Skip to main content

The structure and regulation of Cullin 2 based E3 ubiquitin ligases and their biological functions

Abstract

Background

Cullin-RING E3 ubiquitin ligase complexes play a central role in targeting cellular proteins for ubiquitination-dependent protein turnover through 26S proteasome. Cullin-2 is a member of the Cullin family, and it serves as a scaffold protein for Elongin B and C, Rbx1 and various substrate recognition receptors to form E3 ubiquitin ligases.

Main body of the abstract

First, the composition, structure and the regulation of Cullin-2 based E3 ubiquitin ligases were introduced. Then the targets, the biological functions of complexes that use VHL, Lrr-1, Fem1b, Prame, Zyg-11, BAF250, Rack1 as substrate targeting subunits were described, and their involvement in diseases was discussed. A small molecule inhibitor of Cullins as a potential anti-cancer drug was introduced. Furthermore, proteins with VHL box that might bind to Cullin-2 were described. Finally, how different viral proteins form E3 ubiquitin ligase complexes with Cullin-2 to counter host viral defense were explained.

Conclusions

Cullin-2 based E3 ubiquitin ligases, using many different substrate recognition receptors, recognize a number of substrates and regulate their protein stability. These complexes play critical roles in biological processes and diseases such as cancer, germline differentiation and viral defense. Through the better understanding of their biology, we can devise and develop new therapeutic strategies to treat cancers, inherited diseases and viral infections.

Background

Cullin-RING E3 ubiquitin ligase complexes (CRLs) play a central role in targeting cellular proteins for ubiquitination-dependent protein turnover through 26S proteasome [1]. Cullin-2 (Cul2), a member of Cullin family proteins, is encoded by CUL2. Cul2 functions as a scaffold protein to form CRLs that belong to the Elongin B and C-Cul2 or Cul5-SOCS box protein (ECS) family [2]. In CRL2 complexes, Cul2 assembles with RING protein (Rbx1) (also known as Roc1) as RING finger protein, Elongin B and C proteins as adapter proteins and various substrate recognition receptors [2, 3].

Cul2 is different from other most Cullins, which are evolutionarily conserved from yeast to human. Cul2 is present only in multi-cellular organisms and plays a particular function [4]. The most well-known CRL2 substrate recognition receptor is the tumor suppressor protein VHL that is mutated in von Hippel–Lindau (VHL) syndrome, a rare hereditary cancer syndrome [5]. Germline VHL mutations usually disrupt the interaction between VHL and Elongin B and C, and inactivate the VHL-Elongin B/C-Cullin-2 E3 ligase [6]. CRL2VHL complex-dependent degradation of the α subunits of hypoxia inducible factor (HIFα) is the most studied role of CRL2 ubiquitin ligase in tumorigenesis [7, 8]. In addition, CRL2 ligases are involved in other cellular processes including germline development and viral infection. This review will go over the structure and regulations of CRL2 ligases, their substrate recognition receptors and their numerous substrates, and discuss their involvement in biological processes and diseases.

Main text

Structure and regulation

Similar to other Cullins, Cul2 contains an evolutionary conserved Cullin homology (CH) domain at its C-terminus. The CH domain was found to interact with Rbx1, which further recruits E2 ubiquitin conjugating enzymes [9] (Fig. 1). The N-terminus of Cul2 was responsible for interacting with Elongin B and C and various substrate recognition receptors (Fig. 1). These receptors usually contained a special domain called VHL-box [10].

Fig. 1
figure 1

Structure of CRL2VHL complex and Nedd8-mediated regulation of CRL2 activity. Cul2 is the scaffold protein that binds to Rbx1, Elongin C and VHL directly. Neddylation on lysine 689 of cullin-2 dissociates Cand1, which allows the Cul2 to bind to Elongin B, C and VHL, facilitates appropriate conformation of Rbx1 and promotes ubiquitination on substrate proteins. Ube2m promotes neddylation of Cul2 and increases CRL2 activity, whereas CSN5 and inhibitor inhibit CRL2 activity. Ub ubiquitin, N Nedd8

Elongin B and C proteins were originally found as two regulatory subunits of the Elongin complex, which was a positive regulator of RNA polymerase II and increased the rate of mRNA elongation by suppressing transient pausing along the DNA template. Elongin B and C bound to each other and enhanced the transcriptional activity of the other component of Elongin complex, Elongin A [4–6]. Elongin B and C were later found to bind to Cul2 or Cullin-5 (Cul5) and serve as adapter components of ECS ubiquitin ligases [11–13].

VHL and other Cul2-Rbx1 interacting proteins such as Leucine-Rich Repeat protein-1 (LRR-1) and Feminization-1 (FEM-1) have a region of homology called the VHL box (Fig. 2). This box contained both a BC box [14] (consensus sequence: (S,T,P)LXXX(C,S,A)XXXϕ, with ϕ meaning a hydrophobic amino acid), which bound to Elongin B and C, and a Cullin 2 box (consensus sequence: ϕPXXϕXXXϕ), which was responsible for binding to Cul2. Detailed alignment that defined VHL box could be found in Mahrour et al. [10]. VHL box was very similar to Suppressor Of Cytokine Signaling (SOCS) box, which also contained a BC box and a Cullin 5 box (Fig. 2). Although both VHL box proteins and SOCS box proteins used Elongin B and C as an adaptor, they bound to different Cullins. The different Cullin boxes determined the binding specificity to Cul2 and Cul5 [14–16].

Fig. 2
figure 2

The compositions of VHL box and SOCS box. VHL box is made up of a BC box and a Cullin2 box. SOCS box consists of a BC box and a Cullin5 box

A recent paper described the crystal structure of a CRL2 complex composed of VHL, Elongin B and C and the N-terminus of Cul2 [17]. It showed that in many ways CRL2 structure were different from these of CRL1 or CRL5 complex. The CRL2 complex assumed a tripod shape, with Elongin C located in the center and the other components at the ends. Cul2 bound to the interface between VHL and Elongin C through hydrophobic and electrostatic interactions. Cul2 binding induced a structuring of Elongin C loop (residue 48–57) which made contact with Cul2. The same loop was not structured in the VHL-Elongin BC complex [18]. Different from Cul5, the N-terminal extension of Cul2 played a critical role in binding to Elongin C. For example, residue L3 of Cul2 inserted into a hydrophobic pocket of Elongin C. L3G mutant of Cul2 drastically reduced the interaction between Cul2 and VHL-Elongin BC complex. Consistently, the N-terminal extension was highly conserved across all Cullin-2 orthologs. Importantly, not only the Cullin 2 box was critical for Cul2-VHL interaction [16], the BC box on VHL also made critical contact with Cul2 via hydrogen bonds and salt bridge interactions [17].

Similar to other Cullin family members, Cul2 contained a neddylation site close to the RING protein (Rbx1) binding site [19]. Auto-neddylation of Cullin by Rbx1 induced conformational change at C-terminus, resulting in stabilization of an optimal Rbx1 position and activation of CRL ubiquitin transfer activity [20–22]. NEDD8-conjugating enzyme Ube2m (also known as Ubc12) promoted neddylation of Cullin 1-4 through Rbx1, whereas Ube2f neddylated Cullin 5 through Rbx2 [23]. Conversely, deneddylation by COP9 signalosome complex subunit 5 (Csn5) or a small molecule inhibitor of NEDD8-activating enzyme (MLN4924) [24] led to binding of Cullin-Associated and Neddylation-Dissociated 1 (Cand1) to Cullins. This binding sterically inhibited the interaction between Cullin and adaptor proteins, and impaired Rbx1-mediated E2 ubiquitin activation [24–30] (Fig. 1). Interestingly, engagement of substrates to CRL complex could induce Cullin neddylation [31, 32]. This ‘substrate-mediated neddylation’ was recently reported to be mediated by Defective in Cullin Neddylation 1 (Dcnl1) [33]. Dcnl1 was the human homolog of Dcn1 in Saccharomyces cerevisiae, also known as Sccro or Dcun1d1, which was an E3 Nedd8 ligase that promoted Cullin neddylation with Rbx1 [34–37]. Interaction between VHL and its substrate HIF1α promoted the recruitment of Dcnl1 to trigger Cul2 neddylation, and consequently HIF1α ubiquitination and proteasomal degradation [33].

Different CRL2 E3 ubiquitin ligase complexes

There are a number of CRL2 complexes that are confirmed as functional E3 ubiquitin ligases. They can be divided into two groups: cellular CRL2 complexes that are derived from cellular proteins, and viral CRL2 complexes that contain viral proteins. Known CRL2 E3 complex and their substrates are summarized in Table 1.

Table 1 Known CRL2 complexes and their substrates

CRL2VHL complex

Von Hippel–Lindau (VHL) syndrome was first described separately by von Hippel in 1911 and by Lindau in 1926 [5]. It was characterized by the development of multiple vascular tumors and was caused by a mutation of both alleles of the VHL gene located on the short arm of chromosome 3 [38]. VHL was a 213 amino acid protein product of the VHL tumor suppressor gene. Most germline VHL mutations were missense alterations that produced mutated VHL proteins that lost ability to bind to Elongin B and C [39, 40]. Further study showed that VHL formed a complex with Cul2, Elongin B and C and Rbx1, and had E3 ubiquitin ligase activity [6, 9, 41]. The CRL2 ligase complex could bind to HIFα through the β domain of VHL, promote ubiquitination and proteasomal degradation of HIFα [42, 43]. HIFα family consisted of three members, HIF1α, HIF2α and HIF3α. They were unstable subunit of HIF complex, and formed the HIF transcriptional factor with constitutively expressed HIF1β, also called Aryl Hydrocarbon Receptor Nuclear Translocator (ARNT), to regulate gene expressions [7]. HIF downstream target genes [44] include vascular endothelial growth factor A (VEGFA) [45, 46], solute carrier family 2 member 1 (SLC2A1, which was also called GLUT1), and platelet-derived growth factor-β (PDGFB) [47], which were known to drive cell growth and proliferation of microvascular vessels in VHL syndrome [42].

