Next Article in Journal
Structural Conservation and Functional Diversity of the Poxvirus Immune Evasion (PIE) Domain Superfamily
Previous Article in Journal
Bio-Control of Salmonella Enteritidis in Foods Using Bacteriophages
Previous Article in Special Issue
Hepatitis B Virus HBx Protein Interactions with the Ubiquitin Proteasome System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Viral Mimicry to Usurp Ubiquitin and SUMO Host Pathways

1
Novartis Pharma Germany, Roonstrasse 25, 90429 Nürnberg, Germany
2
Institute of Virology, Technische Universität München, Trogerstrasse 30, 81675 München, Germany
3
Helmholtz Zentrum München, Ingolstädter Landstrasse 1, 85764 Neuherberg/München, Germany
*
Author to whom correspondence should be addressed.
Viruses 2015, 7(9), 4854-4872; https://doi.org/10.3390/v7092849
Submission received: 6 May 2015 / Revised: 16 August 2015 / Accepted: 20 August 2015 / Published: 28 August 2015
(This article belongs to the Special Issue Viruses and the Ubiquitin/Proteasome System)

Abstract

:
Posttranslational modifications (PTMs) of proteins include enzymatic changes by covalent addition of cellular regulatory determinants such as ubiquitin (Ub) and small ubiquitin-like modifier (SUMO) moieties. These modifications are widely used by eukaryotic cells to control the functional repertoire of proteins. Over the last decade, it became apparent that the repertoire of ubiquitiylation and SUMOylation regulating various biological functions is not restricted to eukaryotic cells, but is also a feature of human virus families, used to extensively exploit complex host-cell networks and homeostasis. Intriguingly, besides binding to host SUMO/Ub control proteins and interfering with the respective enzymatic cascade, many viral proteins mimic key regulatory factors to usurp this host machinery and promote efficient viral outcomes. Advanced detection methods and functional studies of ubiquitiylation and SUMOylation during virus-host interplay have revealed that human viruses have evolved a large arsenal of strategies to exploit these specific PTM processes. In this review, we highlight the known viral analogs orchestrating ubiquitin and SUMO conjugation events to subvert and utilize basic enzymatic pathways.

1. Introduction

Ubiquitinylation and SUMOylation are reversible processes that regulate key factors. Consequently, these PTMs (posttranslational modification) contribute to signaling pathways such as gene regulation, epigenetics, differentiation, protein degradation, and tumorigenesis. Phosphorylation, acetylation, and methylation of substrates directly influence whether ubiquitin/SUMO (small ubiquitin-like modifier) moieties are conjugated/deconjugated to proteins, and how the ubiquitin/SUMO chains will be assembled [1,2,3]. Any of these processes can be altered by analogs of viral origin either circumventing or supporting ubiquitinylation/SUMOylation of host or viral substrates. Such events are often beneficial for virus progeny production and/or constitute immune evasion strategies associated with efficient viral gene expression and replication of the viral genome [4]. Therefore, identification and detailed understanding of virus-mediated PTM manipulation by viral analogs infiltrating ubiquitin/SUMO pathways will provide valuable information for future antiviral drug development and/or novel immunotherapies.

1.1. Ubiquitinylation and De-Ubiquitinylation Processes

Ubiquitinylation is an enzymatic cascade involving the covalent linkage of a small protein called ubiquitin to the amino group of a lysine residue in a substrate protein through an isopeptide bond to the C-terminal glycine of ubiquitin [5,6]. Beside canonically transferring ubiquitin to a lysine residue, non-canonically linked poly-ubiquitin chains can arise when ubiquitin is conjugated to the amine group in the N-terminus of the substrate protein [7].
Ubiquitin is a conserved, 76-amino-acid protein (8.5 kDa), containing seven lysine residues: Lys6, Lys11, Lys27, Lys29, Lys33, Lys48, and Lys63 [8]. Ubiquitin is activated by a factor called E1, a ubiquitin-activating enzyme (UBA1, UBA6) in an ATP-dependent manner, when a high-energy thioester bond is formed between the carboxy terminus of ubiquitin and an internal Cys residue of E1. Activated ubiquitin is then transferred and conjugated to the active site cysteine of one of many ubiquitin-conjugating enzymes, called E2. Finally, the ubiquitin ligase, called E3, catalyzes the formation of an isopeptide bond between the target protein and ubiquitin moieties (see Figure 1). In some cases, poly-ubiquitin chains can also be attached to mono-ubiquitinylated lysine residues [7].
Conjugation at Lys63 primarily mediates protein trafficking among subcellular compartments, whereas modification at Lys48 classically designates the substrate as a target for proteasomal degradation [9]. Three classes of E3 ubiquitin ligases are defined depending on their HECT domain, a Cullin-RING, or their U-box motif [10,11]. RBR (ring-between-ring) E3 ligases have also been reported, and these combine the structural features of HECT and RING E3 ligases [12].
De-ubiquitinylation is the reverse process of ubiquitinylation, where ubiquitin is cleaved from the target protein (see Figure 1), catalyzed by de-ubiquitinylation enzymes (DUBs) or ubiquitin-specific proteases (USPs) [13]. Such enzymes hydrolyze ubiquitin precursors directly after translation into free ubiquitin molecules, recycle ubiquitins that are conjugated on target proteins destined for proteasomal degradation, or modulate ubiquitin conjugation of target proteins. DUBs are involved in various biological processes including immune and inflammatory responses, chromatin remodeling, cellular antiviral signaling [14].
Figure 1. Overview of some viruses that modulate host ubiquitinylation pathways. Listed are viruses known to encode proteins exhibiting functions of cellular regulatory proteins involved in ubiquitinylation. Described in detail in the text.
Figure 1. Overview of some viruses that modulate host ubiquitinylation pathways. Listed are viruses known to encode proteins exhibiting functions of cellular regulatory proteins involved in ubiquitinylation. Described in detail in the text.
Viruses 07 02849 g001

1.2. SUMOylation and De-SUMOylation Processes

SUMOylation is an enzymatic cascade involving the covalent linkage of a small ubiquitin-like modifier (SUMO) protein to a lysine residue of a substrate by a three-step enzymatic pathway. SUMO is activated by the E1 enzymes SAE1/SAE2 prior to being conjugated to the unique E2 enzyme Ubc9. Finally, different classes of a very few known SUMO E3 ligases, such as PIAS, RING-domain-containing and TRIM proteins, connect SUMOs to target proteins via an isopeptide bond formed between the C-terminal carboxyl group of SUMO and the amino group of the lysine residue in the substrate (see Figure 2) [4]. SENPs (SUMO-specific proteases) are required for the maturation of SUMO precursors and to reverse SUMOylation.
In contrast to ubiquitinylation, four different SUMO paralogs (~12 kDa) are known (SUMO-1, SUMO-2, SUMO-3, and SUMO-4), which regulate various cellular processes and functions such as gene regulation, cell differentiation, and tumor development [15]. In comparison to ubiquitin, much less is known about the functions that are modulated by SUMO proteins. The covalent attachment of SUMO to target proteins mainly depends on the target consensus motif ψKxE/D (where ψ represents a large hydrophobic amino acid and x represents any amino acid), which is recognized by the E2 enzyme Ubc9 [16,17,18]. However, some reports describe a large number of proteins that lack this consensus sequence but nevertheless can be SUMOylated [3]. Unfortunately, even less is known about the role of multiple SUMO modification by single SUMO moieties at different lysine residues within one target protein, or SUMO chain formation [19]. Additonally, there might be crosstalk between SUMOylation and phosphorylation [20].
Figure 2. Overview of some viruses that modulate host SUMOylation pathways. Listed are viruses known to encode proteins exhibiting functions of cellular regulatory proteins involved in SUMOylation. Described in detail in the text.
Figure 2. Overview of some viruses that modulate host SUMOylation pathways. Listed are viruses known to encode proteins exhibiting functions of cellular regulatory proteins involved in SUMOylation. Described in detail in the text.
Viruses 07 02849 g002

2. Viral Mimicry of Ubiquitin or Ubiquitin E2/E3 Ligases

Many viruses encode viral gene products that mimic the functions of cellular proteins, and thus interfere with key cellular processes and host cell homeostasis. Consistent with this, viruses have acquired strategies to benefit from the eukaryotic ubiquitin system. Here, we summarize the evidence that some viruses encode analogs of enzymes involved in ubiquitinylation events to mimic the conserved functions of host ubiquitin, E2/E3 enzymes, or to directly counteract E3 activity (see table 1).

2.1. Viruses Expressing Viral Ubiquitin (v-ubi) Analogs

AcMNPV (Autographa californica nuclear polyhedrosis virus) infects invertebrates and belongs to the family of baculoviruses with large, rod-shaped, circular DNA. This virus establishes virions, which contain many nucleocapsids within a single viral envelope. AcMNPV encodes the v-ubi protein, which has 76% identity with the eukaryotic protein, but likely retains all the functions of the human homolog [21]. Analysis of infected cells indicated that v-ubi is highly expressed during the late phase of viral infection. Several attempts to establish virus mutants depleted for v-ubi were unsuccessful. Therefore, the expression of this viral gene is essential for the viral life cycle. It is hypothesized that v-ubi might serve as a molecular chaperone for the incorporation of viral proteins into particles [21].

2.2. Viruses Encoding Viral Ubiquitin E2 Enzyme Counterparts

ASFV (African swine fever virus) belongs to the Asfarviridae family and causes African swine fever (ASF). This large, double-stranded DNA virus with a linear genome encodes at least 150 genes and replicates in the cytoplasm of infected cells, often causing a hemorrhagic fever with high mortality rates in pigs [22]. In 1992, a study identified the first viral determinant encoded by ASFV with high homology to cellular UBC (ubiquitin-conjugating/E2) enzymes and similarities in structure and enzyme activity to the yeast ubiquitin-conjugating enzymes UBC2 and UBC3 [23]. The ASFV-encoded enzyme UBCv1 has UBC activity when expressed in E. coli, forming thioester bonds with ubiquitin in the presence of purified ubiquitin-activating enzyme (E1) and ATP, and subsequently transferring ubiquitin to specific protein targets. Mechanistically, the ASFV UBCv1 enzyme can form multi-ubiquitinylated conjugates of histones, di-ubiquitin, and multi-ubiquitinylated UBC/E2 enzymes and also transfers ubiquitin to histones in an E3-independent reaction [24].
It appears that the virus-encoded UBCv1 enzyme is involved in several steps during the virus replication cycle, such as uncoating or assembly of virus particles, transition from early to late gene expression, onset of virus DNA replication, and virus-mediated DNA repair [24].
Interestingly, ASFV UBCv1 has been hypothesized to affect monocyte/macrophage function by modulating ubiquitin-dependent mechanisms, although the molecular details are not known [24]. An additional role for the ASFV UBCv1 enzyme was suggested to be the manipulation of the endosome/lysosome system of monocytes or macrophages to the advantage of the virus. Thus, the acquisition of a UBC enzyme by ASFV enables the virus to exploit the host ubiquitin conjugation system, either to regulate the virus replication cycle or in some other way modify virus-host cell interactions [24].

