Next Article in Journal
Golgi Reassembly Stacking Protein 2 Modulates Myometrial Contractility during Labor by Affecting ATP Production
Next Article in Special Issue
Significance of the cGAS-STING Pathway in Health and Disease
Previous Article in Journal
Management of Neonatal Isolated and Combined Growth Hormone Deficiency: Current Status
Previous Article in Special Issue
The Role of Bacteria–Mitochondria Communication in the Activation of Neuronal Innate Immunity: Implications to Parkinson’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immunomodulatory Role of Interferons in Viral and Bacterial Infections

Department of Experimental Immunology, Medical University of Lublin, 20-093 Lublin, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(12), 10115; https://doi.org/10.3390/ijms241210115
Submission received: 10 May 2023 / Revised: 2 June 2023 / Accepted: 9 June 2023 / Published: 14 June 2023

Abstract

:
Interferons are a group of immunomodulatory substances produced by the human immune system in response to the presence of pathogens, especially during viral and bacterial infections. Their remarkably diverse mechanisms of action help the immune system fight infections by activating hundreds of genes involved in signal transduction pathways. In this review, we focus on discussing the interplay between the IFN system and seven medically important and challenging viruses (herpes simplex virus (HSV), influenza, hepatitis C virus (HCV), lymphocytic choriomeningitis virus (LCMV), human immunodeficiency virus (HIV), Epstein–Barr virus (EBV), and SARS-CoV coronavirus) to highlight the diversity of viral strategies. In addition, the available data also suggest that IFNs play an important role in the course of bacterial infections. Research is currently underway to identify and elucidate the exact role of specific genes and effector pathways in generating the antimicrobial response mediated by IFNs. Despite the numerous studies on the role of interferons in antimicrobial responses, many interdisciplinary studies are still needed to understand and optimize their use in personalized therapeutics.

1. Introduction

Interferons have been the subject of extensive research for several decades. In 1957, two scientists, Alick Isaacs and Jean Lindenmann, discovered interferons by observing a phenomenon called “interference” in viral replication. They found that virus-infected cells contain a soluble agent that can prevent the growth of other viruses, which was later identified as interferon. In the 1960s and 1970s, researchers discovered that interferons are proteins that have antiviral activity and can protect neighboring cells from viral replication. In the 1980s, researchers learned more about the structure and function of interferons thanks to protein isolation and purification methods. The US Food and Drug Administration (FDA) approved the first interferon-based therapy in the 1980s to treat hairy cell leukemia [1,2,3,4,5].
Further research in the 1990s and 2000s using molecular and genetic research led to a better understanding of interferons’ ability to regulate immune responses beyond antiviral activity. Interferons have been used to treat a variety of diseases, including chronic hepatitis B and C, certain cancers, and autoimmune disorders. Developments in recent years and the SARS-CoV-2 pandemic have contributed to the optimization of treatment protocols and deepening knowledge about the role of interferons in various diseases and immune responses [1,2,3,4,5].
These proteins activate the immune system’s defense mechanisms against various pathogens and are considered an essential component of the innate immune response [6,7]. They can activate immune system cells such as natural killer cells and macrophages, while also enhancing the speed of infection recognition by regulating antigen presentation to T lymphocytes. The human body can produce three different types of interferons, categorized based on the type of receptor they communicate with. Type I and III IFNs are induced in virtually all types of immune cells after recognizing viral components, while Type II IFN is restricted to T lymphocytes and NK cells and is induced by cytokines such as IL-12 [8,9,10,11,12,13,14,15]. Interferons not only strengthen healthy cells’ resistance to viral infections but also impair protein synthesis, effectively inhibiting virus multiplication. They also regulate and activate the immune response to bacterial infections [16,17,18]. This review presents an in-depth analysis of the characteristics and roles of the three types of IFN in viral and bacterial infections, including their mechanisms of action, molecular characteristics and signal transduction pathways.

2. Results and Discussion

2.1. Classification, Characteristics, and Mechanisms of Action of Interferons

Interferons, apart from interleukins and chemokines, belong to the family of cytokines, i.e., molecules regulating processes such as proliferation, differentiation, and cell motility. They are also mediators of the body’s defense reactions and participate in the regulation of hematopoiesis, which is why they are often referred to as hormones of the immune system with multifaceted antiviral and cell-modulating properties [19]. The research shows that interferons are mainly responsible for the activation of the immune system response, i.e., they have an immunomodulatory function, activity against a wide range of viruses, and regulation of cell growth and differentiation [20,21,22]. Currently there is a division of IFNs into three types based on affinity to a specific type of receptor, homology of amino acid sequences, and their functional properties.

2.1.1. Type I IFNs

Currently, 21 types of IFN are known, of which 13 subtypes belong to type I IFN. This was confirmed by the team of Wu et al. [23] who showed that there are 13 subtypes of human IFN-α, where IFN-α1 and IFN-α13 share the same amino acid sequences (Table 1). They include IFN-α (alpha), IFN-β (beta), IFN-ɛ (epsilon), IFN-κ (kappa), IFN-ω (omega), IFN-ν (nu), IFN-τ (tau), subtypes IFN-δ (delta), IFN-ζ (zeta) and IFN-χ (chi). These were originally shown to be powerful antiviral immunoregulators, but recent research shows that they also play an important role in bacterial infections [12,24,25].
All types of type I IFN are similar in structure and functioning and they are distinguished from other interferons mainly by the type of stimulated receptor and unique physicochemical properties: the ability to maintain structural stability at 65 °C and pH 2 [26]. In terms of genetic organization, the genes responsible for encoding type I IFNs are located on chromosome 9, region 2, band 1 (9p21). In the human genome, type I IFNs are encoded by a 350 kb multi-gene family: thirteen non-allelic IFN-α genes, at least five pseudogenes, and single genes for interferon β, ω, κ, and ε [19,27]. To date, IFN-α, which is encoded by several genes, and IFN-β have been the best characterized and most extensively described [28]. IFN-ω and IFN-τ are also known, but they remain poorly characterized today due to limited tissue expression, overlapping functions with IFN-α and IFN-β, and species differences [18]. Type I interferon receptors are mainly expressed by T lymphocytes, macrophages, and monocytes [12] where they are induced in infected cells to produce an antiviral state in uninfected cells [18]. A characterization of type I IFN performed by our team showed that most IFN-α contains 52.79-61.78% of hydrophilic amino acids, while IFN-β 60.43%, IFN-ω 56.41%, IFN-ε 57, 69% and IFN-κ 60.39% (Table 1).
Table 1. Molecular characteristics of three types of interferons.
Table 1. Molecular characteristics of three types of interferons.
Types of IFNsSubtypesGensProtein Length (AA)Molecular Mass (kDa)Protein IDPIAmino Acid CompositionSecondary StructureReference
% Hydrophilic% Hydrophobicα-Helixβ-Strand
Type I IFN-α1/13 IFNA1; IFNA13189 21,725 P01562 5.17 57.14% 42.86% 6 0 [29]
IFN-α2 IFNA2188 21,578 P01563 6.05 57.14% 42.86% 2 0 [30]
IFN-α4 IFNA4189 21,808 P05014 5.54 58.20% 41.80% 5 0 [31]
IFN-α5 IFNA518921,942P015695.48 57.14% 42.86% 50[32]
IFN-α6 ITGA61130126,606P232295.9460.88%39.12%561[33]
IFN-α7 CHRNA750256,449P365445.7352.79%47.21%1216[34]
IFN-α8 IFNA818921,989P328815.17 57.14% 42.86% 50[35]
IFN-α10 IFNA1018921,835P015665.7557.67%42.33%50[36]
IFN-α14IFNA1418922,063P015706.4956.61%43.39%50[37]
IFN-α16 IFI1678588,256Q166668.4261.78%38.22%1525[38]
IFN-α17 IFNA1718921,728P015715.5057.67%42.33%50[39]
IFN-α21IFNA2118921,741P015686.0757.14%42.86%50[40]
IFN-βIFNB118722,294P015748.0660.43%39.57%80[41]
IFN-ωIFNW119522,319P050008.3756.41%43.59%50[42]
IFN-ɛIFNE20824,414Q86WN27.9157.69%42.31%60[43]
IFN-κIFNK20725,218Q9P0W07.6660.39%39.61%70[44]
Type II IFN-γ IFNG166 19,348 P01579 8.74 63.85% 36.15% 7 0 [45,46]
Type III IFNλ1/IL-29 IFNL1200 21,898 Q8IU54 8.09 52.50% 47.50% 5 0 [47]
IFNλ2/IL-28A IFNL220022,288Q8IZJ07.1352.00%48.00%60[48]
IFNλ3/IL-28B IFNL319621,706Q8IZI97.5252.04%47.96%60[49]
IFN-λ4 IFNL417919,675K9M1U511.1750.84%49.16%40[50,51]
This group is also characterized by a high diversity of amino acid sequences, ranging from 5.36% to 95.77% similarity (Supplementary Materials Table S2). In addition, analysis of amino acid sequences showed that three subtypes of type I IFNs are characterized by the presence of several isoforms resulting from alternative splicing. These include IFN-α6 in eight additional isoforms; IFN-α7 present in three additional isoforms and IFN-α16 present in four additional isoforms (Table 2).

2.1.2. Type II IFN

The role of IFN type II (IFN-γ) as a key antiviral cytokine was originally discovered in a study of cell resistance to viral infections [52]. IFN-γ plays an important role in the innate and adaptive immune responses against bacteria, cancer, viruses, and pathological inflammatory processes [53]. IFN-γ is produced by various cells of the immune system, including natural killer cells, CD4+ helper T cells, and CD8+ cytotoxic T cells [12,54,55]. IFN-γ is encoded by the IFNG gene, located on chromosome 12 (12q15) [56]. Amino acid sequence analysis showed it consists of 63.85% hydrophilic amino acids (Table 1). IFN-γ is the only type II interferon and is serologically distinct from type I interferons; it is unstable in acid, while type I variants are stable in acid. In addition, analysis of the amino acid similarity of IFN-γ with other types of IFN showed it had relatively low similarity, ranging from 8.79% to 21.28% (Supplementary Materials Table S2). IFN-γ is both an important autocrine signal for APCs in the early innate immune response and an important paracrine signal in the adaptive immune response. IFN-γ expression is induced by the cytokines IL-12, IL-15, IL-18, and type I IFN [57]. In addition, IFN-γ induces the synthesis of other cytokines including interleukin 2, interleukin,6 and tumor necrosis factor TNF-α. Another function of IFN-γ is to stimulate the immune response by activating B and T lymphocytes. The antiviral activity of IFN-γ consists of blocking the synthesis of viral proteins and inhibiting the multiplication of virions. In addition, interferon-gamma enhances the expression of MHC class I and II molecules. This cytokine also has an antiproliferative effect, so it can limit the proliferation of cancer cells. The pro-inflammatory activity of gamma interferon consists of inducing macrophages and endothelial cells [58,59].

2.1.3. Type III IFN

Type III IFNs (IFN-λ) are the least characterized family of interferons involved in the innate immune response. They are structurally related to type I IFN and to the IL-10 family of interleukins [60]. They collectively induce an antiviral state in cells by inducing the expression of INF-stimulated genes (ISGs). Type III IFNs play an important role in the antiviral defense of the mucosal epithelium, which makes type III IFNs the primary factor in the first line of defense against viral infection in places of external contact with the pathogen [61]. According to the literature, the IFN type III group currently includes four subtypes: IFN-λ1 (IL-29), IFN-λ2 (IL-28A), IFN-λ3 (IL-28B), and IFN-λ4. They differ in their biochemical properties, amino acid composition (Table 1) and the degree of amino acid sequence identity with other types of IFN, with values ranging between 7.02% and 95.92% (Supplementary Materials Table S2). Detailed amino acid sequence analysis revealed that IFN-λ4 exists in four additional alternatively spliced isoforms (Table 2).
In type III IFNs, the production and expression of IFN-λ4 are controlled by the double nucleotide polymorphism IFNL4 rs368234815 (ΔG/TT), and the IFN-λ4 protein is produced by individuals carrying a functional genetic variant of the IFNL4-ΔG allele [62].

2.2. The Role of the Signal Transduction Mechanism in the Activity of Interferons

In the human body, there are several mechanisms involved in the innate immune response that protects against pathogens, particularly viruses. One such mechanism is the pattern recognition receptors (PRRs), which detect specific molecular patterns associated with pathogens. These PRRs play a crucial role in initiating the antiviral immune response.
To initiate the antiviral response, PRRs must first detect pathogen-associated molecular patterns (PAMPs) [6,63,64]. Toll-like receptors (TLRs), retinoic acid-induced gene-like receptors (RLRs), and Nod-like receptors (NLRs) are examples of PRRs. These receptors are located on the surface of cells and within cells, and when they detect PAMPs, they activate signaling pathways that result in the production and release of interferons. Researchers are paying particular attention to pattern recognition receptors (PRRs), which detect their presence and bind to pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) [6,63,64] The former group includes viral proteins, viral DNA, and viral RNA, while the latter group includes products resulting from virus-related cell damage and death [65]. Activation of PRR triggers the initiation of IFN production in the cell, which, upon binding to the corresponding receptor, initiates a signal transduction cascade resulting in ISG activation [66,67]. ISG activation induces an antiviral response in the cell. One of the best examples of PRR action is the signaling pathway mediated by Toll-like receptors 3 (TLR3), which, when activated by the presence of HSV, triggers the activation of a MyD88-independent signaling pathway (via the Toll/IL1 receptor domain containing the IFNβ inducing adapter (TRIF) and TRAF3), causing IRF3/7 to be translocated into the nucleus to stimulate type I IFN production. TLR3 activation also triggers the production of inflammatory cytokines and chemokines (via TRIF and TRAF6 activation of NF-κβ) [68].
Other studies have implicated RIG-I-like receptors in pathogenesis, which researchers believe are a different class of viral dsRNA sensors and include molecules such as MDA5 and RIG-I. Upon their activation, they bind to the mitochondrial antiviral signaling protein (MAVS), which activates the IRF3/7 and NF-κβ signaling pathways, resulting in increased production of type I and type III IFNs [69]. Another example of the use of the IRF3/7 signaling pathway for the production of type I IFN is the activity of cGAS, which is a DNA sensor that allows it to bind to viral dsDNA. According to researchers, such a combination induces the synthesis of 2’3’-cGAMP, which after binding to a protein associated with the endoplasmic reticulum, the stimulator of IFN genes (STING) signals through TANK 1 (TBK1) binding kinase, resulting in activation of IRF3/7 [70,71]. As a response to PRR activation, type I interferons, including IFN-alpha and IFN-beta, are produced and released by infected cells and bind to specific receptors on neighboring cells. This binding triggers a cascade of antiviral defense mechanisms that help prevent the spread of infection.
Furthermore, data show that all types of IFN have a separate heterodimeric receptor on the surface of each cell. The structural similarity between type I IFN subtypes is limited, but all share the same heterodimeric IFNA receptor (IFNAR) consisting of the high-affinity receptor IFNAR2 and the low-affinity receptor IFNAR1 [51]. IFNs type II bind to the IFNGR receptor complex (composed of IFNGR1 and IFNGR2), and type III IFNs bind to the IFNLR receptor complex (composed of IFNLR1 and IL-10Rβ), which in turn binds to the Janus kinase receptors JAK1 and JAK2. Activation of these kinases then leads to STAT1 homodimerization, nuclear translocation, and binding to gamma-activated sequences on DNA [72]. Type I and type III IFNs bind to different heterodimeric receptors, activating similar signaling pathways and transcriptional responses (Figure 1) [73]. The binding of type I, II and type III IFNs to their respective receptor complexes results in cross-phosphorylation of JAK1-activated kinase and tyrosine kinase 2 (TYK2) via the cytoplasmic domains of the receptor subunits [54]. However, it has recently been shown that JAK2 can also be phosphorylated by type III IFN but not by IFN type I [74,75]. Following the cross-phosphorylation process, the antiviral signal is transduced by a transcription activator (STAT). This then triggers the phosphorylation of IFN regulatory factor 9 (IRF9) to form IFN 3 (ISGF3) complexes or the phosphorylation of IFN-γ activation factor (GAF), which consists of STAT1 and STAT2 homodimers. These complexes are then translocated to the nucleus where they bind response elements stimulated by IFN regulatory factor (ISRE) or gamma ray sequences (GAS) on gene promoters stimulated by transcription activation of IFN (ISG) and (GAF), causing a response to viral infection [72,76]. Individual binding to these promoter elements results in the transcription of hundreds of genes involved in the antiviral response, including ISG, IFN, IRF, and STAT, OAS (oligoadenylate synthases); GBP (guanylate binding proteins); NOS2 (nitric oxide synthase 2); IFITM (IFN-induced transmembrane proteins); and TRIM (triremembered motif proteins) [2].
Although type I and type III IFNs belong to the innate immune response and show clear similarities in signaling and biological functions, both types of IFNs show differences that affect their biological functions. Type I IFN plays a pathogenic role in autoimmune diseases and is widely expressed in virtually every cell type, whereas IFNLR1 expression is restricted, mainly to epithelial cells lining mucosal surfaces.
Furthermore, while all cells are capable of expressing type I IFN, particularly when infected with a virus, only certain subsets of cells, again particularly epithelial cells, are capable of expressing type III IFN. Therefore, type III IFN is considered to be of paramount importance as a first-line and local innate defense against viral pathogens. The conducted studies do not fully clarify how IFNLR1 expression is limited to cells of epithelial origin, but there are indications that cell type-dependent epigenetic changes of the IFNLR1 gene, including DNA methylation and histone modifications, may be involved [75]. Indeed, DNA methyltransferase and histone deacetylase (HDAC) inhibitors have been reported to cause IFNLR1 gene silencing, increasing receptor expression and restoring sensitivity to IFN-λ in previously unresponsive cells, thereby increasing protection against viral pathogens, including herpes simplex 1 (HSV1) [78]. It now appears that activation of more than one signaling pathway is required to produce different biological properties of IFNs, and the signaling cascade alone is not sufficient to generate any given biological endpoint [74,79].
According to data in the literature, it is estimated that the effects of interferons are mediated by the induction of approximately 2000 ISG gene products. However, according to researchers, most of these ISGs do not directly initiate immunity to viral infection but rather influence the induction of signaling molecules that are responsible for stimulating immune responses to pathogens and disseminating signals received from infection sites [80,81,82,83]. Scientific studies indicate that there are four main effector pathways of the IFN-mediated antiviral response: the Mx antiviral protein pathway, the 2’,5’-oligoadenylate synthetase targeting ribonuclease L pathway, the R protein kinase pathway, and the ISG15-like pathway [2] (Table 3).
By influencing the activity of other cells, interferons contribute not only to the formation of antiviral factors but also to a specific type of antimicrobial readiness, involving many signaling pathways and activation of the body’s immune mechanisms. Activation of selected pathways causes modulation of various parts of the immune response, including non-specific mechanisms (activity of cytotoxic cells, production of complement proteins), the activity of B-line cells (responsible for the production of antibodies), and the expression of class I and II histocompatibility antigens. Interferons α and β are much more potent antivirals than IFN-γ; however, all types of interferons are necessary to activate the body’s full defenses against viruses (Figure 2).
Research is currently underway to identify and elucidate the precise role of individual genes and effector pathways in the generation of interferon-mediated antimicrobial responses. The effector pathways have been shown to block viral transcription, degrade viral RNA, inhibit translation, and modify protein function to control all steps of viral replication. Ongoing research continues to reveal additional activities of these effector proteins and unexpected antiviral response functions [2].

