Next Article in Journal
Feline Panleukopenia Outbreaks and Risk Factors in Cats in Animal Shelters
Previous Article in Journal
Exploring New Routes for Genetic Resistances to Potyviruses: The Case of the Arabidopsis thaliana Phosphoglycerates Kinases (PGK) Metabolic Enzymes
Previous Article in Special Issue
INMI1 Zika Virus NS4B Antagonizes the Interferon Signaling by Suppressing STAT1 Phosphorylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Characterization of SARS-CoV-2 Evasion: Interferon Pathway and Therapeutic Options

by
Mariem Znaidia
,
Caroline Demeret
,
Sylvie van der Werf
and
Anastassia V. Komarova
*
Génétique Moléculaire des Virus à ARN Unit, Department of Virology, Institut Pasteur, Université Paris Cité, CNRS UMR3569, 75015 Paris, France
*
Author to whom correspondence should be addressed.
Viruses 2022, 14(6), 1247; https://doi.org/10.3390/v14061247
Submission received: 28 April 2022 / Revised: 30 May 2022 / Accepted: 2 June 2022 / Published: 8 June 2022
(This article belongs to the Special Issue Viral Evasion of Innate Immunity and Drug Development)

Abstract

:
Severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2) is responsible for the current COVID-19 pandemic. SARS-CoV-2 is characterized by an important capacity to circumvent the innate immune response. The early interferon (IFN) response is necessary to establish a robust antiviral state. However, this response is weak and delayed in COVID-19 patients, along with massive pro-inflammatory cytokine production. This dysregulated innate immune response contributes to pathogenicity and in some individuals leads to a critical state. Characterizing the interplay between viral factors and host innate immunity is crucial to better understand how to manage the disease. Moreover, the constant emergence of new SARS-CoV-2 variants challenges the efficacy of existing vaccines. Thus, to control this virus and readjust the antiviral therapy currently used to treat COVID-19, studies should constantly be re-evaluated to further decipher the mechanisms leading to SARS-CoV-2 pathogenesis. Regarding the role of the IFN response in SARS-CoV-2 infection, in this review we summarize the mechanisms by which SARS-CoV-2 evades innate immune recognition. More specifically, we explain how this virus inhibits IFN signaling pathways (IFN-I/IFN-III) and controls interferon-stimulated gene (ISG) expression. We also discuss the development and use of IFNs and potential drugs controlling the innate immune response to SARS-CoV-2, helping to clear the infection.

1. Introduction

Severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2) is responsible for the current COVID-19 pandemic that emerged in Wuhan, Hubei, China, in December 2019 [1]. SARS-CoV-2 belongs to the Coronaviridae family. It is an enveloped positive-sense single-stranded RNA virus with a genome of 29,903 nucleotides that replicates entirely in the cytoplasm. The SARS-CoV-2 genome is composed of thirteen recognized open reading frames (ORFs) [2,3]. Upon the release of the SARS-CoV-2 genome into the cytoplasm, ORF1ab is translated to polyproteins pp1a and pp1ab. These polyproteins are then cleaved by viral chymotrypsin-like (3CLpro or NSP5) and papain-like (PLpro or NSP3) proteases to give 16 nonstructural proteins (NSP) that encode the RNA-dependent RNA polymerase (RdRp), helicase, and other components required for virus replication. Similar to other coronaviruses, SARS-CoV-2 replication involves the synthesis by RdRp of positive- and negative-sense full-length genomes. During its replication, SARS-CoV-2 also produces a nested set of subgenomic RNAs (sgRNAs). SARS-CoV-2 sgRNAs encode four structural proteins (S, spike; E, envelope; M, membrane; and N, nucleocapsid) and several accessory factors (ORF3a, ORF3b, ORF6, ORF7a, ORF7b, ORF8, and ORF9b) [1,4,5,6,7].
Innate immunity is the first line of defense against pathogens. It plays a critical role in controlling viral infections by mounting a rapid antiviral state and by mediating the adaptive immune response [8]. Interferons (IFNs) are transcriptionally regulated cytokines that play a central role in the innate antiviral immune response and orchestrate adaptive immunity. In some cases, this immune response could participate in the pathology. Indeed, SARS-CoV-2 infection is characterized by a dysregulated innate immune response, which contributes to the pathogenicity, leading in some individuals to a critical state [9,10,11]. An early IFN response is needed to establish a robust antiviral state, but this response is weak and delayed in severe COVID-19 patients, along with massive pro-inflammatory cytokine production [12,13,14].
The first step of innate immunity is the recognition of pathogen associated molecular patterns (PAMPs) by pattern-recognition receptors (PRRs) that initiates signaling cascades leading to the production of IFN, pro-inflammatory cytokines, and chemokines [15]. Among PRRs, Toll-like receptor 3 (TLR3), TLR7, TLR8, RIG-I-like receptors (retinoic acid-inducible gene I (RIG-I)), and melanoma differentiation-associated protein 5 (MDA5) have been described to recognize non-self RNA molecules [16,17]. RIG-I is specialized in the recognition of the 5′-triphosphate end of relatively short double-stranded (ds)RNAs, while MDA5 recognizes long dsRNAs [18]. SARS-CoV-2 RNAs have been described to be recognized by RIG-I and MDA5 [19,20], with MDA5 being the main cytoplasmic PRR capable of detecting SARS-CoV-2 [21,22]. RIG-I and MDA5 are composed of an RNA helicase domain and a C-terminal domain (CTD) involved in viral RNA recognition. The N-terminal caspase activation and recruitment domains (CARDs) of both RIG-I and MDA5 interact with adaptor proteins of the mitochondrial-associated viral signaling proteins (MAVS) to trigger IFN regulatory factor 3 (IRF3) and IRF7 phosphorylation and dimerization, leading to their translocation to the nucleus and the transcription of IFN-I and IFN-III. The NF-κB signaling cascade is also activated through the so-called canonical pathway by a set of receptors that includes TLRs and RLRs, triggering the expression of pro-inflammatory cytokines and chemokines, such as interleukin 6 (IL-6), IL-1B, tumor necrosis factor (TNF)- alpha, and chemokine ligand 2 (CCL2) [23,24]. IFN-I and IFN-III play a central role in antiviral immunity. IFN-α and β belong to IFN-I, while IFN-λ (IFN-λ: λ1, λ2, λ3, and λ4) belongs to IFN-III. Both IFN-I and IFN-III are secreted by infected cells and are involved in inhibiting viral infections. However, their action involves distinct receptors. IFN-α receptor 1 (IFNAR1) and IFNAR2, the two chains forming the heterodimer at the cell membrane surface, are indispensable to IFN-I binding and subsequent signaling. Contrary to IFN-I receptors, which are expressed by most cells, the IFN-λ: λ1, λ2, λ3, and λ4 (IFN-LR1) chain of the IFN-III receptor (IFN-LR1/ IL-10R2 heterodimer) is expressed by a restricted subset of cells, including epithelial cells [25]. IFN-I and IFN-III share an overlapping pathway of activation and functions, but IFN-I has been described to exert a more potent and rapid response, while IFN-III expression is more sustained and restricted to epithelial barriers. Both are secreted by infected cells, leading to an autocrine and paracrine stimulation through the JAK/STAT signaling pathway involving signal transducers and activators of transcription (STAT1 and STAT2), protein phosphorylation by Janus kinase 1 (JAK 1) and tyrosine kinase 2 (TYK2), and dimerization to interact with IRF9. Together, STAT1, STAT2, and IRF9 form transcription factor IFN-stimulated gene factor 3 (ISGF3), which translocates to the nucleus and binds to IFN-stimulated response elements (ISREs) in interferon-stimulated gene (ISG) promoters, resulting in their activation [26,27]. These ISGs play a crucial role in repressing viral replication via various mechanisms. They prevent viral entry into the cell and viral trafficking into the nucleus, inhibit viral replication along with viral transcription/translation, and can also degrade viral nucleic acids and block viral particle assembling [28,29].
SARS-CoV-2 has developed multiple strategies to impede this innate immune response and achieve efficient replication. The knowledge of the mechanisms by which SARS-CoV-2 interferes with innate immunity is crucial to apprehending this newly emerged virus and developing vaccines and therapies. In this review, we describe recent advances in the understanding of the circumvention of the innate immunity by SARS-CoV-2 (summarized in Table 1 and represented in Figure 1) and explore some therapeutic options that were developed and assessed for COVID-19 (Table 2).

2. Evasion of Innate Immunity by SARS-CoV-2

2.1. Evasion of Sensing by Host Innate Immune Receptors

Coronaviruses, including SARS-CoV-2, shield their dsRNA replication intermediates in double-membrane vesicles (DMVs) mediated by NSP3, NSP4, and NSP6, avoiding dsRNA sensing by RIG-I, MDA5, and TLR3 [30,31]. Moreover, viral RNA capping is a critical step to prevent PRR sensing. It is mediated by the RNA triphosphatase activity of NSP13, the RdRp NSP12, the NSP14 N7-MTase activity, NSP10, and NSP16 2′-O-MTase activity [32]. Since RIG-I recognizes RNA 5′-triphosphate, by harboring a ribose 2′-O-methylation 5′-cap structure, SARS-CoV-2 evades innate immune recognition [33,34]. The nucleocapsid (N) has been described to inhibit SARS-CoV-2 RNA recognition by RIG-I. The first molecular mechanism has been proposed where RIG-I inhibition is mediated by a direct interaction through its DExD/H helicase domain with the N protein [35]. This domain has an ATPase activity playing a crucial role in RNA binding, suggesting N directly blocks the RNA recognition.
The activation of PRRs can also be inhibited. ISG15, a ubiquitin-like protein that can be covalently conjugated to lysine residues (ISGylation), plays an essential role in MDA5 activation. NSP3 cleaves ISG15 [36] and consequently antagonizes ISG15-dependent MDA5 activation, evading immune recognition [37]. Additionally, ISG15 acts as an antiviral factor, modifying host and viral proteins, which interferes with viral assembly or function. The other protease, NSP5, further mediates the inhibition of PRR functioning. RIG-I is cleaved by NSP5 at the last 10 N-terminal amino acids, blocking its ability to signal through MAVS [38]. In addition, other studies have described the suppression of IFN-β production in the consequence of TRIM25’s interaction with the N protein [39,40]. The ubiquitination of the RIG-I CARD domain by TRIM25 leads to the activation of RIG-I, leading to downstream signaling. By this N-TRIM25 interplay, the signaling cascade is blocked [40,41].

2.2. Inhibition of Innate Immune Receptor Signaling and IFN Production

Strategies have been mounted by SARS-CoV-2 to inhibit the activation of PRRs downstream of the PAMPs sensing step. IFN-I and IFN-III responses are fundamental defenses of the antiviral innate immunity in the clearance of infection. To establish a successful infection of host cells, SARS-CoV-2 has developed various strategies to antagonize different stages of the signaling leading to IFN production. Particularly, IFN-β production is inhibited by NSP1, NSP3, NSP5, NSP12, NSP13, NSP14, NSP15, ORF3a, ORF3b, ORF6, ORF7a, ORF7b, ORF8, ORF9b, N, and M, as reported by several studies [42,43,44,45,46]. These proteins have been shown to reduce RIG-I-mediated IFN-β promoter activities, suggesting they can suppress RLR-mediated signaling [19,42,43,44]. This is mediated by many molecular strategies (Figure 1 and Table 1). Among them, are the inhibition of IRF3 phosphorylation and its nuclear translocation, which are pivotal steps required for the activation of IFN-I transcription [47,48]. NSP1 [49], NSP12 [50], NSP13, NSP14, NSP15, ORF3b [51], and ORF6 [42] have been reported to inhibit the nuclear translocation of IRF3. For example, the NSP12 inhibition of IRF3 nuclear translocation impairs downstream signaling of type-I IFN, as ISRE and ISG56 promoters have been described to be inhibited in this way [50]. However, in a study by Li et al., no IRF3 nuclear translocation inhibition was shown when NSP12 was overexpressed [52]. This study demonstrated that NSP12 could reduce IFN-β luciferase promoter activity via a yet-to-be-determined mechanism but cannot inhibit the IFN-β production induced by Sendai virus infection (well-known RLR agonist) or another stimulus as well as the downstream signaling. ORF6 has been described to bind importin karyopherin α 2 (KPNA2). Since KPNA2 is an importin involved in importing IRF3, the inhibition of IRF3 translocation in the presence of ORF6 has been suggested to be due to its interaction with KPNA2 [44]. IRF3 activation and subsequent IFN production depend on the assembly of the multiprotein complex containing RIG-I/MDA-5, MAVS, TNF receptor-associated factor 3 (TRAF3), and TANK-binding kinase 1 (TBK1). TBK1 is a protein kinase that is involved in IFN-I transcription by phosphorylating IRF3 [53]. The M protein prevents the formation of this complex [54] by association with TBK1 and its degradation via the ubiquitin pathway, thereby inhibiting IRF3 phosphorylation and IFN production [55,56]. Besides, ORF7a reduces IRF3 nuclear translocation by destabilizing TBK1 [19]. NSP13 also inhibits TBK1 phosphorylation, resulting in decreased IRF3 activation and IFN-β production [57]. NSP6 and NSP15 bind TBK1 to decrease IRF3 phosphorylation as well [44]. ORF9b interacts with the viral RNA sensors RIG-I, MDA-5, and the downstream signaling factors that include, for example, MAVS, and TBK1, leading to TBK1 phosphorylation inhibition, consequently inhibiting IRF3 phosphorylation, its nuclear translocation, and types I and III IFN transcription [58]. Furthermore, ORF9b has been described to associate with TOM70, an adaptor protein of the MAVS complex at the mitochondrial membrane, inhibiting the triggering signal leading to IFN production [59]. Another antiviral role of ORF9b is to interact with NF-κB essential modulator (NEMO) to block its K63-linked polyubiquitination. This ubiquitylation is necessary to activate the IKKα/β kinase phosphorylation function of NEMO, leading to IKKα ubiquitylation and degradation and the subsequent translocation of NF-κB subunit p65 into the nucleus to activate the transcription of pro-inflammatory cytokines [60]. Furthermore, the translocation of p65 is repressed because of NSP9’s interaction with the nucleoporin NUP62 [61]. By its ability to alter nuclear pore complex (NPC) composition, NSP9 is consequently capable of blocking the antiviral immune response. Furthermore, NSP5 inhibits the pro-inflammatory NF-κB pathway [62]. Moreover, NSP5 also prevents the nuclear translocation of IRF3 without direct cleavage [63], while other studies have demonstrated an IRF3 degradation promoted by NSP5 [62,64]. Additionally, NSP5 promotes the ubiquitination and proteasome-mediated degradation of MAVS [38]. Besides, the direct cleavage of IRF3 by NSP3 has been demonstrated by Moustaqil et al. [62]. This huge redundancy in antagonizing IFN production highlights the importance of SARS-CoV-2 to shutting down the IFN response to the success of its infection.