The HIF transcriptional activity is tightly regulated by oxygen concentration. Under normal oxygen tension (normoxia), two key proline residues in the oxygen-dependent degradation domain of HIFα were hydroxylated by HIF prolyl hydroxylases (PHD1–3). Hydroxylated HIFα provided a binding signal for the β-domain of VHL [43, 48–53]. Consequently HIFα was poly-ubiquitinated by CRL2VHL E3 ubiquitin ligase and degraded by the proteasome. When oxygen was taken away (hypoxia), HIFα was produced but not hydroxylated by PHDs, so it escaped recognition by VHL. As a consequence it would accumulate, form HIF, and activate the transcriptional program to respond to hypoxia [8]. Any other conditions that disrupt the functions of PHDs or VHL will also lead to HIF stabilization and the activation of HIF pathway. In hereditary VHL disease, mutations and loss of heterozygosity (LOH) at the VHL locus in the cancer cells inactivates VHL and results in a constitutively high level of HIFα even in the presence of oxygen. The activated HIF targets can have opposing effects on tumor growth [54, 55], but the overall activity of the constitutively active HIF pathway is the major oncogenic force that drives tumorigenesis and tumor growth. It was known to cause many manifestation of VHL disease such as clear cell renal cell carcinoma, hemangioblastoma and pheochromocytoma [7], and the partial blockage of HIF pathway by anti-angiogenesis drugs produced significant clinical benefits [56]. Currently, five drugs targeting VEGFA (bevacizumab) or its receptors (sunitinib, sorafenib, pazopanib and axitinib) were approved by FDA for the treatment of renal cell carcinoma (RCC). The median survival of advanced RCC patients had increased from less than 1 year (receiving cytokine IFN-alpha) to nearly 2 year (receiving targeted therapies) during the last decade [57].

Although HIFα is the most studied VHL substrate and probably the most important one, CRL2VHL also recognizes and mediates the degradation of many other substrates. Similar to HIFα, Sprouty2 (Spry2), a protein that regulates cell migration and proliferation in response to a number of growth factors, was also hydroxylated by PHD at normoxia and recognized by VHL for degradation. Increased cellular level of Spry2 after silencing PHDs or VHL inhibited human fibroblast growth factor-elicited activation of ERK1/2 [58]. Rpb1 is the largest subunit of RNA polymerase II. It is also the enzymatic subunit of the complex, synthesizing cellular mRNAs [59]. Rpb1 contains an LGQLAP motif that bears sequence and structural similarity to a VHL-binding sequence in HIF1α [60]. Similar to HIFα, the proline P1465 within the motif of Rpb1 was hydroxylated under oxidative stress [60]. Proline 1465 hydroxylation by PHD1 and the further recognition by VHL was required for oxidative stress-induced Ser5 phosphorylation of Rpb1, the poly-ubiquitination of Rpb1 and the recruitment of Rpb1 to the DNA, which stimulated formation of tumors by VHL+ cells [59]. In a different cell line, Rpb1 was poly-ubiquitinated by VHL and degraded by proteasome [60]. Since the direct function of CRL2VHL was ubiquitination, the next fate of Rpb1 was probably determined by different cellular context. hsRPB7, another subunit of RNA polymerase II, was also reported to be poly-ubiquitinated by VHL and degraded. Consequently, VHL suppressed hsRPB7-dependent VEGF expression [61]. Atypical protein kinase C (PKC) is made up of two members, PKCλ/ι (PKCι is the human homologue of mouse PKCλ) and PKCζ. In other reports, both PKCλ and PKCζII (a rapidly degraded variant of PKCζ) were poly-ubiquitinated by VHL and degraded [62, 63]. Epidermal growth factor receptor (EGFR) was also reported to be a target of CRL2VHL. VHL limited EGFR signaling by promoting c-Cbl-independent poly-ubiquitination and lysosome-independent degradation of the activated EGFR [64]. In addition, some E3-ligase independent functions of VHL were reported [65–68]. In these cases, VHL interacted with other proteins, regulated their functions, but did not promote their poly-ubiquitination and degradation. For example, VHL bound to NF-kappa B agonist Card9, promoted its phosphorylation by CK2 and inhibited NF-kappa B activity [64]. In particular, Lai et al. performed a series of proteomic analyses that identified many VHL-interacting proteins [66]. It’s a valuable resource for further investigation.

CRL2LRR-1 complex

Leucine-Rich Repeat protein-1 (LRR-1) in worm was found to have a VHL box and functioned as a substrate recognition receptor in a CRL2 complex [16, 69]. In C. elegans, the CRL2LRR-1 complex degraded the Cip/Kip CDK-inhibitor CKI-1 in nucleus to ensure a proper G1-phase cell cycle progression in the germ cells [69–71]. In human cells, the orthologous human CRL2LRR1 complex degraded the CDK-inhibitor p21Cip1, but did not regulate cell cycle because it only did so in the cytoplasm. Consequently, knockdown of Lrr1 resulted in increased cytoplasmic p21. This led to de-phosphorylation of cofilin through the inhibition of Rho/ROCK/LIMK pathway. The de-phosphorylated cofilin activated actin cytoskeleton remodeling and promoted cell motility [69].

CUL2 was highly expressed in the germline and in early embryos in C. elegans [70]. In Drosophila melanogaster, loss of function of CUL2 resulted in defects at the larval neuromuscular junction and aberrations in the development of female germ line [72]. Cul2 was also required to limit the number of motile cells in egg chambers [73] and for germline enclosure in testes [74]. Among several CRL2 complexes that were associated with germline development, CRL2LRR-1 complex was a critical one. LRR-1 null nematodes were defective in germ cell proliferation which resulted in animal sterility. Similar phenotype was observed in CUL2 null animals [71]. Since LRR-1 null germ cells arrested at G2/M stage, it was tested whether suppression of the DNA replication checkpoint would rescue the phenotype. It was discovered that the suppression of CHK-1 (Chk1 in humans for checkpoint kinase 1) or ATL-1 (ATR, Ataxia telangiectasia and Rad3 related) kinases, two core components of the DNA replication checkpoint pathway [75, 76], restored the fertility [71]. However, how LRR-1 or CUL2 deficiency caused hyper-activation of the DNA replication checkpoint pathway was still unknown. As CKI-1 suppression did not rescue the fertility phenotype of LRR-1 null animal, it was not the critical target for this phenotype [71].

In later steps of germ cell development, CRL2LRR-1 regulated the balance between mitotic proliferation and meiotic entry. It was probably because CRL2LRR-1 could regulate the degradation of unidentified meiotic promoting factors in the germline [77]. In nematode, CRL2LRR-1 inhibited the first steps of meiotic prophase through regulating the stability of HORMA-domain protein HTP-3, a key protein for loading synaptonemal complex components onto meiotic chromosomes [77]. Hence, CUL2 played multiple roles in the development of the germline in nematodes. Since Cul2 is conserved in multi-cellular organisms [4], the regulation mechanisms of germline development may be similar in other organisms as well [77, 78].

CRL2FEM1B complex

Feminization-1 (FEM-1) was discovered to regulate apoptosis in the nematode sex determination pathway [79]. FEM-1 and its three homologs, Fem1a, Fem1b and Fem1c were found to contain a VHL-box, so theoretically they could interact with Cul2 [80]. Whereas FEM-1 and Fem1b were shown to target proteins for degradation, Fem1a and Fem1c were not confirmed as a component of CRL2 complex. In nematode, FEM-1 was found to target TRA-1 for ubiquitination [81]. TRA-1 was homologous to the mammalian Gli1 protein, an important transcription factor in Hedgehog signaling. Consistent with the worm data, Fem1b promoted ubiquitination and suppressed transcriptional activity of Gli1 in human [82]. Since Gli1 was an oncoprotein, Fem1b could be a tumor suppressor. Single nucleotide polymorphism (SNP) analysis revealed that Fem1b was associated with polycystic ovary syndrome [83]. In Fem1b-null mice insulin resistance was observed [84]. It was also reported to mediate apoptosis in human colon cancer cells [85] and served as a biomarker in mouse colon cancer model [86]. Fem1a was also implicated in polycystic ovary syndrome [83] and sonic hedgehog pathway hyperactivation in cancer stem cells in gastric cancer [87].

Mouse Fem1b induced ubiquitin-mediated degradation of Ankrd37, a protein that was enriched in mouse testis [88]. In addition, mouse Fem1b interacted with the homeodomain protein Nkx3.1, which was a pivotal regulator of prostate development. Both Fem1b and Nkx3.1 null mice show similar defects in prostate ductal morphogenesis [89]. These data indicate that Fem1b plays a conserved role in the generation of sexual dimorphism.

