Open access peer-reviewed chapter

E3 Ubiquitin Ligases in Cancer and Their Pharmacological Targeting

Written By

Joseph Y. Ong and Jorge Z. Torres

Submitted: 26 June 2018 Reviewed: 04 December 2018 Published: 15 February 2019

DOI: 10.5772/intechopen.82883

From the Edited Volume

Ubiquitin Proteasome System - Current Insights into Mechanism Cellular Regulation and Disease

Edited by Matthew Summers

Chapter metrics overview

1,969 Chapter Downloads

View Full Metrics

Abstract

Ubiquitination plays many critical roles in protein function and regulation. Consequently, mutation and aberrant expression of E3 ubiquitin ligases can drive cancer progression. Identifying key ligase-substrate relationships is crucial to understanding the molecular basis and pathways behind cancer and toward identifying novel targets for cancer therapeutics. Here, we review the importance of E3 ligases in the regulating the hallmarks of cancer, discuss some of the key and novel E3 ubiquitin ligases that drive tumor formation and angiogenesis, and review the clinical development of inhibitors that antagonize their function. We conclude with perspectives on the field and future directions toward understanding ubiquitination and cancer progression.

Keywords

  • E3 ubiquitin ligase
  • cancer
  • pharmacological targeting

1. Introduction

The regulation and turnover of proteins is an essential aspect of cell homeostasis and one that is commonly disrupted in cancer cells [1]. Regulation of a protein’s levels, activity, or localization is affected by ubiquitination, a posttranslational modification that involves the covalent attachment of a 76 amino acid ubiquitin molecule onto a substrate protein [2, 3]. Depending on the cellular context, ubiquitinated proteins can affect a myriad of cellular processes, including signaling [4], epigenetics [5], endosome trafficking [6], DNA repair [7] and protein stability via the 26S-proteasome [8].

The outcome of protein ubiquitination is affected primarily by two properties: what kind of ubiquitin linkage and how many ubiquitin molecules are present [2]. Ubiquitin is usually covalently attached to its substrate via a nucleophilic lysine residue on the substrate and the ubiquitin carboxy terminus. Ubiquitin itself can serve as a nucleophile via one of seven lysine residues (K6, K11, K27, K29, K33, K48, and K63) [9, 10] though K48- and K63-linkages seem to be the most abundant and are the most well-studied. In some cases, the N-terminal amide of the initiator methionine (M1) of the substrate can serve as the nucleophile [11, 12]. If one of the lysine residues or the initiator methionine of ubiquitin serves as the nucleophile for another ubiquitin molecule, a polyubiquitin chain is formed. A K48-linked polyubiquitin chain of four or more ubiquitin molecules is typically enough to target the substrate for 26S-proteasome mediated degradation [13]. Meanwhile, poly-K63 linkages are involved in many processes, including endocytic trafficking, inflammation, and DNA repair [5, 6, 14]. Other ubiquitin linkages [11], combinations of linkages (mixed or branched chains) [15, 16, 17], monoubiquitination [5, 18], and multi-monoubiquitination [19, 20] events have other diverse functions within the cell.

Ubiquitination occurs in three main steps [21, 22]. First, the E1 ubiquitin-activating enzyme (two in the human genome) covalently attaches to a ubiquitin molecule via a thioester bond in an ATP-dependent process. Next, the E1 enzyme transfers ubiquitin onto an E2 ubiquitin-conjugating enzyme (about 40 in the human genome). Finally, the E2 enzyme binds a substrate-bound E3 ligase (about 600 in the human genome) to transfer ubiquitin onto a lysine residue of the substrate. Repeating the cycle creates a polyubiquitin chain.

E3 ligases can function either as single peptides (like Parkin), simple complexes (e.g.: hetero/homodimers, like MDM2/MDMX or XIAP), or as large complexes (like Cullin-RING-ligase complexes or the anaphase promoting complex/cyclosome). There are two main classes of E3 ligases [23]: HECT (about 30 in the human genome) and RING ligases (including RING and RING-like ligases and their accessory proteins, about 600 in the human genome).

HECT ligases contain a C-terminus HECT domain that accepts the ubiquitin molecule from an E2 conjugating enzyme via a thioester bond before transferring the ubiquitin to the substrate [24]. RING ligases contain a zinc finger domain, and these proteins allow the E2 to transfer ubiquitin directly onto the substrate [25]. A subclass of RING ligases known as RING-between-RING (RBR) ligases contain two RING domains that have elements of both HECT and RING ligases: one RING domain binds the charged E2, while the other RING domain accepts the ubiquitin molecule before transferring it onto the substrate [26].

As E3 ligases ultimately determine the target of the ubiquitination machinery, they play a critical role in cell regulation. They regulate key players in processes like apoptosis (caspases), cell senescence and growth (p53, p21, p27; Hippo and Hedgehog signaling), proliferation and genomic stability (c-Myc, cyclins), immune system evasion (PD-L1), inflammation (NFκB), and metastasis and angiogenesis (Wnt signaling) (Figure 1). Misregulation or mutation of E3 ligases can lead to overexpression of oncogenes or downregulation of tumor suppressor genes, leading to cancer progression. Consequently, understanding the molecular targets and functions of E3 ligases serves as the basis for designing new cancer therapies.

Figure 1.

E3 ubiquitin ligases (outer circle) regulate hallmarks of cancer (inner circle) to drive cancer progression.

Here, we describe some central and novel E3 ligases related to cancer development, pharmacological targeting of those ligases, and perspectives on understanding the role of E3 ligases in cancer progression.

Advertisement

2. E3 ligases and cancer progression

2.1. TP53

The tumor protein p53 (TP53) is a transcription factor that serves as one of the principal regulators of cell function and survival (reviewed in [27]), mediating cellular responses to proliferation, cell cycle control, DNA damage response pathways, and apoptosis. Consequently, it is mutated in approximately 50% of all cancer types. Thus, regulators of p53 serve as ideal candidates to understand and address cancer cell progression (Table 1).

E3 ligase Notable substrates and binding partners Expression in cancer Cancer types
TP53 E6AP p53 Gain of function via HPV E6 Cervical, breast [38, 166]
MDM2/X p53 Overexpressed Many; liposarcomas [48, 167]
SCF Skp2 p21, p27 Overexpressed Many [95, 168]
Fbxw7 Cyclin E, mTOR Downregulated or dominant-negative mutant Many; endometrial, cervical, blood [64, 67, 169]
β-TrCP IkB, β-catenin, Wee1, Cdc25a/b Overexpressed (in some tissues) Many [60, 168]
APC/C Cdc20 Cyclin A/B, securin Overexpressed Pancreatic, lung, gastric [95, 168, 170]
Cdh1 Cdc20, Plk1, Aurora kinase A/B Underexpressed Many [171]
Other XIAP Caspases 3, 7, 9 Overexpressed Many [98, 99]
Park2 Cyclin D/E, Cdc20/Cdh1, tubulin Underexpressed Breast, pancreatic, colorectal, ovarian [172]
SPOP PD-L1, androgen and estrogen receptor Downregulated or dominant-negative mutant Prostate, endometrial, kidney [139, 141, 150]

Table 1.

E3 ligases and cancer progression.

E6AP (Ube3a) is a 100 kDa HECT domain ligase discovered for mediating the interaction between human papillomavirus protein E6 and p53 [28]. Neither E6AP nor E6 alone have a strong affinity for p53, but together, the E6/E6AP complex binds to p53 and changes the substrate specificity of E6AP [28], allowing E6AP to ubiquitinate p53 at the N-terminal DNA binding domain and target it for degradation [29]. Consequently, E6AP may play a role in HPV-mediated cervical cancers [30], particularly for those mediated by high-risk HPV16 strain, as E6 proteins from lower-risk strains of HPV lack the ability to degrade p53 [31].

The E6/E6AP complex plays other roles in cancer cell progression. Neither E6 nor E6AP alone can activate the hTERT promoter, but together, the E6/E6AP complex can activate the hTERT promoter, perhaps via interactions with c-Myc and NFX-1 to respectively activate and repress promoter activity [32]. The E6/E6AP complex has also been implicated in the ubiquitination of apoptosis-inducing proteins Bak [33], Fas [34], and TNFR1 [35]. Independent of E6 binding, endogenous E6AP targets include the tumor suppressor PML [36]; cell cycle regulators p27 [36], Cdk1, Cdk4; cell proliferation regulator MAPK1 [37];, and guanine nucleotide exchange factor ECT2 [38]. A published list of 130 likely substrates of E6AP includes β-catenin and PRMT5, proteins involved in cancer progression [37].

