CrossRef]
  282. Graham, B.S.; Perkins, M.D.; Wright, P.F.; Karzon, D.T. Primary respiratory syncytial virus infection in mice. J. Med. Virol. 1988, 26, 153–162. [Google Scholar] [CrossRef] [PubMed]
  283. Openshaw, P.J. The Mouse Model of Respiratory Syncytial Virus Disease. Curr. Top. Microbiol. Immunol. 2013, 372, 359–369. [Google Scholar] [CrossRef]
  284. Openshaw, P.J.M. Immunity and Immunopathology to Respiratory Syncytial Virus. Am. J. Respir. Crit. Care Med. 1995, 152, S59–S62. [Google Scholar] [CrossRef]
  285. Sharma, A.; Wu, W.; Sung, B.; Huang, J.; Tsao, T.; Li, X.; Gomi, R.; Tsuji, M.; Worgall, S. Respiratory Syncytial Virus (RSV) Pulmonary Infection in Humanized Mice Induces Human Anti-RSV Immune Responses and Pathology. J. Virol. 2016, 90, 5068–5074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. You, D.; Siefker, D.T.; Shrestha, B.; Saravia, J.; Cormier, S.A. Building a better neonatal mouse model to understand infant respiratory syncytial virus disease. Respir. Res. 2015, 16, 91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Moore, M.L.; Chi, M.H.; Luongo, C.; Lukacs, N.W.; Polosukhin, V.V.; Huckabee, M.M.; Newcomb, D.C.; Buchholz, U.J.; Crowe, J.E.; Goleniewska, K.; et al. A Chimeric A2 Strain of Respiratory Syncytial Virus (RSV) with the Fusion Protein of RSV Strain Line 19 Exhibits Enhanced Viral Load, Mucus, and Airway Dysfunction. J. Virol. 2009, 83, 4185–4194. [Google Scholar] [CrossRef] [Green Version]
  288. Boyoglu-Barnum, S.; Chirkova, T.; Todd, S.O.; Barnum, T.R.; Gaston, K.A.; Jorquera, P.; Haynes, L.M.; Tripp, R.A.; Moore, M.L.; Anderson, L.J. Prophylaxis with a Respiratory Syncytial Virus (RSV) Anti-G Protein Monoclonal Antibody Shifts the Adaptive Immune Response to RSV rA2-line19F Infection from Th2 to Th1 in BALB/c Mice. J. Virol. 2014, 88, 10569–10583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Johnson, S.; Griego, S.D.; Pfarr, D.S.; Doyle, M.L.; Woods, R.; Carlin, D.; Prince, G.A.; Koenig, S.; Young, J.F.; Dillon, S.B. A Direct Comparison of the Activities of Two Humanized Respiratory Syncytial Virus Monoclonal Antibodies: MEDI-493 and RSHZl9. J. Infect. Dis. 1999, 180, 35–40. [Google Scholar] [CrossRef] [Green Version]
  290. Llorens, X.S.; Castaño, E.; Null, D.; Steichen, J.; Sánchez, P.J.; Ramilo, O.; Top, F.H.; Connor, E. Safety and pharmacokinetics of an intramuscular humanized monoclonal antibody to respiratory syncytial virus in premature infants and infants with bronchopulmonary dysplasia. Pediatr. Infect. Dis. J. 1998, 17, 787–791. [Google Scholar] [CrossRef]
  291. Boukhvalova, M.S.; Blanco, J.C.G. The Cotton Rat Sigmodon Hispidus Model of Respiratory Syncytial Virus Infection. Curr. Top. Microbiol. Immunol. 2013, 372, 347–358. [Google Scholar] [CrossRef]
  292. Boukhvalova, M.S.; Prince, G.A.; Blanco, J.C. The cotton rat model of respiratory viral infections. Biologicals 2009, 37, 152–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Guichelaar, T.; van Erp, E.A.; Hoeboer, J.; Smits, N.A.; van Els, C.A.; Pieren, D.K.; Luytjes, W. Diversity of aging of the immune system classified in the cotton rat (Sigmodon hispidus) model of human infectious diseases. Dev. Comp. Immunol. 2018, 82, 39–48. [Google Scholar] [CrossRef] [PubMed]
  294. Wen, X.; Mo, S.; Chen, S.; Yu, G.; Gao, L.; Chen, S.; Deng, Y.; Xie, X.; Zang, N.; Ren, L.; et al. Pathogenic difference of respiratory syncytial virus infection in cotton rats of different ages. Microb. Pathog. 2019, 137, 103749. [Google Scholar] [CrossRef]