2.3. Inhibition of IFN Signaling and ISG Expression

SARS-CoV-2 has also developed many strategies to antagonize the IFN response, inhibiting IFN-α/β receptor signaling and ISG expression. The very early step of IFN signaling is targeted by SARS-CoV-2. NSP14 mediates IFNAR1 lysosomal degradation, blocking the activation of the transcription factors STAT1 and STAT2 [65]. Additionally, the phosphorylation of STAT1 and STAT2 is counteracted by several viral proteins. Xia et al. have described the capacity of NSP1, NSP6, NSP13, ORF3a, M, ORF6, ORF7a, and ORF7b to suppress >40% of ISRE promoter activity by inhibiting STAT1 phosphorylation for NSP1, NSP6, NSP13, ORF3a, ORF7b, M, ORF6, and ORF7a and STAT2 phosphorylation for NSP1, ORF6, ORF3a, and M [44]. ORF6 impairs IFN production and has also been described to block STAT1/2 nuclear translocation, a crucial step to activate the transcription of ISGs [66]. Indeed, ORF6 interacts with the nucleopore complex Nup98 via its C-terminus, thus blocking mRNA export [67]. This leads to the impairment of IFN mRNA nuclear export and also transcription factor nuclear import. The ORF6 nucleopore complex interaction has been proposed to clog the nuclear pore [68], explaining the inhibition of STAT1 and IRF3 nuclear translocations [43,44]. The N protein has been demonstrated to competitively bind to STAT1/STAT2, thus interfering with the interactions of STAT1 with JAK1 and STAT2 with TYK2, respectively. This inhibits their phosphorylation and subsequent ISGs production [69]. Moreover, ORF7a hijacks the host ubiquitin system. The host ubiquitin system is usurped to form K63-linked ubiquitin chains, enhancing, in this way, its ability to inhibit STAT2 phosphorylation and, hence, blocking ISGF3 complex translocation to the nucleus to activate ISG transcription [70].
Thus, the IFN signaling pathway is blocked at different levels by several SARS-CoV-2 proteins, ensuring an advantage for viral replication. The ISGs produced are also targeted. Viral proteases have been found to cleave IFN-stimulated antiviral proteins as soon as they are produced. For instance, NSP3 cleaves ISG15 [36].

2.4. Inhibition of Host Protein Production by Targeting Post-Transcriptional and Translational Steps

NSP16 interferes with mRNA production. The critical step of mRNA splicing is mediated by the spliceosome, a complex of proteins and non-coding RNAs. NSP16 inhibits the splicing of pre-mRNA by binding to U1 and U2, which are two small nuclear RNAs involved in this complex. The ISG expression is then repressed [71]. mRNA export to the cytoplasm has been described to be blocked by NSP1 [72]. Mechanistically, NSP1 interacts with the host mRNA export receptor heterodimer NXF1-NXT1 [73]. This receptor is responsible for the nuclear export of cellular mRNAs by binding to mRNA and docking at the NPC. NSP1 binds to NXF1, reducing the interaction with the NPC and preventing mRNA nuclear transport. Moreover, the NSP1 C-terminal domain binds to the small ribosomal subunit (40S), more specifically to the 18S rRNA, disrupting 40S mRNA scanning and preventing translation initiation, then blocking host protein production [71]. This impacts IFN production without altering viral mRNA translation. Cryo-electron microscopy analyses of the molecular interaction between NSP1 and the 40S ribosome subunit have revealed that NSP1 is bound to the mRNA entry channel overlapping the mRNA path [74,75,76,77]. However, how viral mRNA containing the leader sequence passes through the translation shut off triggered by NSP1 is not fully understood. Banerjee et al. have observed that the common SARS-CoV-2 5′-leader sequence precisely positions viral mRNAs relative to the NSP1-40S complex, enabling translation [71]. Mendez et al. demonstrated that the alteration of NSP1-mRNA-40S complex by mutating residues within the N-terminal domain of NSP1 makes a weaker association with the ribosome and mRNA, abrogating the translational escape of viral leader-containing mRNAs. They suggested that these NSP1 residues are important to allow the viral leader-sequence-triggered conformational changes to NSP1 that support viral mRNA translation [78]. NSP8 and NSP9 disrupt protein trafficking, leading to the degradation of newly translated proteins. The signal recognition particle (SRP) is a complex that binds the 80S ribosome and identifies co-translationally hydrophobic signal peptides present in secreted and membrane proteins, leading to their translocation to the endoplasmic reticulum and their proper folding and expression at the membrane. NSP8 and NSP9 bind to 7SL RNA, a non-coding RNA that is a part of the SRP complex, leading to the inhibition of signal peptide recognition. This disruption of protein trafficking suppresses the IFN response, as described by Banerjee et al. [71]. Thus, SARS-CoV-2 developed several strategies to alter general cellular processes involved in post-transcriptional and translational processes.

3. Therapeutic Options

SARS-CoV-2 employs many strategies to impede the IFN response, as described above, resulting in an imbalance in the innate responses, with excessive cytokine production [12,13,14]. A weak and delayed IFN response along with an exacerbated inflammatory response has been associated with COVID-19 pathogenesis [9,10,11]. Patients with no IFN-α production present poorer outcomes and show a longer intensive care unit stays [79]. Clinical data support the implication of deficient IFN-I responses in the progression of COVID-19 toward a more severe manifestation [80]. Moreover, studies have shown that COVID-19 severity is enriched in patients with genetic defects in TLR3, IRF7, and TLR7, leading to a deficient induction of IFN-I [81,82,83]. Besides, autoantibodies targeting IFN-I were found in 10% of patients with severe COVID-19, while they were absent in asymptomatic or mildly symptomatic patients. These antibodies were then identified as a risk factor for life-threatening COVID-19 [84,85,86,87] and have been described to increase in patients over 60 years old and cause about 20% of all fatal COVID-19 cases [88]. Many studies have exposed the capacity of both IFN-I and IFN-III to inhibit SARS-CoV-2 in vitro or in vivo [13,89,90,91,92,93]. These observations highlight the crucial role played by the IFN response in controlling SARS-CoV-2 infection and subsequent clinical implications. Therefore, IFN therapy was naturally considered to support innate immunity and control the SARS-CoV-2 infection. Below, we discuss studies that have been performed to test various IFN therapies, which are also summarized in Table 2.

3.1. IFN-α

IFN-α has already been used in the viral infection context. Indeed, pegylated IFN-α2 (PEG-IFN-α2, a long-acting form of IFN-α2b) was used as the standard in the treatment of chronic hepatitis C virus infection until it was replaced by antivirals and is still a therapeutic option in mild to moderate chronic hepatitis B [94,95].
In the context of COVID-19, some studies testing IFN-α efficacy have been conducted. Inhaled IFN-α2b administration significantly reduced the duration of SARS-CoV-2 detection in the upper respiratory tract by a reverse transcription-quantitative PCR (RT-qPCR) test and blood levels of inflammatory markers such as IL-6 and C-reactive protein, as reported in an exploratory study by Zhou et al. [96]. A reduction in viral load and clinical status improvement in moderate COVID-19 cases after subcutaneous administration of pegylated IFN-α2b was also reported in a phase II clinical trial and a phase III clinical trial [97,98], but no difference was observed for inflammatory biomarkers or in the duration of hospitalization. Furthermore, patients who have been administrated subcutaneous IFN-α2b within 72 h following admission had shorter hospital stays compared with late administration, as presented by B.Wang et al. [99]. However, the studies described above are limited by the small number of subjects enrolled (40 in [97] and 41 in [99]). Additionally, in the phase III study from N.Wang et al., a reduction in in-hospital mortality was observed in patients who were given IFN-α2b by inhalation using an air compressor machine (nebulized delivery) within 5 days of admission. However, late initiation of this therapy was associated with increased mortality [100]. Similarly, in the study reported by Yu et al., early administration of IFN-α2b aerosol was associated with improved clinical outcomes (such as a lower risk of disease progression and a shorter hospitalization time), but delayed IFN-α2b intervention was associated with increased probabilities of adverse events [101].
At the molecular level, the SARS-CoV-2 receptor ACE2 was reported as an ISG upregulated by IFN-α [102], raising concerns about potential pro-viral effects of ACE2 induction upon IFN treatment. Later, molecular characteristic studies demonstrated that IFNs only induced the expression of a truncated product of ACE2 that is not able to bind the SARS-CoV-2 spike protein [103,104,105].
Thus, the timing of IFN-α administration seems to play a crucial role in its efficacy, and early administration may lower the risk of worse outcomes. Because of the exclusion of patients at risk in these studies, an understanding of the association between age, comorbidities, the stage of the disease, timing, and the IFN-α treatment is still needed. Furthermore, most clinical studies have been focused on nebulized IFN-α, but this form is not yet approved by the United States Food and Drug Administration (FDA), and its use is not recommended and is restricted in clinical studies by the World Health Organization (WHO). Overall, IFN-α treatment remains controversial, even though the outcomes of early administration of IFN-α are encouraging.

3.2. IFN-β

Since in vitro studies have reported that SARS-CoV-2 is more sensitive to IFN-β than IFN-α [90], and autoantibodies against IFN-α but not against IFN-β have been found in patients with severe disease [84], IFN-β seemed to be a more appropriate therapy than IFN-α.
Different strategies are employed by SARS-CoV-2 proteins to specifically antagonize IFN-β production, as described in Part 2 of this review. The direct delivery of IFN-β could, in principle, overcome these evasion strategies. Indeed, IFN-β binds IFNAR and initiates the JAK-STAT signaling pathway, thereby inducing ISG expression. IFN-β displays anti-inflammatory properties by enhancing IL-10 expression, and it has been reported to upregulate CD73 on endothelial cells, preventing vascular leakage [106] and inhibiting leukocyte recruitment [107]. Besides, a study carried out before the COVID-19 pandemic reported that IFN-β injection improves acute respiratory distress syndrome that occurs in the presence of predisposing factors [108].
Despite all these experimental pieces of evidence, clinical trial results for IFN-β are controversial. Systemic or subcutaneous deliveries of IFN-β have been reported to be efficient in shortening the duration of viral shedding and hospital stay [109], reducing mortality [110,111] when combined with other antiviral drugs but only in the early stage of the disease in mild to moderate cases. However, the SOLIDARITY randomized trial reported that 2000 COVID-19 hospitalized patients who received subcutaneous IFN-β1a or IFN-β1a together with the antivirals lopinavir-ritonavir experienced little or no effect on clinical improvement [112]. In fact, the subcutaneous delivery of IFN-β1a to severe COVID-19 patients failed to reduce mortality according to the WHO SOLIDARITY trial outcomes [112] and as reported by Kalil and colleagues [113]. Worse outcomes after treatment with IFN-β1a have also been reported, suggesting the potential harm of using IFN-β in patients with severe disease, such as those on high-flow oxygen, noninvasive ventilation, or mechanical ventilation [113]. Thus, the COVID-19 Treatment Guidelines Panel does not recommend the use of systemic IFN-β for the treatment of hospitalized patients with COVID-19 [114]. The administration of IFN-β directly in the respiratory tract may result in a robust local anti-viral immune response, avoiding the inconvenience of IFN-β systemic exposure. Inhaled IFN-β was examined in a randomized double-blind placebo-controlled phase II trial [115]. Indeed, patients treated with nebulized IFN-β1a showed significantly greater odds of clinical improvement across the WHO Ordinal Scale for Clinical Improvement than those who received a placebo [115]. Still, in a study performed on adult patients with moderate COVID-19, the administration of IFN-β1b by vapor inhalation combined with an antiviral Favipiravir failed to provide benefits [116].
The combination with other treatments in many IFN-β trials makes it difficult to derive a clear conclusion about the performance of IFN-β. Additionally, the size of the sample and the stage of the disease when patients are recruited have to be considered when analyzing all these clinical studies. More data are needed to settle if inhaled IFN-β could be used to treat COVID-19. However, it has to be noted that nebulized IFN-β is not yet approved in the USA and Europe, and the pharmacokinetics and pharmacodynamics of this mode of administration are not fully known.