MDM2 is best known as a regulator of p53. MDM2 is a RING ligase [39] that forms stable heterodimers with a homolog, MDMX (MDM4), via their RING domains [40]. MDM2 localizes primarily in the nucleus bound to p300/CBP [41]. When complexed to p53, MDM2 inhibits p53 activity in two ways: first, MDM2 binds the N-terminal transactivation domain [42], inhibiting p53-mediated transcription [43]; secondly, MDM2 modulates p53 protein levels via ubiquitination near the C-terminus [44]. After MDM2 monoubiquitinates p53, p300 and CBP catalyze the polyubiquitination of p53, leading to p53 degradation [8, 41, 45]. Overexpression of MDM2 [46, 47], seen in many cancers where p53 is not mutated [48], leads to a loss of p53 activity.

During p53 activation, p53 is phosphorylated by multiple serine/threonine kinases at residues near the N-terminus, disrupting p53/MDM2 binding and stabilizing p53. For example, ATM kinase phosphorylates p53 at S15 [49] to promote p53-mediated transcription. Additionally, ATM phosphorylation of MDM2 on S395 disrupts the MDM2/p53 complex, allowing p53 to accumulate [50].

2.2. SCF complexes

The SCF complex is a multimeric ubiquitination complex with multiple roles in cell regulation (Table 1). The main scaffold of the SCF complex, Cullin 1 (Cul1), recruits the substrate to be ubiquitinated at the N-terminus and the charged ubiquitin at the C-terminus. Rather than bind the substrate directly, Cul1 uses two adaptor proteins: Cul1 binds directly to Skp1, which then binds to one of about 70 F-box proteins [51] that directly bind their substrates. At the C-terminus, Cul1 binds an adaptor protein, either Rbx1 or Rbx2 (also known as Roc1 or Roc2), that will bind a charged E2 ubiquitin conjugating enzyme [52, 53] .

Skp2 (Fbxl1) is a F-box protein that is most active during S-phase [54]. During S phase, Skp2 binds and ubiquitinates phosphorylated p27 [55] by binding the Cdk2-cyclin E complex [56]. Degradation of p27 frees inhibition of Cdk2-cyclinA/E complexes, allowing for progression into S-phase and entry into mitosis [57]. Other targets of Skp2 include p21 [58] and E-cadherin [59]. In some cases, Skp2 requires an accessory protein Cks1 to enhance binding to the substrate [60]. Skp2 both enhances c-Myc transcriptional activity and promotes c-Myc degradation [61]. Interestingly, p300-mediated acetylation of Skp2 changes the localization of Skp2 from nuclear to cytoplasmic, increasing cellular proliferation, motility, and tumorigenesis [59]. Skp2 is commonly overexpressed in a variety of cancers [62], including blood, colorectal, stomach, ovarian, and cervical cancers [60].

Fbxw7 (in yeast, Cdc4) contains a homodimerization domain, an F-box domain that binds Skp1, and eight WD40 repeats that form a beta-propeller structure to bind substrates [63]. Substrate binding is dependent on interaction between the arginine residues of the Fbxw7 WD40 domains and phosphorylated residues of the substrate in a recognition motif termed the Cdc4 phosphodegron (CPD) [63]. Mutations that disrupt substrate binding, especially point mutations of the arginine residues of the WD40 region, are commonly found in tumor samples [64]. Because Fbxw7 homodimerizes, these mutations may have a dominant-negative effect [65], as wild-type Fbxw7-mutant Fbxw7 dimers are able to effectively bind but not ubiquitinate their substrates [66]. Fbxw7 is deleted [67] or mutated in many cancers, with mutations being especially common in cancers of the bile duct and blood [68].

One well-characterized substrate of Fbxw7 is cyclin E [69]. The ubiquitination and degradation of cyclin E is dependent on phosphorylation of by Cdk2 and glycogen synthase kinase 3 (GSK3) [70]. Dimerization of Fbxw7 can also change its affinity for cyclin E as well as other substrates [71]. Other substrates of Fbxw7 include transcription factors c-Myc [72]; c-JUN, Notch 1; DNA-binding protein DEK [73]; and nutrient sensing protein mTOR [74]. Interestingly, the SV40 large T antigen contains a decoy CPD that can mislocalize Fbxw7 and inhibit Fbxw7-mediated degradation of cyclin E [75].

β-TrCP (BTRC), Fbxw1a (β-TrCP1) and Fbxw11 (β-TrCP2) are protein homologs that appear to have redundant roles [76]. These F-box proteins can form homo- and hetereodimers with each other [76] and use WD40 domains to bind a DSG phosphodegron motif (such as DpSGXXpS) [60]. Overexpression of β-TrCP is seen in various types of cancers, including colorectal, pancreatic, breast, ovarian and melanomas [77].

β-TrCP plays an important role as a regulator of Cdk1. One substrate of β-TrCP is Wee1, a kinase that inhibits Cdk1 activity [78]. Phosphorylation of Wee1 at S53 and S123 by Plk1 and Cdk1 respectively allow β-TrCP to bind to and ubiquitinate Wee1, activating Cdk1 during G2 to promote rapid entry into mitosis. Similarly, in prophase, β-TrCP also ubiquitinates Emi1, an inhibitor of the APC/C [79]. Consequently, β-TrCP accelerates mitotic progression both by increasing Cdk1 activity and activating the APC/C. In the case of DNA damage, checkpoint proteins hyperphosphorylate Cdc25a [80], a phosphatase that activates Cdk1 by removing repressive phosphorylation events. β-TrCP binds to and ubiquitinates hyperphosphorylated Cdc25a, deactivating Cdk1 and delaying the cell cycle. β-TrCP also ubiquitinates Cdc25b [81], a phosphatase that activates Cdk2/cyclin A and Cdk1/cyclin B to progress through the G2/M transition [82]. Other β-TrCP substrates that are linked to cancer progression include the IkB family [83], β-catenin [76] and MDM2 [84].

2.3. APC/C

Proper cell cycling and successful mitotic events rely on the coordinated accumulation and destruction of cyclins [85]. Disruption of this coordination can lead to aberrant mitotic events, aneuploidy, and cancer [86] (Table 1). While entry into mitosis is mediated by activation of Cdk1/2, progression through and exit from mitosis is mediated principally by the anaphase promoting complex or cyclosome (APC/C).

The APC/C is a 1.2 megadalton complex whose activity is necessary for entry to and exit from mitosis [87]. The structure of the human APC/C was solved via cryoEM to 7.4 angstrom resolution, allowing for the identification of 20 subunits of the APC/C and a mechanistic understanding of its function [88]. APC/C ubiquitin ligase activity depends on two activating subunits, Cdc20 or Cdh1 (coded by gene FRZ1; not to be confused with the gene CDH1, which codes for E-cadherin), which are necessary for APC/C binding to substrate and subsequent degradation [89] via K11 ubiquitin linkages [90]. In early mitosis, APC/C-Cdc20 degrades proteins such as cyclins A and B and Securin, the inhibitor of separase [91]. In later stages of mitosis and early G1, APC/C-Cdh1 degrades Cdc20, mitotic kinases like Plk1 and Aurora kinases A/B, and the contractile ring protein Anillin to ensure exit from mitosis and proper transition into G1 [92]. Binding of the substrate to APC/C is mediated by two main modalities [93]: for some substrates, Cdc20/Cdh1 binds the substrate through a KEN box motif; for others, both the APC/C subunit Apc10 and Cdc20/Cdh1 “sandwich” the substrate at the substrate’s D box. Some substrates have both and/or additional motifs to bind the APC/C and Cdc20/Cdh1 [92].

Cdc20 is found overexpressed in many cancers, including lung, oral, liver, and colon cancers [94, 95]. Cdh1 is generally a tumor suppressor, as downregulation of Cdh1 is found in some aggressive cancer cell types [95], and loss of Cdh1 sensitizes cells to DNA damage [96].

