Next Article in Journal
Neutrophil Gelatinase-Associated Lipocalin (NGAL) Is Related with the Proteinuria Degree and the Microscopic Kidney Findings in Leishmania-Infected Dogs
Next Article in Special Issue
Deaminase-Independent Mode of Antiretroviral Action in Human and Mouse APOBEC3 Proteins
Previous Article in Journal
Combination of Colistin and Azidothymidine Demonstrates Synergistic Activity against Colistin-Resistant, Carbapenem-Resistant Klebsiella pneumoniae
Previous Article in Special Issue
The Role of APOBECs in Viral Replication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Retroviral Restriction Factors and Their Viral Targets: Restriction Strategies and Evolutionary Adaptations

Laboratory of Molecular Microbiology, National Institute of Allergy and Infectious Diseases, Bethesda, MD 20892-0460, USA
*
Author to whom correspondence should be addressed.
Submission received: 30 October 2020 / Revised: 30 November 2020 / Accepted: 8 December 2020 / Published: 11 December 2020
(This article belongs to the Special Issue Innate Immunity to Retroviral Infection)

Abstract

:
The evolutionary conflict between retroviruses and their vertebrate hosts over millions of years has led to the emergence of cellular innate immune proteins termed restriction factors as well as their viral antagonists. Evidence accumulated in the last two decades has substantially increased our understanding of the elaborate mechanisms utilized by these restriction factors to inhibit retroviral replication, mechanisms that either directly block viral proteins or interfere with the cellular pathways hijacked by the viruses. Analyses of these complex interactions describe patterns of accelerated evolution for these restriction factors as well as the acquisition and evolution of their virus-encoded antagonists. Evidence is also mounting that many restriction factors identified for their inhibition of specific retroviruses have broader antiviral activity against additional retroviruses as well as against other viruses, and that exposure to these multiple virus challenges has shaped their adaptive evolution. In this review, we provide an overview of the restriction factors that interfere with different steps of the retroviral life cycle, describing their mechanisms of action, adaptive evolution, viral targets and the viral antagonists that evolved to counter these factors.

1. Introduction

Retroviruses replicate by converting their single-stranded RNA genome into double-stranded DNA through a virus-encoded reverse transcriptase. These DNA copies are integrated into host chromosomes where they persist as a permanent part of the host cell genome. When retroviruses infect germline cells, integrated copies can be passed on to the next generation. These integrated proviruses are termed endogenous retroviruses (ERVs). As the retroviral genome encodes only a limited number of genes, replication of exogenous viruses and expression of ERVs rely on the host cell machinery. These viruses can be mutagenic and pathogenic, and this puts strong selective evolutionary pressure on the host genes appropriated for virus replication. The result has been the emergence of defensive genes that can inhibit viral infection and that have been derived from cellular genes, or more rarely, from ERVs. These restriction factors comprise the innate immune system that represents the first line of defense against viral attack.
While each restriction factor has followed a unique evolutionary path, some common themes have emerged (Table 1). First, most restriction factors are encoded by interferon-stimulated genes (ISGs) that are upregulated upon cellular exposure to type I or type II interferons (IFNs). Second, phylogenetic studies in the last two decades demonstrated that most restriction factors show signatures of positive or diversifying selection—that is, the rate of nonsynonymous mutations that result in amino acid substitutions is higher than the rate of synonymous mutations. The rapid evolution observed in the genes encoding retroviral restriction factors has often been countered by rapidly evolving viral evasive mechanisms which include virus-encoded antagonists and signatures of positive selection in the viral target of host restriction factors. This evolutionary conflict represents a genetic arms race described by the Red Queen hypothesis [1,2].
Here, we describe a well-characterized set of retroviral restriction factors for which there is substantial information on their restriction mechanisms as well as their evolutionary history. The majority of these factors were discovered in primates or rodents through studies on HIV-1 which causes the acquired immune deficiency syndrome (AIDS) in humans, and on mouse leukemia viruses (MLVs) which induce lymphomas, immunodeficiencies, and neurological diseases. Some of these factors are restricted to specific taxonomic lineages, while others are carried by many mammalian species, showing broad antiviral activity against multiple retroviruses and in some cases against other families of viruses (Table 1). Decades of studies have produced an increased understanding of how these factors work and what stage of the replication cycle is blocked (Figure 1), and, in many cases, also describing the coevolutionary adaptations in the interacting host and virus proteins.

2. Retroviral Restriction Factors

2.1. Binding and Entry

Retroviruses enter susceptible cells through their interactions with specific cell surface receptors. These interactions can be inhibited by receptor polymorphisms as well as by other transmembrane proteins that interfere with binding or membrane fusion.

2.1.1. Receptors

Receptor-mediated restriction of virus entry has been documented for HIV-1 and MLVs. The HIV-1 envelope (Env) interacts with two membrane proteins for entry, a receptor, CD4, and one of two co-receptors, the chemokine receptors CCR5 and CXCR4 (Figure 2A) [3,4,5]. CD4 is a type I transmembrane protein that functions as a co-receptor for the T cell receptor signaling complex in response to antigen presentation by the MHC Class II molecules [6]. These proteins mediate HIV-1 infection of immune system cells, and receptor-functional orthologs are largely restricted to humans and related primates. CD4 binds virus through the D1 Ig domain, and CCR5 has two virus- binding domains in its N-terminus and in its second extracellular loop (Figure 2A) [7]. CD4 is downregulated in infected cells by the HIV-1 proteins Nef and Vpu (Figure 2B), and this prevents superinfection, thus avoiding apoptosis and production of infection-compromised virions, and also reduces sensitivity to inhibition by another restriction factor, SERINC5 [8,9,10]. While there are no known variants of CD4 in humans that affect the efficiency of HIV-1 infection, CD4 is highly variable in chimpanzees, and this variation is responsible for restricted susceptibility to SIV, the progenitor of HIV-1 (Figure 2B) [11,12]. Moreover, CD4 has been molded by positive selection in primates with rapidly evolving residues found in the HIV-1 Env interacting interface of the CD4 protein, but not affecting the sites targeted by Vpu and Nef (Figure 2B) [13,14]. These findings suggest that co-evolution with SIVs has accelerated the evolution of CD4 in primates. In some human populations, HIV-1 entry can be blocked by a variant of the CCR5 coreceptor with a 32-base pair (bp) deletion in the second extracellular loop of the protein leading to the introduction of a premature stop codon which renders it non-functional as a co-receptor [15].
MLVs isolated from laboratory mice have host range subgroups that rely on two receptors, CAT1 for the ecotropic or mouse-tropic MLVs, and XPR1 for MLVs that can also infect other mammalian species (Figure 2A) [20,21]. These host genes function, respectively, as an amino acid transporter and a phosphate exporter [22,23,24]. CAT1 orthologs are functional as receptors only in mice, but wild mice have only recently been exposed to ecotropic MLVs, as these ERVs are found only in Eurasian and some California mice [25]. Only one mouse CAT1 sequence variant has been identified; the Mus dunni CAT1 restricts infection by Moloney MLV [26]. In contrast, the older XPR1-dependent MLV ERVs are found in all house mouse subspecies [25], and this extended exposure to virus challenge was accompanied by the evolution of six functional XPR1 variants in Mus, five of which restrict different subsets of MLVs. These restrictions result from deletions or substitutions in the two receptor determining regions of XPR1 (Figure 2), all of which were acquired by MLV-infected wild mouse populations (reviewed in [17]). This suggests that the mutant XPR1 variants have a survival advantage which is supported by an observed pattern of positive selection (Figure 2B) and also explains the co-evolution of viral Env variants with different receptor usage patterns [19]. Nonpermissive Xpr1 orthologs are rare among mammals and birds but are found in a few mammalian species like hamsters [27], and in chickens, which were domesticated in India where their exposure to MLV-infected mice likely selected for inactivating XPR1 mutations [18]. A third MLV receptor, the Pit2 phosphate transporter [28,29], has no known functional polymorphisms in mice and is used by wild mouse amphotropic MLVs [30,31], a virus subtype that has not endogenized, and is found as infectious virus only in isolated mouse subpopulations in California [32].
Retrovirus entry can also be blocked by factors that interfere with receptor function (reviewed in [16,33]). The mouse genome contains several such resistance genes including Fv4, which blocks ecotropic MLVs, and Rmcf and Rmcf2 which restrict XPR1-dependent MLVs (Figure 1). These genes have all been identified as ERVs that are defective but have intact env genes capable of producing trimeric proteins comprised of extracellular surface (SU) subunits that bind virus and the transmembrane (TM) subunit responsible for fusing host and viral membranes. Fv4, Rcmf, and Rcmf2 are thought to mask or downregulate the activity of their cognate receptors, and Fv4 additionally has a defect in the fusion peptide of the transmembrane domain of env, so virions that incorporate this Env have reduced infectivity [34]. This use of co-opted Env genes to block exogenous infection has also been described in chickens, sheep, and cats (reviewed in [33]).

2.1.2. SERINC5

SERINC5 belongs to the serine incorporator (SERINC) gene family, a highly conserved group of genes found in all eukaryotes that encode 9-11 pass transmembrane proteins. SERINC proteins incorporate the amino acid serine into the lipids of cell membranes [35]. Mammals carry five SERINC genes while some lower eukaryotes, such as C. elegans and S. cerevisiae have only a single SERINC gene [35]. SERINC5 is highly expressed in multiple tissues in humans including lymphoid tissues but is not induced by interferons [36,37].
SERINC3 and SERINC5 restrict replication of HIV-1 variants lacking the viral accessory protein Nef [36,37]. Analysis of transcriptional profiles or virion proteomes in the presence or absence of Nef established SERINC3 and SERINC5 as targets of Nef, and identified SERINC5 as having more impact on infection by Nef-deficient HIV-1 [36,37]. In the absence of Nef, SERINC5 is incorporated into virions in producer cells [36,37,38]. The Nef block to SERINC5 incorporation likely involves vesicular trafficking, endolysosomal degradation, and a role for polyubiquitination in the targeting of SERINC5 to the endolysosomal vesicular bodies [36,37,39,40,41,42]. The sensitivity of SERINC5 to Nef-mediated antagonism maps to the last of the protein’s cytoplasmic loops [43,44].
SERINC5 is also antagonized by the glycogag (g-gag) of MLVs and S2 of EIAV [36,37,40,45,46,47]. G-gag is a Type II transmembrane protein that is expressed via an alternative start codon upstream of gag in some MLVs. Both g-gag and S2-mediated downregulation of SERINC5 levels were shown, like Nef, to function through the endolysosomal system [36,37,40,47]. This convergent evolution of SERINC5 antagonism in different retroviruses is noteworthy in light of the lack of shared sequence or structural homology between Nef, S2, and g-gag.
SERINC5-mediated restriction of HIV-1 is virus Env-type dependent, suggesting that SERINC5 restriction could be at the level of viral entry (Figure 1) [36,37,48,49]. In accordance with this finding, viral reverse transcription in target cells is reduced and viral core delivery to the cytoplasm is blocked in the presence of SERINC5 [36,37,50]. Moreover, fluorescent microscopy of single viral particles as well as super-resolution fluorescent imaging showed that the packaging of SERINC5 into viral particles leads to inhibition of fusion between the virus and the target cell [51,52].
SERINC5 orthologs from multiple mammals as well as amphibians and fish can restrict HIV-1 infection [43,45,53,54,55]. This is in line with the high degree of conservation in the primary amino acid sequence of SERINC5 in vertebrate lineages, and this conservation is reflected in the fact that SERINC5 is under purifying rather than positive selection in primates, suggesting that preservation of its cellular function cannot tolerate alterations [56].

2.2. Post Entry

After traversing the cell membrane, the retroviral capsid begins to uncoat and reverse transcription (RT) produces a DNA copy that is then transported into the nucleus where it integrates into host chromosomes. Multiple host factors can interfere with these processes at or after RT (Figure 1). These factors differ in domain structure and show signatures of positive selection that align with regions that are either important for restriction or are targeted by viral antagonists (Figure 3).

2.2.1. Fv1

Fv1 was the first retroviral restriction factor to be discovered [57]. Pioneering experiments with different isolates of Friend MLV showed that susceptibility to these isolates varies among inbred mouse strains [57,58]. Specifically, NIH Swiss mice with the Fv1n allele are permissive to MLVs classed as N-tropic but not to B-tropic MLVs, while BALB/c mice (Fv1b) are more permissive to B-tropic MLVs [58,59]. Other mouse strains and Mus species carry alternative alleles that restrict other MLV variants such as Fv1nr in 129 mice, Fv1d in DBA, and other restrictive and nonrestrictive variants in wild mouse species [60,61]. Almost three decades after the discovery of Fv1 restriction, the responsible gene was identified using positional cloning and shown to have homology to the gag gene of an ancient ERV family termed ERV-L [62,63].
The Fv1 block occurs after reverse transcription but before the integration of the viral DNA into the host genome (Figure 1) [75]. Constitutive expression of Fv1 is low and not IFN-inducible, and it can be saturated by high virus titers [76]. Fv1 restriction targets the capsid protein of the virus, and specific residues involved in restriction have been identified in the target region of the virus capsid as well as in Fv1 [60,61,77,78]. Fv1 interaction with the capsid requires the assembly of a higher-order capsid structure suggesting interference with the virus uncoating process [79].
Fv1 is found in all but the most basal species in the Mus phylogenetic tree, and is missing in the rat genome, so it was initially thought that Fv1 was acquired shortly after the origins of the Mus genus [62,80,81]. Taking advantage of the availability of whole-genome sequenced species, we showed that Fv1 entered the genome of rodents much earlier than previously thought [64]. These findings indicate that the ERV ancestor of Fv1 was fixed in the common ancestor of the rodent families Muridae, Cricetidae, and Spalacidae at least 45 million years ago [64,82,83]. Although Fv1 is lost or substantially mutated in a variety of rodent lineages, the Fv1 open reading frame (ORF) is present in several branches of the rodent family Muridae [64,82].
The accelerated evolution of Fv1 led to its loss in several lineages but also established signatures of strong positive selection (Figure 3) [64,80,82]. We and others have shown that Fv1 is evolving under positive selection in Mus as well as other lineages in Muridae [64,80,82]. Some of the residues evolving under positive selection determine Fv1 restriction suggesting that exposure to retroviral pathogens contributed to this evolution [64,78,80,81,82]. Fv1 has thus been evolving under positive selection much longer than the exposure of mice to the MLVs that initially defined Fv1 restriction, and subsequent analyses showed that some Fv1 variants have demonstrated restriction activity against retroviruses in other genera including foamy viruses and lentiviruses (Table 1) [25,80,81].

2.2.2. TRIM5

In the late 1990s, Fv1-like early blocks to the replication of HIV-1 and MLV were observed in various primate and other mammalian cell lines [84,85]. The responsible factor was identified as TRIM5α [86]. TRIM5 belongs to the tripartite motif (TRIM) family of genes that encode E3 ubiquitin ligases. The human genome contains more than 80 TRIM family members with roles in various functions ranging from autophagy and innate immunity to cellular differentiation [87,88]. The TRIM genes are named after the three main domains they encode; RING, B-Box, and coiled-coil (Figure 3) [89]. A subset of the TRIM family of proteins, including TRIM5, also contains a fourth domain called SPRY or B30.2 [89]. The restriction factor TRIM5α is encoded by the longest isoform of the primate TRIM5 gene and includes a SPRY domain [86].
TRIM5 contributes to the species-specific post-entry restriction of retroviruses in several mammalian species [90]. Like Fv1, TRIM5 blocks retroviral replication after entry but before integration (Figure 1) and targets at least one of the same capsid residues as Fv1 [84,86]. Phylogenetic analysis of TRIM5 orthologs in primates revealed that this gene has been under positive selection [65]. The rapidly evolving residues are concentrated in a “patch” in the SPRY domain and at least one of those sites was subsequently shown to be critical for restriction [65,91]. In fact, TRIM5 binds to the viral capsids of incoming virus through its SPRY domain [69,70]. This leads to a strong block at reverse transcription (Figure 1) [86]. TRIM5 binding to viral capsid occurs only when the capsid molecules are assembled into a higher-ordered structure, and TRIM5 multimers then form a hexagonal net on the viral capsid core [70,92,93,94,95,96,97]. In contrast to its rhesus macaque ortholog, human TRIM5α is a weak restriction factor against HIV-1, but can become a significant contributor to the IFN-mediated inhibition of HIV-1 which likely involves activation of immunoproteasomes [86,98,99,100,101,102].
Like other TRIM proteins, TRIM5 is an E3 ubiquitin ligase owing to the presence of its RING domain [89]. This TRIM5 feature suggested that the ubiquitin-proteasome pathway may be involved in the TRIM5-mediated restriction of retroviruses and in fact, TRIM5 is a short-lived protein that goes through self-ubiquitination [86,103,104]. Hence one proposed mechanism for TRIM5-mediated restriction involves TRIM5 multimers binding to the viral capsid lattice and recruiting proteasomes that degrade the viral core and its components [104,105,106]. Interestingly, both the addition of proteasome inhibitors and the introduction of point mutations in the RING domain of TRIM5 that disrupt self-ubiquitylation restores reverse transcription, but these additions do not restrict virus infection suggesting that TRIM5 may impose subsequent blocks to viral infection [70,104,106,107].
Due to its ability to recognize retroviral capsid structure, TRIM5 was, until recently, thought to be a retrovirus-specific restriction factor. However, TRIM5α of humans and rhesus macaques can inhibit the replication of some flaviviruses by promoting ubiquitination and subsequent degradation of flavivirus protease [102].
TRIM5 from multiple mammalian species has antiviral activity and presents a complex evolutionary history marked by gene duplications, losses, and fusions, some of which have been linked to retroviral resistance [108,109,110,111,112,113,114,115,116,117,118]. The first such example of fusion associated restriction was identified in owl monkeys, where HIV-1 restriction is caused by the LINE-1-mediated retrotransposition of an intronless copy of the cyclophilin A (CypA) gene into the 3′ end of the TRIM5 gene to generate a TRIMCyp fusion protein [108]. This fusion replaces the capsid-binding SPRY domain with the capsid-binding CypA domain (Figure 3) [108]. Independent CypA insertions creating similar but structurally different TRIMCyp fusion genes are found in several mammalian lineages, including old world monkeys, tree shrews and rodents, examples of the remarkable convergent evolution of retrotransposition-driven gene fusion in disparate lineages, some of which have demonstrated antiviral activity [112,115,116].
In addition to its function as a retroviral restriction factor, TRIM5 can act as a pattern recognition receptor of the retroviral capsid and activate innate immune signaling pathways [119], although this function may not be uniform in different species [116,120,121]. TRIM5α binding to the viral capsid core initiates a cascade that leads to polyubiquitination of TRIM5α and activation of activator protein 1 (AP-1) and nuclear factor kappa B (Nf-κB) pathways as opposed to the monoubiquitination that happens in the absence of viral infection [122]. This distinct ubiquitination pattern is thought to act in a way that specifically activates the innate immune response only in the presence of virus infection [122].