CRL2PRAME complex

The human tumor antigen Preferentially Expressed Antigen in Melanoma (Prame) was frequently overexpressed in various cancers, and the high level expression was usually correlated with advanced stages and poor clinical outcomes in a wide variety of cancers [90]. The consensus LXXLL-binding domain at Prame‘s C terminus mediated interaction with the retinoic acid receptor (RAR), and Prame acted as a dominant repressor of RAR signaling and inhibited retinoic acid induced differentiation, growth arrest, and apoptosis [91]. At the N-terminus of Prame there was a VHL box, and it mediated the interaction with Elongin C and Cul2. Genome-wide chromatin immunoprecipitation experiments revealed that Prame associated with the transcription factor NFY at enhancers and transcriptionally active promoters. In addition, CRL2 complex were present together with Prame on chromatins [92]. Further analysis revealed that Prame interacted with OSGEP and LAGE3, two yeast proteins that were human orthologues of the ancient EKC/KEOPS complex. EKC/KEOPS complex was shown to play a role in telomeres maintenance, transcriptional regulation, and t6A modification of tRNAs [93, 94]. Furthermore, Prame recruited a CRL2 ubiquitin ligase to EKC complex on transcriptionally active chromatin [95]. The substrates of the E3 ligase activity of CRL2PRAME complex are still unknown.

CRL2ZYG-11 complex

ZYG-11 was identified as a gene that contributed to nematode zygote development in C. elegans [96]. Further analysis revealed a VHL box at the N-terminus of ZYG-11, and it was shown to bind to Elongin C and form complex with Cul2 [97]. Although the substrate(s) of CRL2ZYG-11complex was not identified, genetic analysis revealed that the complex was required for many functions of CUL2 in worm, such as the degradation of maternal cyclin B [97]. ZYG-11 homologues are restricted to metazoan. C. elegans has two ZYG11 family members, ZYG-11 and ZER-1. Both contain a VHL box and bind to Elongin C and CUL2. In human there are three ZYG11 family members, Zyg11a, Zyg11b and Zyg11bl. Only Zyg11b and Zyg11bl contained a VHL box and bound to Elongin C and Cul2 [97]. In human, Zyg11bl was found to be highly expressed in skeletal muscle and the testes [98], and it was specifically expressed in the cytoplasm of late pachytene spermatocytes and the round spermatids at meiotic division [98]. Although the substrate(s) was unknown, ZYG11 family members were proposed to function as substrate recognition receptors for CRL2 E3 complexes in the metazoan lineage [97].

CRL2BAF250 complex

Two isoforms of BAF250, BAF250a/ARID1A and BAF250b/ARID1B are defining components of human BAF complex. BAF complex and PBAF complex belong to SWI/SNF chromatin-remodeling complex, which remodels chromatin and facilitates DNA access by transcription factors and the transcription machinery [99]. Both BAF250a and BAF250b contained a BC box. BAF250b was shown to associate with Elongin B and C, Cul2 and Rbx1 to form an E3 ligase, which mono-ubiquitinated histone H2B on lysine 120 site [100]. Konckdown of BAF250a or BAF250b decreased levels of global H2B ubiqitination in human cell line. In addition, the BAF250 Drosophila homolog Osa mutant had reduced levels of mono-ubiquitinated H2B, and functioned synergistically with Cul2 in vivo [100]. These data suggest that BAF250 has an evolutionarily conserved function to regulate H2B ubiqutination as a component of CRL2 E3 ligase to promote transcription.

CRL2RACK1 complex

The Receptor for Activated C Kinase 1 (Rack1), a member of the tryptophan-aspartate repeat (WD-repeat) family proteins, was found to bind the N-terminus of Fem1b and poly-ubiquitinated Fem1b for proteasomal degradation in colon cancer cells [101]. Rack1 was also reported to act as an E3 ligase component to degrade ΔNp63α, a member of the p53 family [102]. Through the WD40 repeats, which contained an amino acid sequence similar to the VHL BC box, Rack1 was reported to bind to Elongin B and C and promoted the degradation of HIF1α in a HSP90-dependent but oxygen-independent manner [103]. In the presence of apoptotic agents, Rack1 mediated the degradation of Bcl-2-interacting mediator of cell death extra long (BimEL) through a CRL2 E3 ligase complex, and inhibited apoptosis in breast cancer cells [104]. The evidence suggests that Rack1 can be a component of CRL2 E3 complex and degrade target proteins via ubiquitin–proteasome pathway.

A CRL2 complex targeting RhoB

Since neddylation on Cullins was required for the activity of CRL complexes [20–22], a small molecule inhibitor of NEDD8-activating enzyme, MLN4924, could induce the accumulation of CRL substrates that lead to DNA damage, cell cycle defects, senescence, apoptosis and autophagy [24, 105–107]. It was tested by several phase I clinical trials because of its significant anticancer activity and relatively low toxicity in preclinical analyses [108–111]. A quantitative proteomic analysis identified RhoB as a target of CUL2-RBX1 complex [112]. The substrate-recognition subunit was not identified in this study. RhoB is a small GTPase and a member of Rho family. It acts as a tumor suppressor and is frequently down-regulated in various cancers. The MLN4924-induced accumulation of RhoB seemed to contribute significantly to the anticancer activity of this drug in liver cancer. A caveat is that MLN4924 impacts on many targets, so it is difficult to pinpoint the contribution of CRL2 to cancer development and treatment. Nonetheless, this highlights the potential therapeutic utility of targeting neddylation-CRL2-RhoB in liver cancer and other cancers.

VHL box proteins

Several proteins have VHL box but were not confirmed as components of CRL2 complex. Their functions are summarized here, and their roles in CRL2 complex await further investigation.

Appbp2, the human homolog of Drosophila PAT1, also known as Ara67, was found to suppress androgen receptor (AR) transactivation through interrupting AR cytoplasmic-nuclear shuttling [113]. Appbp2 was found to be overexpressed through 17q23 amplification in neuroblastoma [114], ovarian clear cell adenocarcinomas [115] and desmoplastic medulloblastomas [116].

Kelch domain containing 2 (Klhdc2), also known as Hclp1, could serve a transcriptional co-repressor through its inhibitory interaction with the Lzip transcription factor [117]. Klhdc3, also known as Peas, is evolutionarily conserved from nematodes to mammals. Mouse Peas was found to be expressed in testis, particularly in the cytoplasm and meiotic chromatin of pachytene spermatocytes. It was suggested that Klhdc3 might be involved in the meiotic recombination process [118].

Zinc finger, SWIM-type containing 2 (Zswim2) was also known as MEKK1-related protein X (Mex), a testis-expressed protein. It contained an N-terminal SWIM (SWI2/SNF2 and MuDR) domain and two RING fingers separated by a ZZ zinc finger domain. Zswim2 was self-ubiquitinated as an E3 ubiquitin ligase and targeted for degradation through the proteasome pathway [119]. The SWIM domain was found to be critical for Zswim2 ubiquitination and was suggested to regulate death receptor-induced apoptosis in the testes. Zswim5 (also known as KIAA1511), Zswim6 and Zswim8 (also known as KIAA0913) all contained a VHL box and might play a similar role in E3 ligase complex, but this was not confirmed. Zswim5 displayed intense staining in gliomas but weak to modest staining in most other neoplasms [120]. Fyn-tyrosine-kinase-deficient mice had increased fearfulness and enhanced excitability. In the amygdala of Fyn-deficient mice, only Zswim6 expression was significantly lowered after administration of N-methyl-D-aspartate (NMDA) when compared with that in Fyn-proficient mice, suggesting that it might be a key mediator of the phenotype [121]. Zswim6 mutations were associated with acromelic frontonasal dysostosis, a rare disorder characterized by the craniofacial, brain and limb malformations. Zswim6 mutations might lead to the phenotypes through the disruption of Hedgehog signaling [122].

Viral CRL2 E3 ligase complex

Viral infection activates host cell defense mechanisms, which will limit viral spread, inhibit viral replication and eliminate virus. Virus has developed various strategies to counter host cell defense and usurp the cellular machinery. One strategy is that viral protein formed E3 ubiquitin ligase complex to destroy host proteins. Several viral proteins that form CRL2 ligase complex had been reported to be indispensable for infection by adenovirus (Ad), Epstein–Barr virus (EBV), human papillomavirus (HPV) and bovine immunodeficiency virus (BIV).

Adenoviruses are linear double-stranded DNA viruses. They infect human and rodent cells, occasionally transform them and cause tumors in animal models [123]. The human adenovirus type 5 (Ad5) early region 4 from open reading frame 6 (E4orf6) contained three BC boxes and formed an E3 ubiqutin ligase complex with Cullin 5 (Cul5) [124, 125], whereas the human adenovirus type 12 (Ad12), type 16 (Ad16), type 40 (Ad40) and type41 (Ad41) formed complex with Cul2 [126, 127]. Adenoviral protein E1B55K associated with the E4orf6 protein and recognized substrate to be degraded by ubiquitin–proteasome pathway [124, 125]. In this complex, E4orf6 was believed to recruit Cul2 or Cul5 as an adaptor protein, whereas E1B55K was believed to act as a substrate recognition receptor. As a result, the E1B55K-E4orf6-Cul2 complex from different types of human adenovirus showed different substrate specificity against p53 [128, 129], Mre11 [130], DNA ligase IV [131] and integrin α3 [126, 127, 132–135]. Among these substrates, DNA Ligase IV was the only universal substrate for all types of adenoviruses tested [126, 133]. In particular, Ad12 E4orf6 not only recruited the Cul2 ubiquitin ligase complex but also acted as a substrate receptor for the ATR activator protein topoisomerase-IIβ–binding protein 1 (TOPBP1). Ad12 E4orf6 could inhibit the ATR-dependent phosphorylation of CHK1 through promoting the proteasomal degradation of TOPBP1 in the absence of E1B55K [133, 136].