2.4. Other

X-linked inhibitor of apoptosis protein (XIAP) is a IAP family E3 ligase characterized by three N-terminal baculovirus IAP repeat domains and a C-terminal RING domain [97]. Like other IAPs, XIAP plays a central role in mediating the cell’s response to apoptosis. XIAP is overexpressed in many cancer cell lines, particularly in kidney and skin cancers [98, 99].

The linker region of XIAP between BIR1 and BIR2 binds to the active site and inhibits caspase 3 and caspase 7 [100]. The BIR3 domain of XIAP also binds to caspase 9, inhibiting caspase 9 dimerization and activity [101]. Moreover, XIAP ubiquitinates caspase 3 [102], caspase 9 [103], and caspase 7 [104] and targets them for degradation. As a final level of regulation, in addition to its ubiquitin E3 ligase role, XIAP can also function as a neddylation E3 ligase, neddylating and inhibiting the activity of caspases [105].

XIAP also plays important roles in cell motility. On one hand, XIAP degrades COMMD1 [106], a regulator of NFκB [107] and copper homeostasis. XIAP also binds to MAP3K7IP1, an event that activates kinase MAP3K7 to phosphorylate substrates leading to removal of NFκB inhibition [108]. XIAP also binds to survivin [109], activating NFκB signaling and encouraging cell metastasis by activating cell motility kinases Fadk1 and Src [110]. Conversely, XIAP has also been show to inhibit cell migration by binding to and ubiquitinating c-RAF to direct another ubiquitin ligase (CHIP) to degrade c-RAF [111]. Under non-stressed conditions, XIAP ubiquitinates and degrades MDM2, stabilizing p53 and inhibiting autophagy [112]. XIAP also binds to and monoubiquitinates TLE3, allowing β-catenin to activate Wnt-mediated transcription [113]. Finally, in addition to inflammation involving the NFκB pathway, XIAP suppresses TLR-based inflammation [114].

Park2 (PARKIN) is an RBR-E3 ligase with both RING and HECT ligase characteristics [115]. The Park2 locus is commonly deleted in cancers [116]. In mouse models, loss of Park2 causes spontaneous liver cancer [117] and contributes to colorectal cancer in mouse models [118]. Additionally, Park2 plays a central role in mitophagy [119], which may affect cell redox state [120], proliferation, and metastasis [121].

Park2 plays a prominent role in regulating cyclin levels. Park2 degrades cyclins D [122] and E [123] in a Cul1-dependent manner [124]. Park2 mutations found in cancer lead to stabilization of these G1/S-phase cyclins, an increase in the number of cells in S and G2/M phase [123, 124], and increased rates of cellular proliferation [122]. Moreover, Park2 associates with Cdc20 and Cdh1 during mitosis in an APC/C-independent manner and regulates the levels of many APC/C substrates including mitotic kinases and mitotic cyclins [125]. Park2 regulates microtubules and the mitotic spindle, cytokinetic bridge [126], cell motility [127], and invasion [128]. Park2 ubiquitinates and degrades HIF-1α to contribute to cell migration, and loss of Park2 leads to tumor metastasis in mouse models [129].

In Park2 knock-out mouse models, the resulting oxidative stress and the Warburg effect [130] caused an increase in the mRNA of Aim2, a protein involved in cytokine production [131]. In these mouse models, activation of Aim2 ultimately led to upregulation of PD-L1 in pancreatic tumors and lower rates of survival, an effect seen in human pancreatic tumors and patients [131]. Thus, Park2’s roles in metabolism may affect the ability of the immune system to regulate cancer progression.

SPOP is a Cul3 substrate adaptor mutated in about 10% of prostate cancers and some kidney cancers [132]. SPOP has three basic domains: an N-terminal MATH domain for substrate recognition [133], a BTB domain for dimerization and interaction with Cul3 [134], and a BACK domain which assembles SPOP dimers into oligomers [134], a mechanism which increases SPOP binding to and ubiquitination of the substrate [135]. As SPOP regulates many proteins responsible for maintaining cell integrity, mutations in the MATH domain that disrupt binding to substrate encourage cancer progression [136].

SPOP plays a role in immunotherapy by ubiquitinating and degrading PD-L1 [137]. SPOP binding mutants cannot ubiquitinate PD-L1, resulting in larger tumor growth and fewer tumor-infiltrating lymphocytes compared to tumors harboring wild-type SPOP in mouse models [137]. Similarly, pancreatic cancer samples with mutant SPOP had higher levels of PD-L1, demonstrating a role for SPOP in immune system invasion [137].

Other notable SPOP substrates include the apoptotic protein Daxx [138, 139], deSUMOlyase SENP7 [140], c-Myc [141], HDAC6 [142], Cdc20 [143], proto-oncogene DEK [144], phosphatases PTEN and Dusp7 [139], hedgehog pathway proteins Gli2 and Gli3 [145, 146], and BET transcriptional coactivators BRD2–4 [147, 148, 149]. SPOP is also closely tied to hormone-activated pathways, as steroid receptor coactivator SRC-3 [150], androgen receptor (AR) [151], enhancer of AR-mediated transcriptional activity TRIM24 [144], and estrogen receptor α (ERα) [136] are all substrates of SPOP. Finally, wild-type, but not mutant SPOP degrades ERG [152]. Interestingly, in some prostate cancer samples, some tumors expressed a fused ERG protein due to genome rearrangements, a phenotype driven by SPOP mutation [153]. Unlike wild-type ERG, these ERG-fusions lack an SPOP binding site, contributing to cancer progression [154].

Advertisement

3. E3 ligases and their inhibitors

One ubiquitin-proteasome inhibitor has already found use in the treatment of cancer: Bortezomib is a 26S-proteasome inhibitor approved for treating certain types of myeloma and lymphoma that binds to and inhibits the proteasome from degrading other proteins [155]. Another compound still in clinical development is MLN4924 (Pevonedistat), an inhibitor of the Nedd8-activating enzyme and thus of Cullin RING ligase complexes [155]. As ubiquitination plays many important roles in cell regulation, these broad inhibitors can affect many cellular pathways, not just those that are therapeutically useful. As E3 ligases are specific for their substrates, E3 ligases serve as precise targets for therapeutic intervention (Table 2). Inhibition of E3 ligases will hopefully minimize off-target effects. Moreover, as some E3 ligases have many oncogenes as their substrates, targeting E3 ligases may serve to be more efficient than targeting individual substrates.

E3 ligase Therapeutic Mechanism Model In clinical trials
In vitro assay Cell culture Mouse model
TP53 E6AP CM-11 peptides [161] Binds HECT domain X X
Compound 9 [173] Binds HPV E6 X X
MDM2/X Nutlins [156], RG7112 [174] Binds p53 binding site X X X
Idasanutlin (RG7388) [175] X X X X
MI-888 [176], SAR405838 [151] Binds p53 binding site X X X X
AMG-232 [177] Binds p53 binding site X X X X
NVP-CGM097 [178], HDM201 [179] Binds p53 binding site X X X X
JNJ-26854165 (Serdemetan) Assumed to bind to RING domain of MDM2 [180] X X X
ALRN-6924 [157] Stapled peptide binds MDM2 and MDMX at p53 binding site X X X X
SCF Skp2 Compound #25 [181] Binds Skp1 binding site X X X
C1, C2, C16, C20 [163, 182] Presumed: Binds Skp2, Cks1 at p27 binding site X X
CpdA [165] Inhibits Skp2-Skp1 binding X X
NSC689857, NSC681152 [164] Inhibits Skp2-Cks1 binding X
Fbxw7 Oridonin [183] Stabilizes Fbxw7, increases the activity of kinase Gsk-3 X X
β-TrCP Erioflorin [184] Inhibits β-TrCP1 binding to substrate X X
GS143 [185] Presumed: Inhibits binding of
β-TrCP1 and p-IkBa
X X
APC/C Cdc20 Apcin [186] Binds to D-box binding site of Cdc20 X X
Cdc20/Cdh1 ProTAME [187] Inhibits formation of APC/C-Cdc20, -Cdh1 X X X
Other XIAP LCL161 [158] Binds to BIR3 domain of XIAP [188] X X X X
AEG 35156 [189] XIAP antisense oligonucleotide X X
SPOP Palbociclib [137] Cdk4 phosphorylates SPOP, destabilizes PD-L1 X X X *
Compound 6b [160] Binds to substrate pocket X X X

Table 2.