2.2.3. APOBEC3G

APOBEC3G belongs to the apolipoprotein B mRNA editing enzyme catalytic polypeptide-like/activation induced cytidine deaminase (APOBEC/AID) family of genes [123]. The 11 members of the APOBEC/AID family in humans include AID, APOBEC1, APOBEC2, APOBEC4, and seven clustered paralogs of APOBEC3A-G. All members of this family except for APOBEC2 and APOBEC4 can catalyze the deamination of cytosine to uracil in single-stranded DNA/RNA [123].
APOBEC3G (A3G) was originally identified as a restriction factor for its ability to inhibit the replication of Vif-deficient HIV-1 [124]. Both human A3G and its mammalian orthologs also restrict other lentiviruses and other retroviruses [125,126,127,128,129,130,131,132,133]. A3G is packaged into viral particles and, in subsequently infected cells, it can catalyze the deamination reaction on newly formed viral single-stranded DNA during reverse transcription [134,135]. This leads to the accumulation of G-to-A mutations on proviral DNA which can produce defective viral proteins and non-infectious viral particles [134,135]. A3G can also block reverse transcription independently of its deaminase function [136,137,138]. Moreover, this deaminase-independent block to reverse transcription of Moloney MLV and mouse mammary tumor virus is the major factor in virus restriction by mouse APOBEC3 (mA3) in vivo [139,140,141,142,143]. mA3 can also block the proteolytic processing of MLV gag and gag-pol and some mA3 proteins incorporate an extra exon that decreases translation efficiency [144,145].
In addition to A3G, several other members of the A3 family can restrict HIV-1 to varying degrees including A3C, A3D, A3F, A3H [125,146,147,148,149,150,151]. Notably, natural polymorphisms of A3C, A3F and A3H in human populations can lead to different restriction profiles against HIV-1 [132,133,147,148,152]. In the case of A3C, a single haplotype found in African populations is the only known variant capable of blocking Vif-deficient HIV-1 [132,148]. Moreover, Both A3F and A3H are highly polymorphic in humans with several variants showing anti-HIV-1 activity [133,146,147,152,153,154,155,156].
Overexpression of APOBEC/AID proteins in transgenic mice can be mutagenic and oncogenic, so expression levels must be regulated [157]. APOBEC3 (A3) genes are expressed at higher levels in hematopoietic cells than the cells of other lineages [158]. Most A3 genes show high expression levels in T cells, the main target of HIV infection, and the expression of multiple members of the A3 gene family can be induced by IFNα [158,159]. In addition, mA3 levels are elevated in mice that have an MLV LTR inserted into this gene, another example of exapted ERVs with an antiviral role [160].
A3G was the first restriction factor shown to be counteracted by a retroviral accessory protein. The antiviral activity of the A3 proteins is antagonized by the lentiviral Vif protein [124]. Vif degrades A3G by recruiting an E3 ligase complex that is composed of Cullin5, Elongin B, Elongin C, and Rbx1 [161]. This leads to polyubiquitination and eventual proteasomal degradation of A3G [161]. Vif recruitment and formation of the E3 ligase complex also require the transcription cofactor CBF-β (core-binding factor subunit beta) which stabilizes Vif binding to A3G and the E3 ligase complex [162,163,164,165]. In addition to Vif as an antagonist of human A3G, the MLV g-gag can counter the mA3 restriction of MLV replication [140,166,167]. Unlike the degradative impact of Vif on A3G, g-gag antagonizes mA3 indirectly by stabilizing the viral core and preventing A3 access to the viral DNA/RNA by shielding the viral reverse transcriptase complex [167]. Apart from g-gag, mA3-mediated restriction of MLV replication is also be counteracted by the viral p50 protein, produced by alternative splicing of gag, which interacts with mA3 and prevents its packaging into newly produced virions [168,169].
While the APOBEC/AID gene family likely originated in early vertebrates, A3 genes are only found in placental mammals [170,171]. The A3 paralogs are tightly clustered in a conserved locus, but the copy number of the A3 genes varies greatly among mammals. For example, while most rodents only have a single A3 gene, the A3 locus saw an expansion in primates and a recent study described the acquisition of new A3 copies in primate genomes through retrotransposition, some of which are active [172,173]. The genomic structure of these A3 paralogs and orthologs also varies as they can contain one or two zinc-coordinating domains, named according to sequence homology as Z1, Z2, and Z3, but only one Z domain per gene has deaminase activity [174]. In the case of human A3G, the C-terminal cytidine deaminase domain (Z1) has catalytic activity [174].
Phylogenetic and computational analyses revealed that A3G has evolved under positive selection in primates and a subsequent analysis of mammalian A3 genes found signatures of positive selection at several sites concentrated in the Vif binding region in loop 7 of the N-terminal A3Z2 domain involved in substrate recognition [66,71,175]. The single copy of mA3 is also under positive selection in Mus, and the positively selected residues in the catalytically active A3Z2 domain line the substrate groove that accommodates nucleic acids [160]. The demonstration that A3 proteins can also more broadly restrict replication of endogenous retroviruses and retrotransposons, indicate that A3 mutators have been in conflict with retroelements throughout mammalian evolution [175,176,177,178,179]. Furthermore, A3G has been demonstrated to have broader antiviral activity that can block HBV, replication of which includes a reverse transcription step [180,181,182].

2.2.4. SAMHD1

Sterile alpha motif and histidine/aspartic acid (HD) domain containing protein 1 (SAMHD1) is an ISG that functions as a triphosphohydrolase and regulates the levels of intracellular deoxynucleoside triphosphates (dNTPs) [183]. SAMHD1 was first identified as a restriction factor that blocks HIV-1 infection in myeloid and dendritic cells and was later linked to HIV-1 restriction in resting CD4-positive T cells [184,185]. SAMHD1 blocks HIV-1 early in the replication cycle (Figure 1), as its dNTPase activity leads to the depletion of the intracellular dNTP pools available for viral reverse transcription [183,184,186]. When SAMHD1 is depleted from myeloid or dendritic cells, there is an increase in late reverse transcription products, consistent with SAMHD1 inhibition of HIV-1 prior to nuclear import [184,185,186].
The HIV-1 restriction activity of SAMHD1 is limited to non-dividing cells, as these cells have lower levels of intracellular dNTPs than dividing cells [183,184,185]. This aligns with the observation that the phosphorylation status of SAMHD1 at residue T592 is linked to cell cycle regulation [187]. Phosphorylation is accomplished upon S phase entry by the cyclin-dependent kinases (CDKs), CDK1 and CDK2 together with cyclin A2 [188]. Following M phase exit, SAMHD1 is dephosphorylated by the phosphatase PP2A-B55α [187]. The phosphorylation status of SAMHD1 is linked to its anti-retroviral activity, although phosphomimetic mutants of SAMHD1 still decrease dNTP pools but lack anti-lentiviral activity suggesting that there may be a dNTPase-independent restriction function by SAMHD1 [189,190,191,192].
SAMHD1 restriction of lentiviruses is counteracted by two related viral accessory proteins found in different lentiviruses: Vpx in HIV-2 and the primate lentiviruses SIVsm, SIVmac, and SIVrcm, and the Vpr variants in some SIVs [67,184,186,193,194]. Vpx mediates degradation of SAMHD1 through a proteasome-mediated mechanism [74,184,186]. Vpx interacts with the C-terminal domain of SAMHD1 (Figure 3) and recruits the Cullin4-DCAF E3 ubiquitin ligase complex [74,184,186,195]. This leads to polyubiquitination and eventual proteasomal degradation of SAMHD1. Some other lentiviruses, like HIV-1 and its close relative SIVcpz, do not encode Vpx and are thus unable to counteract SAMHD1-mediated restriction [184]. The in vivo consequences of this antagonism of SAMHD1 restriction are unclear as the absence of Vpx in HIV-1 and some SIVs may be beneficial to these viruses by providing a mechanism for immune evasion. This is because unlike HIV-2, HIV-1 cannot efficiently infect dendritic cells and therefore these cells fail to induce IFNβ, so no broad antiviral response can be activated [184,196,197]. However, while SIVmac, which expresses Vpx, can readily infect dendritic cells, it causes a pathogenic infection in macaques suggesting that disease induction involves factors other than immune evasion and Vpx interference [198].
SAMHD1 has evolved under positive selection in primates as well as other mammals, and some of these positively selected sites overlap with the Vpx interaction sites (Figure 3) [67,193,194,199]. SAMHD1 orthologs from cats, horses, and cows also show dNTPase activity [200,201]. Both feline and human orthologs inhibit feline immunodeficiency virus (FIV) when overexpressed, and human SAMHD1 also inhibits equine infectious anemia virus (EIAV) and SIV [200,202]. While alpha-, gamma-, and betaretroviruses cannot productively infect non-dividing cells, the addition of exogenous Vpx before infection leads to the increase of late RT products of MLV, Mason Pfizer monkey virus (MPMV) and Rous sarcoma virus (RSV), but not foamy virus, which largely completes reverse transcription prior to target cell entry [202]. In addition to its restriction of lentiviral replication, SAMHD1 inhibits HTLV-1 (human T cell leukemia virus type 1) in non-dividing monocytes [203]. While it remains to be seen whether SAMHD1 has any role in restricting non-lenti retroviruses in vivo, none of the studies performed so far has shown antagonism of SAMHD1 by any of these other retroviruses.
SAMHD1 also restricts HBV, a member of the hepadnavirus family as well as multiple double-stranded DNA viruses, including herpes simplex virus-1 (HSV-1), mouse and human cytomegalovirus, and vaccinia virus [204,205,206,207,208,209]. The dNTPase function of SAMHD1 is required for the restriction of HBV while the inhibition of HSV-1 and vaccinia virus replication was only observed in non-dividing cells [204,205]. Moreover, as shown for retroviruses, phosphorylation at T592 also relieves the inhibitory impact of SAMHD1 on HBV [205,206,210]. This ability to inhibit multiple families of viruses suggests that SAMHD1-mediated depletion of dNTP pools has been adapted to provide a broad innate immune mechanism against viral infection in mammals.

2.2.5. MX2

Myxovirus resistance (MX) proteins are dynamin-like GTPases that are found in most vertebrates [68]. Humans, like most mammals, encode two MX genes: MX1 and MX2 [68]. Structural analyses of human MX1 and MX2 proteins reveal a similar domain architecture, despite having only 63% amino acid identity, with a globular GTPase domain connected to the C-terminal stalk domain via a flexible bundle signaling element (Figure 3) [211,212]. Human MX2 was initially described as an HIV-1 restriction factor activated by IFNα [213,214]. MX2 expression has no impact on late reverse transcription products but results in decreased levels of HIV-1 2-LTR circles and integrated proviral DNA [213,214,215]. These findings place MX2-mediated restriction of viral replication after reverse transcription but before integration (Figure 1) [213,214,215]. While the precise mechanism of MX2 inhibition of HIV-1 infection remains to be determined, the viral capsid is a critical target of MX2 restriction, since capsid-specific replacement mutations can escape MX2-mediated inhibition [213,214,215,216,217]. MX2 interacts with in vitro assembled capsid and capsid-nucleocapsid structures and two binding domains have been identified in the N-terminus and in the GTPase region (Figure 3) [72,73,218,219,220].
The determination that MX2 inhibits viral nuclear import is based on the analysis of the two transcriptional isoforms in humans that result from alternative use of an internal start codon [221]. The shorter isoform lacks the N-terminal nuclear localization signal and does not show any antiviral activity, but interferes with the restrictive activity of the longer isoform through competitive capsid binding [72,213,214]. Moreover, the longer isoform localizes to the nuclear periphery while the shorter isoform is cytoplasmic as is MX1 [222]. These findings, together with the demonstration that MX2 inhibition is after reverse transcription, focused attention on nuclear import and led to the findings that several nuclear import proteins interact with MX2 and are involved in MX2 inhibition of HIV-1 [223,224,225].
MX2 has been under accelerated evolution in primates [68]. Both MX genes have a complex evolutionary history in mammals where they have undergone duplications, losses, and gene conversions between the two MX genes [68,226,227]. While humans and other primates have MX1 and MX2 orthologs, rodents, except for the squirrel clade, have a duplication of Mx1 and lack an Mx2 ortholog [68]. This observation together with the fact that human MX2 does not affect MLV replication suggests that MX2 restriction of retroviruses may be specific to lentiviruses [213,214]. Among the species that are infected by known exogenous lentiviruses, horses carry orthologs of both MX1 and MX2 and equine MX2 blocks EIAV replication at the same point of the replication cycle as the block to HIV-1 [68,228,229]. In contrast, cats, infectible by FIV, and rabbits that harbor an endogenous lentivirus, only contain a single MX gene in their genome, an ortholog of MX1 [68].
Like MX1, MX2 inhibits a variety of other viruses including herpesviruses, HBV, and flaviviruses as well as LINE-1 retrotransposons, suggesting that this restriction factor evolved to play a significant role in the mammalian defense against viral invasion [201,228,230,231,232,233,234]. Notably, while the GTPase activity of MX2 is dispensable for its restriction of HIV-1, MX2 mutants that are deficient in GTP binding or hydrolysis are unable to block the infection of herpesviruses [231]. The fact that sites under positive selection in primate MX2 do not coincide with the residues that bind HIV-1 capsid structures suggests that the diversifying evolution that shaped this gene was influenced by its interaction with a broader set of pathogens (Figure 3) [68,218].

2.3. Post Integration

Restriction factors that operate in late stages of viral replication have been identified in more recent years. While the two highlighted here, ZAP and SLFN11, do not have well-defined mechanisms of action or known viral antagonists, both are under positive selection.

2.3.1. ZAP

Zinc finger antiviral protein (ZAP) is a broad restriction factor that is encoded by the human gene ZC3HAV1 (zinc finger CCCH-type containing, antiviral 1). Originally discovered as an inhibitor of MLV replication, ZAP is a member of the poly ADP ribose polymerase (PARP) family although both isoforms of ZAP (ZAP-L and ZAP-S) lack the poly ADP ribosylation activity [235,236]. The four ZAP zinc fingers form a binding pocket for the CpG dinucleotides in viral mRNAs which leads to either their degradation or translational repression (Figure 1) [235,237,238,239,240,241,242,243,244]. While the exact mechanism of the ZAP-mediated viral RNA inhibition is not known, it has been shown that ZAP recruits host proteins such as TRIM25 and KHNYN known for their antiviral actions [245,246,247]. ZAP has evolved under positive selection in primates with rapidly evolving residues concentrated at the PARP-like domain [248]. ZAP also restricts a variety of other viruses rich in CpG dinucleotides, including alphaviruses, filoviruses, and HBV as well as retrotransposons (Table 1) [238,249,250,251,252,253,254,255].

2.3.2. Schlafen11

Members of the Schlafen family of genes, which includes Schlafen11 (SLFN11), are differentially expressed during thymocyte development [256]. Schlafens are involved in a variety of processes including the regulation of cell cycle, immune cell differentiation, and virus replication [257,258,259]. These genes are found in mammals as well as a few amphibian and fish species [260]. Humans encode six Schlafen genes that are clustered on chromosome 17, and all human Schlafen genes are ISGs [258]. The ten mouse Schlafen genes do not include a SLFN11 ortholog.
Human SLFN11 inhibits HIV-1 replication at the level of protein synthesis (Figure 1) [226]. Evidence suggests that SLFN11 interacts with tRNAs and blocks the shift of the composition of the tRNA pool induced by HIV-1 [226,261]. The impact of SLFN11 on codon usage bias does not seem to be specific to HIV-1 as SLFN11 can also inhibit translation of EIAV, flaviviruses, and even non-codon optimized non-viral genes [262,263,264]. Moreover, SLFN11 has evolved under positive selection in primates, but positively selected residues do not seem to be responsible for the restriction level differences observed among primate orthologs suggesting that either an unknown viral antagonist or a non-viral cellular process may be driving this accelerated evolution [263].

2.4. Envelope Processing and Packaging

2.4.1. GBP5

Guanylate-binding protein-5 (GBP5) is a member of a family of small GTPases that can be induced by IFNγ [265]. GBP5 was identified as a potential restriction factor of HIV-1 in an evolutionary screen of human genes under positive selection [266]. GBP5 and, to a lesser extent, its paralog GBP2, inhibit HIV-1 replication by interfering with the activity of cellular protease furin which leads to defective envelope processing and incorporation [267,268]. Mutations in the GTPase domain of GBP5 had no impact on its ability to restrict HIV-1 [267]. While there is no evidence of viral antagonism against the restrictive action of GBP5, it has been suggested that mutations in the Vpu initiation codon may confer an advantage to the virus against this inhibition [267,269]. Inhibition of furin cleavage by GBP5 is also responsible for the inhibition of other viruses including MLV, influenza virus and measles virus, and GBP5 is a major contributing factor to the IFNγ-mediated restriction of respiratory syncytial virus demonstrating that this restriction factor has broad antiviral activity [268,270].

2.4.2. MARCH8

Membrane-associated RING-CH 8 (MARCH8) is a member of a RING finger E3 ubiquitin ligase family with 11 members in the human genome [271,272]. MARCH8 blocks HIV-1 Env incorporation into viral particles through surface downregulation that depends on a tyrosine motif found in the cytoplasmic tail of the viral Env [272,273]. The antiviral target of MARCH8 is not limited to the HIV-1 envelope since the Vesicular Stomatitis Virus (VSV) and Ebolavirus glycoproteins were also downregulated from the cell surface in the presence of MARCH8 [272,274]. Two other members of the MARCH family, MARCH1 and MARCH2, can also inhibit HIV-1 and VSV envelope incorporation, and, unlike MARCH8, expression of MARCH1 and MARCH2 can be induced by type I IFNs [275,276].

2.4.3. IFITMs

Interferon-induced transmembrane proteins (IFITMs) are a family of small transmembrane proteins upregulated by interferon during virus infection that are evolutionarily conserved among vertebrates [277]. The most well-studied member of this family, IFITM3, restricts the replication of a variety of viruses including influenza virus, flaviviruses and HIV-1 [278,279,280,281,282,283,284,285,286,287,288]. The mechanism of this restriction is not fully understood as inhibition targets two different stages of the viral life cycle (Figure 1); restriction is observed when IFITM3 is expressed in target cells where it interferes with virus entry as well as in producer cells where it can decrease production of infectious virus. [279,280,289,290,291,292,293]. At the entry level, IFITM3 interferes with fusion between the viral and celluIar membranes and can also reduce the fusogenic activity of syncytin proteins responsible for trophoblast fusion in placentation [278,289,294,295,296]. IFITM3 is embedded in the membranes of endocytic vesicles and after membrane fusion, IFITM3 traffics infecting viruses to lysosomes [294].
When IFITM3 is expressed in virus-producing cells, it is incorporated into retroviral particles which then show substantially decreased ability to infect target cells [280,290,291,292]. IFITM3 interferes with the cleavage of the HIV-1 Env protein gp160, leading to a decrease in the amount of mature Env in viral particles [291,293,297]. This inhibition of envelope packaging was also observed for ecotropic and xenotropic variants of MLV [293]. That IFITM3 interferes with envelope processing/packaging is consistent with the observation that CXCR4-tropic strains of HIV-1 are more sensitive to IFITM3 and IFITM2-mediated restriction than CCR5-tropic strains [298]. While there is no evidence of antagonism of IFITM3 by any of the HIV-1 proteins, g-gag of Moloney MLV can counteract the IFITM3 restrictive function [293].
The number of genes in the IFITM cluster varies among vertebrate genomes [299]. For example, while the human IFITM cluster in chromosome 11 is comprised of five genes; IFITM1, IFITM2, IFITM3, IFITM5 and IFITM10, the orthologous cluster in the mouse contains six IFITM genes [299]. Signatures of positive selection were found among the primate IFITM1, 2 and 3 genes located in this cluster, and such sequence variants were also identified in human populations using haplotype analysis [299,300].

2.4.4. Additional Factors

In addition to the restriction factors described above, other proteins have been shown to inhibit retroviral envelope processing and packaging including the mannose receptor and endoplasmic reticulum alpha-mannosidase I (ERManI).
Mannose receptor is a type I transmembrane protein expressed on the surface of macrophages that acts as a pattern recognition receptor with important roles in pathogen internalization by macrophages (reviewed in [301]). This protein inhibits virus egress and also reduces HIV-1 envelope levels [302,303]. While the expression of mannose receptor is not controlled by type I IFNs, HIV-1 accessory proteins Nef and Vpr can counteract the mannose receptor-mediated block to viral release by reducing the surface expression of mannose receptor [303].
ERManI is a member of the glycoside hydrolase family 47 α-mannosidases and plays a critical role in the endoplasmic reticulum-associated protein degradation (ERAD) pathway (reviewed in [304]). ERManI restricts HIV-1 infectivity by initiating the mitochondrial translocator protein (TSPO)-induced ERAD pathway which leads to the degradation of viral Env [305,306]. While the exact mechanism of this restriction is unknown, mutations that abrogate the catalytic activity of ERManI reverse its inhibitory activity on Env suggesting that ERManI may be involved in glycosylation of HIV-1 Env [306]. In addition to its antiviral activity against HIV-1, ERManI is also involved in ERAD-mediated degradation of HA protein of Influenza virus [307].

2.5. Assembly and Release

BST2/Tetherin

Soon after the discovery of Vpu as an HIV-1 accessory gene, it was demonstrated that HIV-1 without Vpu was not efficiently released from certain cell types [308]. Bone marrow stromal cell antigen 2 (BST2) was identified as the responsible ISG by comparing microarray results from various cell types based on their ability to restrict Vpu-deficient HIV-1 [309]. BST2 was originally identified as a surface antigen that is specifically expressed in differentiated B cells [310]. It was later shown that BST2 is expressed in a wide variety of tissues with variable levels of expression [311]. BST2 is a type II single-pass transmembrane protein with a unique domain structure (Figure 4A) [312]. It contains a short N-terminal cytoplasmic tail followed by an alpha-helical transmembrane domain, a coiled-coil ectodomain, and a C-terminal glycosylphosphatidylinositol (GPI) anchor (Figure 4B) [312]. It is the only gene in the human genome that encodes a protein with this structure [313]. This unique topology, with a transmembrane domain at its N-terminus and a GPI anchor at its C-terminus, allows BST2 to act as a bridge between the cell membrane and budding virions, and therefore, BST2 has also been referred to as tetherin [314]. The ability of BST2 to tether budding virions to the cell membrane is not specific to retroviruses, as BST2 blocks release of several other enveloped viruses, including herpesviruses, filoviruses, VSV and SARS coronavirus [315,316,317,318,319,320,321,322,323,324,325].
In addition to cell-free transmission, HIV-1 can also spread between cells via direct contact (reviewed in [326]). While the initial studies on BST2 restriction of retroviruses clearly established a strong block on the cell free transmission of viruses, the role of BST2 on cell-to-cell transmission of HIV-1 is less clear. BST2 can block cell to cell transmission of Vpu deficient HIV-1 from macrophages to CD4 T cells [327], however conflicting results have been reported on BST2-mediated restriction of cell to cell transmission of HIV-1 between T cells [328,329].
The HIV-1 Vpu antagonizes BST2 via multiple mechanisms. Initial studies showed that Vpu interacts with and recruits beta-transducin repeat-containing protein (β-TrCP) to BST2 [330,331,332]. β-TrCP is an F-box protein that makes up the substrate recognition part of the E3 ligase complex with SCF (Skp1-Cullin-F-box) [330,331]. This leads to polyubiquitination and eventual lysosomal degradation of BST2 [330,331]. There is also evidence that Vpu may downregulate BST2 by interfering with its trafficking through the trans-Golgi network [333,334]. In addition to antagonism by the HIV-1 Vpu, the Nef protein of various strains of SIV and the Env of HIV-2 antagonize BST2, and this Nef antagonism of BST2 is important during in vivo infection of rhesus macaques [335,336,337,338,339,340].Unlike their primate counterparts, mouse and rat BST2 orthologs cannot be counteracted by Vpu, Nef or HIV-2 Env [341]. Moreover, mouse BST2 can block spread and pathogenesis of MLV in vivo when induced by type I IFNs or poly (I:C) [342].
Evolutionary studies of BST2 reveal an ancient origin in early vertebrates [313,343]. BST2 is unusual among retroviral restriction factors in that the amino acid sequence of BST2 shows substantial sequence divergence between different mammalian lineages, although its unique structure is retained [313]. Hence, the ability of BST2 to restrict enveloped viruses relies on its specific domain topology rather than primary amino acid sequence [314]. Within the primate lineage, however, there is a distinctive pattern of sequence divergence. The primate BST2 evolved under diversifying selection with the positively selected residues concentrated at the cytoplasmic tail and the transmembrane domain, where the Vpu and Nef interacting residues have been mapped (Figure 4A) [344].