Epstein–Barr virus (EBV) is a human γ-herpesvirus, and it is able to induce several B cell and epithelial-cell malignancies. In viral life cycles, EBV periodically reactivates and replicates in a lytic manner [137]. Induction of the EBV lytic program was found to trigger a cellular DNA damage response via activating the ATM-dependent DNA damage signal transduction pathway [138]. This would induce apoptosis and limit viral replication by Chk2-mediated phosphorylation of p53 at its C-terminus [139, 140]. The EBV virus developed a method to circumvent this limitation. Bzlf1 protein of EBV had Cul2 and Cul5 boxes at its N-terminus and could form complexes with Cul2 and Cul5, and Bzlf1 recognized C-terminal phosphorylated p53 and induced p53 degradation to ensure efficient viral propagation [140, 141].

Human papillomaviruses (HPVs) are DNA viruses that specifically infect squamous epithelial cells Bernard HU2010. Among more than 120 different species identified so far, HPV16 was found in 50 % of cervical cancers [142]. E7 oncoprotein of HPV16 was necessary for the induction and maintenance of the oncogenic transformation [143]. HPV16 E7 was found to form a complex with Cul2 via an incomplete Cul2 box, and it bound and promoted the degradation of a hypophosphorylated form of the retinoblastoma tumor suppressor (RB1) [144–146]. This allowed RB1–E2F complexes to dissociate and the G1-S phase transition to proceed, allowing the replication of the viral DNA in differentiated host cells [147, 148]. In addition, Zyg-11 related cell cycle regulator (Zer1, also known as Zyg11bl) was required for the binding of HPV16 E7 to Cul2 and the destabilization of RB1 in HPV16 E7-expressing cells [149].

The viral infectivity factor (Vif) from human immunodeficiency virus type 1 (HIV-1) and simian immunodeficiency virus (SIV) could form a CRL5 E3 ubiquitin ligase complex to degrade host antiviral APOBEC3 (A3) proteins, so the HIV-1 could escape from A3-mediated host antiviral defense [150]. Similarly, Vif from bovine immunodeficiency virus (BIV) interacted with Cul2, Elongin B/C and Rbx1, instead of Cul5 and Rbx2 in HIV, to form a CRL2 E3 ubiquitin ligase. This complex was reported to degrade the bovine A3 proteins (A3Z2Z3 and A3Z3) [104, 151]. Consistently, BIV Vif with mutations in the BC box or putative VHL box, which failed interact with Elongin B/C or Cul2, respectively, lost the ability to regulate bovine A3 proteins [104].

Conclusions

Among CRLs, Cul2 based E3 ligase complexes had a similar structure and binding partners with Cul5 based E3 ligase complexes, and both belonged to ECS family [3]. The substrate recognition receptor of Cul2 complex generally contained a VHL box, which contained a BC box and a Cullin box, and was very similar to SOCS box in Cul5 complex. Recent crystal structure analysis revealed the differences between CRL2 and CRL5 complexes, and indicated the possibility of fine-tuning CRL2 activity [17]. The activity of CRL2 can be regulated by neddylation on a key residue on Cul2 [24, 26]. Through various substrate receptors, CRL2 complexes recognize a number of substrates and regulate their protein stability and function through polyubiquitination (Table 1). Defects in various CRL2 complexes led to cancer and other human disease through abnormal stabilization and enhanced activity of their protein substrates. Inhibiting the activities of the substrates or those of their downstream effectors have shown clinical efficacy. As different viral proteins co-opt Cul2 to evade host defense, inhibiting their activities might help us fight various viral infections. Thus through better understanding of the biology of CRL2 complexes, we can devise and develop new therapeutic strategies against cancers, inherited diseases and viral infections caused by dysregulated CRL2 complexes.

Abbreviations

CRLs:

Cullin-RING E3 ubiquitin ligase complexes

Cul2:

Cullin-2

Rbx1:

RING protein

ECS:

Elongin B and C-Cul2 or Cul5-SOCS box protein

VHL:

von Hippel–Lindau

HIFα:

α subunits of hypoxia inducible factor

CH:

cullin homology

Cul5:

Cullin-5

LRR-1:

Leucine-Rich Repeat Protein -1

FEM-1:

Feminization-1

SOCS:

Suppressor of Cytokine Signaling

RCC:

renal cell carcinoma

Csn5:

COP9 signalosome complex subunit 5

Cand1:

Cullin-Associated and Neddylation-Dissociated 1

Dcnl1:

Defective in Cullin Neddylation 1

ARNT:

Aryl hydrocarbon Receptor Nuclear Translocator

VEGFA:

vascular endothelial growth factor A

SLC2A1:

solute carrier family 2 member 1, also called GLUT1

PDGFB:

platelet-derived growth factor-β

PHD:

prolyl hydroxylase

LOH:

loss of heterozygosity

Spry2:

Sprouty2

PKC:

protein kinase C

EGFR:

epidermal growth factor receptor

HTP-3:

HORMA-domain protein

SNP:

single nucleotide polymorphism

Prame:

preferentially expressed antigen in melanoma

RAR:

retinoic acid receptor

Rack1:

Receptor for Activated C Kinase 1

WD-repeat:

tryptophan-aspartate repeat

BimEL:

Bcl-2-interacting mediator of cell death extra long

AR:

androgen receptor

Klhdc2:

Kelch domain containing 2

Zswim2:

Zinc finger, SWIM-type containing 2

Mex:

MEKK1-related protein X

SWIM:

SWI2/SNF2 and MuD

NMDA:

N-methyl-D-aspartate

Ad:

adenovirus

EBV:

Epstein–Barr virus

HPV:

human papillomavirus

BIV:

bovine immunodeficiency virus

E4orf6:

early region 4 from open reading frame 6

TOPBP1:

topoisomerase-IIβ–binding protein 1

RB1:

retinoblastoma tumor suppressor

Zer1:

Zyg-11 related cell cycle regulator

Vif:

viral infectivity factor

HIV-1:

human immunodeficiency virus type 1

SIV:

simian immunodeficiency virus

References

  1. Berndsen CE, Wolberger C. New insights into ubiquitin E3 ligase mechanism. Nat Struct Mol Biol. 2014;21:301–7.

    Article  CAS  PubMed  Google Scholar 

  2. Sarikas A, Hartmann T, Pan ZQ. The cullin protein family. Genome Biol. 2011;12:220.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Zimmerman ES, Schulman BA, Zheng N. Structural assembly of cullin-RING ubiquitin ligase complexes. Curr Opin Struct Biol. 2010;20:714–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Petroski MD, Deshaies RJ. Function and regulation of cullin-RING ubiquitin ligases. Nat Rev Mol Cell Biol. 2005;6:9–20.

    Article  CAS  PubMed  Google Scholar 

  5. Linehan WM, Lerman MI, Zbar B. Identification of the von Hippel–Lindau (VHL) gene. Its role in renal cancer. JAMA. 1995;273:564–70.

    Article  CAS  PubMed  Google Scholar 

  6. Pause A, Lee S, Worrell RA, Chen DY, Burgess WH, Linehan WM, Klausner RD. The von Hippel–Lindau tumor-suppressor gene product forms a stable complex with human CUL-2, a member of the Cdc53 family of proteins. Proc Natl Acad Sci USA. 1997;94:2156–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Kaelin WG Jr. Molecular basis of the VHL hereditary cancer syndrome. Nat Rev Cancer. 2002;2:673–82.

    Article  CAS  PubMed  Google Scholar 

  8. Zhang Q, Yang H. The roles of VHL-dependent ubiquitination in signaling and cancer. Front Oncol. 2012;2:35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kamura T, Koepp DM, Conrad MN, Skowyra D, Moreland RJ, Iliopoulos O, Lane WS, Kaelin WG Jr, Elledge SJ, Conaway RC, et al. Rbx1, a component of the VHL tumor suppressor complex and SCF ubiquitin ligase. Science. 1999;284:657–61.

    Article  CAS  PubMed  Google Scholar 

  10. Mahrour N, Redwine WB, Florens L, Swanson SK, Martin-Brown S, Bradford WD, Staehling-Hampton K, Washburn MP, Conaway RC, Conaway JW. Characterization of Cullin-box sequences that direct recruitment of Cul2-Rbx1 and Cul5-Rbx2 modules to Elongin BC-based ubiquitin ligases. J Biol Chem. 2008;283:8005–13.

    Article  CAS  PubMed  Google Scholar 

  11. Garrett KP, Aso T, Bradsher JN, Foundling SI, Lane WS, Conaway RC, Conaway JW. Positive regulation of general transcription factor SIII by a tailed ubiquitin homolog. Proc Natl Acad Sci USA. 1995;92:7172–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Garrett KP, Tan S, Bradsher JN, Lane WS, Conaway JW, Conaway RC. Molecular cloning of an essential subunit of RNA polymerase II elongation factor SIII. Proc Natl Acad Sci USA. 1994;91:5237–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Aso T, Lane WS, Conaway JW, Conaway RC. Elongin (SIII): a multisubunit regulator of elongation by RNA polymerase II. Science. 1995;269:1439–43.

