Next Article in Journal
Association between Maternal and Fetal Genetic Variants and Preeclampsia: Evidence from a Meta-Analysis
Previous Article in Journal
Skin Development and Disease: A Molecular Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

African Swine Fever Virus Immunosuppression and Virulence-Related Gene

1
China/WOAH Reference Laboratory for Classical Swine Fever, China Institute of Veterinary Drug Control, Beijing 100081, China
2
College of Veterinary Medicine, Nanjing Agricultural University, Nanjing 210095, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Curr. Issues Mol. Biol. 2024, 46(8), 8268-8281; https://doi.org/10.3390/cimb46080488
Submission received: 13 June 2024 / Revised: 23 July 2024 / Accepted: 30 July 2024 / Published: 31 July 2024
(This article belongs to the Special Issue Innate Immunity Responds to Virus Infection)

Abstract

:
African swine fever virus (ASFV), a highly contagious pathogen characterized by a complex structure and a variety of immunosuppression proteins, causes hemorrhagic, acute, and aggressive infectious disease that severely injures the pork products and industry. However, there is no effective vaccine or treatment. The main reasons are not only the complex mechanisms that lead to immunosuppression but also the unknown functions of various proteins. This review summarizes the interaction between ASFV and the host immune system, along with the involvement of virulence-related genes and proteins, as well as the corresponding molecular mechanism of immunosuppression of ASFV, encompassing pathways such as cGAS-STING, nuclear factor kappa–light-chain-enhancer of activated B cells (NF-κB), Janus Kinase (JAK) and JAK Signal Transducers and Activators of Transcription (STAT), apoptosis, and other modulation. The aim is to summarize the dynamic process during ASFV infection and entry into the host cell, provide a rational insight into development of a vaccine, and provide a better clear knowledge of how ASFV impacts the host.

1. Introduction

African swine fever virus (ASFV), the pathogeny of African swine fever (ASF), is a 200 nm diameter double-stranded DNA virus in the family Asfarviridae. The viral genome is a linear 170–190 kb and encodes 150–200 viral proteins, but understanding about these proteins is not exactly clear [1,2,3]. The virion exhibits icosahedral symmetry, composed of nucleoid, core shell, inner envelope, capsid, and outer envelope [4]. The capsid proteins account for approximately one third of the total protein of the virus. Surrounding the capsid is the inner envelope derived from the endoplasmic reticulum, which incorporates several significant immune-associated proteins [5,6].
ASFV induces severe clinical symptoms, including fever, redness, viremia, and diarrhea. According to these symptoms, ASF can be divided into three forms, acute, subacute, or chronic [7]. The acute form of ASFV is characterized by a rapid death during the appearance of clinical signs without heavy organ lesion [8]. Infected pigs present diarrhea with blood and organ hemorrhage. Pregnant sows may exhibit miscarriage and stillbirth [9]. The subacute infectious pigs can persist for even 70 days, showing the above symptoms similar to those of acute cases around 6–7 days post-infection; then, some recover health while others experience mild anorexia. Mortality in subacute infectious pigs ranges from 30 to 70% [10]. The chronic infection may last for 2–15 months. During this period, pigs develop fluctuating fever, anorexia, joint swelling, coughing, diarrhea, occasional vomiting, and skin necrosis. The clinical manifestations of ASF closely resemble those of classical swine fever and are also akin to those exhibited in other swine diseases [11]. ASFV-induced outbreaks result in severe adverse economic consequences in affected regions and pose a significant threat to global swine trade due to almost complete mortality and morbidity in susceptible populations [12].
The virus must experience six steps for successfully infecting host cells: adsorption, penetration, uncoating, replication, packaging, and shedding. Upon cellular entry, ASFV sheds its outer membrane and undergoes internalization within endosomes. Subsequent to induction by low pH conditions, fusion between the virus inner membrane and endosomal membrane occurs, leading to the release of the viral core into the cytoplasm [13]. The core is then transported around the nucleus via microtubules and employs its encoded enzymes and cellular factors for early mRNA transcription, translation, and genome replication [14]. The initiation ASFV genome DNA replication commences approximately 5–10 h post-infection of the host cell with the early gene expression. The replication occurs in two stages and its mechanism is similar to that of poxviruses. Initially, a brief replication stage takes place in the nucleus, followed by the synthesis of numerous DNA fragments within the virus factory (VF) located in the perinuclear region. Subsequently, the virus particles undergo assembly in the VF [15]. After replication, many intermediate and late genes express proteins that play a significant role in virus particle assembly and shedding, such as pC962R, pG1211R, pE301R, pE165R, F1055L, and pP1192R, but their exact functions remain barely known [16].
Macrophages are one of the antigen-presenting cells (APCs) and the target cells for ASFV infection and the receptor-mediated endocytosis and clathrin-dependent endocytic pathway are important pathways for ASFV to invade macrophages [17]. While the innate immune response and adaptive immune response are induced, the natural killer cells and dendritic cells are capable of generating significant amounts of interferon to combat ASFV infection, following 8 days of ASFV infection, and then the antibody levels progressively rise, indicating the presence of a humoral immune response in the host. Some studies have reported that specific antibodies are sufficient to protect pigs from lethal challenges of ASFV, but our knowledge is limited for those antibodies [18,19]. Furthermore, cellular immunity, exemplified by CD8a (+) T cells, is crucial in the antiviral defense against ASFV infection. CD4 (+) CD8 (+) double-positive (DP) T cells have the capacity to release perforin and granzyme, which could also contribute to the resistance against ASFV infection [20]. Despite the numerous mechanisms the host immune system possesses to inhibit ASFV invasion, the efficacy is unsatisfactory, as the virus always can easily escape from capture by the host immune system and even suppress it. This review highlights key virulence-associated genes and proteins of ASFV, while also providing an overview of the documented mechanisms for how ASFV influences the host’s innate and adaptive immune responses through regulating the various upstream or downstream signaling factors of signaling pathways and other regulatory processes. It is these intricate mechanisms that elucidate the considerable challenge in developing an effective vaccine.

2. ASFV Virulence-Associated Genes and Key Proteins

The ASFV genome consists of the central conserved region, approximately 125 kb in length, flanked by the variable left region (38–48 kb) and right region (13–22 kb). These variable regions contain five unique multigene families (MGFs), MGF100, MGF110, MGF300, MGF360, and MGF530/505, that exhibit substantial diversity among strains from different sources. Their variability is linked to viral virulence, antigenic variation, and evasion of host immune responses [6] and, especially, the MGF360 and MGF505 are responsible for enhancing infected cell survival and influencing the immune response [21]. Currently, we know that ASFV encodes approximately 50 proteins, including structural proteins of the virus, proteins involved in viral particle assembly, enzymes, and factors required for nucleic acid metabolism, replication, repair, transcription, and processing [22]. The immunosuppression proteins encoded by ASFV are mainly categorized by function as follows: (1) regulation of the host cell protein expression system and transcription, such as DP71L and A238L, blocking protein expression of host cells; (2) inhibition of the type I interferon signaling pathway, including multigene family proteins MGF360, MGF505/530, DP96R, and I329L, suppressing the induction of type I interferon; (3) regulation of programmed cell death, including p54, A179L, A224L, and EP153R, hindering apoptosis in the early stage of infection; and (4) other immunosuppressive proteins, such as CD2v and L83L, which are the obstacles for lymphocyte proliferation and antiviral effects of IL-1β [23,24]. Additionally, there are still some proteins with functions that are not yet clear (Table 1).