3. Concluding Remarks

There has been explosive growth in the identification of retroviral restriction factors since the discovery of Fv1 in 1976, and this set of genes will continue to expand as new screening methods reveal additional cellular proteins that have antiviral activity [99,100,349,350,351]. Most antiretroviral factors are cellular genes that evolved to become part of the innate immune response against viruses, although a subset of restriction factors including Fv1, the MLV LTR embedded in mouse Apobec3, and the Env-producing Fv4, Rmcf and Rmcf2 genes are domesticated ERVs that have no known cellular purpose.
For some of these factors, restriction operates through interference with cellular processes needed for virus replication, for example, depletion of dNTP pools by SAMHD1 or tRNA pools by SLFN11 [183,257]. Other restriction factors interact with viral nucleic acids such as ZAP which targets CpG islands and APOBEC which induces G to A mutations during reverse transcription [135,247]. A larger subset of restriction factors, however, interact with viral proteins (Figure 5). Their interacting interfaces often display signatures of positive selection and in fact, some of the most intense adaptive evolution in mammalian genes is observed among retroviral restriction factors, marking ratchet-like coevolutionary trajectories of sequential adaptations in restriction factors, their viral targets and viral antagonists [14,352]. These interfaces can either represent viral sites targeted directly by host factors such as capsid residues targeted by Fv1 or TRIM5, or restriction factor sites that interact with viral protein antagonists as illustrated by the inhibition of APOBEC3G by Vif or SAMHD1 by Vpx (Figure 5).
There are many restriction factors that have unexpectedly long historical records of intense diversifying selection, and this argues for broader defensive roles, roles that for many of these factors are now known to extend well beyond their initially defined retroviral targets. Thus, Fv1 was initially defined as a Mus-specific factor restricting MLVs but is now known to have inhibitory activity against retroviruses of different genera such as feline foamy virus and some lentiviruses [62,81]. More significantly, many of the restriction factors profiled here like SERINC5, SAMHD1, and BST2 that were originally identified as specifically antiretroviral, show similar inhibitory activity against other viruses including alphaviruses, HBV, and flaviviruses. For at least some of these factors, sites of evolutionary conflict do not necessarily coincide with sites responsible for antiretroviral activity or retroviral antagonist interactions indicating that these adaptations are likely due to challenges by a broader range of viruses. These exceptions include MX2 where positively selected sites are not the ones involved in capsid binding and SFLN11 where positive selection is unconnected to retroviral restriction. This suggests that these rapidly evolving segments result from engagements with multiple viruses.
The rapid evolutionary changes that produce more effective host defenses are not limited to the statistical increase in the rate of non-synonymous mutations, but can also be marked by more dramatic genome reconfigurations including gene duplications and diversifications within specific lineages as seen for TRIM5 which has one copy in human and one to eight copies in rodents, or in APOBEC3 with one gene in mice and seven paralogs in humans, each having different restriction profiles and tissue specific expression patterns [111,116,172]. Genomic insertions can also alter gene expression as shown for mouse Apobec3 and can lead to production of fusion proteins with antiviral activity like TRIMCyp [108,160]. Finally, recurrent gene conversions between related antiviral genes as shown for MX1 and MX2 can also contribute to their diversification [68].
While the level of restriction imposed by these factors varies from total blocks to smaller several fold decreases, even marginal inhibitions are amplified during the spreading infections that precede disease. The knowledge gained by studying the interaction between these host restriction factors and their viral targets helps us understand the barriers to cross-species viral transmission, and should also contribute to the development of novel therapeutics that either exploit the inhibitory impact of these restriction factors or target their viral antagonists.

Author Contributions

G.B. and C.A.K. conceived and drafted the manuscript. G.B. and C.A.K. reviewed and edited the manuscript. Both authors read and approved the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Intramural Research Program of the National Institute of Allergy and Infectious Diseases under grant number AI000300-38 to C.A.K.