E3 ligases and their inhibitors.

Palbociclib is clinically approved for treatment of breast cancer.


While most inhibitors have been identified via high throughput screens, the most clinically relevant inhibitors have been derived from structure–function analyses of E3 ligases complexed to their substrates. For example, the crystal structure of MDM2 bound to p53 allowed for the identification of the MDM2-p53 binding pocket and the design of small molecules [156] (like Nutlins and their derivatives) and stapled peptides [157] that bind to MDM2 and inhibit p53 binding. Similarly, the structure of the IAP family of E3 ligases and their endogenous inhibitors, the SMAC peptides, allowed for the development of higher affinity peptides [158] and peptidomimetics and the discovery of one small molecule inhibitor, Embelin [159]. Of the inhibitors mentioned here, MDM2 and XIAP inhibitors have advanced the farthest in clinical trials. A crystal structure of the SPOP substrate binding domain was also used to develop an SPOP inhibitor, suggesting that structural studies may greatly enhance development of small molecule inhibitors [160].

Most inhibitors disrupt E3 ligase-substrate binding by blocking the binding pocket of the E3 ligase. However, because HECT domains first transfer the ubiquitin molecule to themselves via a thioester bond [24], HECT ligases have an additional mode of pharmacological inhibition. The CM-11 peptides (E6AP inhibitors) are one such therapy that takes advantage of this step to inhibit or disrupt the HECT-Ubiquitin transthiolation reaction [161]. Future work may focus on designing small molecules that disrupt this function of the HECT domain.

To degrade its most clinically relevant targets p21 and p27, Skp2 functions with an adaptor protein, Cks1 [162]. At least two classes of inhibitors (NSC689857/NSC681152 [163] and the C1/2/16/20 compounds [164]) have been developed that disrupt the Skp2-Cks1 interaction. Similarly, the SCF ligase complex is only active upon the binding of an F-box protein to Skp1. CpdA inhibits Skp2-Skp1 binding [165]. These results suggest that another method of inhibitor design may focus on disrupting crucial activators and binding partners of E3 ligases instead of merely disrupting E3 ligase-substrate binding.

Upon phosphorylation by Cdk4, SPOP protein levels are stabilized, and PD-L1 expression levels decrease [137]. To improve the efficiency of anti-PD-L1 immunotherapies, mice treated with both Cdk4/6 inhibitors (to destabilize SPOP and thus stabilize PD-L1) and anti-PD-L1 immunotherapy showed improved survival when compared to untreated mice or mice with each individual treatment [137]. In this case, stabilization of an oncogenic protein led to improved efficacy of a complimentary therapy. Whether a similar combination of therapies can be used to improve the overall survival rate in other pathways remains to be seen.

Advertisement

4. Conclusions and perspectives

Recent research has highlighted the role of ubiquitination in cell regulation, division, and cancer cell progression. While much work has advanced the identification of E3 ubiquitin ligases and their substrates, untangling how these ligases act upon interconnected pathways remains a challenge in cancer cell biology. For example, understanding in which contexts certain E3 ligases are tumor-supportive or tumor-suppressive (like β-TrCP) is still not clear. Genome-wide analyses and advancements in systems biology have aided in and will continue to contribute to addressing these issues.

The tumor microenvironment has established itself as a central component in understanding and treating cancer progression. The macro-level questions of tumors—how cancers induce angiogenesis, interact with the immune system and cytokines, interact with the microbiome, and metastasize—are some questions that are best addressed with research in animal models, not human cell culture models. For example, the recent discoveries that both SPOP and Park2 play a role in mediating PD-L1 stability demonstrate the need to study the roles of E3 ligases in animal models. Given the recent success of immuno-oncology and CAR-T cell therapy, a further understanding how E3 ligases affect macro-level phenotypes like tumor sensitivity to immunotherapies may influence the design of clinical therapies.

While many E3 ligase inhibitors are being identified via high-throughput small molecule screens that assess inhibition of E3 ligase-substrate binding or ubiquitination activity, the most clinically advanced inhibitors have been refined from structural analysis of the E3 ligase binding pocket. The structures of many E3 ligases have already been determined (for example, all 11 ligases discussed here have at least a partial structure), so further pharmacological development may involve identifying binding pockets and designing inhibitors to perturb ligase function, and optimizing already identified inhibitors. On the other hand, E3 ligases are often redundant, so inhibition of one ligase may not completely stabilize a beneficial substrate. Nonetheless, the early clinical success of some E3 ligase inhibitors suggests that ubiquitin ligase inhibition is a promising venue for therapeutic intervention in cancer patients.