2.3. The Role of Interferons in the Course of Viral Infections

The effect of IFN on virus-infected cells and surrounding tissues results in increased expression and antiviral activity of IFN-stimulated genes in the host organism [98]. However, viruses have developed several anti-host mechanisms to promote efficient viral replication, thereby minimizing the antiviral power of IFN [99]. Such mechanisms include the activation of pathways that control IFN signaling, blocking the function of IFN-stimulated gene products, and interfering with various levels of intercommunication between IFN and other cellular pathways [27].
In this review, we will focus on discussing the interrelationships between the IFN system and seven medically important and challenging viruses (Lymphocytic Choriomeningitis Virus (LCMV), hepatitis C virus (HCV), Immunodeficiency Virus (HIV), Epstein–Barr virus (EBV), SARS-CoV coronavirus, influenza and herpes simplex virus (HSV)) to highlight the diversity of viral strategies. Understanding the complex cellular network of antiviral processes and virus–host interactions should help identify new targets for therapeutic intervention in viral infection [10].

2.3.1. Herpes Simplex Virus (HSV)

The herpes simplex virus (HSV) is one of the most widespread viruses in the world population, and humans are its only natural reservoir [98,99]. Epidemiological data indicate that its dissemination in the environment is possible only through direct contact with infected secretions [100,101]. The literature indicates that there are two types of herpes virus (1 and 2), both of which belong to the Herpetiviridae family. In addition to HSV viruses, this family also includes cytomegalovirus (CMV), herpes zoster (VZV), mononucleosis (EBV), human herpesvirus type 6 (HHV-6), human herpesvirus type 7 (HHV-7) and simian herpesvirus type B (HBLV). A characteristic feature of these viruses is their ability to go into latency and undergo reactivation. In clinical terms, HSV infection can take many forms, ranging from asymptomatic infections, through mucocutaneous infections and keratitis, to the most severe, highly lethal inflammations of the central nervous system [102]. Clinical reports indicate that in immunocompetent individuals, reactivation usually occurs within the anatomical innervation and in the vicinity of one dorsal ganglion, and if the primary infection occurs with the first to fourth decades of life, reactivation can occur at any time of the patient’s life [100]. Scientists are still looking for answers regarding the functioning of the mechanisms responsible for the entry latency and reactivation of HSV viruses. One of the demonstrated mechanisms involved in these processes, according to the research, may be granzyme B secreted by CD8+ cells, which degrades the virus ICP-4 protein necessary for its reactivation, keeping it in latency [103,104,105]. Many studies emphasize the important role of the innate immune system in the pathogenesis of HSV infections.
The development of this topic in the context of the role of interferons in response to HSV infection is supported by research conducted by You’s team in 2020. They demonstrated that β-catenin is an important component of the cGAS-STING signaling pathway, and blocking the nuclear translocation of β-catenin resulted in a decrease in the production of type I IFN [106].
In addition, the US3 protein kinase antagonizes IFN production by targeting β-catenin. According to the researchers, these findings shed new light on the mechanisms by which HSV-1 evades host antiviral immunity and will contribute to a better understanding of the interaction between host and HSV-1 infection [106]. Genetic studies show that viral modulation of IFN secretion, in particular by PKR, is the main basis of neurovirulence and pathogenesis associated with HSV infection [107]. In the literature, it is stated that PRR is also classified as ISG, although according to the researchers, they are always present in cells at a basic level of expression, and their expression is increased in response to IFN. This enhances the ability of host cells to recognize HSV PAMPs and thereby enhance IFN responses and the antiviral status mediated by TLR2, TLR3, TLR4, TLR7, cGAS, IFI16, MDA5, RIGI-I and PKR [108,109].
Further investigation of the role of interferon in the pathogenesis of HSV infections will focus on the study of keratinocytes, which researchers believe are the main source of IFN-β, and plasmacytoid dendritic cells (pDCs) infiltrating the dermis as a source of both IFN-α and IFN-β [110,111,112]. According to the literature, pDCs infiltrate the dermis after infection and it is suggested that they interact with other cells of the immune system to control viral replication [113]. Research conducted by Li’s team in 2020 showed that IFN-κ in vitro can be secreted by human keratinocytes [114]. This finding is extremely significant because the researchers showed that blocking the expression of IFN-κ contributes to an increase in HSV-1 replication. This allows us to hypothesize that IFN-κ may play an important role in controlling HSV-1 in herpes lesions [115,116].
Further analysis of the role of interferons in the pathogenesis of HSV showed that type I IFN activates natural killer cells that produce IFN-γ, which plays an important protective role against both HSV-1 and HSV-2 infections [115]. In addition to DCs, CD4+ and CD8+ T cells also contribute to IFN-γ production and also infiltrate the dermis [114]. Research conducted by Kim’s team in 2015 explains that this may be due to the migration of HSV-1-infected cells from the epidermis to the dermis, where they undergo apoptosis and then are taken up by skin DCs, presenting the antigen to CD8+ T cells [116]. This is also confirmed by research by the Harpur team in 2019, which showed that skin DCs and conventional DCs can work together to produce a suitable CD4+ T cell response [117]. Research by Zhu et al. in 2007 and 2009 showed that a subset of HSV-2-specific CD8+ T cells persist in the body long after the virus has been cleared from the lesion. According to these researchers, these lymphocytes may be involved in immune surveillance to control herpes infection [118,119]. Studies have indicated that macrophages play a role in producing IFN-α and IFN-β during HSV infection. As the infection progresses, these cells also contribute to the increase in IFN-γ production by natural killer cells through the secretion of IL-12 [120]. The combination of IFN-γ and IFN-α/β secreted at the site of infection may act synergistically both in vitro and in vivo, which contributes to the reduction in HSV-1 replication [121,122].
Other studies conducted by Peng et al. indicate that the basic mechanisms of interaction between IFN-β1 and IFN-γ function as an independent mechanism to induce separate categories of genes or as a cooperative mechanism (due to the induction of a common set of genes). Researchers showed that genes that were involved in the co-induction of both IFN subtypes led to apoptosis, triggering an inflammatory response and RNA degradation, which resulted in the inhibition of HSV-1 gene expression and increased DNA replication [123].

2.3.2. Influenza Virus

Influenza is a highly contagious disease, making it very important to isolate the sick person as soon as possible. Despite the use of one name, several types of influenza can be distinguished in clinical practice. Depending on the type of virus that causes the disease, influenza type A, B, C and D are mentioned. Depending on the type of virus, the course of influenza is slightly different, and the symptoms, contagiousness, risks, and complications of the disease may also be different. All influenza viruses belong to the family Orthomyxoviridae (ssRNA) [124,125]. According to the data presented by Thompson et al. [126], there are about five million clinical cases of influenza each year, and the highest percentage of deaths is reported in the age group above 65 years. The first line of defense and the way to limit the spread of the influenza virus in the human body are the mechanisms of the innate immune response, both physical barriers and the production of the cytokines IFN and ISG by cells [127,128].
Among the many ISGs activated in influenza infection, the most important are genes from the GTPase Mx family (contributing to the inhibition of nuclear import and replication of viral nucleocapsids) [127,128,129], viperin (which reduces the release of viral particles from infected cells) [130] and IFITM3 (a member of the IFITM family involved in interfering with the fusion between viral and endosomal membranes, thus limiting virus entry) [131,132]. As in the case of HSV, PRRs also play an important role here. There are two main types of PRR playing a role in influenza pathogenesis by inducing IFN [133]. The first is TLR7 expressed on pDC, which recognizes ssRNA molecules as well as incoming virions in endosomal compartments, resulting in the production of high levels of type I IFN [134,135]. Researchers showed that TLR7 is only essential in DC cells as its expression on other immune cell subsets is not required for IFN induction [136,137,138,139]. Therefore, the second important PRR is TLR3, which is responsible for dsRNA recognition due to influenza virus infection [140,141]. Although this receptor is also involved in increasing IFN production in response to other viral infections, in the case of influenza virus it acts by activating pro-inflammatory signaling pathways [140,141]. Moreover, research conducted by Kato et al. and Loo et al. showed that the induction of IFN in response to influenza virus infection is mainly mediated by RIG-I, although these researchers believe that MDA-5 may also be involved in this process as its deficiency in infected cells resulted in a slight decrease in IFN induction and ISG responses [142,143]. However, the influenza virus has developed multiple strategies to escape host immune surveillance to achieve successful replication without IFN induction [125]. One such mechanism is the avoidance of recognition by PRRs. This is possible due to the nuclear nature of replication and transcription of this virus (unlike other RNA viruses that cycle in the cytoplasm). This means that cRNAs and progeny vRNAs, which include RIG-I and MDA-5, are generated in a location inaccessible to PRRs [144]. Another possible mechanism is that influenza virus encodes antagonists of the IFN system [145]. The NS1 protein is the primary antagonist which targets various components of the IFN induction cascade [146,147]. It can reduce RLR activation by sequestering dsRNA, inhibiting RIG-I activation in TRIM25 [148,149,150], and by interfering with processing and nuclear export of cellular mRNAs [151,152].
Innate immune responses, including IFN, play a key role in determining the pathogenicity and outcome of influenza virus infection: effective activation of the IFN system early in the infection effectively restricts viral replication and eliminates the virus, while over-activation of innate immune responses increases host tissue damage [153].

2.3.3. Hepatitis C Virus (HCV)

HCV (Hepatitis C Virus) is a virus that causes hepatitis C. Modern medicine also knows other viruses that cause hepatitis (e.g., HBV, HAV, HEV). A typical feature of infections caused by hepatotropic viruses is damage to the liver parenchyma; however, HCV infection, unlike other infections, in most cases proceeds without characteristic symptoms and may manifest itself after many years in the form of cirrhosis or hepatocellular carcinoma [154]. Epidemiological data estimate that over 58 million people worldwide are infected with HCV, and there are about 1.5 million new infections each year [155,156]. Seven HCV genotypes (1-7) have numerous subtypes that differ in nucleotide sequence [157,158]. The HCV virus can evade the body’s defense mechanisms, including both innate and acquired immune mechanisms, resulting in a wide spectrum of pathogenicity, including the development of hepatocellular carcinoma [159].
The current gold standard of treatment for hepatitis C is weekly administration of pegylated IFN-α along with daily oral ribavirin (RBV) for 24 to 48 weeks. However, as research shows, this therapy is only effective in HCV 2 or 3 patients, causing sustained virological response (SVR) to develop in approximately 80–90% of treated patients. When infected with HCV genotype 1 or 4, only about half of patients can achieve SVR [160]. Clinicians have observed that there is a biphasic decline in HCV RNA levels in HCV infections treated with IFN therapy. This is observed after 8–9 h, which reflects the pharmacokinetics/pharmacodynamics of IFN, as well as the characteristics of the life cycle of HCV [161,162]. In addition to the viral genotype, the researchers also suggested that other factors such as gender or ethnicity are involved in the response to IFN therapy, which significantly indicates the involvement of specific types of genetic factors crucial for viral clearance [163,164]. Genetic studies, including genome-wide association studies, have shown a relationship between the occurrence of single nucleotide polymorphisms (SNPs) in the vicinity of the IFNL3 gene encoding IFN-λ3 and HCV infection and response to therapy. Research to date has focused on four SNPs: rs12979860 and rs8099917 for IFN-λ3, and rs368234815 and rs117648444 for IFN-λ4. Most of the research concerns the first two polymorphisms. The research team of Ge et al. showed that in the case of rs12979860 SNP, the favorable allele (CC genotype) was overrepresented in the European population in the group of people responding positively to interferon treatment, while the unfavorable TT genotype was more common in the African American population [165]. In addition, the researchers showed that the IFN-λ3 rs12979860 genotype was a better predictor of treatment outcome than assumed ethnicity because African Americans with the CC genotype had significantly more SVR than Europeans with the TT genotype. This theory was also confirmed by two other studies for the Australian and Japanese populations [166,167].
However, despite many studies, the exact mechanisms underlying the role of IFN-λ polymorphisms in HCV clearance are not well understood and require further intensive research. As in the case of other HCV viruses, it has also developed defense mechanisms against the immune response of the host organism. HCV can inhibit IFN production by interfering with the JAK-STAT pathway as well as blocking ISG mRNA translation and inhibiting the function of individual ISGs [168,169,170]. The most important mechanism that weakens the immune response is the blockade of translation induced by HCV by PKR phosphorylation and dimerization [171,172]. Phosphorylated PKR phosphorylates the eukaryotic translation initiation factor 2α-GDP (eIF2α-GDP) and prevents it from acting as a substrate for the eIF2B enzyme, which converts eIF2α-GDP to eIF2α-GTP, a key factor in protein translation [173]. ISG production and a translational block mediated by PKR are the mechanisms by which HCV significantly attenuates the interferon response [174]. Studies by Frese et al. [175] showed that MxA inhibited HCV RNA replication and the antiviral activity of IFN-α [175,176]. Another team (Itsui et al., [177]) found that overexpression of some ISGs, including MxA, significantly suppressed the HCV replicon, demonstrating the antiviral activity of MxA against HCV. Therefore, research data indicate that exogenous MxA protein may be involved in enhancing IFN-α and IFN-β activity in chronically HCV-infected patients by inhibiting HCV replication in the JAK-STAT signaling pathway [176].

2.3.4. Lymphocytic Choroid Meningitis Virus (LCMV)

Another virus discussed is the lymphocytic meningitis virus (LCMV), which is a member of the Arenaviridae family. This virus causes lymphocytic meningitis (LCM), an infectious disease transmitted by rodents [178]. According to epidemiological data, the main host of LCMV is the house mouse, of which about 5% in the United States are carriers. Moreover, research reports indicate that the degree of infection of mice with this virus is geographically dependent. In addition to mice, cases of LCMV infection from hamsters and rodents in homes or in pet stores are also increasingly common [179,180]. According to the CDC, the prevalence of LCMV antibodies in human populations ranges from 2% to 5% [178].
Two signaling pathways are involved in LCMV recognition: the first via the TLR and the second via the RIG-I (RLR). Activation of these two signaling pathways is dependent on their expression on immune cells, with TLRs found mainly on macrophages and DCs, while RLRs are widely expressed on most cell types [181]. In the case of LCMV, the virus enters the cell via the endosome, where it is detected by TLR7 and TLR8, which recognize ssRNAs, and membrane-associated TLR2 is also involved in the response to LCMV. In addition to the pro-inflammatory response, TLR2 has been shown to contribute to the production of IFN-I, with TLR2-deficient mice having significantly reduced serum IFN-I bioactivity after infection with LCMV-WE [182]. Researchers also indicate that TLRs expressed on the pDC endosome are also capable of signaling through an additional unique pathway allowing for the production of large amounts of type I IFN. This is possible due to signaling through MyD88 and the IRAK-1 and IKKα kinases which phosphorylate the constitutively expressed factor transcriptional IRF7 that stimulates the expression of type I IFN [183,184,185]. The RIG-I protein is responsible for detecting nucleic acids in the cytosol that are formed as part of the LCMV replication cycle. Upon binding of nucleic acids to RIG-I, ATPase/helicase activity is stimulated, which exposes the caspase recruitment domain (CARD) that in turn binds to the adapter protein, mitochondrial antiviral signaling protein (MAVS). As a result of these processes, a signaling complex containing TRAF3, TBK1, and IKKε is formed, which activates the IRF3 transcription factor responsible for inducing the expression of early types of IFN-I and IFN-β, as well as IFN-α4 [186,187,188]. Multiple studies in mice have shown the importance of type I IFN in controlling LCMV infection. Research data indicate that loss of IFN-I signaling by blocking or deleting the IFNAR receptor may result not only in chronic infection but also in increased susceptibility to slowly replicating strains of LCMV [185,189,190,191]. This virus will also develop specific mechanisms of resistance to host IFNs. One such example is the LCMV Arenavirus nucleoprotein (NP), which blocks the nuclear translocation and transcriptional activity of the regulatory factor IRF-3, resulting in a strong inhibition of type I IFN production. This allows LMCV proliferation to be controlled due to abnormalities in innate immune response pathways [188,189,190,191,192,193].
According to data published by Martinez-Sobrido L. et al., the persistence of LCMV does not prevent type I IFN signaling, but rather suppresses endogenous IFN-β production in response to infection. These data also show that NP is the only LCMV gene product responsible for IFN antagonist activity in LCMV-infected cells and indicate that LCMV-NP can inhibit IRF-3-dependent transcription [194]. Expanding our knowledge on the mechanisms underlying the anti-IFN activity of LCMV NP will contribute to a better understanding of the pathogenesis, and immunogenesis of arenavirus infection and would facilitate the production of recombinant LCMV virus.