2.3. Viruses Synthesizing Ubiquitin E3 Enzyme-Like Proteins

Several reports have demonstrated that the HAdV (Human Adenovirus) early factors E1B-55K and E4orf6 assemble into an E3 ubiquitin ligase complex associating with host cell components Rbx1/ROC1/Hrt1, Cullin2/5, and Elongin B and C in a fashion analogous to the cellular SCF and VBC complexes, which recruit Cullin1 and Cullin2 (summarized in [25]). In the functional E3 ligase complex, E1B-55K serves as the substrate recognition unit while E4orf6 assembles the cellular components. This high molecular mass complex sequesters specific cellular substrates, such as p53, Mre11, DNA ligase IV, Integrin alpha 3, Tip60, ATRX, and SPOC1 into the ubiquitin-dependent proteasomal degradation pathway; this process is essential for HAdV replication [26,27,28,29,30,31,32,33].
It is also becoming increasingly apparent that HAdV possesses additional strategies to hijack cellular E3 ubiquitin ligase functions. There is evidence that E1B-55K-dependent/E4orf6-independent as well as E1B-55K-independent/E4orf6-dependent, Cullin-based E3 ubiquitin ligase activities neutralize host defense processes [34,35]. In line with this, recent work showed that HAdV also assemble E3 ubiquitin ligases in a serotype-dependent manner. These findings indicate that different HAdVs exhibit distinct Cullin specificities, and therefore target a different selection of cellular substrates [36,37]. Taken together, these findings emphasize that further detecting new host substrates during infection, as well as identifying the underlying processes by which different HAdV serotypes modulate host cell pathways, will be important for gaining further insights into how these pathways operate.
KSHV (Kaposi’s sarcoma-associated herpesvirus) is a gammaherpesvirus establishing lifelong infections that are usually asymptomatic in immunocompetent individuals. Like all herpesviruses, KSHV can induce latency and lytic replication. KSHV encodes two membrane-associated RING-CH (MARCH) family E3 ubiquitin ligases, K3 and K5, which direct the ubiquitinylation and subsequent internalization and endolysosomal degradation of several different plasma membrane substrates, including MHC-1 molecules [38,39,40]. K3 preferentially induces Lys63 linked poly-ubiquitinylation, which requires the E2-conjugating enzymes UbcH5b/c and Ubc13 to initiate mono- and subsequent poly-ubiquitinylation of MHC-1 prior to clathrin-dependent internalization mediated by the epsin endocytic adaptor [41]. While a positionally conserved homologue of K3 can be found in several other gammaherpesvirus genomes, the presence of more than one vMARCH gene is a unique feature of KSHV [42,43].
By generating a recombinant KSHV, Brulois and coworkers found that K3- and K5-dependent downregulation of MHC-1 represents stage-specific regulation of immunoevasive activity during infection [44]. Their study uncovered a role for K3 in the removal of MHC-1 from the cell surface during lytic replication. In contrast to K5, this K3 activity was restricted to the later stages of lytic replication. As for some other gammaherpesvirus genes [43,45], it has been suggested that K5 may have originated from a gene duplication event [46]. This may have facilitated the evolution of distinct roles for K3 and K5.
MHV-68 (Murine gammaherpesvirus 68) is closely related to human gammaherpesviruses and naturally infects mice. For MHV-68, the in vivo significance of the virus interfering with antigen presentation is still not clear. Deletion of the viral factor mK3 resulted in low latency during primary infection [47]. However, it did not completely abolish MHV-68-mediated downregulation of MHC-I levels relative to levels in uninfected cells. This suggests that other immunoevasins may be encoded by MHV-68 and that partial recovery of MHC-I expression is sufficient to affect latency establishment [44].
HPV (Human papillomaviruses) infiltrates the ubiquitin pathway through the virus-encoded E3-like enzymes E6 and E7. The HPV E6 gene product, in cooperation with the cellular E6-associated protein (E6AP), binds and induces ubiquitin-dependent degradation of cellular substrates such as p53 [48]. This is mediated by E6/E6AP acting as an ubiquitin E3 ligase in functional cooperation with the E2 conjugating enzymes UbcH5/H6/H7 [49,50]. E6/E6AP was the first viral/cellular complex with E3 enzyme capacity reported to form a ubiquitin thioester as an intermediate step prior to conjugating ubiquitin onto target proteins [51]. Beside p53 as a host substrate [52,53], many other targets are degraded by this complex, such as c-Myc, Mcm7, Bak, E6TP1, and NFX1 [54,55,56,57,58].
The other viral oncoprotein encoded by HPV, E7, competes for binding to the cellular tumor suppressor Rb prior to the release of the cellular E2F transcription factor to promote cell cycle progression. HPV16 E7 was shown to highjack already assembled E3 ligase containing Cullin2/Elongin B/C and Rbx1 in the infected cell to degrade Rb via the host proteasomal pathway [59,60] and to efficiently modulate substrate specificity [61].
HIV (Human immunodeficiency virus) is a lentivirus that causes AIDS, a well-known progressive failure of the immune system, by infecting CD4+ T cells, macrophages, and dendritic cells. The HIV RNA genome harbors several open reading frames in addition to the retroviral elements gag, pol, and env. Rev and Tat polypeptides are involved in viral gene regulation. Vif, Vpu, Vpr, and Nef proteins play essential roles in establishing the virus replicative environment. Among these proteins, at least three (Vif, Vpu, and Vpr) subvert cellular ubiquitin ligases to block the action of anti-viral defenses. Interestingly, investigation of mRNA synthesis revealed that the cellular cytidine deaminase APOBEC3G restricts HIV1 infectivity [62]. This is efficiently counteracted when Vif protein interacts with APOBEC3G. Vif reconstitutes an E3 ubiquitin ligase by hijacking cellular components such as Cullin5, Rbx2, and ElonginB/C [63]. To infiltrate this complex, Vif contains a conserved BC box motif for efficient ElonginB/C binding. Deletion of this BC box in Vif abolishes co-immunoprecipitation with the Elongins [64]. The HCCH zinc motif in Vif is critical for selective Cullin5 binding [65].
HIV Vpr is a 14 kDa virion-associated protein that promotes infection of non-dividing cells [66] and triggers G2 cell cycle arrest in dividing cells [67,68]. Vpr assembles a cellular ubiquitin ligase complex containing DCAF1/DDB1/Cullin4 [69]. Numerous targets have been discovered for this DCAF1/DDB1/Cullin4 ubiquitin ligase complex in the absence of Vpr, such as STAT1, XPC, Chk1, CDT1, Merlin, c-Jun, TSC2, and p27kip [70]. However, Vpr must associate with the DCAF1/DDB1/Cullin4 ubiquitin ligase complex for Vpr-mediated G2 cell cycle arrest and DNA-damage response, macrophage infection, and Vpr-mediated degradation of the uracil-DNA glycosylases UNG2 and SMUG1 in order to increase viral fitness [71]. IRF3 represents another substrate that is degraded in the presence of Vpr [72] to support virus-mediated counteraction of host immune responses in the context of a natural infection.

2.4. Viruses Blocking Host Ubiquitin E3 Enzymes by Virus-Encoded Inhibitors

VACV (Vaccinia virus) belongs to the family of Poxviruses and represents large DNA viruses that replicate in the cytoplasm. VACV was the live vaccine used to eradicate smallpox [73]. Poxviruses express many immunomodulatory factors [74], such as the A49 protein, which is also conserved in other VACV strains and orthopoxviruses including variola, but not in monkeypox, camelpox, and ectromelia viruses [75]. During infection, A49 is transcribed early and late [76]. Mansur and collegues reported that A49 blocks NF-κB activation and thus contributes to virus virulence [77]. A49 exploits molecular mimicry of IκBα to bind the WD40 domain of the E3 ubiquitin ligase β-TrCP via a sequence closely related to motifs in IκBα, IκBβ, HIV Vpu, and other β-TrCP substrates. Consequently, IκBα is not ubiquitinylated or degraded and remains attached to the NF-κB complex in the cytoplasm. Hence, the VACV A49 protein mediates anti-NF-kB activity using molecular mimicry to inhibit β-TrCP E3 ubiquitin ligase function, thereby blocking NF-kB activation to promote immune evasion and enhance virus virulence.
RV (Rotaviruses) are double-stranded RNA viruses of the Reovirus family and are a major cause of pediatric gastroenteritis [78]. The icosahedral RV particle encapsidates a genome encoding six structural and five or six nonstructural proteins [79]. RV replicates primarily in the villous tips of the small intestine [79], where it is sensed by the RLRs (RIG-I like receptors) RIG-I and MDA5, prior to NF-κB and IRF3 activation through MAVS (mitochondrial antiviral-signaling protein). Like VACV A49 above, RV can block NF-κB activation through its nonstructural protein NSP1, which targets the E3 ubiquitin ligase β-TrCP [80,81,82]. Thus, NSP1 mimics the IκB phosphodegron to mediate degradation of β-TrCP or other immunomodulatory proteins, such as IRF3/5/7/9, MAVS, and/or TNF receptor-associated factor 2 (TRAF2) [83,84].

3. Virus-Mediated Cross-Talk between SUMO and Ubiquitin PTM

SIM (SUMO interacting motif)-containing SUMO-targeted ubiquitin ligases (STUbLs) mediate the cross-talk between ubiquitin and SUMO signaling by promoting the ubiquitinylation and proteasome-dependent degradation of poly-SUMOylated substrates [85]. The human RNF4 protein is the most studied STUbL [86,87] and represents the smallest ULS (ubiquitin ligases for SUMO conjugates) characterized so far. RNF4 comprises at least three SIMs, which mediate binding to SUMO-1 and SUMO-2, with a preference for chains of at least three SUMO moieties [86].
Recently, another STUbL protein, Arkadia/RNF111, was identified [88]. Arkadia bears at least three functional SIMs, which mediate strong binding of linear SUMO-1 or SUMO-2 chains, suggesting that the function of Arkadia involves recognizing poly-SUMO signals [88]. Taken together, the ubiquitinylation of SUMOylated proteins mainly sequesters the substrate proteins into the proteasomal pathway to control the levels of SUMOylated proteins.
HSV1 (Human herpes simplex virus type 1) expresses the early viral protein ICP0 (Infected Cell Polypeptide 0), which is a RING finger E3 ubiquitin ligase with STUbL activity, and plays a role in efficiently initiating lytic infection as well as in reactivating latent viral genomes. HSV1 disrupts antiviral PML (promyelocytic leukemia protein) nuclear bodies in the host-cell nucleus mainly via ICP0-mediated proteasomal degradation, by targeting SUMOylated PML and Sp100 [89]. Indeed, ICP0 targets PML more efficiently than SUMO conjugates in general. Whereas PML isoforms IIVI require SUMOylation to be recognized by ICP0, isoform I is also targeted without SUMOylation. Upon infection with ICP0-null mutant viruses, SUMO-1 and SUMO-2/-3 conjugates are seen to accumulate [89,90,91].
VZV (Varicella zoster virus) expresses an ICP0 ortholog called ORF61p, which specifically targets the PML-NB component Sp100 and promotes dispersal of these nuclear aggregates [92]. ORF61p has three SIMs that are required for the viral factor to interact with PML nuclear bodies, whereas its additional RING domain is only required for PML-NB dispersal [93,94]. Unlike ICP0, ORF61p promotes SUMO-independent proteasomal degradation of cellular substrates such as IRF3 [95].
MHV-68 (Murine gammaherpesvirus 68) encodes the vFGARAT protein ORF75c, which mediates proteasome-dependent degradation of the PML protein. While ORF75c does not stimulate the activity of other known regulators of PML protein stability, such as CK2 and E6AP, it possesses self-ubiquitinylation activity in vitro and is sufficient to increase the level of ubiquitinylated PML. These observations suggest that it may be a specific PML E3 ubiquitin ligase, especially since other known PML E3 ubiquitin ligases were not required for ORF75c-mediated PML degradation [96].