  295. Martinez, M.E.; Harder, O.E.; Rosas, L.E.; Joseph, L.; Davis, I.C.; Niewiesk, S. Pulmonary function analysis in cotton rats after respiratory syncytial virus infection. PLoS ONE 2020, 15, e0237404. [Google Scholar] [CrossRef] [PubMed]
  296. Green, M.G.; Petroff, N.; La Perle, K.M.D.; Niewiesk, S. Characterization of Cotton Rat (Sigmodon hispidus) Eosinophils, Including Their Response to Respiratory Syncytial Virus Infection. Comp. Med. 2018, 68, 31–40. [Google Scholar]
  297. Martinez, M.E.; Niewiesk, S.; La Perle, K.M.D. Cotton Rat Placenta Anatomy and Fc Receptor Expression and Their Roles in Maternal Antibody Transfer. Comp. Med. 2020, 70, 510–519. [Google Scholar] [CrossRef] [PubMed]
  298. Blanco, J.C.; Pletneva, L.M.; Otoa, R.O.; Patel, M.C.; Vogel, S.N.; Boukhvalova, M.S. Preclinical assessment of safety of maternal vaccination against respiratory syncytial virus (RSV) in cotton rats. Vaccine 2017, 35, 3951–3958. [Google Scholar] [CrossRef]
  299. Blanco, J.C.G.; Pletneva, L.M.; McGinnes-Cullen, L.; Otoa, R.O.; Patel, M.C.; Fernando, L.R.; Boukhvalova, M.S.; Morrison, T.G. Efficacy of a respiratory syncytial virus vaccine candidate in a maternal immunization model. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef]
  300. Shao, H.-Y.; Chen, Y.-C.; Chung, N.-H.; Lu, Y.-J.; Chang, C.-K.; Yu, S.-L.; Liu, C.-C.; Chow, Y.-H. Maternal immunization with a recombinant adenovirus-expressing fusion protein protects neonatal cotton rats from respiratory syncytia virus infection by transferring antibodies via breast milk and placenta. Virology 2018, 521, 181–189. [Google Scholar] [CrossRef] [PubMed]
  301. Green, G.; Johnson, S.M.; Costello, H.; Brakel, K.; Harder, O.; Oomens, A.G.; Peeples, M.E.; Moulton, H.M.; Niewiesk, S. CX3CR1 Is a Receptor for Human Respiratory Syncytial Virus in Cotton Rats. J. Virol. 2021, 95, e00010-21. [Google Scholar] [CrossRef]
  302. Morris, J.A.; Blount, R.E.; Savage, R.E. Recovery of Cytopathogenic Agent from Chimpanzees with Goryza. Exp. Biol. Med. 1956, 92, 544–549. [Google Scholar] [CrossRef] [PubMed]
  303. Ispas, G.; Koul, A.; Verbeeck, J.; Sheehan, J.; Sanders-Beer, B.; Roymans, D.; Andries, K.; Rouan, M.-C.; De Jonghe, S.; Bonfanti, J.-F.; et al. Antiviral Activity of TMC353121, a Respiratory Syncytial Virus (RSV) Fusion Inhibitor, in a Non-Human Primate Model. PLoS ONE 2015, 10, e0126959. [Google Scholar] [CrossRef] [PubMed]
  304. Marcandalli, J.; Fiala, B.; Ols, S.; Perotti, M.; de van der Schueren, W.; Snijder, J.; Hodge, E.; Benhaim, M.; Ravichandran, R.; Carter, L.; et al. Induction of Potent Neutralizing Antibody Responses by a Designed Protein Nanoparticle Vaccine for Respiratory Syncytial Virus. Cell 2019, 176, 1420–1431.e17. [Google Scholar] [CrossRef] [Green Version]
  305. Swanson, K.A.; Rainho-Tomko, J.N.; Williams, Z.P.; Lanza, L.; Peredelchuk, M.; Kishko, M.; Pavot, V.; Alamares-Sapuay, J.; Adhikarla, H.; Gupta, S.; et al. A respiratory syncytial virus (RSV) F protein nanoparticle vaccine focuses antibody responses to a conserved neutralization domain. Sci. Immunol. 2020, 5, eaba6466. [Google Scholar] [CrossRef]
  306. Sesterhenn, F.; Yang, C.; Bonet, J.; Cramer, J.T.; Wen, X.; Wang, Y.; Chiang, C.-I.; Abriata, L.A.; Kucharska, I.; Castoro, G.; et al. De novo protein design enables the precise induction of RSV-neutralizing antibodies. Science 2020, 368, eaay5051. [Google Scholar] [CrossRef]