    Article  CAS  PubMed  Google Scholar 

  14. Kamura T, Sato S, Haque D, Liu L, Kaelin WG Jr, Conaway RC, Conaway JW. The Elongin BC complex interacts with the conserved SOCS-box motif present in members of the SOCS, ras, WD-40 repeat, and ankyrin repeat families. Genes Dev. 1998;12:3872–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Hilton DJ, Richardson RT, Alexander WS, Viney EM, Willson TA, Sprigg NS, Starr R, Nicholson SE, Metcalf D, Nicola NA. Twenty proteins containing a C-terminal SOCS box form five structural classes. Proc Natl Acad Sci USA. 1998;95:114–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Kamura T, Maenaka K, Kotoshiba S, Matsumoto M, Kohda D, Conaway RC, Conaway JW, Nakayama KI. VHL-box and SOCS-box domains determine binding specificity for Cul2-Rbx1 and Cul5-Rbx2 modules of ubiquitin ligases. Genes Dev. 2004;18:3055–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Nguyen HC, Yang H, Fribourgh JL, Wolfe LS, Xiong Y. Insights into Cullin-RING E3 ubiquitin ligase recruitment: structure of the VHL-EloBC-Cul2 complex. Structure. 2015;23:441–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Stebbins CE, Kaelin WG Jr, Pavletich NP. Structure of the VHL-ElonginC-ElonginB complex: implications for VHL tumor suppressor function. Science. 1999;284:455–61.

    Article  CAS  PubMed  Google Scholar 

  19. Enchev RI, Schulman BA, Peter M. Protein neddylation: beyond cullin-RING ligases. Nat Rev Mol Cell Biol. 2015;16:30–44.

    Article  CAS  PubMed  Google Scholar 

  20. Cope GA, Suh GS, Aravind L, Schwarz SE, Zipursky SL, Koonin EV, Deshaies RJ. Role of predicted metalloprotease motif of Jab1/Csn5 in cleavage of Nedd8 from Cul1. Science. 2002;298:608–11.

    Article  CAS  PubMed  Google Scholar 

  21. Boh BK, Smith PG, Hagen T. Neddylation-induced conformational control regulates cullin RING ligase activity in vivo. J Mol Biol. 2011;409:136–45.

    Article  CAS  PubMed  Google Scholar 

  22. Saha A, Deshaies RJ. Multimodal activation of the ubiquitin ligase SCF by Nedd8 conjugation. Mol Cell. 2008;32:21–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Huang DT, Ayrault O, Hunt HW, Taherbhoy AM, Duda DM, Scott DC, Borg LA, Neale G, Murray PJ, Roussel MF, et al. E2-RING expansion of the NEDD8 cascade confers specificity to cullin modification. Mol Cell. 2009;33:483–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Soucy TA, Smith PG, Milhollen MA, Berger AJ, Gavin JM, Adhikari S, Brownell JE, Burke KE, Cardin DP, Critchley S, et al. An inhibitor of NEDD8-activating enzyme as a new approach to treat cancer. Nature. 2009;458:732–6.

    Article  CAS  PubMed  Google Scholar 

  25. Liu J, Furukawa M, Matsumoto T, Xiong Y. NEDD8 modification of CUL1 dissociates p120(CAND1), an inhibitor of CUL1-SKP1 binding and SCF ligases. Mol Cell. 2002;10:1511–8.

    Article  CAS  PubMed  Google Scholar 

  26. Zhao Y, Morgan MA, Sun Y. Targeting neddylation pathways to inactivate cullin-RING ligases for anticancer therapy. Antioxid Redox Signal. 2014;21:2383–400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Lydeard JR, Schulman BA, Harper JW. Building and remodelling Cullin-RING E3 ubiquitin ligases. EMBO Rep. 2013;14:1050–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Hotton SK, Callis J. Regulation of cullin RING ligases. Annu Rev Plant Biol. 2008;59:467–89.

    Article  CAS  PubMed  Google Scholar 

  29. Echalier A, Pan Y, Birol M, Tavernier N, Pintard L, Hoh F, Ebel C, Galophe N, Claret FX, Dumas C. Insights into the regulation of the human COP9 signalosome catalytic subunit, CSN5/Jab1. Proc Natl Acad Sci USA. 2013;110:1273–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Bennett EJ, Rush J, Gygi SP, Harper JW. Dynamics of cullin-RING ubiquitin ligase network revealed by systematic quantitative proteomics. Cell. 2010;143:951–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sufan RI, Ohh M. Role of the NEDD8 modification of Cul2 in the sequential activation of ECV complex. Neoplasia. 2006;8:956–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Chew EH, Hagen T. Substrate-mediated regulation of cullin neddylation. J Biol Chem. 2007;282:17032–40.

    Article  CAS  PubMed  Google Scholar 

  33. Heir P, Sufan RI, Greer SN, Poon BP, Lee JE, Ohh M. DCNL1 functions as a substrate sensor and activator of cullin 2-RING ligase. Mol Cell Biol. 2013;33:1621–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kurz T, Chou YC, Willems AR, Meyer-Schaller N, Hecht ML, Tyers M, Peter M, Sicheri F. Dcn1 functions as a scaffold-type E3 ligase for cullin neddylation. Mol Cell. 2008;29:23–35.

    Article  CAS  PubMed  Google Scholar 

  35. Kurz T, Ozlu N, Rudolf F, O’Rourke SM, Luke B, Hofmann K, Hyman AA, Bowerman B, Peter M. The conserved protein DCN-1/Dcn1p is required for cullin neddylation in C. elegans and S. cerevisiae. Nature. 2005;435:1257–61.

    Article  CAS  PubMed  Google Scholar 

  36. Kim AY, Bommelje CC, Lee BE, Yonekawa Y, Choi L, Morris LG, Huang G, Kaufman A, Ryan RJ, Hao B, et al. SCCRO (DCUN1D1) is an essential component of the E3 complex for neddylation. J Biol Chem. 2008;283:33211–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Scott DC, Monda JK, Grace CR, Duda DM, Kriwacki RW, Kurz T, Schulman BA. A dual E3 mechanism for Rub1 ligation to Cdc53. Mol Cell. 2010;39:784–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Latif F, Tory K, Gnarra J, Yao M, Duh FM, Orcutt ML, Stackhouse T, Kuzmin I, Modi W, Geil L, et al. Identification of the von Hippel–Lindau disease tumor suppressor gene. Science. 1993;260:1317–20.

    Article  CAS  PubMed  Google Scholar 

  39. Duan DR, Pause A, Burgess WH, Aso T, Chen DY, Garrett KP, Conaway RC, Conaway JW, Linehan WM, Klausner RD. Inhibition of transcription elongation by the VHL tumor suppressor protein. Science. 1995;269:1402–6.

    Article  CAS  PubMed  Google Scholar 

  40. Kishida T, Stackhouse TM, Chen F, Lerman MI, Zbar B. Cellular proteins that bind the von Hippel–Lindau disease gene product: mapping of binding domains and the effect of missense mutations. Cancer Res. 1995;55:4544–8.

    CAS  PubMed  Google Scholar 

  41. Kibel A, Iliopoulos O, DeCaprio JA, Kaelin WG Jr. Binding of the von Hippel–Lindau tumor suppressor protein to Elongin B and C. Science. 1995;269:1444–6.

    Article  CAS  PubMed  Google Scholar 

  42. Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, Wykoff CC, Pugh CW, Maher ER, Ratcliffe PJ. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature. 1999;399:271–5.

    Article  CAS  PubMed  Google Scholar 

  43. Ohh M, Park CW, Ivan M, Hoffman MA, Kim TY, Huang LE, Pavletich N, Chau V, Kaelin WG. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel–Lindau protein. Nat Cell Biol. 2000;2:423–7.

    Article  CAS  PubMed  Google Scholar 

  44. Iliopoulos O, Levy AP, Jiang C, Kaelin WG Jr, Goldberg MA. Negative regulation of hypoxia-inducible genes by the von Hippel–Lindau protein. Proc Natl Acad Sci USA. 1996;93:10595–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Wizigmann-Voos S, Breier G, Risau W, Plate KH. Up-regulation of vascular endothelial growth factor and its receptors in von Hippel–Lindau disease-associated and sporadic hemangioblastomas. Cancer Res. 1995;55:1358–64.

    CAS  PubMed  Google Scholar 

  46. Gnarra JR, Zhou S, Merrill MJ, Wagner JR, Krumm A, Papavassiliou E, Oldfield EH, Klausner RD, Linehan WM. Post-transcriptional regulation of vascular endothelial growth factor mRNA by the product of the VHL tumor suppressor gene. Proc Natl Acad Sci USA. 1996;93:10589–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Kourembanas S, Hannan RL, Faller DV. Oxygen tension regulates the expression of the platelet-derived growth factor-B chain gene in human endothelial cells. J Clin Invest. 1990;86:670–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Epstein AC, Gleadle JM, McNeill LA, Hewitson KS, O’Rourke J, Mole DR, Mukherji M, Metzen E, Wilson MI, Dhanda A, et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell. 2001;107:43–54.

    Article  CAS  PubMed  Google Scholar 

  49. Masson N, Willam C, Maxwell PH, Pugh CW, Ratcliffe PJ. Independent function of two destruction domains in hypoxia-inducible factor-alpha chains activated by prolyl hydroxylation. EMBO J. 2001;20:5197–206.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, von Kriegsheim A, Hebestreit HF, Mukherji M, Schofield CJ, et al. Targeting of HIF-alpha to the von Hippel–Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science. 2001;292:468–72.