2.3.5. Human Immunodeficiency Virus (HIV)

The most effective treatment regimen for the human immunodeficiency virus (HIV) is highly active antiretroviral therapy (HAART) to reduce HIV replication to undetectable levels. IFNs are used as powerful antiviral drugs that can activate the antiviral state in the body during HIV infection. Nevertheless, there is controversy as to whether IFN is an effective treatment for HIV due to its positive effect on viral infection and negative effect on chronic infection. Therefore, the exact role of type I IFN remains unclear and debated [195,196].
Studies by Cheng et al. showed that combination antiretroviral therapies (cART) inhibited HIV replication but failed to completely stop the elevated expression of ISGs, suggesting a prolonged induction of IFN type I [197]. cART stops the virus from making copies of itself in the body, which may result in slowing down the development of Acquired Immune Deficiency Syndrome (AIDS) [198,199]. Other research studies by Zhen A. et al. have shown that elevated type I IFN signaling during chronic HIV infection is a major cause of immune activation, inflammation, and depletion of cytotoxic T (CD8+) lymphocytes. This study confirms that IFN during HIV infection drives sustained immune activation and virus spread [200]. Although treatment with type I IFN is usually necessary to eradicate HIV, several mechanisms have evolved to bypass or suppress the effects of IFN in many situations. The research presented by Gargana et al. in 2018 showed that the mechanisms driven by many accessory proteins that the HIV uses to inhibit the therapeutic and endogenous effects of IFN-α result in a decrease in control over HIV infection [201]. Researchers also found that the HIV protein Vif plays a key role in the process of enzymatic post-translational protein modification (ubiquitination), self-catalysis, and proteosomal degradation of STAT1, STAT3 and monocyte cell lines in the canonical JAK-STAT pathway. All of the above processes allow HIV to inhibit the antiviral effects of type I IFN by inhibiting Vif-mediated STAT1 and STAT3, which consequently reduces the induction of the ISG-15 protein by IFN-α. Multiple mechanisms, which may be cell-specific, may explain the therapeutic failure of type I IFN in the course of HIV infection [201,202,203]. In addition, HIV can block the induction of type II and III IFNs in human dendritic cells and macrophages as a result of inhibition of the phosphorylation process of TANK1 binding kinase (TBK1) [204]. The HIV proteins Vif and Vpr have been shown to bind to TBK1 kinase, leading to inhibition of TBK1 trans-autophosphorylation, subsequent IRF3 phosphorylation, and induction of type I and III IFNs [205,206,207]. In addition, the HIV envelope protein (Gp120) has also been shown to play an important role in the inhibition of IFN-α, which is released and produced during viral infection of pDC. At the same time, interaction of Gp120 with TLR9 was observed in pDC [208]. Another way that HIV can evade IFN therapy is by significantly reducing the level of transcription of many antiviral ISGs, including, i.a., XAF1, OAS1, and AXL, thereby bypassing IFN treatment [209,210]. According to some studies, prolonged exposure to type I IFNs in HIV-infected patients in the chronic phase of infection has been associated with deleterious hyperimmune induction contributing to disease progression [211,212]. Although IFN therapy has been recognized as a promising approach to the treatment of HIV, the role played by type I IFN in the course and control of HIV infection is still unknown. Many questions remain unanswered despite intensive research in this field [196].

2.3.6. Epstein–Barr Virus (EBV)

The Epstein–Barr virus (EBV) is a DNA-type virus belonging to the herpesvirus family that causes a persistent infection that can last a lifetime in a latent state in infected body cells in most of the world’s population. EBV infection activates most cell signaling pathways, leading to the production of type I IFN, which inhibits EBV infection and virus-induced transformation of B cells [213,214,215]. The EBV virus is the first virus with oncogenic potential discovered and its main target cells are B lymphocytes. The EBV virus infects, in particular, epithelial cells of the pharynx and oral cavity through infection with saliva, where EBV implication occurs, which then leads to uncontrolled proliferation and killing of B lymphocytes [216] and natural killer cells, [217,218]. In most cases, EBV infection is asymptomatic, which can cause infectious mononucleosis and lead to serious diseases and carcinogenesis of many malignant tumors. It is associated with cancers of epithelial origins, such as nasopharyngeal cancer, and cancers originating from cells of the hematopoietic and lymphatic systems, such as Burkitt and Hodgkin lymphoma [219,220,221,222]. During the lytic replication and latency phase of the EBV virus, it triggers several viral gene expression models, thanks to which the virus bypasses the reaction of the human immune system. In addition, the virus triggers the induction of many cellular signaling pathways, including TLR pathways [223,224], regulatory factor IRF3 and IRF7, and cellular factor NF-κβ [225], resulting in the formation of type I IFN [224]. However, EBV lytic replication and EBV-depleted B cells show increased resistance to the antiproliferative effects of type I IFNs. This is due to the replication of latency proteins such as EBV-2 nuclear antigen (EBNA-2) and EBV membrane proteins (LMP1), which are considered to be oncoproteins [223,224,225].
Production of type I IFN by EBNA-2 in lymphoblastoid cells (LCL) results in priority expression of type I IFN and induction of STAT1. The LMP1 protein, probably through IRF7 and NF-κβ, activates the secretion of type I IFN, which activates the phosphorylation of STAT1 and IRF7, thanks to which it is possible to determine the course of EBV latency [226,227,228]. IFN-α secretion in EBV infection inhibits EBV early antigen activation when EBV replication is activated by superinfection, but does not inhibit antigen activation in cells undergoing lytic replication [217,226,227].
Data from the literature indicate that IFN-β, unlike IFN-α, does not inhibit EBV early antigen activation in cells unless replication is induced by butyrate [228,229,230], but inhibits early EBV excretion from the throat and mouth in immunocompromised individuals [231,232,233].
Other data indicate that IFN-β also inhibits EBV transformation [234,235]. The EBV envelope protein BGLF2 binds to the tyrosine kinase Tyk2, inhibiting the phosphorylation of STAT1 and STAT3 proteins necessary for immune control of EBV and for maintaining EBV latency [236]. It also blocks IFN signaling type I by counteracting the propensity of IFN-α and IFN-β to inhibit EBV reactivation.
IFN-α is released by infected cells that are triggered by viral DNA and are dependent on endosomal TLR9 signaling. pDCs express TLR7 and TLR9 and function as the primary sources of IFN-α release immediately after EBV infection [237,238,239].

2.3.7. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2)

The human severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) first appeared in December 2019 and has caused many cases of infection and death among the entire world population. The newly discovered SARS-CoV-2 shares about 80% of the nucleotide sequence of its predecessor, the severe respiratory syndrome coronavirus (SARS-CoV) from two decades ago. Interestingly, over 30% of SARS-CoV virus proteins have an inhibitory effect on the immune response mediated by type I IFN [240,241,242,243,244,245]. Studies by Lei et al. indicate that the SARS-CoV-2 virus induces the type I INF response and that therapies with the participation of IFN-β effectively inhibit the replication of the SARS-CoV-2 virus [246]. As with other viruses, the first line of defense against the SARS-CoV-2 virus is host innate immunity, of which the type I IFN system is a part and which is closely related to PAMPs via the PRR [247,248]. However, the exact signaling pathway resulting in the inhibition of IFN production is not yet fully understood. Previous studies indicate that SARS-CoV-2 infection stimulates delayed production of IFN, which suggests that it weakens the immune response as a result of the development of many strategies to antagonize the antiviral response of the human body [249,250,251].
The research team of Kumari et al. [252] showed that the SARS-CoV-2 virus is much more sensitive to type I IFN pretreatment than the SARS-CoV virus. Their study used isolated cell lines (Vero E6) to produce type I IFN that have been successfully used in exogenous therapy [252]. Moreover, infection of respiratory epithelial cells (Calu-3) against type I IFN resulted in reduced SARS-CoV-2 replication compared with SARS-CoV [253]. Studies have shown that there are significant genetic differences in the structure of both epidemic virus strains, which may affect their ability to modulate type I IFN responses. The final study showed that the results in Calu-3 cells are consistent with those of Vero E6 cells and indicate the sensitivity of the SARS-CoV-2 virus to the initial treatment with type I IFN. In addition, research on the SARS-CoV-2 virus shows that type I IFN therapy is promising; however, our knowledge about the novel SARS-CoV-2 coronavirus needs to be verified in terms of operation, infection, and disease caused by the virus [254,255,256,257,258]. Studies conducted by two independent teams [259,260] indicate that the SARS-CoV virus proteins, i.e., ORF3b, Nsp1, and ORF6, not only inhibit the expression of IFN-β in infected cells but additionally prevent ISRE activation and ISG expression in the presence of exogenous recombinant IFNs, indicating that these factors may block the IFNAR signaling pathway. Type III IFNs have been identified as potentially significant in the context of SARS-CoV-2 infection [261]. Clinical trials on IFN-λ for the treatment of patients with COVID-19 are being conducted with the aim of exerting antiviral effects without inflammatory effects. However, given the still relatively low level of knowledge in this area, IFN treatment should be carefully evaluated. Understanding the exact role of type I and III IFNs in SARS-CoV-2 infection should be explored to facilitate future treatment of COVID-19 patients.

2.4. The Role of Interferons in the Course of Bacterial Infections

Type I IFN plays an important role in the course of bacterial infections. Despite this, its role in the course of bacterial infections is not entirely clear and requires extensive research [18]. The signaling of type I IFNs to bacterial host infections is quite diverse, which may be influenced by many factors, e.g., site, route of infection, level of bacterial replication and expression, as well as virulence factors or immune response avoidance strategies. A differentiated effect depending on the route of infection has also been observed in many disease processes [262]. Recent studies indicate that type I IFNs may act as critical immune-promoting cytokines during infection with several species of bacteria. Moreover, the mechanism of action of type I IFNs during bacterial infections may have not only protective functions but also be detrimental to the body’s immune mechanisms (Figure 3). The reasons for such diverse roles of type I IFN in bacterial infections are still unclear and require many interdisciplinary studies.
Based on research, scientists have shown that bacteria trigger the production of type I IFN mainly by recognizing their nucleic acids or lipopolysaccharide (LPS), which is one of the components of the cell wall [263,264,265]. As a result of the recognition of these PAMPs, innate immune response mechanisms are activated. The activation pathways of IFN induction mechanisms have been best described for IFN-β, which, according to studies, is among the first to be produced during infection, along with IFN-α4, and is also the driving force of other type I IFN genes [18,266,267]. Researchers suggest that the induction of IFN-β by bacterial DNA is an extremely complex process and may involve many different pathways. One of the most common methods of inducing IFN-β is the cyclic GMP-AMP synthase with a cytosolic DNA sensor [268,269,270,271]. This mechanism has been described for bacterial infections caused by such pathogens such as Francisella novicida, group B streptococcus (GBS) (Streptococcus agalactiae), Legionella pneumophila, Listeria monocytogenes, Mycobacterium tuberculosis or Neisseria gonorrhoeae [272,273,274,275,276,277,278]. Moreover, the recognition of bacterial DNA (through unmethylated DNA containing the CpG motif) may also occur through the endosomal receptor TLR9, which as a consequence of activation of signal transduction pathways, contributes to the production of IFN-β [279,280,281]. Research in recent years has shown that bacterial RNA may also be a key inducer of IFN-β. This is due to a short and conserved sequence found in bacterial 23S rRNA that is recognized by mouse TLR13. Activation of this receptor contributes to the induction of the Myd88- and IRF5-dependent pathways, which consequently influence the production of IFN-β [282,283,284,285]. In the case of the human body, TLR8 is used as a bacterial RNA sensor, but the exact mechanism of its activation leading to the induction of IFN-β is still not fully understood [286,287,288].
The mechanism of IFN-β induction by LPS occurs after its binding to TLR4 and internalization into the endosome [289,290]. Subsequently, TRIF-dependent activation of TBK1 kinase causes phosphorylation of the transcription factor IRF3 to stimulate transcription of the IFN-β gene. Endosomal signaling is also involved in the induction of IFN-β by TLR2 in response to Gram-positive bacteria, although this mechanism appears to be limited to specific immune cells and/or pathogens [18]. Although the pathways that cause IFN-β induction by bacteria are relatively well understood, further research is needed to assess the importance of individual pathways for overall IFN-β production throughout the body rather than within cells.

3. Materials and Methods

3.1. Search Strategy, Study Selection, and Data Extraction

The analysis of the literature was carried out in the Web of Science, PubMed and Scopus databases, where the search for available articles was made based on the following keywords: “interferon”, “immune system”, “virus infection”, “bacterial infection”, “mechanism of action”, “immune response”. The time range was set to 2002–2022, using filters related to the type of articles (review, systematic review, clinical trials, meta-analysis). The articles produced by this search were assessed for access to the full version. Then, the selected articles were assessed for inclusion in the publication. Duplicates were rejected at each stage of the analysis among the found articles. In the end, 276 articles meeting the criteria were included. The procedure is presented below (Figure 4).

3.2. Bioinformatics Analyses of the Amino Acid Sequences of IFNs

For bioinformatics analyses, the amino acid sequences deposited in the UniProt database [291] were used. The IFNs identification numbers and their amino acid sequences are provided in Supplementary Materials Table S1. These sequences were used to carry out further bioinformatics analyses. The sequence length and molecular weight of individual interleukins were taken from the UniProt database. The determination of the isoelectric point of the tested interleukins and their amino acid compositions was carried out using the IPC isoelectric point calculator software available online [292]. The analysis of the second-order structure of the IFNs was carried out using the NetSurfP-3.0 program available online [293].

3.3. Analysis of the Amino Acid Sequence Identity of IFNs

The amino acid sequences of the selected interleukins from the UniProt databases (Supplementary Materials Table S1) were used for the analyses. The amino acid sequences of individual IFNs were compared with each other using the Clustal Omega program available in Reference [294]. The results of the analyses are presented as the percentage of identical amino acids in the analyzed amino acid sequences and are collected in Supplementary Materials Table S2.

4. Conclusions

The human body’s immune system can generate interferons, which are substances that aid in stimulating and regulating the immune response to viral infections and specific cancers. Interferons have been subject to extensive study and have been employed in medical treatments, including antiviral therapies, cancer treatment, and immunomodulation. Despite their benefits, interferon-based therapies can produce significant side effects, such as flu-like symptoms, fatigue, depression, and hematologic abnormalities, and may exhibit varying effectiveness based on the disease type, patient, and disease stage. Interferons function by activating genes that participate in signaling pathways to assist the immune system in fighting infections, such as by activating immune cells and regulating antigen presentation to T cells. Despite numerous studies, there is still much to learn about interferons, and researchers continue to make progress in understanding their mechanisms of action. Interdisciplinary studies are crucial in optimizing the use of interferons in therapeutic therapies, particularly in the context of combining them with other drugs to improve therapeutic outcomes and patient survival. This review is intended to inspire scientists to explore new research directions and to uncover unknown interactions between interferons and pathogens.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241210115/s1.

Author Contributions

Conceptualization, P.M., S.M., E.G. and K.S.; methodology, P.M. and S.M.; software, P.M.; validation, P.M., S.M. and E.G.; formal analysis, P.M. and S.M.; investigation, P.M. and S.M.; resources, P.M. and S.M.; data curation, E.G.; writing—original draft preparation, P.M.; K.S. and S.M.; writing—review and editing, P.M. and S.M.; visualization, P.M. and K.S.; supervision, E.G.; project administration, E.G.; funding acquisition, E.G. and K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Polish National Science Centre (NCN) grant no. 2021/41/B/NZ6/01765 and Medical University of Lublin grant no. PBsd 167.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Materials necessary for preparing this publication are included as Supplementary Materials. More information can be obtained from the corresponding author of this publication.