References

  1. 1. Hanahan D, Weinberg RA. Hallmarks of cancer: The next generation. Cell. 2011;144:646-674
  2. 2. Komander D, Rape M. The ubiquitin code. Annual Review of Biochemistry. 2012;81:203-229
  3. 3. Yau R, Rape M. The increasing complexity of the ubiquitin code. Nature Cell Biology. 2016;18:579-586
  4. 4. Tao M et al. ITCH K63-ubiquitinates the NOD2 binding protein, RIP2, to influence inflammatory signaling pathways. Current Biology. 2009;19:1255-1263
  5. 5. Sigismund S, Polo S, Di Fiore PP. Signaling through monoubiquitination. Current Topics in Microbiology and Immunology. 2004;286:149-185
  6. 6. Williams RL, Urbé S. The emerging shape of the ESCRT machinery. Nature Reviews. Molecular Cell Biology. 2007;8:355-368
  7. 7. Hoege C, Pfander B, Moldovan G-L, Pyrowolakis G, Jentsch S. RAD6-dependent DNA repair is linked to modification of PCNA by ubiquitin and SUMO. Nature. 2002;419:135-141
  8. 8. Haupt Y, Maya R, Kazaz A, Oren M. Mdm2 promotes the rapid degradation of p 53. Nature. 1997;387:296-299
  9. 9. Xu P et al. Quantitative proteomics reveals the function of unconventional ubiquitin chains in proteasomal degradation. Cell. 2009;137:133-145
  10. 10. Peng J et al. A proteomics approach to understanding protein ubiquitination. Nature Biotechnology. 2003;21:921-926
  11. 11. Emmerich CH et al. Activation of the canonical IKK complex by K63/M1-linked hybrid ubiquitin chains. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:15247-15252
  12. 12. Kirisako T et al. A ubiquitin ligase complex assembles linear polyubiquitin chains. The EMBO Journal. 2006;25:4877-4887
  13. 13. Thrower JS, Hoffman L, Rechsteiner M, Pickart CM. Recognition of the polyubiquitin proteolytic signal. The EMBO Journal. 2000;19:94-102
  14. 14. Chen ZJ, Sun LJ. Nonproteolytic functions of ubiquitin in cell signaling. Molecular Cell. 2009;33:275-286
  15. 15. Ben-Saadon R, Zaaroor D, Ziv T, Ciechanover A. The polycomb protein Ring1B generates self atypical mixed ubiquitin chains required for its In vitro histone H2A ligase activity. Molecular Cell. 2006;24:701-711
  16. 16. Liu C, Liu W, Ye Y, Li W. Ufd2p synthesizes branched ubiquitin chains to promote the degradation of substrates modified with atypical chains. Nature Communications. 2017;8:14274
  17. 17. Ohtake F, Tsuchiya H, Saeki Y, Tanaka K. K63 ubiquitylation triggers proteasomal degradation by seeding branched ubiquitin chains. Proceedings of the National Academy of Sciences of the United States of America. 2018;115:E1401-E1408
  18. 18. Alpi AF, Pace PE, Babu MM, Patel KJ. Mechanistic insight into site-restricted monoubiquitination of FANCD2 by Ube2t, FANCL, and FANCI. Molecular Cell. 2008;32:767-777
  19. 19. Braten O et al. Numerous proteins with unique characteristics are degraded by the 26S proteasome following monoubiquitination. Proceedings of the National Academy of Sciences of the United States of America. 2016;113:E4639-E4647
  20. 20. Kravtsova-Ivantsiv Y, Cohen S, Ciechanover A. Modification by single ubiquitin moieties rather than polyubiquitination is sufficient for proteasomal processing of the p105 NF-κB precursor. Molecular Cell. 2009;33:496-504
  21. 21. George AJ, Hoffiz YC, Charles AJ, Zhu Y, Mabb AM. A comprehensive atlas of E3 ubiquitin ligase mutations in neurological disorders. Frontiers in Genetics. 2018;9:29
  22. 22. Weissman AM, Shabek N, Ciechanover A. The predator becomes the prey: Regulating the ubiquitin system by ubiquitylation and degradation. Nature Reviews. Molecular Cell Biology. 2011;12:605-620
  23. 23. Li W et al. Genome-wide and functional annotation of human E3 ubiquitin ligases identifies MULAN, a mitochondrial E3 that regulates the Organelle’s dynamics and signaling. PLoS One. 2008;3:e1487
  24. 24. Scheffner M, Kumar S. Mammalian HECT ubiquitin-protein ligases: Biological and pathophysiological aspects. Biochimica et Biophysica Acta (BBA)—Molecular Cell Research. 2014;1843:61-74
  25. 25. Lipkowitz S, Weissman AM. RINGs of good and evil: RING finger ubiquitin ligases at the crossroads of tumour suppression and oncogenesis. Nature Reviews. Cancer. 2011;11:629-643
  26. 26. Dove KK, Stieglitz B, Duncan ED, Rittinger K, Klevit RE. Molecular insights into RBR E3 ligase ubiquitin transfer mechanisms. EMBO Reports. 2016;17:1221-1235
  27. 27. Kastenhuber ER, Lowe SW. Putting p53 in context. Cell. 2017;170:1062-1078
  28. 28. Scheffner M, Werness BA, Huibregtse JM, Levine AJ, Howley PM. The E6 oncoprotein encoded by human papillomavirus types 16 and 18 promotes the degradation of p53. Cell. 1990;63:1129-1136
  29. 29. Martinez-Zapien D et al. Structure of the E6/E6AP/p53 complex required for HPV-mediated degradation of p53. Nature. 2016;529:541-545
  30. 30. Alejo M et al. Contribution of human papillomavirus in neuroendocrine tumors from a series of 10,575 invasive cervical cancer cases. Papillomavirus Research. 2018;5:134-142
  31. 31. Elbel M, Carl S, Spaderna S, Iftner T. A comparative analysis of the interactions of the E6 proteins from cutaneous and genital papillomaviruses with p53 and E6AP in correlation to their transforming potential. Virology. 1997;239:132-149
  32. 32. Liu X et al. HPV E6 protein interacts physically and functionally with the cellular telomerase complex. Proceedings of the National Academy of Sciences. 2009;106:18780-18785
  33. 33. Thomas M, Banks L. Inhibition of Bak-induced apoptosis by HPV-18 E6. Oncogene. 1998;17:2943-2954
  34. 34. Filippova M, Parkhurst L, Duerksen-Hughes PJ. The human papillomavirus 16 E6 protein binds to Fas-associated death domain and protects cells from Fas-triggered apoptosis. The Journal of Biological Chemistry. 2004;279:25729-25744
  35. 35. Filippova M, Song H, Connolly JL, Dermody TS, Duerksen-Hughes PJ. The human papillomavirus 16 E6 protein binds to tumor necrosis factor (TNF) R1 and protects cells from TNF-induced apoptosis. The Journal of Biological Chemistry. 2002;277:21730-21739
  36. 36. Raghu D et al. E6AP promotes prostate cancer by reducing p27 expression. Oncotarget. 2017;8:42939-42948
  37. 37. Wang Y et al. Identifying the ubiquitination targets of E6AP by orthogonal ubiquitin transfer. Nature Communications. 2017;8:2232
  38. 38. Mansour M et al. The E3-ligase E6AP represses breast cancer metastasis via regulation of ECT2-rho signaling. Cancer Research. 2016;76:4236-4248
  39. 39. Fang S, Jensen JP, Ludwig RL, Vousden KH, Weissman AM. Mdm2 is a RING finger-dependent ubiquitin protein ligase for itself and p53. The Journal of Biological Chemistry. 2000;275:8945-8951
  40. 40. Wang X, Wang J, Jiang X. MdmX protein is essential for Mdm2 protein-mediated p53 polyubiquitination. The Journal of Biological Chemistry. 2011;286:23725-23734
  41. 41. Ferreon JC et al. Cooperative regulation of p53 by modulation of ternary complex formation with CBP/p300 and HDM2. Proceedings of the National Academy of Sciences. 2009;106:6591-6596
  42. 42. Kussie PH et al. Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain. Science. 1996;274:948-953
  43. 43. Oliner JD et al. Oncoprotein MDM2 conceals the activation domain of tumour suppressor p53. Nature. 1993;362:857-860
  44. 44. Poyurovsky MV et al. The C terminus of p53 binds the N-terminal domain of MDM2. Nature Structural & Molecular Biology. 2010;17:982-989
  45. 45. Grossman SR et al. Polyubiquitination of p53 by a ubiquitin ligase activity of p300. Science (80-). 2003;300:342-344
  46. 46. Zack TI et al. Pan-cancer patterns of somatic copy number alteration. Nature Genetics. 2013;45:1134-1140
  47. 47. Beroukhim R et al. The landscape of somatic copy-number alteration across human cancers. Nature. 2010;463:899-905
  48. 48. Oliner JD, Saiki AY, Caenepeel S. The role of MDM2 amplification and overexpression in tumorigenesis. Cold Spring Harbor Perspectives in Medicine. 2016;6:a026336
  49. 49. Canman CE et al. Activation of the ATM kinase by ionizing radiation and phosphorylation of p53. Science. 1998;281:1677-1679