  307. Pierantoni, A.; Esposito, M.L.; Ammendola, V.; Napolitano, F.; Grazioli, F.; Abbate, A.; del Sorbo, M.; Siani, L.; D’Alise, A.M.; Taglioni, A.; et al. Mucosal delivery of a vectored RSV vaccine is safe and elicits protective immunity in rodents and nonhuman primates. Mol. Ther. Methods Clin. Dev. 2015, 2, 15018. [Google Scholar] [CrossRef]
  308. Mueller, S.; Stauft, C.B.; Kalkeri, R.; Koidei, F.; Kushnir, A.; Tasker, S.; Coleman, J.R. A codon-pair deoptimized live-attenuated vaccine against respiratory syncytial virus is immunogenic and efficacious in non-human primates. Vaccine 2020, 38, 2943–2948. [Google Scholar] [CrossRef]
  309. Buchwald, A.G.; Graham, B.S.; Traore, A.; Haidara, F.C.; Chen, M.; Morabito, K.; Lin, B.C.; Sow, S.O.; Levine, M.M.; Pasetti, M.F.; et al. RSV neutralizing antibodies at birth predict protection from RSV illness in infants in the first three months of life. Clin. Infect. Dis. 2020, 73, e4421–e4427. [Google Scholar] [CrossRef]
  310. Piedra, P.A.; Jewell, A.M.; Cron, S.G.; Atmar, R.L.; Glezen, W.P. Correlates of immunity to respiratory syncytial virus (RSV) associated-hospitalization: Establishment of minimum protective threshold levels of serum neutralizing antibodies. Vaccine 2003, 21, 3479–3482. [Google Scholar] [CrossRef]
  311. Capella, C.; Chaiwatpongsakorn, S.; Gorrell, E.; Risch, Z.A.; Ye, F.; Mertz, S.E.; Johnson, S.M.; Moore-Clingenpeel, M.; Ramilo, O.; Mejias, A.; et al. Prefusion F, Postfusion F, G Antibodies, and Disease Severity in Infants and Young Children with Acute Respiratory Syncytial Virus Infection. J. Infect. Dis. 2017, 216, 1398–1406. [Google Scholar] [CrossRef]
  312. Welliver, R.C.; Papin, J.F.; Preno, A.; Ivanov, V.; Tian, J.-H.; Lu, H.; Guebre-Xabier, M.; Flyer, D.; Massare, M.J.; Glenn, G.; et al. Maternal immunization with RSV fusion glycoprotein vaccine and substantial protection of neonatal baboons against respiratory syncytial virus pulmonary challenge. Vaccine 2019, 38, 1258–1270. [Google Scholar] [CrossRef]
  313. Stensballe, L.G.; Ravn, H.; Kristensen, K.; Agerskov, K.; Meakins, T.; Aaby, P.; Simões, E.A. Respiratory syncytial virus neutralizing antibodies in cord blood, respiratory syncytial virus hospitalization, and recurrent wheeze. J. Allergy Clin. Immunol. 2009, 123, 398–403. [Google Scholar] [CrossRef] [PubMed]
  314. Mazur, N.I.; Horsley, N.M.; Englund, J.A.; Nederend, M.; Magaret, A.; Kumar, A.; Jacobino, S.R.; de Haan, C.; Khatry, S.K.; LeClerq, S.C.; et al. Breast Milk Prefusion F Immunoglobulin G as a Correlate of Protection Against Respiratory Syncytial Virus Acute Respiratory Illness. J. Infect. Dis. 2018, 219, 59–67. [Google Scholar] [CrossRef]
  315. Zambon, M. Active and passive immunisation against respiratory syncytial virus. Rev. Med. Virol. 1999, 9, 227–236. [Google Scholar] [CrossRef]
  316. Kulkarni, P.S.; Hurwitz, J.L.; Simões, E.A.; Piedra, P.A. Establishing Correlates of Protection for Vaccine Development: Considerations for the Respiratory Syncytial Virus Vaccine Field. Viral Immunol. 2018, 31, 195–203. [Google Scholar] [CrossRef]
  317. Vázquez, Y.; González, L.; Noguera, L.; Gonzalez, P.A.; Riedel, C.; Bertrand, P.; Bueno, S.M. Cytokines in the Respiratory Airway as Biomarkers of Severity and Prognosis for Respiratory Syncytial Virus Infection: An Update. Front. Immunol. 2019, 10, 1154. [Google Scholar] [CrossRef]
