Next Article in Journal
The Interplay among Glucocorticoid Therapy, Platelet-Activating Factor and Endocannabinoid Release Influences the Inflammatory Response to COVID-19
Next Article in Special Issue
IFITM2 Presents Antiviral Response through Enhancing Type I IFN Signaling Pathway
Previous Article in Journal
Fragment-Based Approaches Identified Tecovirimat-Competitive Novel Drug Candidate for Targeting the F13 Protein of the Monkeypox Virus
Previous Article in Special Issue
Nonstructural Protein 2 Is Critical to Infection Efficiency of Highly Pathogenic Porcine Reproductive and Respiratory Syndrome Virus on PAMs and Influence Virulence In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antagonisms of ASFV towards Host Defense Mechanisms: Knowledge Gaps in Viral Immune Evasion and Pathogenesis

1
Jiangsu Co-Innovation Center for Prevention and Control of Important Animal Infectious Diseases and Zoonoses, College of Veterinary Medicine, Yangzhou University, Yangzhou 225009, China
2
Joint International Research Laboratory of Agriculture and Agri-Product Safety, Ministry of Education, Yangzhou University, Yangzhou 225009, China
3
Luoyang Putai Biotech Co., Ltd., Luoyang 471003, China
4
College of Animal Science and Veterinary Medicine, Henan Agricultural University, Zhengzhou 450002, China
*
Authors to whom correspondence should be addressed.
Viruses 2023, 15(2), 574; https://doi.org/10.3390/v15020574
Submission received: 28 December 2022 / Revised: 16 February 2023 / Accepted: 16 February 2023 / Published: 19 February 2023
(This article belongs to the Special Issue Porcine Anti-viral Immunity)

Abstract

:
African swine fever (ASF) causes high morbidity and mortality of both domestic pigs and wild boars and severely impacts the swine industry worldwide. ASF virus (ASFV), the etiologic agent of ASF epidemics, mainly infects myeloid cells in swine mononuclear phagocyte system (MPS), including blood-circulating monocytes, tissue-resident macrophages, and dendritic cells (DCs). Since their significant roles in bridging host innate and adaptive immunity, these cells provide ASFV with favorable targets to manipulate and block their antiviral activities, leading to immune escape and immunosuppression. To date, vaccines are still being regarded as the most promising measure to prevent and control ASF outbreaks. However, ASF vaccine development is delayed and limited by existing knowledge gaps in viral immune evasion, pathogenesis, etc. Recent studies have revealed that ASFV can employ diverse strategies to interrupt the host defense mechanisms via abundant self-encoded proteins. Thus, this review mainly focuses on the antagonisms of ASFV-encoded proteins towards IFN-I production, IFN-induced antiviral response, NLRP3 inflammasome activation, and GSDMD-mediated pyroptosis. Additionally, we also make a brief discussion concerning the potential challenges in future development of ASF vaccine.

1. Introduction

African swine fever (ASF) was first discovered in Kenya in the 1920s and gradually spread across many regions covering the Caucasus, sub-Saharan Africa, and Eastern Europe [1]. The first outbreak of ASF in China was reported on 3 August 2018 [2,3]. Subsequently, the disease spread throughout the country with an unprecedented speed, causing huge economic losses to local pig industry [4]. Later, the emergence and coexistence of naturally mutated, low-virulent genotype I and II ASFV field strains has posed more challenges for this disease’s prevention and control [5,6].
African swine fever virus (ASFV) is the causative agent of ASF. It is an enveloped, icosahedral, double-stranded DNA arbovirus with a genomic length ranging from 170 to 193 kilobase (kb) and is the only member of Asfarviridae family [7,8,9]. ASFV exhibits complexity because of its large genome since this virus can encode multifunctional proteins enough for its productive replication [10]. Especially, it is most likely that ASFV possesses an entire transcription machinery of its own to synthesize viral mRNA, including DNA-dependent RNA polymerase subunit (e.g., pEP1242L, pNP1450L), transcription factor (e.g., pI243L), RNA helicase (e.g., pQP509L, pA859L), RNA capping enzyme (e.g., pNP868R), etc. [8]. Moreover, ASFV is known to encode DNA ligase (pNP419L), DNA polymerase X-like (pO174L), lambda-like exonuclease (pD345L), AP endonuclease (pE296R), and PCNA-like (pE301R), which are essential for the base excision repair (BER) [11]. The complicated genomic features make ASFV a difficult opponent.
Vaccines are recognized as the most useful tool to prevent and control viral infection [12,13]. Currently, advances have been made in the research of ASF live attenuated vaccines (LAVs). LAVs can confer some degree of protection against challenge with homologous parental strains, which is proven to be effective. Although these protective LAVs exhibit attenuated characteristics, their safety has always been a controversial issue. As reported, ASFV-G-ΔI177L, a candidate for ASF LAV, is constructed by deleting a single I177L gene from the highly virulent Georgia 2007 isolate. Such experimental vaccine candidates present as completely attenuated at a low or even high dose of inoculation. Vaccinated pigs all remain clinically normal, with low viremia titers and strong virus-specific antibody response [14]. However, assessment on the safety of vaccine candidate ASFV-G-ΔI177L shows that virus shedding from vaccinated pigs can be detected for couple of days [15]. Of interest, ASFV-G-ΔI177L was once approved for commercial use in June 2022 and two months later suspended for its safety issues by authorities in Vietnam [16]. Thus, caution should be maintained about the gene-deleted ASF LAVs.
Although other attempts for developing ASFV vaccines have also been conducted and evaluated in the past decades, no prophylactic and therapeutic vaccines are commercially available to effectively eradicate ASF epidemics [17,18]. ASFV adapts various strategies to suppress host immunity and escape from the innate and adaptive immune responses, which may be responsible for the restricted vaccine development [19,20,21]. For instance, previous studies reported that ASFV could disturb antigen processing and presentation targeting the expression of major histocompatibility complex (MHC) molecules [22]. Furthermore, several studies indicated that ASFV could also induce massive destruction of lymphocyte subsets, which is characterized by the apoptosis of bystander non-infected B and T lymphocytes in vivo [23,24].
To date, a few articles have already highlighted the negative impact of plentiful ASFV proteins on antiviral immune response [25,26,27]. However, several ASFV immunosuppressive proteins (e.g., pE184L, pD345L, pD117L) and novel immune evasion mechanisms have not been summarized. Thus, this review will focus on relevant studies that deeply elucidate the antagonisms of ASFV towards host defense mechanisms in the last five years (2019–2023). In the following sections, we aim to elaborate these mechanisms by which ASFV targets key signals to counteract host antiviral pathway in different manners, mainly analyze the implications that ASFV proteins-mediated antagonisms pose on viral replication, pathogenicity, and virulence in vivo, and briefly discuss some knowledge gaps that should be filled in the future research.

2. Suppression of IFN-I Production and IFN-Induced Antiviral Responses

Type I interferon (IFN-I) serves as critical immune mediators for restricting the spread of viral infection [28,29]. Recent studies have showed that both cGAS-STING and RIG-I-MAVS pathways involve in the production of IFN-I during ASFV infection [30,31]. After secreted outside of cells, the IFN-I will bind their receptors on the cytomembrane and initiate the Janus kinase-signal transducer and activator of transcription (JAK-STAT) signaling cascade, leading to the transcriptional regulation of hundreds of IFN-stimulated genes (ISGs) [32,33,34]. This process presents a remarkable antiviral state in vivo [23]. Although IFN-I-induced immunity provides an effective line of defense, ASFV has evolved several strategies to antagonize it [35,36]. Indeed, ASFV encodes multiple proteins that can manipulate and evade host antiviral response by specific interactions with key elements of the JAK-STAT pathway, as illustrated in Figure 1.