4. Viruses Hijacking De-Ubiquitinylation Processes for Efficient Replication

DUBs are major regulators in various biological processes. Thus, increasing awareness of these enzymes expressed by viruses can not only provide novel information about unknown steps in virus replication (see table 1), as well as further insights into more general aspects of this class of enzymes, but can also highlight potential sites for pharmacological intervention.

4.1. DNA Viruses Synthesize Viral DUBs to Counteract Ubiquitinylation Events

HAdV (Human Adenoviruses) encode the viral protease adenain (Avp) [97]. HAdV virions contain about 10 copies of this cysteine proteinase [98], as well as several other HAdV proteins essential for the formation of mature infectious virions [99]. DNA binding inside the virions activates the viral protease Avp to cleave 11 amino acids (pVIc) from the HAdV capsid pVI protein precursor. pVIc represents a cofactor and results in further Avp activation, which subsequently leads to proteolytic maturation of viral capsid precursor proteins inside the virion. This is a prerequisite for endosomal lysis and cytoplasmic entry of viral capsids during the initial infection cycle [100]. Activation of Avp by the viral DNA and pVIc prevents precursor protein cleavage before virion assembly, thus generating immature capsids.
In 2002, Balakirev and coworkers used a biotinylated form of a DUB inhibitor as a probe to retrieve adenain from a lysate of infected cells and observed a time-dependent increase in de-ubiquitinating activity during the late phase of infection compared to mock infected cells [101]. Moreover, in vitro assays with recombinant purified Avp, substrate proteins, and chemically synthesized pVIc peptides showed that Avp exhibits DUB activity against Lys48-linked poly-ubiquitin, and that the ISG15 precursor protein is a cellular substrate for this adenoviral DUB [102].
Table 1. Host cell SUMO/Ubiquitin processes infiltrated by viral mimicry.
Table 1. Host cell SUMO/Ubiquitin processes infiltrated by viral mimicry.
VirusViral HomologueHost Cell FunctionPutative ConsequenceReferences
Human Adenoviruses (HAdV)AvpDUBefficient viral replication[101]
Human Adenoviruses (HAdV)E1B-55K
E4orf6
E3 ubiquitin ligaseefficient viral replication
oncogenic potential
[25]
Human Adenoviruses (HAdV)AvpSENPefficient viral replication[103,104]
Human Adenoviruses (HAdV)E1B-55KE3 SUMO ligasep53 SUMOylation
oncogenic potential
[105,106,107]
Human Papillomaviruses (HPV)E6E3 ubiquitin ligaseefficient viral replication
oncogenic potential
[48,49,50,51]
Human herpes simplex virus type 1 (HSV1)ICP0E3 ubiquitin ligaseefficient viral replication[89,90]
Human herpes simplex virus type 1 (HSV1)UL36 DUB[108]
Human Cytomegalovirus (HCMV)UL36DUB[109]
Kaposi’s sarcoma associated herpesvirus (KSHV)K3
K5
E3 ubiquitin ligaseimmune evasion[38,39,40]
Kaposi’s sarcoma associated herpesvirus (KSHV)K-bZIPE3 SUMO ligase[110,111]
Epstein-Barr virus (EBV)BPLF1DUB[112]
Murine gammaherpesvirus 68 (MHV68)ORF75cE3 ubiquitin ligaseefficient viral replication[96]
Murine gammaherpesvirus 68 (MHV68)ORF64DUB[113]
Murine gammaherpesvirus 68 (MHV68)mK3E3 ubiquitin ligaseimmune evasion[44,47]
Varicella zoster virus (VZV)ORF61pE3 ubiquitin ligaseNF-κB inhibition
[92,93,95]
Pseudorabies viruses (PrV)pUL36DUB[114]
African swine fever virus (ASFV)S273RSENPefficient viral replication[115]
African swine fever virus (ASFV)UBCv1ubiquitin conjugating activityuncoating/assembly early/late transition virus DNA replication virus-mediated DNA repair[23,24]
Vaccinia virus (VACV)I7SENP[103,104]
Vaccinia virus (VACV)A49E3 ubiquitin ligase inhibitionNF-κB inhibition
viral virulence
immune evasion
[77]
Autographa californica nuclear polyhedrosis virus (AcMNPV)v-ubiubiquitin homologuesubstrate for ubiquitin processing enzymes[21]
Rotavirus (RV)NSP1E3 ubiquitin ligase inhibitionNF-κB inhibition
[80,81,82]
Human immunodeficiency virus (HIV)VifE3 ubiquitin ligase immune evasion[63,64,65]
Human immunodeficiency virus (HIV)VprE3 ubiquitin ligase immune evasion[69]
Crimean–Congo hemorrhagic fever virus (CCHF)LDUBviral replication[116]
Dugbe virus (DUGV)LDUBviral replication[116]
Nairobi sheep disease virus (NSDV)LDUBviral replication[116]
Severe acute respiratory syndrome coronavirus (SARS CoV)PLproDUBde-ISGylation
immune evasion
[117,118]
Middle East respiratory syndrome coronavirus (MERS-CoV)PLproDUBde-ISGylation
immune evasion
[117,119,120,121]
Human coronavirus NL63 (HCoV-NL63)PLP2DUBde-ISGylation
immune evasion
[117,119,120,121]
Porcine epidemic diarrhea virus (PEDV)PLP2DUBde-ISGylation
immune evasion
[117,119,120,121]
Mouse hepatitis virus (MHV)PLP2DUBde-ISGylation
immune evasion
[117,119,120,121]
HSV1 (Herpes simplex virus type 1) tegument protein UL36 (VP1/2; ICP1/2) was reported to exhibit DUB activity in its N-terminal region [108]. Although the reactive sequence shows no similarity to any known DUB protein, the conservation of the Cys and His box motifs is characteristic for Cys protease DUB families [122]. The enzymatic activity of HSV1 UL36 is specific for ubiquitin, but exhibits relatively low catalytic efficiency. It was shown that UL36 disassembled Lys48 but not Lys63 poly-ubiquitin, implicating a role in protein stabilization [108]. In addition to HSV1, the presence of DUB activity with specificity for ubiquitin was also confirmed in recombinant UL36 proteins from EBV (gamma–herpesviruses), MCMV/HCMV (betaherpesviruses) [109,123], and MHV-68 (Murine gammaherpesvirus 68) ORF64 [113].
To also confirm catalytic activity in HCMV (Human Cytomegalovirus) extracellular particles, mutation of the respective Cys and His residues resulted in lower virus yields and delayed development of cytopathic effects. Therefore, the DUB enzymatic activity of UL36 plays a major but not essential role during HCMV replication [109]. In detail, the HCMV homologue of HSV UL36 is the UL48 protein, which is required to cleave and package the HCMV viral genome into capsids [124]. UL48 is involved in connecting incoming viral capsids to the dynein motor during transport along microtubules to the nuclear pores [125] and during retrograde axonal transport [126,127].
EBV (Epstein-Barr virus) is a human gammaherpesvirus that establishes latent infections and is associated with a broad spectrum of malignancies of lymphoid and epithelial cell origin. The viral protein BPLF1, the EBV homologue of HSV1 UL36 and HCMV UL48, was shown to encode viral DUB activity. Using a reporter substrate that carries ubiquitin fused to GFP, the N-terminus of BPLF1 was confirmed as a very potent ubiquitin de-conjugase. EBV BSLF1 and BXLF1 were also capable of cleaving ubiquitin [112]. These novel viral DUBs are expressed in the host nucleus during the early phase of productive EBV infection, in accordance with their reported function of the viral factors in DNA replication and nucleotide metabolism. Interestingly, the regions in BXLF1 that are essential for its DUB activity are not conserved in other herpesvirus analogs. In contrast to other herpesviruses, EBV BXLF1 was shown to localize to the centrosomes, where it encircles the tubulin-rich centrioles in a microtubule-independent manner [128].
PrV (Pseudorabies virus) belongs to the alphaherpesviruses and also encodes for the essential tegument protein pUL36 [114]. Mutational analysis of pUL36-dependent de-ubiquitinylation revealed that the inhibiting active site Cys26 led to defective virion morphogenesis accompanied by impaired neuroinvasion in a mouse infection model. Cellular or viral substrates of the cysteine protease in pUL36 are not known and, therefore, the biological role in herpesvirus replication is not understood in detail [129].
So far, all known herpesviruses encode a large tegument protein containing de-ubiquitinylating activity located within their N-terminal regions [123]. These viral DUBs function as cysteine proteases and have a strictly conserved catalytic triad comprised of Cys, His, and Asp acid residue [123]. Herpesvirus-encoded DUBs do not share sequence homology or conserved domains with known host cell DUBs [108]. It is also not completely understood whether de-ubiquitinylation of viral or cellular proteins by the UL36 DUB activity plays a part in UL36 functions, i.e., nucleocapsid formation, tegumentation, budding, viral egress from and entry into host-cells, nuclear transport and entry of the herpesvirus genome. Notably, the minimal DUB consensus domain is highly conserved among human (HSV, HCMV, EBV) and murine (MCMV, MHV-68) herpesviruses and confers both hydrolytic and ubiquitin specificity [108,123].