Conflicts of Interest

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

Abbreviations

CMVcytomegalovirus
DCsdendritic cells
dsRNAdouble-stranded RNA
EBVEpstein–Barr virus
GAFgamma-interferon activation factor
GASgamma ray sequences
GBPguanylate binding proteins
HBLVHuman B-lymphotropic virus
HCVHepatitis C Virus
HHV-6human herpesvirus type 6
HHV-7human herpesvirus type 7
HIVhuman immunodeficiency virus
HSVherpes simplex virus
IFITMinterferon-induced transmembrane proteins
IFN-ɛinterferon epsilon
IFNsinterferons
IFN-αinterferon alpha
IFN-βinterferon beta
IFN-γinterferon gamma
IFN-δinterferon delta
IFN-ζinterferon zeta
IFN-κinterferon kappa
IFN-νinterferon nu
IFN-τinterferon tau
IFN-χinterferon chi
IFN-ωinterferon omega
ISGsinterferon-stimulated genes
ISREinterferon regulatory factor
JAK2Janus kinase 2
LCMVlymphocytic choriomeningitis virus
LCMVLymphocytic Choriomeningitis Virus
MDA5melanoma-differentiation-associated gene 5
MHCMajor histocompatibility complex
NOS2nitric oxide synthase 2
OASoligoadenylate synthases
pDCsplasmacytoid dendritic cells
RIG-Iretinoic-acid-inducible protein I
SARS-CoV  coronavirus
STATSignal transducer and activator of transcription
TRIMtriremembered motif proteins
TYK2tyrosine kinase 2
VZVherpes zoster