2.1. Impairment on IFN-I Production Targeting cGAS-STING Axis

Recent studies have demonstrated that several ASFV-encoded proteins targeting cGAS-STING pathway can inhibit IFN-I production in diverse manners (Table 1) [25]. In this pathway, cyclic GMP-AMP synthase (cGAS), which was previously identified to be a cytosolic DNA sensor, can directly recognize and bind to the non-self ASFV DNA [37]. After that, cGAS is activated and then catalyzes the synthesis of 2′3′-cyclic GMP-AMP (cGAMP) [37]. As a critical secondary messenger in the cGAS-STING pathway, cGAMP will bind to and activate stimulator of interferon genes (STING) [38]. However, cGAMP is likely to be cleaved by viral nucleases, which certainly restricts the STING-dependent IFN-I production [39,40]. Indeed, ASFV-encoded pEP364R and pC129R, homologs to the nuclease, exhibit strong phosphodiesterase activity [41]. Either pEP364R or pC129R can selectively interact with cGAMP, specifically mediating the cleavage of cGAMP with their enzyme activity [41].
ASFV is also likely to impair the IFN-I production by targeting critical proteins (e.g., STING, IRF3, TBK1) in the downstream cGAS-STING pathway. Firstly, activated STING will further recruit TANK-binding kinase 1 (TBK1) and IFN regulatory factor 3 (IRF3) to form a trimeric complex “STING-IRF3-TBK1” [42]. However, ASFV-encoded pE184L can prevent the formation of “STING-IRF3-TBK1” complex by impairing the oligomerization and dimerization of STING [43]. ASFV structural protein p17 can also disrupt the recruitment of TBK1 and Ikkε by interacting with STING as well although its implications on STING are not clearly elucidated [44]. Secondly, TBK1, a crucial kinase, is required for the IRF3 phosphorylation [45]. The kinase activity of TBK1 is strictly regulated by multiple post-translational modifications (PTMs), including ubiquitination and phosphorylation [46]. As reported, phosphorylation of serine residues at position 172 in the activation loop of TBK1 and K63-linked polyubiquitination of lysine residues at positions 30 and 401 of TBK1 can both activate its kinase activity [47,48]. Nevertheless, early-expressed ASFV pDP96R can significantly suppress the phosphorylation of TBK1 [49]. The only known viral ubiquitin-conjugating enzyme (UBCv1) encoded by the ASFV I215L gene can enhance RING finger protein (RNF)138 to degrade RNF128, which inhibits K63-linked polyubiquitination of TBK1 [50,51]. Except for the negative modulation on PTMs of TBK1, late-expressed pA137R can also mediate the autophagosome and lysosome-dependent degradation of TBK1 to block the STING-IRF3 signaling pathway [52]. Of note, ASFV-encoded pS273R, a member of the SUMO-1-specific protease family, can affect the SUMOylation of inhibitor of nuclear factor kappa-B kinase ε (Ikkε) by its catalytic activity [53]. Ikkε, a homolog of TBK1, seems to involve in the activation of STING-IRF3 signaling pathway although the mechanism is not fully clear [54]. Thirdly, phosphorylated IRF3 will dissociate from the complex and then translocate into the nucleus, which consequently triggers the transcription of interferon genes (e.g., IFN-I) [55]. Of note is that the transcriptional activity of IRF3 is also modulated by posttranslational modifications (e.g., phosphorylation). Nonetheless, ASFV pE301R, pI226R, and pE120R can suppress the phosphorylation of IRF3 to interfere with normal transcriptional function of IRF3 [56,57,58]. ASFV pM1249L, expressed in the late phase of ASFV infection, exhibits dual inhibitory effects on IFN-I production since pM1249L can not only suppress TBK1 phosphorylation but also mediate the lysosome-dependent degradation of IRF3 [59]. Taken together, interference with PTMs of critical signals is one of the major strategies employed by ASFV proteins to impair STING-mediated IFN-I production.
The multigene families (MGFs) of ASFV can exploit cellular protein degradation systems to impair STING-IRF3 signaling pathway mainly by mediating the degradation of critical signals. The MGFs locate at the left terminal 40 kb and right terminal 20 kb variable regions in ASFV genome [60]. They are mainly grouped into MGF-100, MGF-110, MGF-300, MGF-360, and MGF-505/530, whose gain or loss causes variation in the genomes of different ASFV isolates [61]. Among them, pMGF360-11L can mediate the caspase-, proteasome-, and autophagosome-dependent degradation of TBK1 and IRF7 [62]. pMGF505-11R can mediate the lysosomal-, proteasome-, and autophagosome-dependent degradation of STING [63]. The viral non-structural protein pMGF360-14L can mediate the degradation of IRF3 by facilitating E3 ligase TRIM21-mediated K63-linked ubiquitination [64]. Early-expressed pMGF505-7R executes multifaceted inhibition on STING-dependent antiviral responses [65]. pMGF505-7R can mediate proteasome-dependent degradation of TBK1, caspase-, autophagosome-, and proteasome-dependent degradation of IRF7 and autophagosome-dependent degradation of STING [66,67]. Additionally, pMGF-505-7R can also facilitate the degradation of STING by upregulating Unc-51-like autophagy-activating kinase 1 (ULK1) [67,68].
Table 1. Antagonisms of ASFV proteins in latest reports towards IFN-I production by targeting cGAS-STING axis.
Table 1. Antagonisms of ASFV proteins in latest reports towards IFN-I production by targeting cGAS-STING axis.
ASFV ProteinFunctional Site/DomainKey TargetUnderlying MechanismReference
pEP364R, pC129RAmino acids Y76 and N78 in pep364rcGAMPSelectively cleave cgamp[41]
pE184LAmino acids 1–20 in pe184lSTINGImpair STING oligomerization and dimerization[43]
p17 (pD117L)Amino acids 39–59 in p17STINGInterfere with the recruitment of TBK1 and Ikkε[44]
pDP96RAmino acids 30–96 in pdp96rTBK1Suppress the phosphorylation of TBK1[49]
pI215L (UBCv1)UnknownTBK1Inhibit K63-linked polyubiquitination of TBK1[50]
pA137RUnknownTBK1Mediate the degradation of TBK1[52]
pM1249LUnknownTBK1; IRF3Suppress the phosphorylation of TBK1; mediate the degradation of IRF3[59]
pE120RAmino acids 72–73 in pe120rIRF3Suppress the phosphorylation of IRF3[58]
pE301RAmino acids 1–200 in pe301rIRF3Suppress the phosphorylation of IRF3[56]
pI226RUnknownIRF3Suppress the phosphorylation of IRF3[57]
pS273RAmino acids 1–20 and 256–273 in ps273rIkkεAffect the sumoylation of Ikkε[53]
pMGF360-11LAmino acids 167–353 in pmgf360-11LTBK1; IRF7Mediate the degradation of TBK1 and IRF7[62]
pMGF505-11RAmino acids 1–191 and 182–360 in pmgf505-11RSTINGMediate the degradation of STING[63]
pMGF360-14LUnknownIRF3Mediate the degradation of IRF3[64]
pMGF-505-7RUnknownTBK1; IRF7Mediate the degradation of TBK1 and IRF7[66]
STINGMediate the degradation of STING[67]

2.2. Impairment on IFN-I Production Targeting RIG-I-MAVS Axis

During infection, ASFV DNA, as a danger signal, will be released into the cytoplasm. Previous studies showed that DNA-dependent RNA polymerase III (Pol-III) can also detect cytosolic DNA and trigger the production of IFN-I through the RIG-I-MAVS pathway [69]. Pol-III serves AT-rich double-stranded DNA (dsDNA) as the template to transcribe it into double-stranded RNA (dsRNA) containing a 5′-triphosphate end [70]. The newly formed 5′-triphosphate RNA can be detected and tightly bound by RIG-I in the cytoplasm, which initiates the downstream signaling cascade [71]. K63-linked polyubiquitination has proven to be an important regulation for the RIG-I activation [72]. However, some viruses can effectively disturb the RIG-I-mediated antiviral signaling by targeting the process of ubiquitination [73,74]. Ran et al. confirmed that AT-rich regions of ASFV genomes can be recognized and transcribed into AU-rich 5′pppRNA transcripts by Pol-III [31]. They further demonstrated that ASFV virulence factor pI267L originating from either genotype I or II can potently antagonize the RNA Pol-III-RIG-I axis [31]. Mechanistically, pI267L destroys the stabilized and enhanced form of activated RIG-I by impairing Riplet-mediated K63-linked polyubiquitination, a critical step in RIG-I-MAVS signaling pathway [31,75,76].
Furthermore, viruses can exploit the host metabolism to synthesize plentiful metabolites required for their replication, which facilitates viral productive infection [77,78]. To be mentioned, the crosstalk between innate immunity and metabolism is well discussed in several reports [79,80,81]. Recent studies have revealed that ASFV can alter the host cellular metabolism to disrupt innate immune responses for self-replication by impairing RIG-I-mediate IFN-I production [82]. Mechanistically, ASFV infection triggers the increase of pyruvate production, giving rise to an enhanced level of lactate under the action of lactate dehydrogenase (LDH) [82]. Notably, lactate is a natural suppressor of RIG-I-mediated signaling pathway by targeting MAVS, which lowers the expression of beta interferon (IFN-β) [83]. However, the mechanisms by which ASFV can modulate cellular metabolism to promote productive infection have been rarely elucidated. More studies on the metabolomic analysis of ASFV-infected cells maybe provide novel insights into the association between metabolic regulation and immune evasion.

2.3. Impairment on IFN-Induced Antiviral Responses Targeting JAK-STAT Pathway

The activation of the JAK-STAT pathway by IFNs will upregulate the expression of hundreds of ISGs, leading to the remarkable restriction of viral spread and replication [84]. However, to survive and propagate within the host cells, ASFV has encoded multiple proteins to counteract the IFN-induced antiviral responses [36]. Recently, studies have suggested that ASFV proteins mainly block the JAK-STAT pathway by mediating the degradation of critical signals (e.g., JAKs, STATs, IRF9). The IFN-stimulated gene factor 3 (ISG3) complex is formed through the interaction of STAT1/STAT2 heterodimers with IRF9, which plays crucial roles in the IFN-I-triggered JAK-STAT pathway [32]. ASFV pI215L (UBCv1) can interfere with the formation of ISG3 in both ubiquitin-conjugating enzyme-activity-independent and dependent manners. ASFV pI215L (UBCv1) can interact with IRF9 and mediate the degradation of IRF9 via autophagy-lysosome pathway, which is independent of its ubiquitin-conjugating enzyme activity [85]. Additionally, ASFV pI215L (UBCv1) can also mediate the degradation of STAT2 via ubiquitin-proteasome pathway, which is dependent on its ubiquitin-conjugating enzyme activity [86]. In addition, Zhang et al. identified pMGF360-9L as also working as an inhibitor of JAK/STAT pathway [87]. Mechanically, pMGF360-9L inhibits IFN-β-induced ISGs transcription by mediating the degradation of STAT1 and STAT2 through the apoptotic pathway and ubiquitin-proteasome pathway, respectively [87]. Li et al. reported that pMGF-505-7R inhibited the IFN-γ-mediated JAK-STAT axis [88]. Mechanistically, pMGF-505-7R is found to interact with JAK1 and JAK2 and mediates their degradation by upregulating E3 ubiquitin ligase RNF125 expression and inhibiting expression of Hes5, respectively [88].