4.2. DUBs Encoded by RNA Viruses Limiting Host Ubiquitinylation

CCHF (Crimean–Congo hemorrhagic fever virus) is the tick-borne causative agent of severe hemorrhagic fever with outbreaks and high mortality in Asia, Europe, and Africa [130,131]. These viruses have a negative-strand RNA composed of three segments designated S, M, and L [132], where the expressed L protein was shown to resemble the viral polymerase function. For CCHF and other nairoviruses, such as DUGV (Dugbe virus) and NSDV (Nairobi sheep disease virus), an ovarian tumor (OTU)-like protease motif, a predicted papain-like protease with a cysteine protease signature, could be identified located in the amino termini of the L protein [116]. L proteins are polyproteins that are autoproteolytically cleaved or involved in de-ubiquitinylation [116]. Further investigation is needed to determine what role the de-ubiquitinylation activity may play in the virus replication cycle. Intriguingly, the OTU-like protease family also includes proteins encoded by other segmented negative strand viruses (rice stripe virus) and several positive strand RNA viruses (carlaviruses, foveaviruses), which are autoproteolytically cleaved to generate the viral RNA-dependent RNA polymerase and additional proteins such as viral helicases [133,134].
SARS CoV (Severe acute respiratory syndrome coronavirus) is a positive and single-stranded RNA virus with an envelope surrounding the viral particle. SARS CoV is the causative agent for severe acute respiratory syndrome, which frequently leads to outbreaks and high mortality. CoV usually expresses two large polyproteins, which are cleaved into 16 smaller subunits by two coronaviral proteases, 3CLpro (3C-like protease) and PLpro (papain-like protease) [135]. A Ub-like domain was identified at the N-terminus of the PLpro catalytic core domain [136] and studies unraveling the crystal structure of SARS-CoV PLpro established high homology with the host-cell USP family of DUBs [137]. Specifically, SARS-CoV PLpro efficiently debranches Lys48 poly-ubiquitin chains and hydrolyzes general DUB substrates [117,118].
In addition, precursor processing activity was shown for ISG15 [118], an additional ubiquitin-like protein expressed in response to interferon induction, which is covalently conjugated to substrate proteins [138]. It is not yet known whether SARS-CoV PLpro de-ubiquitinates further substrates other than replicase polyprotein sequences in vivo. Since the synthesis of CoV RNA requires ongoing viral protein production, one could speculate whether de-ubiquitinylation by SARS-CoV PLpro protects replicase subunits against proteasomal degradation [139]. In addition, further studies revealed that beside SARS-CoV PLpro, MERS-CoV (Middle East respiratory syndrome coronavirus) PLpro; HCoV-NL63 (Human coronavirus NL63) PLP2; PEDV (Porcine epidemic diarrhea virus) PLP2; and MHV (Mouse hepatitis virus) PLP2 also encode polypeptides with a ubiquitin-like domain located upstream of the protease domain [119], exhibiting protease, DUB, and de-ISGylating activities [117,120,121].

5. Viral Analogs of the Host SUMOylation Pathway

SUMO-1, -2, and -3, are covalently linked to their substrates via an enzymatic cascade that involves the heterodimeric activating enzyme SAE1/SAE2, the conjugating enzyme UBC9, and different classes of SUMO ligases, including the PIAS family, RING-domain-containing proteins, and members of the TRIM protein family. SUMO-specific proteases (SENPs) are required for the maturation of SUMO precursors and to reverse SUMOylation. Various viruses have evolved to take advantage of the conserved host-cell SUMOylation machinery by molecular mimicry of essential components (see table 1).

5.1. Viruses Mimicking Cellular SENPs

The Avp protease of HAdV (Human Adenoviruses) is a viral member of a new class of related cysteine proteases, which include the cellular SUMO proteases [103,104]. A crystal structure of Avp revealed that the residues implicated in catalysis can be superimposed on the corresponding active-site residues of the classical cysteine protease papain, despite the absence of primary sequence similarity between the enzymes (see below). All four key catalytic residues of the adenovirus protease are fully conserved in the cellular SENP enzymes, indicating that both protein types use a similar catalytic mechanism [103].
The core domain of the active site triad of Avp comprises residues His54, Glu71, and Cys122, which, in the crystal structure of this enzyme, have a spatial disposition identical to that of the catalytic triad of the classical cysteine protease papain, although the sequential order in papain is Cys25, His159, and Asn175. In addition, Gln115 is in an equivalent spatial position to the Gln19 of papain involved in forming the oxyanion hole in the active site [140].
Beside the adenoviral Avp protease, the VACV (Vaccinia virus) and Fowlvirus encoded I7 protease also share functional and structural similarities to the SENP proteins, which means they can be grouped in a specific kind of cystein protease family [103,104]. The I7 gene, encoding a 423-amino-acid polypeptide, is expressed late in infection [141] and harbors a C-terminal region of about 200 amino acids containing a 90 amino acid core domain that is also present in the adenovirus protease.
ASFV (African swine fever virus) encodes two polyproteins, pp220 and pp62, which are cleaved to produce six major structural components of the virus particle [142]. All the proteolytic cleavages occur after the second Gly of the consensus sequence Gly-Gly-Xaa, which is also recognized as a cleavage site in the maturation of adenovirus structural proteins and in some cellular proteins [143]. The S273R gene product of ASFV was suggested to be the protease involved in processing the ASFV polyproteins. Inhibitor profiling and site-directed mutagenesis indicated a specific cysteine protease activity [115]. Studies on S273R in cells infected with ASFV support a role for this protein in the processing of the viral polyproteins at late times of infection, when S273R is localized within the cytoplasmic viral factories where morphogenesis occurs. S273R is also present in the core of mature virus particles [115]. Whether the ASFV protease also cleaves other viral or even cellular substrates early in infection is a question that remains to be addressed.

5.2. Viruses Expressing SUMO Ligase-Like Activity

The HAdV (Human Adenoviruses) early viral protein E1B-55K was reported to possess E3 SUMO ligase functions that specifically target p53 when both factors are relocalized into PML nuclear bodies [105,106,107]. Since p53 is one of the most recognized regulators of cell cycle arrest and apoptosis, E1B-55K-induced PTM of p53 with SUMO proteins represents a key event in the oncogenic transformation of primary cells induced by adenoviral oncoproteins [105]. E1B-55K is itself a substrate for the host cell SUMO modification system, harboring a classical SCM (SUMO conjugation motif) around lysine 104 [20,144,145]. Indeed, PTM of E1B-55K is known to be involved in several aspects of viral protein function, such as functional inactivation of the tumor suppressor protein p53, proteasomal degradation of the chromatin remodeling factor Daxx, as well as oncogenic potential.
The K-bZIP protein encoded by KSHV (Kaposi’s sarcoma-associated herpes virus) is another viral protein that utilizes the SUMO pathway to alter the host cell cycle. K-bZIP is expressed after acute infection or reactivation of the latent genome and represents the structural analogue of the EBV transcription factor Zta. Interestingly, K-bZIP appears to be modified by all three SUMO isoforms. Like the adenoviral E1B-55K, K-bZIP is a strong transcriptional repressor whose activity depends on efficient SUMO conjugation [110,146]. It has been proposed that K-bZip functions as a viral SUMO ligase with a preferential selectivity towards SUMO-2/-3 paralogs, allowing recruitment of Ubc9 to KSHV promoter sites followed by transcriptional repression [111]. K-bZip induces SUMO modification of itself as well as on interaction partners such as p53 and pRB. Consistently, K-bZip is recruited to p53 target sites, dependent on the viral protein’s SIM, and affects multiple p53 downstream genes in a SUMO-dependent manner [111].

6. Conclusions

Viruses utilize host ubiquitinylation and SUMOylation processes to enhance viral replication and evade cellular immune responses. These host enzymatic cascades are involved in all physiological processes, and thus influence the balance between normal and pathogenic cellular signaling and determine the final outcome of each viral infection. Beside orchestrating PTMs on cellular and viral proteins, viruses have evolved several strategies to infiltrate such cascades, including molecular mimicry of key molecules such as ubiquitin, ubiquitin ligases, and ubiquitin and SUMO de-conjugating enzymes (see table 1). Ultimately, this modulates the host cell homeostasis and supports enhanced replication or limits an effective host response to prevent virus elimination.
Soon after the discovery of ubiquitin and SUMO, these enzymatic processes were suggested to be connected with virus pathogenesis. However, many molecular events are still not understood in detail. Some 25 years ago, ubiquitin-like genes were discovered in viral genomes that inactivate the ubiquitin-dependent antiviral defense of the host cell by acting as ubiquitin analogs [21,147]. Indeed, manipulation of the ubiquitinylation and SUMOylation systems are emerging as a key theme in viral pathogenesis, and examples of viral proteins reported to mimic or redirect such activities in order to modify the cellular environment in favor of virus persistence or efficient replication are accumulating.
Usurping ubiquitin/SUMO pathways was recently also demonstrated in proteins encoded by some pathogenic bacteria, such as SseL, a Salmonella de-ubiquitinase required for macrophage killing, and Yersinia factor YopJ, which acts as a de-ubiquitinase to inhibit NF-κB activation [148,149,150]. Together these findings imply that such enzymes play general roles in regulating pathogenic infections. However, the systematic identification of viral analogs for ubiquitin or SUMO enzymes is a complicated task due to the wide sequence variation of the known candidates. Moreover, viral proteins are often quite different in sequence and domain organization from their mammalian counterparts, hampering the identification of viral analogs.
Despite the difficulties involved and future work required, the systematic identification of viral factors that interfere with the ubiquitin/SUMO machineries will be rewarding since it will provide novel insights into unknown virus/host interactions and unravel important features of viral pathogenesis. Moreover, the information presented here highlights that ubiquitin/SUMO pathways are promising targets for development of specific antiviral therapeutic interventions.