3. ASFV Interaction with Host Immune System

The host immune system recognizes and targets ASFV for elimination, so interaction between ASFV and the host immune system is a crucial aspect in understanding the pathogenesis of ASF. The co-ordinated efforts of innate and adaptive immune responses are essential for providing protection against a wide range of pathogens. Type I interferons play a key role in priming immune cells for antiviral defense, while effector cells like macrophages and dendritic cells serve as frontline defenders in detecting and combating invading pathogens. This collaborative action ensures a robust and effective immune response to safeguard the body from infections. The analysis of the macrophage transcriptome identified a suppression of immune regulation involving 54 cytokine genes and reduced expression of 13 cytokines [37]. However, ASFV has developed sophisticated mechanisms to evade and manipulate the host immune response by modulating the above cytokines and other unknown targets, allowing it to establish infection and cause disease in pigs.

3.1. ASFV Regulates Innate Immunity

Innate immunity is the body’s first line of defense against pathogens and other foreign invaders. It is a nonspecific immune response that provides immediate protection upon exposure to a threat. APCs detect pathogen-associated molecular patterns (PAMP) through a series of pattern recognition receptors (PRR), including Toll-like receptors, nucleotide oligomerization domain (NOD)-like receptors, and C-type lectin-like receptors, and produce cytokines and chemokines that help in clearing pathogens through phagocytosis. In order to avoid this progress, a multitude of proteins encoded by ASFV genome serve as pivotal regulators in these evasion strategies. For example, E120R, a protein encoded by ASFV, can combine with IRF3 and inhibit the activity of IRF3, resulting in the suppression of IFN-β expression [38]; QP383R, an inhibitor of activation of type I interferons, suppresses cGAS dimerization by replacing the dsDNA to interact with cGAS, thus leading to a reduction in cGAMP production [39]; DP96R, one of the virulence-related genes, can inhibit the activation of the promotor of TBK1 and reduce the level of phosphorylation of TBK1, resulting in the inhibition of Type I interferons and TBK1-induced antiviral effect [40]; MGF360-13L has been identified as the pathogenicity gene that antagonizes the production of Type I interferons and inhibits its mediated signal transduction [41]; I267L, a highly conserved protein among different ASFV strains, effectively impairs Riplet-mediated RIG-I activation by interacting with Riplet and hindering its association with RIG-I. This disruption interferes with Riplet-mediated K63-linked polyubiquitination of RIG-I, a process where ubiquitin molecules are attached to specific proteins after they have been synthesized, and the recruitment of MAVS, the downstream adaptor protein. By inhibiting the activation of the IFN-β promoter triggered by RIG-I and Riplet, I267L hampers the immune response. Moreover, I267L negates Riplet-induced polyubiquitination and the activation of RIG-I, potentially aiding ASFV in evading the host’s innate immunity and enhancing virulence [42]. A part of the specific mechanism will be detailed in the following sections.

3.2. Modulation with Adaptive Immunity

Upon encountering a pathogen, antigen-presenting cells, like dendritic cells, capture and present antigens to T cells through major histocompatibility complex (MHC) molecules, initiating an adaptive immune response. T cells differentiate into effector T cells that can directly kill infected cells or help activate other immune cells. B cells produce antibodies that bind to and neutralize pathogens, marking them for destruction by other immune cells. Nevertheless, certain ASFV strains inhibit MHC-II-mediated antigen processing, activation of M1 macrophages, or macrophage autophagy and apoptosis pathway [40]. Some studies demonstrated that some ASFV strains interfere with the normal expression of MHC-I and ASFV infection also results in the reduced surface expression of CD14 and CD16 on macrophages, potentially compromising their antimicrobial and antiviral capabilities. This finding implies that ASFV may develop strategies to interfere with macrophage function as an approach to evade the host’s immune response [3]. It has been reported that the infected host shows decreased levels of various CD T cells that are essential in mounting the cellular immune defense against ASFV. Furthermore, a reduction in the total count of CD4 (+) T cells in the peripheral blood of infected pigs was detected [43]. Due to a plunge in these effector cells, the purpose of immune evasion is achieved by ASFV. For example, the A179L protein can inhibit sensitized T-cell-induced cell apoptosis by binding to the BH3 domain of the Bcl-2 protein family through a conserved ligand [36]. The ASFV protein DP71L inhibits transcription factor 4 (ATF4) and its downstream target CHOP, thereby suppressing sensitized T-cell-induced cell apoptosis [44]. It recruits PP1 to dephosphorylate eIF2 to promote ASFV replication within the cells [45].

4. ASFV Immunosuppression Mechanisms

ASFV invasion is characterized by immunosuppression; it regulates various signaling pathways to promote viral replication, such as cGAS-STING, nuclear factor kappa–light-chain-enhancer of activated B cells (NF-κB), the Janus Kinase (JAK) and JAK Signal Transducers and Activators of Transcription (STAT), ubiquitination, and apoptosis. These signaling pathways are necessary for the host to bridge the immune system and pathogens so that the host can resist these pathogens. They involve cascading reactions containing numerous regulatory factors, and nodes are significantly influenced by ASFV via phosphorylation or dephosphorylation, degradation or undegradation, and other mechanisms to inhibit the normal function of the host immune system, thus inducing immunosuppression (Figure 1). The specific details are as follows.

