Next Article in Journal
Association of Polygenetic Risk Scores Related to Cell Differentiation and Inflammation with Thyroid Cancer Risk and Genetic Interaction with Dietary Intake
Next Article in Special Issue
The Ubiquitin Proteasome System in Genome Stability and Cancer
Previous Article in Journal
Predicting the Risk of Metastases by PSMA-PET/CT—Evaluation of 335 Men with Treatment-Naïve Prostate Carcinoma
Previous Article in Special Issue
HERC1 Regulates Breast Cancer Cells Migration and Invasion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dysregulation of the Ubiquitin Proteasome System in Human Malignancies: A Window for Therapeutic Intervention

Cancer Science Institute of Singapore, National University of Singapore, Singapore 117599, Singapore
*
Author to whom correspondence should be addressed.
Cancers 2021, 13(7), 1513; https://doi.org/10.3390/cancers13071513
Submission received: 19 February 2021 / Revised: 22 March 2021 / Accepted: 22 March 2021 / Published: 25 March 2021
(This article belongs to the Special Issue The Role of the Ubiquitin-Proteasome-System in Human Cancer)

Abstract

:

Simple Summary

The ubiquitin proteasome system (UPS) governs the non-lysosomal degradation of oxidized, damaged, or misfolded proteins in eukaryotic cells. Dysregulation of the UPS results in loss of ability to maintain protein quality through proteolysis, and is closely related to the development of various malignancies and tumorigenesis. Here, we provide a comprehensive general overview on the regulation and roles of UPS and discuss the mechanisms linking dysregulated UPS to human malignancies. Inhibitors developed against components of the UPS, which include U.S. Food and Drug Administration FDA-approved and those currently undergoing clinical trials, are also presented in this review.

Abstract

The ubiquitin proteasome system (UPS) governs the non-lysosomal degradation of oxidized, damaged, or misfolded proteins in eukaryotic cells. This process is tightly regulated through the activation and transfer of polyubiquitin chains to target proteins which are then recognized and degraded by the 26S proteasome complex. The role of UPS is crucial in regulating protein levels through degradation to maintain fundamental cellular processes such as growth, division, signal transduction, and stress response. Dysregulation of the UPS, resulting in loss of ability to maintain protein quality through proteolysis, is closely related to the development of various malignancies and tumorigenesis. Here, we provide a comprehensive general overview on the regulation and roles of UPS and discuss functional links of dysregulated UPS in human malignancies. Inhibitors developed against components of the UPS, which include U.S. Food and Drug Administration FDA-approved and those currently undergoing clinical trials, are also presented in this review.

1. The Ubiquitin Proteasome System

The ubiquitin proteasome system (UPS) is essential for the regulation of protein homeostasis and control of eukaryotic cellular processes including cell cycle progression, stress response, signal transduction, and transcriptional activation [1,2]. UPS controls the degradation of approximately 80% of intracellular proteins which are oxidized, damaged, or misfolded in eukaryotic cells [3]. Though the UPS and autophagy are both important systems of degradation of proteins, the sizes of substrates critically influence the choice of degradation pathway [4]. The UPS typically degrades single unfolded polypeptides, whereas autophagy deals with larger cytosolic complexes, cellular aggregates, and organelles.
Degradation of targeted proteins involves a tightly coordinated process where ubiquitin is covalently attached to the substrate protein through the sequential action of three enzymes. Ubiquitin is a small protein comprising 76 amino acids found in all eukaryotic cells [5]. The energy derived from ATP hydrolysis initiates the activation of ubiquitin activating enzyme (E1) allowing the formation of thioester bond between E1 and ubiquitin. This is followed by transfer of ubiquitin from E1 to ubiquitin-conjugating enzyme (E2), forming a thioester bond similar to that of E1. The third final step involves the covalent attachment of ubiquitin to lysine residues of target protein, catalyzed by ubiquitin ligase (E3) [6]. The 26S proteasome complex comprises a core 20S proteasome and one or two units of the regulatory 19S proteasome (Figure 1). Once a target protein has been modified with a polyubiquitin chain, it is recognized by the 19S proteosome which removes the polyubiquitin chain and the protein is then unfolded and translocated into the 20S proteasome where it is degraded into short peptides [7]. While polyubiquitination has been associated with protein clearance through proteasomal degradation, mono-ubiquitination, which involves the addition of a single ubiquitin moiety to the substrate protein, is shown to affect a range of cellular processes including kinase activity, epigenetic regulation, protein translocation, and DNA damage signaling [8,9].
Ubiquitin contains seven important lysine residues which can be ubiquitinated (K6, K11, K27, K33, K48, and K63) and can form polyubiquitin chains. The two best characterized ubiquitin linkages are K48 and K63 where K48 polyubiquitination targets proteins for degradation by the 26S proteasome complex [10] and K63 participates in DNA damage signaling and recruits DNA repair proteins to damage sites [11]. Protein ubiquitination can be reversed through the removal of ubiquitin from target proteins by deubiquitinating enzymes (DUBs), and this rescues protein destined for degradation. DUBs have also been implicated in the maturation, recycling, and editing of ubiquitin [12,13,14]. Further, chain configuration and linkage can endow ubiquitin with additional roles through the formation of more complex topologies with unknown activities [15]. Dysregulation or abnormal UPS function is frequently seen in various human malignancies and this identifies the aberrant components of the UPS as potential drug targets [16,17]. This review endeavors to present recent literature on the functional roles of UPS in human cancers. We cover how the dysregulation of UPS components may function either as an oncogene or tumor suppressor and affects cellular signaling in tumors. Further, we present current small inhibitors against the UPS and highlight issues that has severely restricted its development.
Increasing evidence demonstrate ubiquitin enzymes are important in carcinogenesis. Though there are numerous cancer-related studies on these enzymes, a large majority primarily focuses on E3 ligases. Studies on E1-activating enzymes have been largely used to identify potential targets in UPS inhibition in cancer while studies on E2-conjugated enzymes revealed their involvement in cell cycle progression, DNA repair, and regulation of oncogenic signaling pathways during tumorigenesis [18,19]. Further E2 enzymes are often found upregulated and highly correlated with poor prognosis in various malignancies including the pancreas, lung, breast, skin, and thyroid [20]. Currently, eight E1s, > 40 E2s, and > 600 E3s have been identified in the human proteome [21].

2. Ubiquitin-Conjugating Enzymes

The ubiquitin-conjugating E2 family comprises > 40 members, and modulates protein stability and ubiquitination through the conjugation of ubiquitin to target proteins [22]. E2-conjugating enzymes are also found dysregulated in cancers and reported to be potent mediators contributing towards multiple tumorigenic processes including migration/invasion, proliferation, drug resistance, radiation resistance, cell cycle, apoptosis, and stimulation of oncogenic pathways (Table 1). We present selected E2 members with roles in cancer below.

2.1. UBE2C

Ubiquitin-conjugating enzyme E2 C (UBE2C) is expressed at critical points during cell cycle progression, which strongly relates to its role in cell cycle regulation [23]. UBE2C overexpression is associated with various malignancies, and shown to play a role in cancer cell growth and invasive properties [24,25]. UBE2C overexpression has been correlated with poor prognosis in cancers of the stomach, colorectal, thyroid, breast, and pancreas [23,25,26,27,28]. UBE2C promotes drug resistance in breast cancer, as the UBE2C blockade enhances tumor cell sensitivity to doxorubicin [29]. This observation suggests that UBE2C expression can serve as a marker for doxorubicin sensitivity in breast cancer patients. Similarly, silencing of UBE2C enhances sensitivity to cytotoxic drugs in hepatocellular cancer [30]. Upregulated UBE2C plays an oncogenic role in human intestinal-type gastric cancer, where UBE2C silencing in these cells disrupts the cell cycle at the G2/M stage. In gastric cancer, UBE2C overexpression induces chromosomal instability and perturbs the cell cycle [23]. In non-small cell lung cancer (NSCLC), elevated UBE2C expression has been reported in cisplatin-resistant NSCLC cells, which correlates with high proliferation and invasion. Further, UBE2C silencing in high UBE2C expressing NSCLC cells induced the upregulation of E-cadherin with concomitant downregulation of vimentin. UBE2C can also bind to the 5′-UTR of drug resistance-related genes such as ERCC1 and ABCG2. Collectively, these data suggest that targeting UBE2C can suppress cisplatin resistance and reverse epithelial to mesenchymal transition (EMT) in NSCLC [31].
Table 1. Summary of the functions of E2 and E3 enzymes in human cancers described in this review.
Table 1. Summary of the functions of E2 and E3 enzymes in human cancers described in this review.
FamilyNameRoleCancer TypeFunctionTest ModelReference
E2UBE2COncogeneGastricChromosomal stability, Proliferation, Migration, InvasionIn vitro, In vivo[23]
OncogeneColonCell cycle, ProloferationIn vitro[24]
OncogeneColorectalProliferation, InvasionIn vitro[25]
OncogeneThyroidProliferation In vitro[26]
OncogeneBreastProliferation, Drug resistance, Radiation resistanceIn vitro[29]
OncogeneLiverProliferation, Drug resistance, Migration, InvasionIn vitro[30]
OncogeneNon-small cell lungDrug resistanceIn vitro[31]
UBE2Q1OncogeneColorectalProliferation [32]
OncogeneLiverp53 signaling, Cell cycleIn vitro[33]
OncogeneBreastp53 signalingIn vitro[34]
UBE2SOncogeneEndometrialSOX6/β-catenin signaling, ProliferationIn vitro[35]
OncogeneLung adenocarcinomaProliferation, p53 signaling, ApoptosisIn vitro[36]
OncogeneLiverp53 signaling, Cell cycleIn vitro[37]
E3FBW7Tumor suppressorBurkitt’s lymphomac-Myc signalingIn vitro[38,39]
Tumor suppressorChronic myelogenous leukemiac-Myc signalingIn vitro, In vivo[40]
LipogenesisLung, Melanoma, Thyroid, CervicalmTORC2/SREBP1 signalingIn vitro[41]
Tumor suppressorT cell leukemiaNotch signalingIn vitro, In vivo[42]
Tumor suppressorColorectalc-Myc signaling, Cell cycleIn vitro[43]
Tumor suppressorEsophageal squamous cellc-Myc signalingIn vitro[44]
Tumor suppressorColorectal, Cervical, Ovarian, Non-small cell lungApoptosis (via Mcl1)In vitro[45]
MDM2OncogeneNeuroblastomap53 signalingIn vitro, In vivo[46]
OncogeneCervicalCell cycle, ApoptosisIn vitro[47]
OncogeneLiverMetastasis, Drug responseIn vitro, In vivo[48]
Cdc20OncogeneBreastMetastasis, Drug responseIn vitro[49]
Cdh1Tumor suppressorBreastSrc signalingIn vitro[50]
β-TRCPTumor suppressorBreast, ProstateMTSS1 signalingIn vitro[51]
OncogeneLungFOXN2In vitro, In vivo[52]
Tumor suppressorPapillary thyroidVEGFR2 signalingIn vitro, In vivo[53]
E6APOncogeneProstateRadiation responseIn vitro[54]
OncogeneProstatep27 signalingIn vitro, In vivo[55]
OncogeneProstateMetastasisIn vitro, In vivo[56]