3. Inhibition of NLRP3 Inflammasome Activation and GSDMD-Mediated Pyroptosis

The NACHT, LRR, and PYD domains-containing protein 3 (NLRP3) inflammasome has become an indispensable component of host innate immune system, as it can effectively sense viral invasion and trigger strong inflammatory response [89]. Two signals are required for NLRP3 inflammasome activation [90]. The first signal, also called the priming signal, will activate NF-κB and promote the transcription of proinflammatory genes including NLRP3, pro-IL-1β, and pro-IL-18. The second signal, also known as the activation signal, will trigger the assembly of NLRP3 inflammasome [90].
To counteract, ASFV can encode multifunctional proteins to inhibit the activation of NF-κB, restrict the nuclear translocation of NF-κB, and disrupt the assembly of NLRP3 inflammasome by targeting the two signals (Figure 2). Firstly, one of the crucial steps in NF-κB activation is the phosphorylation of inhibitor of NF-κB alpha (IκBα) by canonical IκB kinases (i.e., IKKα and IKKβ) [91]. The canonical IκB kinases (IKKs) associate with an adapter protein NEMO (also known as IKKγ) to form “the canonical IKK complex” through their NEMO-binding domain [92]. NEMO without kinase activity usually acts as a ubiquitin-binding protein whose interaction with polyubiquitin chains is imperative for the canonical IKKs activation [93,94,95]. Nevertheless, ASFV-encoded capsid protein pH240R can mediate proteasome- and lysosome-dependent degradation of NEMO to ultimately block the activation of NF-κB [96]. As reported, phosphorylation of serine residues at positions 176 and 180 in IKKα or at positions 177 and 181 in IKKβ can make their kinase activity activated [97]. ASFV-encoded pF317L can inhibit the phosphorylation of IKKβ and further suppress NF-κB activation by decreasing phosphorylation and degradation of IκBα [98]. Additionally, ASFV pD345L and pMGF505-7R have both proven to interfere with NF-κB activation through their interaction with IKKα and/or IKKβ, but the mechanisms have not been deeply explored [99,100].
Secondly, NF-κB is normally associated with IκBα to form an inactive complex and sequestered in the cytoplasm [101]. In response to viral infection, NF-κB can be activated and translocate from cytoplasm into nucleus to exert its function on the transcription of proinflammatory genes [101]. However, ASFV-encoded pMGF360-12L can interfere with the nuclear import of NF-κB by competitively inhibiting the interaction between p65 and importin-α or karyopherin-α (KPNA) subtypes [102]. Moreover, three multifunctional proteins, namely pI215L (UBCv1), pMGF505-7R (A528R), and pF317L, have all been identified to restrict the nuclear translocation of NF-κB by performing immunofluorescence assay, but the mechanisms are not fully clarified [103,104]. Thirdly, upon activation, the sensor protein NLRP3 will recruit apoptosis-associated speck-like protein containing a CARD (ASC) and pro-caspase-1 to form a multiprotein complex called NLRP3 inflammasome [105]. However, ASFV pH240R can disrupt inflammasome assembly by suppressing NLRP3 oligomerization [96]. In addition, pMGF505-7R can also bind to NLRP3 to disturb the formation of inflammasome, but the mechanism is not further elucidated [100].
Of note, the assembly of NLRP3 inflammasome will further promote the activation of pro-caspase-1 [106]. Activated caspase-1 can cleave gasdermin D (GSDMD) in the interdomain linker to release the N terminal fragment of GSDMD (GSDMD-NT) [107]. GSDMD-NT oligomerizes in the cytomembrane and forms pores, which induces the lytic form of death-termed pyroptosis [107]. Pyroptosis functions as one of host defense mechanisms to restrict viral replication and facilitate the elimination of virus-infected cells [108]. As reported, viruses can utilize viral proteases to regulate pyroptosis [109,110]. Recently, Zhao et al. revealed that the ASFV-encoded pS273R can inhibit pyroptosis through noncanonical cleavage of swine GSDMD [111]. In the late stage of ASFV infection, pS273R cleaves GSDMD at G107-A108 to produce a shorter GSDMD-NT (N1~107) [111]. Unlike the canonical GSDMD-NT (N1~279) produced by caspase-1, GSDMD-NT (N1~107) loses its pore-forming activity on cytomembrane and is unable to induce pyroptosis [111]. Therefore, it is reasonable to speculate that GSDMD-mediated pyroptosis may be radically inhibited during ASFV infection.

4. Effects of Immune Evasion on Viral Replication and Virulence

Innate immunity serves as the crucial line of defense-exerting functions in protection against viral invasion. However, IFN-I activity and inflammatory response, the two principal components of innate immunity, are significantly blocked by ASFV proteins. The relevant mechanisms have already been elaborated above. It is reasonable to presume that antagonisms of ASFV proteins towards innate immunity may promote its own proliferation and alter biological characteristics of virus. Thus, the implications that ASFV proteins-mediated immune evasion pose on viral replication, pathogenicity, and virulence in vivo should be fully investigated, which may provide rational design of LAVs.
Here, immunosuppression-related genes H240R, MGF505-7R, E184L, I226R, and A137R are highlighted, referring to the latest reports on ASF LAVs. After deleting these genes individually from highly virulent ASFV strains, a significant decrease in pathogenicity and virulence can be observed in the mutants (Table 2). Previous studies have confirmed that deletion of the H240R gene will enhance NLRP3-mediated inflammatory responses, resulting in the attenuation of ASFV [112]. Of note, MGF505-7R has proven to be a multifunctional protein, playing crucial roles in suppressing cGAS-STING-mediated IFN-I production, IFN-II-induced antiviral response, and NLRP3 inflammasome activation [65]. Indeed, deletion of the aMGF505-7R gene does make ASFV trigger higher level of IFN-I and IL-1β in pigs, which may contribute to its attenuation. Although ASF LAV candidates, which are constructed by deleting a single gene (e.g., E184L, I226R, A137R), also exhibit attenuated characteristics to different extent, their mechanisms still remain to be further explored.
In addition, ASF LAV candidates SY18ΔI226R, Georgia/2010-ΔA137R, and Georgia/2010-ΔE184L all induce medium-to-high level of viremia (Table 2). The level of viremia reflects the level of viral replication in vivo [113]. Thus, these data demonstrate that single deletion of a gene (e.g., E184L, I226R, A137R) does not significantly affect ASFV replication in vivo. However, E184L, I226R, and A137R can all significantly inhibit viral replication in vitro through interfering with cGAS-STING-mediated IFN-I production. ASFV proteins may mediate antagonisms in vivo in a compensatory but not redundant manner. MGF360-9L and MGF505-7R, two inhibitors of IFN-induced antiviral response, exhibit synergistic restriction on viral replication in vivo. Indeed, combinational deletions of MGF360-9L and MGF505-7R attenuate ASFV and induce a much lower level of viremia than the parental strain does in vivo [114].
Table 2. Viremia, clinical signs, and death in LAVs-inoculated pigs compared with parental strains.
Table 2. Viremia, clinical signs, and death in LAVs-inoculated pigs compared with parental strains.
LAV Candidates vs. Parental StrainsViremia (Replication)Clinical Signs
(Pathogenicity)
Death
(Virulence)
Reference
HLJ/18UnknownYes, 6/6Yes, 6/6[112]
HLJ/18-ΔH240RUnknownNo, 0/6No, 0/6
HLJ/18Yes, highUnknownYes, 5/5[100]
HLJ/18-Δ7RYes, mediumUnknownYes, 2/5
Georgia/2010Yes, highYes, 5/5Yes, 5/5[115]
Georgia/2010-ΔE184LYes, medium to highYes, 2/5Yes, 1/5
SY18Yes, highYes, 5/5Yes, 5/5[116]
SY18-ΔI226RYes, medium to highNo, 0/5No, 0/5
Georgia/2010Yes, highYes, 5/5Yes, 5/5[117]
Georgia/2010-ΔA137RYes, medium to highNo, 0/5No, 0/5
CN/GS/2018Yes, highYes, 6/6Yes, 6/6[114]
CN/GS/2018-Δ9L/Δ7RYes, lowNo, 0/6No, 0/6

5. Future Perspective

The availability of safe and efficient vaccines is required for the control and eradication of ASF epidemics [118]. Thus, works contributing to the rational development of protective ASF vaccines should be a high priority [119]. The prerequisite for developing such effective vaccines is to better understand how ASFV antagonizes host immunity and pathogenesis of ASFV infection [120]. Unfortunately, there are still deep gaps that should be filled in above research fields (Figure 3).
Firstly, the involvement of nucleic acid sensors in ASFV detection has not been thoroughly understood. Although recent studies have highlighted that the cGAS/STING pathway plays predominant roles in resisting ASFV infection, other DNA sensors may be involved in ASFV recognition as well [27]. In particular, the potential functions of toll-like receptor 9 (TLR9) and interferon gamma inducible protein 16 (IFI16) have not yet been well investigated. TLR9 is a DNA-sensing receptor expressed in professional innate immune cells such as DCs and macrophages [121]. It recognizes unmethylated CpG-rich DNA of microbial origins [122]. TLR9 activation promotes the synthesis of proinflammatory cytokines such as IL-12, IL-6, and TNF-α, which is consistent with those induced by ASFV [123,124]. Moreover, recent studies have indicated that knockdown of TLR9 significantly down-regulated ASFV-Δ7R-triggered pro-IL-1β transcription [100]. These data suggest that TLR9 might be involved in the recognition of ASFV invasion. Most importantly, there is abundant unmethylated CpG DNA within the ASFV genome [125]. However, the ASFV genome remains wrapped by the thick core shell until it is released into the cytoplasm. Thus, it is not clear how unmethylated CpG DNA in ASFV genome can be exposed to the TLR9. To the best of our knowledge, endogenous mitochondrial DNA (mtDNA), whose CpG motifs are also unmethylated, is the ligand for TLR9 [126]. Therefore, special cases should be considered in which ASFV infection may indirectly activate the TLR9 pathway by triggering mitophagy-mediated mtDNA release [127]. To date, the relationship between the ASFV and TLR9 signaling pathway has not been elucidated yet. In addition, IFI16 can recognize many DNA viruses and detect genomic lesions following DNA damage [128]. Previous studies confirmed that the host DNA damage response (DDR) is activated from the early stage of ASFV infection [129]. Therefore, it is reasonable to speculate that IFI16 may recognize ASFV through binding to viral DNA or sensing ASFV-induced host DNA damage. More nucleic acid sensors involved in ASFV recognition need to be identified in the near future.
Secondly, the pathogenesis of ASFV has not been clearly elucidated. Hyperactivation of the immune system will result in a sharp and robust increase of proinflammatory cytokines, leading to a “cytokine storm” [130]. Cytokine storm syndrome, defined as a collection of severe clinical manifestations, is characterized by systemic inflammation, multi-organ failure, etc. [130]. As previously reported, the viruses (e.g., SARS-CoV-2, pseudorabies virus)-induced cytokine storm seems to be associated with their highly pathogenic infections, which could even give rise to rapid fatalities [131,132]. Indeed, ASF is also a devastating disease with high mortality in pigs. Recently, Wang et al. confirmed that the cytokine storm is involved in the pathogenesis of ASFV [124]. Upon the infection of type II virulent ASFV SY18 in domestic pigs, they characterized the kinetics of representative cytokines (e.g., interferons, interleukins, growth factors, tumor necrosis factors, and chemokines) circulating in vivo. As a result, they observed ASFV-induced cytokine storm in vivo. ASFV-infected pigs showed severe clinical symptoms from 3 days post inoculation (dpi) and died from 7 to 8 dpi. Of interest, except for IFN-γ, the majority of proinflammatory cytokines had a robust and sustained elevation throughout the ASFV SY18 infection. In addition, a two-time increase of the levels of TNF-α, IL-1β, IL-6, and IL-8 presented an irreversible immune status as well. Notably, Kanneganti et al. found that synergism of TNF-α and IFN-γ induced PANoptosis, and in turn, TNF-α and IFN-γ-mediated PANoptosis maintained the status of “cytokine storm”, which is extremely critical for the pathogenic processes of COVID-19 [133]. PANoptosis is activated by specific stimulus and modulated by its core “PANoptosome” complex that provides a molecular scaffold for extensive crosstalk of key molecules among pyroptosis, apoptosis, and necroptosis [134]. Whether PANoptosis is involved in the maintenance of ASFV-induced “cytokine storm” status and which cytokines synergize to drive the PANoptosis during ASFV infection deserve further investigation.