  50. 50. Gannon HS, Woda BA, Jones SN. ATM phosphorylation of Mdm2 Ser394 regulates the amplitude and duration of the DNA damage response in mice. Cancer Cell. 2012;21:668-679
  51. 51. Lydeard JR, Schulman BA, Harper JW. Building and remodelling Cullin-RING E3 ubiquitin ligases. EMBO Reports. 2013;14:1050-1061
  52. 52. Wei D, Sun Y. Small RING finger proteins RBX1 and RBX2 of SCF E3 ubiquitin ligases: The role in cancer and as cancer targets. Genes & Cancer. 2010;1:700-707
  53. 53. Lyapina SA, Correll CC, Kipreos ET, Deshaies RJ. Human CUL1 forms an evolutionarily conserved ubiquitin ligase complex (SCF) with SKP1 and an F-box protein. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:7451-7456
  54. 54. Nakayama K et al. Skp2-mediated degradation of p27 regulates progression into mitosis. Developmental Cell. 2004;6:661-672
  55. 55. Carrano AC, Eytan E, Hershko A, Pagano M. SKP2 is required for ubiquitin-mediated degradation of the CDK inhibitor p27. Nature Cell Biology. 1999;1:193-199
  56. 56. Ungermannova D, Gao Y, Liu X. Ubiquitination of p27Kip1 requires physical interaction with cyclin E and probable phosphate recognition by SKP2. The Journal of Biological Chemistry. 2005;280:30301-30309
  57. 57. Zhang H, Kobayashi R, Galaktionov K, Beach D. p19Skp1 and p45Skp2 are essential elements of the cyclin A-CDK2 S phase kinase. Cell. 1995;82:915-925
  58. 58. Yu ZK, Gervais JL, Zhang H. Human CUL-1 associates with the SKP1/SKP2 complex and regulates p21(CIP1/WAF1) and cyclin D proteins. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:11324-11329
  59. 59. Inuzuka H et al. Acetylation-dependent regulation of Skp2 function. Cell. 2012;150:179-193
  60. 60. Frescas D, Pagano M. Deregulated proteolysis by the F-box proteins SKP2 and beta-TrCP: Tipping the scales of cancer. Nature Reviews. Cancer. 2008;8:438-449
  61. 61. von der Lehr N et al. The F-box protein Skp2 participates in c-Myc proteosomal degradation and acts as a cofactor for c-Myc-regulated transcription. Molecular Cell. 2003;11:1189-1200
  62. 62. Wei Z et al. Downregulation of Skp2 inhibits the growth and metastasis of gastric cancer cells in vitro and in vivo. Tumor Biology. 2013;34:181-192
  63. 63. Hao B, Oehlmann S, Sowa ME, Harper JW, Pavletich NP. Structure of a Fbw7-Skp1-Cyclin E complex: Multisite-phosphorylated substrate recognition by SCF ubiquitin ligases. Molecular Cell. 2007;26:131-143
  64. 64. Davis RJ, Welcker M, Clurman BE. Tumor suppression by the Fbw7 ubiquitin ligase: Mechanisms and opportunities. Cancer Cell. 2014;26:455-464
  65. 65. Tang X et al. Suprafacial orientation of the SCFCdc4 dimer accommodates multiple geometries for substrate ubiquitination. Cell. 2007;129:1165-1176
  66. 66. O’Neil J et al. FBW7 mutations in leukemic cells mediate NOTCH pathway activation and resistance to gamma-secretase inhibitors. The Journal of Experimental Medicine. 2007;204:1813-1824
  67. 67. Yeh C-H, Bellon M, Nicot C. FBXW7: A critical tumor suppressor of human cancers. Molecular Cancer. 2018;17:115
  68. 68. Akhoondi S et al. FBXW7/hCDC4 is a general tumor suppressor in human cancer. Cancer Research. 2007;67:9006-9012
  69. 69. Strohmaier H et al. Human F-box protein hCdc4 targets cyclin E for proteolysis and is mutated in a breast cancer cell line. Nature. 2001;413:316-322
  70. 70. Welcker M et al. Multisite phosphorylation by Cdk2 and GSK3 controls cyclin E degradation. Molecular Cell. 2003;12:381-392
  71. 71. Welcker M et al. Fbw7 dimerization determines the specificity and robustness of substrate degradation. Genes & Development. 2013;27:2531-2536
  72. 72. Yada M et al. Phosphorylation-dependent degradation of c-Myc is mediated by the F-box protein Fbw7. The EMBO Journal. 2004;23:2116-2125
  73. 73. Babaei-Jadidi R et al. FBXW7 influences murine intestinal homeostasis and cancer, targeting Notch, Jun, and DEK for degradation. The Journal of Experimental Medicine. 2011;208:295-312
  74. 74. Mao J-H et al. FBXW7 targets mTOR for degradation and cooperates with PTEN in tumor suppression. Science. 2008;321:1499-1502
  75. 75. Welcker M, Clurman BE. The SV40 large T antigen contains a decoy phosphodegron that mediates its interactions with Fbw7/hCdc4. The Journal of Biological Chemistry. 2005;280:7654-7658
  76. 76. Nakayama K et al. Impaired degradation of inhibitory subunit of NF-kappa B (I kappa B) and beta-catenin as a result of targeted disruption of the beta-TrCP1 gene. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:8752-8757
  77. 77. Zheng N, Zhou Q , Wang Z, Wei W. Recent advances in SCF ubiquitin ligase complex: Clinical implications. Biochimica et Biophysica Acta. 2016;1866:12-22
  78. 78. Watanabe N et al. M-phase kinases induce phospho-dependent ubiquitination of somatic Wee1 by SCFbeta-TrCP. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:4419-4424
  79. 79. Guardavaccaro D et al. Control of meiotic and mitotic progression by the F box protein beta-Trcp1 in vivo. Developmental Cell. 2003;4:799-812
  80. 80. Busino L et al. Degradation of Cdc25A by β-TrCP during S phase and in response to DNA damage. Nature. 2003;426:87-91
  81. 81. Kanemori Y, Uto K, Sagata N. Beta-TrCP recognizes a previously undescribed nonphosphorylated destruction motif in Cdc25A and Cdc25B phosphatases. Proceedings of the National Academy of Sciences of the United States of America. 2005;102:6279-6284
  82. 82. Lammer C et al. The cdc25B phosphatase is essential for the G2/M phase transition in human cells. Journal of Cell Science. 1998;111(Pt 16):2445-2453
  83. 83. Shirane M, Hatakeyama S, Hattori K, Nakayama K, Nakayama K. Common pathway for the ubiquitination of IkappaBalpha, IkappaBbeta, and IkappaBepsilon mediated by the F-box protein FWD1. The Journal of Biological Chemistry. 1999;274:28169-28174
  84. 84. Inuzuka H et al. Phosphorylation by casein kinase I promotes the turnover of the Mdm2 oncoprotein via the SCF(beta-TRCP) ubiquitin ligase. Cancer Cell. 2010;18:147-159
  85. 85. Satyanarayana A, Kaldis P. Mammalian cell-cycle regulation: Several Cdks, numerous cyclins and diverse compensatory mechanisms. Oncogene. 2009;28:2925-2939
  86. 86. Malumbres M, Barbacid M. Cell cycle, CDKs and cancer: A changing paradigm. Nature Reviews. Cancer. 2009;9:153-166
  87. 87. Zhang J, Wan L, Dai X, Sun Y, Wei W. Functional characterization of anaphase promoting complex/cyclosome (APC/C) E3 ubiquitin ligases in tumorigenesis. Biochimica et Biophysica Acta. 2014;1845:277-293
  88. 88. Chang L, Zhang Z, Yang J, McLaughlin SH, Barford D. Atomic structure of the APC/C and its mechanism of protein ubiquitination. Nature. 2015;522:450-454
  89. 89. Primorac I, Musacchio A. Panta rhei: The APC/C at steady state. The Journal of Cell Biology. 2013;201:177-189
  90. 90. Williamson A et al. Identification of a physiological E2 module for the human anaphase-promoting complex. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:18213-18218
  91. 91. Pines J. Cubism and the cell cycle: The many faces of the APC/C. Nature Reviews. Molecular Cell Biology. 2011;12:427-438
  92. 92. Davey NE, Morgan DO. Building a regulatory network with short linear sequence motifs: Lessons from the degrons of the anaphase-promoting complex. Molecular Cell. 2016;64:12-23
  93. 93. He J et al. Insights into degron recognition by APC/C coactivators from the structure of an Acm1-Cdh1 complex. Molecular Cell. 2013;50:649-660
  94. 94. Wu W et al. CDC20 overexpression predicts a poor prognosis for patients with colorectal cancer. Journal of Translational Medicine. 2013;11:142
  95. 95. Lehman NL et al. Oncogenic regulators and substrates of the anaphase promoting complex/cyclosome are frequently overexpressed in malignant tumors. The American Journal of Pathology. 2007;170:1793-1805
  96. 96. Ishizawa J et al. FZR1 loss increases sensitivity to DNA damage and consequently promotes murine and human B-cell acute leukemia. Blood. 2017;129:1958-1968