Acknowledgments

The work was supported by the Intramural Research Program of the National Institute of Allergy and Infectious Diseases, NIH.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Van Valen, L. A new evolutionary law. Evol. Theory 1973, 1, 1–30. [Google Scholar]
  2. Daugherty, M.D.; Malik, H.S. Rules of engagement: Molecular insights from host-virus arms races. Annu. Rev. Genet. 2012, 46, 677–700. [Google Scholar] [CrossRef]
  3. Feng, Y.; Broder, C.C.; Kennedy, P.E.; Berger, E.A. HIV-1 entry cofactor: Functional cDNA cloning of a seven-transmembrane, G protein-coupled receptor. Science 1996, 272, 872–877. [Google Scholar] [CrossRef]
  4. Alkhatib, G.; Combadiere, C.; Broder, C.C.; Feng, Y.; Kennedy, P.E.; Murphy, P.M.; Berger, E.A. CC CKR5: A RANTES, MIP-1alpha, MIP-1beta receptor as a fusion cofactor for macrophage-tropic HIV-1. Science 1996, 272, 1955–1958. [Google Scholar] [CrossRef]
  5. Dalgleish, A.G.; Beverley, P.C.; Clapham, P.R.; Crawford, D.H.; Greaves, M.F.; Weiss, R.A. The CD4 (T4) antigen is an essential component of the receptor for the AIDS retrovirus. Nature 1984, 312, 763–767. [Google Scholar] [CrossRef]
  6. Gaud, G.; Lesourne, R.; Love, P.E. Regulatory mechanisms in T cell receptor signalling. Nat. Rev. Immunol. 2018, 18, 485–497. [Google Scholar] [CrossRef]
  7. Huang, C.C.; Lam, S.N.; Acharya, P.; Tang, M.; Xiang, S.H.; Hussan, S.S.; Stanfield, R.L.; Robinson, J.; Sodroski, J.; Wilson, I.A.; et al. Structures of the CCR5 N terminus and of a tyrosine-sulfated antibody with HIV-1 gp120 and CD4. Science 2007, 317, 1930–1934. [Google Scholar] [CrossRef] [Green Version]
  8. Willey, R.L.; Buckler-White, A.; Strebel, K. Sequences present in the cytoplasmic domain of CD4 are necessary and sufficient to confer sensitivity to the human immunodeficiency virus type 1 Vpu protein. J. Virol. 1994, 68, 1207–1212. [Google Scholar] [CrossRef] [Green Version]
  9. Grzesiek, S.; Stahl, S.J.; Wingfield, P.T.; Bax, A. The CD4 determinant for downregulation by HIV-1 Nef directly binds to Nef. Mapping of the Nef binding surface by NMR. Biochemistry 1996, 35, 10256–10261. [Google Scholar] [CrossRef]
  10. Ramirez, P.W.; Sharma, S.; Singh, R.; Stoneham, C.A.; Vollbrecht, T.; Guatelli, J. Plasma Membrane-Associated Restriction Factors and Their Counteraction by HIV-1 Accessory Proteins. Cells 2019, 8, 1020. [Google Scholar] [CrossRef] [Green Version]
  11. Bibollet-Ruche, F.; Russell, R.M.; Liu, W.; Stewart-Jones, G.B.E.; Sherrill-Mix, S.; Li, Y.; Learn, G.H.; Smith, A.G.; Gondim, M.V.P.; Plenderleith, L.J.; et al. CD4 receptor diversity in chimpanzees protects against SIV infection. Proc. Natl. Acad. Sci. USA 2019, 116, 3229–3238. [Google Scholar] [CrossRef] [Green Version]
  12. Hvilsom, C.; Carlsen, F.; Siegismund, H.R.; Corbet, S.; Nerrienet, E.; Fomsgaard, A. Genetic subspecies diversity of the chimpanzee CD4 virus-receptor gene. Genomics 2008, 92, 322–328. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Z.D.; Weinstock, G.; Gerstein, M. Rapid evolution by positive Darwinian selection in T-cell antigen CD4 in primates. J. Mol. Evol. 2008, 66, 446–456. [Google Scholar] [CrossRef] [PubMed]
  14. Van der Lee, R.; Wiel, L.; van Dam, T.J.P.; Huynen, M.A. Genome-scale detection of positive selection in nine primates predicts human-virus evolutionary conflicts. Nucleic Acids Res. 2017, 45, 10634–10648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Samson, M.; Libert, F.; Doranz, B.J.; Rucker, J.; Liesnard, C.; Farber, C.M.; Saragosti, S.; Lapoumeroulie, C.; Cognaux, J.; Forceille, C.; et al. Resistance to HIV-1 infection in caucasian individuals bearing mutant alleles of the CCR-5 chemokine receptor gene. Nature 1996, 382, 722–725. [Google Scholar] [CrossRef] [PubMed]
  16. Kozak, C.A. The mouse “xenotropic” gammaretroviruses and their XPR1 receptor. Retrovirology 2010, 7, 101. [Google Scholar] [CrossRef] [Green Version]
  17. Kozak, C.A. Naturally Occurring Polymorphisms of the Mouse Gammaretrovirus Receptors CAT-1 and XPR1 Alter Virus Tropism and Pathogenicity. Adv. Virol. 2011, 2011, 975801. [Google Scholar] [CrossRef] [Green Version]
  18. Martin, C.; Buckler-White, A.; Wollenberg, K.; Kozak, C.A. The avian XPR1 gammaretrovirus receptor is under positive selection and is disabled in bird species in contact with virus-infected wild mice. J. Virol. 2013, 87, 10094–10104. [Google Scholar] [CrossRef] [Green Version]
  19. Yan, Y.; Liu, Q.; Wollenberg, K.; Martin, C.; Buckler-White, A.; Kozak, C.A. Evolution of functional and sequence variants of the mammalian XPR1 receptor for mouse xenotropic gammaretroviruses and the human-derived XMRV. J. Virol. 2010, 84, 11970–11980. [Google Scholar] [CrossRef] [Green Version]
  20. Albritton, L.M.; Tseng, L.; Scadden, D.; Cunningham, J.M. A putative murine ecotropic retrovirus receptor gene encodes a multiple membrane-spanning protein and confers susceptibility to virus infection. Cell 1989, 57, 659–666. [Google Scholar] [CrossRef]
  21. Tailor, C.S.; Nouri, A.; Lee, C.G.; Kozak, C.; Kabat, D. Cloning and characterization of a cell surface receptor for xenotropic and polytropic murine leukemia viruses. Proc. Natl. Acad. Sci. USA 1999, 96, 927–932. [Google Scholar] [CrossRef] [Green Version]
  22. Kim, J.W.; Closs, E.I.; Albritton, L.M.; Cunningham, J.M. Transport of cationic amino acids by the mouse ecotropic retrovirus receptor. Nature 1991, 352, 725–728. [Google Scholar] [CrossRef]
  23. Wang, H.; Kavanaugh, M.P.; North, R.A.; Kabat, D. Cell-surface receptor for ecotropic murine retroviruses is a basic amino-acid transporter. Nature 1991, 352, 729–731. [Google Scholar] [CrossRef]
  24. Giovannini, D.; Touhami, J.; Charnet, P.; Sitbon, M.; Battini, J.L. Inorganic phosphate export by the retrovirus receptor XPR1 in metazoans. Cell Rep. 2013, 3, 1866–1873. [Google Scholar] [CrossRef] [Green Version]
  25. Kozak, C.A.; O’Neill, R.R. Diverse wild mouse origins of xenotropic, mink cell focus-forming, and two types of ecotropic proviral genes. J. Virol. 1987, 61, 3082–3088. [Google Scholar] [CrossRef] [Green Version]
  26. Eiden, M.V.; Farrell, K.; Warsowe, J.; Mahan, L.C.; Wilson, C.A. Characterization of a naturally occurring ecotropic receptor that does not facilitate entry of all ecotropic murine retroviruses. J. Virol. 1993, 67, 4056–4061. [Google Scholar] [CrossRef] [Green Version]
  27. Lu, X.; Kassner, J.; Skorski, M.; Carley, S.; Shaffer, E.; Kozak, C.A. Mutational analysis and glycosylation sensitivity of restrictive XPR1 gammaretrovirus receptors in six mammalian species. Virology 2019, 535, 154–161. [Google Scholar] [CrossRef]
  28. Kavanaugh, M.P.; Miller, D.G.; Zhang, W.; Law, W.; Kozak, S.L.; Kabat, D.; Miller, A.D. Cell-surface receptors for gibbon ape leukemia virus and amphotropic murine retrovirus are inducible sodium-dependent phosphate symporters. Proc. Natl. Acad. Sci. USA 1994, 91, 7071–7075. [Google Scholar] [CrossRef] [Green Version]
  29. Miller, D.G.; Edwards, R.H.; Miller, A.D. Cloning of the cellular receptor for amphotropic murine retroviruses reveals homology to that for gibbon ape leukemia virus. Proc. Natl. Acad. Sci. USA 1994, 91, 78–82. [Google Scholar] [CrossRef] [Green Version]
  30. Hartley, J.W.; Rowe, W.P. Naturally occurring murine leukemia viruses in wild mice: Characterization of a new “amphotropic” class. J. Virol. 1976, 19, 19–25. [Google Scholar] [CrossRef] [Green Version]
  31. Rasheed, S.; Gardner, M.B.; Chan, E. Amphotropic host range of naturally occuring wild mouse leukemia viruses. J. Virol. 1976, 19, 13–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. O’Neill, R.R.; Hartley, J.W.; Repaske, R.; Kozak, C.A. Amphotropic proviral envelope sequences are absent from the Mus germ line. J. Virol. 1987, 61, 2225–2231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Johnson, W.E. Origins and evolutionary consequences of ancient endogenous retroviruses. Nat. Rev. Microbiol. 2019, 17, 355–370. [Google Scholar] [CrossRef] [PubMed]
  34. Taylor, G.M.; Gao, Y.; Sanders, D.A. Fv-4: Identification of the defect in Env and the mechanism of resistance to ecotropic murine leukemia virus. J. Virol. 2001, 75, 11244–11248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Inuzuka, M.; Hayakawa, M.; Ingi, T. Serinc, an activity-regulated protein family, incorporates serine into membrane lipid synthesis. J. Biol. Chem. 2005, 280, 35776–35783. [Google Scholar] [CrossRef]
  36. Rosa, A.; Chande, A.; Ziglio, S.; De Sanctis, V.; Bertorelli, R.; Goh, S.L.; McCauley, S.M.; Nowosielska, A.; Antonarakis, S.E.; Luban, J.; et al. HIV-1 Nef promotes infection by excluding SERINC5 from virion incorporation. Nature 2015, 526, 212–217. [Google Scholar] [CrossRef] [Green Version]
  37. Usami, Y.; Wu, Y.; Gottlinger, H.G. SERINC3 and SERINC5 restrict HIV-1 infectivity and are counteracted by Nef. Nature 2015, 526, 218–223. [Google Scholar] [CrossRef] [Green Version]
  38. Sharma, S.; Lewinski, M.K.; Guatelli, J. An N-Glycosylated Form of SERINC5 Is Specifically Incorporated into HIV-1 Virions. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [Green Version]
  39. Shi, J.; Xiong, R.; Zhou, T.; Su, P.; Zhang, X.; Qiu, X.; Li, H.; Li, S.; Yu, C.; Wang, B.; et al. HIV-1 Nef Antagonizes SERINC5 Restriction by Downregulation of SERINC5 via the Endosome/Lysosome System. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [Green Version]
  40. Ahmad, I.; Li, S.; Li, R.; Chai, Q.; Zhang, L.; Wang, B.; Yu, C.; Zheng, Y.H. The retroviral accessory proteins S2, Nef, and glycoMA use similar mechanisms for antagonizing the host restriction factor SERINC5. J. Biol. Chem. 2019, 294, 7013–7024. [Google Scholar] [CrossRef]
  41. Kmiec, D.; Akbil, B.; Ananth, S.; Hotter, D.; Sparrer, K.M.J.; Stürzel, C.M.; Trautz, B.; Ayouba, A.; Peeters, M.; Yao, Z.; et al. SIVcol Nef counteracts SERINC5 by promoting its proteasomal degradation but does not efficiently enhance HIV-1 replication in human CD4+ T cells and lymphoid tissue. PLoS Pathog. 2018, 14, e1007269. [Google Scholar] [CrossRef] [PubMed]
  42. Staudt, R.P.; Smithgall, T.E. Nef Homodimers Downregulate SERINC5 by AP-2-Mediated Endocytosis to Promote HIV-1 Infectivity. J. Biol. Chem. 2020, 295, 15540–15552. [Google Scholar] [CrossRef] [PubMed]
  43. Dai, W.; Usami, Y.; Wu, Y.; Gottlinger, H. A Long Cytoplasmic Loop Governs the Sensitivity of the Anti-viral Host Protein SERINC5 to HIV-1 Nef. Cell Rep. 2018, 22, 869–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Stoneham, C.A.; Ramirez, P.W.; Singh, R.; Suarez, M.; Debray, A.; Lim, C.; Jia, X.; Xiong, Y.; Guatelli, J. A Conserved Acidic-Cluster Motif in SERINC5 Confers Partial Resistance to Antagonism by HIV-1 Nef. J. Virol. 2020, 94. [Google Scholar] [CrossRef]
  45. Chande, A.; Cuccurullo, E.C.; Rosa, A.; Ziglio, S.; Carpenter, S.; Pizzato, M. S2 from equine infectious anemia virus is an infectivity factor which counteracts the retroviral inhibitors SERINC5 and SERINC3. Proc. Natl. Acad. Sci. USA 2016, 113, 13197–13202. [Google Scholar] [CrossRef] [Green Version]
  46. Ahi, Y.S.; Zhang, S.; Thappeta, Y.; Denman, A.; Feizpour, A.; Gummuluru, S.; Reinhard, B.; Muriaux, D.; Fivash, M.J.; Rein, A. Functional Interplay Between Murine Leukemia Virus Glycogag, Serinc5, and Surface Glycoprotein Governs Virus Entry, with Opposite Effects on Gammaretroviral and Ebolavirus Glycoproteins. mBio 2016, 7, e01985-16. [Google Scholar] [CrossRef] [Green Version]
  47. Li, S.; Ahmad, I.; Shi, J.; Wang, B.; Yu, C.; Zhang, L.; Zheng, Y.H. Murine Leukemia Virus Glycosylated Gag Reduces Murine SERINC5 Protein Expression at Steady-State Levels via the Endosome/Lysosome Pathway to Counteract SERINC5 Antiretroviral Activity. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [Green Version]
  48. Beitari, S.; Pan, Q.; Finzi, A.; Liang, C. Differential pressures of SERINC5 and IFITM3 on HIV-1 envelope glycoprotein over the course of HIV-1 infection. J. Virol. 2020. [Google Scholar] [CrossRef]
  49. Beitari, S.; Ding, S.; Pan, Q.; Finzi, A.; Liang, C. Effect of HIV-1 Env on SERINC5 Antagonism. J. Virol. 2017, 91. [Google Scholar] [CrossRef] [Green Version]
  50. Schulte, B.; Selyutina, A.; Opp, S.; Herschhorn, A.; Sodroski, J.G.; Pizzato, M.; Diaz-Griffero, F. Localization to detergent-resistant membranes and HIV-1 core entry inhibition correlate with HIV-1 restriction by SERINC5. Virology 2018, 515, 52–65. [Google Scholar] [CrossRef]
  51. Sood, C.; Marin, M.; Chande, A.; Pizzato, M.; Melikyan, G.B. SERINC5 protein inhibits HIV-1 fusion pore formation by promoting functional inactivation of envelope glycoproteins. J. Biol. Chem. 2017, 292, 6014–6026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Chen, Y.C.; Sood, C.; Marin, M.; Aaron, J.; Gratton, E.; Salaita, K.; Melikyan, G.B. Super-Resolution Fluorescence Imaging Reveals That Serine Incorporator Protein 5 Inhibits Human Immunodeficiency Virus Fusion by Disrupting Envelope Glycoprotein Clusters. ACS Nano 2020, 14, 10929–10943. [Google Scholar] [CrossRef] [PubMed]
  53. Heigele, A.; Kmiec, D.; Regensburger, K.; Langer, S.; Peiffer, L.; Sturzel, C.M.; Sauter, D.; Peeters, M.; Pizzato, M.; Learn, G.H.; et al. The Potency of Nef-Mediated SERINC5 Antagonism Correlates with the Prevalence of Primate Lentiviruses in the Wild. Cell Host Microbe 2016, 20, 381–391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. De Sousa-Pereira, P.; Abrantes, J.; Bauernfried, S.; Pierini, V.; Esteves, P.J.; Keppler, O.T.; Pizzato, M.; Hornung, V.; Fackler, O.T.; Baldauf, H.M. The antiviral activity of rodent and lagomorph SERINC3 and SERINC5 is counteracted by known viral antagonists. J. Gen. Virol. 2019, 100, 278–288. [Google Scholar] [CrossRef]
  55. Timilsina, U.; Umthong, S.; Lynch, B.; Stablewski, A.; Stavrou, S. SERINC5 Potently Restricts Retrovirus Infection In Vivo. mBio 2020, 11. [Google Scholar] [CrossRef]
  56. Murrell, B.; Vollbrecht, T.; Guatelli, J.; Wertheim, J.O. The Evolutionary Histories of Antiretroviral Proteins SERINC3 and SERINC5 Do Not Support an Evolutionary Arms Race in Primates. J. Virol. 2016, 90, 8085–8089. [Google Scholar] [CrossRef] [Green Version]
  57. Lilly, F. Susceptibility to two strains of Friend leukemia virus in mice. Science 1967, 155, 461–462. [Google Scholar] [CrossRef]
  58. Hartley, J.W.; Rowe, W.P.; Huebner, R.J. Host-range restrictions of murine leukemia viruses in mouse embryo cell cultures. J. Virol. 1970, 5, 221–225. [Google Scholar] [CrossRef] [Green Version]
  59. Rowe, W.P. Studies of genetic transmission of murine leukemia virus by AKR mice. I. Crosses with Fv-1 n strains of mice. J. Exp. Med. 1972, 136, 1272–1285. [Google Scholar] [CrossRef] [Green Version]
  60. Jung, Y.T.; Kozak, C.A. A single amino acid change in the murine leukemia virus capsid gene responsible for the Fv1(nr) phenotype. J. Virol. 2000, 74, 5385–5387. [Google Scholar] [CrossRef]
  61. Kozak, C.A.; Chakraborti, A. Single amino acid changes in the murine leukemia virus capsid protein gene define the target of Fv1 resistance. Virology 1996, 225, 300–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Best, S.; Le Tissier, P.; Towers, G.; Stoye, J.P. Positional cloning of the mouse retrovirus restriction gene Fv1. Nature 1996, 382, 826–829. [Google Scholar] [CrossRef] [PubMed]
  63. Benit, L.; De Parseval, N.; Casella, J.F.; Callebaut, I.; Cordonnier, A.; Heidmann, T. Cloning of a new murine endogenous retrovirus, MuERV-L, with strong similarity to the human HERV-L element and with a gag coding sequence closely related to the Fv1 restriction gene. J. Virol. 1997, 71, 5652–5657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Boso, G.; Buckler-White, A.; Kozak, C.A. Ancient Evolutionary Origin and Positive Selection of the Retroviral Restriction Factor Fv1 in Muroid Rodents. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Sawyer, S.L.; Wu, L.I.; Emerman, M.; Malik, H.S. Positive selection of primate TRIM5alpha identifies a critical species-specific retroviral restriction domain. Proc. Natl. Acad. Sci. USA 2005, 102, 2832–2837. [Google Scholar] [CrossRef] [Green Version]
  66. Sawyer, S.L.; Emerman, M.; Malik, H.S. Ancient adaptive evolution of the primate antiviral DNA-editing enzyme APOBEC3G. PLoS Biol. 2004, 2, E275. [Google Scholar] [CrossRef]
  67. Laguette, N.; Rahm, N.; Sobhian, B.; Chable-Bessia, C.; Munch, J.; Snoeck, J.; Sauter, D.; Switzer, W.M.; Heneine, W.; Kirchhoff, F.; et al. Evolutionary and functional analyses of the interaction between the myeloid restriction factor SAMHD1 and the lentiviral Vpx protein. Cell Host Microbe 2012, 11, 205–217. [Google Scholar] [CrossRef] [Green Version]
  68. Mitchell, P.S.; Young, J.M.; Emerman, M.; Malik, H.S. Evolutionary Analyses Suggest a Function of MxB Immunity Proteins Beyond Lentivirus Restriction. PLoS Pathog. 2015, 11, e1005304. [Google Scholar] [CrossRef]
  69. Sebastian, S.; Luban, J. TRIM5alpha selectively binds a restriction-sensitive retroviral capsid. Retrovirology 2005, 2, 40. [Google Scholar] [CrossRef] [Green Version]
  70. Diaz-Griffero, F.; Vandegraaff, N.; Li, Y.; McGee-Estrada, K.; Stremlau, M.; Welikala, S.; Si, Z.; Engelman, A.; Sodroski, J. Requirements for capsid-binding and an effector function in TRIMCyp-mediated restriction of HIV-1. Virology 2006, 351, 404–419. [Google Scholar] [CrossRef] [Green Version]
  71. Kouno, T.; Luengas, E.M.; Shigematsu, M.; Shandilya, S.M.; Zhang, J.; Chen, L.; Hara, M.; Schiffer, C.A.; Harris, R.S.; Matsuo, H. Structure of the Vif-binding domain of the antiviral enzyme APOBEC3G. Nat. Struct. Mol. Biol. 2015, 22, 485–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Betancor, G.; Dicks, M.D.J.; Jimenez-Guardeno, J.M.; Ali, N.H.; Apolonia, L.; Malim, M.H. The GTPase Domain of MX2 Interacts with the HIV-1 Capsid, Enabling Its Short Isoform to Moderate Antiviral Restriction. Cell Rep. 2019, 29, 1923–1933.e1923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Goujon, C.; Greenbury, R.A.; Papaioannou, S.; Doyle, T.; Malim, M.H. A triple-arginine motif in the amino-terminal domain and oligomerization are required for HIV-1 inhibition by human MX2. J. Virol. 2015, 89, 4676–4680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ahn, J.; Hao, C.; Yan, J.; DeLucia, M.; Mehrens, J.; Wang, C.; Gronenborn, A.M.; Skowronski, J. HIV/simian immunodeficiency virus (SIV) accessory virulence factor Vpx loads the host cell restriction factor SAMHD1 onto the E3 ubiquitin ligase complex CRL4DCAF1. J. Biol. Chem. 2012, 287, 12550–12558. [Google Scholar] [CrossRef] [Green Version]
  75. Jolicoeur, P.; Baltimore, D. Effect of Fv-1 gene product on proviral DNA formation and integration in cells infected with murine leukemia viruses. Proc. Natl. Acad. Sci. USA 1976, 73, 2236–2240. [Google Scholar] [CrossRef] [Green Version]
  76. Li, W.; Yap, M.W.; Voss, V.; Stoye, J.P. Expression levels of Fv1: Effects on retroviral restriction specificities. Retrovirology 2016, 13, 42. [Google Scholar] [CrossRef] [Green Version]
  77. Stevens, A.; Bock, M.; Ellis, S.; LeTissier, P.; Bishop, K.N.; Yap, M.W.; Taylor, W.; Stoye, J.P. Retroviral capsid determinants of Fv1 NB and NR tropism. J. Virol. 2004, 78, 9592–9598. [Google Scholar] [CrossRef] [Green Version]
  78. Bishop, K.N.; Bock, M.; Towers, G.; Stoye, J.P. Identification of the regions of Fv1 necessary for murine leukemia virus restriction. J. Virol. 2001, 75, 5182–5188. [Google Scholar] [CrossRef] [Green Version]
  79. Hilditch, L.; Matadeen, R.; Goldstone, D.C.; Rosenthal, P.B.; Taylor, I.A.; Stoye, J.P. Ordered assembly of murine leukemia virus capsid protein on lipid nanotubes directs specific binding by the restriction factor, Fv1. Proc. Natl. Acad. Sci. USA 2011, 108, 5771–5776. [Google Scholar] [CrossRef] [Green Version]