Acknowledgments

We apologize for all the important publications and observations we could not cite/refer to in detail due to space limitations. SS is supported by grants from the Braun Stiftung, Dräger Stiftung, the Else Kröner-Fresenius-Stiftung and the Deutsche Krebshilfe e. V.. PW is a full time employee of Novartis Germany GmbH.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hunter, T.; Sun, H. Crosstalk between the SUMO and ubiquitin pathways. Ernst Scher. Found. Symp. Proc. 2009, 2008/1, 1–16. [Google Scholar]
  2. Iniguez-Lluhi, J.A. For a healthy histone code, a little SUMO in the tail keeps the acetyl away. ACS Chem. Biol. 2006, 1, 204–206. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, X.J.; Chiang, C.M. SUMOylation in gene regulation, human disease, and therapeutic action. F1000Prime Rep. 2013, 5, e45. [Google Scholar] [CrossRef] [PubMed]
  4. Wimmer, P.; Schreiner, S.; Dobner, T. Human pathogens and the host cell SUMOylation system. J. Virol. 2012, 86, 642–654. [Google Scholar] [CrossRef] [PubMed]
  5. Schlesinger, D.H.; Goldstein, G.; Niall, H.D. The complete amino acid sequence of ubiquitin, an adenylate cyclase stimulating polypeptide probably universal in living cells. Biochemistry 1975, 14, 2214–2218. [Google Scholar] [CrossRef] [PubMed]
  6. Schlesinger, D.H.; Goldstein, G. Molecular conservation of 74 amino acid sequence of ubiquitin between cattle and man. Nature 1975, 255, 423–424. [Google Scholar] [CrossRef] [PubMed]
  7. McDowell, G.S.; Philpott, A. Non-canonical ubiquitylation: Mechanisms and consequences. Int. J. Biochem. Cell Biol. 2013, 45, 1833–1842. [Google Scholar] [CrossRef] [PubMed]
  8. Voutsadakis, I.A. Ubiquitination and the ubiquitin-proteasome system as regulators of transcription and transcription factors in epithelial mesenchymal transition of cancer. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2012, 33, 897–910. [Google Scholar] [CrossRef] [PubMed]
  9. Trempe, J.F. Reading the ubiquitin postal code. Curr. Opin. Struct. Biol. 2011, 21, 792–801. [Google Scholar] [CrossRef] [PubMed]
  10. Metzger, M.B.; Hristova, V.A.; Weissman, A.M. HECT and rRING finger families of E3 ubiquitin ligases at a glance. J. Cell Sci. 2012, 125, 531–537. [Google Scholar] [CrossRef] [PubMed]
  11. Ardley, H.C.; Robinson, P.A. E3 ubiquitin ligases. Essays Biochem. 2005, 41, 15–30. [Google Scholar] [CrossRef] [PubMed]
  12. Spratt, D.E.; Walden, H.; Shaw, G.S. Rbr E3 ubiquitin ligases: New structures, new insights, new questions. Biochem. J. 2014, 458, 421–437. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, W.L.; Jin, G.; Li, C.F.; Jeong, Y.S.; Moten, A.; Xu, D.; Feng, Z.; Chen, W.; Cai, Z.; Darnay, B.; et al. Cycles of ubiquitination and deubiquitination critically regulate growth factor-mediated activation of AKT signaling. Sci. Signal. 2013, 6. [Google Scholar] [CrossRef] [PubMed]
  14. Komander, D.; Clague, M.J.; Urbe, S. Breaking the chains: Structure and function of the deubiquitinases. Nat. Rev. Mol. Cell Biol. 2009, 10, 550–563. [Google Scholar] [CrossRef] [PubMed]
  15. Hannoun, Z.; Greenhough, S.; Jaffray, E.; Hay, R.T.; Hay, D.C. Post-Translational modification by SUMO. Toxicology 2010, 278, 288–293. [Google Scholar] [CrossRef] [PubMed]
  16. Matunis, M.J.; Coutavas, E.; Blobel, G. A novel ubiquitin-like modification modulates the partitioning of the ran-gtpase-activating protein rangap1 between the cytosol and the nuclear pore complex. J. Cell Biol. 1996, 135, 1457–1470. [Google Scholar] [CrossRef] [PubMed]
  17. Tatham, M.H.; Kim, S.; Jaffray, E.; Song, J.; Chen, Y.; Hay, R.T. Unique binding interactions among Ubc9, SUMO and RanBP2 reveal a mechanism for SUMO paralog selection. Nat. Struct. Mol. Biol. 2005, 12, 67–74. [Google Scholar] [CrossRef] [PubMed]
  18. Bernier-Villamor, V.; Sampson, D.A.; Matunis, M.J.; Lima, C.D. Structural basis for E2-mediated SUMO conjugation revealed by a complex between ubiquitin-conjugating enzyme Ubc9 and RanGAP1. Cell 2002, 108, 345–356. [Google Scholar] [CrossRef]
  19. Vertegaal, A.C. Small ubiquitin-related modifiers in chains. Biochem. Soc. Trans. 2007, 35, 1422–1423. [Google Scholar] [CrossRef] [PubMed]
  20. Wimmer, P.; Blanchette, P.; Schreiner, S.; Ching, W.; Groitl, P.; Berscheminski, J.; Branton, P.E.; Will, H.; Dobner, T. Cross-Talk between phosphorylation and SUMOylation regulates transforming activities of an adenoviral oncoprotein. Oncogene 2013, 32, 1626–1637. [Google Scholar] [CrossRef] [PubMed]
  21. Guarino, L.A. Identification of a viral gene encoding a ubiquitin-like protein. Proc. Natl. Acad. Sci. USA 1990, 87, 409–413. [Google Scholar] [CrossRef] [PubMed]
  22. Dixon, L.K.; Chapman, D.A.; Netherton, C.L.; Upton, C. African swine fever virus replication and genomics. Virus Res. 2013, 173, 3–14. [Google Scholar] [CrossRef] [PubMed]
  23. Hingamp, P.M.; Arnold, J.E.; Mayer, R.J.; Dixon, L.K. A ubiquitin conjugating enzyme encoded by african swine fever virus. EMBO J. 1992, 11, 361–366. [Google Scholar] [PubMed]
  24. Hingamp, P.M.; Leyland, M.L.; Webb, J.; Twigger, S.; Mayer, R.J.; Dixon, L.K. Characterization of a ubiquitinated protein which is externally located in african swine fever virions. J. Virol. 1995, 69, 1785–1793. [Google Scholar] [PubMed]
  25. Schreiner, S.; Wimmer, P.; Dobner, T. Adenovirus degradation of cellular proteins. Future Microbiol. 2012, 7, 211–225. [Google Scholar] [CrossRef] [PubMed]
  26. Baker, A.; Rohleder, K.J.; Hanakahi, L.A.; Ketner, G. Adenovirus E4 34k and E1b 55k oncoproteins target host DNA ligase IV for proteasomal degradation. J. Virol. 2007, 81, 7034–7040. [Google Scholar] [CrossRef] [PubMed]
  27. Querido, E.; Blanchette, P.; Yan, Q.; Kamura, T.; Morrison, M.; Boivin, D.; Kaelin, W.G.; Conaway, R.C.; Conaway, J.W.; Branton, P.E. Degradation of p53 by adenovirus E4orf6 and E1B55K proteins occurs via a novel mechanism involving a cullin-containing complex. Genes Dev. 2001, 15, 3104–3117. [Google Scholar] [CrossRef] [PubMed]
  28. Stracker, T.H.; Carson, C.T.; Weitzman, M.D. Adenovirus oncoproteins inactivate the Mre11 Rad50 NBS1 DNA repair complex. Nature 2002, 418, 348–352. [Google Scholar] [CrossRef] [PubMed]
  29. Orazio, N.I.; Naeger, C.M.; Karlseder, J.; Weitzman, M.D. The adenovirus e1b55k/e4orf6 complex induces degradation of the bloom helicase during infection. J. Virol. 2011, 85, 1887–1892. [Google Scholar] [CrossRef] [PubMed]
  30. Dallaire, F.; Blanchette, P.; Groitl, P.; Dobner, T.; Branton, P.E. Identification of integrin alpha3 as a new substrate of the adenovirus E4orf6/E1B 55-kilodalton e3 ubiquitin ligase complex. J. Virol. 2009, 83, 5329–5338. [Google Scholar] [CrossRef] [PubMed]
  31. Schreiner, S.; Kinkley, S.; Bürck, C.; Mund, A.; Wimmer, P.; Schubert, T.; Groitl, P.; Will, H.; Dobner, T. SPOC1-mediated antiviral host cell response is antagonized early in human adenovirus type 5 infection. PLoS Pathog. 2013, 9, e1003775. [Google Scholar] [CrossRef] [PubMed]
  32. Schreiner, S.; Bürck, C.; Glass, M.; Groitl, P.; Wimmer, P.; Kinkley, S.; Mund, A.; Everett, R.D.; Dobner, T. Control of human adenovirus type 5 (Ad5) gene expression by cellular Daxx/ATRX chromatin-associated complexes. Nucleic Acids Res. 2013, 41, 3532–3550. [Google Scholar] [CrossRef] [PubMed]
  33. Gupta, A.; Jha, S.; Engel, D.A.; Ornelles, D.A.; Dutta, A. Tip60 degradation by adenovirus relieves transcriptional repression of viral transcriptional activator E1A. Oncogene 2013, 32, 5017–5025. [Google Scholar] [CrossRef] [PubMed]
  34. Blackford, A.N.; Patel, R.N.; Forrester, N.A.; Theil, K.; Groitl, P.; Stewart, G.S.; Taylor, A.M.; Morgan, I.M.; Dobner, T.; Grand, R.J.; et al. Adenovirus 12 E4orf6 inhibits atr activation by promoting topbp1 degradation. Proc. Natl. Acad. Sci. USA 2010, 107, 12251–12256. [Google Scholar] [CrossRef] [PubMed]
  35. Schreiner, S.; Wimmer, P.; Sirma, H.; Everett, R.D.; Blanchette, P.; Groitl, P.; Dobner, T. Proteasome-dependent degradation of daxx by the viral E1b-55K protein in human adenovirus-infected cells. J. Virol. 2010, 84, 7029–7038. [Google Scholar] [CrossRef] [PubMed]
  36. Cheng, C.Y.; Gilson, T.; Dallaire, F.; Ketner, G.; Branton, P.E.; Blanchette, P. The E4orf6/E1B55k E3 ubiquitin ligase complexes of human adenoviruses exhibit heterogenity in composition and substrate specificity. J. Virol. 2011, 85, 765–775. [Google Scholar] [CrossRef] [PubMed]
  37. Forrester, N.A.; Sedgwick, G.G.; Thomas, A.; Blackford, A.N.; Speiseder, T.; Dobner, T.; Byrd, P.J.; Stewart, G.S.; Turnell, A.S.; Grand, R.J. Serotype-specific inactivation of the cellular DNA damage response during adenovirus infection. J. Virol. 2010, 85, 2201–2211. [Google Scholar] [CrossRef] [PubMed]
  38. Coscoy, L.; Ganem, D. Kaposi's sarcoma-associated herpesvirus encodes two proteins that block cell surface display of MHC class I chains by enhancing their endocytosis. Proc. Natl. Acad. Sci. USA 2000, 97, 8051–8056. [Google Scholar] [CrossRef] [PubMed]
  39. Haque, M.; Ueda, K.; Nakano, K.; Hirata, Y.; Parravicini, C.; Corbellino, M.; Yamanishi, K. Major histocompatibility complex class i molecules are down-regulated at the cell surface by the K5 protein encoded by kaposi's sarcoma-associated herpesvirus/human herpesvirus-8. J. Gen. Virol. 2001, 82, 1175–1180. [Google Scholar] [PubMed]