2.2. UBE2Q1

Ubiquitin-conjugating enzyme E2 Q1 (UBE2Q1) dysregulation has been observed in several cancers including acute lymphocytic leukemia (ALL), breast, hepatocellular, and colorectal [32]. Higher UBE2Q1 levels have been detected in hepatocellular carcinoma (HCC) tumors compared to adjacent normal tissues. Further, elevated UBE2Q1 is significantly correlated with advanced stages and confers poor prognosis in HCC. Silencing of UBE2Q1 by siRNAs reduces proliferation, induces cell cycle arrest, and upregulates tumor suppressor protein 53 (p53) and cyclin-dependent kinase inhibitor 1 (p21) in HCC cells [33]. In breast cancer cells, UBE2Q1 overexpression downregulates p53 and increases cell resistance to apoptosis, indicating that UBE2Q1 offers a protective role in tumor cells through ubiquitination and proteasome degradation of p53 [34].

2.3. UBE2S

Ubiquitin-conjugating enzyme E2 S (UBE2S) overexpression enhances migration and growth of endometrial cancer cells, while silencing it can reverse these effects. Upregulation of UBE2S in these cells can induce nuclear translocation of β-Catenin through SOX6 suppression, and cyclin D1 and c-Myc up regulation [35]. UBE2S increases β-catenin stability by inhibiting its degradation. The β-catenin blockade inhibits UBE2S-induced cancer cell expansion, suggesting that components of the UBE2S-SOX6/β-catenin axis are possible therapeutic targets of endometrial cancer. Overexpression of UBE2S has also been observed in lung cancer cell lines where UBE2S silencing reduces colony formation and increased apoptosis [36]. UBE2S is also reported to be a prognostic factor of hepatocellular carcinoma, and targeting UBE2S in high UBE2S-expressing cells restricts the proliferation and migration of tumor cells where increased expression of p53, p21, and cyclin D1 is seen after UBE2S suppression [37].

3. Ubiquitin Ligases

E3 ubiquitin ligases are a large family of enzymes that promote ubiquitin transfer to proteins or polyubiquitin chains [21]. E3 ligases play an important role in the ubiquitin-mediated proteolytic cascade and are classified into four main classes, according to their domain structure and substrate recognition. The four E3 classes are the homologous to the E3 ubiquitin ligase E6-associated protein (E6AP) C-terminus (HECT), really interesting new gene (RING)-finger, U-box, and plant homeodomain (PHD)-finger. Depending on the substrate targets, E3 ligases can function either as a tumor suppressor or oncogene and can participate in various cellular processes including cell cycle, apoptosis, drug response, metastasis, radiation response, and oncogenic signaling (Table 1). We present selected E3 ligases with roles in cancer in the section below.

3.1. FBW7

F-box and WD repeat domain-containing 7 (FBW7) is a substrate recognition component of the Skp, Cullin, F-box (SCF) complex that regulates multiple pro-oncogenic pathways including c-Myc, Cyclin E, mTOR, and Notch [57,58,59]. FBW7 regulates the proteasomal degradation of c-Myc through phosphorylation of Thr58 and Ser62 residues, and mutation occurring at the Thr58 residue can promote Burkitt’s lymphoma progression [38,39]. In chronic myelogenous leukemia (CML), FBW7 deletion enhances c-Myc expression and can induce p53-dependent apoptosis in human leukemia-initiating cells [40]. FBW7 is shown to negatively regulate lipogenesis where the inhibition of FBW7 suppressor, mTORC2, stabilizes sterol regulatory element-binding protein 1 (SREBP1). FBW7 expression is shown to decelerate tumor progression in lung, thyroid, melanoma, and cervical cancer [41]. Further, tumor suppressor’s function of FBW7 is demonstrated where it suppresses Notch signaling and inhibits adult T-cell leukemia lymphoma progression [42]. More importantly, reduced FBW7 has been reported in various malignancies including hepatocellular, colorectal, and esophageal squamous cell carcinoma [43,44,60]. Though numerous studies have shown a protective role of FBW7 in cancer, there is controversy if FBW7 inhibition promotes or inhibits cancer as its impairment in cancer cells can induce chemoresistance by stabilizing oncogenic proteins [45]. This controversy can be attributed to the presence of FBW7 mutations which occur in ~10% of human cancers and the protein substrate which FBW7 is associated with. Currently, more than 20 substrates of FBW7 have been identified [61,62].

3.2. MDM2

Under normal physiological conditions, tumor suppressor protein 53 (p53) phosphorylation restrains mouse double minute 2 (MDM2)–p53 binding and this induces p53 overexpression in cells. In tumor cells, MDM2 can polyubiquitinate p53, leading to its proteasomal degradation, which is a major pathway altered in cancers [63,64]. MDM2 upregulation has been reported in various cancers, which renders it an attractive drug target for cancer therapy [46,47,48,65]. However, p53 mutations are frequently observed in tumors. These mutants are shown to bind to MDM2 with higher avidity than their wild-type counterpart, thus rendering them resistant to MDM2 inhibitors [66,67]. More importantly, MDM2 has been shown to be functionally linked with the regulation of cancer metabolism in animal models. In the mouse model, p53-independent roles of MDM2 have been reported where they mediate serine/glycine metabolism, which is an important oncogenic signaling in MDM2 overexpressing tumors [68].

3.3. Cdc20 and Cdh1

The anaphase promoting complex/cyclosome (APC/C) has two co-activators: Cdc20 and Cdh1. Cdc20 mRNA levels are elevated and are positively correlated with tumor size in breast cancer patients, suggesting its prognostic value [69]. Further, Cdc20 inhibition prevents breast cancer cell line migration and Cdc20 overexpression accelerates metastases of cancer cells in vitro [49]. Cdh1 is found dysregulated in breast cancer and melanoma, and triggers genome instability and weakens DNA damage response [50]. Cdh1 is also shown to delay G1 to S transition in cell cycle regulation, suggesting its role in tumorigenesis [70].

3.4. βTrCP

Beta-transducin repeats-containing protein (βTrCP) is overexpressed in breast and prostate cancer and the underlying molecular mechanism is that βTrCP promotes tumorigenesis by promoting the degradation of metastasis suppressor 1 (MTSS1) protein by UPS. Blocking MTSS1 degradation is thus a potential treatment for aggressive breast and prostate cancer [51]. βTrCP regulates the Forkhead box transcription factor (FOXN2) ubiquitination, and affects cell proliferation and radiosensitivity in lung cancer [52]. Further, βTRCP is shown to stimulate vascular endothelial growth factor receptor 2 (VEGFR2) ubiquitination and degradation, which inhibit angiogenesis and migration of papillary thyroid cancer cells [53]. These lines of evidence indicate that roles of βTRCP in cancer regulation are context-dependent.