References

  1. Borden, E.C.; Sen, G.C.; Uze, G.; Silverman, R.H.; Ransohoff, R.M.; Foster, G.R.; Stark, G.R. Interferons at age 50: Past, Current and Future Impact on Biomedicine. Nat. Rev. Drug Discov. 2007, 6, 975–990. [Google Scholar] [CrossRef] [PubMed]
  2. Sadler, A.J.; Williams, B.R.G. Interferon-inducible antiviral effectors. Nat. Rev. Immunol. 2008, 8, 559–568. [Google Scholar] [CrossRef] [PubMed]
  3. Fitzgerald-Bocarsly, P.; Dai, J.; Singh, S. Plasmacytoid dendritic cells and type I IFN: 50 years of convergent history. Cytokine Growth Factor Rev. 2008, 19, 3–19. [Google Scholar] [CrossRef] [Green Version]
  4. Billiau, A.; Matthys, P. Interferon-γ: A historical perspective. Cytokine Growth Factor Rev. 2009, 20, 97–113. [Google Scholar] [CrossRef] [PubMed]
  5. De Andrea, M.; Ravera, R.; Gioia, D.; Gariglio, M.; Landolfo, S. The Interferon System: An Overview. Eur. J. Paediatric Neurol. 2002, 6, A41–A46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Zhang, K.; Wang, H.; Hu, H. Overview of Interferon: Characteristics, signaling and anti-cancer effect. Arch. Biotechnol. Biomed. 2017, 1, 001–016. [Google Scholar] [CrossRef]
  7. Walker, F.C.; Sridhar, P.R.; Baldridge, M.T. Differential roles of interferons in innate responses to mucosal viral infections. Trends Immunol. 2021, 42, 1009–1023. [Google Scholar] [CrossRef]
  8. Zitvogel, L.; Galluzzi, L.; Kepp, O.; Smyth, M.J.; Kroemer, G. Type I interferons in anticancer immunity. Nat. Rev. Immunol. 2015, 15, 405–414. [Google Scholar] [CrossRef]
  9. Kopitar-Jerala, N. The Role of Interferons in Inflammation and Inflammasome Activation. Front. Immunol. 2017, 8, 873. [Google Scholar] [CrossRef] [Green Version]
  10. Katze, M.G.; He, Y.; Gale, M., Jr. Viruses and interferon: A fight for supremacy. Nat. Rev. Immunol. 2002, 2, 675–687. [Google Scholar] [CrossRef]
  11. Zhao, F.-R.; Wang, W.; Zheng, Q.; Zhang, Y.-G.; Chen, J. The regulation of antiviral activity of interferon epsilon. Front. Microbiol. 2022, 13, 6481. [Google Scholar] [CrossRef] [PubMed]
  12. Negishi, H.; Taniguchi, T.; Yanai, H. The Interferon (IFN) Class of Cytokines and the IFN Regulatory Factor (IRF) Transcription Factor Family. Cold Spring Harb. Perspect. Biol. 2017, 10, a028423. [Google Scholar] [CrossRef] [PubMed]
  13. Wack, A. Interfering with transmission. Elife 2018, 7, e37552. [Google Scholar] [CrossRef] [PubMed]
  14. Hervas-Stubbs, S.; Perez-Gracia, J.L.; Rouzaut, A.; Sanmamed, M.F.; Le Bon, A.; Melero, I. Direct Effects of Type I Interferons on Cells of the Immune System. Clin. Cancer Res. 2011, 17, 2619–2627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Bedsaul, J.R.; Zaritsky, L.A.; Zoon, K.C. Type I Interferon-Mediated Induction of Antiviral Genes and Proteins Fails to Protect Cells from the Cytopathic Effects of Sendai Virus Infection. J. Interf. Cytokine Res. 2016, 36, 652–665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Shibabaw, T.; Molla, M.D.; Teferi, B.; Ayelign, B. Role of IFN and Complements System: Innate Immunity in SARS-CoV-2. J. Inflamm. Res. 2020, 13, 507–518. [Google Scholar] [CrossRef]
  17. Boxx, G.M.; Cheng, G. The Roles of Type I Interferon in Bacterial Infection. Cell Host Microbe 2016, 19, 760–769. [Google Scholar] [CrossRef] [Green Version]
  18. Capobianchi, M.R.; Uleri, E.; Caglioti, C.; Dolei, A. Type I IFN family members: Similarity, differences and interaction. Cytokine Growth Factor Rev. 2014, 26, 103–111. [Google Scholar] [CrossRef]
  19. Cao, L.; Zhang, L.; Zhang, X.; Liu, J.; Jia, M.-A.; Zhang, J.; Liu, J.; Wang, F. Types of Interferons and Their Expression in Plant Systems. J. Interf. Cytokine Res. 2022, 42, 62–71. [Google Scholar] [CrossRef]
  20. Borden, E.C. Interferons α and β in cancer: Therapeutic opportunities from new insights. Nat. Rev. Drug Discov. 2019, 18, 219–234. [Google Scholar] [CrossRef]
  21. Fox, L.E.; Locke, M.C.; Lenschow, D.J. Context Is Key: Delineating the Unique Functions of IFNα and IFNβ in Disease. Front. Immunol. 2020, 11, 606874. [Google Scholar] [CrossRef]
  22. Wu, Y.; Ma, L.; Zhuang, Z.; Cai, S.; Zhao, Z.; Zhou, L.; Zhang, J.; Wang, P.-H.; Zhao, J.; Cui, J. Main protease of SARS-CoV-2 serves as a bifunctional molecule in restricting type I interferon antiviral signaling. Signal Transduct. Target. Ther. 2020, 5, 221. [Google Scholar] [CrossRef] [PubMed]
  23. de Jong, T.D.; Vosslamber, S.; Mantel, E.; de Ridder, S.; Wesseling, J.G.; van der Pouw Kraan, T.C.T.M.; Leurs, C.; Hegen, H.; Deisenhammer, F.; Killestein, J.; et al. Physiological evidence for diversification of IFNα- and IFNβ-mediated response programs in different autoimmune diseases. Arthritis Res. Ther. 2016, 18, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pestka, S.; Krause, C.D.; Walter, M.R. Interferons, interferon-like cytokines, and their receptors. Immunol. Rev. 2004, 202, 8–32. [Google Scholar] [CrossRef] [PubMed]
  25. Walter, M.R. The Role of Structure in the Biology of Interferon Signaling. Front. Immunol. 2020, 11, 606489. [Google Scholar] [CrossRef]
  26. Manry, J.; Laval, G.; Patin, E.; Fornarino, S.; Itan, Y.; Fumagalli, M.; Sironi, M.; Tichit, M.; Bouchier, C.; Casanova, J.-L.; et al. Evolutionary genetic dissection of human interferons. J. Exp. Med. 2011, 208, 2747–2759. [Google Scholar] [CrossRef] [Green Version]
  27. Decker, T.; Müller, M.; Stockinger, S. The Yin and Yang of type I interferon activity in bacterial infection. Nat. Rev. Immunol. 2005, 5, 675–687. [Google Scholar] [CrossRef]
  28. Aggarwal, A.; Stella, A.O.; Walker, G.; Akerman, A.; Esneau, C.; Milogiannakis, V.; Burnett, D.L.; McAllery, S.; Silva, M.R.; Lu, Y.; et al. Platform for isolation and characterization of SARS-CoV-2 variants enables rapid characterization of Omicron in Australia. Nat. Microbiol. 2022, 7, 896–908. [Google Scholar] [CrossRef]
  29. IFNA1—Interferon Alpha-1/13—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01562/entry (accessed on 16 February 2023).
  30. Interferon Alpha-2′—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q14645/entry (accessed on 16 February 2023).
  31. IFNA4—Interferon Alpha-4—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P05014/entry (accessed on 16 February 2023).
  32. IFNA5—Interferon Alpha-5—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01569/entry (accessed on 16 February 2023).
  33. ITGA6—Integrin Alpha-6—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P23229/entry (accessed on 16 February 2023).
  34. CHRNA7—Neuronal Acetylcholine Receptor Subunit Alpha-7—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P36544/entry (accessed on 16 February 2023).
  35. IFNA8—Interferon Alpha-8—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P32881/entry (accessed on 16 February 2023).
  36. IFNA10—Interferon Alpha-10—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01566/entry (accessed on 16 February 2023).
  37. IFNA14—Interferon Alpha-14—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01570/entry (accessed on 16 February 2023).
  38. IFI16—Gamma-Interferon-Inducible Protein 16—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q16666/entry (accessed on 16 February 2023).
  39. IFNA17—Interferon Alpha-17—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01571/entry (accessed on 16 February 2023).
  40. Virion Membrane Protein A21—Variola Virus|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P0DSR8/entry (accessed on 16 February 2023).
  41. IFNB1—Interferon Beta—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01574/entry (accessed on 16 February 2023).
  42. IFNW1—Interferon Omega-1—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P05000/entry (accessed on 16 February 2023).
  43. IFNE—Interferon Epsilon—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q86WN2/entry (accessed on 16 February 2023).
  44. IFNK—Interferon Kappa—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q9P0W0/entry (accessed on 16 February 2023).
  45. IFNG—Interferon Gamma—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/P01579/entry (accessed on 16 February 2023).
  46. Razaghi, A.; Owens, L.; Heimann, K. Review of the recombinant human interferon gamma as an immunotherapeutic: Impacts of production platforms and glycosylation. J. Biotechnol. 2016, 240, 48–60. [Google Scholar] [CrossRef]
  47. IFNL1—Interferon Lambda-1—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q8IU54/entry (accessed on 16 February 2023).
  48. IFNL2—Interferon Lambda-2—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q8IZJ0/entry (accessed on 16 February 2023).
  49. IFNL3—Interferon Lambda-3—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/Q8IZI9/entry (accessed on 16 February 2023).
  50. IFNL4—Interferon Lambda-4—Homo Sapiens (Human)|UniProtKB|UniProt. Available online: https://www.uniprot.org/uniprotkb/K9M1U5/entry (accessed on 16 February 2023).
  51. Takaoka, A.; Yanai, H. Interferon signalling network in innate defence. Cell. Microbiol. 2006, 8, 907–922. [Google Scholar] [CrossRef]
  52. Alspach, E.; Lussier, D.M.; Schreiber, R.D. Interferon γ and Its Important Roles in Promoting and Inhibiting Spontaneous and Therapeutic Cancer Immunity. Cold Spring Harb. Perspect. Biol. 2019, 11, a028480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. 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]
  54. 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]
  55. Lee, A.J.; Ashkar, A.A. The Dual Nature of Type I and Type II Interferons. Front. Immunol. 2018, 9, 2061. [Google Scholar] [CrossRef] [Green Version]
  56. IFNG Gene—GeneCards|IFNG Protein|IFNG Antibody. Available online: https://www.genecards.org/cgi-bin/carddisp.pl?gene=IFNG (accessed on 16 February 2023).
  57. Castro, F.; Cardoso, A.P.; Gonçalves, R.M.; Serre, K.; Oliveira, M.J. Interferon-Gamma at the Crossroads of Tumor Immune Surveillance or Evasion. Front. Immunol. 2018, 9, 847. [Google Scholar] [CrossRef] [Green Version]
  58. Kang, S.; Brown, H.M.; Hwang, S. Direct Antiviral Mechanisms of Interferon-Gamma. Immune Netw. 2018, 18, e33. [Google Scholar] [CrossRef]
  59. Ping, Z.; Qi, J.; Sun, Y.; Lu, G.; Shi, Y.; Wang, X.; Gao, G.F.; Wang, M. Crystal Structure of the Interferon Gamma Receptor Alpha Chain from Chicken Reveals an Undetected Extra Helix Compared with the Human Counterparts. J. Interf. Cytokine Res. 2014, 34, 41–51. [Google Scholar] [CrossRef] [Green Version]
  60. Gad, H.H.; Dellgren, C.; Hamming, O.J.; Vends, S.; Paludan, S.R.; Hartmann, R. Interferon-λ Is Functionally an Interferon but Structurally Related to the Interleukin-10 Family. J. Biol. Chem. 2009, 284, 20869–20875. [Google Scholar] [CrossRef] [Green Version]
  61. Kotenko, S.V.; Rivera, A.; Parker, D.; Durbin, J.E. Type III IFNs: Beyond antiviral protection. Semin. Immunol. 2019, 43, 101303. [Google Scholar] [CrossRef]
  62. Prokunina-Olsson, L.; Muchmore, B.; Tang, W.; Pfeiffer, R.M.; Park, H.; Dickensheets, H.; Hergott, D.; Porter-Gill, P.; Mumy, A.; Kohaar, I.; et al. A variant upstream of IFNL3 (IL28B) creating a new interferon gene IFNL4 is associated with impaired clearance of hepatitis C virus. Nat. Genet. 2013, 45, 164–171. [Google Scholar] [CrossRef]
  63. Ivashkiv, L.B.; Donlin, L.T. Regulation of type I interferon responses. Nat. Rev. Immunol. 2014, 14, 36–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Kienes, I.; Weidl, T.; Mirza, N.; Chamaillard, M.; Kufer, T.A. Role of NLRs in the Regulation of Type I Interferon Signaling, Host Defense and Tolerance to Inflammation. Int. J. Mol. Sci. 2021, 22, 1301. [Google Scholar] [CrossRef] [PubMed]
  65. Tang, D.; Kang, R.; Coyne, C.B.; Zeh, H.J.; Lotze, M.T. PAMPs and DAMPs: Signal 0s that spur autophagy and immunity. Immunol. Rev. 2012, 249, 158–175. [Google Scholar] [CrossRef]
  66. Zhao, J.; Qin, C.; Liu, Y.; Rao, Y.; Feng, P. Herpes Simplex Virus and Pattern Recognition Receptors: An Arms Race. Front. Immunol. 2021, 11, 3799. [Google Scholar] [CrossRef] [PubMed]
  67. Jounai, N.; Kobiyama, K.; Takeshita, F.; Ishii, K.J. Recognition of damage-associated molecular patterns related to nucleic acids during inflammation and vaccination. Front. Cell Infect. Microbiol. 2013, 2, 168. [Google Scholar] [CrossRef] [Green Version]
  68. Danastas, K.; Miranda-Saksena, M.; Cunningham, A.L. Herpes Simplex Virus Type 1 Interactions with the Interferon System. Int. J. Mol. Sci. 2020, 21, 5150. [Google Scholar] [CrossRef]
  69. Schenten, D.; Medzhitov, R. The Control of Adaptive Immune Responses by the Innate Immune System. Adv. Immunol. 2011, 109, 87–124. [Google Scholar] [CrossRef]
  70. Ablasser, A.; Goldeck, M.; Cavlar, T.; Deimling, T.; Witte, G.; Röhl, I.; Hopfner, K.-P.; Ludwig, J.; Hornung, V. cGAS produces a 2′-5′-linked cyclic dinucleotide second messenger that activates STING. Nature 2013, 498, 380–384. [Google Scholar] [CrossRef] [Green Version]
  71. Abe, T.; Barber, G.N. Cytosolic-DNA-Mediated, STING-Dependent Proinflammatory Gene Induction Necessitates Canonical NF-ΚB Activation through TBK1. J. Virol. 2014, 88, 5328–5341. [Google Scholar] [CrossRef] [Green Version]
  72. Platanias, L.C. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat. Rev. Immunol. 2005, 5, 375–386. [Google Scholar] [CrossRef]
  73. 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] [PubMed]
  74. Odendall, C.; Dixit, E.; Stavru, F.; Bierne, H.; Franz, K.M.; Durbin, A.; Boulant, S.; Gehrke, L.; Cossart, P.; Kagan, J.C. Diverse intracellular pathogens activate type III interferon expression from peroxisomes. Nat. Immunol. 2014, 15, 717–726. [Google Scholar] [CrossRef] [PubMed]
  75. Yin, Y.; Favoreel, H.W. Herpesviruses and the Type III Interferon System. Virol. Sin. 2021, 36, 577–587. [Google Scholar] [CrossRef] [PubMed]
  76. Teijaro, J.R. Type I interferons in viral control and immune regulation. Curr. Opin. Virol. 2016, 16, 31–40. [Google Scholar] [CrossRef] [PubMed]
  77. Stein, B.L.; Tiu, R.V. Biological Rationale and Clinical Use of Interferon in the Classical BCR-ABL-Negative Myeloproliferative Neoplasms. J. Interf. Cytokine Res. 2013, 33, 145–153. [Google Scholar] [CrossRef]
  78. Ding, S.; Khoury-Hanold, W.; Iwasaki, A.; Robek, M.D. Epigenetic Reprogramming of the Type III Interferon Response Potentiates Antiviral Activity and Suppresses Tumor Growth. PLoS Biol. 2014, 12, e1001758. [Google Scholar] [CrossRef] [Green Version]
  79. Wu, W.; Metcalf, J.P. The Role of Type I IFNs in Influenza: Antiviral Superheroes or Immunopathogenic Villains? J. Innate Immun. 2020, 12, 437–447. [Google Scholar] [CrossRef]
  80. 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]
  81. Rojas, M.; Luz-Crawford, P.; Soto-Rifo, R.; Reyes-Cerpa, S.; Toro-Ascuy, D. The Landscape of IFN/ISG Signaling in HIV-1-Infected Macrophages and Its Possible Role in the HIV-1 Latency. Cells 2021, 10, 2378. [Google Scholar] [CrossRef]
  82. Yang, E.; Li, M.M.H. All About the RNA: Interferon-Stimulated Genes That Interfere With Viral RNA Processes. Front. Immunol. 2020, 11, 5024. [Google Scholar] [CrossRef]
  83. Schoggins, J.W. Interferon-Stimulated Genes: What Do They All Do? Annu. Rev. Virol. 2019, 6, 567–584. [Google Scholar] [CrossRef] [PubMed]
  84. Haller, O.; Stertz, S.; Kochs, G. The Mx GTPase family of interferon-induced antiviral proteins. Microbes Infect. 2007, 9, 1636–1643. [Google Scholar] [CrossRef] [PubMed]
  85. Verhelst, J.; Hulpiau, P.; Saelens, X. Mx Proteins: Antiviral Gatekeepers That Restrain the Uninvited. Microbiol. Mol. Biol. Rev. 2013, 77, 551–566. [Google Scholar] [CrossRef] [Green Version]
  86. Braun, B.A.; Marcovitz, A.; Camp, J.G.; Jia, R.; Bejerano, G. Mx1 and Mx2 key antiviral proteins are surprisingly lost in toothed whales. Proc. Natl. Acad. Sci. USA 2015, 112, 8036–8040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lee, S.-H.; Vidal, S.M. Functional Diversity of Mx Proteins: Variations on a Theme of Host Resistance to Infection. Genome Res. 2002, 12, 527–530. [Google Scholar] [CrossRef] [Green Version]
  88. Silverman, R.H. Viral Encounters with 2′,5′-Oligoadenylate Synthetase and RNase L during the Interferon Antiviral Response. J. Virol. 2007, 81, 12720–12729. [Google Scholar] [CrossRef] [Green Version]
  89. Ogden, K.M.; Hu, L.; Jha, B.K.; Sankaran, B.; Weiss, S.R.; Silverman, R.H.; Patton, J.T.; Prasad, B.V.V. Structural Basis for 2′-5′-Oligoadenylate Binding and Enzyme Activity of a Viral RNase L Antagonist. J. Virol. 2015, 89, 6633–6645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Pindel, A.; Sadler, A. The Role of Protein Kinase R in the Interferon Response. J. Interf. Cytokine Res. 2011, 31, 59–70. [Google Scholar] [CrossRef] [PubMed]
  91. McAllister, C.S.; Taghavi, N.; Samuel, C.E. Protein Kinase PKR Amplification of Interferon β Induction Occurs through Initiation Factor eIF-2α-mediated Translational Control. J. Biol. Chem. 2012, 287, 36384–36392. [Google Scholar] [CrossRef] [Green Version]
  92. Gal-Ben-Ari, S.; Barrera, I.; Ehrlich, M.; Rosenblum, K. PKR: A Kinase to Remember. Front. Mol. Neurosci. 2019, 11, 480. [Google Scholar] [CrossRef] [Green Version]
  93. Zhao, C.; Collins, M.N.; Hsiang, T.-Y.; Krug, R.M. Interferon-induced ISG15 pathway: An ongoing virus–host battle. Trends Microbiol. 2013, 21, 181–186. [Google Scholar] [CrossRef] [PubMed]
  94. Chen, L.; Li, S.; McGilvray, I. The ISG15/USP18 ubiquitin-like pathway (ISGylation system) in Hepatitis C Virus infection and resistance to interferon therapy. Int. J. Biochem. Cell Biol. 2011, 43, 1427–1431. [Google Scholar] [CrossRef] [PubMed]
  95. Chang, Y.-G.; Yan, X.-Z.; Xie, Y.-Y.; Gao, X.-C.; Song, A.-X.; Zhang, D.-E.; Hu, H.-Y. Different Roles for Two Ubiquitin-like Domains of ISG15 in Protein Modification*. J. Biol. Chem. 2008, 283, 13370–13377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Lin, F.-C.; Young, H.A. Interferons: Success in anti-viral immunotherapy. Cytokine Growth Factor Rev. 2014, 25, 369–376. [Google Scholar] [CrossRef] [Green Version]