    Article  CAS  PubMed  Google Scholar 

  51. Ivan M, Kondo K, Yang H, Kim W, Valiando J, Ohh M, Salic A, Asara JM, Lane WS, Kaelin WG Jr. HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science. 2001;292:464–8.

    Article  CAS  PubMed  Google Scholar 

  52. Hon WC, Wilson MI, Harlos K, Claridge TD, Schofield CJ, Pugh CW, Maxwell PH, Ratcliffe PJ, Stuart DI, Jones EY. Structural basis for the recognition of hydroxyproline in HIF-1 alpha by pVHL. Nature. 2002;417:975–8.

    Article  CAS  PubMed  Google Scholar 

  53. Min JH, Yang H, Ivan M, Gertler F, Kaelin WG Jr, Pavletich NP. Structure of an HIF-1alpha -pVHL complex: hydroxyproline recognition in signaling. Science. 2002;296:1886–9.

    Article  CAS  PubMed  Google Scholar 

  54. Zhang T, Niu X, Liao L, Cho EA, Yang H. The contributions of HIF-target genes to tumor growth in RCC. PLoS One. 2013;8:e80544.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Niu X, Zhang T, Liao L, Zhou L, Lindner DJ, Zhou M, Rini B, Yan Q, Yang H. The von Hippel–Lindau tumor suppressor protein regulates gene expression and tumor growth through histone demethylase JARID1C. Oncogene. 2012;31:776–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Rini BI. VEGF-targeted therapy in metastatic renal cell carcinoma. Oncologist. 2005;10:191–7.

    Article  CAS  PubMed  Google Scholar 

  57. Vano YA, Tartour E, Fournier LS, Beuselinck B, Mejean A, Oudard S. Prognostic factors in patients with advanced renal cell carcinoma treated with VEGF-targeted agents. Expert Rev Anticancer Ther. 2014;14:523–42.

    Article  CAS  PubMed  Google Scholar 

  58. Anderson K, Nordquist KA, Gao X, Hicks KC, Zhai B, Gygi SP, Patel TB. Regulation of cellular levels of Sprouty2 protein by prolyl hydroxylase domain and von Hippel–Lindau proteins. J Biol Chem. 2011;286:42027–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Mikhaylova O, Ignacak ML, Barankiewicz TJ, Harbaugh SV, Yi Y, Maxwell PH, Schneider M, Van Geyte K, Carmeliet P, Revelo MP, et al. The von Hippel–Lindau tumor suppressor protein and Egl-9-Type proline hydroxylases regulate the large subunit of RNA polymerase II in response to oxidative stress. Mol Cell Biol. 2008;28:2701–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Kuznetsova AV, Meller J, Schnell PO, Nash JA, Ignacak ML, Sanchez Y, Conaway JW, Conaway RC, Czyzyk-Krzeska MF. von Hippel–Lindau protein binds hyperphosphorylated large subunit of RNA polymerase II through a proline hydroxylation motif and targets it for ubiquitination. Proc Natl Acad Sci USA. 2003;100:2706–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Na X, Duan HO, Messing EM, Schoen SR, Ryan CK, di Sant’Agnese PA, Golemis EA, Wu G. Identification of the RNA polymerase II subunit hsRPB7 as a novel target of the von Hippel–Lindau protein. EMBO J. 2003;22:4249–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Iturrioz X, Parker PJ. PKCzetaII is a target for degradation through the tumour suppressor protein pVHL. FEBS Lett. 2007;581:1397–402.

    Article  CAS  PubMed  Google Scholar 

  63. Okuda H, Saitoh K, Hirai S, Iwai K, Takaki Y, Baba M, Minato N, Ohno S, Shuin T. The von Hippel–Lindau tumor suppressor protein mediates ubiquitination of activated atypical protein kinase C. J Biol Chem. 2001;276:43611–7.

    Article  CAS  PubMed  Google Scholar 

  64. Zhou L, Yang H. The von Hippel–Lindau tumor suppressor protein promotes c-Cbl-independent poly-ubiquitylation and degradation of the activated EGFR. PLoS One. 2011;6:e23936.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Yang H, Minamishima YA, Yan Q, Schlisio S, Ebert BL, Zhang X, Zhang L, Kim WY, Olumi AF, Kaelin WG Jr. pVHL acts as an adaptor to promote the inhibitory phosphorylation of the NF-kappaB agonist Card9 by CK2. Mol Cell. 2007;28:15–27.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Ohh M, Yauch RL, Lonergan KM, Whaley JM, Stemmer-Rachamimov AO, Louis DN, Gavin BJ, Kley N, Kaelin WG Jr, Iliopoulos O. The von Hippel–Lindau tumor suppressor protein is required for proper assembly of an extracellular fibronectin matrix. Mol Cell. 1998;1:959–68.

    Article  CAS  PubMed  Google Scholar 

  67. Lai Y, Song M, Hakala K, Weintraub ST, Shiio Y. Proteomic dissection of the von Hippel–Lindau (VHL) interactome. J Proteome Res. 2011;10:5175–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Lai Y, Song M, Hakala K, Weintraub ST, Shiio Y. The interaction of the von Hippel–Lindau tumor suppressor and heterochromatin protein 1. Arch Biochem Biophys. 2012;518:103–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Starostina NG, Simpliciano JM, McGuirk MA, Kipreos ET. CRL2(LRR-1) targets a CDK inhibitor for cell cycle control in C. elegans and actin-based motility regulation in human cells. Dev Cell. 2010;19:753–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Feng H, Zhong W, Punkosdy G, Gu S, Zhou L, Seabolt EK, Kipreos ET. CUL-2 is required for the G1-to-S-phase transition and mitotic chromosome condensation in Caenorhabditis elegans. Nat Cell Biol. 1999;1:486–92.

    Article  CAS  PubMed  Google Scholar 

  71. Merlet J, Burger J, Tavernier N, Richaudeau B, Gomes JE, Pintard L. The CRL2LRR-1 ubiquitin ligase regulates cell cycle progression during C. elegans development. Development. 2010;137:3857–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Ayyub C. Cullin-5 and cullin-2 play a role in the development of neuromuscular junction and the female germ line of Drosophila. J Genet. 2011;90:239–49.

    Article  CAS  PubMed  Google Scholar 

  73. Monahan AJ, Starz-Gaiano M. Socs36E limits STAT signaling via Cullin2 and a SOCS-box independent mechanism in the Drosophila egg chamber. Mech Dev. 2015;138(Pt 3):313–27.

    Article  CAS  PubMed  Google Scholar 

  74. Qian Y, Ng CL, Schulz C. CSN maintains the germline cellular microenvironment and controls the level of stem cell genes via distinct CRLs in testes of Drosophila melanogaster. Dev Biol. 2015;398:68–79.

    Article  CAS  PubMed  Google Scholar 

  75. Brauchle M, Baumer K, Gonczy P. Differential activation of the DNA replication checkpoint contributes to asynchrony of cell division in C. elegans embryos. Curr Biol. 2003;13:819–27.

    Article  CAS  PubMed  Google Scholar 

  76. Kalogeropoulos N, Christoforou C, Green AJ, Gill S, Ashcroft NR. chk-1 is an essential gene and is required for an S-M checkpoint during early embryogenesis. Cell Cycle. 2004;3:1196–200.

    Article  CAS  PubMed  Google Scholar 

  77. Burger J, Merlet J, Tavernier N, Richaudeau B, Arnold A, Ciosk R, Bowerman B, Pintard L. CRL2(LRR-1) E3-ligase regulates proliferation and progression through meiosis in the Caenorhabditis elegans germline. PLoS Genet. 2013;9:e1003375.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Merlet J, Pintard L. Role of the CRL2(LRR-1) E3 ubiquitin-ligase in the development of the germline in C. elegans. Worm. 2013;2:e25716.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Hodgkin J, Doniach T, Shen M. The sex determination pathway in the nematode Caenorhabditis elegans: variations on a theme. Cold Spring Harb Symp Quant Biol. 1985;50:585–93.

    Article  CAS  PubMed  Google Scholar 

  80. Krakow D, Sebald E, King LM, Cohn DH. Identification of human FEM1A, the ortholog of a C. elegans sex-differentiation gene. Gene. 2001;279:213–9.

    Article  CAS  PubMed  Google Scholar 

  81. Starostina NG, Lim JM, Schvarzstein M, Wells L, Spence AM, Kipreos ET. A CUL-2 ubiquitin ligase containing three FEM proteins degrades TRA-1 to regulate C. elegans sex determination. Dev Cell. 2007;13:127–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Gilder AS, Chen YB, Jackson RJ 3rd, Jiang J, Maher JF. Fem1b promotes ubiquitylation and suppresses transcriptional activity of Gli1. Biochem Biophys Res Commun. 2013;440:431–6.

    Article  CAS  PubMed  Google Scholar 

  83. Goodarzi MO, Maher JF, Cui J, Guo X, Taylor KD, Azziz R. FEM1A and FEM1B: novel candidate genes for polycystic ovary syndrome. Hum Reprod. 2008;23:2842–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Lu D, Ventura-Holman T, Li J, McMurray RW, Subauste JS, Maher JF. Abnormal glucose homeostasis and pancreatic islet function in mice with inactivation of the Fem1b gene. Mol Cell Biol. 2005;25:6570–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Subauste MC, Sansom OJ, Porecha N, Raich N, Du L, Maher JF. Fem1b, a proapoptotic protein, mediates proteasome inhibitor-induced apoptosis of human colon cancer cells. Mol Carcinog. 2010;49:105–13.