4.1. Cyclic GMP-AMP Synthase Signaling Pathway

The cGAS/STING signaling pathway plays a crucial role as a key immune regulator in response to pathogens [46]. Extensive research has focused on cGAS’s ability to detect cytosolic or viral DNA. Monomeric cGAS is normally found in the cytoplasm of healthy cells where it cannot bind to DNA or function as an enzyme [47]. However, once cGAS binds to the pathogen’s dsDNA, it triggers the production of cyclic GMP-AMP (cGAMP). This cGAMP molecule then acts as a second messenger by activating the STING adaptor protein, which recruits a series of signaling events involving TBK1 and IRF3 [48], ultimately resulting in their translocation to the nucleus [49]. Activation of this pathway induces the production of type I interferons and proinflammatory cytokines, promoting an effective antiviral immune response [38].
A part of the proteins of ASFV are able to inhibit the expression of type I interferons and antagonize their antiviral effect, which allows ASFV to induce immunosuppression. Overexpression of QP383R inhibited the activation of type I interferons triggered by dsDNA and cGAS/STING pathways. QP383R was shown to directly interact with cGAS, interfering with its DNA binding ability and dimerization process. This interaction results in the inhibition of cGAS enzymatic functions, ultimately reducing the production of cGAMP, a critical signaling molecule involved in the host immune response against viral infections [50]. ASFV DP96R, a kind of conserved early expressed protein, suppresses the activation of the promoter of IFN-β and ISRE, mediated by GAS-STING and phosphorylation of TBK1 [51]. In the MGF360 multi-gene family, A276R has been identified to suppress the production and induction of type I interferon through the IRF3 signaling pathway [52]. Similarly, A528R from MGF505 has been shown to inhibit the type I interferon signaling pathway [53]. Furthermore, deletion of MGF360-18R (DP148R) results in reduced viral virulence without affecting viral replication in porcine alveolar macrophages (PAM) [54]. MGF360-13L inhibits the activation of the cGAS-STING-mediated IFN signaling pathway. It disrupts the assembly of the STING–TBK1–IRF3 complex, leading to the inhibition of TBK1 and IRF3 phosphorylation and dimerization. Ultimately, this restriction limits the nuclear translocation of IRF3 [41].
Gene deletion is a promising approach for developing effective attenuated vaccines, especially after elucidating the specific mechanism of a gene. In order to explore how various genes influence type I IFN via the signaling pathway in vivo, the MGF505-2R gene, an ASFV immunomodulator gene involved in controlling the innate immune response through regulation of the cGAS/STING pathway, was deleted from wild-type ASFV. The result showed that the recombinant MGF505-2R deletion virus demonstrated characteristics of a potential live attenuated vaccine (LAV), providing approximately 60–70% protection, and may serve as a strong candidate for developing enhanced LAVs, because it lost its ability to interact with STING and inhibit the phosphorylation of TBK1.

4.2. NF-κB Signaling Pathway

The NF-κB pathway is a critical signaling pathway involved in the regulation of immune responses, inflammation, cell proliferation, and survival [55]. Upon activation by corresponding factors, the IκB proteins are phosphorylated by the IκB kinase (IKK) complex, releasing NF-κB dimers (typically p50 and RelA or p65) into the nucleus to regulate immune reaction [56]. The ASFV gene-encoded proteins have the capacity of influencing this signaling pathway [57].
Recently, studies have demonstrated that MGF360-12L, belonging to the member of ASFV MGF that encodes the most effective antagonist for host immunity defense, can inhibit the production of type I interferons by interacting with KPNA2, KPNA3, and KPNA4, thereby preventing the interaction between importinα and the NF-κB signaling pathway [58]. pMGF505-7R, also encoded by ASFV MGF, inhibits NF-κB activity by binding to IKK, blocking its translocation to the nucleus [59]. This inhibition prevents the phosphorylation of NF-κB p65 and IκB proteins, subsequently leading to the suppression of IL-1β production [60]. The homolog IkBα A238L shares a similar Ankyrin repeat sequence with IkBα. The modified form of A238L interacts directly with the subunit of NF-κB, forming the A238L–p65 complex in the cytoplasm. This complex inhibits the entry of NF-κB p65 into the nucleus and its binding to DNA, thereby resulting in the inhibition of NF-κB pathway [61]. UBCv1, which is the only known conjugating enzyme encoded by the ASFV, has been recognized as a novel inhibitor of NF-κB activation that impedes the movement of p65 into the nucleus [62]. F317L is a protein of ASFV that consists of 317 amino acids and can combine with IκB kinase β (IKKβ) and prevent its own phosphorylating so that it limits the level of IκB ubiquitination and phosphorylation, thereby increasing the IκB stability and suppressing the host immune system, but not all proteins express inhibitive function; K205R and A224L, two exceptions, have been discovered to contribute P65 to translocate into the nucleus, resulting in the activation of NF-κB [63,64].
The activated NF-κB signaling pathway initiates the release of immune-associated molecules, including IFNs and proinflammatory cytokines. Previous studies have elucidated the specific inhibitory mechanism of MGF300-4L on this signaling pathway. Moreover, upon deletion of this gene to investigate its function in vivo, it was observed that the recombinant virus induced a greater proinflammatory response in pigs when compared to wild-type ASFV. Additionally, they also observed that it was capable of eliciting a faster humoral and cellular immune response

4.3. JAK/STAT Signaling Pathway

JAK/STAT signaling pathway is a crucial intracellular signaling cascade that plays a significant role in immune response. Once the activated kinases, JAK and tyrosine kinase2 (TYK2), phosphorylate STAT1 and STAT2, the phosphorylated proteins subsequently form a complex with IFN Regulatory Factor 9 (IRF9), creating the Interferon-Stimulated Gene Factor (ISGF) 3. Finally, the ISGF3 induces a series of signaling activation to initiate the host’s antiviral defense mechanisms [65].
Protein degradation serves as a key mechanism for modulating protein function within biological processes, and the primary strategies employed for protein degradation and regulation include the ubiquitin–proteasome system, autophagy-lysosomal pathway, and apoptosis pathways. MGF-360-9L, a virulence factor of the ASFV, binds to and degrades STAT1 and STAT2 through apoptosis and ubiquitin-proteasome pathways, leading to the suppression of IFN-β signaling [66]. MGF360-10L specifically interacts with JAK1, resulting in a notable rise in its ubiquitination levels in both HEK293T cells and PAMs. This elevates ubiquitination then triggers the degradation of JAK1 through the proteasome-mediated pathway. Subsequently, through RNA-seq analysis, MGF-360-10L was identified to facilitate the ubiquitination of JAK1 by recruiting HERC5 [67]. The CD2v contributes to STAT3 transcription and translocation into the nucleus and prevents the apoptosis of JAK2-STAT3 pathway to facilitate viral replication [25]. In another experiment, researchers found that overexpression of MGF360-12L was shown to reduce the expression of IRF9 while not impacting the levels of STAT1 and STAT2. Their analysis of truncated mutations revealed that the C-terminal region of MGF360-12L was responsible for suppressing immune responses related to TBK1 and IKKβ. Furthermore, complete sequences of MGF360-12L might be necessary for its inhibitory effects on IRF3-5D and IRF9 [68].
Due to the importance of JAK/STAT signaling pathway in vivo and the antagonism of MGF360-9L on this signaling pathway, some researchers have found that deletion of MGF-360-10L from wild-type ASFV infection causes milder pathological injury in pigs and weakens the virulence compared to ASFV-WT. Thus, the recombinant virus attenuates the virulence of ASFV, providing guidance for the development of safer and more reliable attenuated recombinant ASFV vaccines [66].