3.5. E6AP

E6AP ubiquitin-protein ligase (E6AP) is recruited by the human papilloma virus E6 protein to promote p53 degradation, which is a key step in human papilloma virus-induced cancers [71]. E6AP is shown to slow down the growth of prostate cancer cell lines, and induces cell senescence in vivo. E6AP knockdown is also shown to increase sensitivity of cells to radiation-induced death [54]. In a separate study, E6AP was reported to inhibit the cell cycle regulator protein 27 (p27), acting through the E2F1-dependent pathway in prostate cancer [55]. Further, E6AP can suppress the metastasis suppressor n-Myc downstream regulated 1 (NDRG1) in mesenchymal phenotypes of prostate cancer where stabilization of NDRG1, via pharmacological inhibition of E6AP, has been proposed for prostate cancer therapy [56].

4. Deubiquitinases (DUBs)

DUBs rescue ubiquitin chains before the substrate protein is degraded by the 26S proteasome. Unbound ubiquitin chains are then processed by other DUBs to restore the free cellular ubiquitin pool. DUB genes are classified into two main groups– cysteine proteases and zinc metalloproteases [13]. DUBs can be categorized into five subclasses: Ubiquitin-specific proteases (USPs), ovarian tumor proteases (OTUs), Machado–Joseph domain proteases (MJDs), ubiquitin C-terminal hydrolases (UCHs), and Jab1/Mov34/Mpr1 Pad1 N-terminal+ (MPN+) (JAMM) domain metalloproteases [72]. Post-translational modification of proteins by ubiquitination affects the targeted protein in several ways: (i) It determines if the ubiquitinated proteins remain in the cell or to be degraded by the proteasome, (ii) it affects the cellular localization of proteins, (iii) it determines protein activity or inactivity, and (iv) it facilitates protein–protein interactions [73,74,75].
Currently, about >100 DUB genes have been identified where the biological functions for the majority are still unknown [13,14]. USPs are, by far, the largest class of DUBs, comprising ~60 human proteases with most containing several domains apart from the catalytic domain. Conserved sequences among these proteases are restricted to the catalytic domain which is designated by the catalytic motif containing Cys, His, and Asp (or Asn) residues. Conserved catalytic domains are thought to be important for substrate specificity, catalytic activity regulation, and mediating protein–protein interaction to each USP. Dysregulated DUBs in cancer are hence potential drug targets. The challenge in the development of drugs is the difficulty in designing a specific inhibitor for a single DUB. 3D crystallography structures of catalytic domains of USP2, USP7, USP8, and USP14 reveal a remarkable structural conservation of their active site shared among these enzymes, thus providing evidence that the development of inhibitors may prove to be challenging [76,77,78,79]. Further, the crystal structures show that the catalytic domains are in inactive conformation prior ubiquitin binding suggesting an alternative target for intervention. Apart from deubiquitination, DUBs have been shown to modulate cellular processes in human malignancies including DNA damage response, oncogenic signaling cascades, drug resistance, apoptosis, cell cycle, immunomodulation, and invasion/migration (Table 2). We will discuss several selected DUBs with roles in cancer below.
Table 2. Summary of the functions of DUB enzymes described in this review.
Table 2. Summary of the functions of DUB enzymes described in this review.
NameRoleCancer TypeFunctionTest ModelReference
BAP1Tumor suppressorLung, Osteosarcoma, ColonDNA double-strand repairIn vitro[80,81,82]
Tumor suppressorRenalFerroptosis signalingIn vitro[83]
USP7OncogeneCervicalSelf-renewal; Foxp3 signalingIn vitro[84]
OncogeneNon-small cell lungImmune Response; Foxp3 signalingIn vitro[85]
OncogeneLungp53 signalingIn vitro, in vivo[86]
USP22OncogeneLungCell CycleIn vitro[87]
OncogeneLung adenocarcinomaEGFR-TKI resistanceIn vitro, in vivo[88]
OncogeneColonCCNB1 signalingIn vitro, in vivo[89]
OncogeneGlioblastomaKDM1A signalingIn vitro, in vivo[90]
UCHL1OncogeneBreastDrug resistance; Invasion/migrationIn vitro[91]
Ataxin 3OncogeneBreast, Osteosarcoma, Cervical, ColorectalDNA In vitro[92]
OncogeneTesticularmTOR/Akt signalingIn vitro[93]
PSMD11OncogeneCervical. OsteosarcomaDNA damage responseIn vitro[94]
OncogeneLung, Prostate, Colorectal, Breast, CervixCell cycleIn vitro[95]
OncogeneLiverE2F1 signalingIn vitro, in vivo[96]
A20Tumor suppressorColorectalApoptosis signalingIn vitro[97]
Tumor suppressorDiffuse large B-cell lymphomaNF-κβ signalingIn vitro[98]
Tumor suppressorSarcomaNF-κβ signalingIn vitro[99]

4.1. BAP1

BRCA1-associated protein 1 (BAP1), a ubiquitin carboxy-terminal hydrolase (UCH), was identified in 1998 as a nuclear protein that bound to the RING-finger domain of the BRCA1 protein [100]. Subsequent studies proposed that BAP1 influences BRCA1 activity through direct binding to BRCA1-associated RING domain 1 (BARD1) and disrupts formation of the BRCA1–BARD1 complex, thus inhibiting BRCA1′s E3 ligase function [101]. However, a recent interactome study based on an affinity purification mass spectrometry approach showed that neither BRCA1 nor BARD1 was found associated with BAP1 [102]. Therefore, the association of BAP1 with BRCA1 and BARD1 is currently unclear. The tumor suppressor functions of BAP1 have been linked to its dual activity in the nucleus, where it has roles in DNA repair [80,81] and transcription [82,103,104,105], and regulation of cell death [83,106] and mitochondrial metabolism [107,108] in the cytoplasm. Being a tumor suppressor, loss of BAP1 through germline and somatic mutations has been directly linked to the predisposition of BAP1-mutated individuals to various malignancies including mesothelioma [109,110,111,112,113,114], uveal melanoma [115,116], cutaneous melanoma [117], and clear cell renal cell carcinoma [118,119,120]. Importantly, inherited BAP1 mutations in the form of truncating mutations have been detected in families with members diagnosed with mesothelioma, uveal melanoma, or breast cancer, suggesting a BAP1 familial cancer syndrome due to germline BAP1 mutations [109,121,122,123].

4.2. USP7

Identified in 1997, ubiquitin-specific protease 7 (USP7) was first characterized to interact with herpes simplex virus type 1 (HSV-1) immediate-early protein (Vmw110) of the HSV-1 regulatory protein and other viral proteins [124,125]. USP7 is the most widely studied DUB and considered to be an oncogene as it promotes tumor growth and negatively affects patient’s immune response to tumors [126,127]. Identified substrates and binding partners of USP7 are found to play crucial roles in immune response and tumor suppression, among others in a variety of human malignancies, which are discussed below.
In immune response, emerging data have shown on a functional link between USP7 and T regulatory cells (Treg) functions. Treg cells express the forkhead transcription factor Foxp3, which is important for Treg cell development [128]. Foxp3 is also shown to possess oncogenic functions including self-renewal and immunomodulation properties in tumor cells [84,129]. Foxp3 can be ubiquitinated at five different lysine residues (K249, K251, K263, K267, and K393) and can be stabilized by USP7 mediated deubiquitination, resulting in Treg number and function maintenance [130]. Aberrant USP7 overexpression reduces Foxp3 polyubiquitination and prevents its degradation. This, in turn, suppresses Treg function and promotes tumor growth [128]. Subsequent studies reveal that Foxp3 can be ubiquitinated and undergoes degradation by E3 ubiquitin ligase stress inducible protein 1 homology and U-Box containing protein 1 (STUB1). Further, heat shock protein 70 (Hsp70), STUB1, and Foxp3 form a complex, suggesting that these proteins bind and promote Foxp3 ubiquitination [85].
USP7 plays conflicting roles in p53 regulation through various mechanisms. p53 first interacts with the tumor necrosis factor (TNF) receptor-associated factor (TRAF) domain and C-terminal (amino acids 880–1050) of USP7, where, by acting as a tumor suppressor, USP7 then ubiquitinates p53 directly and prevents its degradation. Conversely, the TRAF domain and C-terminal (amino acids 801–1050) of USP7 can interact with MDM2 and increase its stability by removing ubiquitin from MDM2, an E3 ligase of p53, preventing its degradation [131]. MDM2 then ubiquitinates p53 and promotes its proteasomal degradation, thus leading to low p53 levels in tumor cells [86]. MDM2 is also shown to inhibit p53 transcription [64]. Further, disruption of MDM2 and p53 interaction induces p53 stabilization [132]. Crystallography and binding studies reveal that both MDM2 and p53 peptides bind to the same surface groove in USP7 [133].