  318. Rajan, D.; O’Keefe, E.L.; Travers, C.; McCracken, C.; Geoghegan, S.; Caballero, M.T.; Acosta, P.L.; Polack, F.; Anderson, L.J. MUC5AC Levels Associated with Respiratory Syncytial Virus Disease Severity. Clin. Infect. Dis. 2018, 67, 1441–1444. [Google Scholar] [CrossRef] [PubMed]
  319. Han, Z.; Rao, J.; Xie, Z.; Wang, C.; Xu, B.; Qian, S.; Wang, Y.; Zhu, J.; Yang, B.; Xu, F.; et al. Chemokine (C-X-C Motif) Ligand 4 Is a Restrictor of Respiratory Syncytial Virus Infection and an Indicator of Clinical Severity. Am. J. Respir. Crit. Care Med. 2020, 202, 717–729. [Google Scholar] [CrossRef]
  320. Wynn, T.A.; Chawla, A.; Pollard, J.W. Macrophage biology in development, homeostasis and disease. Nature 2013, 496, 445–455. [Google Scholar] [CrossRef]
  321. Mills, C.D.; Kincaid, K.; Alt, J.M.; Heilman, M.J.; Hill, A.M. M-1/M-2 Macrophages and the Th1/Th2 Paradigm. J. Immunol. 2000, 164, 6166–6173. [Google Scholar] [CrossRef] [Green Version]
  322. Ruytinx, P.; Proost, P.; Van Damme, J.; Struyf, S. Chemokine-Induced Macrophage Polarization in Inflammatory Conditions. Front. Immunol. 2018, 9, 1930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Nikonova, A.A.; Pichugin, A.V.; Chulkina, M.; Lebedeva, E.S.; Gaisina, A.R.; Shilovskiy, I.P.; Ataullakhanov, R.I.; Khaitov, M.R.; Khaitov, R.M. The TLR4 Agonist Immunomax Affects the Phenotype of Mouse Lung Macrophages during Respiratory Syncytial Virus Infection. Acta Nat. 2018, 10, 95–99. [Google Scholar] [CrossRef]
  324. Keegan, A.D.; Shirey, K.A.; Bagdure, D.; Blanco, J.; Viscardi, R.M.; Vogel, S.N. Enhanced allergic responsiveness after early childhood infection with respiratory viruses: Are long-lived alternatively activated macrophages the missing link? Pathog. Dis. 2016, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  325. Srinivasa, B.T.; Restori, K.H.; Shan, J.; Cyr, L.; Xing, L.; Lee, S.; Ward, B.J.; Fixman, E.D. STAT6 inhibitory peptide given during RSV infection of neonatal mice reduces exacerbated airway responses upon adult reinfection. J. Leukoc. Biol. 2016, 101, 519–529. [Google Scholar] [CrossRef]
  326. Shirey, K.A.; Pletneva, L.M.; Puche, A.C.; Keegan, A.D.; Prince, G.A.; Blanco, J.; Vogel, S.N. Control of RSV-induced lung injury by alternatively activated macrophages is IL-4Rα-, TLR4-, and IFN-β-dependent. Mucosal Immunol. 2010, 3, 291–300. [Google Scholar] [CrossRef]
  327. Shirey, K.A.; Lai, W.; Pletneva, L.M.; Finkelman, F.D.; Feola, D.J.; Blanco, J.C.G.; Vogel, S.N. Agents that increase AAM differentiation blunt RSV-mediated lung pathology. J. Leukoc. Biol. 2014, 96, 951–955. [Google Scholar] [CrossRef] [Green Version]
  328. Shirey, K.A.; Lai, W.; Pletneva, L.M.; Karp, C.; Divanovic, S.; Blanco, J.; Vogel, S.N. Role of the lipoxygenase pathway in RSV-induced alternatively activated macrophages leading to resolution of lung pathology. Mucosal Immunol. 2013, 7, 549–557. [Google Scholar] [CrossRef] [Green Version]
  329. Tripp, R.A. Respiratory Syncytial Virus (RSV) Modulation at the Virus-Host Interface Affects Immune Outcome and Disease Pathogenesis. Immune Netw. 2013, 13, 163–167. [Google Scholar] [CrossRef] [Green Version]
  330. Tripp, R.A. Pathogenesis of Respiratory Syncytial Virus Infection. Viral Immunol. 2004, 17, 165–181. [Google Scholar] [CrossRef] [PubMed]