  97. Subramanian, G.; Kuzmanovic, T.; Zhang, Y.; Peter, C.B.; Veleeparambil, M.; Chakravarti, R.; Sen, G.C.; Chattopadhyay, S. A new mechanism of interferon’s antiviral action: Induction of autophagy, essential for paramyxovirus replication, is inhibited by the interferon stimulated gene, TDRD7. PLoS Pathog. 2018, 14, e1006877. [Google Scholar] [CrossRef] [Green Version]
  98. Fritsch, S.D.; Weichhart, T. Effects of Interferons and Viruses on Metabolism. Front. Immunol. 2016, 7, 630. [Google Scholar] [CrossRef] [Green Version]
  99. Fensterl, V.; Sen, G.C. Interferons and viral infections. Biofactors 2009, 35, 14–20. [Google Scholar] [CrossRef]
  100. Herpes Simplex Virus. Available online: https://www.who.int/news-room/fact-sheets/detail/herpes-simplex-virus (accessed on 16 February 2023).
  101. Herpes Simplex Virus (HSV) Infections. Available online: https://empendium.com/mcmtextbook/social/chapter/B31.II.18.1.8. (accessed on 16 February 2023).
  102. Gatherer, D.; Depledge, D.P.; Hartley, C.A.; Szpara, M.L.; Vaz, P.K.; Benkő, M.; Brandt, C.R.; Bryant, N.A.; Dastjerdi, A.; Doszpoly, A.; et al. ICTV Virus Taxonomy Profile: Herpesviridae 2021. J. Gen. Virol. 2021, 102, 001673. [Google Scholar] [CrossRef]
  103. Gerada, C.; Steain, M.; Campbell, T.M.; McSharry, B.; Slobedman, B.; Abendroth, A. Granzyme B Cleaves Multiple Herpes Simplex Virus 1 and Varicella-Zoster Virus (VZV) Gene Products, and VZV ORF4 Inhibits Natural Killer Cell Cytotoxicity. J. Virol. 2019, 93, e01140-19. [Google Scholar] [CrossRef]
  104. Nagda, M. The Role of Granzyme B in the Pathogenesis of Ocular HSV—1 Infection. Dissertation Thesis, University of nevada Honors College, Las Vegas, NV, USA, 2018. [Google Scholar]
  105. Cartier, A.; Broberg, E.; Komai, T.; Henriksson, M.; Masucci, M.G. The herpes simplex virus-1 Us3 protein kinase blocks CD8T cell lysis by preventing the cleavage of Bid by granzyme B. Cell Death Differ. 2003, 10, 1320–1328. [Google Scholar] [CrossRef]
  106. You, H.; Lin, Y.; Lin, F.; Yang, M.; Li, J.; Zhang, R.; Huang, Z.; Shen, Q.; Tang, R.; Zheng, C. β-Catenin Is Required for the cGAS/STING Signaling Pathway but Antagonized by the Herpes Simplex Virus 1 US3 Protein. J. Virol. 2020, 94, e01847-19. [Google Scholar] [CrossRef] [PubMed]
  107. Li, S.-F.; Gong, M.-J.; Zhao, F.-R.; Shao, J.-J.; Xie, Y.-L.; Zhang, Y.-G.; Chang, H.-Y. Type I Interferons: Distinct Biological Activities and Current Applications for Viral Infection. Cell Physiol. Biochem. 2018, 51, 2377–2396. [Google Scholar] [CrossRef] [PubMed]
  108. Ma, F.; Li, B.; Liu, S.-Y.; Iyer, S.S.; Yu, Y.; Wu, A.; Cheng, G. Positive Feedback Regulation of Type I IFN Production by the IFN-Inducible DNA Sensor cGAS. J. Immunol. 2015, 194, 1545–1554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Sirén, J.; Pirhonen, J.; Julkunen, I.; Matikainen, S. IFN-Alpha Regulates TLR-Dependent Gene Expression of IFN-Alpha, IFN-Beta, IL-28, and IL-29. J. Immunol. 2005, 174, 1932–1937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Gary, H.; Lebon, P.; Dalloul, A.H. Type I Interferon Production by Plasmacytoid Dendritic Cells and Monocytes Is Triggered by Viruses, but the Level of Production Is Controlled by Distinct Cytokines. J. Interf. Cytokine Res. 2002, 22, 653–659. [Google Scholar] [CrossRef]
  111. Swiecki, M.; Wang, Y.; Gilfillan, S.; Colonna, M. Plasmacytoid Dendritic Cells Contribute to Systemic but Not Local Antiviral Responses to HSV Infections. PLoS Pathog. 2013, 9, e1003728. [Google Scholar] [CrossRef] [Green Version]
  112. Schuster, P.; Thomann, S.; Werner, M.; Vollmer, J.; Schmidt, B. A Subset of Human Plasmacytoid Dendritic Cells Expresses CD8α upon Exposure to Herpes Simplex Virus Type 1. Front. Microbiol. 2015, 6, 557. [Google Scholar] [CrossRef] [Green Version]
  113. Donaghy, H.; Bosnjak, L.; Harman, A.N.; Marsden, V.; Tyring, S.K.; Meng, T.-C.; Cunningham, A.L. Role for Plasmacytoid Dendritic Cells in the Immune Control of Recurrent Human Herpes Simplex Virus Infection. J. Virol. 2009, 83, 1952–1961. [Google Scholar] [CrossRef] [Green Version]
  114. Li, Y.; Song, Y.; Zhu, L.; Wang, X.; Richers, B.; Leung, D.Y.M.; Bin, L. Interferon Kappa Is Important for Keratinocyte Host Defense against Herpes Simplex Virus-1. J. Immunol. Res. 2020, 2020, 5084682. [Google Scholar] [CrossRef] [Green Version]
  115. Posavad, C.M.; Zhao, L.; Mueller, D.E.; Stevens, C.E.; Huang, M.L.; Wald, A.; Corey, L. Persistence of Mucosal T-Cell Responses to Herpes Simplex Virus Type 2 in the Female Genital Tract. Mucosal Immunol. 2014, 8, 115–126. [Google Scholar] [CrossRef] [Green Version]
  116. Kim, M.; Truong, N.R.; James, V.; Bosnjak, L.; Sandgren, K.J.; Harman, A.N.; Nasr, N.; Bertram, K.M.; Olbourne, N.; Sawleshwarkar, S.; et al. Relay of Herpes Simplex Virus between Langerhans Cells and Dermal Dendritic Cells in Human Skin. PLoS Pathog. 2015, 11, e1004812. [Google Scholar] [CrossRef] [PubMed]
  117. Harpur, C.M.; Kato, Y.; Dewi, S.T.; Stankovic, S.; Johnson, D.N.; Bedoui, S.; Whitney, P.G.; Lahoud, M.H.; Caminschi, I.; Heath, W.R.; et al. Classical Type 1 Dendritic Cells Dominate Priming of Th1 Responses to Herpes Simplex Virus Type 1 Skin Infection. J. Immunol. 2019, 202, 653–663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zhu, J.; Koelle, D.; Cao, J.; Vazquez, J.; Huang, M.L.; Hladik, F.; Wald, A.; Corey, L. Virus-specific CD8+ T cells accumulate near sensory nerve endings in genital skin during subclinical HSV-2 reactivation. J. Exp. Med. 2007, 204, 595–603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Zhu, J.; Hladik, F.; Woodward, A.L.; Klock, A.; Peng, T.; Johnston, C.; Remington, M.; Magaret, A.; Koelle, D.; Wald, A.; et al. Persistence of HIV-1 receptor–positive cells after HSV-2 reactivation is a potential mechanism for increased HIV-1 acquisition. Nat. Med. 2009, 15, 886–892. [Google Scholar] [CrossRef] [Green Version]
  120. Ellermann-Eriksen, S. Macrophages and cytokines in the early defence against herpes simplex virus. Virol. J. 2005, 2, 59. [Google Scholar] [CrossRef] [Green Version]
  121. Sainz, B.; Halford, W.P. Alpha/Beta Interferon and Gamma Interferon Synergize To Inhibit the Replication of Herpes Simplex Virus Type 1. J. Virol. 2002, 76, 11541–11550. [Google Scholar] [CrossRef] [Green Version]
  122. Mikloska, Z.; Cunningham, A.L. Alpha and Gamma Interferons Inhibit Herpes Simplex Virus Type 1 Infection and Spread in Epidermal Cells after Axonal Transmission. J. Virol. 2001, 75, 11821–11826. [Google Scholar] [CrossRef] [Green Version]
  123. Peng, T.; Zhu, J.; Hwangbo, Y.; Corey, L.; Bumgarner, R.E. Independent and Cooperative Antiviral Actions of Beta Interferon and Gamma Interferon against Herpes Simplex Virus Replication in Primary Human Fibroblasts. J. Virol. 2008, 82, 1934–1945. [Google Scholar] [CrossRef] [Green Version]
  124. Nturibi, E.; Bhagwat, A.R.; Coburn, S.; Myerburg, M.M.; Lakdawala, S.S. Intracellular Colocalization of Influenza Viral RNA and Rab11A Is Dependent upon Microtubule Filaments. J. Virol. 2017, 91, e01179-17. [Google Scholar] [CrossRef] [Green Version]
  125. Chen, X.; Liu, S.; Goraya, M.U.; Maarouf, M.; Huang, S.; Chen, J.-L. Host Immune Response to Influenza A Virus Infection. Front. Immunol. 2018, 9, 320. [Google Scholar] [CrossRef] [Green Version]
  126. Thompson, W.W.; Shay, D.K.; Weintraub, E.; Brammer, L.; Cox, N.; Anderson, L.J.; Fukuda, K. Mortality Associated With Influenza and Respiratory Syncytial Virus in the United States. JAMA J. Am. Med. Assoc. 2003, 289, 179–186. [Google Scholar] [CrossRef] [PubMed]
  127. Gao, S.; von der Malsburg, A.; Dick, A.; Faelber, K.; Schröder, G.F.; Haller, O.; Kochs, G.; Daumke, O. Structure of Myxovirus Resistance Protein A Reveals Intra- and Intermolecular Domain Interactions Required for the Antiviral Function. Immunity 2011, 35, 514–525. [Google Scholar] [CrossRef] [Green Version]
  128. Turan, K.; Mibayashi, M.; Sugiyama, K.; Saito, S.; Numajiri, A.; Nagata, K. Nuclear MxA proteins form a complex with influenza virus NP and inhibit the transcription of the engineered influenza virus genome. Nucleic Acids Res. 2004, 32, 643–652. [Google Scholar] [CrossRef] [PubMed]
  129. Xiao, H.; Killip, M.J.; Staeheli, P.; Randall, R.; Jackson, D. The Human Interferon-Induced MxA Protein Inhibits Early Stages of Influenza A Virus Infection by Retaining the Incoming Viral Genome in the Cytoplasm. J. Virol. 2013, 87, 13053–13058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Wang, X.; Hinson, E.R.; Cresswell, P. The Interferon-Inducible Protein Viperin Inhibits Influenza Virus Release by Perturbing Lipid Rafts. Cell Host Microbe 2007, 2, 96–105. [Google Scholar] [CrossRef] [Green Version]
  131. Li, K.; Markosyan, R.M.; Zheng, Y.-M.; Golfetto, O.; Bungart, B.; Li, M.; Ding, S.; He, Y.; Liang, C.; Lee, J.C.; et al. IFITM Proteins Restrict Viral Membrane Hemifusion. PLoS Pathog. 2013, 9, e1003124. [Google Scholar] [CrossRef] [Green Version]
  132. Everitt, A.R.; Clare, S.; Pertel, T.; John, S.P.; Wash, R.S.; Smith, S.E.; Chin, C.R.; Feeley, E.M.; Sims, J.S.; Adams, D.J.; et al. IFITM3 restricts the morbidity and mortality associated with influenza. Nature 2012, 484, 519–523. [Google Scholar] [CrossRef] [Green Version]
  133. Lester, S.N.; Li, K. Toll-Like Receptors in Antiviral Innate Immunity. J. Mol. Biol. 2014, 426, 1246–1264. [Google Scholar] [CrossRef]
  134. Diebold, S.S.; Kaisho, T.; Hemmi, H.; Akira, S.; Reis E Sousa, C. Innate Antiviral Responses by Means of TLR7-Mediated Recognition of Single-Stranded RNA. Science 2004, 303, 1529–1531. [Google Scholar] [CrossRef]
  135. Lund, J.M.; Alexopoulou, L.; Sato, A.; Karow, M.; Adams, N.C.; Gale, N.W.; Iwasaki, A.; Flavell, R.A. Recognition of single-stranded RNA viruses by Toll-like receptor 7. Proc. Natl. Acad. Sci. USA 2004, 101, 5598–5603. [Google Scholar] [CrossRef] [Green Version]
  136. Kato, H.; Sato, S.; Yoneyama, M.; Yamamoto, M.; Uematsu, S.; Matsui, K.; Tsujimura, T.; Takeda, K.; Fujita, T.; Takeuchi, O.; et al. Cell Type-Specific Involvement of RIG-I in Antiviral Response. Immunity 2005, 23, 19–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Koyama, S.; Ishii, K.J.; Kumar, H.; Tanimoto, T.; Coban, C.; Uematsu, S.; Kawai, T.; Akira, S. Differential Role of TLR- and RLR-Signaling in the Immune Responses to Influenza A Virus Infection and Vaccination. J. Immunol. 2007, 179, 4711–4720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Yoneyama, M.; Kikuchi, M.; Natsukawa, T.; Shinobu, N.; Imaizumi, T.; Miyagishi, M.; Taira, K.; Akira, S.; Fujita, T. The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nat. Immunol. 2004, 5, 730–737. [Google Scholar] [CrossRef] [PubMed]
  139. Mayer, A.K.; Muehmer, M.; Mages, J.; Gueinzius, K.; Hess, C.; Heeg, K.; Bals, R.; Lang, R.; Dalpke, A.H. Differential Recognition of TLR-Dependent Microbial Ligands in Human Bronchial Epithelial Cells. J. Immunol. 2007, 178, 3134–3142. [Google Scholar] [CrossRef] [Green Version]
  140. Guillot, L.; Le Goffic, R.; Bloch, S.; Escriou, N.; Akira, S.; Chignard, M.; Si-Tahar, M. Involvement of Toll-like Receptor 3 in the Immune Response of Lung Epithelial Cells to Double-stranded RNA and Influenza A Virus. J. Biol. Chem. 2005, 280, 5571–5580. [Google Scholar] [CrossRef] [Green Version]
  141. Le Goffic, R.; Pothlichet, J.; Vitour, D.; Fujita, T.; Meurs, E.; Chignard, M.; Si-Tahar, M. Cutting Edge: Influenza A Virus Activates TLR3-Dependent Inflammatory and RIG-I-Dependent Antiviral Responses in Human Lung Epithelial Cells. J. Immunol. 2007, 178, 3368–3372. [Google Scholar] [CrossRef] [Green Version]
  142. Kato, H.; Takeuchi, O.; Sato, S.; Yoneyama, M.; Yamamoto, M.; Matsui, K.; Uematsu, S.; Jung, A.; Kawai, T.; Ishii, K.J.; et al. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature 2006, 441, 101–105. [Google Scholar] [CrossRef]
  143. Loo, Y.-M.; Fornek, J.; Crochet, N.; Bajwa, G.; Perwitasari, O.; Martinez-Sobrido, L.; Akira, S.; Gill, M.A.; García-Sastre, A.; Katze, M.G.; et al. Distinct RIG-I and MDA5 Signaling by RNA Viruses in Innate Immunity. J. Virol. 2008, 82, 335–345. [Google Scholar] [CrossRef] [Green Version]
  144. Ye, Q.; Krug, R.M.; Tao, Y.J. The mechanism by which influenza A virus nucleoprotein forms oligomers and binds RNA. Nature 2006, 444, 1078–1082. [Google Scholar] [CrossRef]
  145. García-Sastre, A. Induction and evasion of type I interferon responses by influenza viruses. Virus Res. 2011, 162, 12–18. [Google Scholar] [CrossRef] [Green Version]
  146. Engel, D.A. The influenza virus NS1 protein as a therapeutic target. Antivir. Res. 2013, 99, 409–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Hale, B.G.; Randall, R.E.; Ortín, J.; Jackson, D. The multifunctional NS1 protein of influenza A viruses. J. Gen. Virol. 2008, 89, 2359–2376. [Google Scholar] [CrossRef] [PubMed]
  148. Donelan, N.R.; Basler, C.F.; García-Sastre, A. A Recombinant Influenza A Virus Expressing anRNA-Binding-Defective NS1 Protein Induces High Levels of BetaInterferon and Is Attenuated inMice. J. Virol. 2003, 77, 13257–13266. [Google Scholar] [CrossRef] [Green Version]
  149. Steidle, S.; Martínez-Sobrido, L.; Mordstein, M.; Lienenklaus, S.; García-Sastre, A.; Stäheli, P.; Kochs, G. Glycine 184 in Nonstructural Protein NS1 Determines the Virulence of Influenza A Virus Strain PR8 without Affecting the Host Interferon Response. J. Virol. 2010, 84, 12761–12770. [Google Scholar] [CrossRef] [Green Version]
  150. Gack, M.U.; Albrecht, R.A.; Urano, T.; Inn, K.-S.; Huang, I.-C.; Carnero, E.; Farzan, M.; Inoue, S.; Jung, J.U.; García-Sastre, A. Influenza A Virus NS1 Targets the Ubiquitin Ligase TRIM25 to Evade Recognition by the Host Viral RNA Sensor RIG-I. Cell Host Microbe 2009, 5, 439–449. [Google Scholar] [CrossRef] [Green Version]
  151. Hayman, A.; Comely, S.; Lackenby, A.; Murphy, S.; McCauley, J.; Goodbourn, S.; Barclay, W. Variation in the Ability of Human Influenza A Viruses to Induce and Inhibit the IFN-Beta Pathway. Virology 2006, 347, 52–64. [Google Scholar] [CrossRef] [Green Version]
  152. Satterly, N.; Tsai, P.-L.; van Deursen, J.; Nussenzveig, D.R.; Wang, Y.; Faria, P.A.; Levay, A.; Levy, D.E.; Fontoura, B.M.A. Influenza virus targets the mRNA export machinery and the nuclear pore complex. Proc. Natl. Acad. Sci. USA 2007, 104, 1853–1858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Killip, M.J.; Fodor, E.; Randall, R.E. Influenza virus activation of the interferon system. Virus Res. 2015, 209, 11–22. [Google Scholar] [CrossRef] [Green Version]
  154. Blach, S.; A Terrault, N.; Tacke, F.; Gamkrelidze, I.; Craxi, A.; Tanaka, J.; Waked, I.; Dore, G.J.; Abbas, Z.; Abdallah, A.R.; et al. Global change in hepatitis C virus prevalence and cascade of care between 2015 and 2020: A modelling study. Lancet Gastroenterol. Hepatol. 2022, 7, 396–415. [Google Scholar] [CrossRef]
  155. Hepatitis, C. Available online: https://www.who.int/news-room/fact-sheets/detail/hepatitis-c (accessed on 16 February 2023).
  156. Alter, M.J. Epidemiology of hepatitis C virus infection. World J. Gastroenterol. 2007, 13, 2436–2441. [Google Scholar] [CrossRef] [Green Version]
  157. Simmonds, P. The Origin of Hepatitis C Virus. In Hepatitis C Virus: From Molecular Virology to Antiviral Therapy; Bartenschlager, R., Ed.; Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2013; pp. 1–15. ISBN 978-3-642-27340-7. [Google Scholar]
  158. Afzal, M.S.; Iqbal, M.A.; Ramamoorthy, V.; Campa, A.; Rubens, M.; Martinez, S.S.; Fleetwood, C.; Stewart, T.; Liuzzi, J.P.; George, F.; et al. Hepatitis C Virus and Interferon-Free Antiviral Therapeutics Revolution: Implications for Pakistan. Viral Immunol. 2017, 30, 252–257. [Google Scholar] [CrossRef] [PubMed]
  159. El-Shamy, A.; Hotta, H. Impact of hepatitis C virus heterogeneity on interferon sensitivity: An overview. World J. Gastroenterol. 2014, 20, 7555–7569. [Google Scholar] [CrossRef] [PubMed]
  160. Rong, L.; Perelson, A.S. Treatment of Hepatitis C Virus Infection With Interferon and Small Molecule Direct Antivirals: Viral Kinetics and Modeling. Crit. Rev. Immunol. 2010, 30, 131–148. [Google Scholar] [CrossRef] [Green Version]
  161. Mazana, J. Interferon Treatment of Chronic Hepatitis C. Basic Imunobiology and Imunopathology of the Liver. Rev. Española De Sanid. Penit. 2013, 15, 16–22. [Google Scholar] [CrossRef] [Green Version]
  162. Badr, G.; Bédard, N.; Abdel-Hakeem, M.S.; Trautmann, L.; Willems, B.; Villeneuve, J.-P.; Haddad, E.K.; Sékaly, R.P.; Bruneau, J.; Shoukry, N.H. Early Interferon Therapy for Hepatitis C Virus Infection Rescues Polyfunctional, Long-Lived CD8+ Memory T Cells. J. Virol. 2008, 82, 10017–10031. [Google Scholar] [CrossRef] [Green Version]
  163. Chen, S.L.; Morgan, T.R. The Natural History of Hepatitis C Virus (HCV) Infection. Int. J. Med. Sci. 2006, 3, 47–52. [Google Scholar] [CrossRef] [Green Version]
  164. Rauch, A.; Gaudieri, S.; Thio, C.; Bochud, P.-Y. Host genetic determinants of spontaneous hepatitis C clearance. Pharmacogenomics 2009, 10, 1819–1837. [Google Scholar] [CrossRef]
  165. Ge, D.; Fellay, J.; Thompson, A.J.; Simon, J.S.; Shianna, K.V.; Urban, T.J.; Heinzen, E.L.; Qiu, P.; Bertelsen, A.H.; Muir, A.J.; et al. Genetic variation in IL28B predicts hepatitis C treatment-induced viral clearance. Nature 2009, 461, 399–401. [Google Scholar] [CrossRef]
  166. Suppiah, V.; Moldovan, M.; Ahlenstiel, G.; Berg, T.; Weltman, M.; Abate, M.L.; Bassendine, M.; Spengler, U.; Dore, G.J.; Powell, E.; et al. IL28B is associated with response to chronic hepatitis C interferon-α and ribavirin therapy. Nat. Genet. 2009, 41, 1100–1104. [Google Scholar] [CrossRef]
  167. Tanaka, Y.; Nishida, N.; Sugiyama, M.; Kurosaki, M.; Matsuura, K.; Sakamoto, N.; Nakagawa, M.; Korenaga, M.; Hino, K.; Hige, S.; et al. Genome-wide association of IL28B with response to pegylated interferon-α and ribavirin therapy for chronic hepatitis C. Nat. Genet. 2009, 41, 1105–1109. [Google Scholar] [CrossRef]