4.4. Inhibition of Apoptosis

Apoptosis, also known as programmed cell death, is a fundamental biological process in multicellular organisms. It plays a critical role in maintaining tissue homeostasis by eliminating unwanted, damaged, or infected cells without inducing an inflammatory response [69]. Apoptosis is tightly regulated by a balance of proapoptotic and antiapoptotic factors [70].
Proteins of the Bcl-2 family regulate mitochondrial outer membrane permeabilization and play crucial roles in cellular apoptosis pathways, where they can be categorized into two main groups: antiapoptotic members such as Bcl-2, Bcl-XL, Bcl-W, Mcl-1, and CED9; and proapoptotic members, including Bax, Bak, Bcl-XS, Bad, Bik, and Bid [52] (Figure 2). The A179L protein, with a molecular weight of approximately 18 kDa and consisting of 179 amino acids, belongs to the Bcl-2 family and possesses conserved structural domains BH1, BH2, BH3, and BH4, while lacking the corresponding transmembrane domain [71]. The A179L is expressed in both early and late stages of viral infection of ASFV and its sequence shows high conservation across different ASFV strains. Previous studies have shown that the A179L can suppress cell apoptosis induced by p68 in HeLa, as well as apoptosis induced by large molecules in BSC-40 cells, and interacts with proapoptotic proteins such as Bid p13 and p15, which express only the BH3 domain, inhibiting their activity. The A179L also interacts with other members of the proapoptotic protein family, including Bad, Bmf, Bik, and Bim, forming heterodimers that prevent cells from entering the apoptotic pathway [65]. Additionally, it regulates cellular autophagy by interacting with Beclin-1, modulating starvation-induced autophagosome formation [72].

4.5. Other Modulations

Given that the inflammatory responses serve as a potent host antiviral defense mechanism in combating viral infection, ASFV has employed evolved mechanisms to inhibit these defense pathways. The host regulates inflammatory responses by controlling the production of proinflammatory factors, inflammatory mediators, and proteins generated in the downstream cascades of the NF-κB. The NLRP3 inflammasome is a cytoplasmic multiprotein complex that assembles and triggers inflammatory immune responses upon exposure to pathogens and tissue damage; predominantly found in macrophages, it comprises NLRP3 protein, ASC adaptor protein, and pro-Caspase-1. Activation of NLRP3 inflammasome-induced IL-1β and pyroptosis requires two priming signals: (1) NF-κB activation, stimulating factors like LPS or viruses to activate NF-κB, thereby inducing upregulation of proinflammatory cytokines and transcriptional levels of NLRP3, pro-IL-1β, etc.; (2) NLRP3 activation, where external pathogens induce assembly of the NLRP3 inflammasome complex and trigger self-cleavage activation of pro-Caspase-1, ultimately leading to maturation of IL-1β within cells and formation of pores on the cell membrane by N-GSDMD, resulting in pyroptosis [36]. ASFV-C84L significantly promotes NLRP3-mediated Caspase-1 activation and expression of proinflammatory cytokines, thus aiding in initiating the inflammatory response.
Ubiquitination is a process where ubiquitin molecules are attached to specific proteins after they have been synthesized [44]. This modification of the ubiquitination serves various purposes, such as directing proteins to different parts within the cell, altering their functions, promoting or inhibiting interactions with other proteins, and labeling them for degradation through the proteasome [73]. ASFV pI215L triggers the interaction between RNF138 and RNF128. This interaction results in an increase in RNF138 levels, causing it to break down RNF128. As a result, the K63-linked ubiquitination of tank-binding kinase 1 (TBK1) is inhibited. Thereby, the downstream of TBK1 is negatively regulated [74]. The immunosuppression protein S273R of ASFV interacts with STAT2 and recruits the E3 ubiquitin ligase DCS1, leading to K48-linked polyubiquitin chains at K55 of STAT2, subsequently prompting the degradation of STAT2 via the proteasome pathway [75]. The ubiquitination serves as a modification mechanism utilized by various viruses, which influences the normal functions of signaling pathways by interacting with specific proteins.
Autophagy is a highly conserved cellular process that involves the degradation and recycling of damaged organelles, misfolded proteins, and other cellular components. The process of autophagy is tightly regulated and involves the formation of double-membrane vesicles called autophagosomes that engulf cytoplasmic cargo targeted for degradation. These autophagosomes then fuse with lysosomes to form autolysosomes, where the cargo is broken down by lysosomal enzymes. It plays a critical role in various physiological processes. The early expressed MGF505-7R exhibits diverse inhibitory actions on STIN-dependent antiviral reaction. It is capable of facilitating the proteasome-mediated degradation of TBK1, as well as the degradation of caspase, IRF7 via autophagosome pathways, and STING through autophagosome-dependent mechanisms. Studies have identified that MGF505-7R enhances the levels of the autophagy-related protein ULK1, leading to the degradation of STING. Additionally, it suppresses IFN-γ-induced signaling pathways mediated by JAK1 and JAK2 [76,77,78].

5. Conclusions and Perspective

The immune modulation of ASFV on the host organism consists of manipulating the host cell’s innate immune responses through various encoded proteins. Additionally, ASFV impacts lymphocyte development by hindering antigen presentation and modulating chemokine expression to prevent adaptive immunity activation. Indeed, various structural proteins encoded by ASFV play crucial roles in virus attachment, entry, and replication. In particular, MGF proteins are essential at multiple stages of virus infection in host cells, including transcription, translation, virulence, and immune evasion. For example, MGF-360-10L was identified to facilitate the ubiquitination of JAK1 by recruiting HERC5. MGF-360-9L binds to and degrades STAT1 and STAT2, leading to the suppression of IFN-β signaling and protein degradation. pMGF505-7R inhibits NF-κB activity by binding to IKK, blocking its translocation to the nucleus. MGF360-13L inhibits the activation of the cGAS-STING-mediated IFN signaling pathway. However, our current understanding of the functions and characteristics of ASFV proteins are not fully understood, and the complex and diverse immune regulation mechanisms after invading the host remain unclear. Therefore, a significant effort is required to gain a clear understanding of the specific functions of over 150 nonstructural proteins and structural proteins.
In conclusion, the intricate network of pathways and distinct stages involved in immune modulation poses challenges for the development of effective vaccines and treatments, particularly due to issues regarding safety and stability. These obstacles stem from the multifaceted nature of the process, which includes limited comprehension of virus–host interactions and the complex structure of the ASFV itself. While immunization with high-dose inactivated ASFV or γ-irradiated ASFV is considered safe, the immune protection efficacy remains suboptimal. Meanwhile, the nucleic acid vaccines, subunit vaccines, and viral vector vaccines only offer partial immunity. Therefore, studies on the functions of ASFV-encoded proteins, particularly the proteins mediating ASFV immune escape and immunosuppression, are very important for the vaccine development and design. Attenuated ASFV strains derived from either naturally weakened variants or genetically engineered constructs can confer full protection upon immunized pigs. However, viral presence persists in blood and certain organs, leading to intermittent shedding and posing a risk of horizontal transmission (Table 2).
Through modern virology and immunology techniques, identifying the ASFV immunomodulatory genes and subsequently discovering and characterizing essential virulence genes of ASFV becomes imperative. This will elucidate their roles in virus infection, immune evasion, and pathogenesis. Utilizing techniques such as CRISPR, genomics, transcriptomics, proteomics, and metabolomics can help unravel the specific molecular mechanisms underlying the ASFV lifecycle or immune modulation. This review summarized the ASFV virulence-associated genes and proteins while providing an overview of the current research on how ASFV suppresses the innate and adaptive immune responses in the host by its different proteins. It points out areas of our knowledge of how ASFV impacts the cellular and humoral immune systems of the host. The process of immunosuppression of ASFV is characterized by diversity and complexity and more mechanisms underlying the interaction between ASFV and the host await discovery. By combining classical molecular biology research methods, critical replication or immune suppression mechanisms of ASFV can be further identified. This could provide theoretic support for the discovery of suitable drugs or the design of stable and safe vaccines to make ASF treatable, controllable, and preventable.