4.3. USP22

Ubiquitin-specific protease 22 (USP22) is putated to be an oncogene. It is highly expressed in multiple malignancies and has been associated with poor overall survival (OS). Aberrant USP22 expression plays crucial roles in regulating DNA transcription, cell cycle transformation and maintaining genomic stability of tumors [87,134]. Apart from activating oncogenes such as BMI-1 and c-MYC, USP22 can inhibit the expression of tumor suppressors such as TP53 through ubiquitination, thus promoting proliferation of tumors [87]. USP22 is shown to regulate EGFR endocytosis through deubiquitination, resulting in sustained activation of the EGFR-dependent signaling pathway and promoting resistance to EGFR-TKIs in EGFR-mutated lung adenocarcinoma [88]. Further, USP22 is shown to facilitate cell-cycle progression and colorectal tumorigenesis by targeting CCNB1 while in glioblastoma, USP22 promotes tumorigenesis via stabilizing KDM1A [89,90]. Together, the evidence above clearly shows the roles of USP22 in driving tumor growth.

4.4. UCHL1

Ubiquitin C-terminal hydrolase L1 (UCHL1) belongs to the ubiquitin C-terminal hydrolases (UCHs) sub-family. Tissue UCHL1 distribution analysis shows that it is predominantly expressed in neuronal tissue, suggesting a role in the central nervous system [135]. UCHL1 overexpression is observed in some cancers including pancreatic, myeloma, prostate, osteosarcoma, and lung [136,137,138,139,140]. Further, UCHL1 expression positively correlates with chemotherapy resistance [91] and metastasis [141]. It has been suggested that UCHL1 is a tumor antigen capable of triggering a humoral immune response in lung cancer, thus pointing its potential as a biomarker in lung cancer [142].

4.5. Ataxin 3

Ataxin 3 (ATX3) belongs to the Josephin family of DUBs and it is mutated in spinocerebellar ataxia type 3 or Machado–Joseph disease [143]. ATX3 functions as a polyubiquitin chain-editing enzyme that controls protein folding and stability [144]. Apart from the Josephin domain, ATX3 contains other domains like the ubiquitin-interacting motif, suggesting that ATX3 is capable of interacting with two distal ubiquitins on a macromolecule [145]. ATX3 forms a complex with the p97/vasolin-containing protein (VCP) ubiquitin-dependent unfoldase/segregase and interacts with Ring finger protein 8 (RNF8) to form a physical and functional complex to regulate proteasome-dependent homeostasis [92]. RNF8 is rapidly recruited to DNA lesion sites during genotoxic stress where p97-ATX3 stimulates RNF8 removal from the chromatin to provide DNA repair pathway stability and promote cell survival after ionizing radiation (IR). Inactivation of the p97-ATX3 complex is further shown to sensitize cancer cells to IR. In testicular cancer (TC), ATX3 is overexpressed and correlates with tumor stages [93].

4.6. PSMD14

The regulatory particle non-ATPase 11 (PSMD14) is a member of the JAB1/MPN/Mov34 metalloenzyme (JAMM) family of metalloproteases, and is responsible for providing isopeptidase activities to reverse the effects of the ubiquitin-proteasome degradation system. PSMD14 plays a diverse role in various biological processes, including DNA repair, embryonic cell development, cell differentiation, programmed cell death and resistance to drugs [94,146,147]. Further, PSMD14 modulates tumor cell proliferation, through regulating the retinoblastoma protein (Rb) phosphorylation and cyclin-dependent kinases (CDKs) [95]. Upregulation of PSMD14, in hepatocellular carcinoma, has been associated with tumor promotion through E2F transcription factor 1 (E2F1) stabilization and its target genes [96].

4.7. A20

Tumor necrosis factor inducible protein A20 (A20) is a ubiquitin-editing enzyme that removes K63-linked ubiquitin chains from adaptor proteins, such as receptor-interacting serine/threonine-protein kinase 1 protein (RIPK1), and subsequently adds K48-linked polyubiquitin chains to trigger degradation. RIPK1 degradation triggers termination of nuclear factor Kappa light-chain enhancer of activated B (NF-κB) signaling in response to tumor necrosis factor (TNF), and prevents TNF-mediated apoptosis in a cell-dependent manner [148]. Though the cell death mechanism exerted by A20 is unclear, several mechanisms have been proposed. It has been suggested that A20 inhibits TNF-induced cell death by preventing interaction of RIPK1 and tumor necrosis factor receptor type 1-associated death domain protein (TRADD) to TNF receptor 1 (TNFR1), thereby inhibiting subsequent recruitment of FAS-associated death domain protein (FADD) and caspase-8 [149]. In a different study, A20 acts by impeding Cullin 3 (CUL3)-dependent ubiquitination and caspase-8 activation [97]. A20 dysregulation, in the form of promoter methylation and inactivating mutation, has been observed in lymphomas [150,151]. A20 expression can be regulated by specific miRNAs in a cellular context manner. MicroRNAs miR125a and miR125b can suppress A20 expression in diffuse large B-cell lymphoma, thus contributing to tumor cell proliferation through constitutive NF-κB signaling [98]. In contrast, miR29 is shown to protect A20 transcripts from degradation by preventing RNA-binding protein HuR binding and the subsequent recruitment of the RNA degradation complex in sarcoma cells [99].

5. Other Ubiquitin Modifiers

The E1-activating enzymes are a group of proteins also known as ubiquitin-like proteins (UBLs) as they share sequence homology and possess a similar three-dimensional structure to ubiquitin [152]. The eight human E1 enzymes identified are ubiquitin-like modifier activating enzyme 1 (UBA1), neural precursor cell expressed, developmentally downregulated 8 (NEDD8)-activating enzyme (NAE), small ubiquitin-like modifier (SUMO)-activating enzyme (SAE), UBA4, UBA5, UBA6, UBA7, and autophagy-related protein 7 (ATG7) [21,153,154]. E1s can be subdivided into canonical and noncanonical based on their structural and biochemical properties.
Among the eight E1 enzymes members identified, UBA1 has been shown to be most important where loss-of-function exerts the most unfavorable effect on the survival and growth of tumor cells [21,155]. In a genome-wide study to identify genes which are important for cancer survival through RNAi screening, UBA1 and UBA2 (encoding SAE active subunit) genes have been identified to be essential for survival after knockdown in large pan-cancer RNA interference screens in cancer cell lines [156]. Cancer cells can be identified according to stress phenotypes which include oxidative, metabolic, mitotic, and DNA damage stress [157]. Tumor cells rely on normal cellular machineries, such as the UPS system, to cope with these stresses. There is a tendency for tumor cells to rely on these machineries to a greater extent versus normal cells. This provides an avenue of therapeutic opportunities through stress overload and stress sensitization [157,158]. Therapeutic inhibition of E1 enzymes, UBA1 and NAE, induces tumor cell death through unfolded protein response (UPR) and proteotoxic stress in hematologic malignancies [159,160]. Neddylation of Cullin RING ligases (CRLs) by NAE enhances ubiquitination of CRL-bound protein substrates [161,162]. NAE inhibition has been shown to be a promising approach to treat cancers with the anti-E1 inhibitor, MLN4924. MLN4924 prevents CRL neddylation, leading to CRL substrate accumulation and DNA re-replication prior to tumor cell apoptosis [159]. Importantly, MLN4924 has been shown to reduce tumor growth in vivo. A single-nucleotide transition of alanine 171 to threonine of the NAE subunit ubiquitin-activating enzyme (UBA3), however, has been reported to reduce the enzyme’s affinity for MLN4924 and ATP while increasing NEDD8 activation at physiological ATP concentrations, which suggests a potential pitfall of this strategy [163].

6. UPS Inhibitors in Cancer Therapy

The progress in targeting the UPS has been slow and this delay has been attributed to the following reasons. First, most components of the ubiquitin system do not possess a well-defined catalytic pocket to allow binding of small inhibitors. Second, the ubiquitination process relies on the dynamic rearrangement of multiple protein–protein interactions that traditionally have been challenging to disrupt with small molecule inhibitors. Third, components of the UPS are shown to possess both oncogenic and tumor suppressor properties due to the complexity of their regulatory cellular processes. Despite these challenges, components of the UPS have been considered as attractive targets for cancer treatment. In the following sections, we introduce some inhibitors against components of the UPS that have been tested in preclinical and clinical studies (Table 3).