  97. 97. Gyrd-Hansen M, Meier P. IAPs: From caspase inhibitors to modulators of NF-κB, inflammation and cancer. Nature Reviews. Cancer. 2010;10:561-574
  98. 98. Fong WG et al. Expression and genetic analysis of XIAP-associated factor 1 (XAF1) in cancer cell lines. Genomics. 2000;70:113-122
  99. 99. Tamm I et al. Expression and prognostic significance of IAP-family genes in human cancers and myeloid leukemias. Clinical Cancer Research. 2000;6:1796-1803
  100. 100. Paulsen M et al. Interaction with XIAP prevents full caspase-3/-7 activation in proliferating human tlymphocytes. European Journal of Immunology. 2008;38:1979-1987
  101. 101. Shiozaki EN et al. Mechanism of XIAP-mediated inhibition of caspase-9. Molecular Cell. 2003;11:519-527
  102. 102. Suzuki Y, Nakabayashi Y, Takahashi R. Ubiquitin-protein ligase activity of X-linked inhibitor of apoptosis protein promotes proteasomal degradation of caspase-3 and enhances its anti-apoptotic effect in Fas-induced cell death. Proceedings of the National Academy of Sciences. 2001;98:8662-8667
  103. 103. Morizane Y, Honda R, Fukami K, Yasuda H. X-linked inhibitor of apoptosis functions as ubiquitin ligase toward mature caspase-9 and cytosolic Smac/DIABLO. Journal of Biochemistry. 2005;137:125-132
  104. 104. Creagh EM, Murphy BM, Duriez PJ, Duckett CS, Martin SJ. Smac/Diablo antagonizes ubiquitin ligase activity of inhibitor of apoptosis proteins. The Journal of Biological Chemistry. 2004;279:26906-26914
  105. 105. Broemer M et al. Systematic In vivo RNAi analysis identifies IAPs as NEDD8-E3 ligases. Molecular Cell. 2010;40:810-822
  106. 106. Burstein E et al. A novel role for XIAP in copper homeostasis through regulation of MURR1. The EMBO Journal. 2004;23:244-254
  107. 107. Ganesh L et al. The gene product Murr1 restricts HIV-1 replication in resting CD4+ lymphocytes. Nature. 2003;426:853-857
  108. 108. Lu M et al. XIAP induces NF-κB activation via the BIR1/TAB1 interaction and BIR1 dimerization. Molecular Cell. 2007;26:689-702
  109. 109. Arora V et al. Degradation of survivin by the X-linked inhibitor of apoptosis (XIAP)-XAF1 complex. The Journal of Biological Chemistry. 2007;282:26202-26209
  110. 110. Mehrotra S et al. IAP regulation of metastasis. Cancer Cell. 2010;17:53-64
  111. 111. Dogan T et al. X-linked and cellular IAPs modulate the stability of C-RAF kinase and cell motility. Nature Cell Biology. 2008;10:1447-1455
  112. 112. Huang X, Wu Z, Mei Y, Wu M. XIAP inhibits autophagy via XIAP-Mdm2-p53 signalling. The EMBO Journal. 2013;32:2204-2216
  113. 113. Hanson AJ et al. XIAP monoubiquitylates Groucho/TLE to promote canonical Wnt signaling. Molecular Cell. 2012;45:619-628
  114. 114. Lawlor KE et al. XIAP loss triggers RIPK3- and caspase-8-driven IL-1β activation and cell death as a consequence of TLR-MyD88-induced cIAP1-TRAF2 degradation. Cell Reports. 2017;20:668-682
  115. 115. Riley BE et al. Structure and function of Parkin E3 ubiquitin ligase reveals aspects of RING and HECT ligases. Nature Communications. 2013;4:1982
  116. 116. Cesari R et al. Parkin, a gene implicated in autosomal recessive juvenile parkinsonism, is a candidate tumor suppressor gene on chromosome 6q25-q27. Proceedings of the National Academy of Sciences. 2003;100:5956-5961
  117. 117. Fujiwara M et al. Parkin as a tumor suppressor gene for hepatocellular carcinoma. Oncogene. 2008;27:6002-6011
  118. 118. Poulogiannis G et al. PARK2 deletions occur frequently in sporadic colorectal cancer and accelerate adenoma development in Apc mutant mice. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:15145-15150
  119. 119. Chourasia AH, Boland ML, Macleod KF. Mitophagy and cancer. Cancer & Metabolism. 2015;3:4
  120. 120. Suen D-F, Narendra DP, Tanaka A, Manfredi G, Youle RJ. Parkin overexpression selects against a deleterious mtDNA mutation in heteroplasmic cybrid cells. Proceedings of the National Academy of Sciences. 2010;107:11835-11840
  121. 121. Gupta A et al. PARK2 depletion connects energy and oxidative stress to PI3K/Akt activation via PTEN S-nitrosylation. Molecular Cell. 2017;65:999-1013.e7
  122. 122. Yeo CWS et al. Parkin pathway activation mitigates glioma cell proliferation and predicts patient survival. Cancer Research. 2012;72:2543-2553
  123. 123. Staropoli JF et al. Parkin is a component of an SCF-like ubiquitin ligase complex and protects postmitotic neurons from kainate excitotoxicity. Neuron. 2003;37:735-749
  124. 124. Gong Y et al. Pan-cancer genetic analysis identifies PARK2 as a master regulator of G1/S cyclins. Nature Genetics. 2014;46:588-594
  125. 125. Lee SB et al. Parkin regulates mitosis and genomic stability through Cdc20/Cdh1. Molecular Cell. 2015;60:21-34
  126. 126. Sun X et al. Parkin deficiency contributes to pancreatic tumorigenesis by inducing spindle multipolarity and misorientation. Cell Cycle. 2013;12:1133-1141
  127. 127. Tay S-P et al. Parkin enhances the expression of cyclin-dependent kinase 6 and negatively regulates the proliferation of breast cancer cells. The Journal of Biological Chemistry. 2010;285:29231-29238
  128. 128. Wang H et al. PARK2 negatively regulates the metastasis and epithelial-mesenchymal transition of glioblastoma cells via ZEB1. Oncology Letters. 2017;14:2933-2939
  129. 129. Liu J et al. Parkin targets HIF-1α for ubiquitination and degradation to inhibit breast tumor progression. Nature Communications. 2017;8:1823
  130. 130. Zhang C et al. Parkin, a p53 target gene, mediates the role of p53 in glucose metabolism and the Warburg effect. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:16259-16264
  131. 131. Li C et al. PINK1 and PARK2 suppress pancreatic tumorigenesis through control of mitochondrial iron-mediated immunometabolism. Developmental Cell. 2018;46:441-455.e8
  132. 132. Frank S, Nelson P, Vasioukhin V. Recent advances in prostate cancer research: Large-scale genomic analyses reveal novel driver mutations and DNA repair defects. F1000 Research. 2018;7:1173
  133. 133. Zhuang M et al. Structures of SPOP-substrate complexes: Insights into molecular architectures of BTB-Cul3 ubiquitin ligases. Molecular Cell. 2009;36:39-50
  134. 134. van Geersdaele LK et al. Structural basis of high-order oligomerization of the cullin-3 adaptor SPOP. Acta Crystallographica. Section D, Biological Crystallography. 2013;69:1677-1684
  135. 135. Pierce WK et al. Multiple weak linear motifs enhance recruitment and processivity in SPOP-mediated substrate ubiquitination. Journal of Molecular Biology. 2016;428:1256-1271
  136. 136. Zhang P et al. Endometrial cancer-associated mutants of SPOP are defective in regulating estrogen receptor-α protein turnover. Cell Death & Disease. 2015;6:e1687
  137. 137. Zhang J et al. Cyclin D-CDK4 kinase destabilizes PD-L1 via cullin 3-SPOP to control cancer immune surveillance. Nature. 2017;553:91-95
  138. 138. Kwon JE et al. BTB domain-containing speckle-type POZ protein (SPOP) serves as an adaptor of Daxx for ubiquitination by Cul3-based ubiquitin ligase. The Journal of Biological Chemistry. 2006;281:12664-12672
  139. 139. Li G et al. SPOP promotes tumorigenesis by acting as a key regulatory hub in kidney cancer. Cancer Cell. 2014;25:455-468
  140. 140. Zhu H et al. SPOP E3 ubiquitin ligase adaptor promotes cellular senescence by degrading the SENP7 deSUMOylase. Cell Reports. 2015;13:1183-1193
  141. 141. Geng C et al. SPOP regulates prostate epithelial cell proliferation and promotes ubiquitination and turnover of c-MYC oncoprotein. Oncogene. 2017;36:4767-4777
  142. 142. Tan Y et al. Cullin 3 SPOP ubiquitin E3 ligase promotes the poly-ubiquitination and degradation of HDAC6. Oncotarget. 2017;8:47890-47901
  143. 143. Wu F et al. Prostate cancer-associated mutation in SPOP impairs its ability to target Cdc20 for poly-ubiquitination and degradation. Cancer Letters. 2017;385:207-214