3.3. IFN-λ

IFN-I is able to provide a systemic response along with a local one, whereas IFN-III delivers a protection restricted to epithelial surfaces. Especially in the context of influenza virus infection, type-III IFNs appear to be essential in controlling viral infection in the upper respiratory tract, while IFN-I and IFN-III dispense functional redundancy in the lower respiratory tract [117]. IFN-I triggers systemic and local inflammation by activating and recruiting immune cells to the site of infection. In contrast, IFN-III has been shown to limit viral activities at the local level without being associated with inflammation [118,119,120], conversely restricting the neutrophil inflammatory functions then limiting tissue damage [120,121,122]. In the context of SARS-CoV-2 infection, in vitro and in vivo studies have demonstrated antiviral IFN-III activities in epithelial cells without excessive inflammation [89,91,92,123,124]. Furthermore, the protein expression of IFN-λ2/λ3 were reduced in nasopharyngeal swab samples from COVID-19-positive individuals [125]. Additionally, studies have demonstrated a downregulation of systemic IFN-λ1/λ2 in patients with severe COVID-19, suggesting that a reduced level of IFN-III may impact the severity of the disease [126,127,128].
The abovementioned studies highlight the potential benefits of using IFN-λ2 as an antiviral treatment for SARS-CoV-2. IFN-λ has been proposed to help clear the infection and minimize the severity of COVID-19 [129]. Despite of all these promising benefits, as of today only two clinical trials have been finalized [130,131], while few others are still recruiting. In a randomized placebo-controlled trial reported by Jagannathan et al. on 120 patients with mild to moderate COVID-19, 60 received subcutaneous doses of Peginterferon IFN-λ1a within 72 h of diagnosis. No difference has been observed between the Peginterferon IFN-λ1a and placebo groups in SARS-CoV-2 viral shedding or in the improvement of symptoms [131]. In contrast, the study by Feld et al. described the benefits of subcutaneous Peginterferon IFN-λ in treating mild to moderate cases of COVID-19 who were within 7 days of symptom onset or the first positive test if asymptomatic [130]. Further studies are needed to validate the efficacy of IFN-III treatment of COVID-19.
It should be pointed out that in the context of influenza infection IFN-λ has been reported to contribute to bacterial superinfections [132,133]. Since critically ill COVID-19 patients suffer from bacterial superinfections, caution should be exercised in using IFN-λ as a COVID-19 treatment. Additionally, IFN-λ has not been approved by the FDA for any use. These facts could explain the precautionary approach to limit the use of IFN-λ to clinical trials.

4. Conclusions and Perspective

COVID-19 has been characterized by a dysregulation of the immune response that can lead to severe acute respiratory distress syndrome in some people. Multiple strategies have been evolved by SARS-CoV-2 structural, nonstructural, and accessory proteins to counteract viral sensing, IFN production, and IFN signaling pathways. This redundancy in evading innate immunity characterizes SARS-CoV-2 infection (Table 1) and underlines the adaptation to the human host. The inhibition of the IFN response by SARS-CoV-2 viral factors, especially NSPs that are expressed early upon virus cell entry, promotes viral replication and participates in the delay in the IFN response observed in COVID-19 patients.
The direct use of IFN therapy in the context of COVID-19 has been expected to resolve a weak early innate immune reaction and restore a strong and well-timed IFN response capable of inhibiting the virus replication before the immunopathology takes place. However, the IFN response is a fine-tuned process that is not fully understood, and clinical trials reveal disparate results depending on time of treatment initiation and severity state. These contradictory results of IFN therapy’s benefits in handling COVID-19 underline that two central factors should be considered: the time from the beginning of the treatment and the route of administration. Some in vitro and in vivo studies have also described differences in efficacy upon IFN-I pretreatment and post-infection treatment, underlying the interest in using models to calibrate IFN therapies [134,135,136]. Besides, a study showed that high IFN-III expression at the protein level in the upper respiratory tract is associated with a mild disease and, hence, could be protective [137], while another study demonstrated an increase at the transcriptional level for IFN-I/IFN-III mRNA but a decrease of the corresponding proteins [125]. Along the same line, single-cell RNA sequencing on nasopharyngeal swabs of COVID-19 patients has revealed a robust IFN-I-specific gene signature in patients with mild to moderate disease in contrast to an impaired one in severe cases [138]. Conversely, some studies have reported a potent IFN response in critical cases. Elevated IFN-I and III in the lower airways at both the transcriptional and protein levels have been reported in patients with severe COVID-19 [137]. Moreover, a sustained production of IFN-α in the blood has been reported in severe COVID-19 cases in a longitudinal study [139], along with an IFN-I transcriptome signature co-existing with the inflammatory one [140]. Thus, IFNs could play a direct role in exacerbating the inflammatory response to SARS-CoV-2 infection. One hypothesis that could explain the opposite roles played by IFN-I/III during COVID-19 pathogenesis is that the IFN response is blocked by the SARS-CoV-2 counteracting strategy in infected epithelial cells, leading to a massive replication and the subsequent high viral load observed. Then, the produced viral particles could stimulate innate immune cells (monocytes, dendritic cells, and macrophages), which produce substantial amounts of IFNs. Thus, the side-effect observed by late IFN administration could be explained by a potential contribution to the cytokine storm, emphasizing the inflammatory state. Furthermore, the ambivalent properties that depend on the anatomical location of various IFN-producing cells and viral tropism to the upper and low respiratory tracts have to be taken into consideration. This underlines the complex role of IFNs in COVID-19 pathogenesis, and further investigations are needed to decipher their exact implication in the immunopathology and to state the appropriate conditions for the therapy.
The understanding of the mechanisms underlying the host and virus interactions is crucial to the development of drugs and vaccines. In particular, apprehending the process of innate immune response establishment and SARS-CoV-2 evasion could help the development of drugs such as immunomodulators or antiviral drugs that directly target viral immune counteractors. The inhibition of those viral components may attenuate pathogenicity by increasing the naturally early host antiviral response. However, a particularity of SARS-CoV-2 among other coronaviruses is to entertain high mutation capacities [141] that give rise to variants, thus challenging drug development and vaccine efficacy. Targeting conserved proteins is needed because they are less prone to undergo mutations. Conducting a computational screening of FDA-approved drugs against such conserved proteins makes a rapid and efficient method to identify effective treatments that can be tested in vitro and in vivo. In this regard, NSP1 appears to be an attractive target since it is a very conserved protein [142] playing a central role in antagonizing the innate immune response, as described in this review [46,57,73,76]. Montelukast sodium hydrate (which is an FDA-approved drug that stably binds to NSP1) has shown antiviral activity against SARS-CoV-2, with reduced viral replication in cell lines in vitro [143]. GRL0617, another FDA-approved drug, which is a non-covalent inhibitor of SARS-CoV-2’s NSP3 [144], displays promising issues on SARS-CoV-2′s NSP3 [145]. GRL0617 has been demonstrated to block the binding of SARS-CoV-2 NSP3 to ISG15 [36], inhibiting its deubiquitinase activity [146]. These examples underline the benefits of investigating the area of SARS-CoV-2’s conserved proteins that are implicated in innate immunity evasion.
Currently, few antiviral drugs and immunomodulators are approved by the WHO Therapeutics and COVID-19: living guideline in the treatment of COVID-19 [114]. Among them, remdesivir, an analog of adenine inhibiting RdRp was first recommended to treat severe cases and has been described to show a higher efficacy early in the course of the disease, during the viral replication phase of COVID-19, and to not be so useful during the immunopathology phase [114]. In contrast, immunomodulators were shown to be effective to harness inflammation. IL-6 blockers such as the JAK1/JAK2 inhibitor Baricitinib or the monoclonal antibody Tocilizumab, which antagonizes the membrane-bound and soluble forms of the IL-6 receptor, have been recommended to treat severe or critical COVID-19 patients [114]. Nevertheless, ideal drugs for the treatment of COVID-19 are yet to be discovered. Regarding the emergence of several variants of concern since the beginning of the COVID-19 pandemic, challenging the efficacy of currently developed vaccines, therapies targeting conserved viral proteins and adjusting the effective innate immune response to SARS-CoV-2 should be priorities.

Author Contributions

M.Z. drafted the manuscript. All authors discussed, corrected, edited, and approved the text. All authors have read and agreed to the published version of the manuscript.

Funding

A.V.K. acknowledges the ANR (Agence Nationale de la Recherche) and FRM (Fondation de la Recherche Médicale) for funding the AAP Flash-COVID 19 project SARS-CoV-2immunRNAs and Institut Pasteur for the «URGENCE COVID-19» fundraising campaign for the project CoV-2SENSING and the Fondation Entreprise Michelin for the PFR5 SARS-CoV-2 program CoV-2_CVnet. C.D. acknowledges the ANR DARWIN (ANR-20-COVI-0063), and Institut Pasteur for the «URGENCE COVID-19» fundraising campaign for the project PSII. S.v.d.W. acknowledges the funding of the RECOVER project (ANR-10-LABX-62-IBEID) and the European Union’s Horizon 2020 research and innovation program RECOVER (No. 101003589).

Acknowledgments

The authors would like to thank members of Sylvie van der Werf’s laboratory and the National Reference Center (CNR) for Respiratory Viruses at the Institut Pasteur for support and valuable discussions. The figures were created with BioRender.com.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CARDcaspase activation and recruitment domain
CCL2chemokine ligand 2
3CLproprotease 3C-like
DMVdouble-membrane vesicle
dsRNAdouble-stranded RNA
Eenvelope
FDAUnited States Food and Drug Administration
IFNinterferon
IFNARIFN-α receptor
ILinterleukin
IFN-LR1IFN-l receptor 1
IKKα/ßinhibitor of nuclear factor kappa-B kinase
IRFIFN-regulatory factor
ISGIFN-stimulated gene
ISREIFN-stimulated response element
ISGFIFN-stimulated gene factor
JAKJanus-associated kinase
KPNA2importin karyopherin α 2
Mmembrane
MAVSmitochondrial-associated viral signaling protein
MDA5melanoma differentiation-associated protein 5
Nnucleocapsid
NEMONF-κB essential modulator
NF-kBnuclear factor kB
NPCnuclear pore complex
NSPnonstructural protein
ORFopen reading frame
PAMPpathogen-associated molecular patterns
PLpropapain-like protease
PRRpattern-recognition receptors
RIG-Iretinoic acid-inducible gene I protein
RdRpRNA-dependent RNA polymerase
RT-qPCRreverse transcription-quantitative PCR
SARS-CoV-2severe acute respiratory syndrome coronavirus 2
sgRNAsubgenomic RNA
Sspike
SREBP1sterol regulatory element-binding protein 1
SRPsignal recognition particle
STATtransducer and activator of transcription
TBK1TANK-binding kinase 1
TLRToll-like receptor
TNFtumor necrosis factor
TRAFTNF receptor-associated factor
TYKtyrosine kinase
WHOWorld Health Organization