6.1. Targeting the Proteasome

Proteasome inhibitors were first developed for the treatment of cachexia in patients with advanced cancers. Cachexia is a condition characterized by a catabolic state of progressive wasting [69]. It was hypothesized that a proteasome inhibitor could prevent protein degradation and muscle wasting [164]. MG132, a peptide aldehyde, is a well-known potent proteasome inhibitor commonly used in research. Though MG132 has been proven to be a valuable research tool, it has never been tested clinically due to its rapid oxidation [165].
Bortezomib (velcade), a peptide boronate, was the first U.S. Food and Drug Administration (FDA)-approved proteasome inhibitor. It can reversibly form tetrahedral adducts with Thr1 residues in the catalytic β5 subunits of 19S, and exhibits phenomenal potency in vitro. Bortezomib initially showed only unremarkable activity in solid cancer patients in phase I trials. Subsequently, it was reported to promote disease regression in multiple myeloma patients [166] through stabilization of I-κB, thus suppressing NF-κB signaling [167]. Further, bortezomib induces accumulation of two negative regulators of the cell cycle, p27KIP1 and p53 [168]. Additional anticancer properties of bortezomib are that it can induce pro-apoptotic protein BAX and oxidative stress which tilt the balance towards apoptosis [169,170]. However, bortezomib exhibits dose-limiting toxicities and multiple side effects have been reported including peripheral neuropathy, myelosuppression, and cardiotoxicity. Approximately 60% of patients treated with bortezomib developed resistance to this drug, and the FDA approved the use of bortezomib, in combination with chemotherapeutic agents such as doxorubicin and dexamethasone, to overcome resistance in clinical trials [171,172].
Carfilzomib (kyprolis), a tetrapeptide epoxyketone, is the second proteasome inhibitor approved by the FDA [76]. Carfilzomib binds to both hydroxyl and the free α-amino group of Thr1 in the catalytic β5 subunits and blocks the binding of substrate proteins to β5 [173,174,175]. Carfilzomib can irreversibly inhibit proteasomal activity to less than 20%, and has been shown to induce antitumor response in bortezomib-resistant multiple myeloma [175]. It possesses a more selective mechanism of action and exhibits high chemical stability, which is shown by fewer side effects when compared to bortezomib [176,177].
Ixazomib (Ninlaro) was developed as a second-generation proteasome inhibitor and the first one FDA-approved, as an oral administered drug, for multiple myeloma [178]. Similar to the first generation, ixazomib triggers apoptotic cell death in multiple myeloma cell lines and enhances the expression of pro-apoptotic genes [178]. Unlike bortezomib, ixazomib does not affect the mitochondrial serine protease Htr/Omi, which is a neuropathy-related off-target of bortezomib [172]. Other proteasome inhibitors currently undergoing clinical trials include delanzomib, marizomib, and oprozomib (Table 3).

6.2. Targeting Ubiquitinases

MLN4924 inhibits the NEDD8-activating enzyme (NAE) E1 enzyme where it forms an adduct with NAE and prevents Cullin-RING ligases (CRLs) neddylation [159,179]. Blocking CRL neddylation restricts tumor cell growth, resulting in elevated substrate levels, which then triggers cell death through DNA damage response, autophagy, cell-cycle arrest, and apoptosis [180]. MLN4924 has shown clinical activity in a phase I clinical trial of acute myelogenous leukemia (AML) [181]. It is also shown to inhibit angiogenesis during tumor development [182]. TAK981 (SAE) and TAS4464 (NAE) are two other E1 inhibitors currently undergoing clinical trials against various cancers (Table 3).
Table 3. Summary of UPS inhibitors which are FDA-approved and/or tested in clinical trials described in this review.
Table 3. Summary of UPS inhibitors which are FDA-approved and/or tested in clinical trials described in this review.
InhibitorTargetCancer TypeClinical Trial Reference
BortezomibProteasomal inhibitorMultiple myeloma, Mantle cell lymphoma, Leukemia, Neuroblastoma,
Head and Neck, Thyroid, Hepatocellular
FDA approvedwww.clinicaltrials.gov
[126,166,167,169,170,171,172,177]
CarfilzomibProteasomal inhibitorMultiple myeloma, Lymphoma, Relapsed and/or refractory multiple myeloma,
Leukemia, Lung, Thyroid, Refractory renal cell carcinoma
FDA approvedwww.clinicaltrials.gov
[54,173,174,175,176,177]
IxazomibProteasomal inhibitorMultiple myeloma, Relapsed and/or refractory multiple myeloma, Lymphoma,
Leukemia, Breast, Glioblastoma, Renal cell carcinoma, Hodgkin and T cell
lymphoma
FDA approvedwww.clinicaltrials.gov
[178]
DelanzomibProteasomal inhibitorNon-Hodgkin’s lymphomaPhase Iwww.clinicaltrials.gov
MarizomibProteasomal inhibitorMultiple myeloma, Advanced solid tumorsPhase I/IIwww.clinicaltrials.gov
OprozomibProteasomal inhibitorMultiple myeloma, Glioma, Pancreatic, Lung, Melanoma, Lymphoma, GlipblastomaPhase I/II/IIIwww.clinicaltrials.gov
MLN4924NAE and UBA1(E1)Advanced malignant solid tumors, Melanoma, Hepatocellular, B cell lymphoma,
Hematologic malignancies, Acute myelocytic leukemia
Phase I/II/IIIwww.clinicaltrials.gov
TAK981SAE (E1)B cell lymphoma, colorectal, non-Hodgkin’s, Advcnced/metasiatic solid tumorsPhase I/IIwww.clinicaltrials.gov
TAS4464NAE (E1)Multiple myeloma, non-Hodgkin lymphomaPhase I/IIwww.clinicaltrials.gov
SAR-405838MDM2 (E2)Solid tumorsPhase Iwww.clinicaltrials.gov
[172,183]
CGM-097MDM2 (E2)Advanced p53 wildtype solid tumorsPhase I www.clinicaltrials.gov
[184,185]
DS-3032bMDM2 (E2)Acute myelocytic leukemiaPhase I/IIwww.clinicaltrials.gov
[186,187]
Debio1143 (AT-406)cIAP1/2 (E3)Acute myeloid leukemiaPhase I www.clinicaltrials.gov
[188]
LC-161IAP (E3)Advanced solid tumorsPhase I www.clinicaltrials.gov
[189]
BirinapantIAP (E3)Solid tumorsPhase I/IIwww.clinicaltrials.gov
[190]
PimozideUSP1Glioma, Non-small cell lung cancerFDA approced for Tourette’s syndrome; Preclinical[191,192]
MitoxantroneUSP11Metastatic crastrate -resistant prostate, Acute myeloid leukemia, Advanced breast cancer,
non-Hodgkin’s lymphoma, Primary liver
FDA approved[193,194,195,196,197,198,199,200,201]
Nutlin-3a is the first small molecule inhibitor of the oncogene MDM2, a negative regulator of p53, and this prevents interaction between MDM2 and p53 [132]. RO-5503781 (idasanutlin), a Nutlin-3a derivative, is designed and developed as an oral inhibitor for the treatment of various malignancies including glioblastoma, diffuse large B cell lymphoma, multiple myeloma, prostate, and AML in preclinical studies [172,183]. However, RO-5503781 was terminated in phase II clinical trials. Similar to RO-5503781, Nutlin-3a-derivative MK-8242 was terminated in phase I clinical trials for advanced solid tumors and recurrent AML and liposarcomas. Other Nutlin-3a derivatives such as SAR-405838, CGM-097, and DS-3032b are currently in phase I clinical trials for patients with advanced solid tumors and lymphoma, either alone or in combination with chemotherapy [184,185,186,187,202].
The inhibitors of apoptosis (IAP) proteins are ubiquitin E3 ligases, responsible for cell survival and oncogenesis. cIAP1 and cIAP2 paralogs contain three N-terminal baculoviral IAP repeat (BIR) domains and a C-terminal E3 ligase really interesting new gene (RING) domain. IAP antagonist compounds, known as second mitochondria-derived activator of caspase (SMAC) mimetics, bind to BIR domains and activate RING-dependent autoubiquitylation which leads to IAP destruction [203]. Debio1143 (AT-406) is a small molecule SMAC mimetic IAP antagonist that has entered clinical development. Debio1143 is reported to inhibit the cellular inhibitor of apoptosis 1 (cIAP-1) in acute myeloid leukemia where about a third of patients enrolled achieved complete remission within the study period [188]. Elevated plasma TNFα and IL-8 are associated with responders and patients showed good tolerability in combination with chemotherapy. Other SMAC mimetic IAP antagonists such as LCL161 and birinapant have entered phase II trials [189,190,204]. Debio1143 and birinapant can preferentially target cIAP1 and cIAP2 rather than X-linked inhibitor of apoptosis protein (XIAP), while LCL161 is a pan-IAP inhibitor that possesses similar affinity to XIAP, cIAP1, and cIAP2 [205,206,207,208]. Hence, the potency of SMAC mimetics relies on their specificity to IAPs. LCL161 and birinapant are among the most commonly used SMAC mimetics due their therapeutic effectiveness and known mechanism of actions demonstrated in both preclinical and clinical studies.