    CAS  PubMed  Google Scholar 

  86. Subauste MC, Ventura-Holman T, Lu D, Du L, Sansom OJ, Maher JF. Fem1b antigen in the stool of ApcMin mice as a biomarker of early Wnt signaling activation in intestinal neoplasia. Cancer Epidemiol. 2011;35:97–100.

    Article  CAS  PubMed  Google Scholar 

  87. Samadani AA, Akhavan-Niaki H. Interaction of sonic hedgehog (SHH) pathway with cancer stem cell genes in gastric cancer. Med Oncol. 2015;32:48.

    Article  PubMed  CAS  Google Scholar 

  88. Shi YQ, Liao SY, Zhuang XJ, Han CS. Mouse Fem1b interacts with and induces ubiquitin-mediated degradation of Ankrd37. Gene. 2011;485:153–9.

    Article  CAS  PubMed  Google Scholar 

  89. Wang X, Desai N, Hu YP, Price SM, Abate-Shen C, Shen MM. Mouse Fem1b interacts with the Nkx3.1 homeoprotein and is required for proper male secondary sexual development. Dev Dyn. 2008;237:2963–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Wadelin F, Fulton J, McEwan PA, Spriggs KA, Emsley J, Heery DM. Leucine-rich repeat protein PRAME: expression, potential functions and clinical implications for leukaemia. Mol Cancer. 2010;9:226.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Epping MT, Wang L, Edel MJ, Carlee L, Hernandez M, Bernards R. The human tumor antigen PRAME is a dominant repressor of retinoic acid receptor signaling. Cell. 2005;122:835–47.

    Article  CAS  PubMed  Google Scholar 

  92. Costessi A, Mahrour N, Tijchon E, Stunnenberg R, Stoel MA, Jansen PW, Sela D, Martin-Brown S, Washburn MP, Florens L, et al. The tumour antigen PRAME is a subunit of a Cul2 ubiquitin ligase and associates with active NFY promoters. EMBO J. 2011;30:3786–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Downey M, Houlsworth R, Maringele L, Rollie A, Brehme M, Galicia S, Guillard S, Partington M, Zubko MK, Krogan NJ, et al. A genome-wide screen identifies the evolutionarily conserved KEOPS complex as a telomere regulator. Cell. 2006;124:1155–68.

    Article  CAS  PubMed  Google Scholar 

  94. Kisseleva-Romanova E, Lopreiato R, Baudin-Baillieu A, Rousselle JC, Ilan L, Hofmann K, Namane A, Mann C, Libri D. Yeast homolog of a cancer-testis antigen defines a new transcription complex. EMBO J. 2006;25:3576–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Costessi A, Mahrour N, Sharma V, Stunnenberg R, Stoel MA, Tijchon E, Conaway JW, Conaway RC, Stunnenberg HG. The human EKC/KEOPS complex is recruited to Cullin2 ubiquitin ligases by the human tumour antigen PRAME. PLoS One. 2012;7:e42822.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Kemphues KJ, Wolf N, Wood WB, Hirsh D. Two loci required for cytoplasmic organization in early embryos of Caenorhabditis elegans. Dev Biol. 1986;113:449–60.

    Article  CAS  PubMed  Google Scholar 

  97. Vasudevan S, Starostina NG, Kipreos ET. The Caenorhabditis elegans cell-cycle regulator ZYG-11 defines a conserved family of CUL-2 complex components. EMBO Rep. 2007;8:279–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Feral C, Wu YQ, Pawlak A, Guellaen G. Meiotic human sperm cells express a leucine-rich homologue of Caenorhabditis elegans early embryogenesis gene, Zyg-11. Mol Hum Reprod. 2001;7:1115–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Wu JI, Lessard J, Crabtree GR. Understanding the words of chromatin regulation. Cell. 2009;136:200–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Li XS, Trojer P, Matsumura T, Treisman JE, Tanese N. Mammalian SWI/SNF—a subunit BAF250/ARID1 is an E3 ubiquitin ligase that targets histone H2B. Mol Cell Biol. 2010;30:1673–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Subauste MC, Ventura-Holman T, Du L, Subauste JS, Chan SL, Yu VC, Maher JF. RACK1 downregulates levels of the pro-apoptotic protein Fem1b in apoptosis-resistant colon cancer cells. Cancer Biol Ther. 2009;8:2297–305.

    Article  PubMed  Google Scholar 

  102. Fomenkov A, Zangen R, Huang YP, Osada M, Guo Z, Fomenkov T, Trink B, Sidransky D, Ratovitski EA. RACK1 and stratifin target DeltaNp63alpha for a proteasome degradation in head and neck squamous cell carcinoma cells upon DNA damage. Cell Cycle. 2004;3:1285–95.

    Article  CAS  PubMed  Google Scholar 

  103. Liu YV, Baek JH, Zhang H, Diez R, Cole RN, Semenza GL. RACK1 competes with HSP90 for binding to HIF-1alpha and is required for O(2)-independent and HSP90 inhibitor-induced degradation of HIF-1alpha. Mol Cell. 2007;25:207–17.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  104. Zhang W, Wang H, Li Z, Liu X, Liu G, Harris RS, Yu XF. Cellular requirements for bovine immunodeficiency virus Vif-mediated inactivation of bovine APOBEC3 proteins. J Virol. 2014;88:12528–40.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Jia L, Li H, Sun Y. Induction of p21-dependent senescence by an NAE inhibitor, MLN4924, as a mechanism of growth suppression. Neoplasia. 2011;13:561–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Luo Z, Yu G, Lee HW, Li L, Wang L, Yang D, Pan Y, Ding C, Qian J, Wu L, et al. The Nedd8-activating enzyme inhibitor MLN4924 induces autophagy and apoptosis to suppress liver cancer cell growth. Cancer Res. 2012;72:3360–71.

    Article  CAS  PubMed  Google Scholar 

  107. Milhollen MA, Traore T, Adams-Duffy J, Thomas MP, Berger AJ, Dang L, Dick LR, Garnsey JJ, Koenig E, Langston SP, et al. MLN4924, a NEDD8-activating enzyme inhibitor, is active in diffuse large B-cell lymphoma models: rationale for treatment of NF-{kappa}B-dependent lymphoma. Blood. 2010;116:1515–23.

    Article  CAS  PubMed  Google Scholar 

  108. Nawrocki ST, Griffin P, Kelly KR, Carew JS. MLN4924: a novel first-in-class inhibitor of NEDD8-activating enzyme for cancer therapy. Expert Opin Investig Drugs. 2012;21:1563–73.

    Article  CAS  PubMed  Google Scholar 

  109. Soucy TA, Dick LR, Smith PG, Milhollen MA, Brownell JE. The NEDD8 conjugation pathway and its relevance in cancer biology and therapy. Genes Cancer. 2010;1:708–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Tanaka T, Nakatani T, Kamitani T. Inhibition of NEDD8-conjugation pathway by novel molecules: potential approaches to anticancer therapy. Mol Oncol. 2012;6:267–75.

    Article  CAS  PubMed  Google Scholar 

  111. Tanaka T, Nakatani T, Kamitani T. Negative regulation of NEDD8 conjugation pathway by novel molecules and agents for anticancer therapy. Curr Pharm Des. 2013;19:4131–9.

    Article  CAS  PubMed  Google Scholar 

  112. Xu J, Li L, Yu G, Ying W, Gao Q, Zhang W, Li X, Ding C, Jiang Y, Wei D, et al. The neddylation-cullin 2-RBX1 E3 ligase axis targets tumor suppressor RhoB for degradation in liver cancer. Mol Cell Proteomics. 2015;14:499–509.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Zhang Y, Yang Y, Yeh S, Chang C. ARA67/PAT1 functions as a repressor to suppress androgen receptor transactivation. Mol Cell Biol. 2004;24:1044–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Saito-Ohara F, Imoto I, Inoue J, Hosoi H, Nakagawara A, Sugimoto T, Inazawa J. PPM1D is a potential target for 17q gain in neuroblastoma. Cancer Res. 2003;63:1876–83.

    CAS  PubMed  Google Scholar 

  115. Hirasawa A, Saito-Ohara F, Inoue J, Aoki D, Susumu N, Yokoyama T, Nozawa S, Inazawa J, Imoto I. Association of 17q21–q24 gain in ovarian clear cell adenocarcinomas with poor prognosis and identification of PPM1D and APPBP2 as likely amplification targets. Clin Cancer Res. 2003;9:1995–2004.

    CAS  PubMed  Google Scholar 

  116. Ehrbrecht A, Muller U, Wolter M, Hoischen A, Koch A, Radlwimmer B, Actor B, Mincheva A, Pietsch T, Lichter P, et al. Comprehensive genomic analysis of desmoplastic medulloblastomas: identification of novel amplified genes and separate evaluation of the different histological components. J Pathol. 2006;208:554–63.

    Article  CAS  PubMed  Google Scholar 

  117. Zhou HJ, Wong CM, Chen JH, Qiang BQ, Yuan JG, Jin DY. Inhibition of LZIP-mediated transcription through direct interaction with a novel host cell factor-like protein. J Biol Chem. 2001;276:28933–8.

    Article  CAS  PubMed  Google Scholar 

  118. Ohinata Y, Sutou S, Mitsui Y. A novel testis-specific RAG2-like protein, Peas: its expression in pachytene spermatocyte cytoplasm and meiotic chromatin. FEBS Lett. 2003;537:1–5.