References

  1. Wu, F.; Zhao, S.; Yu, B.; Chen, Y.-M.; Wang, W.; Song, Z.-G.; Hu, Y.; Tao, Z.-W.; Tian, J.-H.; Pei, Y.-Y.; et al. A New Coronavirus Associated with Human Respiratory Disease in China. Nature 2020, 579, 265–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Khailany, R.A.; Safdar, M.; Ozaslan, M. Genomic Characterization of a Novel SARS-CoV-2. Gene Rep. 2020, 19, 100682. [Google Scholar] [CrossRef] [PubMed]
  3. Malone, B.; Urakova, N.; Snijder, E.J.; Campbell, E.A. Structures and Functions of Coronavirus Replication–Transcription Complexes and Their Relevance for SARS-CoV-2 Drug Design. Nat. Rev. Mol. Cell Biol. 2022, 23, 21–39. [Google Scholar] [CrossRef] [PubMed]
  4. Davidson, A.D.; Williamson, M.K.; Lewis, S.; Shoemark, D.; Carroll, M.W.; Heesom, K.J.; Zambon, M.; Ellis, J.; Lewis, P.A.; Hiscox, J.A.; et al. Characterisation of the Transcriptome and Proteome of SARS-CoV-2 Reveals a Cell Passage Induced in-Frame Deletion of the Furin-like Cleavage Site from the Spike Glycoprotein. Genome Med. 2020, 12, 68. [Google Scholar] [CrossRef]
  5. Kim, D.; Lee, J.-Y.; Yang, J.-S.; Kim, J.W.; Kim, V.N.; Chang, H. The Architecture of SARS-CoV-2 Transcriptome. Cell 2020, 181, 914–921.e10. [Google Scholar] [CrossRef]
  6. Zhou, P.; Yang, X.-L.; Wang, X.-G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.-R.; Zhu, Y.; Li, B.; Huang, C.-L.; et al. A Pneumonia Outbreak Associated with a New Coronavirus of Probable Bat Origin. Nature 2020, 579, 270–273. [Google Scholar] [CrossRef] [Green Version]
  7. Hartenian, E.; Nandakumar, D.; Lari, A.; Ly, M.; Tucker, J.M.; Glaunsinger, B.A. The Molecular Virology of Coronaviruses. J. Biol. Chem. 2020, 295, 12910–12934. [Google Scholar] [CrossRef]
  8. Seth, R.B.; Sun, L.; Chen, Z.J. Antiviral Innate Immunity Pathways. Cell Res. 2006, 16, 141–147. [Google Scholar] [CrossRef] [Green Version]
  9. Huang, C.; Wang, Y.; Li, X.; Ren, L.; Zhao, J.; Hu, Y.; Zhang, L.; Fan, G.; Xu, J.; Gu, X.; et al. Clinical Features of Patients Infected with 2019 Novel Coronavirus in Wuhan, China. Lancet 2020, 395, 497–506. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, Q.; Meng, Y.; Wang, K.; Zhang, X.; Chen, W.; Sheng, J.; Qiu, Y.; Diao, H.; Li, L. Inflammation and Antiviral Immune Response Associated with Severe Progression of COVID-19. Front. Immunol. 2021, 12, 631226. [Google Scholar] [CrossRef]
  11. Zhou, F.; Yu, T.; Du, R.; Fan, G.; Liu, Y.; Liu, Z.; Xiang, J.; Wang, Y.; Song, B.; Gu, X.; et al. Clinical Course and Risk Factors for Mortality of Adult Inpatients with COVID-19 in Wuhan, China: A Retrospective Cohort Study. Lancet 2020, 395, 1054–1062. [Google Scholar] [CrossRef]
  12. Hadjadj, J.; Yatim, N.; Barnabei, L.; Corneau, A.; Boussier, J.; Smith, N.; Péré, H.; Charbit, B.; Bondet, V.; Chenevier-Gobeaux, C.; et al. Impaired Type I Interferon Activity and Inflammatory Responses in Severe COVID-19 Patients. Science 2020, 369, 718–724. [Google Scholar] [CrossRef] [PubMed]
  13. Blanco-Melo, D.; Nilsson-Payant, B.E.; Liu, W.-C.; Uhl, S.; Hoagland, D.; Møller, R.; Jordan, T.X.; Oishi, K.; Panis, M.; Sachs, D.; et al. Imbalanced Host Response to SARS-CoV-2 Drives Development of COVID-19. Cell 2020, 181, 1036–1045. [Google Scholar] [CrossRef] [PubMed]
  14. Galani, I.-E.; Rovina, N.; Lampropoulou, V.; Triantafyllia, V.; Manioudaki, M.; Pavlos, E.; Koukaki, E.; Fragkou, P.C.; Panou, V.; Rapti, V.; et al. Untuned Antiviral Immunity in COVID-19 Revealed by Temporal Type I/III Interferon Patterns and Flu Comparison. Nat. Immunol. 2021, 22, 32–40. [Google Scholar] [CrossRef] [PubMed]
  15. Brubaker, S.W.; Bonham, K.S.; Zanoni, I.; Kagan, J.C. Innate Immune Pattern Recognition: A Cell Biological Perspective. Annu. Rev. Immunol. 2015, 33, 257–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Goubau, D.; Deddouche, S.; Reis e Sousa, C. Cytosolic Sensing of Viruses. Immunity 2013, 38, 855–869. [Google Scholar] [CrossRef] [Green Version]
  17. Carty, M.; Guy, C.; Bowie, A.G. Detection of Viral Infections by Innate Immunity. Biochem. Pharmacol. 2021, 183, 114316. [Google Scholar] [CrossRef]
  18. Reikine, S.; Nguyen, J.B.; Modis, Y. Pattern Recognition and Signaling Mechanisms of RIG-I and MDA5. Front. Immunol. 2014, 5, 342. [Google Scholar] [CrossRef] [Green Version]
  19. Kouwaki, T.; Nishimura, T.; Wang, G.; Oshiumi, H. RIG-I-Like Receptor-Mediated Recognition of Viral Genomic RNA of Severe Acute Respiratory Syndrome Coronavirus-2 and Viral Escape from the Host Innate Immune Responses. Front. Immunol. 2021, 12, 700926. [Google Scholar] [CrossRef]
  20. Thorne, L.G.; Reuschl, A.; Zuliani-Alvarez, L.; Whelan, M.V.X.; Turner, J.; Noursadeghi, M.; Jolly, C.; Towers, G.J. SARS-CoV-2 Sensing by RIG-I and MDA5 Links Epithelial Infection to Macrophage Inflammation. EMBO J. 2021, 40, e107826. [Google Scholar] [CrossRef]
  21. Yin, X.; Riva, L.; Pu, Y.; Martin-Sancho, L.; Kanamune, J.; Yamamoto, Y.; Sakai, K.; Gotoh, S.; Miorin, L.; De Jesus, P.D.; et al. MDA5 Governs the Innate Immune Response to SARS-CoV-2 in Lung Epithelial Cells. Cell Rep. 2021, 34, 108628. [Google Scholar] [CrossRef] [PubMed]
  22. Sampaio, N.G.; Chauveau, L.; Hertzog, J.; Bridgeman, A.; Fowler, G.; Moonen, J.P.; Dupont, M.; Russell, R.A.; Noerenberg, M.; Rehwinkel, J. The RNA Sensor MDA5 Detects SARS-CoV-2 Infection. Sci. Rep. 2021, 11, 13638. [Google Scholar] [CrossRef] [PubMed]
  23. Takaoka, A.; Yamada, T. Regulation of Signaling Mediated by Nucleic Acid Sensors for Innate Interferon-Mediated Responses during Viral Infection. Int. Immunol. 2019, 31, 477–488. [Google Scholar] [CrossRef] [PubMed]
  24. Anaeigoudari, A.; Mollaei, H.R.; Arababadi, M.K.; Nosratabadi, R. Severe Acute Respiratory Syndrome Coronavirus 2: The Role of the Main Components of the Innate Immune System. Inflammation 2021, 44, 2151–2169. [Google Scholar] [CrossRef]
  25. 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]
  26. Schindler, C.; Levy, D.E.; Decker, T. JAK-STAT Signaling: From Interferons to Cytokines. J. Biol. Chem. 2007, 282, 20059–20063. [Google Scholar] [CrossRef] [Green Version]
  27. Mesev, E.V.; LeDesma, R.A.; Ploss, A. Decoding Type I and III Interferon Signalling during Viral Infection. Nat. Microbiol. 2019, 4, 914–924. [Google Scholar] [CrossRef]
  28. Schneider, W.M.; Chevillotte, M.D.; Rice, C.M. Interferon-Stimulated Genes: A Complex Web of Host Defenses. Annu. Rev. Immunol. 2014, 32, 513–545. [Google Scholar] [CrossRef] [Green Version]
  29. Schoggins, J.W. Interferon-Stimulated Genes: What Do They All Do? Annu. Rev. Virol. 2019, 6, 567–584. [Google Scholar] [CrossRef]
  30. Hagemeijer, M.C.; Monastyrska, I.; Griffith, J.; van der Sluijs, P.; Voortman, J.; van Bergen en Henegouwen, P.M.; Vonk, A.M.; Rottier, P.J.M.; Reggiori, F.; de Haan, C.A.M. Membrane Rearrangements Mediated by Coronavirus Nonstructural Proteins 3 and 4. Virology 2014, 458, 125–135. [Google Scholar] [CrossRef] [Green Version]
  31. Klein, S.; Cortese, M.; Winter, S.L.; Wachsmuth-Melm, M.; Neufeldt, C.J.; Cerikan, B.; Stanifer, M.L.; Boulant, S.; Bartenschlager, R.; Chlanda, P. SARS-CoV-2 Structure and Replication Characterized by in Situ Cryo-Electron Tomography. Nat. Commun. 2020, 11, 5885. [Google Scholar] [CrossRef] [PubMed]
  32. Romano, M.; Ruggiero, A.; Squeglia, F.; Maga, G.; Berisio, R. A Structural View of SARS-CoV-2 RNA Replication Machinery: RNA Synthesis, Proofreading and Final Capping. Cells 2020, 9, 1267. [Google Scholar] [CrossRef] [PubMed]
  33. Züst, R.; Cervantes-Barragan, L.; Habjan, M.; Maier, R.; Neuman, B.W.; Ziebuhr, J.; Szretter, K.J.; Baker, S.C.; Barchet, W.; Diamond, M.S.; et al. Ribose 2′-O-Methylation Provides a Molecular Signature for the Distinction of Self and Non-Self MRNA Dependent on the RNA Sensor Mda5. Nat. Immunol. 2011, 12, 137–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Leung, D.W.; Amarasinghe, G.K. When Your Cap Matters: Structural Insights into Self vs Non-Self Recognition of 5′ RNA by Immunomodulatory Host Proteins. Curr. Opin. Struct. Biol. 2016, 36, 133–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Chen, K.; Xiao, F.; Hu, D.; Ge, W.; Tian, M.; Wang, W.; Pan, P.; Wu, K.; Wu, J. SARS-CoV-2 Nucleocapsid Protein Interacts with RIG-I and Represses RIG-Mediated IFN-β Production. Viruses 2020, 13, 47. [Google Scholar] [CrossRef] [PubMed]
  36. Shin, D.; Mukherjee, R.; Grewe, D.; Bojkova, D.; Baek, K.; Bhattacharya, A.; Schulz, L.; Widera, M.; Mehdipour, A.R.; Tascher, G.; et al. Papain-like Protease Regulates SARS-CoV-2 Viral Spread and Innate Immunity. Nature 2020, 587, 657–662. [Google Scholar] [CrossRef]
  37. Liu, G.; Lee, J.-H.; Parker, Z.M.; Acharya, D.; Chiang, J.J.; van Gent, M.; Riedl, W.; Davis-Gardner, M.E.; Wies, E.; Chiang, C.; et al. ISG15-Dependent Activation of the Sensor MDA5 Is Antagonized by the SARS-CoV-2 Papain-like Protease to Evade Host Innate Immunity. Nat. Microbiol. 2021, 6, 467–478. [Google Scholar] [CrossRef]
  38. Liu, Y.; Qin, C.; Rao, Y.; Ngo, C.; Feng, J.J.; Zhao, J.; Zhang, S.; Wang, T.-Y.; Carriere, J.; Savas, A.C.; et al. SARS-CoV-2 Nsp5 Demonstrates Two Distinct Mechanisms Targeting RIG-I and MAVS To Evade the Innate Immune Response. mBio 2021, 12, e0233521. [Google Scholar] [CrossRef]
  39. Gori Savellini, G.; Anichini, G.; Gandolfo, C.; Cusi, M.G. SARS-CoV-2 N Protein Targets TRIM25-Mediated RIG-I Activation to Suppress Innate Immunity. Viruses 2021, 13, 1439. [Google Scholar] [CrossRef]
  40. Oh, S.J.; Shin, O.S. SARS-CoV-2 Nucleocapsid Protein Targets RIG-I-Like Receptor Pathways to Inhibit the Induction of Interferon Response. Cells 2021, 10, 530. [Google Scholar] [CrossRef]
  41. Zhao, Y.; Sui, L.; Wu, P.; Wang, W.; Wang, Z.; Yu, Y.; Hou, Z.; Tan, G.; Liu, Q.; Wang, G. A Dual-Role of SARS-CoV-2 Nucleocapsid Protein in Regulating Innate Immune Response. Signal Transduct. Target. Ther. 2021, 6, 331. [Google Scholar] [CrossRef] [PubMed]
  42. Yuen, C.-K.; Lam, J.-Y.; Wong, W.-M.; Mak, L.-F.; Wang, X.; Chu, H.; Cai, J.-P.; Jin, D.-Y.; To, K.K.-W.; Chan, J.F.-W.; et al. SARS-CoV-2 Nsp13, Nsp14, Nsp15 and Orf6 Function as Potent Interferon Antagonists. Emerg. Microbes Infect. 2020, 9, 1418–1428. [Google Scholar] [CrossRef] [PubMed]