6.3. Targeting DUBs

Proteasome-associated DUBs are responsible for deubiquitination of target proteins. Many deubiquitinases are dysregulated and their roles in cancer rely on the substrates they deubiquitinate. Although potent irreversible inhibitors of DUBs such as ubiquitin aldehyde or ubiquitin vinyl sulfone have been widely used as research tools, only a few DUB inhibitors have entered clinical trials for certain cancers. The development of selective DUB inhibitors has, however, been hampered by limited knowledge of DUB biology and the pleiotropic properties of small-molecule DUB inhibitors.
Pimozide was developed as an antipsychotic drug and approved for the treatment of Tourette’s syndrome by the FDA. It is a reversible inhibitor of the enzymatic ubiquitin-specific protease 1/ USP1-associated factor 1 (USP1/UAF1) complex activity. The USP1/UAF1 complex plays an important role in the regulation of homologous recombination (HR)-mediated double-strand break (DSB) repair [209]. In cisplatin-resistant NSCLC cells, pimozide inhibits USP1/UAF1, and acts synergistically with cisplatin to inhibit cell proliferation [191]. Further, pimozide is shown to resensitize glioma tumors to radiation therapy [192].
Mitoxantrone (novantrone) was approved by the FDA for use in metastatic castrate-resistant prostate cancer [193]. Mitoxantrone inhibits USP11 catalytic activity and impairs the DNA-damage repair (DDR) system in pancreatic ductal adenocarcinoma (PDA) in vitro [210]. USP 11 is a tumor suppressor where it is shown to interact with ubiquitinated BRCA2 and exerts a prosurvival function in response to cellular DNA damage [211]. Further, it is reported to inhibit transforming growth factor (TGF)-β1-induced phosphorylation of SMAD2/3 and type I TGF-β receptor by destabilizing the type II TGF-β receptor in lung fibroblasts [212]. Mitoxantrone has shown clinical activity in relapsed/refractory acute myeloid leukemia either administered alone or in combination with chemotherapy [194,195]. Other cancer types which show response and are well-tolerated to mitoxantrone include advanced breast cancer, non-Hodgkin’s lymphoma, and primary liver cancer [196,197,198,199,200,201].
P5091 and b-AP15 are two other DUB inhibitors which have shown promise in preclinical studies. P5091 is shown to selectively inhibit USP7 in multiple myeloma both in vitro and in vivo [213]. USP7 plays a crucial role in p53 regulation, and P5091 shows weaker activity after USP7 inhibition, thus demonstrating its specificity. Importantly, P5091 induces apoptotic cell death in patient multiple myeloma cells resistance to bortezomib, lenalidomide, and dexamethasone [213]. Nevertheless, P5091 shares the same limitation as Nutlins as they only stabilize wild-type p53, while the majority of tumors are mutated p53. b-AP15 inhibits both UCH37 and USP14, and induces accumulation of ubiquitinated substrates [214]. B-AP-15 demonstrates outstanding efficacy in solid tumor and acute myeloid leukemia models in vivo.

7. Conclusions

Frequent aberrant UPS activity seen in human malignancies indicates that the proteasome and components of the UPS are attractive therapeutic targets. Targeting the proteasome, in the clinic, has achieved success with FDA-approved proteasome inhibitors such as bortezomib, carfilzomib, and ixazomib. Being the last step in the UPS, the use of proteasome inhibitors has shown undesirable side effects arising from the action of upstream UPS components. This shows that there is an untapped potential for the development of drugs against other components of the UPS. Thus, ubiquitin-activating steps, E2, E3, and DUBs can be exploited for inhibition [215,216]. Unfortunately, most of these inhibitors show good efficacy in culture models but less so in animal models and clinical trials [217,218,219]. Traditionally, the ubiquitin activating steps and degradation possess the greatest potential due to presence of well-defined activity pockets but face issues of substrate specificity. The other UPS components, however, do not possess defined pockets for targeting with small inhibitors. Hence, delay in the development of successful UPS inhibitors can be attributed to the lack of knowledge of target protein structures and identifiable activity pockets for inhibitor binding. Advances in technology such as computer-aided design, mass spectrophotometry, and high throughput screening may aid in the identification of suitable candidates. Further, the occurrence of oncogenic signaling together with aberrant UPS activity may affect the success of future UPS inhibitors. Nevertheless, a greater effort is required to elucidate the functions of aberrant UPS at both preclinical and clinical levels to better understand their roles in human malignancies to develop alternative paradigms for therapeutic intervention.

Author Contributions

C.W.F. and A.A., conceptualization, writing, and editing. A.A., funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by R713-000-216-720.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