Author Contributions

Conceptualization, L.Y., X.L. and K.T.; writing, L.Y., Z.Z. and X.L.; manuscript revision and supervision, L.Y., J.D. and X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Jiangsu Agricultural Science and Technology Independent Innovation Fund Project (CX(21)2035), Jiangsu Provincial Key R&D plan (BE2020398), the National Key R&D Program of China (2021YFD1801302-1), the 111 Project D18007, and the Project of the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qu, H.; Ge, S.; Zhang, Y.; Wu, X.; Wang, Z. A systematic review of genotypes and serogroups of African swine fever virus. Virus Genes 2022, 58, 77–87. [Google Scholar] [CrossRef]
  2. Zhou, X.; Li, N.; Luo, Y.; Liu, Y.; Miao, F.; Chen, T.; Zhang, S.; Cao, P.; Li, X.; Tian, K.; et al. Emergence of African Swine Fever in China, 2018. Transbound. Emerg. Dis. 2018, 65, 1482–1484. [Google Scholar] [CrossRef] [Green Version]
  3. Ge, S.; Li, J.; Fan, X.; Liu, F.; Li, L.; Wang, Q.; Ren, W.; Bao, J.; Liu, C.; Wang, H.; et al. Molecular Characterization of African Swine Fever Virus, China, 2018. Emerg. Infect. Dis. 2018, 24, 2131–2133. [Google Scholar] [CrossRef] [Green Version]
  4. Zhao, D.; Liu, R.; Zhang, X.; Li, F.; Wang, J.; Zhang, J.; Liu, X.; Wang, L.; Zhang, J.; Wu, X.; et al. Replication and virulence in pigs of the first African swine fever virus isolated in China. Emerg. Microbes. Infect. 2019, 8, 438–447. [Google Scholar] [CrossRef] [Green Version]
  5. Sun, E.; Zhang, Z.; Wang, Z.; He, X.; Zhang, X.; Wang, L.; Wang, W.; Huang, L.; Xi, F.; Huangfu, H.; et al. Emergence and prevalence of naturally occurring lower virulent African swine fever viruses in domestic pigs in China in 2020. Sci. China Life Sci. 2021, 64, 752–765. [Google Scholar] [CrossRef]
  6. Sun, E.; Huang, L.; Zhang, X.; Zhang, J.; Shen, D.; Zhang, Z.; Wang, Z.; Huo, H.; Wang, W.; Huangfu, H.; et al. Genotype I African swine fever viruses emerged in domestic pigs in China and caused chronic infection. Emerg. Microbes. Infect. 2021, 10, 2183–2193. [Google Scholar] [CrossRef]
  7. Wang, N.; Zhao, D.; Wang, J.; Zhang, Y.; Wang, M.; Gao, Y.; Li, F.; Wang, J.; Bu, Z.; Rao, Z.; et al. Architecture of African swine fever virus and implications for viral assembly. Science 2019, 366, 640–644. [Google Scholar] [CrossRef]
  8. Cackett, G.; Matelska, D.; Sykora, M.; Portugal, R.; Malecki, M.; Bahler, J.; Dixon, L.; Werner, F. The African Swine Fever Virus Transcriptome. J. Virol. 2020, 94, e00119-20. [Google Scholar] [CrossRef] [Green Version]
  9. Gaudreault, N.N.; Madden, D.W.; Wilson, W.C.; Trujillo, J.D.; Richt, J.A. African Swine Fever Virus: An Emerging DNA Arbovirus. Front. Vet. Sci. 2020, 7, 215. [Google Scholar] [CrossRef]
  10. Duan, X.; Ru, Y.; Yang, W.; Ren, J.; Hao, R.; Qin, X.; Li, D.; Zheng, H. Research progress on the proteins involved in African swine fever virus infection and replication. Front. Immunol. 2022, 13, 947180. [Google Scholar] [CrossRef]
  11. Wang, G.; Xie, M.; Wu, W.; Chen, Z. Structures and Functional Diversities of ASFV Proteins. Viruses 2021, 13, 2124. [Google Scholar] [CrossRef]
  12. Rappuoli, R. Vaccines: Science, health, longevity, and wealth. Proc. Natl. Acad. Sci. USA 2014, 111, 12282. [Google Scholar] [CrossRef] [Green Version]
  13. Urbano, A.C.; Ferreira, F. African swine fever control and prevention: An update on vaccine development. Emerg. Microbes. Infect. 2022, 11, 2021–2033. [Google Scholar] [CrossRef]
  14. Borca, M.V.; Ramirez-Medina, E.; Silva, E.; Vuono, E.; Rai, A.; Pruitt, S.; Holinka, L.G.; Velazquez-Salinas, L.; Zhu, J.; Gladue, D.P. Development of a Highly Effective African Swine Fever Virus Vaccine by Deletion of the I177L Gene Results in Sterile Immunity against the Current Epidemic Eurasia Strain. J. Virol. 2020, 94, e02017-19. [Google Scholar] [CrossRef]
  15. Tran, X.H.; Phuong, L.T.T.; Huy, N.Q.; Thuy, D.T.; Nguyen, V.D.; Quang, P.H.; Ngon, Q.V.; Rai, A.; Gay, C.G.; Gladue, D.P.; et al. Evaluation of the Safety Profile of the ASFV Vaccine Candidate ASFV-G-DeltaI177L. Viruses 2022, 14, 896. [Google Scholar] [CrossRef]
  16. Vietnam suspends African swine fever vaccine after pig deaths. Available online: https://www.reuters.com/world/asia-pacific/vietnam-suspends-african-swine-fever-vaccine-after-pig-deaths-2022-08-24/ (accessed on 27 December 2022).
  17. Wang, T.; Sun, Y.; Huang, S.; Qiu, H.J. Multifaceted Immune Responses to African Swine Fever Virus: Implications for Vaccine Development. Vet. Microbiol. 2020, 249, 108832. [Google Scholar] [CrossRef]
  18. Zhu, J.J. African Swine Fever Vaccinology: The Biological Challenges from Immunological Perspectives. Viruses 2022, 14, 2021. [Google Scholar] [CrossRef]
  19. Wu, L.; Yang, B.; Yuan, X.; Hong, J.; Peng, M.; Chen, J.L.; Song, Z. Regulation and Evasion of Host Immune Response by African Swine Fever Virus. Front. Microbiol. 2021, 12, 698001. [Google Scholar] [CrossRef]
  20. Zhu, J.J.; Ramanathan, P.; Bishop, E.A.; O’Donnell, V.; Gladue, D.P.; Borca, M.V. Mechanisms of African swine fever virus pathogenesis and immune evasion inferred from gene expression changes in infected swine macrophages. PLoS ONE 2019, 14, e0223955. [Google Scholar] [CrossRef]
  21. Chathuranga, K.; Lee, J.-S. African Swine Fever Virus (ASFV): Immunity and Vaccine Development. Vaccines 2023, 11, 199. [Google Scholar] [CrossRef]
  22. Franzoni, G.; Dei Giudici, S.; Oggiano, A. Infection, modulation and responses of antigen-presenting cells to African swine fever viruses. Virus Res. 2018, 258, 73–80. [Google Scholar] [CrossRef]
  23. Dixon, L.K.; Islam, M.; Nash, R.; Reis, A.L. African swine fever virus evasion of host defences. Virus Res. 2019, 266, 25–33. [Google Scholar] [CrossRef]
  24. Reis, A.L.; Netherton, C.; Dixon, L.K. Unraveling the Armor of a Killer: Evasion of Host Defenses by African Swine Fever Virus. J. Virol. 2017, 91, e02338-16. [Google Scholar] [CrossRef] [Green Version]
  25. He, W.R.; Yuan, J.; Ma, Y.H.; Zhao, C.Y.; Yang, Z.Y.; Zhang, Y.; Han, S.; Wan, B.; Zhang, G.P. Modulation of Host Antiviral Innate Immunity by African Swine Fever Virus: A Review. Animals 2022, 12, 2935. [Google Scholar] [CrossRef]
  26. Zheng, X.; Nie, S.; Feng, W.H. Regulation of antiviral immune response by African swine fever virus (ASFV). Virol. Sin. 2022, 37, 157–167. [Google Scholar] [CrossRef]
  27. Ayanwale, A.; Trapp, S.; Guabiraba, R.; Caballero, I.; Roesch, F. New Insights in the Interplay Between African Swine Fever Virus and Innate Immunity and Its Impact on Viral Pathogenicity. Front. Microbiol. 2022, 13, 958307. [Google Scholar] [CrossRef]
  28. Seo, Y.J.; Hahm, B. Type I interferon modulates the battle of host immune system against viruses. Adv. Appl. Microbiol. 2010, 73, 83–101. [Google Scholar] [CrossRef]
  29. Fan, W.; Jiao, P.; Zhang, H.; Chen, T.; Zhou, X.; Qi, Y.; Sun, L.; Shang, Y.; Zhu, H.; Hu, R.; et al. Inhibition of African Swine Fever Virus Replication by Porcine Type I and Type II Interferons. Front. Microbiol. 2020, 11, 1203. [Google Scholar] [CrossRef]
  30. Garcia-Belmonte, R.; Perez-Nunez, D.; Pittau, M.; Richt, J.A.; Revilla, Y. African Swine Fever Virus Armenia/07 Virulent Strain Controls Interferon Beta Production through the cGAS-STING Pathway. J. Virol. 2019, 93, e02298-18. [Google Scholar] [CrossRef] [Green Version]
  31. Ran, Y.; Li, D.; Xiong, M.G.; Liu, H.N.; Feng, T.; Shi, Z.W.; Li, Y.H.; Wu, H.N.; Wang, S.Y.; Zheng, H.X.; et al. African swine fever virus I267L acts as an important virulence factor by inhibiting RNA polymerase III-RIG-I-mediated innate immunity. PLoS Pathog. 2022, 18, e1010270. [Google Scholar] [CrossRef]
  32. 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]