  168. Heim, M.H. Innate immunity and HCV. J. Hepatol. 2013, 58, 564–574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Thimme, R.; Binder, M.; Bartenschlager, R. Failure of innate and adaptive immune responses in controlling hepatitis C virus infection. FEMS Microbiol. Rev. 2012, 36, 663–683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Raja, R.; Baral, S.; Dixit, N.M. Interferon at the cellular, individual, and population level in hepatitis C virus infection: Its role in the interferon-free treatment era. Immunol. Rev. 2018, 285, 55–71. [Google Scholar] [CrossRef] [Green Version]
  171. Wieland, S.; Makowska, Z.; Campana, B.; Calabrese, D.; Dill, M.T.; Chung, J.; Chisari, F.V.; Heim, M.H. Simultaneous detection of hepatitis C virus and interferon stimulated gene expression in infected human liver. Hepatology 2013, 59, 2121–2130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Perales, C.; Beach, N.M.; Gallego, I.; Soria, M.E.; Quer, J.; Esteban, J.I.; Rice, C.; Domingo, E.; Sheldon, J. Response of Hepatitis C Virus to Long-Term Passage in the Presence of Alpha Interferon: Multiple Mutations and a Common Phenotype. J. Virol. 2013, 87, 7593–7607. [Google Scholar] [CrossRef] [Green Version]
  173. Rowlands, A.G.; Panniers, R.; Henshaw, E.C. The catalytic mechanism of guanine nucleotide exchange factor action and competitive inhibition by phosphorylated eukaryotic initiation factor 2. J. Biol. Chem. 1988, 263, 5526–5533. [Google Scholar] [CrossRef]
  174. Kau, A.; Vermehren, J.; Sarrazin, C. Treatment predictors of a sustained virologic response in hepatitis B and C. J. Hepatol. 2008, 49, 634–651. [Google Scholar] [CrossRef] [Green Version]
  175. Frese, M.; Pietschmann, T.; Moradpour, D.; Haller, O.; Bartenschlager, R. Interferon-α inhibits hepatitis C virus subgenomic RNA replication by an MxA-independent pathway. J. Gen. Virol. 2001, 82, 723–733. [Google Scholar] [CrossRef] [Green Version]
  176. Shi, X.; Jiao, B.; Chen, Y.; Li, S.; Chen, L. MxA is a positive regulator of type I IFN signaling in HCV infection. J. Med. Virol. 2017, 89, 2173–2180. [Google Scholar] [CrossRef]
  177. Itsui, Y.; Sakamoto, N.; Kurosaki, M.; Kanazawa, N.; Tanabe, Y.; Koyama, T.; Takeda, Y.; Nakagawa, M.; Kakinuma, S.; Sekine, Y.; et al. Expressional screening of interferon-stimulated genes for antiviral activity against hepatitis C virus replication. J. Viral Hepat. 2006, 13, 690–700. [Google Scholar] [CrossRef]
  178. Lymphocytic Choriomeningitis (LCM)|CDC. Available online: https://www.cdc.gov/vhf/lcm/index.html (accessed on 16 February 2023).
  179. Knust, B.; Holman, R.C.; Redd, J.; Mehal, J.M.; Grube, S.M.; MacNeil, A.; Cheek, J.; Rollin, P. Lymphocytic Choriomeningitis Virus Infections among American Indians. Emerg. Infect. Dis. 2013, 19, 328–329. [Google Scholar] [CrossRef]
  180. Ledesma, J.; Fedele, C.G.; Carro, F.; Lledó, L.; Sánchez-Seco, M.P.; Tenorio, A.; Soriguer, R.C.; Saz, J.V.; Domínguez, G.; Rosas, M.F.; et al. Independent Lineage of Lymphocytic Choriomeningitis Virus in Wood Mice (Apodemus sylvaticus), Spain. Emerg. Infect. Dis. 2009, 15, 1677–1680. [Google Scholar] [CrossRef] [PubMed]
  181. Clingan, J.M.; Ostrow, K.; Hosiawa, K.A.; Chen, Z.J.; Matloubian, M. Differential Roles for RIG-I–like Receptors and Nucleic Acid-Sensing TLR Pathways in Controlling a Chronic Viral Infection. J. Immunol. 2012, 188, 4432–4440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Zhou, S.; Kurt-Jones, E.A.; Mandell, L.; Cerny, A.; Chan, M.; Golenbock, D.T.; Finberg, R.W. MyD88 is critical for the development of innate and adaptive immunity during acute lymphocytic choriomeningitis virus infection. Eur. J. Immunol. 2005, 35, 822–830. [Google Scholar] [CrossRef] [PubMed]
  183. Blasius, A.L.; Beutler, B. Intracellular Toll-like Receptors. Immunity 2010, 32, 305–315. [Google Scholar] [CrossRef] [Green Version]
  184. Kawai, T.; Sato, S.; Ishii, K.J.; Coban, C.; Hemmi, H.; Yamamoto, M.; Terai, K.; Matsuda, M.; Inoue, J.-I.; Uematsu, S.; et al. Interferon-Alpha Induction through Toll-like Receptors Involves a Direct Interaction of IRF7 with MyD88 and TRAF6. Nat. Immunol. 2004, 5, 1061–1068. [Google Scholar] [CrossRef]
  185. Gilliet, M.; Cao, W.; Liu, Y.-J. Plasmacytoid dendritic cells: Sensing nucleic acids in viral infection and autoimmune diseases. Nat. Rev. Immunol. 2008, 8, 594–606. [Google Scholar] [CrossRef]
  186. Zhou, S.; Cerny, A.M.; Zacharia, A.; Fitzgerald, K.A.; Kurt-Jones, E.A.; Finberg, R.W. Induction and Inhibition of Type I Interferon Responses by Distinct Components of Lymphocytic Choriomeningitis Virus. J. Virol. 2010, 84, 9452–9462. [Google Scholar] [CrossRef] [Green Version]
  187. Marié, I.; Durbin, J.E.; Levy, D.E. Differential viral induction of distinct interferon-alpha genes by positive feedback through interferon regulatory factor-7. EMBO J. 1998, 17, 6660–6669. [Google Scholar] [CrossRef] [Green Version]
  188. Suprunenko, T.; Hofer, M.J. Complexities of Type I Interferon Biology: Lessons from LCMV. Viruses 2019, 11, 172. [Google Scholar] [CrossRef] [Green Version]
  189. Ni Lee, L.; Burke, S.; Montoya, M.; Borrow, P. Multiple Mechanisms Contribute to Impairment of Type 1 Interferon Production during Chronic Lymphocytic Choriomeningitis Virus Infection of Mice1. J. Immunol. 2009, 182, 7178–7189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Daugan, M.; Murira, A.; Mindt, B.C.; Germain, A.; Tarrab, E.; Lapierre, P.; Fritz, J.H.; Lamarre, A. Type I Interferon Impairs Specific Antibody Responses Early during Establishment of LCMV Infection. Front. Immunol. 2016, 7, 564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Sullivan, B.; Teijaro, J.R.; De La Torre, J.C.; Oldstone, M.B.A. Early Virus-Host Interactions Dictate the Course of a Persistent Infection. PLoS Pathog. 2015, 11, e1004588. [Google Scholar] [CrossRef] [PubMed]
  192. Ortiz-Riaño, E.; Cheng, B.Y.; de la Torre, J.C.; Martínez-Sobrido, L. D471G Mutation in LCMV-NP Affects Its Ability to Self-associate and Results in a Dominant Negative Effect in Viral RNA Synthesis. Viruses 2012, 4, 2137–2161. [Google Scholar] [CrossRef] [Green Version]
  193. Martínez-Sobrido, L.; Emonet, S.; Giannakas, P.; Cubitt, B.; García-Sastre, A.; de la Torre, J.C. Identification of Amino Acid Residues Critical for the Anti-Interferon Activity of the Nucleoprotein of the Prototypic Arenavirus Lymphocytic Choriomeningitis Virus. J. Virol. 2009, 83, 11330–11340. [Google Scholar] [CrossRef] [Green Version]
  194. Doyle, T.; Goujon, C.; Malim, M.H. HIV-1 and interferons: Who’s interfering with whom? Nat. Rev. Genet. 2015, 13, 403–413. [Google Scholar] [CrossRef]
  195. Bosinger, S.E.; Utay, N.S. Type I Interferon: Understanding Its Role in HIV Pathogenesis and Therapy. Curr. HIV/AIDS Rep. 2015, 12, 41–53. [Google Scholar] [CrossRef]
  196. Cheng, L.; Ma, J.; Li, J.; Li, D.; Li, G.; Li, F.; Zhang, Q.; Yu, H.; Yasui, F.; Ye, C.; et al. Blocking type I interferon signaling enhances T cell recovery and reduces HIV-1 reservoirs. J. Clin. Investig. 2016, 127, 269–279. [Google Scholar] [CrossRef] [Green Version]
  197. Collaboration, T.H.-C. The effect of combined antiretroviral therapy on the overall mortality of HIV-infected individuals. Aids 2010, 24, 123–137. [Google Scholar] [CrossRef] [Green Version]
  198. Ortega, M.; Brier, M.R.; Ances, B.M. Effects of HIV and combination antiretroviral therapy on cortico-striatal functional connectivity. Aids 2015, 29, 703–712. [Google Scholar] [CrossRef] [Green Version]
  199. Zhen, A.; Rezek, V.; Youn, C.; Lam, B.; Chang, N.; Rick, J.; Carrillo, M.; Martin, H.; Kasparian, S.; Syed, P.; et al. Targeting type I interferon–mediated activation restores immune function in chronic HIV infection. J. Clin. Investig. 2016, 127, 260–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Gargan, S.; Ahmed, S.; Mahony, R.; Bannan, C.; Napoletano, S.; O’Farrelly, C.; Borrow, P.; Bergin, C.; Stevenson, N.J. HIV-1 Promotes the Degradation of Components of the Type 1 IFN JAK/STAT Pathway and Blocks Anti-viral ISG Induction. Ebiomedicine 2018, 30, 203–216. [Google Scholar] [CrossRef] [Green Version]
  201. Nguyen, N.V.; Tran, J.T.; Sanchez, D.J. HIV blocks Type I IFN signaling through disruption of STAT1 phosphorylation. J. Endotoxin Res. 2018, 24, 490–500. [Google Scholar] [CrossRef] [Green Version]
  202. Ezeonwumelu, I.J.; Garcia-Vidal, E.; Ballana, E. JAK-STAT Pathway: A Novel Target to Tackle Viral Infections. Viruses 2021, 13, 2379. [Google Scholar] [CrossRef] [PubMed]
  203. Chistiakov, D.A.; Grechko, A.V.; Orekhov, A.N.; Bobryshev, Y.V. An immunoregulatory role of dendritic cell-derived exosomes versus HIV-1 infection: Take it easy but be warned. Ann. Transl. Med. 2017, 5, 362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Harman, A.N.; Nasr, N.; Feetham, A.; Galoyan, A.; Alshehri, A.A.; Rambukwelle, D.; Botting, R.A.; Hiener, B.M.; Diefenbach, E.; Diefenbach, R.J.; et al. HIV Blocks Interferon Induction in Human Dendritic Cells and Macrophages by Dysregulation of TBK1. J. Virol. 2015, 89, 6575–6584. [Google Scholar] [CrossRef] [Green Version]
  205. Okumura, A.; Alce, T.; Lubyova, B.; Ezelle, H.; Strebel, K.; Pitha, P.M. HIV-1 accessory proteins VPR and Vif modulate antiviral response by targeting IRF-3 for degradation. Virology 2008, 373, 85–97. [Google Scholar] [CrossRef] [Green Version]
  206. Khan, H.; Sumner, R.P.; Rasaiyaah, J.; Tan, C.P.; Rodriguez-Plata, M.T.; Van Tulleken, C.; Fink, D.; Zuliani-Alvarez, L.; Thorne, L.; Stirling, D.; et al. HIV-1 Vpr antagonizes innate immune activation by targeting karyopherin-mediated NF-κB/IRF3 nuclear transport. Elife 2020, 9, e60821. [Google Scholar] [CrossRef]
  207. Martinelli, E.; Cicala, C.; Van Ryk, D.; Goode, D.J.; Macleod, K.; Arthos, J.; Fauci, A.S. HIV-1 gp120 inhibits TLR9-mediated activation and IFN-α secretion in plasmacytoid dendritic cells. Proc. Natl. Acad. Sci. USA 2007, 104, 3396–3401. [Google Scholar] [CrossRef] [Green Version]
  208. Wie, S.-H.; Du, P.; Luong, T.Q.; Rought, S.E.; Beliakova-Bethell, N.; Lozach, J.; Corbeil, J.; Kornbluth, R.S.; Richman, D.D.; Woelk, C.H. HIV Downregulates Interferon-Stimulated Genes in Primary Macrophages. J. Interf. Cytokine Res. 2013, 33, 90–95. [Google Scholar] [CrossRef] [Green Version]
  209. He, B.; Tran, J.T.; Sanchez, D.J. Manipulation of Type I Interferon Signaling by HIV and AIDS-Associated Viruses. J. Immunol. Res. 2019, 2019, 8685312. [Google Scholar] [CrossRef]
  210. McNab, F.; Mayer-Barber, K.; Sher, A.; Wack, A.; O’Garra, A. Type I interferons in infectious disease. Nat. Rev. Immunol. 2015, 15, 87–103. [Google Scholar] [CrossRef]
  211. Scagnolari, C.; Antonelli, G. Type I interferon and HIV: Subtle balance between antiviral activity, immunopathogenesis and the microbiome. Cytokine Growth Factor Rev. 2018, 40, 19–31. [Google Scholar] [CrossRef] [PubMed]
  212. Bouvet, M.; Voigt, S.; Tagawa, T.; Albanese, M.; Chen, Y.-F.A.; Chen, Y.; Fachko, D.N.; Pich, D.; Göbel, C.; Skalsky, R.L.; et al. Multiple Viral microRNAs Regulate Interferon Release and Signaling Early during Infection with Epstein-Barr Virus. mBio 2021, 12, e03440-20. [Google Scholar] [CrossRef] [PubMed]
  213. Raab-Traub, N. EBV-Induced Oncogenesis. In Human Herpesviruses: Biology, Therapy, and Immunoprophylaxis; Arvin, A., Campadelli-Fiume, G., Mocarski, E., Moore, P.S., Roizman, B., Whitley, R., Yamanishi, K., Eds.; Cambridge University Press: Cambridge, UK, 2007; ISBN 978-0-521-82714-0. [Google Scholar]
  214. Saha, A.; Robertson, E.S. Mechanisms of B-Cell Oncogenesis Induced by Epstein-Barr Virus. J. Virol. 2019, 93, e00238-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Mui, U.N.; Haley, C.T.; Tyring, S.K. Viral Oncology: Molecular Biology and Pathogenesis. J. Clin. Med. 2017, 6, 111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Cai, Q.; Chen, K.; Young, K.H. Epstein–Barr virus-positive T/NK-cell lymphoproliferative disorders. Exp. Mol. Med. 2015, 47, e133. [Google Scholar] [CrossRef] [Green Version]
  217. Liu, X.; Sadaoka, T.; Krogmann, T.; Cohen, J.I. Epstein-Barr Virus (EBV) Tegument Protein BGLF2 Suppresses Type I Interferon Signaling To Promote EBV Reactivation. J. Virol. 2020, 94, e00258-20. [Google Scholar] [CrossRef]
  218. Mui, U.N.; Haley, C.T.; Vangipuram, R.; Tyring, S.K. Human oncoviruses: Mucocutaneous manifestations, pathogenesis, therapeutics, and prevention: Hepatitis viruses, human T-cell leukemia viruses, herpesviruses, and Epstein-Barr virus. J. Am. Acad. Dermatol. 2019, 81, 23–41. [Google Scholar] [CrossRef]
  219. El-Sharkawy, A.; Al Zaidan, L.; Malki, A. Epstein–Barr Virus-Associated Malignancies: Roles of Viral Oncoproteins in Carcinogenesis. Front. Oncol. 2018, 8, 265. [Google Scholar] [CrossRef] [Green Version]
  220. Yin, H.; Qu, J.; Peng, Q.; Gan, R. Molecular mechanisms of EBV-driven cell cycle progression and oncogenesis. Med. Microbiol. Immunol. 2018, 208, 573–583. [Google Scholar] [CrossRef] [Green Version]
  221. Luo, Y.; Liu, Y.; Wang, C.; Gan, R. Signaling pathways of EBV-induced oncogenesis. Cancer Cell Int. 2021, 21, 93. [Google Scholar] [CrossRef]
  222. O’Neill, L.A.; Bowie, A.G. Sensing and Signaling in Antiviral Innate Immunity. Curr. Biol. 2010, 20, R328–R333. [Google Scholar] [CrossRef] [Green Version]
  223. Barber, G.N. Innate immune DNA sensing pathways: STING, AIMII and the regulation of interferon production and inflammatory responses. Curr. Opin. Immunol. 2011, 23, 10–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Eliopoulos, A.G.; Young, L.S. LMP1 structure and signal transduction. Semin. Cancer Biol. 2001, 11, 435–444. [Google Scholar] [CrossRef] [PubMed]
  225. Ikushima, H.; Negishi, H.; Taniguchi, T. The IRF Family Transcription Factors at the Interface of Innate and Adaptive Immune Responses. Cold Spring Harb. Symp. Quant. Biol. 2013, 78, 105–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Xu, D.; Brumm, K.; Zhang, L. The Latent Membrane Protein 1 of Epstein-Barr Virus (EBV) Primes EBV Latency Cells for Type I Interferon Production. J. Biol. Chem. 2006, 281, 9163–9169. [Google Scholar] [CrossRef] [Green Version]
  227. Zhang, L.; Pagano, J.S.; Servant, M.J.; Tenoever, B.; Lin, R.; Barnes, B.; Lubyova, B.; Pitha, P.M.; Levy, D.E.; Marié, I.; et al. Review: Structure and Function of IRF-7. J. Interf. Cytokine Res. 2002, 22, 95–101. [Google Scholar] [CrossRef] [PubMed]
  228. Quan, T.E.; Roman, R.M.; Rudenga, B.J.; Holers, V.M.; Craft, J.E. Epstein-Barr virus promotes interferon-α production by plasmacytoid dendritic cells. Arthritis Rheum. 2010, 62, 1693–1701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Jangra, S.; Yuen, K.-S.; Botelho, M.G.; Jin, D.-Y. Epstein–Barr Virus and Innate Immunity: Friends or Foes? Microorganisms 2019, 7, 183. [Google Scholar] [CrossRef] [Green Version]
  230. Chemudupati, M.; Kenney, A.D.; Smith, A.C.; Fillinger, R.J.; Zhang, L.; Zani, A.; Liu, S.-L.; Anderson, M.Z.; Sharma, A.; Yount, J.S. Butyrate Reprograms Expression of Specific Interferon-Stimulated Genes. J. Virol. 2020, 94, e00326-20. [Google Scholar] [CrossRef] [PubMed]
  231. Nanbo, A.; Inoue, K.; Adachi-Takasawa, K.; Takada, K. Epstein-Barr virus RNA confers resistance to interferon-alpha-induced apoptosis in Burkitt’s lymphoma. EMBO J. 2002, 21, 954–965. [Google Scholar] [CrossRef] [PubMed]
  232. Rafiee, N.; Ravanshad, M.; Asadi, B.; Kianfar, R.; Maleki, A. Investigation of IL-2 and IFN-γ to EBV Peptides in Stimulated Whole Blood among Multiple Sclerosis Patients and Healthy Individuals. Intervirology 2021, 64, 203–208. [Google Scholar] [CrossRef] [PubMed]
  233. Shi, Y.; Lutz, C.T.; Jones, K.; Wockner, L.; Brennan, R.M.; Keane, C.; Chattopadhyay, P.K.; Roederer, M.; Price, D.A.; Cole, D.K.; et al. Interferon-γControl of EBV-Transformed B Cells: A Role for CD8+T Cells That Poorly Kill EBV-Infected Cells. Viral Immunol. 2002, 15, 213–225. [Google Scholar] [CrossRef]
  234. Christian, M. Dendritic cells during Epstein Barr virus infection. Front. Microbiol. 2014, 5, 308. [Google Scholar] [CrossRef]
  235. Petersen, T.; Møller-Larsen, A.; Ellermann-Eriksen, S.; Thiel, S.; Christensen, T. Effects of interferon-beta therapy on elements in the antiviral immune response towards the human herpesviruses EBV, HSV, and VZV, and to the human endogenous retroviruses HERV-H and HERV-W in multiple sclerosis. J. Neuroimmunol. 2012, 249, 105–108. [Google Scholar] [CrossRef]
  236. Siegel, A.M.; Heimall, J.; Freeman, A.F.; Hsu, A.P.; Brittain, E.; Brenchley, J.M.; Douek, D.C.; Fahle, G.H.; Cohen, J.I.; Holland, S.M.; et al. A Critical Role for STAT3 Transcription Factor Signaling in the Development and Maintenance of Human T Cell Memory. Immunity 2011, 35, 806–818. [Google Scholar] [CrossRef] [Green Version]
  237. Gaglia, M.M. Anti-viral and pro-inflammatory functions of Toll-like receptors during gamma-herpesvirus infections. Virol. J. 2021, 18, 218. [Google Scholar] [CrossRef]
  238. Valente, R.M.; Ehlers, E.; Xu, D.; Ahmad, H.; Steadman, A.; Blasnitz, L.; Zhou, Y.; Kastanek, L.; Meng, B.; Zhang, L. Toll-like Receptor 7 Stimulates the Expression of Epstein-Barr Virus Latent Membrane Protein 1. PLoS ONE 2012, 7, e43317. [Google Scholar] [CrossRef] [Green Version]
  239. Severa, M.; Giacomini, E.; Gafa, V.; Anastasiadou, E.; Rizzo, F.; Corazzari, M.; Romagnoli, A.; Trivedi, P.; Fimia, G.M.; Coccia, E.M. EBV stimulates TLR- and autophagy-dependent pathways and impairs maturation in plasmacytoid dendritic cells: Implications for viral immune escape. Eur. J. Immunol. 2012, 43, 147–158. [Google Scholar] [CrossRef]
  240. Lim, Y.X.; Ng, Y.L.; Tam, J.P.; Liu, D.X. Human Coronaviruses: A Review of Virus–Host Interactions. Diseases 2016, 4, 26. [Google Scholar] [CrossRef] [Green Version]
  241. Ribero, M.S.; Jouvenet, N.; Dreux, M.; Nisole, S. Interplay between SARS-CoV-2 and the type I interferon response. PLoS Pathog. 2020, 16, e1008737. [Google Scholar] [CrossRef]
  242. Min, Y.-Q.; Huang, M.; Sun, X.; Deng, F.; Wang, H.; Ning, Y.-J. Immune evasion of SARS-CoV-2 from interferon antiviral system. Comput. Struct. Biotechnol. J. 2021, 19, 4217–4225. [Google Scholar] [CrossRef]
  243. 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] [Green Version]