  80. Yan, Y.; Buckler-White, A.; Wollenberg, K.; Kozak, C.A. Origin, antiviral function and evidence for positive selection of the gammaretrovirus restriction gene Fv1 in the genus Mus. Proc. Natl. Acad. Sci. USA 2009, 106, 3259–3263. [Google Scholar] [CrossRef] [Green Version]
  81. Yap, M.W.; Colbeck, E.; Ellis, S.A.; Stoye, J.P. Evolution of the retroviral restriction gene Fv1: Inhibition of non-MLV retroviruses. PLoS Pathog. 2014, 10, e1003968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Young, G.R.; Yap, M.W.; Michaux, J.R.; Steppan, S.J.; Stoye, J.P. Evolutionary journey of the retroviral restriction gene Fv1. Proc. Natl. Acad. Sci. USA 2018, 115, 10130–10135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Steppan, S.J.; Schenk, J.J. Muroid rodent phylogenetics: 900-species tree reveals increasing diversification rates. PLoS ONE 2017, 12, e0183070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Towers, G.; Bock, M.; Martin, S.; Takeuchi, Y.; Stoye, J.P.; Danos, O. A conserved mechanism of retrovirus restriction in mammals. Proc. Natl. Acad. Sci. USA 2000, 97, 12295–12299. [Google Scholar] [CrossRef] [Green Version]
  85. Hatziioannou, T.; Cowan, S.; Goff, S.P.; Bieniasz, P.D.; Towers, G.J. Restriction of multiple divergent retroviruses by Lv1 and Ref1. EMBO J. 2003, 22, 385–394. [Google Scholar] [CrossRef] [Green Version]
  86. Stremlau, M.; Owens, C.M.; Perron, M.J.; Kiessling, M.; Autissier, P.; Sodroski, J. The cytoplasmic body component TRIM5alpha restricts HIV-1 infection in Old World monkeys. Nature 2004, 427, 848–853. [Google Scholar] [CrossRef]
  87. Hatakeyama, S. TRIM Family Proteins: Roles in Autophagy, Immunity, and Carcinogenesis. Trends Biochem. Sci. 2017, 42, 297–311. [Google Scholar] [CrossRef]
  88. Van Gent, M.; Sparrer, K.M.J.; Gack, M.U. TRIM Proteins and Their Roles in Antiviral Host Defenses. Annu. Rev. Virol. 2018, 5, 385–405. [Google Scholar] [CrossRef]
  89. Sardiello, M.; Cairo, S.; Fontanella, B.; Ballabio, A.; Meroni, G. Genomic analysis of the TRIM family reveals two groups of genes with distinct evolutionary properties. BMC Evol. Biol. 2008, 8, 225. [Google Scholar] [CrossRef] [Green Version]
  90. Hatziioannou, T.; Perez-Caballero, D.; Yang, A.; Cowan, S.; Bieniasz, P.D. Retrovirus resistance factors Ref1 and Lv1 are species-specific variants of TRIM5alpha. Proc. Natl. Acad. Sci. USA 2004, 101, 10774–10779. [Google Scholar] [CrossRef] [Green Version]
  91. Yap, M.W.; Nisole, S.; Stoye, J.P. A single amino acid change in the SPRY domain of human Trim5alpha leads to HIV-1 restriction. Curr. Biol. 2005, 15, 73–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Black, L.R.; Aiken, C. TRIM5alpha disrupts the structure of assembled HIV-1 capsid complexes in vitro. J. Virol. 2010, 84, 6564–6569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ganser-Pornillos, B.K.; Chandrasekaran, V.; Pornillos, O.; Sodroski, J.G.; Sundquist, W.I.; Yeager, M. Hexagonal assembly of a restricting TRIM5alpha protein. Proc. Natl. Acad. Sci. USA 2011, 108, 534–539. [Google Scholar] [CrossRef] [Green Version]
  94. Li, Y.L.; Chandrasekaran, V.; Carter, S.D.; Woodward, C.L.; Christensen, D.E.; Dryden, K.A.; Pornillos, O.; Yeager, M.; Ganser-Pornillos, B.K.; Jensen, G.J.; et al. Primate TRIM5 proteins form hexagonal nets on HIV-1 capsids. eLife 2016, 5, e16269. [Google Scholar] [CrossRef] [PubMed]
  95. Wagner, J.M.; Roganowicz, M.D.; Skorupka, K.; Alam, S.L.; Christensen, D.; Doss, G.; Wan, Y.; Frank, G.A.; Ganser-Pornillos, B.K.; Sundquist, W.I.; et al. Mechanism of B-box 2 domain-mediated higher-order assembly of the retroviral restriction factor TRIM5alpha. eLife 2016, 5, e16309. [Google Scholar] [CrossRef] [PubMed]
  96. Roganowicz, M.D.; Komurlu, S.; Mukherjee, S.; Plewka, J.; Alam, S.L.; Skorupka, K.A.; Wan, Y.; Dawidowski, D.; Cafiso, D.S.; Ganser-Pornillos, B.K.; et al. TRIM5alpha SPRY/coiled-coil interactions optimize avid retroviral capsid recognition. PLoS Pathog. 2017, 13, e1006686. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Morger, D.; Zosel, F.; Bühlmann, M.; Züger, S.; Mittelviefhaus, M.; Schuler, B.; Luban, J.; Grütter, M.G. The Three-Fold Axis of the HIV-1 Capsid Lattice Is the Species-Specific Binding Interface for TRIM5α. J. Virol. 2018, 92, e01541–e01617. [Google Scholar] [CrossRef] [Green Version]
  98. Stremlau, M.; Perron, M.; Welikala, S.; Sodroski, J. Species-specific variation in the B30.2(SPRY) domain of TRIM5alpha determines the potency of human immunodeficiency virus restriction. J. Virol. 2005, 79, 3139–3145. [Google Scholar] [CrossRef] [Green Version]
  99. Kane, M.; Zang, T.M.; Rihn, S.J.; Zhang, F.; Kueck, T.; Alim, M.; Schoggins, J.; Rice, C.M.; Wilson, S.J.; Bieniasz, P.D. Identification of Interferon-Stimulated Genes with Antiretroviral Activity. Cell Host Microbe 2016, 20, 392–405. [Google Scholar] [CrossRef] [Green Version]
  100. OhAinle, M.; Helms, L.; Vermeire, J.; Roesch, F.; Humes, D.; Basom, R.; Delrow, J.J.; Overbaugh, J.; Emerman, M. A virus-packageable CRISPR screen identifies host factors mediating interferon inhibition of HIV. eLife 2018, 7, e39823. [Google Scholar] [CrossRef]
  101. Jimenez-Guardeno, J.M.; Apolonia, L.; Betancor, G.; Malim, M.H. Immunoproteasome activation enables human TRIM5alpha restriction of HIV-1. Nat. Microbiol. 2019, 4, 933–940. [Google Scholar] [CrossRef] [PubMed]
  102. Chiramel, A.I.; Meyerson, N.R.; McNally, K.L.; Broeckel, R.M.; Montoya, V.R.; Mendez-Solis, O.; Robertson, S.J.; Sturdevant, G.L.; Lubick, K.J.; Nair, V.; et al. TRIM5alpha Restricts Flavivirus Replication by Targeting the Viral Protease for Proteasomal Degradation. Cell Rep. 2019, 27, 3269–3283.e3266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Diaz-Griffero, F.; Li, X.; Javanbakht, H.; Song, B.; Welikala, S.; Stremlau, M.; Sodroski, J. Rapid turnover and polyubiquitylation of the retroviral restriction factor TRIM5. Virology 2006, 349, 300–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Campbell, E.M.; Weingart, J.; Sette, P.; Opp, S.; Sastri, J.; O’Connor, S.K.; Talley, S.; Diaz-Griffero, F.; Hirsch, V.; Bouamr, F. TRIM5α-Mediated Ubiquitin Chain Conjugation Is Required for Inhibition of HIV-1 Reverse Transcription and Capsid Destabilization. J. Virol. 2016, 90, 1849–1857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Kutluay, S.B.; Perez-Caballero, D.; Bieniasz, P.D. Fates of retroviral core components during unrestricted and TRIM5-restricted infection. PLoS Pathog. 2013, 9, e1003214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Fletcher, A.J.; Christensen, D.E.; Nelson, C.; Tan, C.P.; Schaller, T.; Lehner, P.J.; Sundquist, W.I.; Towers, G.J. TRIM5alpha requires Ube2W to anchor Lys63-linked ubiquitin chains and restrict reverse transcription. EMBO J. 2015, 34, 2078–2095. [Google Scholar] [CrossRef]
  107. Roa, A.; Hayashi, F.; Yang, Y.; Lienlaf, M.; Zhou, J.; Shi, J.; Watanabe, S.; Kigawa, T.; Yokoyama, S.; Aiken, C.; et al. RING domain mutations uncouple TRIM5alpha restriction of HIV-1 from inhibition of reverse transcription and acceleration of uncoating. J. Virol. 2012, 86, 1717–1727. [Google Scholar] [CrossRef] [Green Version]
  108. Sayah, D.M.; Sokolskaja, E.; Berthoux, L.; Luban, J. Cyclophilin A retrotransposition into TRIM5 explains owl monkey resistance to HIV-1. Nature 2004, 430, 569–573. [Google Scholar] [CrossRef]
  109. Si, Z.; Vandegraaff, N.; O’Huigin, C.; Song, B.; Yuan, W.; Xu, C.; Perron, M.; Li, X.; Marasco, W.A.; Engelman, A.; et al. Evolution of a cytoplasmic tripartite motif (TRIM) protein in cows that restricts retroviral infection. Proc. Natl. Acad. Sci. USA 2006, 103, 7454–7459. [Google Scholar] [CrossRef] [Green Version]
  110. Ylinen, L.M.; Keckesova, Z.; Webb, B.L.; Gifford, R.J.; Smith, T.P.; Towers, G.J. Isolation of an active Lv1 gene from cattle indicates that tripartite motif protein-mediated innate immunity to retroviral infection is widespread among mammals. J. Virol. 2006, 80, 7332–7338. [Google Scholar] [CrossRef] [Green Version]
  111. Sawyer, S.L.; Emerman, M.; Malik, H.S. Discordant evolution of the adjacent antiretroviral genes TRIM22 and TRIM5 in mammals. PLoS Pathog. 2007, 3, e197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Wilson, S.J.; Webb, B.L.; Ylinen, L.M.; Verschoor, E.; Heeney, J.L.; Towers, G.J. Independent evolution of an antiviral TRIMCyp in rhesus macaques. Proc. Natl. Acad. Sci. USA 2008, 105, 3557–3562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Tareen, S.U.; Sawyer, S.L.; Malik, H.S.; Emerman, M. An expanded clade of rodent Trim5 genes. Virology 2009, 385, 473–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Malfavon-Borja, R.; Wu, L.I.; Emerman, M.; Malik, H.S. Birth, decay, and reconstruction of an ancient TRIMCyp gene fusion in primate genomes. Proc. Natl. Acad. Sci. USA 2013, 110, E583–E592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Mu, D.; Yang, H.; Zhu, J.W.; Liu, F.L.; Tian, R.R.; Zheng, H.Y.; Han, J.B.; Shi, P.; Zheng, Y.T. Independent birth of a novel TRIMCyp in Tupaia belangeri with a divergent function from its paralog TRIM5. Mol. Biol. Evol. 2014, 31, 2985–2997. [Google Scholar] [CrossRef] [Green Version]
  116. Boso, G.; Shaffer, E.; Liu, Q.; Cavanna, K.; Buckler-White, A.; Kozak, C.A. Evolution of the rodent Trim5 cluster is marked by divergent paralogous expansions and independent acquisitions of TrimCyp fusions. Sci. Rep. 2019, 9, 11263. [Google Scholar] [CrossRef]
  117. Águeda-Pinto, A.; Lemos de Matos, A.; Pinheiro, A.; Neves, F.; de Sousa-Pereira, P.; Esteves, P.J. Not so unique to Primates: The independent adaptive evolution of TRIM5 in Lagomorpha lineage. PLoS ONE 2019, 14, e0226202. [Google Scholar] [CrossRef] [Green Version]
  118. Morrison, J.H.; Miller, C.; Bankers, L.; Crameri, G.; Wang, L.F.; Poeschla, E.M. A Potent Postentry Restriction to Primate Lentiviruses in a Yinpterochiropteran Bat. mBio 2020, 11. [Google Scholar] [CrossRef]
  119. Pertel, T.; Hausmann, S.; Morger, D.; Zuger, S.; Guerra, J.; Lascano, J.; Reinhard, C.; Santoni, F.A.; Uchil, P.D.; Chatel, L.; et al. TRIM5 is an innate immune sensor for the retrovirus capsid lattice. Nature 2011, 472, 361–365. [Google Scholar] [CrossRef] [Green Version]
  120. Lascano, J.; Uchil, P.D.; Mothes, W.; Luban, J. TRIM5 Retroviral Restriction Activity Correlates with the Ability to Induce Innate Immune Signaling. J. Virol. 2016, 90, 308–316. [Google Scholar] [CrossRef] [Green Version]
  121. Zhu, J.W.; Mu, D.; Liu, F.L.; Luo, M.T.; Luo, R.H.; Zheng, Y.T. Activation of NF-kappaB induced by TRIMCyp showing a discrepancy between owl monkey and northern pig-tailed macaque. Mol. Immunol. 2018, 101, 627–634. [Google Scholar] [CrossRef] [PubMed]
  122. Fletcher, A.J.; Vaysburd, M.; Maslen, S.; Zeng, J.; Skehel, J.M.; Towers, G.J.; James, L.C. Trivalent RING Assembly on Retroviral Capsids Activates TRIM5 Ubiquitination and Innate Immune Signaling. Cell Host Microbe 2018, 24, 761–775.e766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Salter, J.D.; Bennett, R.P.; Smith, H.C. The APOBEC Protein Family: United by Structure, Divergent in Function. Trends Biochem. Sci. 2016, 41, 578–594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sheehy, A.M.; Gaddis, N.C.; Choi, J.D.; Malim, M.H. Isolation of a human gene that inhibits HIV-1 infection and is suppressed by the viral Vif protein. Nature 2002, 418, 646–650. [Google Scholar] [CrossRef]
  125. Larue, R.S.; Lengyel, J.; Jonsson, S.R.; Andresdottir, V.; Harris, R.S. Lentiviral Vif degrades the APOBEC3Z3/APOBEC3H protein of its mammalian host and is capable of cross-species activity. J. Virol. 2010, 84, 8193–8201. [Google Scholar] [CrossRef] [Green Version]
  126. Yoshikawa, R.; Izumi, T.; Nakano, Y.; Yamada, E.; Moriwaki, M.; Misawa, N.; Ren, F.; Kobayashi, T.; Koyanagi, Y.; Sato, K. Small ruminant lentiviral Vif proteins commonly utilize cyclophilin A, an evolutionarily and structurally conserved protein, to degrade ovine and caprine APOBEC3 proteins. Microbiol. Immunol. 2016, 60, 427–436. [Google Scholar] [CrossRef]
  127. Su, X.; Wang, H.; Zhou, X.; Li, Z.; Zheng, B.; Zhang, W. Jembrana disease virus Vif antagonizes the inhibition of bovine APOBEC3 proteins through ubiquitin-mediate protein degradation. Virology 2018, 519, 53–63. [Google Scholar] [CrossRef]
  128. Zhao, Z.; Li, Z.; Huan, C.; Wang, H.; Su, X.; Zhang, W. CAEV Vif Hijacks ElonginB/C, CYPA and Cullin5 to Assemble the E3 Ubiquitin Ligase Complex Stepwise to Degrade oaA3Z2-Z3. Front. Microbiol. 2019, 10, 565. [Google Scholar] [CrossRef]
  129. Konno, Y.; Nagaoka, S.; Kimura, I.; Yamamoto, K.; Kagawa, Y.; Kumata, R.; Aso, H.; Ueda, M.T.; Nakagawa, S.; Kobayashi, T.; et al. New World feline APOBEC3 potently controls inter-genus lentiviral transmission. Retrovirology 2018, 15, 31. [Google Scholar] [CrossRef] [Green Version]
  130. Adolph, M.B.; Ara, A.; Feng, Y.; Wittkopp, C.J.; Emerman, M.; Fraser, J.S.; Chelico, L. Cytidine deaminase efficiency of the lentiviral viral restriction factor APOBEC3C correlates with dimerization. Nucleic Acids Res. 2017, 45, 3378–3394. [Google Scholar] [CrossRef]
  131. Nakano, Y.; Misawa, N.; Juarez-Fernandez, G.; Moriwaki, M.; Nakaoka, S.; Funo, T.; Yamada, E.; Soper, A.; Yoshikawa, R.; Ebrahimi, D.; et al. HIV-1 competition experiments in humanized mice show that APOBEC3H imposes selective pressure and promotes virus adaptation. PLoS Pathog. 2017, 13, e1006348. [Google Scholar] [CrossRef] [Green Version]
  132. Anderson, B.D.; Ikeda, T.; Moghadasi, S.A.; Martin, A.S.; Brown, W.L.; Harris, R.S. Natural APOBEC3C variants can elicit differential HIV-1 restriction activity. Retrovirology 2018, 15, 78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Mohammadzadeh, N.; Follack, T.B.; Love, R.P.; Stewart, K.; Sanche, S.; Chelico, L. Polymorphisms of the cytidine deaminase APOBEC3F have different HIV-1 restriction efficiencies. Virology 2019, 527, 21–31. [Google Scholar] [CrossRef] [PubMed]
  134. Harris, R.S.; Bishop, K.N.; Sheehy, A.M.; Craig, H.M.; Petersen-Mahrt, S.K.; Watt, I.N.; Neuberger, M.S.; Malim, M.H. DNA deamination mediates innate immunity to retroviral infection. Cell 2003, 113, 803–809. [Google Scholar] [CrossRef] [Green Version]
  135. Mangeat, B.; Turelli, P.; Caron, G.; Friedli, M.; Perrin, L.; Trono, D. Broad antiretroviral defence by human APOBEC3G through lethal editing of nascent reverse transcripts. Nature 2003, 424, 99–103. [Google Scholar] [CrossRef]
  136. Iwatani, Y.; Chan, D.S.; Wang, F.; Maynard, K.S.; Sugiura, W.; Gronenborn, A.M.; Rouzina, I.; Williams, M.C.; Musier-Forsyth, K.; Levin, J.G. Deaminase-independent inhibition of HIV-1 reverse transcription by APOBEC3G. Nucleic Acids Res. 2007, 35, 7096–7108. [Google Scholar] [CrossRef]
  137. Wang, X.; Ao, Z.; Chen, L.; Kobinger, G.; Peng, J.; Yao, X. The cellular antiviral protein APOBEC3G interacts with HIV-1 reverse transcriptase and inhibits its function during viral replication. J. Virol. 2012, 86, 3777–3786. [Google Scholar] [CrossRef] [Green Version]
  138. Morse, M.; Huo, R.; Feng, Y.; Rouzina, I.; Chelico, L.; Williams, M.C. Dimerization regulates both deaminase-dependent and deaminase-independent HIV-1 restriction by APOBEC3G. Nat. Commun. 2017, 8, 597. [Google Scholar] [CrossRef] [Green Version]
  139. Takeda, E.; Tsuji-Kawahara, S.; Sakamoto, M.; Langlois, M.A.; Neuberger, M.S.; Rada, C.; Miyazawa, M. Mouse APOBEC3 restricts friend leukemia virus infection and pathogenesis in vivo. J. Virol. 2008, 82, 10998–11008. [Google Scholar] [CrossRef] [Green Version]
  140. Rulli, S.J.; Mirro, J.; Hill, S.A.; Lloyd, P.; Gorelick, R.J.; Coffin, J.M.; Derse, D.; Rein, A. Interactions of murine APOBEC3 and human APOBEC3G with murine leukemia viruses. J. Virol. 2008, 82, 6566–6575. [Google Scholar] [CrossRef] [Green Version]
  141. MacMillan, A.L.; Kohli, R.M.; Ross, S.R. APOBEC3 inhibition of mouse mammary tumor virus infection: The role of cytidine deamination versus inhibition of reverse transcription. J. Virol. 2013, 87, 4808–4817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Stavrou, S.; Zhao, W.; Blouch, K.; Ross, S.R. Deaminase-Dead Mouse APOBEC3 Is an In Vivo Retroviral Restriction Factor. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Stavrou, S.; Crawford, D.; Blouch, K.; Browne, E.P.; Kohli, R.M.; Ross, S.R. Different modes of retrovirus restriction by human APOBEC3A and APOBEC3G in vivo. PLoS Pathog. 2014, 10, e1004145. [Google Scholar] [CrossRef]
  144. Hakata, Y.; Li, J.; Fujino, T.; Tanaka, Y.; Shimizu, R.; Miyazawa, M. Mouse APOBEC3 interferes with autocatalytic cleavage of murine leukemia virus Pr180gag-pol precursor and inhibits Pr65gag processing. PLoS Pathog. 2019, 15, e1008173. [Google Scholar] [CrossRef] [Green Version]
  145. Li, J.; Hakata, Y.; Takeda, E.; Liu, Q.; Iwatani, Y.; Kozak, C.A.; Miyazawa, M. Two genetic determinants acquired late in Mus evolution regulate the inclusion of exon 5, which alters mouse APOBEC3 translation efficiency. PLoS Pathog. 2012, 8, e1002478. [Google Scholar] [CrossRef]
  146. Hultquist, J.F.; Lengyel, J.A.; Refsland, E.W.; LaRue, R.S.; Lackey, L.; Brown, W.L.; Harris, R.S. Human and rhesus APOBEC3D, APOBEC3F, APOBEC3G, and APOBEC3H demonstrate a conserved capacity to restrict Vif-deficient HIV-1. J. Virol. 2011, 85, 11220–11234. [Google Scholar] [CrossRef] [Green Version]
  147. Krisko, J.F.; Begum, N.; Baker, C.E.; Foster, J.L.; Garcia, J.V. APOBEC3G and APOBEC3F Act in Concert To Extinguish HIV-1 Replication. J. Virol. 2016, 90, 4681–4695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Wittkopp, C.J.; Adolph, M.B.; Wu, L.I.; Chelico, L.; Emerman, M. A Single Nucleotide Polymorphism in Human APOBEC3C Enhances Restriction of Lentiviruses. PLoS Pathog. 2016, 12, e1005865. [Google Scholar] [CrossRef] [PubMed]
  149. Ayyappan Jaguva Vasudevan, A.; Balakrishnan, K.; Gertzen, C.W.G.; Borveto, F.; Zhang, Z.; Sangwiman, A.; Held, U.; Kustermann, C.; Banerjee, S.; Schumann, G.G.; et al. Loop 1 of APOBEC3C regulates its antiviral activity against HIV-1. J. Mol. Biol. 2020. [Google Scholar] [CrossRef]
  150. Zheng, Y.H.; Irwin, D.; Kurosu, T.; Tokunaga, K.; Sata, T.; Peterlin, B.M. Human APOBEC3F is another host factor that blocks human immunodeficiency virus type 1 replication. J. Virol. 2004, 78, 6073–6076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Wiegand, H.L.; Doehle, B.P.; Bogerd, H.P.; Cullen, B.R. A second human antiretroviral factor, APOBEC3F, is suppressed by the HIV-1 and HIV-2 Vif proteins. EMBO J. 2004, 23, 2451–2458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Harari, A.; Ooms, M.; Mulder, L.C.; Simon, V. Polymorphisms and splice variants influence the antiretroviral activity of human APOBEC3H. J. Virol. 2009, 83, 295–303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Mulder, L.C.; Ooms, M.; Majdak, S.; Smedresman, J.; Linscheid, C.; Harari, A.; Kunz, A.; Simon, V. Moderate influence of human APOBEC3F on HIV-1 replication in primary lymphocytes. J. Virol. 2010, 84, 9613–9617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Wang, X.; Abudu, A.; Son, S.; Dang, Y.; Venta, P.J.; Zheng, Y.H. Analysis of human APOBEC3H haplotypes and anti-human immunodeficiency virus type 1 activity. J. Virol. 2011, 85, 3142–3152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Duggal, N.K.; Fu, W.; Akey, J.M.; Emerman, M. Identification and antiviral activity of common polymorphisms in the APOBEC3 locus in human populations. Virology 2013, 443, 329–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Chesarino, N.M.; Emerman, M. Polymorphisms in Human APOBEC3H Differentially Regulate Ubiquitination and Antiviral Activity. Viruses 2020, 12, 378. [Google Scholar] [CrossRef] [Green Version]