    Article  CAS  PubMed  Google Scholar 

  119. Nishito Y, Hasegawa M, Inohara N, Nunez G. MEX is a testis-specific E3 ubiquitin ligase that promotes death receptor-induced apoptosis. Biochem J. 2006;396:411–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Meyer MA. Highly expressed genes in human high grade gliomas: immunohistochemical analysis of data from the human protein atlas. Neurol Int. 2014;6:5348.

    PubMed  PubMed Central  Google Scholar 

  121. Kai N, Iwase K, Imai K, Nakahira E, Soma M, Ohtsuka S, Yagi T, Kobayashi K, Koga H, Takiguchi M, et al. Altered gene expression in the subdivisions of the amygdala of Fyn-deficient mice as revealed by laser capture microdissection and mKIAA cDNA array analysis. Brain Res. 2006;1073–1074:60–70.

    Article  PubMed  CAS  Google Scholar 

  122. Smith JD, Hing AV, Clarke CM, Johnson NM, Perez FA, Park SS, Horst JA, Mecham B, Maves L, Nickerson DA, et al. Exome sequencing identifies a recurrent de novo ZSWIM6 mutation associated with acromelic frontonasal dysostosis. Am J Hum Genet. 2014;95:235–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Moran E. Interaction of adenoviral proteins with pRB and p53. FASEB J. 1993;7:880–5.

    CAS  PubMed  Google Scholar 

  124. Blanchette P, Cheng CY, Yan Q, Ketner G, Ornelles DA, Dobner T, Conaway RC, Conaway JW, Branton PE. Both BC-box motifs of adenovirus protein E4orf6 are required to efficiently assemble an E3 ligase complex that degrades p53. Mol Cell Biol. 2004;24:9619–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Cheng CY, Blanchette P, Branton PE. The adenovirus E4orf6 E3 ubiquitin ligase complex assembles in a novel fashion. Virology. 2007;364:36–44.

    Article  CAS  PubMed  Google Scholar 

  126. Cheng CY, Gilson T, Dallaire F, Ketner G, Branton PE, Blanchette P. The E4orf6/E1B55K E3 ubiquitin ligase complexes of human adenoviruses exhibit heterogeneity in composition and substrate specificity. J Virol. 2011;85:765–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Zou XH, Li WJ, Guo XJ, Qu JG, Wang M, Si HL, Lu ZZ, Hung T. Inefficient export of viral late mRNA contributes to fastidiousness of human adenovirus type 41 (HAdV-41) in 293 cells. Virology. 2014;468–470:388–96.

    Article  PubMed  CAS  Google Scholar 

  128. Moore M, Horikoshi N, Shenk T. Oncogenic potential of the adenovirus E4orf6 protein. Proc Natl Acad Sci USA. 1996;93:11295–301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Querido E, Marcellus RC, Lai A, Charbonneau R, Teodoro JG, Ketner G, Branton PE. Regulation of p53 levels by the E1B 55-kilodalton protein and E4orf6 in adenovirus-infected cells. J Virol. 1997;71:3788–98.

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Stracker TH, Carson CT, Weitzman MD. Adenovirus oncoproteins inactivate the Mre11-Rad50-NBS1 DNA repair complex. Nature. 2002;418:348–52.

    Article  CAS  PubMed  Google Scholar 

  131. Baker A, Rohleder KJ, Hanakahi LA, Ketner G. Adenovirus E4 34k and E1b 55k oncoproteins target host DNA ligase IV for proteasomal degradation. J Virol. 2007;81:7034–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Dallaire F, Blanchette P, Groitl P, Dobner T, Branton PE. Identification of integrin alpha3 as a new substrate of the adenovirus E4orf6/E1B 55-kilodalton E3 ubiquitin ligase complex. J Virol. 2009;83:5329–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Forrester NA, Sedgwick GG, Thomas A, Blackford AN, Speiseder T, Dobner T, Byrd PJ, Stewart GS, Turnell AS, Grand RJ. Serotype-specific inactivation of the cellular DNA damage response during adenovirus infection. J Virol. 2011;85:2201–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Cheng CY, Gilson T, Wimmer P, Schreiner S, Ketner G, Dobner T, Branton PE, Blanchette P. Role of E1B55K in E4orf6/E1B55K E3 ligase complexes formed by different human adenovirus serotypes. J Virol. 2013;87:6232–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Gilson T, Cheng CY, Hur WS, Blanchette P, Branton PE. Analysis of the Cullin binding sites of the E4orf6 proteins of human adenovirus E3 ubiquitin ligases. J Virol. 2014;88:3885–97.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  136. Blackford AN, Patel RN, Forrester NA, Theil K, Groitl P, Stewart GS, Taylor AM, Morgan IM, Dobner T, Grand RJ, et al. Adenovirus 12 E4orf6 inhibits ATR activation by promoting TOPBP1 degradation. Proc Natl Acad Sci USA. 2010;107:12251–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Tsurumi T. EBV replication enzymes. Curr Top Microbiol Immunol. 2001;258:65–87.

    CAS  PubMed  Google Scholar 

  138. Kudoh A, Fujita M, Zhang L, Shirata N, Daikoku T, Sugaya Y, Isomura H, Nishiyama Y, Tsurumi T. Epstein–Barr virus lytic replication elicits ATM checkpoint signal transduction while providing an S-phase-like cellular environment. J Biol Chem. 2005;280:8156–63.

    Article  CAS  PubMed  Google Scholar 

  139. Ou YH, Chung PH, Sun TP, Shieh SY. p53 C-terminal phosphorylation by CHK1 and CHK2 participates in the regulation of DNA-damage-induced C-terminal acetylation. Mol Biol Cell. 2005;16:1684–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Sato Y, Kamura T, Shirata N, Murata T, Kudoh A, Iwahori S, Nakayama S, Isomura H, Nishiyama Y, Tsurumi T. Degradation of phosphorylated p53 by viral protein-ECS E3 ligase complex. PLoS Pathog. 2009;5:e1000530.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  141. Sato Y, Shirata N, Kudoh A, Iwahori S, Nakayama S, Murata T, Isomura H, Nishiyama Y, Tsurumi T. Expression of Epstein–Barr virus BZLF1 immediate-early protein induces p53 degradation independent of MDM2, leading to repression of p53-mediated transcription. Virology. 2009;388:204–11.

    Article  CAS  PubMed  Google Scholar 

  142. Lowy DR, Solomon D, Hildesheim A, Schiller JT, Schiffman M. Human papillomavirus infection and the primary and secondary prevention of cervical cancer. Cancer. 2008;113:1980–93.

    Article  PubMed  Google Scholar 

  143. Munger K, Baldwin A, Edwards KM, Hayakawa H, Nguyen CL, Owens M, Grace M, Huh K. Mechanisms of human papillomavirus-induced oncogenesis. J Virol. 2004;78:11451–60.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  144. Huh K, Zhou X, Hayakawa H, Cho JY, Libermann TA, Jin J, Harper JW, 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;81:9737–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Dyson N, Howley PM, Munger K, Harlow E. The human papilloma virus-16 E7 oncoprotein is able to bind to the retinoblastoma gene product. Science. 1989;243:934–7.

    Article  CAS  PubMed  Google Scholar 

  146. Munger K, Werness BA, Dyson N, Phelps WC, Harlow E, Howley PM. Complex formation of human papillomavirus E7 proteins with the retinoblastoma tumor suppressor gene product. EMBO J. 1989;8:4099–105.

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Dyson N, Guida P, Munger K, Harlow E. Homologous sequences in adenovirus E1A and human papillomavirus E7 proteins mediate interaction with the same set of cellular proteins. J Virol. 1992;66:6893–902.

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Jones DL, Munger K. Analysis of the p53-mediated G1 growth arrest pathway in cells expressing the human papillomavirus type 16 E7 oncoprotein. J Virol. 1997;71:2905–12.

    CAS  PubMed  PubMed Central  Google Scholar 

  149. White EA, Sowa ME, Tan MJ, Jeudy S, Hayes SD, Santha S, Munger K, Harper JW, Howley PM. Systematic identification of interactions between host cell proteins and E7 oncoproteins from diverse human papillomaviruses. Proc Natl Acad Sci USA. 2012;109:E260–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Conticello SG, Harris RS, Neuberger MS. The Vif protein of HIV triggers degradation of the human antiretroviral DNA deaminase APOBEC3G. Curr Biol. 2003;13:2009–13.

    Article  CAS  PubMed  Google Scholar 

  151. Zhang J, Wu J, Wang W, Wu H, Yu B, Wang J, Lv M, Wang X, Zhang H, Kong W, et al. Role of cullin-elonginB-elonginC E3 complex in bovine immunodeficiency virus and maedi-visna virus Vif-mediated degradation of host A3Z2-Z3 proteins. Retrovirology. 2014;11:77.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Authors’ contributions

WC wrote the first draft of the manuscript. HY revised the manuscript and provided the guidance for revision. Both authors read and approved the final manuscript.

Acknowledgements

None.

Competing interests

The authors declare that they have no competing interests.

Funding

Haifeng Yang is funded by Grants 5R01CA155015 and 5P30CA056036-14.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Weijia Cai or Haifeng Yang.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cai, W., Yang, H. The structure and regulation of Cullin 2 based E3 ubiquitin ligases and their biological functions. Cell Div 11, 7 (2016). https://doi.org/10.1186/s13008-016-0020-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13008-016-0020-7

Keywords