  331. Piedra, F.-A.; Mei, M.; Avadhanula, V.; Mehta, R.; Aideyan, L.; Garofalo, R.P.; Piedra, P.A. The interdependencies of viral load, the innate immune response, and clinical outcome in children presenting to the emergency department with respiratory syncytial virus-associated bronchiolitis. PLoS ONE 2017, 12, e0172953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Lui, G.; Wong, C.; Chan, M.; Chong, K.; Wong, R.; Chu, I.; Zhang, M.; Li, T.; Hui, D.; Lee, N.; et al. Host inflammatory response is the major marker of severe respiratory syncytial virus infection in older adults. J. Infect. 2021. [Google Scholar] [CrossRef] [PubMed]
  333. Taleb, S.A.; Al-Ansari, K.; Nasrallah, G.K.; Elrayess, M.A.; Al-Thani, A.A.; Derrien-Colemyn, A.; Ruckwardt, T.J.; Graham, B.S.; Yassine, H.M. Level of maternal respiratory syncytial virus (RSV) F antibodies in hospitalized children and correlates of protection. Int. J. Infect. Dis. 2021, 109, 56–62. [Google Scholar] [CrossRef]
  334. Van Erp, E.A.; Luytjes, W.; Ferwerda, G.; Van Kasteren, P.B. Fc-Mediated Antibody Effector Functions During Respiratory Syncytial Virus Infection and Disease. Front. Immunol. 2019, 10, 548. [Google Scholar] [CrossRef] [Green Version]
  335. Polack, F.P.; Teng, M.N.; Collins, P.L.; Prince, G.A.; Exner, M.; Regele, H.; Lirman, D.D.; Rabold, R.; Hoffman, S.J.; Karp, C.; et al. A Role for Immune Complexes in Enhanced Respiratory Syncytial Virus Disease. J. Exp. Med. 2002, 196, 859–865. [Google Scholar] [CrossRef]
  336. Tripp, R.A.; Power, U.F. Original Antigenic Sin and Respiratory Syncytial Virus Vaccines. Vaccines 2019, 7, 107. [Google Scholar] [CrossRef] [Green Version]
  337. Feng, S.; Zeng, D.; Zheng, J.; Zhao, D. MicroRNAs: Mediators and Therapeutic Targets to Airway Hyper Reactivity After Respiratory Syncytial Virus Infection. Front. Microbiol. 2018, 9, 2177. [Google Scholar] [CrossRef]
  338. Atherton, L.J.; Jorquera, P.A.; Bakre, A.A.; Tripp, R.A. Determining Immune and miRNA Biomarkers Related to Respiratory Syncytial Virus (RSV) Vaccine Types. Front. Immunol. 2019, 10, 2323. [Google Scholar] [CrossRef] [PubMed]
  339. Eilam-Frenkel, B.; Naaman, H.; Brkic, G.; Veksler-Lublinsky, I.; Rall, G.; Shemer-Avni, Y.; Gopas, J. MicroRNA 146-5p, miR-let-7c-5p, miR-221 and miR-345-5p are differentially expressed in Respiratory Syncytial Virus (RSV) persistently infected HEp-2 cells. Virus Res. 2018, 251, 34–39. [Google Scholar] [CrossRef]
  340. Zhang, Y.; Shao, L. Decreased microRNA-140-5p contributes to respiratory syncytial virus disease through targeting Toll-like receptor 4. Exp. Ther. Med. 2018, 16, 993–999. [Google Scholar] [CrossRef] [PubMed]
  341. Du, X.; Yang, Y.; Xiao, G.; Yang, M.; Yuan, L.; Qin, L.; He, R.; Wang, L.; Wu, M.; Wu, S.; et al. Respiratory syncytial virus infection-induced mucus secretion by down-regulation of miR-34b/c-5p expression in airway epithelial cells. J. Cell. Mol. Med. 2020, 24, 12694–12705. [Google Scholar] [CrossRef]
  342. Coultas, J.A.; Smyth, R.; Openshaw, P.J. Respiratory syncytial virus (RSV): A scourge from infancy to old age. Thorax 2019, 74, 986–993. [Google Scholar] [CrossRef] [Green Version]
  343. Tahamtan, A.; Samadizadeh, S.; Rastegar, M.; Nakstad, B.; Salimi, V. Respiratory syncytial virus infection: Why does disease severity vary among individuals? Expert Rev. Respir. Med. 2020, 14, 415–423. [Google Scholar] [CrossRef]