  157. Covino, D.A.; Gauzzi, M.C.; Fantuzzi, L. Understanding the regulation of APOBEC3 expression: Current evidence and much to learn. J. Leukoc. Biol. 2018, 103, 433–444. [Google Scholar] [CrossRef]
  158. Refsland, E.W.; Stenglein, M.D.; Shindo, K.; Albin, J.S.; Brown, W.L.; Harris, R.S. Quantitative profiling of the full APOBEC3 mRNA repertoire in lymphocytes and tissues: Implications for HIV-1 restriction. Nucleic Acids Res. 2010, 38, 4274–4284. [Google Scholar] [CrossRef] [Green Version]
  159. Koning, F.A.; Newman, E.N.; Kim, E.Y.; Kunstman, K.J.; Wolinsky, S.M.; Malim, M.H. Defining APOBEC3 expression patterns in human tissues and hematopoietic cell subsets. J. Virol. 2009, 83, 9474–9485. [Google Scholar] [CrossRef] [Green Version]
  160. Sanville, B.; Dolan, M.A.; Wollenberg, K.; Yan, Y.; Martin, C.; Yeung, M.L.; Strebel, K.; Buckler-White, A.; Kozak, C.A. Adaptive evolution of Mus Apobec3 includes retroviral insertion and positive selection at two clusters of residues flanking the substrate groove. PLoS Pathog. 2010, 6, e1000974. [Google Scholar] [CrossRef]
  161. Yu, X.; Yu, Y.; Liu, B.; Luo, K.; Kong, W.; Mao, P.; Yu, X.F. Induction of APOBEC3G ubiquitination and degradation by an HIV-1 Vif-Cul5-SCF complex. Science 2003, 302, 1056–1060. [Google Scholar] [CrossRef] [PubMed]
  162. Jager, S.; Kim, D.Y.; Hultquist, J.F.; Shindo, K.; LaRue, R.S.; Kwon, E.; Li, M.; Anderson, B.D.; Yen, L.; Stanley, D.; et al. Vif hijacks CBF-beta to degrade APOBEC3G and promote HIV-1 infection. Nature 2011, 481, 371–375. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, W.; Du, J.; Evans, S.L.; Yu, Y.; Yu, X.F. T-cell differentiation factor CBF-beta regulates HIV-1 Vif-mediated evasion of host restriction. Nature 2011, 481, 376–379. [Google Scholar] [CrossRef] [PubMed]
  164. Hu, Y.; Desimmie, B.A.; Nguyen, H.C.; Ziegler, S.J.; Cheng, T.C.; Chen, J.; Wang, J.; Wang, H.; Zhang, K.; Pathak, V.K.; et al. Structural basis of antagonism of human APOBEC3F by HIV-1 Vif. Nat. Struct. Mol. Biol. 2019, 26, 1176–1183. [Google Scholar] [CrossRef] [PubMed]
  165. Miyagi, E.; Welbourn, S.; Sukegawa, S.; Fabryova, H.; Kao, S.; Strebel, K. Inhibition of Vif-Mediated Degradation of APOBEC3G through Competitive Binding of Core-Binding Factor Beta. J. Virol. 2020, 94. [Google Scholar] [CrossRef] [PubMed]
  166. Kolokithas, A.; Rosenke, K.; Malik, F.; Hendrick, D.; Swanson, L.; Santiago, M.L.; Portis, J.L.; Hasenkrug, K.J.; Evans, L.H. The glycosylated Gag protein of a murine leukemia virus inhibits the antiretroviral function of APOBEC3. J. Virol. 2010, 84, 10933–10936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Stavrou, S.; Nitta, T.; Kotla, S.; Ha, D.; Nagashima, K.; Rein, A.R.; Fan, H.; Ross, S.R. Murine leukemia virus glycosylated Gag blocks apolipoprotein B editing complex 3 and cytosolic sensor access to the reverse transcription complex. Proc. Natl. Acad. Sci. USA 2013, 110, 9078–9083. [Google Scholar] [CrossRef] [Green Version]
  168. Zhao, W.; Akkawi, C.; Mougel, M.; Ross, S.R. Murine Leukemia Virus P50 Protein Counteracts APOBEC3 by Blocking Its Packaging. J. Virol. 2020, 94. [Google Scholar] [CrossRef]
  169. Houzet, L.; Battini, J.L.; Bernard, E.; Thibert, V.; Mougel, M. A new retroelement constituted by a natural alternatively spliced RNA of murine replication-competent retroviruses. EMBO J. 2003, 22, 4866–4875. [Google Scholar] [CrossRef] [Green Version]
  170. LaRue, R.S.; Jonsson, S.R.; Silverstein, K.A.; Lajoie, M.; Bertrand, D.; El-Mabrouk, N.; Hotzel, I.; Andresdottir, V.; Smith, T.P.; Harris, R.S. The artiodactyl APOBEC3 innate immune repertoire shows evidence for a multi-functional domain organization that existed in the ancestor of placental mammals. BMC Mol. Biol. 2008, 9, 104. [Google Scholar] [CrossRef] [Green Version]
  171. Hirano, M. Evolution of vertebrate adaptive immunity: Immune cells and tissues, and AID/APOBEC cytidine deaminases. Bioessays 2015, 37, 877–887. [Google Scholar] [CrossRef] [PubMed]
  172. Munk, C.; Willemsen, A.; Bravo, I.G. An ancient history of gene duplications, fusions and losses in the evolution of APOBEC3 mutators in mammals. BMC Evol. Biol. 2012, 12, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Yang, L.; Emerman, M.; Malik, H.S.; McLaughlin, R.N.J. Retrocopying expands the functional repertoire of APOBEC3 antiviral proteins in primates. eLife 2020, 9, e58436. [Google Scholar] [CrossRef] [PubMed]
  174. LaRue, R.S.; Andresdottir, V.; Blanchard, Y.; Conticello, S.G.; Derse, D.; Emerman, M.; Greene, W.C.; Jonsson, S.R.; Landau, N.R.; Lochelt, M.; et al. Guidelines for naming nonprimate APOBEC3 genes and proteins. J. Virol. 2009, 83, 494–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Ito, J.; Gifford, R.J.; Sato, K. Retroviruses drive the rapid evolution of mammalian APOBEC3 genes. Proc. Natl. Acad. Sci. USA 2020, 117, 610–618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Jern, P.; Stoye, J.P.; Coffin, J.M. Role of APOBEC3 in genetic diversity among endogenous murine leukemia viruses. PLoS Genet. 2007, 3, 2014–2022. [Google Scholar] [CrossRef] [Green Version]
  177. Knisbacher, B.A.; Levanon, E.Y. DNA Editing of LTR Retrotransposons Reveals the Impact of APOBECs on Vertebrate Genomes. Mol. Biol. Evol. 2016, 33, 554–567. [Google Scholar] [CrossRef] [PubMed]
  178. Renner, T.M.; Belanger, K.; Goodwin, L.R.; Campbell, M.; Langlois, M.A. Characterization of molecular attributes that influence LINE-1 restriction by all seven human APOBEC3 proteins. Virology 2018, 520, 127–136. [Google Scholar] [CrossRef]
  179. Treger, R.S.; Tokuyama, M.; Dong, H.; Salas-Briceno, K.; Ross, S.R.; Kong, Y.; Iwasaki, A. Human APOBEC3G Prevents Emergence of Infectious Endogenous Retrovirus in Mice. J. Virol. 2019, 93, e00728–e00819. [Google Scholar] [CrossRef] [Green Version]
  180. Turelli, P.; Mangeat, B.; Jost, S.; Vianin, S.; Trono, D. Inhibition of hepatitis B virus replication by APOBEC3G. Science 2004, 303, 1829. [Google Scholar] [CrossRef]
  181. Chen, Z.; Eggerman, T.L.; Bocharov, A.V.; Baranova, I.N.; Vishnyakova, T.G.; Kurlander, R.; Patterson, A.P. Heat shock proteins stimulate APOBEC-3-mediated cytidine deamination in the hepatitis B virus. J. Biol. Chem. 2017, 292, 13459–13479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Kanagaraj, A.; Sakamoto, N.; Que, L.; Li, Y.; Mohiuddin, M.; Koura, M.; Wakae, K.; Kurachi, M.; Muramatsu, M.; Kitamura, K. Different antiviral activities of natural APOBEC3C, APOBEC3G, and APOBEC3H variants against hepatitis B virus. Biochem. Biophys. Res. Commun. 2019, 518, 26–31. [Google Scholar] [CrossRef] [PubMed]
  183. Goldstone, D.C.; Ennis-Adeniran, V.; Hedden, J.J.; Groom, H.C.; Rice, G.I.; Christodoulou, E.; Walker, P.A.; Kelly, G.; Haire, L.F.; Yap, M.W.; et al. HIV-1 restriction factor SAMHD1 is a deoxynucleoside triphosphate triphosphohydrolase. Nature 2011, 480, 379–382. [Google Scholar] [CrossRef] [PubMed]
  184. Laguette, N.; Sobhian, B.; Casartelli, N.; Ringeard, M.; Chable-Bessia, C.; Segeral, E.; Yatim, A.; Emiliani, S.; Schwartz, O.; Benkirane, M. SAMHD1 is the dendritic- and myeloid-cell-specific HIV-1 restriction factor counteracted by Vpx. Nature 2011, 474, 654–657. [Google Scholar] [CrossRef] [PubMed]
  185. Baldauf, H.M.; Pan, X.; Erikson, E.; Schmidt, S.; Daddacha, W.; Burggraf, M.; Schenkova, K.; Ambiel, I.; Wabnitz, G.; Gramberg, T.; et al. SAMHD1 restricts HIV-1 infection in resting CD4(+) T cells. Nat. Med. 2012, 18, 1682–1687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Hrecka, K.; Hao, C.; Gierszewska, M.; Swanson, S.K.; Kesik-Brodacka, M.; Srivastava, S.; Florens, L.; Washburn, M.P.; Skowronski, J. Vpx relieves inhibition of HIV-1 infection of macrophages mediated by the SAMHD1 protein. Nature 2011, 474, 658–661. [Google Scholar] [CrossRef] [Green Version]
  187. Schott, K.; Fuchs, N.V.; Derua, R.; Mahboubi, B.; Schnellbacher, E.; Seifried, J.; Tondera, C.; Schmitz, H.; Shepard, C.; Brandariz-Nunez, A.; et al. Dephosphorylation of the HIV-1 restriction factor SAMHD1 is mediated by PP2A-B55alpha holoenzymes during mitotic exit. Nat. Commun. 2018, 9, 2227. [Google Scholar] [CrossRef]
  188. Cribier, A.; Descours, B.; Valadao, A.L.; Laguette, N.; Benkirane, M. Phosphorylation of SAMHD1 by cyclin A2/CDK1 regulates its restriction activity toward HIV-1. Cell Rep. 2013, 3, 1036–1043. [Google Scholar] [CrossRef] [Green Version]
  189. White, T.E.; Brandariz-Nunez, A.; Valle-Casuso, J.C.; Amie, S.; Nguyen, L.A.; Kim, B.; Tuzova, M.; Diaz-Griffero, F. The retroviral restriction ability of SAMHD1, but not its deoxynucleotide triphosphohydrolase activity, is regulated by phosphorylation. Cell Host Microbe 2013, 13, 441–451. [Google Scholar] [CrossRef] [Green Version]
  190. Welbourn, S.; Dutta, S.M.; Semmes, O.J.; Strebel, K. Restriction of virus infection but not catalytic dNTPase activity is regulated by phosphorylation of SAMHD1. J. Virol. 2013, 87, 11516–11524. [Google Scholar] [CrossRef] [Green Version]
  191. Arnold, L.H.; Groom, H.C.; Kunzelmann, S.; Schwefel, D.; Caswell, S.J.; Ordonez, P.; Mann, M.C.; Rueschenbaum, S.; Goldstone, D.C.; Pennell, S.; et al. Phospho-dependent Regulation of SAMHD1 Oligomerisation Couples Catalysis and Restriction. PLoS Pathog. 2015, 11, e1005194. [Google Scholar] [CrossRef] [PubMed]
  192. Wang, F.; St Gelais, C.; de Silva, S.; Zhang, H.; Geng, Y.; Shepard, C.; Kim, B.; Yount, J.S.; Wu, L. Phosphorylation of mouse SAMHD1 regulates its restriction of human immunodeficiency virus type 1 infection, but not murine leukemia virus infection. Virology 2016, 487, 273–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Lim, E.S.; Fregoso, O.I.; McCoy, C.O.; Matsen, F.A.; Malik, H.S.; Emerman, M. The ability of primate lentiviruses to degrade the monocyte restriction factor SAMHD1 preceded the birth of the viral accessory protein Vpx. Cell Host Microbe 2012, 11, 194–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Fregoso, O.I.; Ahn, J.; Wang, C.; Mehrens, J.; Skowronski, J.; Emerman, M. Evolutionary toggling of Vpx/Vpr specificity results in divergent recognition of the restriction factor SAMHD1. PLoS Pathog. 2013, 9, e1003496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Guo, H.; Zhang, N.; Shen, S.; Yu, X.F.; Wei, W. Determinants of lentiviral Vpx-CRL4 E3 ligase-mediated SAMHD1 degradation in the substrate adaptor protein DCAF1. Biochem. Biophys. Res. Commun. 2019, 513, 933–939. [Google Scholar] [CrossRef]
  196. Hosmalin, A.; McIlroy, D.; Cheynier, R.; Clauvel, J.P.; Oksenhendler, E.; Wain-Hobson, S.; Debré, P.; Autran, B. Splenic interdigitating dendritic cells in humans: Characterization and HIV infection frequency in vivo. Adv. Exp. Med. Biol. 1995, 378, 439–441. [Google Scholar] [CrossRef]
  197. Banchereau, J.; Steinman, R.M. Dendritic cells and the control of immunity. Nature 1998, 392, 245–252. [Google Scholar] [CrossRef]
  198. Schmitz, J.E.; Korioth-Schmitz, B. Immunopathogenesis of simian immunodeficiency virus infection in nonhuman primates. Curr. Opin. HIV AIDS 2013, 8, 273–279. [Google Scholar] [CrossRef] [Green Version]
  199. Monit, C.; Morris, E.R.; Ruis, C.; Szafran, B.; Thiltgen, G.; Tsai, M.C.; Mitchison, N.A.; Bishop, K.N.; Stoye, J.P.; Taylor, I.A.; et al. Positive selection in dNTPase SAMHD1 throughout mammalian evolution. Proc. Natl. Acad. Sci. USA 2019, 116, 18647–18654. [Google Scholar] [CrossRef] [Green Version]
  200. Mereby, S.A.; Maehigashi, T.; Holler, J.M.; Kim, D.H.; Schinazi, R.F.; Kim, B. Interplay of ancestral non-primate lentiviruses with the virus-restricting SAMHD1 proteins of their hosts. J. Biol. Chem. 2018, 293, 16402–16412. [Google Scholar] [CrossRef] [Green Version]
  201. Wang, C.; Zhang, K.; Meng, L.; Zhang, X.; Song, Y.; Zhang, Y.; Gai, Y.; Zhang, Y.; Yu, B.; Wu, J.; et al. The C-terminal domain of feline and bovine SAMHD1 proteins has a crucial role in lentiviral restriction. J. Biol. Chem. 2020, 295, 4252–4264. [Google Scholar] [CrossRef] [PubMed]
  202. Gramberg, T.; Kahle, T.; Bloch, N.; Wittmann, S.; Mullers, E.; Daddacha, W.; Hofmann, H.; Kim, B.; Lindemann, D.; Landau, N.R. Restriction of diverse retroviruses by SAMHD1. Retrovirology 2013, 10, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Sze, A.; Belgnaoui, S.M.; Olagnier, D.; Lin, R.; Hiscott, J.; van Grevenynghe, J. Host restriction factor SAMHD1 limits human T cell leukemia virus type 1 infection of monocytes via STING-mediated apoptosis. Cell Host Microbe 2013, 14, 422–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Hollenbaugh, J.A.; Gee, P.; Baker, J.; Daly, M.B.; Amie, S.M.; Tate, J.; Kasai, N.; Kanemura, Y.; Kim, D.H.; Ward, B.M.; et al. Host factor SAMHD1 restricts DNA viruses in non-dividing myeloid cells. PLoS Pathog. 2013, 9, e1003481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Jeong, G.U.; Park, I.H.; Ahn, K.; Ahn, B.Y. Inhibition of hepatitis B virus replication by a dNTPase-dependent function of the host restriction factor SAMHD1. Virology 2016, 495, 71–78. [Google Scholar] [CrossRef]
  206. Sommer, A.F.; Riviere, L.; Qu, B.; Schott, K.; Riess, M.; Ni, Y.; Shepard, C.; Schnellbacher, E.; Finkernagel, M.; Himmelsbach, K.; et al. Restrictive influence of SAMHD1 on Hepatitis B Virus life cycle. Sci. Rep. 2016, 6, 26616. [Google Scholar] [CrossRef] [Green Version]
  207. Businger, R.; Deutschmann, J.; Gruska, I.; Milbradt, J.; Wiebusch, L.; Gramberg, T.; Schindler, M. Human cytomegalovirus overcomes SAMHD1 restriction in macrophages via pUL97. Nat. Microbiol. 2019, 4, 2260–2272. [Google Scholar] [CrossRef]
  208. Sliva, K.; Martin, J.; von Rhein, C.; Herrmann, T.; Weyrich, A.; Toda, M.; Schnierle, B.S. Interference with SAMHD1 Restores Late Gene Expression of Modified Vaccinia Virus Ankara in Human Dendritic Cells and Abrogates Type I Interferon Expression. J. Virol. 2019, 93, e01097–e01119. [Google Scholar] [CrossRef]
  209. Zhang, K.; Lv, D.W.; Li, R. Conserved Herpesvirus Protein Kinases Target SAMHD1 to Facilitate Virus Replication. Cell Rep. 2019, 28, 449–459. [Google Scholar] [CrossRef] [Green Version]
  210. Hu, J.; Qiao, M.; Chen, Y.; Tang, H.; Zhang, W.; Tang, D.; Pi, S.; Dai, J.; Tang, N.; Huang, A.; et al. Cyclin E2-CDK2 mediates SAMHD1 phosphorylation to abrogate its restriction of HBV replication in hepatoma cells. FEBS Lett. 2018, 592, 1893–1904. [Google Scholar] [CrossRef] [Green Version]
  211. Gao, S.; von der Malsburg, A.; Dick, A.; Faelber, K.; Schroder, G.F.; Haller, O.; Kochs, G.; Daumke, O. Structure of myxovirus resistance protein a reveals intra- and intermolecular domain interactions required for the antiviral function. Immunity 2011, 35, 514–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Alvarez, F.J.D.; He, S.; Perilla, J.R.; Jang, S.; Schulten, K.; Engelman, A.N.; Scheres, S.H.W.; Zhang, P. CryoEM structure of MxB reveals a novel oligomerization interface critical for HIV restriction. Sci. Adv. 2017, 3, e1701264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Goujon, C.; Moncorge, O.; Bauby, H.; Doyle, T.; Ward, C.C.; Schaller, T.; Hue, S.; Barclay, W.S.; Schulz, R.; Malim, M.H. Human MX2 is an interferon-induced post-entry inhibitor of HIV-1 infection. Nature 2013, 502, 559–562. [Google Scholar] [CrossRef] [PubMed]
  214. Kane, M.; Yadav, S.S.; Bitzegeio, J.; Kutluay, S.B.; Zang, T.; Wilson, S.J.; Schoggins, J.W.; Rice, C.M.; Yamashita, M.; Hatziioannou, T.; et al. MX2 is an interferon-induced inhibitor of HIV-1 infection. Nature 2013, 502, 563–566. [Google Scholar] [CrossRef] [Green Version]
  215. Liu, Z.; Pan, Q.; Ding, S.; Qian, J.; Xu, F.; Zhou, J.; Cen, S.; Guo, F.; Liang, C. The interferon-inducible MxB protein inhibits HIV-1 infection. Cell Host Microbe 2013, 14, 398–410. [Google Scholar] [CrossRef] [Green Version]
  216. Bulli, L.; Apolonia, L.; Kutzner, J.; Pollpeter, D.; Goujon, C.; Herold, N.; Schwarz, S.M.; Giernat, Y.; Keppler, O.T.; Malim, M.H.; et al. Complex Interplay between HIV-1 Capsid and MX2-Independent Alpha Interferon-Induced Antiviral Factors. J. Virol. 2016, 90, 7469–7480. [Google Scholar] [CrossRef] [Green Version]
  217. Xu, F.; Zhao, F.; Zhao, X.; Zhang, D.; Liu, X.; Hu, S.; Mei, S.; Fan, Z.; Huang, Y.; Sun, H.; et al. Pro-515 of the dynamin-like GTPase MxB contributes to HIV-1 inhibition by regulating MxB oligomerization and binding to HIV-1 capsid. J. Biol. Chem. 2020, 295, 6447–6456. [Google Scholar] [CrossRef] [Green Version]
  218. Fribourgh, J.L.; Nguyen, H.C.; Matreyek, K.A.; Alvarez, F.J.D.; Summers, B.J.; Dewdney, T.G.; Aiken, C.; Zhang, P.; Engelman, A.; Xiong, Y. Structural insight into HIV-1 restriction by MxB. Cell Host Microbe 2014, 16, 627–638. [Google Scholar] [CrossRef] [Green Version]
  219. Fricke, T.; White, T.E.; Schulte, B.; de Souza Aranha Vieira, D.A.; Dharan, A.; Campbell, E.M.; Brandariz-Nunez, A.; Diaz-Griffero, F. MxB binds to the HIV-1 core and prevents the uncoating process of HIV-1. Retrovirology 2014, 11, 68. [Google Scholar] [CrossRef]
  220. Smaga, S.S.; Xu, C.; Summers, B.J.; Digianantonio, K.M.; Perilla, J.R.; Xiong, Y. MxB Restricts HIV-1 by Targeting the Tri-hexamer Interface of the Viral Capsid. Structure 2019, 27, 1234–1245.e1235. [Google Scholar] [CrossRef]
  221. Melen, K.; Keskinen, P.; Ronni, T.; Sareneva, T.; Lounatmaa, K.; Julkunen, I. Human MxB protein, an interferon-alpha-inducible GTPase, contains a nuclear targeting signal and is localized in the heterochromatin region beneath the nuclear envelope. J. Biol. Chem. 1996, 271, 23478–23486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Goujon, C.; Moncorge, O.; Bauby, H.; Doyle, T.; Barclay, W.S.; Malim, M.H. Transfer of the amino-terminal nuclear envelope targeting domain of human MX2 converts MX1 into an HIV-1 resistance factor. J. Virol. 2014, 88, 9017–9026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Dicks, M.D.J.; Betancor, G.; Jimenez-Guardeno, J.M.; Pessel-Vivares, L.; Apolonia, L.; Goujon, C.; Malim, M.H. Multiple components of the nuclear pore complex interact with the amino-terminus of MX2 to facilitate HIV-1 restriction. PLoS Pathog. 2018, 14, e1007408. [Google Scholar] [CrossRef] [PubMed]
  224. Kane, M.; Rebensburg, S.V.; Takata, M.A.; Zang, T.M.; Yamashita, M.; Kvaratskhelia, M.; Bieniasz, P.D. Nuclear pore heterogeneity influences HIV-1 infection and the antiviral activity of MX2. eLife 2018, 7, e35738. [Google Scholar] [CrossRef] [PubMed]
  225. Xie, L.; Chen, L.; Zhong, C.; Yu, T.; Ju, Z.; Wang, M.; Xiong, H.; Zeng, Y.; Wang, J.; Hu, H.; et al. MxB impedes the NUP358-mediated HIV-1 pre-integration complex nuclear import and viral replication cooperatively with CPSF6. Retrovirology 2020, 17, 16. [Google Scholar] [CrossRef] [PubMed]
  226. Mitchell, P.S.; Patzina, C.; Emerman, M.; Haller, O.; Malik, H.S.; Kochs, G. Evolution-guided identification of antiviral specificity determinants in the broadly acting interferon-induced innate immunity factor MxA. Cell Host Microbe 2012, 12, 598–604. [Google Scholar] [CrossRef] [Green Version]
  227. Braun, B.A.; Marcovitz, A.; Camp, J.G.; Jia, R.; Bejerano, G. Mx1 and Mx2 key antiviral proteins are surprisingly lost in toothed whales. Proc. Natl. Acad. Sci. USA 2015, 112, 8036–8040. [Google Scholar] [CrossRef] [Green Version]
  228. Meier, K.; Jaguva Vasudevan, A.A.; Zhang, Z.; Bahr, A.; Kochs, G.; Haussinger, D.; Munk, C. Equine MX2 is a restriction factor of equine infectious anemia virus (EIAV). Virology 2018, 523, 52–63. [Google Scholar] [CrossRef]