Funding

This research was funded by [National Key Research and Development Program of China] grant number [2021YFD1800105] and The APC was funded by [National Key Research and Development Program of China].

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ASFVAfrican swine fever virus
ASFAfrican swine fever
VFvirus factory
APCsAntigen-presenting cells
MGFmultigene family
TBKTANK-binding kinase
IRFIFN regulatory factor
PAMPPathogen-associated molecular patterns
PRRPattern recognition receptors
NODNucleotide oligomerization domain
MHCmajor histocompatibility complex
JAKJanus Kinase
STATSignal Transducers and Activators of Transcription
cGAMPcyclic GMP-AMP
ISGFIFN-stimulated gene factor
ISREIFN-stimulated response element
ISGsIFN-stimulated genes
IKKIκB kinase
IKKβIκB kinase β
TYK2tyrosine kinase2
ISGFInterferon-Stimulated Gene Factor

References

  1. Alejo, A.; Matamoros, T.; Guerra, M.; Andrés, G.A. Proteomic Atlas of the African Swine Fever Virus Particle. J. Virol. 2018, 92, e01293-18. [Google Scholar] [CrossRef] [PubMed]
  2. Revilla, Y.; Perez-Nunez, D.; Richt, J.A. African Swine Fever Virus Biology and Vaccine Approaches. Adv. Virus Res. 2018, 100, 41–74. [Google Scholar] [PubMed]
  3. Karger, A.; Pérez-Núñez, D.; Urquiza, J.; Hinojar, P.; Alonso, C.; Freitas, F.B.; Revilla, Y.; Le Potier, M.-F.; Montoya, M. An Update on African Swine Fever Virology. Viruses 2019, 11, 864. [Google Scholar] [CrossRef] [PubMed]
  4. Li, H.; Liu, Q.; Shao, L.; Xiang, Y. Structural Insights into the Assembly of the African Swine Fever Virus Inner Capsid. J. Virol. 2023, 97, e26823. [Google Scholar] [CrossRef]
  5. Salas, M.L.; Andres, G. African swine fever virus morphogenesis. Virus Res. 2013, 173, 29–41. [Google Scholar] [CrossRef]
  6. Dixon, L.K.; Chapman, D.A.; Netherton, C.L.; Upton, C. African swine fever virus replication and genomics. Virus Res. 2013, 173, 3–14. [Google Scholar] [CrossRef] [PubMed]
  7. Gabriel, C.; Blome, S.; Malogolovkin, A.; Parilov, S.; Kolbasov, D.; Teifke, J.P.; Beer, M. Characterization of African swine fever virus Caucasus isolate in European wild boars. Emerg. Infect. Dis. 2011, 17, 2342–2345. [Google Scholar] [CrossRef]
  8. Sánchez-Vizcaíno, J.M.; Mur, L.; Gomez-Villamandos, J.C.; Carrasco, L. An update on the epidemiology and pathology of African swine fever. J. Comp. Pathol. 2015, 152, 9–21. [Google Scholar] [CrossRef]
  9. Guinat, C.; Reis, A.L.; Netherton, C.L.; Goatley, L.; Pfeiffer, D.U.; Dixon, L. Dynamics of African swine fever virus shedding and excretion in domestic pigs infected by intramuscular inoculation and contact transmission. Vet. Res. 2014, 45, 93. [Google Scholar] [CrossRef]
  10. 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]
  11. 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] [PubMed]
  12. Li, Z.; Chen, W.; Qiu, Z.; Li, Y.; Fan, J.; Wu, K.; Chen, J. African Swine Fever Virus: A Review. Life 2022, 12, 1255. [Google Scholar] [CrossRef]
  13. Netherton, C.L.; Shimmon, G.L.; Hui, J.Y.; Connell, S.; Reis, A.L. African Swine Fever Virus Host-Pathogen Interactions. Subcell Biochem. 2023, 106, 283–331. [Google Scholar]
  14. Yang, S.; Miao, C.; Liu, W.; Zhang, G.; Shao, J.; Chang, H. Structure and function of African swine fever virus proteins: Current understanding. Front. Microbiol. 2023, 14, 1043129. [Google Scholar] [CrossRef]
  15. Frost, L.; Batten, C. African Swine Fever Virus Plaque Assay and Disinfectant Testing. Methods Mol. Biol. 2022, 2503, 187–194. [Google Scholar] [PubMed]
  16. Wang, Y.; Kang, W.; Yang, W.; Zhang, J.; Li, D.; Zheng, H. Structure of African Swine Fever Virus and Associated Molecular Mechanisms Underlying Infection and Immunosuppression: A Review. Front. Immunol. 2021, 12, 715582. [Google Scholar] [CrossRef] [PubMed]
  17. Forth, J.H.; Calvelage, S.; Fischer, M.; Hellert, J.; Sehl-Ewert, J.; Roszyk, H.; Deutschmann, P.; Reichold, A.; Lange, M.; Thulke, H.-H.; et al. African swine fever virus—Variants on the rise. Emerg. Microbes Infect. 2023, 12, 2146537. [Google Scholar] [CrossRef]
  18. Dixon, L.; Islam, M.; Nash, R.; Reis, A. African swine fever virus evasion of host defences. Virus Res. 2019, 266, 25–33. [Google Scholar] [CrossRef] [PubMed]
  19. 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]
  20. Cabezón, O.; Muñoz-González, S.; Colom-Cadena, A.; Pérez-Simó, M.; Rosell, R.; Lavín, S.; Marco, I.; Fraile, L.; de la Riva, P.M.; Rodríguez, F.; et al. African swine fever virus infection in Classical swine fever subclinically infected wild boars. BMC Vet. Res. 2017, 13, 227. [Google Scholar] [CrossRef]
  21. O'Donnell, V.; Holinka, L.G.; Sanford, B.; Krug, P.W.; Carlson, J.; Pacheco, J.M.; Reese, B.; Risatti, G.R.; Gladue, D.P.; Borca, M.V. African swine fever virus Georgia isolate harboring deletions of 9GL and MGF360/505 genes is highly attenuated in swine but does not confer protection against parental virus challenge. Virus Res. 2016, 221, 8–14. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, F.; Zhang, H.; Hou, L.; Yang, C.; Wen, Y. Advance of African swine fever virus in recent years. Res. Vet. Sci. 2021, 136, 535–539. [Google Scholar] [CrossRef]
  23. Neilan, J.G.; Zsak, L.; Lu, Z.; Kutish, G.F.; Afonso, C.L.; Rock, D.L. Novel swine virulence determinant in the left variable region of the African swine fever virus genome. J. Virol. 2002, 76, 3095–3104. [Google Scholar] [CrossRef] [PubMed]
  24. Miao, C.; Yang, S.; Shao, J.; Zhou, G.; Ma, Y.; Wen, S.; Hou, Z.; Peng, D.; Guo, H.; Liu, W.; et al. Identification of p72 epitopes of African swine fever virus and preliminary application. Front. Microbiol. 2023, 14, 1126794. [Google Scholar] [CrossRef]