In this section, you can acknowledge any support given which is not covered by the author contribution or funding sections. This may include administrative and technical support, or donations in kind (e.g., materials used for experiments).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ciechanover, A. Intracellular protein degradation: From a vague idea thru the lysosome and the ubiquitin-proteasome system and onto human diseases and drug targeting. Best Pract. Res. Clin. Haematol. 2017, 30, 341–355. [Google Scholar] [CrossRef]
  2. Myung, J.; Kim, K.B.; Crews, C.M. The ubiquitin-proteasome pathway and proteasome inhibitors. Med. Res. Rev. 2001, 21, 245–273. [Google Scholar] [CrossRef]
  3. Yang, H.; Chen, X.; Li, K.; Cheaito, H.; Yang, Q.; Wu, G.; Liu, J.; Dou, Q.P. Repurposing old drugs as new inhibitors of the ubiquitin-proteasome pathway for cancer treatment. Semin. Cancer Biol. 2021, 68, 105–122. [Google Scholar] [CrossRef] [PubMed]
  4. Dikic, I. Proteasomal and Autophagic Degradation Systems. Annu. Rev. Biochem. 2017, 86, 193–224. [Google Scholar] [CrossRef] [PubMed]
  5. Nandi, D.; Tahiliani, P.; Kumar, A.; Chandu, D. The ubiquitin-proteasome system. J. Biosci. 2006, 31, 137–155. [Google Scholar] [CrossRef]
  6. Fang, S.; Weissman, A.M. Ubiquitin-proteasome system. Cell. Mol. Life Sci. 2004, 61, 1546–1561. [Google Scholar] [CrossRef]
  7. Adams, J. The proteasome: Structure, function, and role in the cell. Cancer Treat. Rev. 2003, 29 (Suppl. S1), 3–9. [Google Scholar] [CrossRef]
  8. Wang, G.; Gao, Y.; Li, L.; Jin, G.; Cai, Z.; Chao, J.-I.; Lin, H.-K. K63-Linked Ubiquitination in Kinase Activation and Cancer. Front. Oncol. 2012, 2, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Miranda, M.; Sorkin, A. Regulation of Receptors and Transporters by Ubiquitination: New Insights into Surprisingly Similar Mechanisms. Mol. Interv. 2007, 7, 157–167. [Google Scholar] [CrossRef]
  10. Hochstrasser, M. Ubiquitin-Dependent protein degradation. Annu. Rev. Genet. 1996, 30, 405–439. [Google Scholar] [CrossRef] [PubMed]
  11. Panier, S.; Durocher, D. Regulatory ubiquitylation in response to DNA double-strand breaks. DNA Repair 2009, 8, 436–443. [Google Scholar] [CrossRef]
  12. Kim, J.H.; Park, K.C.; Chung, S.S.; Bang, O.; Chung, C.H. Deubiquitinating Enzymes as Cellular Regulators. J. Biochem. 2003, 134, 9–18. [Google Scholar] [CrossRef] [PubMed]
  13. Nijman, S.M.; Luna-Vargas, M.P.; Velds, A.; Brummelkamp, T.R.; Dirac, A.M.; Sixma, T.K.; Bernards, R. A Genomic and Functional Inventory of Deubiquitinating Enzymes. Cell 2005, 123, 773–786. [Google Scholar] [CrossRef] [Green Version]
  14. Clague, M.J.; Barsukov, I.; Coulson, J.M.; Liu, H.; Rigden, D.J.; Urbé, S. Deubiquitylases From Genes to Organism. Physiol. Rev. 2013, 93, 1289–1315. [Google Scholar] [CrossRef]
  15. Komander, D. The emerging complexity of protein ubiquitination. Biochem. Soc. Trans. 2009, 37 Pt 5, 937–953. [Google Scholar] [CrossRef] [Green Version]
  16. Ciechanover, A.; Brundin, P. The Ubiquitin Proteasome System in Neurodegenerative Diseases. Neuron 2003, 40, 427–446. [Google Scholar] [CrossRef] [Green Version]
  17. Ciechanover, A.; Schwartz, A.L. The ubiquitin system: Pathogenesis of human diseases and drug targeting. Biochim. Biophys. Acta 2004, 1695, 3–17. [Google Scholar] [CrossRef] [Green Version]
  18. Yang, Y.; Kitagaki, J.; Dai, R.-M.; Tsai, Y.C.; Lorick, K.L.; Ludwig, R.L.; Pierre, S.A.; Jensen, J.P.; Davydov, I.V.; Oberoi, P.; et al. Inhibitors of Ubiquitin-Activating Enzyme (E1), a New Class of Potential Cancer Therapeutics. Cancer Res. 2007, 67, 9472–9481. [Google Scholar] [CrossRef] [Green Version]
  19. Liu, J.; Shaik, S.; Dai, X.; Wu, Q.; Zhou, X.; Wang, Z.; Wei, W. Targeting the ubiquitin pathway for cancer treatment. Biochim. Biophys. Acta 2015, 1855, 50–60. [Google Scholar] [CrossRef] [Green Version]
  20. Hosseini, S.M.; Okoye, I.; Chaleshtari, M.G.; Hazhirkarzar, B.; Mohamadnejad, J.; Azizi, G.; Hojjat-Farsangi, M.; Mohammadi, H.; Shotorbani, S.S.; Jadidi-Niaragh, F. E2 ubiquitin-conjugating enzymes in cancer: Implications for immunotherapeutic interventions. Clin. Chim. Acta 2019, 498, 126–134. [Google Scholar] [CrossRef]
  21. Clague, M.J.; Heride, C.; Urbé, S. The demographics of the ubiquitin system. Trends Cell Biol. 2015, 25, 417–426. [Google Scholar] [CrossRef] [PubMed]
  22. Stewart, M.D.; Ritterhoff, T.; Klevit, R.E.; Brzovic, P.S. E2 enzymes: More than just middle men. Cell Res. 2016, 26, 423–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, J.; Liu, X.; Yu, G.; Liu, L.; Wang, J.; Chen, X.; Bian, Y.; Ji, Y.; Zhou, X.; Chen, Y.; et al. UBE2C Is a Potential Biomarker of Intestinal-Type Gastric Cancer With Chromosomal Instability. Front. Pharmacol. 2018, 9, 847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Fujita, T.; Ikeda, H.; Taira, N.; Hatoh, S.; Naito, M.; Doihara, H. Overexpression of UbcH10 alternates the cell cycle profile and accelerate the tumor proliferation in colon cancer. BMC Cancer 2009, 9, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chen, S.; Chen, Y.; Hu, C.; Jing, H.; Cao, Y.; Liu, X. Association of clinicopathological features with UbcH10 expression in colorectal cancer. J. Cancer Res. Clin. Oncol. 2009, 136, 419–426. [Google Scholar] [CrossRef] [PubMed]
  26. Pallante, P.; Berlingieri, M.; Troncone, G.; Kruhoffer, M.; Orntoft, T.; Viglietto, G.; Caleo, A.; Migliaccio, I.; Decaussin-Petrucci, M.; Santoro, M.; et al. UbcH10 overexpression may represent a marker of anaplastic thyroid carcinomas. Br. J. Cancer 2005, 93, 464–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Psyrri, A.; Kalogeras, K.T.; Kronenwett, R.; Wirtz, R.M.; Batistatou, A.; Bournakis, E.; Timotheadou, E.; Gogas, H.; Aravantinos, G.; Christodoulou, C.; et al. Prognostic significance of UBE2C mRNA expression in high-risk early breast cancer. A Hellenic Cooperative Oncology Group (HeCOG) Study. Ann. Oncol. 2012, 23, 1422–1427. [Google Scholar] [CrossRef] [PubMed]
  28. Zhao, Z.-K.; Wu, W.-G.; Chen, L.; Dong, P.; Gu, J.; Mu, J.-S.; Yang, J.-H.; Liu, Y.-B. Expression of UbcH10 in pancreatic ductal adenocarcinoma and its correlation with prognosis. Tumor Biol. 2013, 34, 1473–1477. [Google Scholar] [CrossRef]
  29. Rawat, A.; Gopal, G.; Selvaluxmy, G.; Rajkumar, T. Inhibition of ubiquitin conjugating enzyme UBE2C reduces proliferation and sensitizes breast cancer cells to radiation, doxorubicin, tamoxifen and letrozole. Cell. Oncol. 2013, 36, 459–467. [Google Scholar] [CrossRef]
  30. Xiong, Y.; Lu, J.; Fang, Q.; Lu, Y.; Xie, C.; Wu, H.; Yin, Z. UBE2C functions as a potential oncogene by enhancing cell proliferation, migration, invasion, and drug resistance in hepatocellular carcinoma cells. Biosci. Rep. 2019, 39, BSR20182384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Guo, J.; Jin, D.; Wu, Y.; Yang, L.; Du, J.; Gong, K.; Chen, W.; Dai, J.; Miao, S.; Xi, S. The miR 495-UBE2C-ABCG2/ERCC1 axis reverses cisplatin resistance by downregulating drug resistance genes in cisplatin-resistant non-small cell lung cancer cells. EBioMedicine 2018, 35, 204–221. [Google Scholar] [CrossRef]
  32. Fahmidehkar, M.A.; Shafiee, S.M.; Eftekhar, E.; Mahbudi, L.; Seghatoleslam, A. Induction of cell proliferation, clonogenicity and cell accumulation in S phase as a consequence of human UBE2Q1 overexpression. Oncol. Lett. 2016, 12, 2169–2174. [Google Scholar] [CrossRef] [Green Version]
  33. Chang, R.; Wei, L.; Lu, Y.; Cui, X.; Lu, C.; Liu, L.; Jiang, D.; Xiong, Y.; Wang, G.; Wan, C.; et al. Upregulated expression of ubiquitin-conjugating enzyme E2Q1 (UBE2Q1) is associated with enhanced cell proliferation and poor prognosis in human hapatocellular carcinoma. J. Mol. Histol. 2014, 46, 45–56. [Google Scholar] [CrossRef] [PubMed]
  34. Shafiee, S.M.; Rasti, M.; Seghatoleslam, A.; Azimi, T.; Owji, A.A. UBE2Q1 in a Human Breast Carcinoma Cell Line: Overexpression and Interaction with p53. Asian Pac. J. Cancer Prev. 2015, 16, 3723–3727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lin, M.; Lei, T.; Zheng, J.; Chen, S.; Du, L.; Xie, H. UBE2S mediates tumor progression via SOX6/β-Catenin signaling in endometrial cancer. Int. J. Biochem. Cell Biol. 2019, 109, 17–22. [Google Scholar] [CrossRef]
  36. Liu, Z.; Xu, L. UBE2S promotes the proliferation and survival of human lung adenocarcinoma cells. BMB Rep. 2018, 51, 642–647. [Google Scholar] [CrossRef] [PubMed]
  37. Pan, Y.-H.; Yang, M.; Liu, L.-P.; Wu, D.-C.; Li, M.-Y.; Su, S.-G. UBE2S enhances the ubiquitination of p53 and exerts oncogenic activities in hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2018, 503, 895–902. [Google Scholar] [CrossRef]
  38. Bahram, F.; von der Lehr, N.; Cetinkaya, C.; Larsson, L.G. c-Myc hot spot mutations in lymphomas result in inefficient ubiquitination and decreased proteasome-mediated turnover. Blood 2000, 95, 2104–2110. [Google Scholar] [CrossRef] [PubMed]