  43. Lei, X.; Dong, X.; Ma, R.; Wang, W.; Xiao, X.; Tian, Z.; Wang, C.; Wang, Y.; Li, L.; Ren, L.; et al. Activation and Evasion of Type I Interferon Responses by SARS-CoV-2. Nat. Commun. 2020, 11, 3810. [Google Scholar] [CrossRef] [PubMed]
  44. Xia, H.; Cao, Z.; Xie, X.; Zhang, X.; Chen, J.Y.-C.; Wang, H.; Menachery, V.D.; Rajsbaum, R.; Shi, P.-Y. Evasion of Type I Interferon by SARS-CoV-2. Cell Rep. 2020, 33, 108234. [Google Scholar] [CrossRef] [PubMed]
  45. Li, J.-Y.; Liao, C.-H.; Wang, Q.; Tan, Y.-J.; Luo, R.; Qiu, Y.; Ge, X.-Y. The ORF6, ORF8 and Nucleocapsid Proteins of SARS-CoV-2 Inhibit Type I Interferon Signaling Pathway. Virus Res. 2020, 286, 198074. [Google Scholar] [CrossRef]
  46. Shemesh, M.; Aktepe, T.E.; Deerain, J.M.; McAuley, J.L.; Audsley, M.D.; David, C.T.; Purcell, D.F.J.; Urin, V.; Hartmann, R.; Moseley, G.W.; et al. SARS-CoV-2 Suppresses IFNβ Production Mediated by NSP1, 5, 6, 15, ORF6 and ORF7b but Does Not Suppress the Effects of Added Interferon. PLoS Pathog. 2021, 17, e1009800. [Google Scholar] [CrossRef]
  47. Sato, M.; Suemori, H.; Hata, N.; Asagiri, M.; Ogasawara, K.; Nakao, K.; Nakaya, T.; Katsuki, M.; Noguchi, S.; Tanaka, N.; et al. Distinct and Essential Roles of Transcription Factors IRF-3 and IRF-7 in Response to Viruses for IFNa/b Gene Induction. Immunity 2000, 13, 538–548. [Google Scholar] [CrossRef] [Green Version]
  48. Sato, M.; Hata, N.; Asagiri, M.; Nakaya, T.; Taniguchi, T.; Tanaka, N. Positive Feedback Regulation of Type I IFN Genes by the IFN-Inducible Transcription Factor IRF-7. FEBS Lett. 1998, 441, 106–110. [Google Scholar] [CrossRef] [Green Version]
  49. Kumar, A.; Ishida, R.; Strilets, T.; Cole, J.; Lopez-Orozco, J.; Fayad, N.; Felix-Lopez, A.; Elaish, M.; Evseev, D.; Magor, K.E.; et al. SARS-CoV-2 Nonstructural Protein 1 Inhibits the Interferon Response by Causing Depletion of Key Host Signaling Factors. J. Virol. 2021, 95, e00266-21. [Google Scholar] [CrossRef]
  50. Wang, W.; Zhou, Z.; Xiao, X.; Tian, Z.; Dong, X.; Wang, C.; Li, L.; Ren, L.; Lei, X.; Xiang, Z.; et al. SARS-CoV-2 Nsp12 Attenuates Type I Interferon Production by Inhibiting IRF3 Nuclear Translocation. Cell. Mol. Immunol. 2021, 18, 945–953. [Google Scholar] [CrossRef]
  51. Konno, Y.; Kimura, I.; Uriu, K.; Fukushi, M.; Irie, T.; Koyanagi, Y.; Sauter, D.; Gifford, R.J.; Nakagawa, S.; Sato, K. SARS-CoV-2 ORF3b Is a Potent Interferon Antagonist Whose Activity Is Increased by a Naturally Occurring Elongation Variant. Cell Rep. 2020, 32, 108185. [Google Scholar] [CrossRef] [PubMed]
  52. Li, A.; Zhao, K.; Zhang, B.; Hua, R.; Fang, Y.; Jiang, W.; Zhang, J.; Hui, L.; Zheng, Y.; Li, Y.; et al. SARS-CoV-2 NSP12 Protein Is Not an Interferon-β Antagonist. J. Virol. 2021, 95, e00747-21. [Google Scholar] [CrossRef] [PubMed]
  53. Loo, Y.-M.; Gale, M. Immune Signaling by RIG-I-like Receptors. Immunity 2011, 34, 680–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Fu, Y.-Z.; Wang, S.-Y.; Zheng, Z.-Q.; Huang, Y.; Li, W.-W.; Xu, Z.-S.; Wang, Y.-Y. SARS-CoV-2 Membrane Glycoprotein M Antagonizes the MAVS-Mediated Innate Antiviral Response. Cell. Mol. Immunol. 2021, 18, 613–620. [Google Scholar] [CrossRef]
  55. Sui, L.; Zhao, Y.; Wang, W.; Wu, P.; Wang, Z.; Yu, Y.; Hou, Z.; Tan, G.; Liu, Q. SARS-CoV-2 Membrane Protein Inhibits Type I Interferon Production Through Ubiquitin-Mediated Degradation of TBK1. Front. Immunol. 2021, 12, 662989. [Google Scholar] [CrossRef]
  56. Zheng, Y.; Zhuang, M.-W.; Han, L.; Zhang, J.; Nan, M.-L.; Zhan, P.; Kang, D.; Liu, X.; Gao, C.; Wang, P.-H. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) Membrane (M) Protein Inhibits Type I and III Interferon Production by Targeting RIG-I/MDA-5 Signaling. Signal Transduct. Target. Ther. 2020, 5, 299. [Google Scholar] [CrossRef]
  57. Vazquez, C.; Swanson, S.E.; Negatu, S.G.; Dittmar, M.; Miller, J.; Ramage, H.R.; Cherry, S.; Jurado, K.A. SARS-CoV-2 Viral Proteins NSP1 and NSP13 Inhibit Interferon Activation through Distinct Mechanisms. PLoS ONE 2021, 16, e0253089. [Google Scholar] [CrossRef]
  58. Han, L.; Zhuang, M.-W.; Deng, J.; Zheng, Y.; Zhang, J.; Nan, M.-L.; Zhang, X.-J.; Gao, C.; Wang, P.-H. SARS-CoV-2 ORF9b Antagonizes Type I and III Interferons by Targeting Multiple Components of the RIG-I/MDA-5-MAVS, TLR3-TRIF, and CGAS-STING Signaling Pathways. J. Med. Virol. 2021, 93, 5376–5389. [Google Scholar] [CrossRef]
  59. Jiang, H.-W.; Zhang, H.-N.; Meng, Q.-F.; Xie, J.; Li, Y.; Chen, H.; Zheng, Y.-X.; Wang, X.-N.; Qi, H.; Zhang, J.; et al. SARS-CoV-2 Orf9b Suppresses Type I Interferon Responses by Targeting TOM70. Cell. Mol. Immunol. 2020, 17, 998–1000. [Google Scholar] [CrossRef]
  60. Wu, J.; Shi, Y.; Pan, X.; Wu, S.; Hou, R.; Zhang, Y.; Zhong, T.; Tang, H.; Du, W.; Wang, L.; et al. SARS-CoV-2 ORF9b Inhibits RIG-I-MAVS Antiviral Signaling by Interrupting K63-Linked Ubiquitination of NEMO. Cell Rep. 2021, 34, 108761. [Google Scholar] [CrossRef]
  61. Makiyama, K.; Hazawa, M.; Kobayashi, A.; Lim, K.; Voon, D.C.; Wong, R.W. NSP9 of SARS-CoV-2 Attenuates Nuclear Transport by Hampering Nucleoporin 62 Dynamics and Functions in Host Cells. Biochem. Biophys. Res. Commun. 2022, 586, 137–142. [Google Scholar] [CrossRef]
  62. Moustaqil, M.; Ollivier, E.; Chiu, H.-P.; Van Tol, S.; Rudolffi-Soto, P.; Stevens, C.; Bhumkar, A.; Hunter, D.J.B.; Freiberg, A.N.; Jacques, D.; et al. SARS-CoV-2 Proteases PLpro and 3CLpro Cleave IRF3 and Critical Modulators of Inflammatory Pathways (NLRP12 and TAB1): Implications for Disease Presentation across Species. Emerg. Microbes Infect. 2021, 10, 178–195. [Google Scholar] [CrossRef] [PubMed]
  63. Fung, S.-Y.; Siu, K.-L.; Lin, H.; Yeung, M.L.; Jin, D.-Y. SARS-CoV-2 Main Protease Suppresses Type I Interferon Production by Preventing Nuclear Translocation of Phosphorylated IRF3. Int. J. Biol. Sci. 2021, 17, 1547–1554. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, W.; Ma, Z.; Wu, Y.; Shi, X.; Zhang, Y.; Zhang, M.; Zhang, M.; Wang, L.; Liu, W. SARS-CoV-2 3C-like Protease Antagonizes Interferon-Beta Production by Facilitating the Degradation of IRF3. Cytokine 2021, 148, 155697. [Google Scholar] [CrossRef] [PubMed]
  65. Hayn, M.; Hirschenberger, M.; Koepke, L.; Nchioua, R.; Straub, J.H.; Klute, S.; Hunszinger, V.; Zech, F.; Prelli Bozzo, C.; Aftab, W.; et al. Systematic Functional Analysis of SARS-CoV-2 Proteins Uncovers Viral Innate Immune Antagonists and Remaining Vulnerabilities. Cell Rep. 2021, 35, 109126. [Google Scholar] [CrossRef]
  66. Kimura, I.; Konno, Y.; Uriu, K.; Hopfensperger, K.; Sauter, D.; Nakagawa, S.; Sato, K. Sarbecovirus ORF6 Proteins Hamper Induction of Interferon Signaling. Cell Rep. 2021, 34, 108916. [Google Scholar] [CrossRef]
  67. Miorin, L.; Kehrer, T.; Sanchez-Aparicio, M.T.; Zhang, K.; Cohen, P.; Patel, R.S.; Cupic, A.; Makio, T.; Mei, M.; Moreno, E.; et al. SARS-CoV-2 Orf6 Hijacks Nup98 to Block STAT Nuclear Import and Antagonize Interferon Signaling. Proc. Natl. Acad. Sci. USA 2020, 117, 28344–28354. [Google Scholar] [CrossRef]
  68. Addetia, A.; Lieberman, N.A.P.; Phung, Q.; Hsiang, T.-Y.; Xie, H.; Roychoudhury, P.; Shrestha, L.; Loprieno, M.A.; Huang, M.-L.; Gale, M.; et al. SARS-CoV-2 ORF6 Disrupts Bidirectional Nucleocytoplasmic Transport through Interactions with Rae1 and Nup98. mBio 2021, 12, e00065-21. [Google Scholar] [CrossRef]
  69. Mu, J.; Fang, Y.; Yang, Q.; Shu, T.; Wang, A.; Huang, M.; Jin, L.; Deng, F.; Qiu, Y.; Zhou, X. SARS-CoV-2 N Protein Antagonizes Type I Interferon Signaling by Suppressing Phosphorylation and Nuclear Translocation of STAT1 and STAT2. Cell Discov. 2020, 6, 65. [Google Scholar] [CrossRef]
  70. Cao, Z.; Xia, H.; Rajsbaum, R.; Xia, X.; Wang, H.; Shi, P.-Y. Ubiquitination of SARS-CoV-2 ORF7a Promotes Antagonism of Interferon Response. Cell. Mol. Immunol. 2021, 18, 746–748. [Google Scholar] [CrossRef]
  71. Banerjee, A.K.; Blanco, M.R.; Bruce, E.A.; Honson, D.D.; Chen, L.M.; Chow, A.; Bhat, P.; Ollikainen, N.; Quinodoz, S.A.; Loney, C.; et al. SARS-CoV-2 Disrupts Splicing, Translation, and Protein Trafficking to Suppress Host Defenses. Cell 2020, 183, 1325–1339. [Google Scholar] [CrossRef] [PubMed]
  72. Burke, J.M.; Clair, L.A.S.; Perera, R.; Parker, R. SARS-CoV-2 Infection Triggers Widespread Host MRNA Decay Leading to an MRNA Export Block. RNA 2021, 27, 1318–1329. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, K.; Miorin, L.; Makio, T.; Dehghan, I.; Gao, S.; Xie, Y.; Zhong, H.; Esparza, M.; Kehrer, T.; Kumar, A.; et al. Nsp1 Protein of SARS-CoV-2 Disrupts the MRNA Export Machinery to Inhibit Host Gene Expression. Sci. Adv. 2021, 7, eabe7386. [Google Scholar] [CrossRef] [PubMed]
  74. Yuan, S.; Peng, L.; Park, J.J.; Hu, Y.; Devarkar, S.C.; Dong, M.B.; Shen, Q.; Wu, S.; Chen, S.; Lomakin, I.B.; et al. Nonstructural Protein 1 of SARS-CoV-2 Is a Potent Pathogenicity Factor Redirecting Host Protein Synthesis Machinery toward Viral RNA. Mol. Cell 2020, 80, 1055–1066. [Google Scholar] [CrossRef] [PubMed]
  75. Thoms, M.; Buschauer, R.; Ameismeier, M.; Koepke, L.; Denk, T.; Hirschenberger, M.; Kratzat, H.; Hayn, M.; Mackens-Kiani, T.; Cheng, J.; et al. Structural Basis for Translational Shutdown and Immune Evasion by the Nsp1 Protein of SARS-CoV-2. Science 2020, 369, 1249–1255. [Google Scholar] [CrossRef] [PubMed]
  76. Schubert, K.; Karousis, E.D.; Jomaa, A.; Scaiola, A.; Echeverria, B.; Gurzeler, L.-A.; Leibundgut, M.; Thiel, V.; Mühlemann, O.; Ban, N. SARS-CoV-2 Nsp1 Binds the Ribosomal MRNA Channel to Inhibit Translation. Nat. Struct. Mol. Biol. 2020, 27, 959–966. [Google Scholar] [CrossRef]
  77. Lapointe, C.P.; Grosely, R.; Johnson, A.G.; Wang, J.; Fernández, I.S.; Puglisi, J.D. Dynamic Competition between SARS-CoV-2 NSP1 and MRNA on the Human Ribosome Inhibits Translation Initiation. Proc. Natl. Acad. Sci. USA 2021, 118, e2017715118. [Google Scholar] [CrossRef]
  78. Mendez, A.S.; Ly, M.; González-Sánchez, A.M.; Hartenian, E.; Ingolia, N.T.; Cate, J.H.; Glaunsinger, B.A. The N-Terminal Domain of SARS-CoV-2 Nsp1 Plays Key Roles in Suppression of Cellular Gene Expression and Preservation of Viral Gene Expression. Cell Rep. 2021, 37, 109841. [Google Scholar] [CrossRef]