  25. Gao, Q.; Yang, Y.; Luo, Y.; Chen, X.; Gong, T.; Wu, D.; Feng, Y.; Zheng, X.; Wang, H.; Zhang, G.; et al. African Swine Fever Virus Envelope Glycoprotein CD2v Interacts with Host CSF2RA to Regulate the JAK2-STAT3 Pathway and Inhibit Apoptosis to Facilitate Virus Replication. J. Virol. 2023, 97, e188922. [Google Scholar] [CrossRef]
  26. Hu, B.; Zhong, G.; Ding, S.; Xu, K.; Peng, X.; Dong, W.; Zhou, J. African swine fever virus protein p17 promotes mitophagy by facilitating the interaction of SQSTM1 with TOMM70. Virulence 2023, 14, 2232707. [Google Scholar] [CrossRef]
  27. 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]
  28. Yang, K.; Huang, Q.; Wang, R.; Zeng, Y.; Cheng, M.; Xue, Y.; Shi, C.; Ye, L.; Yang, W.; Jiang, Y.; et al. Corrigendum to “African swine fever virus MGF505-11R inhibits type I interferon production by negatively regulating the cGAS-STING-mediated signaling pathway” [Vet. Microbiol. 263, 2021, 109265]. Vet. Microbiol. 2023, 285, 109723. [Google Scholar] [CrossRef]
  29. 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] [PubMed]
  30. 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]
  31. Petrovan, V.; Rathakrishnan, A.; Islam, M.; Goatley, L.C.; Moffat, K.; Sanchez-Cordon, P.J.; Reis, A.L.; Dixon, L.K. Role of African Swine Fever Virus Proteins EP153R and EP402R in Reducing Viral Persistence in Blood and Virulence in Pigs Infected with BeninDeltaDP148R. J. Virol. 2022, 96, e134021. [Google Scholar] [CrossRef] [PubMed]
  32. Qi, X.; Feng, T.; Ma, Z.; Zheng, L.; Liu, H.; Shi, Z.; Shen, C.; Li, P.; Wu, P.; Ru, Y.; et al. Deletion of DP148R, DP71L, and DP96R Attenuates African Swine Fever Virus, and the Mutant Strain Confers Complete Protection against Homologous Challenges in Pigs. J. Virol. 2023, 97, e24723. [Google Scholar] [CrossRef] [PubMed]
  33. Li, T.; Zhao, G.; Zhang, T.; Zhang, Z.; Chen, X.; Song, J.; Wang, X.; Li, J.; Huang, L.; Wen, L.; et al. African Swine Fever Virus pE199L Induces Mitochondrial-Dependent Apoptosis. Viruses 2021, 13, 2240. [Google Scholar] [CrossRef] [PubMed]
  34. 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, e195721. [Google Scholar] [CrossRef]
  35. 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, e94422. [Google Scholar] [CrossRef] [PubMed]
  36. Shi, J.; Liu, W.; Zhang, M.; Sun, J.; Xu, X. The A179L Gene of African Swine Fever Virus Suppresses Virus-Induced Apoptosis but Enhances Necroptosis. Viruses 2021, 13, 2490. [Google Scholar] [CrossRef] [PubMed]
  37. Niu, S.; Guo, Y.; Wang, X.; Wang, Z.; Sun, L.; Dai, H.; Peng, G. Innate immune escape and adaptive immune evasion of African swine fever virus: A review. Virology 2023, 587, 109878. [Google Scholar] [CrossRef] [PubMed]
  38. 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, e82421. [Google Scholar] [CrossRef] [PubMed]
  39. 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]
  40. 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, e223955. [Google Scholar] [CrossRef]
  41. Luo, J.; Cheng, M.; Duan, Y.; Xing, X.; Lu, M.; Sun, Y.; Shi, C.; Wang, J.; Lu, Y.; Li, X.; et al. African swine fever virus encoded protein MGF360-13L inhibits cGAS-STING-mediated IFN-I signaling pathway. Gene 2023, 874, 147490. [Google Scholar] [CrossRef]
  42. 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]
  43. Hühr, J.; Schäfer, A.; Schwaiger, T.; Zani, L.; Sehl, J.; Mettenleiter, T.C.; Blome, S.; Blohm, U. Impaired T-cell responses in domestic pigs and wild boar upon infection with a highly virulent African swine fever virus strain. Transbound. Emerg. Dis 2020, 67, 3016–3032. [Google Scholar] [CrossRef]
  44. Banjara, S.; Caria, S.; Dixon, L.K.; Hinds, M.G.; Kvansakul, M. Structural Insight into African Swine Fever Virus A179L-Mediated Inhibition of Apoptosis. J. Virol. 2017, 91, 10–1128. [Google Scholar] [CrossRef]
  45. Sun, W.; Zhang, H.; Fan, W.; He, L.; Chen, T.; Zhou, X.; Qi, Y.; Sun, L.; Hu, R.; Luo, T.; et al. Evaluation of Cellular Immunity with ASFV Infection by Swine Leukocyte Antigen (SLA)-Peptide Tetramers. Viruses 2021, 13, 2264. [Google Scholar] [CrossRef]
  46. 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]
  47. Luo, W.W.; Li, S.; Li, C.; Lian, H.; Yang, Q.; Zhong, B.; Shu, H.-B. iRhom2 is essential for innate immunity to DNA viruses by mediating trafficking and stability of the adaptor STING. Nat. Immunol. 2016, 17, 1057–1066. [Google Scholar] [CrossRef]
  48. Hopfner, K.P.; Hornung, V. Molecular mechanisms and cellular functions of cGAS-STING signalling. Nat. Rev. Mol. Cell Biol. 2020, 21, 501–521. [Google Scholar] [CrossRef]
  49. Dobbs, N.; Burnaevskiy, N.; Chen, D.; Gonugunta, V.K.; Alto, N.M.; Yan, N. STING Activation by Translocation from the ER Is Associated with Infection and Autoinflammatory Disease. Cell Host Microbe 2015, 18, 157–168. [Google Scholar] [CrossRef]
  50. Song, J.; Li, K.; Li, T.; Zhao, G.; Zhou, S.; Li, H.; Li, J.; Weng, C. Screening of PRRSV- and ASFV-encoded proteins involved in the inflammatory response using a porcine iGLuc reporter. J. Virol. Methods 2020, 285, 113958. [Google Scholar] [CrossRef]
  51. 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]
  52. Gallardo, C.; Sanchez, E.G.; Perez-Nunez, D.; Nogal, M.; de Leon, P.; Carrascosa, Á.L.; Nieto, R.; Soler, A.; Arias, M.L.; Revilla, Y. African swine fever virus (ASFV) protection mediated by NH/P68 and NH/P68 recombinant live-attenuated viruses. Vaccine 2018, 36, 2694–2704. [Google Scholar] [CrossRef]