  229. Ji, S.; Na, L.; Ren, H.; Wang, Y.; Wang, X. Equine Myxovirus Resistance Protein 2 Restricts Lentiviral Replication by Blocking Nuclear Uptake of Capsid Protein. J. Virol. 2018, 92, e00499–e00518. [Google Scholar] [CrossRef] [Green Version]
  230. Crameri, M.; Bauer, M.; Caduff, N.; Walker, R.; Steiner, F.; Franzoso, F.D.; Gujer, C.; Boucke, K.; Kucera, T.; Zbinden, A.; et al. MxB is an interferon-induced restriction factor of human herpesviruses. Nat. Commun. 2018, 9, 1980. [Google Scholar] [CrossRef] [Green Version]
  231. Schilling, M.; Bulli, L.; Weigang, S.; Graf, L.; Naumann, S.; Patzina, C.; Wagner, V.; Bauersfeld, L.; Goujon, C.; Hengel, H.; et al. Human MxB Protein Is a Pan-herpesvirus Restriction Factor. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [Green Version]
  232. Yi, D.R.; An, N.; Liu, Z.L.; Xu, F.W.; Raniga, K.; Li, Q.J.; Zhou, R.; Wang, J.; Zhang, Y.X.; Zhou, J.M.; et al. Human MxB Inhibits the Replication of Hepatitis C Virus. J. Virol. 2019, 93, e01285–e01318. [Google Scholar] [CrossRef] [Green Version]
  233. Jaguva Vasudevan, A.A.; Bahr, A.; Grothmann, R.; Singer, A.; Haussinger, D.; Zimmermann, A.; Munk, C. MXB inhibits murine cytomegalovirus. Virology 2018, 522, 158–167. [Google Scholar] [CrossRef]
  234. Wang, Y.X.; Niklasch, M.; Liu, T.; Wang, Y.; Shi, B.; Yuan, W.; Baumert, T.F.; Yuan, Z.; Tong, S.; Nassal, M.; et al. Interferon-inducible MX2 is a host restriction factor of hepatitis B virus replication. J. Hepatol. 2020, 72, 865–876. [Google Scholar] [CrossRef] [PubMed]
  235. Gao, G.; Guo, X.; Goff, S.P. Inhibition of retroviral RNA production by ZAP, a CCCH-type zinc finger protein. Science 2002, 297, 1703–1706. [Google Scholar] [CrossRef] [PubMed]
  236. Karlberg, T.; Klepsch, M.; Thorsell, A.G.; Andersson, C.D.; Linusson, A.; Schuler, H. Structural basis for lack of ADP-ribosyltransferase activity in poly(ADP-ribose) polymerase-13/zinc finger antiviral protein. J. Biol. Chem. 2015, 290, 7336–7344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Guo, X.; Carroll, J.W.; Macdonald, M.R.; Goff, S.P.; Gao, G. The zinc finger antiviral protein directly binds to specific viral mRNAs through the CCCH zinc finger motifs. J. Virol. 2004, 78, 12781–12787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Muller, S.; Moller, P.; Bick, M.J.; Wurr, S.; Becker, S.; Gunther, S.; Kummerer, B.M. Inhibition of filovirus replication by the zinc finger antiviral protein. J. Virol. 2007, 81, 2391–2400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Guo, X.; Ma, J.; Sun, J.; Gao, G. The zinc-finger antiviral protein recruits the RNA processing exosome to degrade the target mRNA. Proc. Natl. Acad. Sci. USA 2007, 104, 151–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Zhu, Y.; Chen, G.; Lv, F.; Wang, X.; Ji, X.; Xu, Y.; Sun, J.; Wu, L.; Zheng, Y.T.; Gao, G. Zinc-finger antiviral protein inhibits HIV-1 infection by selectively targeting multiply spliced viral mRNAs for degradation. Proc. Natl. Acad. Sci. USA 2011, 108, 15834–15839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Zhu, Y.; Wang, X.; Goff, S.P.; Gao, G. Translational repression precedes and is required for ZAP-mediated mRNA decay. EMBO J. 2012, 31, 4236–4246. [Google Scholar] [CrossRef] [PubMed]
  242. Takata, M.A.; Goncalves-Carneiro, D.; Zang, T.M.; Soll, S.J.; York, A.; Blanco-Melo, D.; Bieniasz, P.D. CG dinucleotide suppression enables antiviral defence targeting non-self RNA. Nature 2017, 550, 124–127. [Google Scholar] [CrossRef] [PubMed]
  243. Meagher, J.L.; Takata, M.; Goncalves-Carneiro, D.; Keane, S.C.; Rebendenne, A.; Ong, H.; Orr, V.K.; MacDonald, M.R.; Stuckey, J.A.; Bieniasz, P.D.; et al. Structure of the zinc-finger antiviral protein in complex with RNA reveals a mechanism for selective targeting of CG-rich viral sequences. Proc. Natl. Acad. Sci. USA 2019, 116, 24303–24309. [Google Scholar] [CrossRef] [PubMed]
  244. Ficarelli, M.; Antzin-Anduetza, I.; Hugh-White, R.; Firth, A.E.; Sertkaya, H.; Wilson, H.; Neil, S.J.D.; Schulz, R.; Swanson, C.M. CpG Dinucleotides Inhibit HIV-1 Replication through Zinc Finger Antiviral Protein (ZAP)-Dependent and -Independent Mechanisms. J. Virol. 2020, 94, e01337–e01419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Li, M.M.; Lau, Z.; Cheung, P.; Aguilar, E.G.; Schneider, W.M.; Bozzacco, L.; Molina, H.; Buehler, E.; Takaoka, A.; Rice, C.M.; et al. TRIM25 Enhances the Antiviral Action of Zinc-Finger Antiviral Protein (ZAP). PLoS Pathog. 2017, 13, e1006145. [Google Scholar] [CrossRef] [PubMed]
  246. Zheng, X.; Wang, X.; Tu, F.; Wang, Q.; Fan, Z.; Gao, G. TRIM25 Is Required for the Antiviral Activity of Zinc Finger Antiviral Protein. J. Virol. 2017, 91. [Google Scholar] [CrossRef] [Green Version]
  247. Ficarelli, M.; Wilson, H.; Pedro Galao, R.; Mazzon, M.; Antzin-Anduetza, I.; Marsh, M.; Neil, S.J.; Swanson, C.M. KHNYN is essential for the zinc finger antiviral protein (ZAP) to restrict HIV-1 containing clustered CpG dinucleotides. eLife 2019, 8, e46767. [Google Scholar] [CrossRef]
  248. Kerns, J.A.; Emerman, M.; Malik, H.S. Positive selection and increased antiviral activity associated with the PARP-containing isoform of human zinc-finger antiviral protein. PLoS Genet. 2008, 4, e21. [Google Scholar] [CrossRef]
  249. Bick, M.J.; Carroll, J.W.; Gao, G.; Goff, S.P.; Rice, C.M.; MacDonald, M.R. Expression of the zinc-finger antiviral protein inhibits alphavirus replication. J. Virol. 2003, 77, 11555–11562. [Google Scholar] [CrossRef] [Green Version]
  250. Mao, R.; Nie, H.; Cai, D.; Zhang, J.; Liu, H.; Yan, R.; Cuconati, A.; Block, T.M.; Guo, J.T.; Guo, H. Inhibition of hepatitis B virus replication by the host zinc finger antiviral protein. PLoS Pathog. 2013, 9, e1003494. [Google Scholar] [CrossRef]
  251. Goodier, J.L.; Pereira, G.C.; Cheung, L.E.; Rose, R.J.; Kazazian, H.H., Jr. The Broad-Spectrum Antiviral Protein ZAP Restricts Human Retrotransposition. PLoS Genet. 2015, 11, e1005252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Li, M.; Yan, K.; Wei, L.; Yang, J.; Lu, C.; Xiong, F.; Zheng, C.; Xu, W. Zinc finger antiviral protein inhibits coxsackievirus B3 virus replication and protects against viral myocarditis. Antiviral Res. 2015, 123, 50–61. [Google Scholar] [CrossRef] [PubMed]
  253. Zhu, M.; Ma, X.; Cui, X.; Zhou, J.; Li, C.; Huang, L.; Shang, Y.; Cheng, Z. Inhibition of avian tumor virus replication by CCCH-type zinc finger antiviral protein. Oncotarget 2017, 8, 58865–58871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Miyazato, P.; Matsuo, M.; Tan, B.J.Y.; Tokunaga, M.; Katsuya, H.; Islam, S.; Ito, J.; Murakawa, Y.; Satou, Y. HTLV-1 contains a high CG dinucleotide content and is susceptible to the host antiviral protein ZAP. Retrovirology 2019, 16, 38. [Google Scholar] [CrossRef] [PubMed]
  255. Chiu, H.P.; Chiu, H.; Yang, C.F.; Lee, Y.L.; Chiu, F.L.; Kuo, H.C.; Lin, R.J.; Lin, Y.L. Inhibition of Japanese encephalitis virus infection by the host zinc-finger antiviral protein. PLoS Pathog. 2018, 14, e1007166. [Google Scholar] [CrossRef]
  256. Schwarz, D.A.; Katayama, C.D.; Hedrick, S.M. Schlafen, a new family of growth regulatory genes that affect thymocyte development. Immunity 1998, 9, 657–668. [Google Scholar] [CrossRef] [Green Version]
  257. Li, M.; Kao, E.; Gao, X.; Sandig, H.; Limmer, K.; Pavon-Eternod, M.; Jones, T.E.; Landry, S.; Pan, T.; Weitzman, M.D.; et al. Codon-usage-based inhibition of HIV protein synthesis by human schlafen 11. Nature 2012, 491, 125–128. [Google Scholar] [CrossRef] [Green Version]
  258. Mavrommatis, E.; Fish, E.N.; Platanias, L.C. The schlafen family of proteins and their regulation by interferons. J. Interferon Cytokine Res. 2013, 33, 206–210. [Google Scholar] [CrossRef] [Green Version]
  259. Liu, F.; Zhou, P.; Wang, Q.; Zhang, M.; Li, D. The Schlafen family: Complex roles in different cell types and virus replication. Cell Biol. Int. 2018, 42, 2–8. [Google Scholar] [CrossRef]
  260. Bustos, O.; Naik, S.; Ayers, G.; Casola, C.; Perez-Lamigueiro, M.A.; Chippindale, P.T.; Pritham, E.J.; de la Casa-Esperón, E. Evolution of the Schlafen genes, a gene family associated with embryonic lethality, meiotic drive, immune processes and orthopoxvirus virulence. Gene 2009, 447, 1–11. [Google Scholar] [CrossRef]
  261. Van Weringh, A.; Ragonnet-Cronin, M.; Pranckeviciene, E.; Pavon-Eternod, M.; Kleiman, L.; Xia, X. HIV-1 modulates the tRNA pool to improve translation efficiency. Mol. Biol. Evol. 2011, 28, 1827–1834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Lin, Y.Z.; Sun, L.K.; Zhu, D.T.; Hu, Z.; Wang, X.F.; Du, C.; Wang, Y.H.; Wang, X.J.; Zhou, J.H. Equine schlafen 11 restricts the production of equine infectious anemia virus via a codon usage-dependent mechanism. Virology 2016, 495, 112–121. [Google Scholar] [CrossRef] [PubMed]
  263. Stabell, A.C.; Hawkins, J.; Li, M.; Gao, X.; David, M.; Press, W.H.; Sawyer, S.L. Non-human Primate Schlafen11 Inhibits Production of Both Host and Viral Proteins. PLoS Pathog. 2016, 12, e1006066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Valdez, F.; Salvador, J.; Palermo, P.M.; Mohl, J.E.; Hanley, K.A.; Watts, D.; Llano, M. Schlafen 11 Restricts Flavivirus Replication. J. Virol. 2019, 93, e00104–e00119. [Google Scholar] [CrossRef] [Green Version]
  265. Olszewski, M.A.; Gray, J.; Vestal, D.J. In silico genomic analysis of the human and murine guanylate-binding protein (GBP) gene clusters. J. Interferon Cytokine Res. 2006, 26, 328–352. [Google Scholar] [CrossRef]
  266. McLaren, P.J.; Gawanbacht, A.; Pyndiah, N.; Krapp, C.; Hotter, D.; Kluge, S.F.; Gotz, N.; Heilmann, J.; Mack, K.; Sauter, D.; et al. Identification of potential HIV restriction factors by combining evolutionary genomic signatures with functional analyses. Retrovirology 2015, 12, 41. [Google Scholar] [CrossRef] [Green Version]
  267. Krapp, C.; Hotter, D.; Gawanbacht, A.; McLaren, P.J.; Kluge, S.F.; Sturzel, C.M.; Mack, K.; Reith, E.; Engelhart, S.; Ciuffi, A.; et al. Guanylate Binding Protein (GBP) 5 Is an Interferon-Inducible Inhibitor of HIV-1 Infectivity. Cell Host Microbe 2016, 19, 504–514. [Google Scholar] [CrossRef] [Green Version]
  268. Braun, E.; Hotter, D.; Koepke, L.; Zech, F.; Gross, R.; Sparrer, K.M.J.; Muller, J.A.; Pfaller, C.K.; Heusinger, E.; Wombacher, R.; et al. Guanylate-Binding Proteins 2 and 5 Exert Broad Antiviral Activity by Inhibiting Furin-Mediated Processing of Viral Envelope Proteins. Cell Rep. 2019, 27, 2092–2104.e2010. [Google Scholar] [CrossRef] [Green Version]
  269. Hotter, D.; Sauter, D.; Kirchhoff, F. Guanylate binding protein 5: Impairing virion infectivity by targeting retroviral envelope glycoproteins. Small GTPases 2017, 8, 31–37. [Google Scholar] [CrossRef] [Green Version]
  270. Li, Z.; Qu, X.; Liu, X.; Huan, C.; Wang, H.; Zhao, Z.; Yang, X.; Hua, S.; Zhang, W. GBP5 is an interferon-induced inhibitor of respiratory syncytial virus. J. Virol. 2020, 94, e01407–e01420. [Google Scholar] [CrossRef]
  271. Ohmura-Hoshino, M.; Goto, E.; Matsuki, Y.; Aoki, M.; Mito, M.; Uematsu, M.; Hotta, H.; Ishido, S. A novel family of membrane-bound E3 ubiquitin ligases. J. Biochem. 2006, 140, 147–154. [Google Scholar] [CrossRef] [PubMed]
  272. Tada, T.; Zhang, Y.; Koyama, T.; Tobiume, M.; Tsunetsugu-Yokota, Y.; Yamaoka, S.; Fujita, H.; Tokunaga, K. MARCH8 inhibits HIV-1 infection by reducing virion incorporation of envelope glycoproteins. Nat. Med. 2015, 21, 1502–1507. [Google Scholar] [CrossRef] [PubMed]
  273. Zhang, Y.; Tada, T.; Ozono, S.; Kishigami, S.; Fujita, H.; Tokunaga, K. MARCH8 inhibits viral infection by two different mechanisms. eLife 2020, 9. [Google Scholar] [CrossRef] [PubMed]
  274. Yu, C.; Li, S.; Zhang, X.; Khan, I.; Ahmad, I.; Zhou, Y.; Li, S.; Shi, J.; Wang, Y.; Zheng, Y.H. MARCH8 Inhibits Ebola Virus Glycoprotein, Human Immunodeficiency Virus Type 1 Envelope Glycoprotein, and Avian Influenza Virus H5N1 Hemagglutinin Maturation. mBio 2020, 11. [Google Scholar] [CrossRef] [PubMed]
  275. Zhang, Y.; Lu, J.; Liu, X. MARCH2 is upregulated in HIV-1 infection and inhibits HIV-1 production through envelope protein translocation or degradation. Virology 2018, 518, 293–300. [Google Scholar] [CrossRef]
  276. Zhang, Y.; Tada, T.; Ozono, S.; Yao, W.; Tanaka, M.; Yamaoka, S.; Kishigami, S.; Fujita, H.; Tokunaga, K. Membrane-associated RING-CH (MARCH) 1 and 2 are MARCH family members that inhibit HIV-1 infection. J. Biol. Chem. 2019, 294, 3397–3405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Zhao, X.; Li, J.; Winkler, C.A.; An, P.; Guo, J.T. IFITM Genes, Variants, and Their Roles in the Control and Pathogenesis of Viral Infections. Front. Microbiol. 2018, 9, 3228. [Google Scholar] [CrossRef] [Green Version]
  278. Brass, A.L.; Huang, I.C.; Benita, Y.; John, S.P.; Krishnan, M.N.; Feeley, E.M.; Ryan, B.J.; Weyer, J.L.; van der Weyden, L.; Fikrig, E.; et al. The IFITM proteins mediate cellular resistance to influenza A H1N1 virus, West Nile virus, and dengue virus. Cell 2009, 139, 1243–1254. [Google Scholar] [CrossRef] [Green Version]
  279. Perreira, J.M.; Chin, C.R.; Feeley, E.M.; Brass, A.L. IFITMs restrict the replication of multiple pathogenic viruses. J. Mol. Biol. 2013, 425, 4937–4955. [Google Scholar] [CrossRef]
  280. Compton, A.A.; Bruel, T.; Porrot, F.; Mallet, A.; Sachse, M.; Euvrard, M.; Liang, C.; Casartelli, N.; Schwartz, O. IFITM proteins incorporated into HIV-1 virions impair viral fusion and spread. Cell Host Microbe 2014, 16, 736–747. [Google Scholar] [CrossRef] [Green Version]
  281. Savidis, G.; Perreira, J.M.; Portmann, J.M.; Meraner, P.; Guo, Z.; Green, S.; Brass, A.L. The IFITMs Inhibit Zika Virus Replication. Cell Rep. 2016, 15, 2323–2330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Weston, S.; Czieso, S.; White, I.J.; Smith, S.E.; Wash, R.S.; Diaz-Soria, C.; Kellam, P.; Marsh, M. Alphavirus Restriction by IFITM Proteins. Traffic 2016, 17, 997–1013. [Google Scholar] [CrossRef] [PubMed]
  283. Gorman, M.J.; Poddar, S.; Farzan, M.; Diamond, M.S. The Interferon-Stimulated Gene Ifitm3 Restricts West Nile Virus Infection and Pathogenesis. J. Virol. 2016, 90, 8212–8225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Poddar, S.; Hyde, J.L.; Gorman, M.J.; Farzan, M.; Diamond, M.S. The Interferon-Stimulated Gene IFITM3 Restricts Infection and Pathogenesis of Arthritogenic and Encephalitic Alphaviruses. J. Virol. 2016, 90, 8780–8794. [Google Scholar] [CrossRef] [Green Version]
  285. McMichael, T.M.; Zhang, Y.; Kenney, A.D.; Zhang, L.; Zani, A.; Lu, M.; Chemudupati, M.; Li, J.; Yount, J.S. IFITM3 Restricts Human Metapneumovirus Infection. J. Infect. Dis. 2018, 218, 1582–1591. [Google Scholar] [CrossRef] [Green Version]
  286. Li, C.; Du, S.; Tian, M.; Wang, Y.; Bai, J.; Tan, P.; Liu, W.; Yin, R.; Wang, M.; Jiang, Y.; et al. The Host Restriction Factor Interferon-Inducible Transmembrane Protein 3 Inhibits Vaccinia Virus Infection. Front. Immunol. 2018, 9, 228. [Google Scholar] [CrossRef]
  287. Li, C.; Zheng, H.; Wang, Y.; Dong, W.; Liu, Y.; Zhang, L.; Zhang, Y. Antiviral Role of IFITM Proteins in Classical Swine Fever Virus Infection. Viruses 2019, 11, 126. [Google Scholar] [CrossRef] [Green Version]
  288. Londrigan, S.L.; Wakim, L.M.; Smith, J.; Haverkate, A.J.; Brooks, A.G.; Reading, P.C. IFITM3 and type I interferons are important for the control of influenza A virus replication in murine macrophages. Virology 2020, 540, 17–22. [Google Scholar] [CrossRef]
  289. Lu, J.; Pan, Q.; Rong, L.; He, W.; Liu, S.L.; Liang, C. The IFITM proteins inhibit HIV-1 infection. J. Virol. 2011, 85, 2126–2137. [Google Scholar] [CrossRef] [Green Version]
  290. Tartour, K.; Appourchaux, R.; Gaillard, J.; Nguyen, X.N.; Durand, S.; Turpin, J.; Beaumont, E.; Roch, E.; Berger, G.; Mahieux, R.; et al. IFITM proteins are incorporated onto HIV-1 virion particles and negatively imprint their infectivity. Retrovirology 2014, 11, 103. [Google Scholar] [CrossRef] [Green Version]
  291. Yu, J.; Li, M.; Wilkins, J.; Ding, S.; Swartz, T.H.; Esposito, A.M.; Zheng, Y.M.; Freed, E.O.; Liang, C.; Chen, B.K.; et al. IFITM Proteins Restrict HIV-1 Infection by Antagonizing the Envelope Glycoprotein. Cell Rep. 2015, 13, 145–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Appourchaux, R.; Delpeuch, M.; Zhong, L.; Burlaud-Gaillard, J.; Tartour, K.; Savidis, G.; Brass, A.; Etienne, L.; Roingeard, P.; Cimarelli, A. Functional Mapping of Regions Involved in the Negative Imprinting of Virion Particle Infectivity and in Target Cell Protection by Interferon-Induced Transmembrane Protein 3 against HIV-1. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Ahi, Y.S.; Yimer, D.; Shi, G.; Majdoul, S.; Rahman, K.; Rein, A.; Compton, A.A. IFITM3 Reduces Retroviral Envelope Abundance and Function and Is Counteracted by glycoGag. mBio 2020, 11, e03088–e03119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Spence, J.S.; He, R.; Hoffmann, H.H.; Das, T.; Thinon, E.; Rice, C.M.; Peng, T.; Chandran, K.; Hang, H.C. IFITM3 directly engages and shuttles incoming virus particles to lysosomes. Nat. Chem. Biol. 2019, 15, 259–268. [Google Scholar] [CrossRef] [PubMed]
  295. Buchrieser, J.; Degrelle, S.A.; Couderc, T.; Nevers, Q.; Disson, O.; Manet, C.; Donahue, D.A.; Porrot, F.; Hillion, K.H.; Perthame, E.; et al. IFITM proteins inhibit placental syncytiotrophoblast formation and promote fetal demise. Science 2019, 365, 176–180. [Google Scholar] [CrossRef] [PubMed]
  296. Zani, A.; Zhang, L.; McMichael, T.M.; Kenney, A.D.; Chemudupati, M.; Kwiek, J.J.; Liu, S.L.; Yount, J.S. Interferon-induced transmembrane proteins inhibit cell fusion mediated by trophoblast syncytins. J. Biol. Chem. 2019, 294, 19844–19851. [Google Scholar] [CrossRef]
  297. Wang, Y.; Pan, Q.; Ding, S.; Wang, Z.; Yu, J.; Finzi, A.; Liu, S.L.; Liang, C. The V3 Loop of HIV-1 Env Determines Viral Susceptibility to IFITM3 Impairment of Viral Infectivity. J. Virol. 2017, 91. [Google Scholar] [CrossRef] [Green Version]
  298. Foster, T.L.; Wilson, H.; Iyer, S.S.; Coss, K.; Doores, K.; Smith, S.; Kellam, P.; Finzi, A.; Borrow, P.; Hahn, B.H.; et al. Resistance of Transmitted Founder HIV-1 to IFITM-Mediated Restriction. Cell Host Microbe 2016, 20, 429–442. [Google Scholar] [CrossRef] [Green Version]
  299. Zhang, Z.; Liu, J.; Li, M.; Yang, H.; Zhang, C. Evolutionary dynamics of the interferon-induced transmembrane gene family in vertebrates. PLoS ONE 2012, 7, e49265. [Google Scholar] [CrossRef] [Green Version]
  300. Everitt, A.R.; Clare, S.; Pertel, T.; John, S.P.; Wash, R.S.; Smith, S.E.; Chin, C.R.; Feeley, E.M.; Sims, J.S.; Adams, D.J.; et al. IFITM3 restricts the morbidity and mortality associated with influenza. Nature 2012, 484, 519–523. [Google Scholar] [CrossRef] [Green Version]
  301. Martinez-Pomares, L. The mannose receptor. J. Leukoc. Biol. 2012, 92, 1177–1186. [Google Scholar] [CrossRef] [PubMed]
  302. Sukegawa, S.; Miyagi, E.; Bouamr, F.; Farkasova, H.; Strebel, K. Mannose Receptor 1 Restricts HIV Particle Release from Infected Macrophages. Cell Rep. 2018, 22, 786–795. [Google Scholar] [CrossRef] [Green Version]
  303. Lubow, J.; Virgilio, M.C.; Merlino, M.; Collins, D.R.; Mashiba, M.; Peterson, B.G.; Lukic, Z.; Painter, M.M.; Gomez-Rivera, F.; Terry, V.; et al. Mannose receptor is an HIV restriction factor counteracted by Vpr in macrophages. eLife 2020, 9, e51035. [Google Scholar] [CrossRef]
  304. Olzmann, J.A.; Kopito, R.R.; Christianson, J.C. The mammalian endoplasmic reticulum-associated degradation system. Cold Spring Harb. Perspect Biol. 2013, 5, a013185. [Google Scholar] [CrossRef] [Green Version]