  79. Trouillet-Assant, S.; Viel, S.; Gaymard, A.; Pons, S.; Richard, J.-C.; Perret, M.; Villard, M.; Brengel-Pesce, K.; Lina, B.; Mezidi, M.; et al. Type I IFN Immunoprofiling in COVID-19 Patients. J. Allergy Clin. Immunol. 2020, 146, 206–208. [Google Scholar] [CrossRef]
  80. Contoli, M.; Papi, A.; Tomassetti, L.; Rizzo, P.; Vieceli Dalla Sega, F.; Fortini, F.; Torsani, F.; Morandi, L.; Ronzoni, L.; Zucchetti, O.; et al. Blood Interferon-α Levels and Severity, Outcomes, and Inflammatory Profiles in Hospitalized COVID-19 Patients. Front. Immunol. 2021, 12, 648004. [Google Scholar] [CrossRef]
  81. Zhang, Q.; Bastard, P.; Liu, Z.; Le Pen, J.; Moncada-Velez, M.; Chen, J.; Ogishi, M.; Sabli, I.K.D.; Hodeib, S.; Korol, C.; et al. Inborn Errors of Type I IFN Immunity in Patients with Life-Threatening COVID-19. Science 2020, 370, eabd4570. [Google Scholar] [CrossRef] [PubMed]
  82. Van der Made, C.I.; Simons, A.; Schuurs-Hoeijmakers, J.; van den Heuvel, G.; Mantere, T.; Kersten, S.; van Deuren, R.C.; Steehouwer, M.; van Reijmersdal, S.V.; Jaeger, M.; et al. Presence of Genetic Variants Among Young Men with Severe COVID-19. JAMA 2020, 324, 663–673. [Google Scholar] [CrossRef] [PubMed]
  83. Abolhassani, H.; Vosughimotlagh, A.; Asano, T.; Landegren, N.; Boisson, B.; Delavari, S.; Bastard, P.; Aranda-Guillén, M.; Wang, Y.; Zuo, F.; et al. X-Linked TLR7 Deficiency Underlies Critical COVID-19 Pneumonia in a Male Patient with Ataxia-Telangiectasia. J. Clin. Immunol. 2022, 42, 1–9. [Google Scholar] [CrossRef] [PubMed]
  84. Bastard, P.; Rosen, L.B.; Zhang, Q.; Michailidis, E.; Hoffmann, H.-H.; Zhang, Y.; Dorgham, K.; Philippot, Q.; Rosain, J.; Béziat, V.; et al. Autoantibodies against Type I IFNs in Patients with Life-Threatening COVID-19. Science 2020, 370, eabd4585. [Google Scholar] [CrossRef]
  85. Koning, R.; Bastard, P.; Casanova, J.-L.; Brouwer, M.C.; van de Beek, D. Autoantibodies against Type I Interferons Are Associated with Multi-Organ Failure in COVID-19 Patients. Intensive Care Med. 2021, 47, 704–706. [Google Scholar] [CrossRef]
  86. Troya, J.; Bastard, P.; Planas-Serra, L.; Ryan, P.; Ruiz, M.; de Carranza, M.; Torres, J.; Martínez, A.; Abel, L.; Casanova, J.-L.; et al. Neutralizing Autoantibodies to Type I IFNs in >10% of Patients with Severe COVID-19 Pneumonia Hospitalized in Madrid, Spain. J. Clin. Immunol. 2021, 41, 914–922. [Google Scholar] [CrossRef]
  87. Van der Wijst, M.G.P.; Vazquez, S.E.; Hartoularos, G.C.; Bastard, P.; Grant, T.; Bueno, R.; Lee, D.S.; Greenland, J.R.; Sun, Y.; Perez, R.; et al. Type I Interferon Autoantibodies Are Associated with Systemic Immune Alterations in Patients with COVID-19. Sci. Transl. Med. 2021, 13, eabh2624. [Google Scholar] [CrossRef]
  88. Bastard, P.; Gervais, A.; Le Voyer, T.; Rosain, J.; Philippot, Q.; Manry, J.; Michailidis, E.; Hoffmann, H.-H.; Eto, S.; Garcia-Prat, M.; et al. Autoantibodies Neutralizing Type I IFNs Are Present in ~4% of Uninfected Individuals over 70 Years and Account for ~20% of COVID-19 Deaths. Sci. Immunol. 2021, 6, eabl4340. [Google Scholar] [CrossRef]
  89. Felgenhauer, U.; Schoen, A.; Gad, H.H.; Hartmann, R.; Schaubmar, A.R.; Failing, K.; Drosten, C.; Weber, F. Inhibition of SARS-CoV-2 by Type I and Type III Interferons. J. Biol. Chem. 2020, 295, 13958–13964. [Google Scholar] [CrossRef]
  90. Mantlo, E.; Bukreyeva, N.; Maruyama, J.; Paessler, S.; Huang, C. Antiviral Activities of Type I Interferons to SARS-CoV-2 Infection. Antiviral Res. 2020, 179, 104811. [Google Scholar] [CrossRef]
  91. Dinnon, K.H.; Leist, S.R.; Schäfer, A.; Edwards, C.E.; Martinez, D.R.; Montgomery, S.A.; West, A.; Yount, B.L.; Hou, Y.J.; Adams, L.E.; et al. A Mouse-Adapted Model of SARS-CoV-2 to Test COVID-19 Countermeasures. Nature 2020, 586, 560–566. [Google Scholar] [CrossRef] [PubMed]
  92. Vanderheiden, A.; Ralfs, P.; Chirkova, T.; Upadhyay, A.A.; Zimmerman, M.G.; Bedoya, S.; Aoued, H.; Tharp, G.M.; Pellegrini, K.L.; Manfredi, C.; et al. Type I and Type III Interferons Restrict SARS-CoV-2 Infection of Human Airway Epithelial Cultures. J. Virol. 2020, 94, e00985-20. [Google Scholar] [CrossRef] [PubMed]
  93. Hoagland, D.A.; Møller, R.; Uhl, S.A.; Oishi, K.; Frere, J.; Golynker, I.; Horiuchi, S.; Panis, M.; Blanco-Melo, D.; Sachs, D.; et al. Leveraging the Antiviral Type I Interferon System as a First Line of Defense against SARS-CoV-2 Pathogenicity. Immunity 2021, 54, 557–570. [Google Scholar] [CrossRef] [PubMed]
  94. Aghemo, A.; Rumi, M.G.; Colombo, M. Pegylated Interferons A2a and A2b in the Treatment of Chronic Hepatitis C. Nat. Rev. Gastroenterol. Hepatol. 2010, 7, 485–494. [Google Scholar] [CrossRef]
  95. Broquetas, T.; Garcia-Retortillo, M.; Micó, M.; Canillas, L.; Puigvehí, M.; Cañete, N.; Coll, S.; Viu, A.; Hernandez, J.J.; Bessa, X.; et al. Hepatitis B Surface Antigen and Hepatitis B Core-Related Antigen Kinetics after Adding Pegylated-Interferon to Nucleos(t)Ids Analogues in Hepatitis B e Antigen-Negative Patients. World J. Hepatol. 2020, 12, 1076–1088. [Google Scholar] [CrossRef]
  96. Zhou, Q.; Chen, V.; Shannon, C.P.; Wei, X.-S.; Xiang, X.; Wang, X.; Wang, Z.-H.; Tebbutt, S.J.; Kollmann, T.R.; Fish, E.N. Interferon-A2b Treatment for COVID-19. Front. Immunol. 2020, 11, 1061. [Google Scholar] [CrossRef]
  97. Pandit, A.; Bhalani, N.; Bhushan, B.L.S.; Koradia, P.; Gargiya, S.; Bhomia, V.; Kansagra, K. Efficacy and Safety of Pegylated Interferon Alfa-2b in Moderate COVID-19: A Phase II, Randomized, Controlled, Open-Label Study. Int. J. Infect. Dis. 2021, 105, 516–521. [Google Scholar] [CrossRef]
  98. Bhushan, B.L.S.; Wanve, S.; Koradia, P.; Bhomia, V.; Soni, P.; Chakraborty, S.; Khobragade, A.; Joshi, S.; Mendiratta, S.K.; Kansagra, K.K.; et al. Efficacy and Safety of Pegylated Interferon-A2b in Moderate COVID-19: A Phase 3, Randomized, Comparator-Controlled, Open-Label Study. Int. J. Infect. Dis. 2021, 111, 281–287. [Google Scholar] [CrossRef]
  99. Wang, B.; Li, D.; Liu, T.; Wang, H.; Luo, F.; Liu, Y. Subcutaneous Injection of IFN Alpha-2b for COVID-19: An Observational Study. BMC Infect. Dis. 2020, 20, 723. [Google Scholar] [CrossRef]
  100. Wang, N.; Zhan, Y.; Zhu, L.; Hou, Z.; Liu, F.; Song, P.; Qiu, F.; Wang, X.; Zou, X.; Wan, D.; et al. Retrospective Multicenter Cohort Study Shows Early Interferon Therapy Is Associated with Favorable Clinical Responses in COVID-19 Patients. Cell Host Microbe 2020, 28, 455–464. [Google Scholar] [CrossRef]
  101. Yu, J.; Lu, X.; Tong, L.; Shi, X.; Ma, J.; Lv, F.; Wu, J.; Pan, Q.; Yang, J.; Cao, H.; et al. Interferon-α-2b Aerosol Inhalation Is Associated with Improved Clinical Outcomes in Patients with Coronavirus Disease-2019. Br. J. Clin. Pharmacol. 2021, 87, 4737–4746. [Google Scholar] [CrossRef] [PubMed]
  102. Ziegler, C.G.K.; Allon, S.J.; Nyquist, S.K.; Mbano, I.M.; Miao, V.N.; Tzouanas, C.N.; Cao, Y.; Yousif, A.S.; Bals, J.; Hauser, B.M.; et al. SARS-CoV-2 Receptor ACE2 Is an Interferon-Stimulated Gene in Human Airway Epithelial Cells and Is Detected in Specific Cell Subsets across Tissues. Cell 2020, 181, 1016–1035. [Google Scholar] [CrossRef]
  103. Blume, C.; Jackson, C.L.; Spalluto, C.M.; Legebeke, J.; Nazlamova, L.; Conforti, F.; Perotin, J.-M.; Frank, M.; Butler, J.; Crispin, M.; et al. A Novel ACE2 Isoform Is Expressed in Human Respiratory Epithelia and Is Upregulated in Response to Interferons and RNA Respiratory Virus Infection. Nat. Genet. 2021, 53, 205–214. [Google Scholar] [CrossRef] [PubMed]
  104. Ng, K.W.; Attig, J.; Bolland, W.; Young, G.R.; Major, J.; Wrobel, A.G.; Gamblin, S.; Wack, A.; Kassiotis, G. Tissue-Specific and Interferon-Inducible Expression of Nonfunctional ACE2 through Endogenous Retroelement Co-Option. Nat. Genet. 2020, 52, 1294–1302. [Google Scholar] [CrossRef] [PubMed]
  105. Onabajo, O.O.; Banday, A.R.; Yan, W.; Obajemu, A.; Stanifer, M.L.; Santer, D.M.; Florez-Vargas, O.; Piontkivska, H.; Vargas, J.; Kee, C.; et al. Interferons and Viruses Induce a Novel Truncated ACE2 Isoform and Not the full-length SARS-CoV-2 Receptor. Nat. Genet. 2020, 52, 1283–1293. [Google Scholar] [CrossRef] [PubMed]
  106. Kiss, J.; Yegutkin, G.G.; Koskinen, K.; Savunen, T.; Jalkanen, S.; Salmi, M. IFN-β Protects from Vascular Leakage via up-Regulation of CD73. Eur. J. Immunol. 2007, 37, 3334–3338. [Google Scholar] [CrossRef]
  107. Aeffner, F.; Woods, P.S.; Davis, I.C. Activation of A1-Adenosine Receptors Promotes Leukocyte Recruitment to the Lung and Attenuates Acute Lung Injury in Mice Infected with Influenza A/WSN/33 (H1N1) Virus. J. Virol. 2014, 88, 10214–10227. [Google Scholar] [CrossRef] [Green Version]
  108. Ranieri, V.M.; Pettilä, V.; Karvonen, M.K.; Jalkanen, J.; Nightingale, P.; Brealey, D.; Mancebo, J.; Ferrer, R.; Mercat, A.; Patroniti, N.; et al. Effect of Intravenous Interferon β-1a on Death and Days Free from Mechanical Ventilation Among Patients with Moderate to Severe Acute Respiratory Distress Syndrome: A Randomized Clinical Trial. JAMA 2020, 323, 725–733. [Google Scholar] [CrossRef]
  109. Hung, I.F.-N.; Lung, K.-C.; Tso, E.Y.-K.; Liu, R.; Chung, T.W.-H.; Chu, M.-Y.; Ng, Y.-Y.; Lo, J.; Chan, J.; Tam, A.R.; et al. Triple Combination of Interferon Beta-1b, Lopinavir–Ritonavir, and Ribavirin in the Treatment of Patients Admitted to Hospital with COVID-19: An Open-Label, Randomised, Phase 2 Trial. Lancet 2020, 395, 1695–1704. [Google Scholar] [CrossRef]
  110. Alavi Darazam, I.; Shokouhi, S.; Pourhoseingholi, M.A.; Naghibi Irvani, S.S.; Mokhtari, M.; Shabani, M.; Amirdosara, M.; Torabinavid, P.; Golmohammadi, M.; Hashemi, S.; et al. Role of Interferon Therapy in Severe COVID-19: The COVIFERON Randomized Controlled Trial. Sci. Rep. 2021, 11, 8059. [Google Scholar] [CrossRef]
  111. Davoudi-Monfared, E.; Rahmani, H.; Khalili, H.; Hajiabdolbaghi, M.; Salehi, M.; Abbasian, L.; Kazemzadeh, H.; Yekaninejad, M.S. A Randomized Clinical Trial of the Efficacy and Safety of Interferon β-1a in Treatment of Severe COVID-19. Antimicrob. Agents Chemother. 2020, 64, e01061-20. [Google Scholar] [CrossRef] [PubMed]
  112. WHO Solidarity Trial Consortium. Repurposed Antiviral Drugs for COVID-19—Interim WHO Solidarity Trial Results. N. Engl. J. Med. 2021, 384, 497–511. [Google Scholar] [CrossRef] [PubMed]