  53. Correia, S.; Ventura, S.; Parkhouse, R.M. Identification and utility of innate immune system evasion mechanisms of ASFV. Virus Res. 2013, 173, 87–100. [Google Scholar] [CrossRef]
  54. Reis, A.L.; Goatley, L.C.; Jabbar, T.; Sanchez-Cordon, P.J.; Netherton, C.L.; Chapman, D.A.; Dixon, L.K. Deletion of the African Swine Fever Virus Gene DP148R Does Not Reduce Virus Replication in Culture but Reduces Virus Virulence in Pigs and Induces High Levels of Protection against Challenge. J. Virol. 2017, 91, e01428-17. [Google Scholar] [CrossRef]
  55. Sunwoo, S.Y.; Garcia-Belmonte, R.; Walczak, M.; Vigara-Astillero, G.; Kim, D.-M.; Szymankiewicz, K.; Kochanowski, M.; Liu, L.; Tark, D.; Podgórska, K.; et al. Deletion of MGF505-2R Gene Activates the cGAS-STING Pathway Leading to Attenuation and Protection against Virulent African Swine Fever Virus. Vaccines 2024, 12, 407. [Google Scholar] [CrossRef]
  56. Zhang, T.; Ma, C.; Zhang, Z.; Zhang, H.; Hu, H. NF-kappaB signaling in inflammation and cancer. MedComm 2021, 2, 618–653. [Google Scholar] [CrossRef]
  57. Dorrington, M.G.; Fraser, I. NF-kappaB Signaling in Macrophages: Dynamics, Crosstalk, and Signal Integration. Front. Immunol. 2019, 10, 705. [Google Scholar] [CrossRef]
  58. Gao, Q.; Yang, Y.; Feng, Y.; Quan, W.; Luo, Y.; Wang, H.; Zheng, J.; Chen, X.; Huang, Z.; Chen, X.; et al. Effects of the NF-kappaB Signaling Pathway Inhibitor BAY11-7082 in the Replication of ASFV. Viruses 2022, 14, 297. [Google Scholar] [CrossRef]
  59. 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]
  60. 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]
  61. Barrado-Gil, L.; Del, P.A.; Galindo, I.; Cuesta-Geijo, M.Á.; García-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]
  62. Granja, A.G.; Perkins, N.D.; Revilla, Y. A238L inhibits NF-ATc2, NF-kappa B, and c-Jun activation through a novel mechanism involving protein kinase C-theta-mediated up-regulation of the amino-terminal transactivation domain of p300. J. Immunol. 2008, 180, 2429–2442. [Google Scholar] [CrossRef]
  63. 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, e65821. [Google Scholar] [CrossRef]
  64. Wang, Q.; Zhou, L.; Wang, J.; Su, D.; Li, D.; Du, Y.; Yang, G.; Zhang, G.; Chu, B. African Swine Fever Virus K205R Induces ER Stress and Consequently Activates Autophagy and the NF-kappaB Signaling Pathway. Viruses 2022, 14, 394. [Google Scholar] [CrossRef]
  65. Hernaez, B.; Cabezas, M.; Munoz-Moreno, R.; Galindo, I.; ACuesta-Geijo, M.; Alonso, C. A179L, a new viral Bcl2 homolog targeting Beclin 1 autophagy related protein. Curr. Mol. Med. 2013, 13, 305–316. [Google Scholar] [CrossRef]
  66. Wang, T.; Luo, R.; Zhang, J.; Lan, J.; Lu, Z.; Zhai, H.; Li, L.F.; Sun, Y.; Qiu, H.J. The African swine fever virus MGF300-4L protein is associated with viral pathogenicity by promoting the autophagic degradation of IKKβ and increasing the stability of IκBα. Emerg. Microbes Infect. 2024, 13, 2333381. [Google Scholar] [CrossRef]
  67. Stark, G.R.; Darnell, J.J. The JAK-STAT pathway at twenty. Immunity 2012, 36, 503–514. [Google Scholar] [CrossRef]
  68. 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, e233021. [Google Scholar] [CrossRef]
  69. Li, D.; Peng, J.; Wu, J.; Yi, J.; Wu, P.; Qi, X.; Ren, J.; Peng, G.; Duan, X.; Ru, Y.; et al. African swine fever virus MGF-360-10L is a novel and crucial virulence factor that mediates ubiquitination and degradation of JAK1 by recruiting the E3 ubiquitin ligase HERC5. MBio 2023, 14, e60623. [Google Scholar] [CrossRef]
  70. Chen, Q.; Wang, X.X.; Jiang, S.W.; Gao, X.; Huang, S.; Liang, Y.; Jia, H.; Zhu, H. MGF360-12L of ASFV-SY18 is an immune-evasion protein that inhibits host type I IFN, NF-kappaB, and JAK/STAT pathways. Pol. J. Vet. Sci. 2023, 26, 119–130. [Google Scholar] [CrossRef]
  71. Tian, Y.; Wang, D.; He, S.; Cao, Z.; Li, W.; Jiang, F.; Shi, Y.; Hao, Y.; Wei, X.; Wang, Q.; et al. Immune cell early activation, apoptotic kinetic, and T-cell functional impairment in domestic pigs after ASFV CADC_HN09 strain infection. Front. Microbiol. 2024, 15, 1328177. [Google Scholar] [CrossRef]
  72. Galindo, I.; Hernaez, B.; Diaz-Gil, G.; Escribano, J.M.; Alonso, C. A179L, a viral Bcl-2 homologue, targets the core Bcl-2 apoptotic machinery and its upstream BH3 activators with selective binding restrictions for Bid and Noxa. Virology 2008, 375, 561–572. [Google Scholar] [CrossRef]
  73. Guo, H.; Callaway, J.B.; Ting, J.P. Inflammasomes: Mechanism of action, role in disease, and therapeutics. Nat. Med. 2015, 21, 677–687. [Google Scholar] [CrossRef]
  74. Barrado-Gil, L.; Galindo, I.; Martinez-Alonso, D.; Viedma, S.; Alonso, C. The ubiquitin-proteasome system is required for African swine fever replication. PLoS ONE 2017, 12, e189741. [Google Scholar] [CrossRef]
  75. Mukhopadhyay, D.; Riezman, H. Proteasome-independent functions of ubiquitin in endocytosis and signaling. Science 2007, 315, 201–205. [Google Scholar] [CrossRef]
  76. Huang, L.; Xu, W.; Liu, H.; Xue, M.; Liu, X.; Zhang, K.; Hu, L.; Li, J.; Liu, X.; Xiang, Z.; et al. Correction: African Swine Fever Virus pI215L Negatively Regulates cGAS-STING Signaling Pathway through Recruiting RNF138 to Inhibit K63-Linked Ubiquitination of TBK1. J. Immunol. 2022, 208, 1510–1511. [Google Scholar] [CrossRef]
  77. Li, Y.H.; Peng, J.L.; Xu, Z.S.; Xiong, M.G.; Wu, H.N.; Wang, S.Y.; Li, D.; Zhu, G.Q.; Ran, Y.; Wang, Y.Y.; et al. African Swine Fever Virus Cysteine Protease pS273R Inhibits Type I Interferon Signaling by Mediating STAT2 Degradation. J. Virol. 2023, 97, e194222. [Google Scholar] [CrossRef]