  144. 144. Theurillat J-PP et al. Ubiquitylome analysis identifies dysregulation of effector substrates in SPOP-mutant prostate cancer. Science. 2014;346(80):85-89
  145. 145. Chen M-H et al. Cilium-independent regulation of Gli protein function by Sufu in hedgehog signaling is evolutionarily conserved. Genes & Development. 2009;23:1910-1928
  146. 146. Cai H, Liu A. Spop promotes skeletal development and homeostasis by positively regulating Ihh signaling. Proceedings of the National Academy of Sciences. 2016;113:14751-14756
  147. 147. Dai X et al. Prostate cancer-associated SPOP mutations confer resistance to BET inhibitors through stabilization of BRD4. Nature Medicine. 2017;23:1063-1071
  148. 148. Zhang P et al. Intrinsic BET inhibitor resistance in SPOP-mutated prostate cancer is mediated by BET protein stabilization and AKT–mTORC1 activation. Nature Medicine. 2017;23:1055-1062
  149. 149. Janouskova H et al. Opposing effects of cancer-type-specific SPOP mutants on BET protein degradation and sensitivity to BET inhibitors. Nature Medicine. 2017;23:1046-1054
  150. 150. Geng C et al. Prostate cancer-associated mutations in speckle-type POZ protein (SPOP) regulate steroid receptor coactivator 3 protein turnover. Proceedings of the National Academy of Sciences. 2013;110:6997-7002
  151. 151. Wang S et al. SAR405838: An optimized inhibitor of MDM2-p53 interaction that induces complete and durable tumor regression. Cancer Research. 2014;74:5855-5865
  152. 152. Gan W et al. SPOP promotes ubiquitination and degradation of the ERG oncoprotein to suppress prostate cancer progression. Molecular Cell. 2015;59:917-930
  153. 153. Boysen G et al. SPOP mutation leads to genomic instability in prostate cancer. eLife. 2015;4:1-18. e09207
  154. 154. An J et al. Truncated ERG oncoproteins from TMPRSS2-ERG fusions are resistant to SPOP-mediated proteasome degradation. Molecular Cell. 2015;59:904-916
  155. 155. Zhao Y, Sun Y. Cullin-RING ligases as attractive anti-cancer targets. Current Pharmaceutical Design. 2013;19:3215-3225
  156. 156. Vassilev LT et al. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science (80-). 2004;303:844-848
  157. 157. Carvajal LA et al. Dual inhibition of MDMX and MDM2 as a therapeutic strategy in leukemia. Science Translational Medicine. 2018;10:eaao3003
  158. 158. Weisberg E et al. Smac mimetics: Implications for enhancement of targeted therapies in leukemia. Leukemia. 2010;24:2100-2109
  159. 159. Nikolovska-Coleska Z et al. Discovery of embelin as a cell-permeable, small-molecular weight inhibitor of XIAP through structure-based computational screening of a traditional herbal medicine three-dimensional structure database. Journal of Medicinal Chemistry. 2004;47:2430-2440
  160. 160. Guo Z-Q et al. Small-molecule targeting of E3 ligase adaptor SPOP in kidney cancer. Cancer Cell. 2016;30:474-484
  161. 161. Yamagishi Y et al. Natural product-like macrocyclic N-methyl-peptide inhibitors against a ubiquitin ligase uncovered from a ribosome-expressed de novo library. Chemistry & Biology. 2011;18:1562-1570
  162. 162. Ganoth D et al. The cell-cycle regulatory protein Cks1 is required for SCFSkp2-mediated ubiquitinylation of p27. Nature Cell Biology. 2001;3:321-324
  163. 163. Pavlides SC et al. Inhibitors of SCF-Skp2/Cks1 E3 ligase block estrogen-induced growth stimulation and degradation of nuclear p27 kip1 : Therapeutic potential for endometrial cancer. Endocrinology. 2013;154:4030-4045
  164. 164. Ungermannova D et al. High-throughput screening AlphaScreen assay for identification of small-molecule inhibitors of ubiquitin E3 ligase SCFSkp2−Cks1. Journal of Biomolecular Screening. 2013;18:910-920
  165. 165. Chen Q et al. Targeting the p27 E3 ligase SCF(Skp2) results in p27- and Skp2-mediated cell-cycle arrest and activation of autophagy. Blood. 2008;111:4690-4699
  166. 166. Beaudenon S, Huibregtse JM. HPV E6, E6AP and cervical cancer. BMC Biochemistry. 2008;9:S4
  167. 167. Wade M, Li Y-C, Wahl GM. MDM2, MDMX and p53 in oncogenesis and cancer therapy. Nature Reviews. Cancer. 2013;13:83-96
  168. 168. Nakayama KI, Nakayama K. Ubiquitin ligases: Cell-cycle control and cancer. Nature Reviews. Cancer. 2006;6:369-381
  169. 169. Lau AW, Fukushima H, Wei W. The Fbw7 and betaTRCP E3 ubiquitin ligases and their roles in tumorigenesis. Frontiers in Bioscience (Landmark Ed.). 2012;17:2197-2212
  170. 170. Manchado E et al. Targeting mitotic exit leads to tumor regression In vivo: Modulation by Cdk1, Mastl, and the PP2A/B55α,δ phosphatase. Cancer Cell. 2010;18:641-654
  171. 171. Wäsch R, Robbins JA, Cross FR. The emerging role of APC/CCdh1 in controlling differentiation, genomic stability and tumor suppression. Oncogene. 2010;29:1-10
  172. 172. Xu L, Lin D, Yin D, Koeffler HP. An emerging role of PARK2 in cancer. Journal of Molecular Medicine. 2014;92:31-42
  173. 173. Baleja JD et al. Identification of inhibitors to papillomavirus typ. 16 E6 protein based on three-dimensional structures of interacting proteins. Antiviral Research. 2006;72:49-59
  174. 174. Vu B et al. Discovery of RG7112: A small-molecule MDM2 inhibitor in clinical development. ACS Medicinal Chemistry Letters. 2013;4:466-469
  175. 175. Ding Q et al. Discovery of RG7388, a potent and selective p53–MDM2 inhibitor in clinical development. Journal of Medicinal Chemistry. 2013;56:5979-5983
  176. 176. Zhao Y et al. A potent small-molecule inhibitor of the MDM2–p53 interaction (MI-888) achieved complete and durable tumor regression in mice. Journal of Medicinal Chemistry. 2013;56:5553-5561
  177. 177. Sun D et al. Discovery of AMG 232, a potent, selective, and orally bioavailable MDM2–p53 inhibitor in clinical development. Journal of Medicinal Chemistry. 2014;57:1454-1472
  178. 178. Jeay S et al. A distinct p53 target gene set predicts for response to the selective p53-HDM2 inhibitor NVP-CGM097. eLife. 2015;4:1-23. e06498
  179. 179. Furet P et al. Discovery of a novel class of highly potent inhibitors of the p53–MDM2 interaction by structure-based design starting from a conformational argument. Bioorganic & Medicinal Chemistry Letters. 2016;26:4837-4841
  180. 180. Yuan Y, Liao Y-M, Hsueh C-T, Mirshahidi HR. Novel targeted therapeutics: Inhibitors of MDM2, ALK and PARP. Journal of Hematology & Oncology. 2011;4:16
  181. 181. Chan C-H et al. Pharmacological inactivation of Skp2 SCF ubiquitin ligase restricts cancer stem cell traits and cancer progression. Cell. 2013;154:556-568
  182. 182. Wu L et al. Specific small molecule inhibitors of Skp2-mediated p27 degradation. Chemistry & Biology. 2012;19:1515-1524
  183. 183. Huang H-L et al. Triggering Fbw7-mediated proteasomal degradation of c-Myc by oridonin induces cell growth inhibition and apoptosis. Molecular Cancer Therapeutics. 2012;11:1155-1165
  184. 184. Blees JS et al. Erioflorin stabilizes the tumor suppressor Pdcd4 by inhibiting its interaction with the E3-ligase β-TrCP1. PLoS One. 2012;7:e46567
  185. 185. Nakajima H, Fujiwara H, Furuichi Y, Tanaka K, Shimbara N. A novel small-molecule inhibitor of NF-κB signaling. Biochemical and Biophysical Research Communications. 2008;368:1007-1013
  186. 186. Sackton KL et al. Synergistic blockade of mitotic exit by two chemical inhibitors of the APC/C. Nature. 2014;514:646-649
  187. 187. Zeng X et al. Pharmacologic inhibition of the anaphase-promoting complex induces a spindle checkpoint-dependent mitotic arrest in the absence of spindle damage. Cancer Cell. 2010;18:382-395
  188. 188. Sharma SK, Straub C, Zawel L. Development of peptidomimetics targeting IAPs. International Journal of Peptide Research and Therapeutics. 2006;12:21-32
  189. 189. McManus DC et al. Loss of XIAP protein expression by RNAi and antisense approaches sensitizes cancer cells to functionally diverse chemotherapeutics. Oncogene. 2004;23:8105-8117

Written By

Joseph Y. Ong and Jorge Z. Torres

Submitted: 26 June 2018 Reviewed: 04 December 2018 Published: 15 February 2019