  305. Zhou, T.; Dang, Y.; Zheng, Y.H. The mitochondrial translocator protein, TSPO, inhibits HIV-1 envelope glycoprotein biosynthesis via the endoplasmic reticulum-associated protein degradation pathway. J. Virol. 2014, 88, 3474–3484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Zhou, T.; Frabutt, D.A.; Moremen, K.W.; Zheng, Y.H. ERManI (Endoplasmic Reticulum Class I alpha-Mannosidase) Is Required for HIV-1 Envelope Glycoprotein Degradation via Endoplasmic Reticulum-associated Protein Degradation Pathway. J. Biol. Chem. 2015, 290, 22184–22192. [Google Scholar] [CrossRef] [Green Version]
  307. Frabutt, D.A.; Wang, B.; Riaz, S.; Schwartz, R.C.; Zheng, Y.H. Innate Sensing of Influenza A Virus Hemagglutinin Glycoproteins by the Host Endoplasmic Reticulum (ER) Stress Pathway Triggers a Potent Antiviral Response via ER-Associated Protein Degradation. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Strebel, K.; Klimkait, T.; Maldarelli, F.; Martin, M.A. Molecular and biochemical analyses of human immunodeficiency virus type 1 vpu protein. J. Virol. 1989, 63, 3784–3791. [Google Scholar] [CrossRef] [Green Version]
  309. Neil, S.J.; Zang, T.; Bieniasz, P.D. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nature 2008, 451, 425–430. [Google Scholar] [CrossRef] [Green Version]
  310. Goto, T.; Kennel, S.J.; Abe, M.; Takishita, M.; Kosaka, M.; Solomon, A.; Saito, S. A novel membrane antigen selectively expressed on terminally differentiated human B cells. Blood 1994, 84, 1922–1930. [Google Scholar] [CrossRef] [Green Version]
  311. Erikson, E.; Adam, T.; Schmidt, S.; Lehmann-Koch, J.; Over, B.; Goffinet, C.; Harter, C.; Bekeredjian-Ding, I.; Sertel, S.; Lasitschka, F.; et al. In vivo expression profile of the antiviral restriction factor and tumor-targeting antigen CD317/BST-2/HM1.24/tetherin in humans. Proc. Natl. Acad. Sci. USA 2011, 108, 13688–13693. [Google Scholar] [CrossRef] [Green Version]
  312. Kupzig, S.; Korolchuk, V.; Rollason, R.; Sugden, A.; Wilde, A.; Banting, G. Bst-2/HM1.24 is a raft-associated apical membrane protein with an unusual topology. Traffic 2003, 4, 694–709. [Google Scholar] [CrossRef]
  313. Blanco-Melo, D.; Venkatesh, S.; Bieniasz, P.D. Origins and Evolution of tetherin, an Orphan Antiviral Gene. Cell Host Microbe 2016, 20, 189–201. [Google Scholar] [CrossRef] [Green Version]
  314. Perez-Caballero, D.; Zang, T.; Ebrahimi, A.; McNatt, M.W.; Gregory, D.A.; Johnson, M.C.; Bieniasz, P.D. Tetherin inhibits HIV-1 release by directly tethering virions to cells. Cell 2009, 139, 499–511. [Google Scholar] [CrossRef] [Green Version]
  315. Sakuma, T.; Noda, T.; Urata, S.; Kawaoka, Y.; Yasuda, J. Inhibition of Lassa and Marburg virus production by tetherin. J. Virol. 2009, 83, 2382–2385. [Google Scholar] [CrossRef] [Green Version]
  316. Weidner, J.M.; Jiang, D.; Pan, X.B.; Chang, J.; Block, T.M.; Guo, J.T. Interferon-induced cell membrane proteins, IFITM3 and tetherin, inhibit vesicular stomatitis virus infection via distinct mechanisms. J. Virol. 2010, 84, 12646–12657. [Google Scholar] [CrossRef] [Green Version]
  317. Blondeau, C.; Pelchen-Matthews, A.; Mlcochova, P.; Marsh, M.; Milne, R.S.; Towers, G.J. Tetherin restricts herpes simplex virus 1 and is antagonized by glycoprotein M. J. Virol. 2013, 87, 13124–13133. [Google Scholar] [CrossRef] [Green Version]
  318. Taylor, J.K.; Coleman, C.M.; Postel, S.; Sisk, J.M.; Bernbaum, J.G.; Venkataraman, T.; Sundberg, E.J.; Frieman, M.B. Severe Acute Respiratory Syndrome Coronavirus ORF7a Inhibits Bone Marrow Stromal Antigen 2 Virion Tethering through a Novel Mechanism of Glycosylation Interference. J. Virol. 2015, 89, 11820–11833. [Google Scholar] [CrossRef] [Green Version]
  319. Li, M.; Wang, P.; Zheng, Z.; Hu, K.; Zhang, M.; Guan, X.; Fu, M.; Zhang, D.; Wang, W.; Xiao, G.; et al. Japanese encephalitis virus counteracts BST2 restriction via its envelope protein E. Virology 2017, 510, 67–75. [Google Scholar] [CrossRef]
  320. Hoffmann, M.; Nehlmeier, I.; Brinkmann, C.; Krahling, V.; Behner, L.; Moldenhauer, A.S.; Kruger, N.; Nehls, J.; Schindler, M.; Hoenen, T.; et al. Tetherin Inhibits Nipah Virus but Not Ebola Virus Replication in Fruit Bat Cells. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [Green Version]
  321. Wan, J.J.; Ooi, Y.S.; Kielian, M. Mechanism of Tetherin Inhibition of Alphavirus Release. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Kelly, J.T.; Human, S.; Alderman, J.; Jobe, F.; Logan, L.; Rix, T.; Gonçalves-Carneiro, D.; Leung, C.; Thakur, N.; Birch, J.; et al. BST2/Tetherin Overexpression Modulates Morbillivirus Glycoprotein Production to Inhibit Cell-Cell Fusion. Viruses 2019, 11, 692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Zadeh, V.R.; Urata, S.; Sakaguchi, M.; Yasuda, J. Human BST-2/tetherin inhibits Junin virus release from host cells and its inhibition is partially counteracted by viral nucleoprotein. J. Gen. Virol. 2020, 101. [Google Scholar] [CrossRef]
  324. Wang, S.M.; Huang, K.J.; Wang, C.T. Severe acute respiratory syndrome coronavirus spike protein counteracts BST2-mediated restriction of virus-like particle release. J. Med. Virol. 2019, 91, 1743–1750. [Google Scholar] [CrossRef] [Green Version]
  325. González-Hernández, M.; Hoffmann, M.; Brinkmann, C.; Nehls, J.; Winkler, M.; Schindler, M.; Pöhlmann, S. A GXXXA Motif in the Transmembrane Domain of the Ebola Virus Glycoprotein Is Required for Tetherin Antagonism. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [Green Version]
  326. Bracq, L.; Xie, M.; Benichou, S.; Bouchet, J. Mechanisms for Cell-to-Cell Transmission of HIV-1. Front. Immunol. 2018, 9, 260. [Google Scholar] [CrossRef]
  327. Giese, S.; Marsh, M. Tetherin can restrict cell-free and cell-cell transmission of HIV from primary macrophages to T cells. PLoS Pathog. 2014, 10, e1004189. [Google Scholar] [CrossRef] [Green Version]
  328. Casartelli, N.; Sourisseau, M.; Feldmann, J.; Guivel-Benhassine, F.; Mallet, A.; Marcelin, A.G.; Guatelli, J.; Schwartz, O. Tetherin restricts productive HIV-1 cell-to-cell transmission. PLoS Pathog. 2010, 6, e1000955. [Google Scholar] [CrossRef] [Green Version]
  329. Jolly, C.; Booth, N.J.; Neil, S.J. Cell-cell spread of human immunodeficiency virus type 1 overcomes tetherin/BST-2-mediated restriction in T cells. J. Virol. 2010, 84, 12185–12199. [Google Scholar] [CrossRef] [Green Version]
  330. Mitchell, R.S.; Katsura, C.; Skasko, M.A.; Fitzpatrick, K.; Lau, D.; Ruiz, A.; Stephens, E.B.; Margottin-Goguet, F.; Benarous, R.; Guatelli, J.C. Vpu antagonizes BST-2-mediated restriction of HIV-1 release via beta-TrCP and endo-lysosomal trafficking. PLoS Pathog. 2009, 5, e1000450. [Google Scholar] [CrossRef] [Green Version]
  331. Douglas, J.L.; Viswanathan, K.; McCarroll, M.N.; Gustin, J.K.; Fruh, K.; Moses, A.V. Vpu directs the degradation of the human immunodeficiency virus restriction factor BST-2/Tetherin via a {beta}TrCP-dependent mechanism. J. Virol. 2009, 83, 7931–7947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Song, Y.E.; Cyburt, D.; Lucas, T.M.; Gregory, D.A.; Lyddon, T.D.; Johnson, M.C. βTrCP is Required for HIV-1 Vpu Modulation of CD4, GaLV Env, and BST-2/Tetherin. Viruses 2018, 10, 573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Dube, M.; Roy, B.B.; Guiot-Guillain, P.; Binette, J.; Mercier, J.; Chiasson, A.; Cohen, E.A. Antagonism of tetherin restriction of HIV-1 release by Vpu involves binding and sequestration of the restriction factor in a perinuclear compartment. PLoS Pathog. 2010, 6, e1000856. [Google Scholar] [CrossRef] [Green Version]
  334. Andrew, A.J.; Miyagi, E.; Strebel, K. Differential effects of human immunodeficiency virus type 1 Vpu on the stability of BST-2/tetherin. J. Virol. 2011, 85, 2611–2619. [Google Scholar] [CrossRef] [Green Version]
  335. Zhang, F.; Wilson, S.J.; Landford, W.C.; Virgen, B.; Gregory, D.; Johnson, M.C.; Munch, J.; Kirchhoff, F.; Bieniasz, P.D.; Hatziioannou, T. Nef proteins from simian immunodeficiency viruses are tetherin antagonists. Cell Host Microbe 2009, 6, 54–67. [Google Scholar] [CrossRef] [Green Version]
  336. Zhang, F.; Landford, W.N.; Ng, M.; McNatt, M.W.; Bieniasz, P.D.; Hatziioannou, T. SIV Nef proteins recruit the AP-2 complex to antagonize Tetherin and facilitate virion release. PLoS Pathog. 2011, 7, e1002039. [Google Scholar] [CrossRef]
  337. Chen, C.Y.; Shingai, M.; Welbourn, S.; Martin, M.A.; Borrego, P.; Taveira, N.; Strebel, K. Antagonism of BST-2/Tetherin Is a Conserved Function of the Env Glycoprotein of Primary HIV-2 Isolates. J. Virol. 2016, 90, 11062–11074. [Google Scholar] [CrossRef] [Green Version]
  338. Janaka, S.K.; Tavakoli-Tameh, A.; Neidermyer, W.J.; Serra-Moreno, R.; Hoxie, J.A.; Desrosiers, R.C.; Johnson, R.P.; Lifson, J.D.; Wolinsky, S.M.; Evans, D.T. Polymorphisms in Rhesus Macaque Tetherin Are Associated with Differences in Acute Viremia in Simian Immunodeficiency Virus Δ. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [Green Version]
  339. Buffalo, C.Z.; Sturzel, C.M.; Heusinger, E.; Kmiec, D.; Kirchhoff, F.; Hurley, J.H.; Ren, X. Structural Basis for Tetherin Antagonism as a Barrier to Zoonotic Lentiviral Transmission. Cell Host Microbe 2019, 26, 359–368.e358. [Google Scholar] [CrossRef]
  340. Tavakoli-Tameh, A.; Janaka, S.K.; Zarbock, K.; O’Connor, S.; Crosno, K.; Capuano, S., 3rd; Uno, H.; Lifson, J.D.; Evans, D.T. Loss of tetherin antagonism by Nef impairs SIV replication during acute infection of rhesus macaques. PLoS Pathog. 2020, 16, e1008487. [Google Scholar] [CrossRef]
  341. Goffinet, C.; Schmidt, S.; Kern, C.; Oberbremer, L.; Keppler, O.T. Endogenous CD317/Tetherin limits replication of HIV-1 and murine leukemia virus in rodent cells and is resistant to antagonists from primate viruses. J. Virol. 2010, 84, 11374–11384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Liberatore, R.A.; Bieniasz, P.D. Tetherin is a key effector of the antiretroviral activity of type I interferon in vitro and in vivo. Proc. Natl. Acad. Sci. USA 2011, 108, 18097–18101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Heusinger, E.; Kluge, S.F.; Kirchhoff, F.; Sauter, D. Early Vertebrate Evolution of the Host Restriction Factor Tetherin. J. Virol. 2015, 89, 12154–12165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Gupta, R.K.; Hue, S.; Schaller, T.; Verschoor, E.; Pillay, D.; Towers, G.J. Mutation of a single residue renders human tetherin resistant to HIV-1 Vpu-mediated depletion. PLoS Pathog. 2009, 5, e1000443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Lim, E.S.; Malik, H.S.; Emerman, M. Ancient adaptive evolution of tetherin shaped the functions of Vpu and Nef in human immunodeficiency virus and primate lentiviruses. J. Virol. 2010, 84, 7124–7134. [Google Scholar] [CrossRef] [Green Version]
  346. Liu, J.; Chen, K.; Wang, J.H.; Zhang, C. Molecular evolution of the primate antiviral restriction factor tetherin. PLoS ONE 2010, 5, e11904. [Google Scholar] [CrossRef] [Green Version]
  347. McNatt, M.W.; Zang, T.; Hatziioannou, T.; Bartlett, M.; Fofana, I.B.; Johnson, W.E.; Neil, S.J.; Bieniasz, P.D. Species-specific activity of HIV-1 Vpu and positive selection of tetherin transmembrane domain variants. PLoS Pathog. 2009, 5, e1000300. [Google Scholar] [CrossRef]
  348. Arias, J.F.; Colomer-Lluch, M.; von Bredow, B.; Greene, J.M.; MacDonald, J.; O’Connor, D.H.; Serra-Moreno, R.; Evans, D.T. Tetherin Antagonism by HIV-1 Group M Nef Proteins. J. Virol. 2016, 90, 10701–10714. [Google Scholar] [CrossRef] [Green Version]
  349. Jager, S.; Cimermancic, P.; Gulbahce, N.; Johnson, J.R.; McGovern, K.E.; Clarke, S.C.; Shales, M.; Mercenne, G.; Pache, L.; Li, K.; et al. Global landscape of HIV-human protein complexes. Nature 2012, 481, 365–370. [Google Scholar] [CrossRef]
  350. Matheson, N.J.; Sumner, J.; Wals, K.; Rapiteanu, R.; Weekes, M.P.; Vigan, R.; Weinelt, J.; Schindler, M.; Antrobus, R.; Costa, A.S.; et al. Cell Surface Proteomic Map of HIV Infection Reveals Antagonism of Amino Acid Metabolism by Vpu and Nef. Cell Host Microbe 2015, 18, 409–423. [Google Scholar] [CrossRef] [Green Version]
  351. Jain, P.; Boso, G.; Langer, S.; Soonthornvacharin, S.; De Jesus, P.D.; Nguyen, Q.; Olivieri, K.C.; Portillo, A.J.; Yoh, S.M.; Pache, L.; et al. Large-Scale Arrayed Analysis of Protein Degradation Reveals Cellular Targets for HIV-1 Vpu. Cell Rep. 2018, 22, 2493–2503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Nielsen, R.; Bustamante, C.; Clark, A.G.; Glanowski, S.; Sackton, T.B.; Hubisz, M.J.; Fledel-Alon, A.; Tanenbaum, D.M.; Civello, D.; White, T.J.; et al. A scan for positively selected genes in the genomes of humans and chimpanzees. PLoS Biol. 2005, 3, e170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Restriction factors block specific stages of the retroviral life cycle.
Figure 1. Restriction factors block specific stages of the retroviral life cycle.
Microorganisms 08 01965 g001
Figure 2. Structure and functional features of the HIV-1 and MLV cell surface receptors. (A) Schematic diagrams of the receptors for HIV-1 (CD4 and CCR5) and for different MLV subtypes (XPR1 and CAT1). Red bars indicate regions that bind virus envelope [7,16,17]. (B) XPR1 and CD4 receptor proteins. Blocks identify the transmembrane domain of CD4 and the extracellular loops (ECLs) in XPR1. Positively selected residues are marked with red arrows [13,17,18,19]. Receptor critical sites are marked with blue arrows and black arrows identify polymorphic sites in chimpanzee CD4 that influence SIV binding. Green bars identify CD4 sites susceptible to downregulation by Nef and Vpu [8,9,11,12].
Figure 2. Structure and functional features of the HIV-1 and MLV cell surface receptors. (A) Schematic diagrams of the receptors for HIV-1 (CD4 and CCR5) and for different MLV subtypes (XPR1 and CAT1). Red bars indicate regions that bind virus envelope [7,16,17]. (B) XPR1 and CD4 receptor proteins. Blocks identify the transmembrane domain of CD4 and the extracellular loops (ECLs) in XPR1. Positively selected residues are marked with red arrows [13,17,18,19]. Receptor critical sites are marked with blue arrows and black arrows identify polymorphic sites in chimpanzee CD4 that influence SIV binding. Green bars identify CD4 sites susceptible to downregulation by Nef and Vpu [8,9,11,12].
Microorganisms 08 01965 g002
Figure 3. Domain organization of retroviral post-entry restriction factors. Schematic representation identifies positively selected residues (down arrows) found through analyses of Fv1 in rodents and primate genes for TRIM5α, APOBEC3G, SAMHD1 and MX2 [64,65,66,67,68]. The red bars indicate the binding regions for Vif in APOBEC3G, Vpx in SAMHDI and capsid in TRIM5α, TRIMCyp and MX2 [69,70,71,72,73,74]. MHR, major homology region; CypA, cyclophilin A; BSE, bundle signaling element; NLS, nuclear localization signal; N, unstructured amino terminal domain.
Figure 3. Domain organization of retroviral post-entry restriction factors. Schematic representation identifies positively selected residues (down arrows) found through analyses of Fv1 in rodents and primate genes for TRIM5α, APOBEC3G, SAMHD1 and MX2 [64,65,66,67,68]. The red bars indicate the binding regions for Vif in APOBEC3G, Vpx in SAMHDI and capsid in TRIM5α, TRIMCyp and MX2 [69,70,71,72,73,74]. MHR, major homology region; CypA, cyclophilin A; BSE, bundle signaling element; NLS, nuclear localization signal; N, unstructured amino terminal domain.
Microorganisms 08 01965 g003
Figure 4. BST2 domain organization and membrane association. (A) Schematic diagram of the domain organization of BST2. Positively selected residues are identified by down arrows [345,346]. Regions of HIV-1 Vpu and Nef interaction are indicated with red bars [347,348]. (B) Structural representation of BST2 as a transmembrane protein with a GPI anchor. TM, transmembrane domain; GPI, glycosylphosphatidylinositol anchor.
Figure 4. BST2 domain organization and membrane association. (A) Schematic diagram of the domain organization of BST2. Positively selected residues are identified by down arrows [345,346]. Regions of HIV-1 Vpu and Nef interaction are indicated with red bars [347,348]. (B) Structural representation of BST2 as a transmembrane protein with a GPI anchor. TM, transmembrane domain; GPI, glycosylphosphatidylinositol anchor.
Microorganisms 08 01965 g004
Figure 5. Retroviral targets and virus-encoded antagonists of restriction factors. Viral antagonists and the restriction factors they target are indicated above each schematically represented MLV and HIV-1 genome. Restriction factors and their known viral protein targets are shown below each genome. LTR: long terminal repeat, MA: matrix, CA: capsid, NC: nucleocapsid, PR: protease, RT: reverse transcriptase, IN: integrase, RBD: receptor binding domain, SU: surface protein, TM: transmembrane protein. Vif, Vpu, Vpx/r and Nef are lentiviral accessory proteins.
Figure 5. Retroviral targets and virus-encoded antagonists of restriction factors. Viral antagonists and the restriction factors they target are indicated above each schematically represented MLV and HIV-1 genome. Restriction factors and their known viral protein targets are shown below each genome. LTR: long terminal repeat, MA: matrix, CA: capsid, NC: nucleocapsid, PR: protease, RT: reverse transcriptase, IN: integrase, RBD: receptor binding domain, SU: surface protein, TM: transmembrane protein. Vif, Vpu, Vpx/r and Nef are lentiviral accessory proteins.
Microorganisms 08 01965 g005
Table 1. Properties of Retroviral Restriction Factors.
Table 1. Properties of Retroviral Restriction Factors.
Restriction FactorViral Antagonist *IFN InducedRestricted VirusesPositive Selection *
RetrovirusesOther *
CAT1-NoEcotropic MLV-No
XPR1-NoNonecotropic MLV-Yes
CD4Vpu, NefNoHIV, SIV-Yes
CCR5Vpu, NefNoHIV--
Fv4-NoEcotropic MLV--
Rmcf, Rmcf2-NoNonecotropic MLVs--
SERINC5Nef, Vpu
S2 (EIAV) Glycogag (MLV)
NoHIV, SIVs, EIAV, MLV Alphaviruses, FilovirusesNo
Fv1-NoMLV, EIAV, Feline Foamy Virus-Yes
TRIM5-YesRetroviruses FlavivirusesYes
huAPOBEC3GVifYesHIV, SIVs, MLV HBVYes
mApobec3Glycogag, p50 (MLV)MLV-
SAMHD1Vpx, VprYesHIV, SIVs, EIAV, FIVHBV, Herpesvirus, VacciniaYes
MX2-YesHIV, SIVs, EIAV, FIVHerpesvirus, HBV, FlavivirusYes
ZAP-YesHIV, MLV HBV, Alphaviruses, FilovirusesYes
SLFN11-YesHIV, EIAV, MLVFlavivirusesYes
BST2/TetherinVpu, NefYesAll known retrovirusesEnveloped VirusesYes
GBP5-YesHIV, MLV Influenza, Zika, MeaslesYes
MARCH8-NoHIVVesicular Stomatitis Virus, Ebola Virus-
IFITM3Glycogag (MLV)YesHIV, MLV Influenza, MeaslesYes
*-, none or unknown. HIV: Human Immunodeficiency Virus; MLV: Murine Leukemia Virus; SIV: Simian Immunodeficiency Virus; EIAV: Equine Infectious Anemia Virus; FIV: Feline Immunodeficiency Virus; HBV: Hepatitis B Virus.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Boso, G.; Kozak, C.A. Retroviral Restriction Factors and Their Viral Targets: Restriction Strategies and Evolutionary Adaptations. Microorganisms 2020, 8, 1965. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms8121965

AMA Style

Boso G, Kozak CA. Retroviral Restriction Factors and Their Viral Targets: Restriction Strategies and Evolutionary Adaptations. Microorganisms. 2020; 8(12):1965. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms8121965

Chicago/Turabian Style

Boso, Guney, and Christine A. Kozak. 2020. "Retroviral Restriction Factors and Their Viral Targets: Restriction Strategies and Evolutionary Adaptations" Microorganisms 8, no. 12: 1965. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms8121965

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