  78. 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]
Figure 1. The relations among cGAS-STING, NF-κB, and JAK/STAT signaling pathways. After invasion of ASFV into host cells, the dsDNA of ASFV binds to and activates cGAS, along with the cGAS-STING pathway, and TBKI is activated to phosphorylate IRF3 and initiate the NF-κB pathway, resulting in production of type I interferons. These type I interferons then activate and phosphorylate JAK1 and tyrosine kinase 2, subsequently phosphorylating STAT1 and STAT2. The phosphorylated STAT1 and STAT2 then bind to IRF9 to form the IFN-stimulated gene factor (ISGF) 3 complex; the ISGF3 complex can translocate into the nucleus and boost the activity of IFN-stimulated response element (ISRE), thereby increasing the expression of IFN-stimulated genes (ISGs), which play important roles in the cell’s response to viruses and pathogens. During the above processes, the ASFV is able to influence the normal function of various regulatory factors and promotes its pathogenicity.
Figure 1. The relations among cGAS-STING, NF-κB, and JAK/STAT signaling pathways. After invasion of ASFV into host cells, the dsDNA of ASFV binds to and activates cGAS, along with the cGAS-STING pathway, and TBKI is activated to phosphorylate IRF3 and initiate the NF-κB pathway, resulting in production of type I interferons. These type I interferons then activate and phosphorylate JAK1 and tyrosine kinase 2, subsequently phosphorylating STAT1 and STAT2. The phosphorylated STAT1 and STAT2 then bind to IRF9 to form the IFN-stimulated gene factor (ISGF) 3 complex; the ISGF3 complex can translocate into the nucleus and boost the activity of IFN-stimulated response element (ISRE), thereby increasing the expression of IFN-stimulated genes (ISGs), which play important roles in the cell’s response to viruses and pathogens. During the above processes, the ASFV is able to influence the normal function of various regulatory factors and promotes its pathogenicity.
Cimb 46 00488 g001
Figure 2. The molecular mechanisms of apoptosis, including intrinsic and extrinsic pathways. Even though these two pathways have distinct starting points, their processes interact with each other and ultimately converge to induce a common apoptotic response.
Figure 2. The molecular mechanisms of apoptosis, including intrinsic and extrinsic pathways. Even though these two pathways have distinct starting points, their processes interact with each other and ultimately converge to induce a common apoptotic response.
Cimb 46 00488 g002
Table 1. Immunosuppression proteins of ASFV.
Table 1. Immunosuppression proteins of ASFV.
Viral ProteinFunctionReferences
pC129RTarget Cyclic GMP-AMP To Inhibit the
cGAS-STING Signaling Pathway
[25]
P17Inhibits cGAS-STING signaling pathway
through interacting with STING
[26]
pM1249LSuppress phosphorylation of TANK-binding kinase (TBK) 1 and
degrading IFN regulatory
factor 3 (IRF3)
[27]
pMGF505-11RBinds to STING and promotes its
degradation through the lysosomal and
autophagy mechanisms
[28]
pMGF360-11LInhibits IL-1, IL-6, and IFN-β secretion[29]
pA528RInhibits phosphorylation of p65[30]
pEP153RPrevent apoptosis via activating the p53
and caspase 3 pathways
[31]
DP71LInhibits apoptosis at early infection[32]
pE199LPromotes cell autophagy through the
interaction of PYCR2
[33]
pA137RInhibited the nuclear import of IRF3[34]
pI215LEncode the ubiquitin-conjugating
enzyme making
[35]
P54, pA179LRegulate the programmed cell death[36]
CD2vHinder lymphocyte proliferation[25]
Table 2. The disadvantages and targets of various vaccines.
Table 2. The disadvantages and targets of various vaccines.
Type of VaccineDisadvantageAdvantage and Target
Traditional vaccineCannot provide strong immunity and has poor long-term immunogenicity.The simple preparation process, high safety.
Subunit vaccineCannot effectively induce the body to generate a cellular immune response.Immunize animals with an adjuvant, high safety.
Nucleic vaccineCan induce the body to generate a cellular immune responseThe spectrum of antigens is not broad enough.
Live attenuated vaccineThe risk of increased virulence and the potential for recombination between different strains, resulting in the emergence of highly virulent strains.Avoiding key issues associated with inactivated vaccines, subunit vaccines, and DNA vaccines, it can replicate continuously in the animal body, mimicking natural infection pathways to effectively induce both humoral and cellular immune responses. Additionally, it does not require adjuvants to enhance the immune response.
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

Huang, T.; Li, F.; Xia, Y.; Zhao, J.; Zhu, Y.; Liu, Y.; Qian, Y.; Zou, X. African Swine Fever Virus Immunosuppression and Virulence-Related Gene. Curr. Issues Mol. Biol. 2024, 46, 8268-8281. https://doi.org/10.3390/cimb46080488

AMA Style

Huang T, Li F, Xia Y, Zhao J, Zhu Y, Liu Y, Qian Y, Zou X. African Swine Fever Virus Immunosuppression and Virulence-Related Gene. Current Issues in Molecular Biology. 2024; 46(8):8268-8281. https://doi.org/10.3390/cimb46080488

Chicago/Turabian Style

Huang, Tao, Fangtao Li, Yingju Xia, Junjie Zhao, Yuanyuan Zhu, Yebing Liu, Yingjuan Qian, and Xingqi Zou. 2024. "African Swine Fever Virus Immunosuppression and Virulence-Related Gene" Current Issues in Molecular Biology 46, no. 8: 8268-8281. https://doi.org/10.3390/cimb46080488

Article Metrics

Back to TopTop