  244. Zhu, N.; Zhang, D.; Wang, W.; Li, X.; Yang, B.; Song, J.; Zhao, X.; Huang, B.; Shi, W.; Lu, R.; et al. A Novel Coronavirus from Patients with Pneumonia in China, 2019. N. Engl. J. Med. 2020, 382, 727–733. [Google Scholar] [CrossRef] [PubMed]
  245. 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] [PubMed] [Green Version]
  246. 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]
  247. Akira, S.; Uematsu, S.; Takeuchi, O. Pathogen recognition and innate immunity. Cell 2006, 124, 783–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Pichlmair, A.; Schulz, O.; Tan, C.P.; Näslundslund, T.I.; Liljeström, P.; Weber, F.; Reis e Sousa, C. RIG-I-Mediated Antiviral Responses to Single-Stranded RNA Bearing 5’-Phosphates. Science 2006, 314, 997–1001. [Google Scholar] [CrossRef] [Green Version]
  249. Setaro, A.C.; Gaglia, M.M. All hands on deck: SARS-CoV-2 proteins that block early anti-viral interferon responses. Curr. Res. Virol. Sci. 2021, 2, 100015. [Google Scholar] [CrossRef]
  250. Rashid, F.; Xie, Z.; Suleman, M.; Shah, A.; Khan, S.; Luo, S. Roles and Functions of SARS-CoV-2 Proteins in Host Immune Evasion. Front. Immunol. 2022, 13, 4122. [Google Scholar] [CrossRef] [PubMed]
  251. Raj, R. Analysis of non-structural proteins, NSPs of SARS-CoV-2 as targets for computational drug designing. Biochem. Biophys. Rep. 2020, 25, 100847. [Google Scholar] [CrossRef] [PubMed]
  252. 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] [PubMed]
  253. Baczenas, J.J.; Andersen, H.; Rashid, S.; Yarmosh, D.; Puthuveetil, N.; Parker, M.; Bradford, R.; Florence, C.; Stemple, K.J.; Lewis, M.G.; et al. Propagation of SARS-CoV-2 in Calu-3 Cells to Eliminate Mutations in the Furin Cleavage Site of Spike. Viruses 2021, 13, 2434. [Google Scholar] [CrossRef]
  254. Busnadiego, I.; Fernbach, S.; Pohl, M.O.; Karakus, U.; Huber, M.; Trkola, A.; Stertz, S.; Hale, B.G. Antiviral Activity of Type I, II, and III Interferons Counterbalances ACE2 Inducibility and Restricts SARS-CoV-2. mBio 2020, 11, e01928-20. [Google Scholar] [CrossRef]
  255. Rebendenne, A.; Valadão, A.L.C.; Tauziet, M.; Maarifi, G.; Bonaventure, B.; McKellar, J.; Planès, R.; Nisole, S.; Arnaud-Arnould, M.; Moncorgé, O.; et al. SARS-CoV-2 Triggers an MDA-5-Dependent Interferon Response Which Is Unable To Control Replication in Lung Epithelial Cells. J. Virol. 2021, 95, e02415-20. [Google Scholar] [CrossRef]
  256. Grossegesse, M.; Bourquain, D.; Neumann, M.; Schaade, L.; Schulze, J.; Mache, C.; Wolff, T.; Nitsche, A.; Doellinger, J. Deep Time Course Proteomics of SARS-CoV- and SARS-CoV-2-Infected Human Lung Epithelial Cells (Calu-3) Reveals Strong Induction of Interferon-Stimulated Gene Expression by SARS-CoV-2 in Contrast to SARS-CoV. J. Proteome Res. 2022, 21, 459–469. [Google Scholar] [CrossRef]
  257. 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]
  258. Schroeder, S.; Pott, F.; Niemeyer, D.; Veith, T.; Richter, A.; Muth, D.; Goffinet, C.; A Müller, M.; Drosten, C. Interferon antagonism by SARS-CoV-2: A functional study using reverse genetics. Lancet Microbe 2021, 2, e210–e218. [Google Scholar] [CrossRef]
  259. Taylor, J.K.; Coleman, C.M.; Postel, S.; Sisk, J.M.; Bernbaum, J.G.; Venkataraman, T.; Sundberg, E.J.; Frieman, M.B. Severe Acute Respiratory Syndrome Coronavirus ORF7a Inhibits Bone Marrow Stromal Antigen 2 Virion Tethering through a Novel Mechanism of Glycosylation Interference. J. Virol. 2015, 89, 11820–11833. [Google Scholar] [CrossRef] [Green Version]
  260. Kopecky-Bromberg, S.A.; Martínez-Sobrido, L.; Frieman, M.; Baric, R.A.; Palese, P. Severe Acute Respiratory Syndrome Coronavirus Open Reading Frame (ORF) 3b, ORF 6, and Nucleocapsid Proteins Function as Interferon Antagonists. J. Virol. 2007, 81, 548–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Kim, Y.-M.; Shin, E.-C. Type I and III interferon responses in SARS-CoV-2 infection. Exp. Mol. Med. 2021, 53, 750–760. [Google Scholar] [CrossRef] [PubMed]
  262. Kovarik, P.; Castiglia, V.; Ivin, M.; Ebner, F. Type I Interferons in Bacterial Infections: A Balancing Act. Front. Immunol. 2016, 7, 652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Alphonse, N.; Dickenson, R.E.; Odendall, C. Interferons: Tug of War Between Bacteria and Their Host. Front. Cell Infect. Microbiol. 2021, 11, 624094. [Google Scholar] [CrossRef] [PubMed]
  264. Ayala, A.V.; Hsu, C.-Y.; Matsuo, K.; Buzun, E.; Terrazas, M.C.; Loomis, L.R.; Lu, H.-H.; Park, J.H.; Rivaud, P.; Thomson, M.; et al. Commensal Bacteria Promote Type I Interferon Signaling to Maintain Immune Tolerance. bioRxiv 2021, 2021, 464743. [Google Scholar]
  265. Alphonse, N.; Dickenson, R.E.; Alrehaili, A.; Odendall, C. Functions of IFNλs in Anti-Bacterial Immunity at Mucosal Barriers. Front. Immunol. 2022, 13, 7639. [Google Scholar] [CrossRef]
  266. Peignier, A.; Parker, D. Impact of Type I Interferons on Susceptibility to Bacterial Pathogens. Trends Microbiol. 2021, 29, 823–835. [Google Scholar] [CrossRef]
  267. Perry, A.K.; Chen, G.; Zheng, D.; Tang, H.; Cheng, G. The host type I interferon response to viral and bacterial infections. Cell Res. 2005, 15, 407–422. [Google Scholar] [CrossRef] [Green Version]
  268. An, J.; Woodward, J.J.; Sasaki, T.; Minie, M.; Elkon, K.B. Cutting Edge: Antimalarial Drugs Inhibit IFN-β Production through Blockade of Cyclic GMP-AMP Synthase–DNA Interaction. J. Immunol. 2015, 194, 4089–4093. [Google Scholar] [CrossRef] [Green Version]
  269. Fu, Q.; He, Q.; Dong, Q.; Xie, J.; Geng, Y.; Han, H.; Huang, Y.; Lu, J.; Zeng, Z.; Wang, W.; et al. The role of cyclic GMP-AMP synthase and Interferon-I-inducible protein 16 as candidatebiomarkers of systemic lupus erythematosus. Clin. Chim. Acta 2021, 524, 69–77. [Google Scholar] [CrossRef]
  270. Nagarajan, U. Induction and Function of IFNβ During Viral and Bacterial Infection. Crit. Rev. Immunol. 2011, 31, 459–474. [Google Scholar] [CrossRef] [PubMed]
  271. Hu, T.; Pan, M.; Yin, Y.; Wang, C.; Cui, Y.; Wang, Q. The Regulatory Network of Cyclic GMP-AMP Synthase-Stimulator of Interferon Genes Pathway in Viral Evasion. Front. Microbiol. 2021, 12, 3635. [Google Scholar] [CrossRef] [PubMed]
  272. Andrade, W.A.; Agarwal, S.; Mo, S.; Shaffer, S.A.; Dillard, J.P.; Schmidt, T.; Hornung, V.; Fitzgerald, K.A.; Kurt-Jones, E.A.; Golenbock, D.T. Type I Interferon Induction by Neisseria gonorrhoeae: Dual Requirement of Cyclic GMP-AMP Synthase and Toll-like Receptor 4. Cell Rep. 2016, 15, 2438–2448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Andrade, W.A.; Firon, A.; Schmidt, T.; Hornung, V.; Fitzgerald, K.A.; Kurt-Jones, E.A.; Trieu-Cuot, P.; Golenbock, D.T.; Kaminski, P.-A. Group B Streptococcus Degrades Cyclic-di-AMP to Modulate STING-Dependent Type I Interferon Production. Cell Host Microbe 2016, 20, 49–59. [Google Scholar] [CrossRef] [Green Version]
  274. Collins, A.C.; Cai, H.; Li, T.; Franco, L.H.; Li, X.-D.; Nair, V.R.; Scharn, C.R.; Stamm, C.E.; Levine, B.; Chen, Z.J.; et al. Cyclic GMP-AMP Synthase Is an Innate Immune DNA Sensor for Mycobacterium tuberculosis. Cell Host Microbe 2015, 17, 820–828. [Google Scholar] [CrossRef] [Green Version]
  275. Hansen, K.; Prabakaran, T.; Laustsen, A.; E Jørgensen, S.; Rahbæk, S.H.; Jensen, S.B.; Nielsen, R.; Leber, J.H.; Decker, T.; A Horan, K.; et al. Listeria monocytogenes induces IFNβ expression through an IFI16-, cGAS- and STING-dependent pathway. EMBO J. 2014, 33, 1654–1666. [Google Scholar] [CrossRef] [Green Version]
  276. Storek, K.M.; Gertsvolf, N.A.; Ohlson, M.B.; Monack, D.M. cGAS and Ifi204 Cooperate To Produce Type I IFNs in Response to Francisella Infection. J. Immunol. 2015, 194, 3236–3245. [Google Scholar] [CrossRef] [Green Version]
  277. Wassermann, R.; Gulen, M.F.; Sala, C.; Perin, S.G.; Lou, Y.; Rybniker, J.; Schmid-Burgk, J.L.; Schmidt, T.; Hornung, V.; Cole, S.T.; et al. Mycobacterium tuberculosis Differentially Activates cGAS- and Inflammasome-Dependent Intracellular Immune Responses through ESX-1. Cell Host Microbe 2015, 17, 799–810. [Google Scholar] [CrossRef] [Green Version]
  278. Watson, R.O.; Bell, S.L.; MacDuff, D.A.; Kimmey, J.M.; Diner, E.J.; Olivas, J.; Vance, R.E.; Stallings, C.L.; Virgin, H.W.; Cox, J.S. The Cytosolic Sensor cGAS Detects Mycobacterium tuberculosis DNA to Induce Type I Interferons and Activate Autophagy. Cell Host Microbe 2015, 17, 811–819. [Google Scholar] [CrossRef] [Green Version]
  279. Kim, T.H.; Kim, D.; Gautam, A.; Lee, H.; Kwak, M.H.; Park, M.C.; Park, S.; Wu, G.; Lee, B.L.; Lee, Y.; et al. CpG-DNA exerts antibacterial effects by protecting immune cells and producing bacteria-reactive antibodies. Sci. Rep. 2018, 8, 16236. [Google Scholar] [CrossRef] [Green Version]
  280. Van Seventer, J. Type I Interferon in LPS/CpG-Induced Lethal Toxic Shock. Available online: https://grantome.com/grant/NIH/R21-AI061433-02 (accessed on 16 February 2023).
  281. Carrero, J.A. Confounding roles for type I interferons during bacterial and viral pathogenesis. Int. Immunol. 2013, 25, 663–669. [Google Scholar] [CrossRef] [Green Version]
  282. Castiglia, V.; Piersigilli, A.; Ebner, F.; Janos, M.; Goldmann, O.; Damböck, U.; Kröger, A.; Weiss, S.; Knapp, S.; Jamieson, A.M.; et al. Type I Interferon Signaling Prevents IL-1β-Driven Lethal Systemic Hyperinflammation during Invasive Bacterial Infection of Soft Tissue. Cell Host Microbe 2016, 19, 375–387. [Google Scholar] [CrossRef] [Green Version]
  283. Fieber, C.; Janos, M.; Koestler, T.; Gratz, N.; Li, X.-D.; Castiglia, V.; Aberle, M.; Sauert, M.; Wegner, M.; Alexopoulou, L.; et al. Innate Immune Response to Streptococcus pyogenes Depends on the Combined Activation of TLR13 and TLR2. PLoS ONE 2015, 10, e0119727. [Google Scholar] [CrossRef] [Green Version]
  284. Hidmark, A.; Paul, A.V.S.; Dalpke, A.H. Cutting Edge: TLR13 Is a Receptor for Bacterial RNA. J. Immunol. 2012, 189, 2717–2721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Li, X.-D.; Chen, Z.J. Sequence specific detection of bacterial 23S ribosomal RNA by TLR13. Elife 2012, 1, e00102. [Google Scholar] [CrossRef] [PubMed]
  286. Oldenburg, M.; Krüger, A.; Ferstl, R.; Kaufmann, A.; Nees, G.; Sigmund, A.; Bathke, B.; Lauterbach, H.; Suter, M.; Dreher, S.; et al. TLR13 Recognizes Bacterial 23S RRNA Devoid of Erythromycin Resistance–Forming Modification. Science 2012, 337, 1111–1115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Bergstrøm, B.; Aune, M.H.; Awuh, J.A.; Kojen, J.F.; Blix, K.J.; Ryan, L.; Flo, T.H.; Mollnes, T.E.; Espevik, T.; Stenvik, J. TLR8 Senses Staphylococcus aureus RNA in Human Primary Monocytes and Macrophages and Induces IFN-β Production via a TAK1–IKKβ–IRF5 Signaling Pathway. J. Immunol. 2015, 195, 1100–1111. [Google Scholar] [CrossRef] [Green Version]
  288. Krüger, A.; Oldenburg, M.; Chebrolu, C.; Beisser, D.; Kolter, J.; Sigmund, A.M.; Steinmann, J.; Schäfer, S.; Hochrein, H.; Rahmann, S.; et al. Human TLR 8 senses UR/URR motifs in bacterial and mitochondrial RNA. EMBO Rep. 2015, 16, 1656–1663. [Google Scholar] [CrossRef]
  289. Kagan, J.C.; Su, T.; Horng, T.; Chow, A.; Akira, S.; Medzhitov, R. TRAM couples endocytosis of Toll-like receptor 4 to the induction of interferon-β. Nat. Immunol. 2008, 9, 361–368. [Google Scholar] [CrossRef] [Green Version]
  290. Aubry, C.; Corr, S.C.; Wienerroither, S.; Goulard, C.; Jones, R.; Jamieson, A.M.; Decker, T.; O’Neill, L.A.J.; Dussurget, O.; Cossart, P. Both TLR2 and TRIF Contribute to Interferon-β Production during Listeria Infection. PLoS ONE 2012, 7, e33299. [Google Scholar] [CrossRef]
  291. UniProt. Available online: https://www.uniprot.org/ (accessed on 16 February 2023).
  292. Kozlowski, L.P. IPC—Isoelectric Point Calculator. Biol. Direct 2016, 11, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Services. Available online: https://services.healthtech.dtu.dk (accessed on 16 February 2023).
  294. Align|UniProt. Available online: https://www.uniprot.org/align (accessed on 16 February 2023).
Figure 1. Mechanisms of signaling in the cell by different types of interferons (based on [6,77]). IFN—interferon; IFNAR—IFNA receptor; IFNAR1—IFNα receptor 1; IFNAR2—IFNα receptor 2; IRF—regulatory factor; ISG—IFN stimulated gene; JAK-STAT—Janus kinase/signal transducer and activator of transcription; STAT1—signal transducer and activator of transcription 1; STAT2—signal transducer and activator of transcription 2; TYK2—tyrosine kinase 2; GAF—IFNγ activation factor; ISGF3—IFN-stimulated gene factor 3; IRF9—IFN-regulatory factor 9; ISRE—IFN-stimulated response elements; IFNGR—IFNγ receptor; IFNGR1—IFNγ receptor 1; IFNGR2—IFNγ receptor 2; IFNLR1—IFNλ receptor 1 (also known as IL-28Ra) and IL-10R2 (also known as IL-10Rβ); GAS—gamma-activated sequences; P—phosphorus.
Figure 1. Mechanisms of signaling in the cell by different types of interferons (based on [6,77]). IFN—interferon; IFNAR—IFNA receptor; IFNAR1—IFNα receptor 1; IFNAR2—IFNα receptor 2; IRF—regulatory factor; ISG—IFN stimulated gene; JAK-STAT—Janus kinase/signal transducer and activator of transcription; STAT1—signal transducer and activator of transcription 1; STAT2—signal transducer and activator of transcription 2; TYK2—tyrosine kinase 2; GAF—IFNγ activation factor; ISGF3—IFN-stimulated gene factor 3; IRF9—IFN-regulatory factor 9; ISRE—IFN-stimulated response elements; IFNGR—IFNγ receptor; IFNGR1—IFNγ receptor 1; IFNGR2—IFNγ receptor 2; IFNLR1—IFNλ receptor 1 (also known as IL-28Ra) and IL-10R2 (also known as IL-10Rβ); GAS—gamma-activated sequences; P—phosphorus.
Ijms 24 10115 g001
Figure 2. Scheme of the antiviral effect of interferons in a cell (based on [55,96,97]).
Figure 2. Scheme of the antiviral effect of interferons in a cell (based on [55,96,97]).
Ijms 24 10115 g002
Figure 3. Effects of type I IFN signaling in bacterial infection (based on [18,262,263,264]).
Figure 3. Effects of type I IFN signaling in bacterial infection (based on [18,262,263,264]).
Ijms 24 10115 g003
Figure 4. Search strategy, study selection, and data extraction according to the PRISMA statement.
Figure 4. Search strategy, study selection, and data extraction according to the PRISMA statement.
Ijms 24 10115 g004
Table 2. Characteristics of the occurrence of interferon isoforms.
Table 2. Characteristics of the occurrence of interferon isoforms.
NameIsoformProtein IDCharacteristicReference
IFN-α6 Alpha-6X1AP23229-2259-297: Missing
1084-1130:RYDDSVPRYHAVRIRKEEREIKDEKYIDNLEKKQWITKWNENESYS → NKKDHYDATYHKAEIHAQPSDKERLTSDA
[33]
Alpha-6X1BP23229-3259-297: Missing
Alpha-6X2AP23229-4215-258: Missing
1084-1130:SRYDDSVPRYHAVRIRKEEREIKDEKYIDNLEKKQWITKWNENESYS → NKKDHYDATYHKAEIHAQPSDKERLTSDA
Alpha-6X2BP23229-5215-258: Missing
Alpha-6X1X2AP23229-61084-1130:SRYDDSVPRYHAVRIRKEEREIKDEKYIDNLEKKQWITKWNENESYS → NKKDHYDATYHKAEIHAQPSDKERLTSDA
7P23229-71-114: Missing
215-258: Missing
1084-1130:SRYDDSVPRYHAVRIRKEEREIKDEKYIDNLEKKQWITKWNENESYS → NKKDHYDATYHKAEIHAQPSDKERLTSDA
9P23229-9918-932: Missing
1084-1130:SRYDDSVPRYHAVRIRKEEREIKDEKYIDNLEKKQWITKWNENESYS → NKKDHYDATYHKAEIHAQPSDKERLTSDA
IFN-α72P36544-218-18: H → HGKATASPPSTPPWDPGHIPGASVRPAPGP[33]
3P36544-381-102: SWTDHYLQWNVSEYPGVKTVRF → AYSRVPATSMYAGFPLMCSTAN
103-502: Missing
IFN-α16 IFI 16BQ16666-2444-499: Missing[38]
IFI 16CQ16666-3444-555: Missing
4Q16666-6128-183: Missing
IFN-λ4p170K9M1U5-2123-179:LELARPGSSRKVPGAQKRRHKPRRADSPRCRKASVVFNLLRLLTWELRLAAHSGPCL → VSDGRAPPPLSPASFSASSGPRRAPALCQSVLLSGKTHPDRSRVLWVS[50]
p131K9M1U5-375-122: Missing
p107K9M1U5-451-122: Missing
Table 3. Effector pathways of the antiviral response in which interferons are involved.
Table 3. Effector pathways of the antiviral response in which interferons are involved.
NameCharacteristicsReference
Mx antiviral protein pathway
  • Mx proteins are evolutionarily conserved large dynamin-like GTPases, and GTPase activity is required for their antiviral activity;
  • They are key mediators of innate antiviral resistance induced in cells by type I (α/β) and type III (λ) interferons (IFNs);
  • Inhibits a wide range of viruses by blocking the early stage of the replication cycle.
[84,85,86,87]
Ribonuclease L pathway targeting 2’,5’-oligoadenylate synthetase
  • The 2′,5′-oligoadenylate synthetase (OAS)/RNase L pathway is a component of innate immunity that responds to a pathogen-associated molecular pattern (dsRNA) to degrade viral and cellular RNAs, thereby blocking the infection;
  • The 2-5A molecule is a unique ligand for RNase L and is believed to be a trigger for the signaling of innate antiviral immunity mechanisms by activating RNase L.
[88,89]
Protein kinase R pathway
  • Protein Kinase R (PKR) is one of the best-characterized IFN effector molecules;
  • Mediates responses by phosphorylation of protein substrates and promotes signal transduction pathways to maintain homeostasis;
  • Involved in activating immune mechanisms and promoting apoptosis in cells.
[90,91,92]
Pathway similar to ISG15
  • ISG15 is a member of the ubiquitin family, which includes ubiquitin and ubiquitin-like modifiers;
  • This protein is strongly induced by type I interferons;
  • ISG15 may compete with ubiquitin for ubiquitin-binding sites on the protein, thereby indirectly regulating protein degradation;
  • This protein acts as a cytokine and can also interact with various intracellular protein partners.
[93,94,95]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mertowska, P.; Smolak, K.; Mertowski, S.; Grywalska, E. Immunomodulatory Role of Interferons in Viral and Bacterial Infections. Int. J. Mol. Sci. 2023, 24, 10115. https://doi.org/10.3390/ijms241210115

AMA Style

Mertowska P, Smolak K, Mertowski S, Grywalska E. Immunomodulatory Role of Interferons in Viral and Bacterial Infections. International Journal of Molecular Sciences. 2023; 24(12):10115. https://doi.org/10.3390/ijms241210115

Chicago/Turabian Style

Mertowska, Paulina, Konrad Smolak, Sebastian Mertowski, and Ewelina Grywalska. 2023. "Immunomodulatory Role of Interferons in Viral and Bacterial Infections" International Journal of Molecular Sciences 24, no. 12: 10115. https://doi.org/10.3390/ijms241210115

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