  113. Kalil, A.C.; Mehta, A.K.; Patterson, T.F.; Erdmann, N.; Gomez, C.A.; Jain, M.K.; Wolfe, C.R.; Ruiz-Palacios, G.M.; Kline, S.; Regalado Pineda, J.; et al. Efficacy of Interferon Beta-1a plus Remdesivir Compared with Remdesivir Alone in Hospitalised Adults with COVID-19: A Double-Bind, Randomised, Placebo-Controlled, Phase 3 Trial. Lancet Respir. Med. 2021, 9, 1365–1376. [Google Scholar] [CrossRef]
  114. National Institutes of Health. Coronavirus Disease 2019 (COVID-19) Treatment Guidelines; National Institutes of Health: Bethesda, MD, USA, 2022; p. 394. [Google Scholar]
  115. Monk, P.D.; Marsden, R.J.; Tear, V.J.; Brookes, J.; Batten, T.N.; Mankowski, M.; Gabbay, F.J.; Davies, D.E.; Holgate, S.T.; Ho, L.-P.; et al. Safety and Efficacy of Inhaled Nebulised Interferon Beta-1a (SNG001) for Treatment of SARS-CoV-2 Infection: A Randomised, Double-Blind, Placebo-Controlled, Phase 2 Trial. Lancet Respir. Med. 2021, 9, 196–206. [Google Scholar] [CrossRef]
  116. Khamis, F.; Al Naabi, H.; Al Lawati, A.; Ambusaidi, Z.; Al Sharji, M.; Al Barwani, U.; Pandak, N.; Al Balushi, Z.; Al Bahrani, M.; Al Salmi, I.; et al. Randomized Controlled Open Label Trial on the Use of Favipiravir Combined with Inhaled Interferon Beta-1b in Hospitalized Patients with Moderate to Severe COVID-19 Pneumonia. Int. J. Infect. Dis. 2021, 102, 538–543. [Google Scholar] [CrossRef] [PubMed]
  117. Klinkhammer, J.; Schnepf, D.; Ye, L.; Schwaderlapp, M.; Gad, H.H.; Hartmann, R.; Garcin, D.; Mahlakõiv, T.; Staeheli, P. IFN-λ Prevents Influenza Virus Spread from the Upper Airways to the Lungs and Limits Virus Transmission. eLife 2018, 7, e33354. [Google Scholar] [CrossRef]
  118. Davidson, S.; McCabe, T.M.; Crotta, S.; Gad, H.H.; Hessel, E.M.; Beinke, S.; Hartmann, R.; Wack, A. IFNλ Is a Potent Anti-influenza Therapeutic without the Inflammatory Side Effects of IFNα Treatment. EMBO Mol. Med. 2016, 8, 1099–1112. [Google Scholar] [CrossRef]
  119. Forero, A.; Ozarkar, S.; Li, H.; Lee, C.H.; Hemann, E.A.; Nadjsombati, M.S.; Hendricks, M.R.; So, L.; Green, R.; Roy, C.N.; et al. Differential Activation of the Transcription Factor IRF1 Underlies the Distinct Immune Responses Elicited by Type I and Type III Interferons. Immunity 2019, 51, 451–464. [Google Scholar] [CrossRef]
  120. Galani, I.E.; Triantafyllia, V.; Eleminiadou, E.-E.; Koltsida, O.; Stavropoulos, A.; Manioudaki, M.; Thanos, D.; Doyle, S.E.; Kotenko, S.V.; Thanopoulou, K.; et al. Interferon-λ Mediates Non-Redundant Front-Line Antiviral Protection against Influenza Virus Infection without Compromising Host Fitness. Immunity 2017, 46, 875–890.e6. [Google Scholar] [CrossRef]
  121. Broggi, A.; Tan, Y.; Granucci, F.; Zanoni, I. IFN-λ Suppresses Intestinal Inflammation by Non-Translational Regulation of Neutrophil Function. Nat. Immunol. 2017, 18, 1084–1093. [Google Scholar] [CrossRef]
  122. Blazek, K.; Eames, H.L.; Weiss, M.; Byrne, A.J.; Perocheau, D.; Pease, J.E.; Doyle, S.; McCann, F.; Williams, R.O.; Udalova, I.A. IFN-λ Resolves Inflammation via Suppression of Neutrophil Infiltration and IL-1β Production. J. Exp. Med. 2015, 212, 845–853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Stanifer, M.L.; Kee, C.; Cortese, M.; Zumaran, C.M.; Triana, S.; Mukenhirn, M.; Kraeusslich, H.-G.; Alexandrov, T.; Bartenschlager, R.; Boulant, S. Critical Role of Type III Interferon in Controlling SARS-CoV-2 Infection in Human Intestinal Epithelial Cells. Cell Rep. 2020, 32, 107863. [Google Scholar] [CrossRef] [PubMed]
  124. Sohn, S.-Y.; Hearing, J.; Mugavero, J.; Kirillov, V.; Gorbunova, E.; Helminiak, L.; Mishra, S.; Mackow, E.; Hearing, P.; Reich, N.C.; et al. Interferon-Lambda Intranasal Protection and Differential Sex Pathology in a Murine Model of SARS-CoV-2 Infection. mBio 2021, 12, e02756-21. [Google Scholar] [CrossRef] [PubMed]
  125. Liou, T.G.; Adler, F.R.; Cahill, B.C.; Cox, D.R.; Cox, J.E.; Grant, G.J.; Hanson, K.E.; Hartsell, S.C.; Hatton, N.D.; Helms, M.N.; et al. SARS-CoV-2 Innate Effector Associations and Viral Load in Early Nasopharyngeal Infection. Physiol. Rep. 2021, 9, e14761. [Google Scholar] [CrossRef] [PubMed]
  126. Fukuda, Y.; Homma, T.; Inoue, H.; Onitsuka, C.; Ikeda, H.; Goto, Y.; Sato, Y.; Kimura, T.; Hirai, K.; Ohta, S.; et al. Downregulation of Type III Interferons in Patients with Severe COVID-19. J. Med. Virol. 2021, 93, 4559–4563. [Google Scholar] [CrossRef]
  127. Shahbazi, M.; Amri Maleh, P.; Bagherzadeh, M.; Moulana, Z.; Sepidarkish, M.; Rezanejad, M.; Mirzakhani, M.; Ebrahimpour, S.; Ghorbani, H.; Ahmadnia, Z.; et al. Linkage of Lambda Interferons in Protection Against Severe COVID-19. J. Interferon Cytokine Res. 2021, 41, 149–152. [Google Scholar] [CrossRef]
  128. Fukuda, Y.; Homma, T.; Inoue, H.; Goto, Y.; Sato, Y.; Ikeda, H.; Onitsuka, C.; Sato, H.; Akimoto, K.; Ebato, T.; et al. Serum IL-28A/IFN-Λ2 Is Linked to Disease Severity of COVID-19. Sci. Rep. 2022, 12, 5458. [Google Scholar] [CrossRef]
  129. O’Brien, T.R.; Thomas, D.L.; Jackson, S.S.; Prokunina-Olsson, L.; Donnelly, R.P.; Hartmann, R. Weak Induction of Interferon Expression by Severe Acute Respiratory Syndrome Coronavirus 2 Supports Clinical Trials of Interferon-λ to Treat Early Coronavirus Disease 2019. Clin. Infect. Dis. 2020, 71, 1410–1412. [Google Scholar] [CrossRef]
  130. Feld, J.J.; Kandel, C.; Biondi, M.J.; Kozak, R.A.; Zahoor, M.A.; Lemieux, C.; Borgia, S.M.; Boggild, A.K.; Powis, J.; McCready, J.; et al. Peginterferon Lambda for the Treatment of Outpatients with COVID-19: A Phase 2, Placebo-Controlled Randomised Trial. Lancet Respir. Med. 2021, 9, 498–510. [Google Scholar] [CrossRef]
  131. Jagannathan, P.; Andrews, J.R.; Bonilla, H.; Hedlin, H.; Jacobson, K.B.; Balasubramanian, V.; Purington, N.; Kamble, S.; de Vries, C.R.; Quintero, O.; et al. Peginterferon Lambda-1a for Treatment of Outpatients with Uncomplicated COVID-19: A Randomized Placebo-Controlled Trial. Nat. Commun. 2021, 12, 1967. [Google Scholar] [CrossRef]
  132. Planet, P.J.; Parker, D.; Cohen, T.S.; Smith, H.; Leon, J.D.; Ryan, C.; Hammer, T.J.; Fierer, N.; Chen, E.I.; Prince, A.S. Lambda Interferon Restructures the Nasal Microbiome and Increases Susceptibility to Staphylococcus Aureus Superinfection. mBio 2016, 7, e01939-15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Rich, H.E.; McCourt, C.C.; Zheng, W.Q.; McHugh, K.J.; Robinson, K.M.; Wang, J.; Alcorn, J.F. Interferon Lambda Inhibits Bacterial Uptake during Influenza Superinfection. Infect. Immun. 2019, 87, e00114-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Chen, X.; Saccon, E.; Appelberg, K.S.; Mikaeloff, F.; Rodriguez, J.E.; Vinhas, B.S.; Frisan, T.; Végvári, Á.; Mirazimi, A.; Neogi, U.; et al. Type-I Interferon Signatures in SARS-CoV-2 Infected Huh7 Cells. Cell Death Discov. 2021, 7, 114. [Google Scholar] [CrossRef] [PubMed]
  135. Lokugamage, K.G.; Hage, A.; de Vries, M.; Valero-Jimenez, A.M.; Schindewolf, C.; Dittmann, M.; Rajsbaum, R.; Menachery, V.D. Type I Interferon Susceptibility Distinguishes SARS-CoV-2 from SARS-CoV. J. Virol. 2020, 94, e01410-20. [Google Scholar] [CrossRef]
  136. Bessière, P.; Wasniewski, M.; Picard-Meyer, E.; Servat, A.; Figueroa, T.; Foret-Lucas, C.; Coggon, A.; Lesellier, S.; Boué, F.; Cebron, N.; et al. Intranasal Type I Interferon Treatment Is Beneficial Only When Administered before Clinical Signs Onset in the SARS-CoV-2 Hamster Model. PLOS Pathog. 2021, 17, e1009427. [Google Scholar] [CrossRef]
  137. Sposito, B.; Broggi, A.; Pandolfi, L.; Crotta, S.; Clementi, N.; Ferrarese, R.; Sisti, S.; Criscuolo, E.; Spreafico, R.; Long, J.M.; et al. The Interferon Landscape along the Respiratory Tract Impacts the Severity of COVID-19. Cell 2021, 184, 4953–4968. [Google Scholar] [CrossRef]
  138. Ziegler, C.G.K.; Miao, V.N.; Owings, A.H.; Navia, A.W.; Tang, Y.; Bromley, J.D.; Lotfy, P.; Sloan, M.; Laird, H.; Williams, H.B.; et al. Impaired Local Intrinsic Immunity to SARS-CoV-2 Infection in Severe COVID-19. Cell 2021, 184, 4713–4733. [Google Scholar] [CrossRef]
  139. Lucas, C.; Wong, P.; Klein, J.; Castro, T.B.R.; Silva, J.; Sundaram, M.; Ellingson, M.K.; Mao, T.; Oh, J.E.; Israelow, B.; et al. Longitudinal Analyses Reveal Immunological Misfiring in Severe COVID-19. Nature 2020, 584, 463–469. [Google Scholar] [CrossRef]
  140. Lee, J.S.; Park, S.; Jeong, H.W.; Ahn, J.Y.; Choi, S.J.; Lee, H.; Choi, B.; Nam, S.K.; Sa, M.; Kwon, J.-S.; et al. Immunophenotyping of COVID-19 and Influenza Highlights the Role of Type I Interferons in Development of Severe COVID-19. Sci. Immunol. 2020, 5, eabd1554. [Google Scholar] [CrossRef]
  141. Amoutzias, G.D.; Nikolaidis, M.; Tryfonopoulou, E.; Chlichlia, K.; Markoulatos, P.; Oliver, S.G. The Remarkable Evolutionary Plasticity of Coronaviruses by Mutation and Recombination: Insights for the COVID-19 Pandemic and the Future Evolutionary Paths of SARS-CoV-2. Viruses 2022, 14, 78. [Google Scholar] [CrossRef]
  142. Min, Y.-Q.; Mo, Q.; Wang, J.; Deng, F.; Wang, H.; Ning, Y.-J. SARS-CoV-2 Nsp1: Bioinformatics, Potential Structural and Functional Features, and Implications for Drug/Vaccine Designs. Front. Microbiol. 2020, 11, 587317. [Google Scholar] [CrossRef] [PubMed]
  143. Afsar, M.; Narayan, R.; Akhtar, M.N.; Das, D.; Rahil, H.; Nagaraj, S.K.; Eswarappa, S.M.; Tripathi, S.; Hussain, T. Drug Targeting Nsp1-Ribosomal Complex Shows Antiviral Activity against SARS-CoV-2. eLife 2022, 11, e74877. [Google Scholar] [CrossRef] [PubMed]
  144. Ratia, K.; Pegan, S.; Takayama, J.; Sleeman, K.; Coughlin, M.; Baliji, S.; Chaudhuri, R.; Fu, W.; Prabhakar, B.S.; Johnson, M.E.; et al. A Noncovalent Class of Papain-like Protease/Deubiquitinase Inhibitors Blocks SARS Virus Replication. Proc. Natl. Acad. Sci. USA 2008, 105, 16119–16124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Fu, Z.; Huang, B.; Tang, J.; Liu, S.; Liu, M.; Ye, Y.; Liu, Z.; Xiong, Y.; Zhu, W.; Cao, D.; et al. The Complex Structure of GRL0617 and SARS-CoV-2 PLpro Reveals a Hot Spot for Antiviral Drug Discovery. Nat. Commun. 2021, 12, 488. [Google Scholar] [CrossRef] [PubMed]