  33. Cai, S.; Zheng, Z.; Cheng, J.; Zhong, L.; Shao, R.; Zheng, F.; Lai, Z.; Ou, J.; Xu, L.; Zhou, P.; et al. Swine Interferon-Inducible Transmembrane Proteins Potently Inhibit African Swine Fever Virus Replication. Front. Immunol. 2022, 13, 827709. [Google Scholar] [CrossRef]
  34. Munoz-Moreno, R.; Cuesta-Geijo, M.A.; Martinez-Romero, C.; Barrado-Gil, L.; Galindo, I.; Garcia-Sastre, A.; Alonso, C. Antiviral Role of IFITM Proteins in African Swine Fever Virus Infection. PLoS ONE 2016, 11, e0154366. [Google Scholar] [CrossRef] [Green Version]
  35. Razzuoli, E.; Franzoni, G.; Carta, T.; Zinellu, S.; Amadori, M.; Modesto, P.; Oggiano, A. Modulation of Type I Interferon System by African Swine Fever Virus. Pathogens 2020, 9, 361. [Google Scholar] [CrossRef]
  36. Portugal, R.; Leitao, A.; Martins, C. Modulation of type I interferon signaling by African swine fever virus (ASFV) of different virulence L60 and NHV in macrophage host cells. Vet. Microbiol. 2018, 216, 132–141. [Google Scholar] [CrossRef]
  37. Sun, L.; Wu, J.; Du, F.; Chen, X.; Chen, Z.J. Cyclic GMP-AMP synthase is a cytosolic DNA sensor that activates the type I interferon pathway. Science 2013, 339, 786–791. [Google Scholar] [CrossRef] [Green Version]
  38. Shang, G.; Zhang, C.; Chen, Z.J.; Bai, X.C.; Zhang, X. Cryo-EM structures of STING reveal its mechanism of activation by cyclic GMP-AMP. Nature 2019, 567, 389–393. [Google Scholar] [CrossRef]
  39. Eaglesham, J.B.; Pan, Y.; Kupper, T.S.; Kranzusch, P.J. Viral and metazoan poxins are cGAMP-specific nucleases that restrict cGAS-STING signalling. Nature 2019, 566, 259–263. [Google Scholar] [CrossRef]
  40. Hernaez, B.; Alonso, G.; Georgana, I.; El-Jesr, M.; Martin, R.; Shair, K.H.Y.; Fischer, C.; Sauer, S.; Maluquer de Motes, C.; Alcami, A. Viral cGAMP nuclease reveals the essential role of DNA sensing in protection against acute lethal virus infection. Sci. Adv. 2020, 6, eabb4565. [Google Scholar] [CrossRef]
  41. Dodantenna, N.; Ranathunga, L.; Chathuranga, W.A.G.; Weerawardhana, A.; Cha, J.W.; Subasinghe, A.; Gamage, N.; Haluwana, D.K.; Kim, Y.; Jheong, W.; et al. African Swine Fever Virus EP364R and C129R Target Cyclic GMP-AMP To Inhibit the cGAS-STING Signaling Pathway. J. Virol. 2022, 96, e0102222. [Google Scholar] [CrossRef]
  42. Ghosh, M.; Saha, S.; Bettke, J.; Nagar, R.; Parrales, A.; Iwakuma, T.; van der Velden, A.W.M.; Martinez, L.A. Mutant p53 suppresses innate immune signaling to promote tumorigenesis. Cancer Cell 2021, 39, 494–508.e5. [Google Scholar] [CrossRef]
  43. Zhu, Z.; Li, S.; Ma, C.; Yang, F.; Cao, W.; Liu, H.; Chen, X.; Feng, T.; Shi, Z.; Tian, H.; et al. African Swine Fever Virus E184L Protein Interacts with Innate Immune Adaptor STING to Block IFN Production for Viral Replication and Pathogenesis. J. Immunol. 2023, 210, 442–458. [Google Scholar] [CrossRef]
  44. Zheng, W.; Xia, N.; Zhang, J.; Cao, Q.; Jiang, S.; Luo, J.; Wang, H.; Chen, N.; Zhang, Q.; Meurens, F.; et al. African Swine Fever Virus Structural Protein p17 Inhibits cGAS-STING Signaling Pathway Through Interacting With STING. Front. Immunol. 2022, 13, 941579. [Google Scholar] [CrossRef]
  45. Liu, S.; Cai, X.; Wu, J.; Cong, Q.; Chen, X.; Li, T.; Du, F.; Ren, J.; Wu, Y.T.; Grishin, N.V.; et al. Phosphorylation of innate immune adaptor proteins MAVS, STING, and TRIF induces IRF3 activation. Science 2015, 347, aaa2630. [Google Scholar] [CrossRef] [Green Version]
  46. Zhou, R.; Zhang, Q.; Xu, P. TBK1, a central kinase in innate immune sensing of nucleic acids and beyond. Acta. Biochim. Biophys. Sin. 2020, 52, 757–767. [Google Scholar] [CrossRef]
  47. Larabi, A.; Devos, J.M.; Ng, S.L.; Nanao, M.H.; Round, A.; Maniatis, T.; Panne, D. Crystal structure and mechanism of activation of TANK-binding kinase 1. Cell Rep. 2013, 3, 734–746. [Google Scholar] [CrossRef] [Green Version]
  48. Tu, D.; Zhu, Z.; Zhou, A.Y.; Yun, C.H.; Lee, K.E.; Toms, A.V.; Li, Y.; Dunn, G.P.; Chan, E.; Thai, T.; et al. Structure and ubiquitination-dependent activation of TANK-binding kinase 1. Cell Rep. 2013, 3, 747–758. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, X.; Wu, J.; Wu, Y.; Chen, H.; Zhang, S.; Li, J.; Xin, T.; Jia, H.; Hou, S.; Jiang, Y.; et al. Inhibition of cGAS-STING-TBK1 signaling pathway by DP96R of ASFV China 2018/1. Biochem. Biophys. Res. Commun. 2018, 506, 437–443. [Google Scholar] [CrossRef]
  50. Huang, L.; Xu, W.; Liu, H.; Xue, M.; Liu, X.; Zhang, K.; Hu, L.; Li, J.; Liu, X.; Xiang, Z.; et al. African Swine Fever Virus pI215L Negatively Regulates cGAS-STING Signaling Pathway through Recruiting RNF138 to Inhibit K63-Linked Ubiquitination of TBK1. J. Immunol. 2021, 207, 2754–2769. [Google Scholar] [CrossRef]
  51. Song, G.; Liu, B.; Li, Z.; Wu, H.; Wang, P.; Zhao, K.; Jiang, G.; Zhang, L.; Gao, C. E3 ubiquitin ligase RNF128 promotes innate antiviral immunity through K63-linked ubiquitination of TBK1. Nat. Immunol. 2016, 17, 1342–1351. [Google Scholar] [CrossRef]
  52. Sun, M.; Yu, S.; Ge, H.; Wang, T.; Li, Y.; Zhou, P.; Pan, L.; Han, Y.; Yang, Y.; Sun, Y.; et al. The A137R Protein of African Swine Fever Virus Inhibits Type I Interferon Production via the Autophagy-Mediated Lysosomal Degradation of TBK1. J. Virol. 2022, 96, e0195721. [Google Scholar] [CrossRef]
  53. Luo, J.; Zhang, J.; Ni, J.; Jiang, S.; Xia, N.; Guo, Y.; Shao, Q.; Cao, Q.; Zheng, W.; Chen, N.; et al. The African swine fever virus protease pS273R inhibits DNA sensing cGAS-STING pathway by targeting IKKepsilon. Virulence 2022, 13, 740–756. [Google Scholar] [CrossRef]
  54. Balka, K.R.; Louis, C.; Saunders, T.L.; Smith, A.M.; Calleja, D.J.; D’Silva, D.B.; Moghaddas, F.; Tailler, M.; Lawlor, K.E.; Zhan, Y.; et al. TBK1 and IKKepsilon Act Redundantly to Mediate STING-Induced NF-kappaB Responses in Myeloid Cells. Cell Rep. 2020, 31, 107492. [Google Scholar] [CrossRef]
  55. Luo, W.W.; Tong, Z.; Cao, P.; Wang, F.B.; Liu, Y.; Zheng, Z.Q.; Wang, S.Y.; Li, S.; Wang, Y.Y. Transcription-independent regulation of STING activation and innate immune responses by IRF8 in monocytes. Nat. Commun. 2022, 13, 4822. [Google Scholar] [CrossRef]
  56. Liu, X.; Liu, H.; Ye, G.; Xue, M.; Yu, H.; Feng, C.; Zhou, Q.; Liu, X.; Zhang, L.; Jiao, S.; et al. African swine fever virus pE301R negatively regulates cGAS-STING signaling pathway by inhibiting the nuclear translocation of IRF3. Vet. Microbiol. 2022, 274, 109556. [Google Scholar] [CrossRef]
  57. Hong, J.; Chi, X.; Yuan, X.; Wen, F.; Rai, K.R.; Wu, L.; Song, Z.; Wang, S.; Guo, G.; Chen, J.L. I226R Protein of African Swine Fever Virus Is a Suppressor of Innate Antiviral Responses. Viruses 2022, 14, 575. [Google Scholar] [CrossRef]
  58. Liu, H.; Zhu, Z.; Feng, T.; Ma, Z.; Xue, Q.; Wu, P.; Li, P.; Li, S.; Yang, F.; Cao, W.; et al. African Swine Fever Virus E120R Protein Inhibits Interferon Beta Production by Interacting with IRF3 To Block Its Activation. J. Virol. 2021, 95, e0082421. [Google Scholar] [CrossRef]
  59. Cui, S.; Wang, Y.; Gao, X.; Xin, T.; Wang, X.; Yu, H.; Chen, S.; Jiang, Y.; Chen, Q.; Jiang, F.; et al. African swine fever virus M1249L protein antagonizes type I interferon production via suppressing phosphorylation of TBK1 and degrading IRF3. Virus Res. 2022, 319, 198872. [Google Scholar] [CrossRef]
  60. Zhu, Z.; Chen, H.; Liu, L.; Cao, Y.; Jiang, T.; Zou, Y.; Peng, Y. Classification and characterization of multigene family proteins of African swine fever viruses. Brief. Bioinform. 2021, 22, bbaa380. [Google Scholar] [CrossRef]
  61. Cackett, G.; Sykora, M.; Werner, F. Transcriptome view of a killer: African swine fever virus. Biochem. Soc. Trans. 2020, 48, 1569–1581. [Google Scholar] [CrossRef]
  62. Yang, K.; Xue, Y.; Niu, H.; Shi, C.; Cheng, M.; Wang, J.; Zou, B.; Wang, J.; Niu, T.; Bao, M.; et al. African swine fever virus MGF360–11L negatively regulates cGAS-STING-mediated inhibition of type I interferon production. Vet. Res. 2022, 53, 7. [Google Scholar] [CrossRef]