  344. Adkins, B.; Leclerc, C.; Marshall-Clarke, S. Neonatal adaptive immunity comes of age. Nat. Rev. Immunol. 2004, 4, 553–564. [Google Scholar] [CrossRef]
  345. Puthothu, B.; Krueger, M.; Forster, J.; Heinzmann, A. Association between Severe Respiratory Syncytial Virus Infection and IL13/IL4 Haplotypes. J. Infect. Dis. 2006, 193, 438–441. [Google Scholar] [CrossRef]
  346. Caballero, M.T.; Serra, M.E.; Acosta, P.L.; Marzec, J.; Gibbons, L.; Salim, M.; Rodríguez, A.; Reynaldi, A.; Garcia, A.; Bado, D.; et al. TLR4 genotype and environmental LPS mediate RSV bronchiolitis through Th2 polarization. J. Clin. Investig. 2015, 125, 571–582. [Google Scholar] [CrossRef] [Green Version]
  347. Gui, J.; Mustachio, L.; Su, N.-M.; Craig, R.W. Thymus Size and Age-related Thymic Involution: Early Programming, Sexual Dimorphism, Progenitors and Stroma. Aging Dis. 2012, 3, 280–290. [Google Scholar] [PubMed]
  348. Carvajal, J.J.; Avellaneda, A.M.; Salazar-Ardiles, C.; Maya, J.E.; Kalergis, A.; Lay, M.K. Host Components Contributing to Respiratory Syncytial Virus Pathogenesis. Front. Immunol. 2019, 10, 2152. [Google Scholar] [CrossRef] [PubMed]
  349. Cusi, M.G.; Martorelli, B.; Di Genova, G.; Terrosi, C.; Campoccia, G.; Correale, P. Age related changes in T cell mediated immune response and effector memory to Respiratory Syncytial Virus (RSV) in healthy subjects. Immun. Ageing 2010, 7, 14–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  350. De Waal, L.; Power, U.F.; Yüksel, S.; Van Amerongen, G.; Nguyen, T.N.; Niesters, H.G.; De Swart, R.L.; Osterhaus, A.D. Evaluation of BBG2Na in infant macaques: Specific immune responses after vaccination and RSV challenge. Vaccine 2004, 22, 915–922. [Google Scholar] [CrossRef]
  351. Murphy, S. Nirsevimab reduces medically attended RSV-associated lower respiratory tract infection and hospitalisations in healthy pre-term infants. Arch. Dis. Child. Educ. Pract. Ed. 2021. [Google Scholar] [CrossRef] [PubMed]
  352. Stobart, C.; Rostad, C.; Ke, Z.; Dillard, R.S.; Hampton, C.; Strauss, J.D.; Yi, H.; Hotard, A.L.; Meng, J.; Pickles, R.J.; et al. A live RSV vaccine with engineered thermostability is immunogenic in cotton rats despite high attenuation. Nat. Commun. 2016, 7, 13916. [Google Scholar] [CrossRef] [PubMed]
  353. Jorquera, P.A.; Mathew, C.; Pickens, J.; Williams, C.; Luczo, J.M.; Tamir, S.; Ghildyal, R.; Tripp, R.A. Verdinexor (KPT-335), a Selective Inhibitor of Nuclear Export, Reduces Respiratory Syncytial Virus Replication In Vitro. J. Virol. 2019, 93, e01684-18. [Google Scholar] [CrossRef] [Green Version]
  354. Sung, R.Y.; Yin, J.; Oppenheimer, S.J.; Tam, J.S.; Lau, J. Treatment of respiratory syncytial virus infection with recombinant interferon alfa-2a. Arch. Dis. Child. 1993, 69, 440–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bergeron, H.C.; Tripp, R.A. Immunopathology of RSV: An Updated Review. Viruses 2021, 13, 2478. https://0-doi-org.brum.beds.ac.uk/10.3390/v13122478

AMA Style

Bergeron HC, Tripp RA. Immunopathology of RSV: An Updated Review. Viruses. 2021; 13(12):2478. https://0-doi-org.brum.beds.ac.uk/10.3390/v13122478

Chicago/Turabian Style

Bergeron, Harrison C., and Ralph A. Tripp. 2021. "Immunopathology of RSV: An Updated Review" Viruses 13, no. 12: 2478. https://0-doi-org.brum.beds.ac.uk/10.3390/v13122478

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