  40. Boname, J.M.; Lehner, P.J. What has the study of the K3 and K5 viral ubiquitin E3 ligases taught us about ubiquitin-mediated receptor regulation? Viruses 2011, 3, 118–131. [Google Scholar] [CrossRef] [PubMed]
  41. Duncan, L.M.; Piper, S.; Dodd, R.B.; Saville, M.K.; Sanderson, C.M.; Luzio, J.P.; Lehner, P.J. Lysine-63-linked ubiquitination is required for endolysosomal degradation of class I molecules. EMBO J. 2006, 25, 1635–1645. [Google Scholar] [CrossRef] [PubMed]
  42. Bruce, A.G.; Ryan, J.T.; Thomas, M.J.; Peng, X.; Grundhoff, A.; Tsai, C.C.; Rose, T.M. Next-generation sequence analysis of the genome of rfhvmn, the macaque homolog of Kaposi's Sarcoma (KS)-associated herpesvirus, from a KS-like tumor of a pig-tailed macaque. J. Virol. 2013, 87, 13676–13693. [Google Scholar] [CrossRef] [PubMed]
  43. Virgin, H.W.T.; Latreille, P.; Wamsley, P.; Hallsworth, K.; Weck, K.E.; Dal Canto, A.J.; Speck, S.H. Complete sequence and genomic analysis of murine gammaherpesvirus 68. J. Virol. 1997, 71, 5894–5904. [Google Scholar] [PubMed]
  44. Brulois, K.; Jung, J.U. Interplay between kaposi's sarcoma-associated herpesvirus and the innate immune system. Cytokine Growth Factor Rev. 2014, 25, 597–609. [Google Scholar] [CrossRef] [PubMed]
  45. Searles, R.P.; Bergquam, E.P.; Axthelm, M.K.; Wong, S.W. Sequence and genomic analysis of a rhesus macaque rhadinovirus with similarity to kaposi's sarcoma-associated herpesvirus/human herpesvirus 8. J. Virol. 1999, 73, 3040–3053. [Google Scholar] [PubMed]
  46. Ishido, S.; Choi, J.K.; Lee, B.S.; Wang, C.; DeMaria, M.; Johnson, R.P.; Cohen, G.B.; Jung, J.U. Inhibition of natural killer cell-mediated cytotoxicity by kaposi's sarcoma-associated herpesvirus K5 protein. Immunity 2000, 13, 365–374. [Google Scholar] [CrossRef]
  47. Stevenson, P.G.; May, J.S.; Smith, X.G.; Marques, S.; Adler, H.; Koszinowski, U.H.; Simas, J.P.; Efstathiou, S. K3-mediated evasion of CD8(+) T cells aids amplification of a latent gamma-herpesvirus. Nat. Immunol. 2002, 3, 733–740. [Google Scholar] [PubMed]
  48. Huibregtse, J.M.; Scheffner, M.; Howley, P.M. A cellular protein mediates association of p53 with the E6 oncoprotein of human papillomavirus types 16 or 18. EMBO J. 1991, 10, 4129–4135. [Google Scholar] [PubMed]
  49. Scheffner, M.; Huibregtse, J.M.; Howley, P.M. Identification of a human ubiquitin-conjugating enzyme that mediates the E6-AP-dependent ubiquitination of p53. Proc. Natl. Acad. Sci. USA 1994, 91, 8797–8801. [Google Scholar] [CrossRef] [PubMed]
  50. Nuber, U.; Schwarz, S.; Kaiser, P.; Schneider, R.; Scheffner, M. Cloning of human ubiquitin-conjugating enzymes Ubch6 and Ubch7 (e2-f1) and characterization of their interaction with E6-AP and Rsp5. J. Biol. Chem. 1996, 271, 2795–2800. [Google Scholar] [CrossRef] [PubMed]
  51. Scheffner, M.; Nuber, U.; Huibregtse, J.M. Protein ubiquitination involving an E1-E2-E3 enzyme ubiquitin thioester cascade. Nature 1995, 373, 81–83. [Google Scholar] [CrossRef] [PubMed]
  52. Scheffner, M.; Takahashi, T.; Huibregtse, J.M.; Minna, J.D.; Howley, P.M. Interaction of the human papillomavirus type 16 E6 oncoprotein with wild-type and mutant human p53 proteins. J. Virol. 1992, 66, 5100–5105. [Google Scholar] [PubMed]
  53. Band, V.; De Caprio, J.A.; Delmolino, L.; Kulesa, V.; Sager, R. Loss of p53 protein in human papillomavirus type 16 E6-immortalized human mammary epithelial cells. J. Virol. 1991, 65, 6671–6676. [Google Scholar] [PubMed]
  54. Gross-Mesilaty, S.; Reinstein, E.; Bercovich, B.; Tobias, K.E.; Schwartz, A.L.; Kahana, C.; Ciechanover, A. Basal and human papillomavirus e6 oncoprotein-induced degradation of Myc proteins by the ubiquitin pathway. Proc. Natl. Acad. Sci. USA 1998, 95, 8058–8063. [Google Scholar] [CrossRef] [PubMed]
  55. Kuhne, C.; Banks, L. E3-ubiquitin ligase/E6-AP links multicopy maintenance protein 7 to the ubiquitination pathway by a novel motif, the l2g box. J. Biol. Chem. 1998, 273, 34302–34309. [Google Scholar] [CrossRef] [PubMed]
  56. Thomas, M.; Banks, L. Inhibition of Bak-induced apoptosis by HPV-18 E6. Oncogene 1998, 17, 2943–2954. [Google Scholar] [CrossRef] [PubMed]
  57. Gao, Q.; Srinivasan, S.; Boyer, S.N.; Wazer, D.E.; Band, V. The E6 oncoproteins of high-risk papillomaviruses bind to a novel putative gap protein, E6tp1, and target it for degradation. Mol. Cell. Biol. 1999, 19, 733–744. [Google Scholar] [PubMed]
  58. Gewin, L.; Myers, H.; Kiyono, T.; Galloway, D.A. Identification of a novel telomerase repressor that interacts with the human papillomavirus type-16 E6/E6-AP complex. Genes Dev. 2004, 18, 2269–2282. [Google Scholar] [CrossRef] [PubMed]
  59. Berezutskaya, E.; Yu, B.; Morozov, A.; Raychaudhuri, P.; Bagchi, S. Differential regulation of the pocket domains of the retinoblastoma family proteins by the HPV 16 E7 oncoprotein. Cell. Growth Differ. 1997, 8, 1277–1286. [Google Scholar] [PubMed]
  60. Boyer, S.N.; Wazer, D.E.; Band, V. E7 protein of human papilloma virus-16 induces degradation of retinoblastoma protein through the ubiquitin-proteasome pathway. Cancer Res. 1996, 56, 4620–4624. [Google Scholar] [PubMed]
  61. Huh, K.; Zhou, X.; Hayakawa, H.; Cho, J.Y.; Libermann, T.A.; Jin, J.; Harper, J.W.; Munger, K. Human papillomavirus type 16 E7 oncoprotein associates with the cullin 2 ubiquitin ligase complex, which contributes to degradation of the retinoblastoma tumor suppressor. J. Virol. 2007, 81, 9737–9747. [Google Scholar] [CrossRef] [PubMed]
  62. 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] [PubMed]
  63. 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]
  64. Mehle, A.; Strack, B.; Ancuta, P.; Zhang, C.; McPike, M.; Gabuzda, D. Vif overcomes the innate antiviral activity of APOBEC3G by promoting its degradation in the ubiquitin-proteasome pathway. J. Biol. Chem. 2004, 279, 7792–7798. [Google Scholar] [CrossRef] [PubMed]
  65. Mehle, A.; Thomas, E.R.; Rajendran, K.S.; Gabuzda, D. A zinc-binding region in vif binds Cul5 and determines cullin selection. J. Biol. Chem. 2006, 281, 17259–17265. [Google Scholar] [CrossRef] [PubMed]
  66. Balliet, J.W.; Kolson, D.L.; Eiger, G.; Kim, F.M.; McGann, K.A.; Srinivasan, A.; Collman, R. Distinct effects in primary macrophages and lymphocytes of the human immunodeficiency virus type 1 accessory genes vpr, vpu, and nef: Mutational analysis of a primary HIV-1 isolate. Virology 1994, 200, 623–631. [Google Scholar] [CrossRef] [PubMed]
  67. Jowett, J.B.; Planelles, V.; Poon, B.; Shah, N.P.; Chen, M.L.; Chen, I.S. The human immunodeficiency virus type 1 Vpr gene arrests infected t cells in the G2 + M phase of the cell cycle. J. Virol. 1995, 69, 6304–6313. [Google Scholar] [PubMed]
  68. Rogel, M.E.; Wu, L.I.; Emerman, M. The human immunodeficiency virus type 1 Vpr gene prevents cell proliferation during chronic infection. J. Virol. 1995, 69, 882–888. [Google Scholar] [PubMed]
  69. Casey, L.; Wen, X.; de Noronha, C.M. The functions of the HIV1 protein Vpr and its action through the DCAF1.DDB1.Cullin4 ubiquitin ligase. Cytokine 2010, 51, 1–9. [Google Scholar] [CrossRef] [PubMed]
  70. Shiyanov, P.; Nag, A.; Raychaudhuri, P. Cullin 4a associates with the uv-damaged DNA-binding protein DDB. J. Biol. Chem. 1999, 274, 35309–35312. [Google Scholar] [CrossRef] [PubMed]
  71. Schrofelbauer, B.; Yu, Q.; Zeitlin, S.G.; Landau, N.R. Human immunodeficiency virus type 1 Vpr induces the degradation of the UNG and SMUG uracil-DNA glycosylases. J. Virol. 2005, 79, 10978–10987. [Google Scholar] [CrossRef] [PubMed]
  72. Okumura, A.; Alce, T.; Lubyova, B.; Ezelle, H.; Strebel, K.; Pitha, P.M. HIV-1 accessory proteins Vpr and Vif modulate antiviral response by targeting IRF-3 for degradation. Virology 2008, 373, 85–97. [Google Scholar] [CrossRef] [PubMed]
  73. Fenner, F. A successful eradication campaign. Global eradication of smallpox. Rev. Infect. Dis 1982, 4, 916–930. [Google Scholar] [CrossRef] [PubMed]
  74. Seet, B.T.; Johnston, J.B.; Brunetti, C.R.; Barrett, J.W.; Everett, H.; Cameron, C.; Sypula, J.; Nazarian, S.H.; Lucas, A.; McFadden, G. Poxviruses and immune evasion. Annu. Rev. Immunol 2003, 21, 377–423. [Google Scholar] [CrossRef] [PubMed]
  75. Aguado, B.; Selmes, I.P.; Smith, G.L. Nucleotide sequence of 21.8 kbp of variola major virus strain harvey and comparison with vaccinia virus. J. Gen. Virol. 1992, 73, 2887–2902. [Google Scholar] [CrossRef] [PubMed]
  76. Assarsson, E.; Greenbaum, J.A.; Sundstrom, M.; Schaffer, L.; Hammond, J.A.; Pasquetto, V.; Oseroff, C.; Hendrickson, R.C.; Lefkowitz, E.J.; Tscharke, D.C.; et al. Kinetic analysis of a complete poxvirus transcriptome reveals an immediate-early class of genes. Proc. Natl. Acad. Sci. USA 2008, 105, 2140–2145. [Google Scholar] [CrossRef] [PubMed]
  77. Mansur, D.S.; Maluquer de Motes, C.; Unterholzner, L.; Sumner, R.P.; Ferguson, B.J.; Ren, H.; Strnadova, P.; Bowie, A.G.; Smith, G.L. Poxvirus targeting of E3 ligase beta-trcp by molecular mimicry: A mechanism to inhibit NFkB activation and promote immune evasion and virulence. PLoS Pathog. 2013, 9, e1003183. [Google Scholar] [CrossRef] [PubMed]