  63. Yang, K.; Huang, Q.; Wang, R.; Zeng, Y.; Cheng, M.; Xue, Y.; Shi, C.; Ye, L.; Yang, W.; Jiang, Y.; et al. African swine fever virus MGF505–11R inhibits type I interferon production by negatively regulating the cGAS-STING-mediated signaling pathway. Vet. Microbiol. 2021, 263, 109265. [Google Scholar] [CrossRef]
  64. Wang, Y.; Cui, S.; Xin, T.; Wang, X.; Yu, H.; Chen, S.; Jiang, Y.; Gao, X.; Jiang, Y.; Guo, X.; et al. African Swine Fever Virus MGF360–14L Negatively Regulates Type I Interferon Signaling by Targeting IRF3. Front. Cell Infect. Microbiol. 2021, 11, 818969. [Google Scholar] [CrossRef]
  65. Huang, L.; Li, J.; Zheng, J.; Li, D.; Weng, C. Multifunctional pMGF505–7R Is a Key Virulence-Related Factor of African Swine Fever Virus. Front. Microbiol. 2022, 13, 852431. [Google Scholar] [CrossRef]
  66. Yang, K.; Xue, Y.; Niu, T.; Li, X.; Cheng, M.; Bao, M.; Zou, B.; Shi, C.; Wang, J.; Yang, W.; et al. African swine fever virus MGF505–7R protein interacted with IRF7and TBK1 to inhibit type I interferon production. Virus Res. 2022, 322, 198931. [Google Scholar] [CrossRef]
  67. Li, D.; Yang, W.; Li, L.; Li, P.; Ma, Z.; Zhang, J.; Qi, X.; Ren, J.; Ru, Y.; Niu, Q.; et al. African Swine Fever Virus MGF-505–7R Negatively Regulates cGAS-STING-Mediated Signaling Pathway. J. Immunol. 2021, 206, 1844–1857. [Google Scholar] [CrossRef]
  68. Konno, H.; Konno, K.; Barber, G.N. Cyclic dinucleotides trigger ULK1 (ATG1) phosphorylation of STING to prevent sustained innate immune signaling. Cell 2013, 155, 688–698. [Google Scholar] [CrossRef] [Green Version]
  69. Chiu, Y.H.; Macmillan, J.B.; Chen, Z.J. RNA polymerase III detects cytosolic DNA and induces type I interferons through the RIG-I pathway. Cell 2009, 138, 576–591. [Google Scholar] [CrossRef] [Green Version]
  70. Ablasser, A.; Bauernfeind, F.; Hartmann, G.; Latz, E.; Fitzgerald, K.A.; Hornung, V. RIG-I-dependent sensing of poly(dA:dT) through the induction of an RNA polymerase III-transcribed RNA intermediate. Nat. Immunol. 2009, 10, 1065–1072. [Google Scholar] [CrossRef] [Green Version]
  71. Hornung, V.; Ellegast, J.; Kim, S.; Brzozka, K.; Jung, A.; Kato, H.; Poeck, H.; Akira, S.; Conzelmann, K.K.; Schlee, M.; et al. 5′-Triphosphate RNA is the ligand for RIG-I. Science 2006, 314, 994–997. [Google Scholar] [CrossRef] [Green Version]
  72. Okamoto, M.; Kouwaki, T.; Fukushima, Y.; Oshiumi, H. Regulation of RIG-I Activation by K63-Linked Polyubiquitination. Front Immunol 2017, 8, 1942. [Google Scholar] [CrossRef] [Green Version]
  73. Ban, J.; Lee, N.R.; Lee, N.J.; Lee, J.K.; Quan, F.S.; Inn, K.S. Human Respiratory Syncytial Virus NS 1 Targets TRIM25 to Suppress RIG-I Ubiquitination and Subsequent RIG-I-Mediated Antiviral Signaling. Viruses 2018, 10, 716. [Google Scholar] [CrossRef] [Green Version]
  74. Sanchez-Aparicio, M.T.; Feinman, L.J.; Garcia-Sastre, A.; Shaw, M.L. Paramyxovirus V Proteins Interact with the RIG-I/TRIM25 Regulatory Complex and Inhibit RIG-I Signaling. J. Virol. 2018, 92, e01950-17. [Google Scholar] [CrossRef] [Green Version]
  75. Hu, H.; Sun, S.C. Ubiquitin signaling in immune responses. Cell Res. 2016, 26, 457–483. [Google Scholar] [CrossRef] [Green Version]
  76. Maelfait, J.; Beyaert, R. Emerging role of ubiquitination in antiviral RIG-I signaling. Microbiol. Mol. Biol. Rev. 2012, 76, 33–45. [Google Scholar] [CrossRef] [Green Version]
  77. Zhang, Y.; Guo, R.; Kim, S.H.; Shah, H.; Zhang, S.; Liang, J.H.; Fang, Y.; Gentili, M.; Leary, C.N.O.; Elledge, S.J.; et al. SARS-CoV-2 hijacks folate and one-carbon metabolism for viral replication. Nat. Commun. 2021, 12, 1676. [Google Scholar] [CrossRef]
  78. Sanchez-Garcia, F.J.; Perez-Hernandez, C.A.; Rodriguez-Murillo, M.; Moreno-Altamirano, M.M.B. The Role of Tricarboxylic Acid Cycle Metabolites in Viral Infections. Front. Cell Infect. Microbiol. 2021, 11, 725043. [Google Scholar] [CrossRef]
  79. Yang, S.; Jin, S.; Xian, H.; Zhao, Z.; Wang, L.; Wu, Y.; Zhou, L.; Li, M.; Cui, J. Metabolic enzyme UAP1 mediates IRF3 pyrophosphorylation to facilitate innate immune response. Mol. Cell 2023, 83, 298–313.e8. [Google Scholar] [CrossRef]
  80. Gonzalez Plaza, J.J.; Hulak, N.; Kausova, G.; Zhumadilov, Z.; Akilzhanova, A. Role of metabolism during viral infections, and crosstalk with the innate immune system. Intractable Rare Dis. Res. 2016, 5, 90–96. [Google Scholar] [CrossRef] [Green Version]
  81. Ye, L.; Jiang, Y.; Zhang, M. Crosstalk between glucose metabolism, lactate production and immune response modulation. Cytokine Growth Factor Rev. 2022, 68, 81–92. [Google Scholar] [CrossRef]
  82. Xue, Q.; Liu, H.; Zhu, Z.; Yang, F.; Song, Y.; Li, Z.; Xue, Z.; Cao, W.; Liu, X.; Zheng, H. African Swine Fever Virus Regulates Host Energy and Amino Acid Metabolism To Promote Viral Replication. J. Virol. 2022, 96, e0191921. [Google Scholar] [CrossRef]
  83. Zhang, W.; Wang, G.; Xu, Z.G.; Tu, H.; Hu, F.; Dai, J.; Chang, Y.; Chen, Y.; Lu, Y.; Zeng, H.; et al. Lactate Is a Natural Suppressor of RLR Signaling by Targeting MAVS. Cell 2019, 178, 176–189.e15. [Google Scholar] [CrossRef]
  84. Schoggins, J.W.; Rice, C.M. Interferon-stimulated genes and their antiviral effector functions. Curr. Opin. Virol. 2011, 1, 519–525. [Google Scholar] [CrossRef]
  85. Li, L.; Fu, J.; Li, J.; Guo, S.; Chen, Q.; Zhang, Y.; Liu, Z.; Tan, C.; Chen, H.; Wang, X. African Swine Fever Virus pI215L Inhibits Type I Interferon Signaling by Targeting Interferon Regulatory Factor 9 for Autophagic Degradation. J. Virol. 2022, 96, e0094422. [Google Scholar] [CrossRef]
  86. Riera, E.; Garcia-Belmonte, R.; Madrid, R.; Perez-Nunez, D.; Revilla, Y. African swine fever virus ubiquitin-conjugating enzyme pI215L inhibits IFN-I signaling pathway through STAT2 degradation. Front. Microbiol. 2022, 13, 1081035. [Google Scholar] [CrossRef]
  87. Zhang, K.; Yang, B.; Shen, C.; Zhang, T.; Hao, Y.; Zhang, D.; Liu, H.; Shi, X.; Li, G.; Yang, J.; et al. MGF360–9L Is a Major Virulence Factor Associated with the African Swine Fever Virus by Antagonizing the JAK/STAT Signaling Pathway. mBio 2022, 13, e0233021. [Google Scholar] [CrossRef]
  88. Li, D.; Zhang, J.; Yang, W.; Li, P.; Ru, Y.; Kang, W.; Li, L.; Ran, Y.; Zheng, H. African swine fever virus protein MGF-505–7R promotes virulence and pathogenesis by inhibiting JAK1- and JAK2-mediated signaling. J. Biol. Chem. 2021, 297, 101190. [Google Scholar] [CrossRef]
  89. Zhao, C.; Zhao, W. NLRP3 Inflammasome-A Key Player in Antiviral Responses. Front. Immunol. 2020, 11, 211. [Google Scholar] [CrossRef] [Green Version]
  90. Jo, E.K.; Kim, J.K.; Shin, D.M.; Sasakawa, C. Molecular mechanisms regulating NLRP3 inflammasome activation. Cell. Mol. Immunol. 2016, 13, 148–159. [Google Scholar] [CrossRef] [Green Version]
  91. Zandi, E.; Rothwarf, D.M.; Delhase, M.; Hayakawa, M.; Karin, M. The IkappaB kinase complex (IKK) contains two kinase subunits, IKKalpha and IKKbeta, necessary for IkappaB phosphorylation and NF-kappaB activation. Cell 1997, 91, 243–252. [Google Scholar] [CrossRef] [Green Version]
  92. Israel, A. The IKK complex, a central regulator of NF-kappaB activation. Cold Spring Harb. Perspect Biol. 2010, 2, a000158. [Google Scholar] [CrossRef] [Green Version]
  93. Clark, K.; Nanda, S.; Cohen, P. Molecular control of the NEMO family of ubiquitin-binding proteins. Nat. Rev. Mol. Cell Biol. 2013, 14, 673–685. [Google Scholar] [CrossRef]
  94. Hadian, K.; Griesbach, R.A.; Dornauer, S.; Wanger, T.M.; Nagel, D.; Metlitzky, M.; Beisker, W.; Schmidt-Supprian, M.; Krappmann, D. NF-kappaB essential modulator (NEMO) interaction with linear and lys-63 ubiquitin chains contributes to NF-kappaB activation. J. Biol. Chem. 2011, 286, 26107–26117. [Google Scholar] [CrossRef] [Green Version]