  146. Pitsillou, E.; Liang, J.; Ververis, K.; Lim, K.W.; Hung, A.; Karagiannis, T.C. Identification of Small Molecule Inhibitors of the Deubiquitinating Activity of the SARS-CoV-2 Papain-Like Protease: In Silico Molecular Docking Studies and in Vitro Enzymatic Activity Assay. Front. Chem. 2020, 8, 623971. [Google Scholar] [CrossRef]
Figure 1. Model describing the pathways targeted by SARS-CoV-2 to antagonize the innate immune response.
Figure 1. Model describing the pathways targeted by SARS-CoV-2 to antagonize the innate immune response.
Viruses 14 01247 g001
Table 1. Summary of mechanisms applied by SARS-CoV-2 to antagonize innate immune responses. The SARS-CoV-2 genome organization is represented, and proteins involved in host innate immunity pathway regulation are colored in red.
Table 1. Summary of mechanisms applied by SARS-CoV-2 to antagonize innate immune responses. The SARS-CoV-2 genome organization is represented, and proteins involved in host innate immunity pathway regulation are colored in red.
Mechanism of InhibitionViral ProteinsRefs.
Evasion of sensing by host innate immune receptors
Viruses 14 01247 i001
Formation of DMVsNSP3
NSP4
NSP6
[30,31]
CappingNSP10
NSP12
NSP13
NSP14
NSP16
[32,33,34]
Blockage of RIG-I RNA recognition
Interaction with its DExD/H helicase domain
N[35]
Cleavage of ISG15
Antagonism of ISG15-dependent MDA5 activation
NSP3[36,37]
Cleavage of RIG-I at the last 10 N-terminal amino acids
Blockage of its ability to signal through MAVS
Promotion of the ubiquitination and proteosome-mediated degradation of MAVS
NSP5[38]
Inhibition of RIG-I CARD domain activation
Interaction with Trim25
N[39,40,41]
Inhibition of innate immune receptor signaling and IFNs production
Viruses 14 01247 i002
Reduction of MAVS mediated IFN-β promoter activitiesNSP1
NSP3
NSP5
NSP12
NSP13
NSP14
M
N
ORF3a
ORF3b
ORF6
ORF7a
ORF7b
ORF8
ORF9b
[42,43,44,45,46]
Inhibition of IRF3 phosphorylation/nuclear translocationNSP1
NSP5
NSP6
NSP12
NSP13
NSP14
NSP15
ORF6
ORF3b
[42,43,44,49,50,51,62,63,64]
Cleavage of IRF3NSP3
NSP5
[62,64]
Inhibition of MAVS signaling complex
Interaction with TBK1
Interaction with TOM70
NSP13
M
ORF7a
ORF9b
[19,44,54,55,56,57,58,59]
Inhibition of NF-κB pathway
Interaction with Nup69 to block p65 translocation
Blockage of Nemo K63-linked polyubiquitination
Cleavage of TAB1
NSP5
NSP9
ORF9b
[60,61,62]
Inhibition of IFNs signaling and ISGs expression
Viruses 14 01247 i003
Inhibition of STAT1/STAT2 phosphorylationNSP1
NSP6
NSP13
M
N
ORF3a
ORF6
ORF7a
ORF7b
[44,69,70]
Blockage of STAT1/STAT2 nuclear translocation
Interaction with the nucleopore Nup98
ORF6[66,67,68]
IFNAR1 lysosomal degradationNSP14[65]
Inhibition of ISGsNSP3
NSP5
[36,43,76]
Inhibition of protein production by targeting post-transcription and translation steps
Viruses 14 01247 i004
Inhibition of pre-mRNA splicing
Binding to U1 and U2
NSP16[71]
Disruption of protein trafficking
Binding to SRP complex leading to the inhibition of signal peptide recognition
NSP8
NSP9
[71]
Blockage of mRNAs export
Binding to the mRNA entry channel overlapping mRNA path
Interaction with export receptor heterodimer NXF1-NXT
NSP1[71,72,73,74,75]
Table 2. Summary of discussed clinical trials with IFN therapies and their outcomes.
Table 2. Summary of discussed clinical trials with IFN therapies and their outcomes.
Viruses 14 01247 i005
Clinical Trial NameType of TrialType of PatientsOutcomesRefs.
IFN-α
Nebulized IFN-α2b with or
without Arbidol
Uncontrolled, exploratory study77 patients hospitalized with confirmed COVID-19 diagnosis (7 received IFN-α2b only, 46 IFN-α2b+ Arbidol)Time to negative RT-qPCR significantly shorter in patients receiving inhaled IFN-α2b.
Significant reduction in the duration of detectable virus in the upper respiratory tract.
Reduced blood levels of inflammatory markers (IL-6, CRP).
[94]
Nebulized IFN-α2b with or without UmifenovirRetrospective multicenter study446 patients with confirmed COVID-19 diagnosis (242 received IFN-α2b, 216 early and 26 late)Nebulized IFN-α2b initiation within 5 days of admission:
-
Associated with reduced mortality.
-
Not associated with hospital discharge or computed tomography (CT) scan improvement.
Late initiation of nebulized IFN-α2b therapy:
-
Associated with delayed recovery.
-
Associated with increased mortality.
[98]
Inhaled IFN-α2bRetrospective multicenter study1401 patients with confirmed COVID-19 diagnosis (852 received IFN-α2b)Early administration (3–5 days after symptom onset), associated with improved clinical outcomes:
-
Lower risk of disease progression.
-
Shorter hospitalization time.
Delayed IFN-α2b intervention (6–8 days after symptom onset): associated with increased probabilities of risk events.
[99]
PEG IFN-α2b (subcutaneous injection)
in addition to standard of care [antipyretics, cough suppressants, antibiotics, steroids, vitamins, anticoagulants, hydroxychloroquine and antivirals (e.g., Remdesivir)]
Multicenter, randomized, comparator-controlled, open-label phase 3 study250 patients with moderate COVID-19 (120 received PEG IFN-α2b + standard of care)Early viral clearance.
Clinical status improvement.
Decreased duration of supplemental oxygen.
[96]
IFN-α2b (subcutaneous injection) combined with Lopinavir/RitonavirExploratory study41 patients with confirmed COVID-19 diagnosis (19 received IFN-α2b + lopinavir/ritonavir)Early administration of IFN-α2b within 72 h following admission- resulted in shorter hospital stay, (10 days compared with late administration - after 72 h following admission).[97]
IFN-β
Nebulized IFN-β1a (inhalation) - SNG001
ClinicalTrials.gov Identifier: NCT04385095
Double-blind randomized, placebo-controlled, phase 2 pilot study101 patients with confirmed COVID-19 diagnosis (50 received IFN-β1a SNG001)Significantly greater odds of clinical improvement across the WHO Ordinal Scale for Clinical Improvement.
Reduction of the odds of developing severe disease or dying.
[113]
Nebulized IFN-β1b (inhalation through a vibrating mesh aerogen nebulizer-Aerogen Solo) combined with Favipiravir
ClinicalTrials.gov Identifier: NCT04385095
Randomized controlled open label study89 adult patients hospitalized with moderate to severe COVID-19 (44 received IFN-β1b)No significant differences in the inflammatory biomarkers at hospital discharge, in the overall lower hospital stay, transfers to the intensive care unit, neither in overall mortality.[114]
IFN-β1b (subcutaneous injection) combined with Lopinavir, Ritonavir, RibavirinMulticenter, prospective, open label, randomized, phase 2 study127 patients with mild to moderate COVID-19 (86 received IFN-β1b)Administration within 7 days of symptom onset:
-
Suppression of the shedding of SARS-CoV-2.
-
Significant reductions in duration of RT-qPCR positivity and viral load (RT-qPCR negative by day 8).
-
Shorter time in complete alleviation of symptoms
reduction in duration of hospital stay.
[107]
IFN-β1a or IFN-β1b (subcutaneous injections) combined with Lopinavir/Ritonavir
COVIFERON trial
ClinicalTrials.gov Identifier: NCT04343769
Randomized, open-label, controlled study60 severely ill patients with positive RT-qPCR and Chest CT scans (20 patients assigned to IFN-β1a and 20 to IFN-β1b)IFN-β1a: significant shorter time to clinical improvement.
IFN-β1b: no significant difference.
Lower numerically mortality both of the intervention groups (20% in the IFN-β1a group, 30% in the IFN-β1b group vs 45% in the control group) but not statically significant.
[108]
IFN-β1a (subcutaneous injection) in addition to the national protocol medications (Hydroxychloroquine plus Lopinavir- Ritonavir or Atazanavir-Ritonavir)
Clinical Identifier: IRCT20100228003449N28
Randomized, open-label, controlled study92 patients with severe COVID-19 (42 received IFN-β1a)No change the time to reach the clinical response.
Length of intensive care unit and hospital stays and duration of mechanical ventilation not statistically different.
Significantly increased discharge rate on day 14.
Early administration significantly reduced mortality.
[109]
IFN-β1a (subcutaneous injection) combined with Remdesivir
ClinicalTrials.gov Identifier: NCT04492475
Randomized, double-blind, placebo-controlled study969 patients hospitalized COVID-19 patients with presence of radiographic infiltrates on imaging, a peripheral oxygen saturation on room air of 94% or less, or requiring supplemental oxygen (487 received IFN-β1a)No clinical improvement.
Worse outcomes after treatment IFN-β1a in patients who required high-flow oxygen at baseline.
[111]
IFN-β1a (subcutaneous injection). For patients receiving high-flow oxygen, ventilation, or extra- corporeal membrane oxygenation: intravenous interferon.
ClinicalTrials.gov Identifier: NCT04315948
Randomized, double-blind, placebo-controlled study 4127 (2063 received IFN-β1a)No effect on hospitalized patients (based on overall mortality, initiation of ventilation, and duration of hospital stay).[110]
IFN-λ
PEG IFN-λ (subcutaneous injection)
ClinicalTrials.gov Identifier: NCT04354259
Randomized, double-blind, placebo-controlled study60 outpatients with COVID-19 (30 received PEG IFN-λ)Greater decline in RT-qPCR with viral clearance by day 7.
Prevent clinical deterioration and shorten duration of viral shedding.
[128]
PEG IFN-λ (subcutaneous injection)
ClinicalTrials.gov Identifier: NCT04331899
Randomized, double-blind, placebo-controlled phase 2 study120 outpatients with mild to moderate COVID-19 (30 received PEG IFN-λ)No shortened duration of SARS-CoV-2 viral shedding.
No improved of symptoms.
[129]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Znaidia, M.; Demeret, C.; van der Werf, S.; Komarova, A.V. Characterization of SARS-CoV-2 Evasion: Interferon Pathway and Therapeutic Options. Viruses 2022, 14, 1247. https://doi.org/10.3390/v14061247

AMA Style

Znaidia M, Demeret C, van der Werf S, Komarova AV. Characterization of SARS-CoV-2 Evasion: Interferon Pathway and Therapeutic Options. Viruses. 2022; 14(6):1247. https://doi.org/10.3390/v14061247

Chicago/Turabian Style

Znaidia, Mariem, Caroline Demeret, Sylvie van der Werf, and Anastassia V. Komarova. 2022. "Characterization of SARS-CoV-2 Evasion: Interferon Pathway and Therapeutic Options" Viruses 14, no. 6: 1247. https://doi.org/10.3390/v14061247

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