  78. Tate, J.E.; Burton, A.H.; Boschi-Pinto, C.; Steele, A.D.; Duque, J.; Parashar, U.D.; WHO-coordinated Global Rotavirus Surveillance Network. 2008 estimate of worldwide rotavirus-associated mortality in children younger than 5 years before the introduction of universal rotavirus vaccination programmes: A systematic review and meta-analysis. Lancet Infect. Dis. 2012, 12, 136–141. [Google Scholar] [CrossRef]
  79. Desselberger, U. Rotaviruses. Virus Res. 2014, 190, 75–96. [Google Scholar] [CrossRef] [PubMed]
  80. Broquet, A.H.; Hirata, Y.; McAllister, C.S.; Kagnoff, M.F. RIG-I/MDA5/MAVS are required to signal a protective ifn response in rotavirus-infected intestinal epithelium. J. Immunol. 2011, 186, 1618–1626. [Google Scholar] [CrossRef] [PubMed]
  81. Uzri, D.; Greenberg, H.B. Characterization of rotavirus RNAs that activate innate immune signaling through the RIG-I-like receptors. PLoS ONE 2013, 8, e69825. [Google Scholar] [CrossRef] [PubMed]
  82. Sen, A.; Pruijssers, A.J.; Dermody, T.S.; Garcia-Sastre, A.; Greenberg, H.B. The early interferon response to rotavirus is regulated by PKR and depends on MAVS/IPS-1, RIG-I, MDA5, and IRF3. J. Virol. 2011, 85, 3717–3732. [Google Scholar] [CrossRef] [PubMed]
  83. Graff, J.W.; Ettayebi, K.; Hardy, M.E. Rotavirus NSP1 inhibits nfkappab activation by inducing proteasome-dependent degradation of beta-TRCP: A novel mechanism of ifn antagonism. PLoS Pathog. 2009, 5, e1000280. [Google Scholar] [CrossRef] [PubMed]
  84. Bagchi, P.; Bhowmick, R.; Nandi, S.; Kant Nayak, M.; Chawla-Sarkar, M. Rotavirus nsp1 inhibits interferon induced non-canonical NFkB activation by interacting with TNF receptor associated factor 2. Virology 2013, 444, 41–44. [Google Scholar] [CrossRef] [PubMed]
  85. Sriramachandran, A.M.; Dohmen, R.J. SUMO-targeted ubiquitin ligases. Biochim. Biophys. Acta 2014, 1843, 75–85. [Google Scholar] [CrossRef] [PubMed]
  86. Tatham, M.H.; Geoffroy, M.C.; Shen, L.; Plechanovova, A.; Hattersley, N.; Jaffray, E.G.; Palvimo, J.J.; Hay, R.T. RNF4 is a poly-SUMO-specific E3 ubiquitin ligase required for arsenic-induced PML degradation. Nat. Cell Biol. 2008, 10, 538–546. [Google Scholar] [CrossRef] [PubMed]
  87. Sun, H.; Leverson, J.D.; Hunter, T. Conserved function of RNF4 family proteins in eukaryotes: Targeting a ubiquitin ligase to SUMOylated proteins. EMBO J. 2007, 26, 4102–4112. [Google Scholar] [CrossRef] [PubMed]
  88. Sun, H.; Hunter, T. Poly-small ubiquitin-like modifier (polysumo)-binding proteins identified through a string search. J. Biol. Chem. 2012, 287, 42071–42083. [Google Scholar] [CrossRef] [PubMed]
  89. Boutell, C.; Orr, A.; Everett, R.D. PML residue lysine 160 is required for the degradation of PML induced by herpes simplex virus type 1 regulatory protein ICP0. J. Virol. 2003, 77, 8686–8694. [Google Scholar] [CrossRef] [PubMed]
  90. Boutell, C.; Cuchet-Lourenco, D.; Vanni, E.; Orr, A.; Glass, M.; McFarlane, S.; Everett, R.D. A viral ubiquitin ligase has substrate preferential SUMO targeted ubiquitin ligase activity that counteracts intrinsic antiviral defence. PLoS Pathog. 2011, 7, e1002245. [Google Scholar] [CrossRef] [PubMed]
  91. Chaurushiya, M.S.; Lilley, C.E.; Aslanian, A.; Meisenhelder, J.; Scott, D.C.; Landry, S.; Ticau, S.; Boutell, C.; Yates, J.R., 3rd; Schulman, B.A.; et al. Viral E3 ubiquitin ligase-mediated degradation of a cellular E3: Viral mimicry of a cellular phosphorylation mark targets the RNF8 fha domain. Mol. Cell 2012, 46, 79–90. [Google Scholar] [CrossRef] [PubMed]
  92. Kyratsous, C.A.; Silverstein, S.J. Components of nuclear domain 10 bodies regulate varicella-zoster virus replication. J. Virol. 2009, 83, 4262–4274. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, L.; Oliver, S.L.; Sommer, M.; Rajamani, J.; Reichelt, M.; Arvin, A.M. Disruption of PML nuclear bodies is mediated by orf61 SUMO-interacting motifs and required for varicella-zoster virus pathogenesis in skin. PLoS Pathog. 2011, 7, e1002157. [Google Scholar] [CrossRef] [PubMed]
  94. Walters, M.S.; Kyratsous, C.A.; Silverstein, S.J. The ring finger domain of varicella-zoster virus orf61p has E3 ubiquitin ligase activity that is essential for efficient autoubiquitination and dispersion of sp100-containing nuclear bodies. J. Virol. 2010, 84, 6861–6865. [Google Scholar] [CrossRef] [PubMed]
  95. Zhu, H.; Zheng, C.; Xing, J.; Wang, S.; Li, S.; Lin, R.; Mossman, K.L. Varicella-zoster virus immediate-early protein orf61 abrogates the IRF3-mediated innate immune response through degradation of activated IRF3. J. Virol. 2011, 85, 11079–11089. [Google Scholar] [CrossRef] [PubMed]
  96. Ling, P.D.; Tan, J.; Sewatanon, J.; Peng, R. Murine gammaherpesvirus 68 open reading frame 75c tegument protein induces the degradation of PML and is essential for production of infectious virus. J. Virol. 2008, 82, 8000–8012. [Google Scholar] [CrossRef] [PubMed]
  97. Mangel, W.F.; Baniecki, M.L.; McGrath, W.J. Specific interactions of the adenovirus proteinase with the viral DNA, an 11-amino-acid viral peptide, and the cellular protein actin. Cell. Mol. Life Sci. 2003, 60, 2347–2355. [Google Scholar] [CrossRef] [PubMed]
  98. Anderson, C.W. The proteinase polypeptide of adenovirus serotype 2 virions. Virology 1990, 177, 259–272. [Google Scholar] [CrossRef]
  99. Weber, J.M. Adenovirus endopeptidase and its role in virus infection. Curr. Top. Microbiol. Immunol. 1995, 199, 227–235. [Google Scholar] [PubMed]
  100. Cotten, M.; Weber, J.M. The adenovirus protease is required for virus entry into host cells. Virology 1995, 213, 494–502. [Google Scholar] [CrossRef] [PubMed]
  101. Balakirev, M.Y.; Jaquinod, M.; Haas, A.L.; Chroboczek, J. Deubiquitinating function of adenovirus proteinase. J. Virol. 2002, 76, 6323–6331. [Google Scholar] [CrossRef] [PubMed]
  102. Chen, P.H.; Ornelles, D.A.; Shenk, T. The adenovirus l3 23-kilodalton proteinase cleaves the amino-terminal head domain from cytokeratin 18 and disrupts the cytokeratin network of hela cells. J. Virol. 1993, 67, 3507–3514. [Google Scholar] [PubMed]
  103. Li, S.J.; Hochstrasser, M. A new protease required for cell-cycle progression in yeast. Nature 1999, 398, 246–251. [Google Scholar] [PubMed]
  104. Mukhopadhyay, D.; Dasso, M. Modification in reverse: The SUMO proteases. Trends Biochem. Sci. 2007, 32, 286–295. [Google Scholar] [CrossRef] [PubMed]
  105. Wimmer, P.; Berscheminski, J.; Blanchette, P.; Groitl, P.; Branton, P.E.; Hay, R.T.; Dobner, D.; Schreiner, S. Pml isoforms IV and V contribute to adenovirus-mediated oncogenic transformation by functional inhibition of the tumor suppressor p53. Oncogene 2015. manuscript accepted. [Google Scholar] [CrossRef] [PubMed]
  106. Pennella, M.A.; Liu, Y.; Woo, J.L.; Kim, C.A.; Berk, A.J. Adenovirus e1b 55-kilodalton protein is a p53-SUMO1 E3 ligase that represses p53 and stimulates its nuclear export through interactions with promyelocytic leukemia nuclear bodies. J. Virol. 2010, 84, 12210–12225. [Google Scholar] [CrossRef] [PubMed]
  107. Muller, S.; Dobner, T. The adenovirus E1B-55K oncoprotein induces sumo modification of p53. Cell Cycle 2008, 7, 754–758. [Google Scholar] [CrossRef] [PubMed]
  108. Kattenhorn, L.M.; Korbel, G.A.; Kessler, B.M.; Spooner, E.; Ploegh, H.L. A deubiquitinating enzyme encoded by HSV-1 belongs to a family of cysteine proteases that is conserved across the family herpesviridae. Mol. Cell 2005, 19, 547–557. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, J.; Loveland, A.N.; Kattenhorn, L.M.; Ploegh, H.L.; Gibson, W. High-molecular-weight protein (pul48) of human cytomegalovirus is a competent deubiquitinating protease: Mutant viruses altered in its active-site cysteine or histidine are viable. J. Virol. 2006, 80, 6003–6012. [Google Scholar] [CrossRef] [PubMed]
  110. Izumiya, Y.; Ellison, T.J.; Yeh, E.T.; Jung, J.U.; Luciw, P.A.; Kung, H.J. Kaposi's sarcoma-associated herpesvirus K-bZip represses gene transcription via sumo modification. J. Virol. 2005, 79, 9912–9925. [Google Scholar] [CrossRef] [PubMed]
  111. Chang, P.C.; Izumiya, Y.; Wu, C.Y.; Fitzgerald, L.D.; Campbell, M.; Ellison, T.J.; Lam, K.S.; Luciw, P.A.; Kung, H.J. Kaposi's sarcoma-associated herpesvirus (KSHV) encodes a SUMO E3 ligase that is SIM-dependent and SUMO-2/3-specific. J. Biol. Chem. 2010, 285, 5266–5273. [Google Scholar] [CrossRef] [PubMed]
  112. Sompallae, R.; Gastaldello, S.; Hildebrand, S.; Zinin, N.; Hassink, G.; Lindsten, K.; Haas, J.; Persson, B.; Masucci, M.G. Epstein-barr virus encodes three bona fide ubiquitin-specific proteases. J. Virol. 2008, 82, 10477–10486. [Google Scholar] [CrossRef] [PubMed]
  113. Gredmark, S.; Schlieker, C.; Quesada, V.; Spooner, E.; Ploegh, H.L. A functional ubiquitin-specific protease embedded in the large tegument protein (orf64) of murine gammaherpesvirus 68 is active during the course of infection. J. Virol. 2007, 81, 10300–10309. [Google Scholar] [CrossRef] [PubMed]
  114. Fuchs, W.; Klupp, B.G.; Granzow, H.; Mettenleiter, T.C. Essential function of the pseudorabies virus UL36 gene product is independent of its interaction with the UL37 protein. J. Virol. 2004, 78, 11879–11889. [Google Scholar] [CrossRef] [PubMed]