  95. Du, M.; Ea, C.K.; Fang, Y.; Chen, Z.J. Liquid phase separation of NEMO induced by polyubiquitin chains activates NF-kappaB. Mol. Cell 2022, 82, 2415–2426.e5. [Google Scholar] [CrossRef]
  96. Zhou, P.; Dai, J.; Zhang, K.; Wang, T.; Li, L.F.; Luo, Y.; Sun, Y.; Qiu, H.J.; Li, S. The H240R Protein of African Swine Fever Virus Inhibits Interleukin 1beta Production by Inhibiting NEMO Expression and NLRP3 Oligomerization. J. Virol. 2022, 96, e0095422. [Google Scholar] [CrossRef]
  97. Kwak, Y.T.; Guo, J.; Shen, J.; Gaynor, R.B. Analysis of domains in the IKKalpha and IKKbeta proteins that regulate their kinase activity. J. Biol. Chem. 2000, 275, 14752–14759. [Google Scholar] [CrossRef] [Green Version]
  98. Yang, J.; Li, S.; Feng, T.; Zhang, X.; Yang, F.; Cao, W.; Chen, H.; Liu, H.; Zhang, K.; Zhu, Z.; et al. African Swine Fever Virus F317L Protein Inhibits NF-kappaB Activation To Evade Host Immune Response and Promote Viral Replication. mSphere 2021, 6, e0065821. [Google Scholar] [CrossRef]
  99. Chen, H.; Wang, Z.; Gao, X.; Lv, J.; Hu, Y.; Jung, Y.S.; Zhu, S.; Wu, X.; Qian, Y.; Dai, J. ASFV pD345L protein negatively regulates NF-kappaB signalling by inhibiting IKK kinase activity. Vet. Res. 2022, 53, 32. [Google Scholar] [CrossRef]
  100. Li, J.; Song, J.; Kang, L.; Huang, L.; Zhou, S.; Hu, L.; Zheng, J.; Li, C.; Zhang, X.; He, X.; et al. pMGF505–7R determines pathogenicity of African swine fever virus infection by inhibiting IL-1beta and type I IFN production. PloS Pathog. 2021, 17, e1009733. [Google Scholar] [CrossRef]
  101. Liu, T.; Zhang, L.; Joo, D.; Sun, S.C. NF-kappaB signaling in inflammation. Signal Transduct Target Ther. 2017, 2, 17023. [Google Scholar] [CrossRef] [Green Version]
  102. Zhuo, Y.; Guo, Z.; Ba, T.; Zhang, C.; He, L.; Zeng, C.; Dai, H. African Swine Fever Virus MGF360–12L Inhibits Type I Interferon Production by Blocking the Interaction of Importin alpha and NF-kappaB Signaling Pathway. Virol. Sin. 2021, 36, 176–186. [Google Scholar] [CrossRef]
  103. Barrado-Gil, L.; Del Puerto, A.; Galindo, I.; Cuesta-Geijo, M.A.; Garcia-Dorival, I.; de Motes, C.M.; Alonso, C. African Swine Fever Virus Ubiquitin-Conjugating Enzyme Is an Immunomodulator Targeting NF-kappaB Activation. Viruses 2021, 13, 1160. [Google Scholar] [CrossRef]
  104. Liu, X.; Ao, D.; Jiang, S.; Xia, N.; Xu, Y.; Shao, Q.; Luo, J.; Wang, H.; Zheng, W.; Chen, N.; et al. African Swine Fever Virus A528R Inhibits TLR8 Mediated NF-kappaB Activity by Targeting p65 Activation and Nuclear Translocation. Viruses 2021, 13, 2046. [Google Scholar] [CrossRef]
  105. Kelley, N.; Jeltema, D.; Duan, Y.; He, Y. The NLRP3 Inflammasome: An Overview of Mechanisms of Activation and Regulation. Int. J. Mol. Sci. 2019, 20, 3328. [Google Scholar] [CrossRef] [Green Version]
  106. Huang, Y.; Xu, W.; Zhou, R. NLRP3 inflammasome activation and cell death. Cell Mol. Immunol. 2021, 18, 2114–2127. [Google Scholar] [CrossRef]
  107. Liu, X.; Zhang, Z.; Ruan, J.; Pan, Y.; Magupalli, V.G.; Wu, H.; Lieberman, J. Inflammasome-activated gasdermin D causes pyroptosis by forming membrane pores. Nature 2016, 535, 153–158. [Google Scholar] [CrossRef] [Green Version]
  108. Kuriakose, T.; Kanneganti, T.D. Pyroptosis in Antiviral Immunity. Curr. Top Microbiol. Immunol. 2019. [Google Scholar] [CrossRef]
  109. Wen, W.; Li, X.; Wang, H.; Zhao, Q.; Yin, M.; Liu, W.; Chen, H.; Qian, P. Seneca Valley Virus 3C Protease Induces Pyroptosis by Directly Cleaving Porcine Gasdermin D. J. Immunol. 2021, 207, 189–199. [Google Scholar] [CrossRef]
  110. Yamaoka, Y.; Matsunaga, S.; Jeremiah, S.S.; Nishi, M.; Miyakawa, K.; Morita, T.; Khatun, H.; Shimizu, H.; Okabe, N.; Kimura, H.; et al. Zika virus protease induces caspase-independent pyroptotic cell death by directly cleaving gasdermin D. Biochem. Biophys. Res. Commun. 2021, 534, 666–671. [Google Scholar] [CrossRef]
  111. Zhao, G.; Li, T.; Liu, X.; Zhang, T.; Zhang, Z.; Kang, L.; Song, J.; Zhou, S.; Chen, X.; Wang, X.; et al. African swine fever virus cysteine protease pS273R inhibits pyroptosis by noncanonically cleaving gasdermin D. J. Biol. Chem. 2022, 298, 101480. [Google Scholar] [CrossRef]
  112. Huang, L.; Liu, H.; Ye, G.; Liu, X.; Chen, W.; Wang, Z.; Zhao, D.; Zhang, Z.; Feng, C.; Hu, L.; et al. Deletion of African Swine Fever Virus (ASFV) H240R Gene Attenuates the Virulence of ASFV by Enhancing NLRP3-Mediated Inflammatory Responses. J. Virol. 2023, e01227-22. [Google Scholar] [CrossRef]
  113. Christopher, J.; Burrell, C.R.H.; Murpy, F.A. Pathogenesis of Virus Infections. In Fenner and White’s Medical Virology, 5th ed.; Academic Press: Salt Lake City, UT, USA, 2016; p. 87. [Google Scholar]
  114. Ding, M.; Dang, W.; Liu, H.; Xu, F.; Huang, H.; Sunkang, Y.; Li, T.; Pei, J.; Liu, X.; Zhang, Y.; et al. Combinational Deletions of MGF360–9L and MGF505–7R Attenuated Highly Virulent African Swine Fever Virus and Conferred Protection against Homologous Challenge. J. Virol. 2022, 96, e0032922. [Google Scholar] [CrossRef]
  115. Ramirez-Medina, E.; Vuono, E.; Rai, A.; Pruitt, S.; Espinoza, N.; Velazquez-Salinas, L.; Pina-Pedrero, S.; Zhu, J.; Rodriguez, F.; Borca, M.V.; et al. Deletion of E184L, a Putative DIVA Target from the Pandemic Strain of African Swine Fever Virus, Produces a Reduction in Virulence and Protection against Virulent Challenge. J. Virol. 2022, 96, e0141921. [Google Scholar] [CrossRef]
  116. Zhang, Y.; Ke, J.; Zhang, J.; Yang, J.; Yue, H.; Zhou, X.; Qi, Y.; Zhu, R.; Miao, F.; Li, Q.; et al. African Swine Fever Virus Bearing an I226R Gene Deletion Elicits Robust Immunity in Pigs to African Swine Fever. J. Virol. 2021, 95, e0119921. [Google Scholar] [CrossRef]
  117. Gladue, D.P.; Ramirez-Medina, E.; Vuono, E.; Silva, E.; Rai, A.; Pruitt, S.; Espinoza, N.; Velazquez-Salinas, L.; Borca, M.V. Deletion of the A137R Gene from the Pandemic Strain of African Swine Fever Virus Attenuates the Strain and Offers Protection against the Virulent Pandemic Virus. J. Virol. 2021, 95, e0113921. [Google Scholar] [CrossRef]
  118. Wang, T.; Sun, Y.; Luo, Y.; Qiu, H.J. Prevention, control and vaccine development of African swine fever: Challenges and countermeasures. Sheng Wu Gong Cheng Xue Bao 2018, 34, 1931–1942. [Google Scholar] [CrossRef]
  119. Arias, M.; de la Torre, A.; Dixon, L.; Gallardo, C.; Jori, F.; Laddomada, A.; Martins, C.; Parkhouse, R.M.; Revilla, Y.; Rodriguez, F.A.J.; et al. Approaches and Perspectives for Development of African Swine Fever Virus Vaccines. Vaccines 2017, 5, 35. [Google Scholar] [CrossRef]
  120. Wang, Z.; Ai, Q.; Huang, S.; Ou, Y.; Gao, Y.; Tong, T.; Fan, H. Immune Escape Mechanism and Vaccine Research Progress of African Swine Fever Virus. Vaccines 2022, 10, 344. [Google Scholar] [CrossRef]
  121. Latz, E.; Schoenemeyer, A.; Visintin, A.; Fitzgerald, K.A.; Monks, B.G.; Knetter, C.F.; Lien, E.; Nilsen, N.J.; Espevik, T.; Golenbock, D.T. TLR9 signals after translocating from the ER to CpG DNA in the lysosome. Nat. Immunol. 2004, 5, 190–198. [Google Scholar] [CrossRef]
  122. Fiola, S.; Gosselin, D.; Takada, K.; Gosselin, J. TLR9 contributes to the recognition of EBV by primary monocytes and plasmacytoid dendritic cells. J. Immunol. 2010, 185, 3620–3631. [Google Scholar] [CrossRef] [Green Version]
  123. Gil, S.; Sepulveda, N.; Albina, E.; Leitao, A.; Martins, C. The low-virulent African swine fever virus (ASFV/NH/P68) induces enhanced expression and production of relevant regulatory cytokines (IFNalpha, TNFalpha and IL12p40) on porcine macrophages in comparison to the highly virulent ASFV/L60. Arch. Virol. 2008, 153, 1845–1854. [Google Scholar] [CrossRef] [Green Version]