  115. Andres, G.; Alejo, A.; Simon-Mateo, C.; Salas, M.L. African swine fever virus protease, a new viral member of the SUMO-1-specific protease family. J. Biol. Chem. 2001, 276, 780–787. [Google Scholar] [CrossRef] [PubMed]
  116. Honig, J.E.; Osborne, J.C.; Nichol, S.T. Crimean-congo hemorrhagic fever virus genome l RNA segment and encoded protein. Virology 2004, 321, 29–35. [Google Scholar] [CrossRef] [PubMed]
  117. Barretto, N.; Jukneliene, D.; Ratia, K.; Chen, Z.; Mesecar, A.D.; Baker, S.C. The papain-like protease of severe acute respiratory syndrome coronavirus has deubiquitinating activity. J. Virol. 2005, 79, 15189–15198. [Google Scholar] [CrossRef] [PubMed]
  118. Lindner, H.A.; Fotouhi-Ardakani, N.; Lytvyn, V.; Lachance, P.; Sulea, T.; Menard, R. The papain-like protease from the severe acute respiratory syndrome coronavirus is a deubiquitinating enzyme. J. Virol. 2005, 79, 15199–15208. [Google Scholar] [CrossRef] [PubMed]
  119. Kass, R.S.; Arena, J.P.; Chin, S. Block of l-type calcium channels by charged dihydropyridines. Sensitivity to side of application and calcium. J. Gen. Physiol. 1991, 98, 63–75. [Google Scholar] [CrossRef] [PubMed]
  120. Clementz, M.A.; Chen, Z.; Banach, B.S.; Wang, Y.; Sun, L.; Ratia, K.; Baez-Santos, Y.M.; Wang, J.; Takayama, J.; Ghosh, A.K.; et al. Deubiquitinating and interferon antagonism activities of coronavirus papain-like proteases. J. Virol. 2010, 84, 4619–4629. [Google Scholar] [CrossRef] [PubMed]
  121. Mielech, A.M.; Kilianski, A.; Baez-Santos, Y.M.; Mesecar, A.D.; Baker, S.C. Mers-CoV papain-like protease has deisgylating and deubiquitinating activities. Virology 2014, 450, 64–70. [Google Scholar] [CrossRef] [PubMed]
  122. Amerik, A.Y.; Hochstrasser, M. Mechanism and function of deubiquitinating enzymes. Biochim. Biophys. Acta 2004, 1695, 189–207. [Google Scholar] [CrossRef] [PubMed]
  123. Schlieker, C.; Korbel, G.A.; Kattenhorn, L.M.; Ploegh, H.L. A deubiquitinating activity is conserved in the large tegument protein of the herpesviridae. J. Virol. 2005, 79, 15582–15585. [Google Scholar] [CrossRef] [PubMed]
  124. Das, S.; Ortiz, D.A.; Gurczynski, S.J.; Khan, F.; Pellett, P.E. Identification of human cytomegalovirus genes important for biogenesis of the cytoplasmic virion assembly complex. J. Virol. 2014, 88, 9086–9099. [Google Scholar] [CrossRef] [PubMed]
  125. Granzow, H.; Klupp, B.G.; Mettenleiter, T.C. Entry of pseudorabies virus: An immunogold-labeling study. J. Virol. 2005, 79, 3200–3205. [Google Scholar] [CrossRef] [PubMed]
  126. Antinone, S.E.; Smith, G.A. Two modes of herpesvirus trafficking in neurons: Membrane acquisition directs motion. J. Virol. 2006, 80, 11235–11240. [Google Scholar] [CrossRef] [PubMed]
  127. Luxton, G.W.; Haverlock, S.; Coller, K.E.; Antinone, S.E.; Pincetic, A.; Smith, G.A. Targeting of herpesvirus capsid transport in axons is coupled to association with specific sets of tegument proteins. Proc. Natl. Acad. Sci. USA 2005, 102, 5832–5837. [Google Scholar] [CrossRef] [PubMed]
  128. Gill, M.B.; Kutok, J.L.; Fingeroth, J.D. Epstein-barr virus thymidine kinase is a centrosomal resident precisely localized to the periphery of centrioles. J. Virol. 2007, 81, 6523–6535. [Google Scholar] [CrossRef] [PubMed]
  129. Bottcher, S.; Maresch, C.; Granzow, H.; Klupp, B.G.; Teifke, J.P.; Mettenleiter, T.C. Mutagenesis of the active-site cysteine in the ubiquitin-specific protease contained in large tegument protein pUL36 of pseudorabies virus impairs viral replication in vitro and neuroinvasion in vivo. J. Virol. 2008, 82, 6009–6016. [Google Scholar] [CrossRef] [PubMed]
  130. Burt, F.J.; Leman, P.A.; Abbott, J.C.; Swanepoel, R. Serodiagnosis of crimean-congo haemorrhagic fever. Epidemiol. Infect. 1994, 113, 551–562. [Google Scholar] [CrossRef] [PubMed]
  131. Deyde, V.M.; Khristova, M.L.; Rollin, P.E.; Ksiazek, T.G.; Nichol, S.T. Crimean-congo hemorrhagic fever virus genomics and global diversity. J. Virol. 2006, 80, 8834–8842. [Google Scholar] [CrossRef] [PubMed]
  132. Marriott, A.C.; Nuttall, P.A. Large rna segment of dugbe nairovirus encodes the putative RNA polymerase. J. Gen. Virol. 1996, 77, 1775–1780. [Google Scholar] [CrossRef] [PubMed]
  133. Lawrence, D.M.; Rozanov, M.N.; Hillman, B.I. Autocatalytic processing of the 223-kDa protein of blueberry scorch carlavirus by a papain-like proteinase. Virology 1995, 207, 127–135. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, H.M.; Yang, J.; Sun, H.R.; Xin, X.; Wang, H.D.; Chen, J.P.; Adams, M.J. Genomic analysis of rice stripe virus zhejiang isolate shows the presence of an OTU-like domain in the RNA1 protein and a novel sequence motif conserved within the intergenic regions of ambisense segments of tenuiviruses. Arch. Virol. 2007, 152, 1917–1923. [Google Scholar] [CrossRef] [PubMed]
  135. Snijder, E.J.; Bredenbeek, P.J.; Dobbe, J.C.; Thiel, V.; Ziebuhr, J.; Poon, L.L.; Guan, Y.; Rozanov, M.; Spaan, W.J.; Gorbalenya, A.E. Unique and conserved features of genome and proteome of SARS-coronavirus, an early split-off from the coronavirus group 2 lineage. J. Mol. Biol. 2003, 331, 991–1004. [Google Scholar] [CrossRef]
  136. Kiel, C.; Serrano, L. The ubiquitin domain superfold: Structure-based sequence alignments and characterization of binding epitopes. J. Mol. Biol. 2006, 355, 821–844. [Google Scholar] [CrossRef] [PubMed]
  137. Ratia, K.; Saikatendu, K.S.; Santarsiero, B.D.; Barretto, N.; Baker, S.C.; Stevens, R.C.; Mesecar, A.D. Severe acute respiratory syndrome coronavirus papain-like protease: Structure of a viral deubiquitinating enzyme. Proc. Natl. Acad. Sci. USA 2006, 103, 5717–5722. [Google Scholar] [CrossRef] [PubMed]
  138. Morales, D.J.; Lenschow, D.J. The antiviral activities of isg15. J. Mol. Biol. 2013, 425, 4995–5008. [Google Scholar] [CrossRef] [PubMed]
  139. Sawicki, S.G.; Sawicki, D.L. The effect of overproduction of nonstructural proteins on alphavirus plus-strand and minus-strand RNA synthesis. Virology 1986, 152, 507–512. [Google Scholar] [CrossRef]
  140. Yamamoto, D.; Ishida, T.; Inoue, M. A comparison between the binding modes of a substrate and inhibitor to papain as observed in complex crystal structures. Biochem. Biophys. Res. Commun. 1990, 171, 711–716. [Google Scholar] [CrossRef]
  141. Kane, E.M.; Shuman, S. Vaccinia virus morphogenesis is blocked by a temperature-sensitive mutation in the I7 gene that encodes a virion component. J. Virol. 1993, 67, 2689–2698. [Google Scholar] [PubMed]
  142. Simon-Mateo, C.; Andres, G.; Vinuela, E. Polyprotein processing in african swine fever virus: A novel gene expression strategy for a DNA virus. EMBO J. 1993, 12, 2977–2987. [Google Scholar] [PubMed]
  143. Lopez-Otin, C.; Simon-Mateo, C.; Martinez, L.; Vinuela, E. Gly-Gly-x, a novel consensus sequence for the proteolytic processing of viral and cellular proteins. J. Biol. Chem. 1989, 264, 9107–9110. [Google Scholar] [PubMed]
  144. Endter, C.; Kzhyshkowska, J.; Stauber, R.; Dobner, T. Sumo-1 modification required for transformation by adenovirus type 5 early region 1b 55-kDa oncoprotein. Proc. Natl. Acad. Sci. USA 2001, 98, 11312–11317. [Google Scholar] [CrossRef] [PubMed]
  145. Kindsmuller, K.; Groitl, P.; Hartl, B.; Blanchette, P.; Hauber, J.; Dobner, T. Intranuclear targeting and nuclear export of the adenovirus E1B-55K protein are regulated by SUMO1 conjugation. Proc. Natl. Acad. Sci. USA 2007, 104, 6684–6689. [Google Scholar] [CrossRef] [PubMed]
  146. Kobayashi, S.; Shibata, H.; Kurihara, I.; Yokota, K.; Suda, N.; Saito, I.; Saruta, T. Ubc9 interacts with chicken ovalbumin upstream promoter-transcription factor i and represses receptor-dependent transcription. J. Mol. Endocrinol. 2004, 32, 69–86. [Google Scholar] [CrossRef] [PubMed]
  147. Meyers, G.; Tautz, N.; Dubovi, E.J.; Thiel, H.J. Viral cytopathogenicity correlated with integration of ubiquitin-coding sequences. Virology 1991, 180, 602–616. [Google Scholar] [CrossRef]
  148. Rytkonen, A.; Holden, D.W. Bacterial interference of ubiquitination and deubiquitination. Cell Host Microbe 2007, 1, 13–22. [Google Scholar] [CrossRef] [PubMed]
  149. Rytkonen, A.; Poh, J.; Garmendia, J.; Boyle, C.; Thompson, A.; Liu, M.; Freemont, P.; Hinton, J.C.; Holden, D.W. Ssel, a salmonella deubiquitinase required for macrophage killing and virulence. Proc. Natl. Acad. Sci. USA 2007, 104, 3502–3507. [Google Scholar] [CrossRef] [PubMed]
  150. Zhou, H.; Monack, D.M.; Kayagaki, N.; Wertz, I.; Yin, J.; Wolf, B.; Dixit, V.M. Yersinia virulence factor yopj acts as a deubiquitinase to inhibit NF-kappa b activation. J. Exp. Med. 2005, 202, 1327–1332. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Wimmer, P.; Schreiner, S. Viral Mimicry to Usurp Ubiquitin and SUMO Host Pathways. Viruses 2015, 7, 4854-4872. https://doi.org/10.3390/v7092849

AMA Style

Wimmer P, Schreiner S. Viral Mimicry to Usurp Ubiquitin and SUMO Host Pathways. Viruses. 2015; 7(9):4854-4872. https://doi.org/10.3390/v7092849

Chicago/Turabian Style

Wimmer, Peter, and Sabrina Schreiner. 2015. "Viral Mimicry to Usurp Ubiquitin and SUMO Host Pathways" Viruses 7, no. 9: 4854-4872. https://doi.org/10.3390/v7092849

Article Metrics

Back to TopTop