  124. Wang, S.; Zhang, J.; Zhang, Y.; Yang, J.; Wang, L.; Qi, Y.; Han, X.; Zhou, X.; Miao, F.; Chen, T.; et al. Cytokine Storm in Domestic Pigs Induced by Infection of Virulent African Swine Fever Virus. Front. Vet. Sci. 2020, 7, 601641. [Google Scholar] [CrossRef]
  125. Weber, S.; Hakobyan, A.; Zakaryan, H.; Doerfler, W. Intracellular African swine fever virus DNA remains unmethylated in infected Vero cells. Epigenomics 2018, 10, 289–299. [Google Scholar] [CrossRef]
  126. Bao, W.; Xia, H.; Liang, Y.; Ye, Y.; Lu, Y.; Xu, X.; Duan, A.; He, J.; Chen, Z.; Wu, Y.; et al. Toll-like Receptor 9 Can be Activated by Endogenous Mitochondrial DNA to Induce Podocyte Apoptosis. Sci. Rep. 2016, 6, 22579. [Google Scholar] [CrossRef] [Green Version]
  127. Jing, R.; Hu, Z.K.; Lin, F.; He, S.; Zhang, S.S.; Ge, W.Y.; Dai, H.J.; Du, X.K.; Lin, J.Y.; Pan, L.H. Mitophagy-Mediated mtDNA Release Aggravates Stretching-Induced Inflammation and Lung Epithelial Cell Injury via the TLR9/MyD88/NF-kappaB Pathway. Front. Cell Dev. Biol. 2020, 8, 819. [Google Scholar] [CrossRef]
  128. Taffoni, C.; Steer, A.; Marines, J.; Chamma, H.; Vila, I.K.; Laguette, N. Nucleic Acid Immunity and DNA Damage Response: New Friends and Old Foes. Front. Immunol. 2021, 12, 660560. [Google Scholar] [CrossRef]
  129. Simoes, M.; Martins, C.; Ferreira, F. Host DNA damage response facilitates African swine fever virus infection. Vet. Microbiol. 2013, 165, 140–147. [Google Scholar] [CrossRef]
  130. Karki, R.; Kanneganti, T.D. The ‘cytokine storm’: Molecular mechanisms and therapeutic prospects. Trends Immunol. 2021, 42, 681–705. [Google Scholar] [CrossRef]
  131. Kalinina, O.; Golovkin, A.; Zaikova, E.; Aquino, A.; Bezrukikh, V.; Melnik, O.; Vasilieva, E.; Karonova, T.; Kudryavtsev, I.; Shlyakhto, E. Cytokine Storm Signature in Patients with Moderate and Severe COVID-19. Int. J. Mol. Sci. 2022, 23, 8879. [Google Scholar] [CrossRef]
  132. Sun, W.; Liu, S.; Huang, X.; Yuan, R.; Yu, J. Cytokine storms and pyroptosis are primarily responsible for the rapid death of mice infected with pseudorabies virus. R. Soc. Open Sci. 2021, 8, 210296. [Google Scholar] [CrossRef]
  133. Karki, R.; Sharma, B.R.; Tuladhar, S.; Williams, E.P.; Zalduondo, L.; Samir, P.; Zheng, M.; Sundaram, B.; Banoth, B.; Malireddi, R.K.S.; et al. Synergism of TNF-alpha and IFN-gamma Triggers Inflammatory Cell Death, Tissue Damage, and Mortality in SARS-CoV-2 Infection and Cytokine Shock Syndromes. Cell 2021, 184, 149–168.e17. [Google Scholar] [CrossRef]
  134. Pandian, N.; Kanneganti, T.D. PANoptosis: A Unique Innate Immune Inflammatory Cell Death Modality. J. Immunol. 2022, 209, 1625–1633. [Google Scholar] [CrossRef]
Figure 1. Schematic overview of impairment on IFN-I production targeting RIG-MAVS axis and IFN-induced antiviral responses by ASFV proteins in the latest reports. After ASFV infection, viral pI267L and increased lactate inhibit IFN-I production. Mechanistically, pI267L destroys the stabilized form of activated RIG-I and lactate to prevent the aggregation of MAVS. ASFV pI215L and pMGF360-9L inhibit IFN-I-induced antiviral response, while pMGF505-7R inhibit IFN-II-induced antiviral response. Mechanistically, pI267L mediates the degradation of IRF9 and STAT2, pMGF360-9L mediates the degradation of STAT1 and STAT2, and pMGF505-7R mediates the degradation of JAK1 and JAK2.
Figure 1. Schematic overview of impairment on IFN-I production targeting RIG-MAVS axis and IFN-induced antiviral responses by ASFV proteins in the latest reports. After ASFV infection, viral pI267L and increased lactate inhibit IFN-I production. Mechanistically, pI267L destroys the stabilized form of activated RIG-I and lactate to prevent the aggregation of MAVS. ASFV pI215L and pMGF360-9L inhibit IFN-I-induced antiviral response, while pMGF505-7R inhibit IFN-II-induced antiviral response. Mechanistically, pI267L mediates the degradation of IRF9 and STAT2, pMGF360-9L mediates the degradation of STAT1 and STAT2, and pMGF505-7R mediates the degradation of JAK1 and JAK2.
Viruses 15 00574 g001
Figure 2. Schematic overview of inhibition of NLRP3 inflammasome activation and GSDMD-mediated pyroptosis by ASFV proteins in the latest reports. (A) ASFV pF317L, pH240R, pMGF505-7R, pD345L, pI215L, and pMGF360-12L inhibit the transcription of proinflammatory genes by targeting canonical NF-κB pathway. Mechanistically, pF317L inhibits the phosphorylation of IKKβ and the nuclear translocation of NF-κB, pH240R mediates the degradation of NEMO, pMGF505-7R and pI215L inhibit the nuclear translocation of NF-κB, pD345L directly interacts with IKKα and IKKβ, and pMGF360-12L competitively inhibits the interaction between p65 and importin-α. (B) ASFV pH240R and pMGF505-7R inhibit the assembly of inflammasome, and pS273R inhibits the GSDMD-mediated pyroptosis. Mechanistically, pH240R suppresses NLRP3 oligomerization, pMGF505-7R directly interacts with NLRP3, and S273R cleaves N-terminal of GSDMD.
Figure 2. Schematic overview of inhibition of NLRP3 inflammasome activation and GSDMD-mediated pyroptosis by ASFV proteins in the latest reports. (A) ASFV pF317L, pH240R, pMGF505-7R, pD345L, pI215L, and pMGF360-12L inhibit the transcription of proinflammatory genes by targeting canonical NF-κB pathway. Mechanistically, pF317L inhibits the phosphorylation of IKKβ and the nuclear translocation of NF-κB, pH240R mediates the degradation of NEMO, pMGF505-7R and pI215L inhibit the nuclear translocation of NF-κB, pD345L directly interacts with IKKα and IKKβ, and pMGF360-12L competitively inhibits the interaction between p65 and importin-α. (B) ASFV pH240R and pMGF505-7R inhibit the assembly of inflammasome, and pS273R inhibits the GSDMD-mediated pyroptosis. Mechanistically, pH240R suppresses NLRP3 oligomerization, pMGF505-7R directly interacts with NLRP3, and S273R cleaves N-terminal of GSDMD.
Viruses 15 00574 g002
Figure 3. Schematic overview of scientific hypotheses proposed in this review. (A) Other nucleic acid sensors may be involved in ASFV detection. Mechanistically, TLR9 may directly recognize the unmethylated CpG DNA in the ASFV genome somehow or by indirectly sensing the ASFV-induced mitophagy-mediated mtDNA release. IFI16 may recognize ASFV through binding to viral DNA or sensing ASFV-induced host DNA damage. (B) Synergism of multiple cytokines may facilitate the PANoptosis; in turn, PANoptosis may sustain the ASFV-induced cytokine storm status.
Figure 3. Schematic overview of scientific hypotheses proposed in this review. (A) Other nucleic acid sensors may be involved in ASFV detection. Mechanistically, TLR9 may directly recognize the unmethylated CpG DNA in the ASFV genome somehow or by indirectly sensing the ASFV-induced mitophagy-mediated mtDNA release. IFI16 may recognize ASFV through binding to viral DNA or sensing ASFV-induced host DNA damage. (B) Synergism of multiple cytokines may facilitate the PANoptosis; in turn, PANoptosis may sustain the ASFV-induced cytokine storm status.
Viruses 15 00574 g003
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

Yu, L.; Zhu, Z.; Deng, J.; Tian, K.; Li, X. Antagonisms of ASFV towards Host Defense Mechanisms: Knowledge Gaps in Viral Immune Evasion and Pathogenesis. Viruses 2023, 15, 574. https://doi.org/10.3390/v15020574

AMA Style

Yu L, Zhu Z, Deng J, Tian K, Li X. Antagonisms of ASFV towards Host Defense Mechanisms: Knowledge Gaps in Viral Immune Evasion and Pathogenesis. Viruses. 2023; 15(2):574. https://doi.org/10.3390/v15020574

Chicago/Turabian Style

Yu, Liangzheng, Zhenbang Zhu, Junhua Deng, Kegong Tian, and Xiangdong Li. 2023. "Antagonisms of ASFV towards Host Defense Mechanisms: Knowledge Gaps in Viral Immune Evasion and Pathogenesis" Viruses 15, no. 2: 574. https://doi.org/10.3390/v15020574

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