  39. Yeh, C.-H.; Bellon, M.; Nicot, C. FBXW7: A critical tumor suppressor of human cancers. Mol. Cancer 2018, 17, 115. [Google Scholar] [CrossRef]
  40. Reavie, L.; Buckley, S.M.; Loizou, E.; Takeishi, S.; Aranda-Orgilles, B.; Ndiaye-Lobry, D.; Abdel-Wahab, O.; Ibrahim, S.; Nakayama, K.I.; Aifantis, I. Regulation of c-Myc Ubiquitination Controls Chronic Myelogenous Leukemia Initiation and Progression. Cancer Cell 2013, 23, 362–375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Li, S.; Oh, Y.T.; Yue, P.; Khuri, F.R.; Sun, S.-Y. Inhibition of mTOR complex 2 induces GSK3/FBXW7-dependent degradation of sterol regulatory element-binding protein 1 (SREBP1) and suppresses lipogenesis in cancer cells. Oncogene 2016, 35, 642–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Yeh, C.-H.; Bellon, M.; Pancewicz-Wojtkiewicz, J.; Nicot, C. Oncogenic mutations in the FBXW7 gene of adult T-cell leukemia patients. Proc. Natl. Acad. Sci. USA 2016, 113, 6731–6736. [Google Scholar] [CrossRef] [Green Version]
  43. Iwatsuki, M.; Mimori, K.; Ishii, H.; Yokobori, T.; Takatsuno, Y.; Sato, T.; Toh, H.; Onoyama, I.; Nakayama, K.I.; Baba, H.; et al. Loss of FBXW7, a cell cycle regulating gene, in colorectal cancer: Clinical significance. Int. J. Cancer 2010, 126, 1828–1837. [Google Scholar] [CrossRef]
  44. Mori, M.; Yokobori, T.; Mimori, K.; Iwatsuki, M.; Ishii, H.; Tanaka, F.; Sato, T.; Toh, H.; Sudo, T.; Iwaya, T.; et al. Copy number loss of FBXW7 is related to gene expression and poor prognosis in esophageal squamous cell carcinoma. Int. J. Oncol. 2012, 41, 253–259. [Google Scholar] [CrossRef] [Green Version]
  45. Wertz, I.E.; Kusam, S.; Lam, C.; Okamoto, T.; Sandoval, W.; Anderson, D.J.; Helgason, E.; Ernst, J.A.; Eby, M.; Liu, J.; et al. Sensitivity to antitubulin chemotherapeutics is regulated by MCL1 and FBW7. Nature 2011, 471, 110–114. [Google Scholar] [CrossRef]
  46. Wang, W.; Wang, X.; Rajaei, M.; Youn, J.Y.; Zafar, A.; Deokar, H.; Buolamwini, J.K.; Yang, J.; Foster, J.H.; Zhou, J.; et al. Targeting MDM2 for Neuroblastoma Therapy: In Vitro and In Vivo Anticancer Activity and Mechanism of Action. Cancers 2020, 12, 3651. [Google Scholar] [CrossRef] [PubMed]
  47. Xu, L.; Wang, J.; Yuan, X.; Yang, S.; Xu, X.; Li, K.; He, Y.; Wei, L.; Zhang, J.; Tian, Y. IU1 suppresses proliferation of cervical cancer cells through MDM2 degradation. Int. J. Biol. Sci. 2020, 16, 2951–2963. [Google Scholar] [CrossRef]
  48. Wang, W.; Hu, B.; Qin, J.-J.; Cheng, J.-W.; Li, X.; Rajaei, M.; Fan, J.; Yang, X.-R.; Zhang, R. A novel inhibitor of MDM2 oncogene blocks metastasis of hepatocellular carcinoma and overcomes chemoresistance. Genes Dis. 2019, 6, 419–430. [Google Scholar] [CrossRef] [PubMed]
  49. Cheng, S.; Castillo, V.; Sliva, D. CDC20 associated with cancer metastasis and novel mushroom-derived CDC20 inhibitors with antimetastatic activity. Int. J. Oncol. 2019, 54, 2250–2256. [Google Scholar] [CrossRef]
  50. Han, T.; Jiang, S.; Zheng, H.; Yin, Q.; Xie, M.; Little, M.R.; Yin, X.; Chen, M.; Song, S.J.; Beg, A.A.; et al. Interplay between c-Src and the APC/C co-activator Cdh1 regulates mammary tumorigenesis. Nat. Commun. 2019, 10, 3716. [Google Scholar] [CrossRef]
  51. Zhong, J.; Shaik, S.; Wan, L.; Tron, A.E.; Wang, Z.; Sun, L.; Inuzuka, H.; Wei, W. SCFβ-TRCP targets MTSS1 for ubiquitination-mediated destruction to regulate cancer cell proliferation and migration. Oncotarget 2013, 4, 2339–2353. [Google Scholar] [CrossRef] [Green Version]
  52. Ma, J.; Lu, Y.; Zhang, S.; Li, Y.; Huang, J.; Yin, Z.; Ren, J.; Huang, K.; Liu, L.; Yang, K.; et al. β-Trcp ubiquitin ligase and RSK2 kinase-mediated degradation of FOXN2 promotes tumorigenesis and radioresistance in lung cancer. Cell Death Differ. 2018, 25, 1473–1485. [Google Scholar] [CrossRef]
  53. Shaik, S.; Nucera, C.; Inuzuka, H.; Gao, D.; Garnaas, M.; Frechette, G.; Harris, L.; Wan, L.; Fukushima, H.; Husain, A.; et al. SCFβ-TRCP suppresses angiogenesis and thyroid cancer cell migration by promoting ubiquitination and destruction of VEGF receptor 2. J. Exp. Med. 2012, 209, 1289–1307. [Google Scholar] [CrossRef]
  54. Paul, P.J.; Raghu, D.; Chan, A.-L.; Gulati, T.; Lambeth, L.; Takano, E.; Herold, M.J.; Hagekyriakou, J.; Vessella, R.L.; Fedele, C.; et al. Restoration of tumor suppression in prostate cancer by targeting the E3 ligase E6AP. Oncogene 2016, 35, 6235–6245. [Google Scholar] [CrossRef] [PubMed]
  55. Raghu, D.; Paul, P.J.; Gulati, T.; Deb, S.; Khoo, C.; Russo, A.; Gallo, E.; Blandino, G.; Chan, A.-L.; Takano, E.; et al. E6AP promotes prostate cancer by reducing p27 expression. Oncotarget 2017, 8, 42939–42948. [Google Scholar] [CrossRef] [PubMed]
  56. Gamell, C.; Bandilovska, I.; Gulati, T.; Kogan, A.; Lim, S.C.; Kovacevic, Z.; Takano, E.A.; Timpone, C.; Agupitan, A.D.; Litchfield, C.; et al. E6AP Promotes a Metastatic Phenotype in Prostate Cancer. iScience 2019, 22, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Xu, W.; Taranets, L.; Popov, N. Regulating Fbw7 on the road to cancer. Semin. Cancer Biol. 2016, 36, 62–70. [Google Scholar] [CrossRef]
  58. Yumimoto, K.; Nakayama, K.I. Recent insight into the role of FBXW7 as a tumor suppressor. Semin. Cancer Biol. 2020, 67 Pt 2, 1–15. [Google Scholar] [CrossRef]
  59. Yada, M.; Hatakeyama, S.; Kamura, T.; Nishiyama, M.; Tsunematsu, R.; Imaki, H.; Ishida, N.; Okumura, F.; Nakayama, K.; Nakayama, K.I. Phosphorylation-dependent degradation of c-Myc is mediated by the F-box protein Fbw7. EMBO J. 2004, 23, 2116–2125. [Google Scholar] [CrossRef] [Green Version]
  60. Imura, S.; Tovuu, L.-O.; Utsunomiya, T.; Morine, Y.; Ikemoto, T.; Arakawa, Y.; Kanamoto, M.; Iwahashi, S.; Saito, Y.; Takasu, C.; et al. The role of Fbxw7 expression in hepatocellular carcinoma and adjacent non-tumor liver tissue. J. Gastroenterol. Hepatol. 2014, 29, 1822–1829. [Google Scholar] [CrossRef]
  61. Welcker, M.; Clurman, B.E. FBW7 ubiquitin ligase: A tumour suppressor at the crossroads of cell division, growth and differentiation. Nat. Rev. Cancer 2008, 8, 83–93. [Google Scholar] [CrossRef]
  62. Akhoondi, S.; Sun, D.; Von Der Lehr, N.; Apostolidou, S.; Klotz, K.; Maljukova, A.; Cepeda, D.; Fiegl, H.; Dofou, D.; Marth, C.; et al. FBXW7/hCDC4 Is a General Tumor Suppressor in Human Cancer. Cancer Res. 2007, 67, 9006–9012. [Google Scholar] [CrossRef] [Green Version]
  63. DeVine, T.; Dai, M.S. Targeting the Ubiquitin-Mediated Proteasome Degradation of p53 for Cancer Therapy. Curr. Pharm. Des. 2013, 19, 3248–3262. [Google Scholar] [CrossRef] [Green Version]
  64. Wade, M.; Li, Y.-C.; Wahl, G.M. MDM2, MDMX and p53 in oncogenesis and cancer therapy. Nat. Rev. Cancer 2013, 13, 83–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Shaikh, M.F.; Morano, W.F.; Lee, J.; Gleeson, E.; Babcock, B.D.; Michl, J.; Sarafraz-Yazdi, E.; Pincus, M.R.; Bowne, W.B. Emerging Role of MDM2 as Target for Anti-Cancer Therapy: A Review. Ann. Clin. Lab. Sci. 2016, 46, 627–634. [Google Scholar] [PubMed]
  66. Yang, L.; Song, T.; Cheng, Q.; Chen, L.; Chen, J. Mutant p53 Sequestration of the MDM2 Acidic Domain Inhibits E3 Ligase Activity. Mol. Cell. Biol. 2019, 39, e00375-18. [Google Scholar] [CrossRef] [Green Version]
  67. Wu, C.-E.; Koay, T.S.; Ho, Y.-H.; Lovat, P.; Lunec, J. TP53 mutant cell lines selected for resistance to MDM2 inhibitors retain growth inhibition by MAPK pathway inhibitors but a reduced apoptotic response. Cancer Cell Int. 2019, 19, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Riscal, R.; Schrepfer, E.; Arena, G.; Cissé, M.Y.; Bellvert, F.; Heuillet, M.; Rambow, F.; Bonneil, E.; Sabourdy, F.; Vincent, C.; et al. Chromatin-Bound MDM2 Regulates Serine Metabolism and Redox Homeostasis Independently of p53. Mol. Cell 2016, 62, 890–902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Alfarsi, L.H.; El Ansari, R.; Craze, M.L.; Toss, M.S.; Masisi, B.; Ellis, I.O.; Rakha, E.A.; Green, A.R. CDC20 expression in oestrogen receptor positive breast cancer predicts poor prognosis and lack of response to endocrine therapy. Breast Cancer Res. Treat. 2019, 178, 535–544. [Google Scholar] [CrossRef] [PubMed]
  70. Schrock, M.S.; Stromberg, B.R.; Scarberry, L.; Summers, M.K. APC/C ubiquitin ligase: Functions and mechanisms in tumorigenesis. Semin. Cancer Biol. 2020, 67 Pt 2, 80–91. [Google Scholar] [CrossRef]
  71. Beaudenon, S.; Huibregtse, J.M. HPV E6, E6AP and cervical cancer. BMC Biochem. 2008, 9 (Suppl. S1), S4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Amerik, A.Y.; Hochstrasser, M. Mechanism and function of deubiquitinating enzymes. Biochim. Biophys. Acta 2004, 1695, 189–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Glickman, M.H.; Ciechanover, A. The Ubiquitin-Proteasome Proteolytic Pathway: Destruction for the Sake of Construction. Physiol. Rev. 2002, 82, 373–428. [Google Scholar] [CrossRef]
  74. Schnell, J.D.; Hicke, L. Non-traditional Functions of Ubiquitin and Ubiquitin-binding Proteins. J. Biol. Chem. 2003, 278, 35857–35860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Mukhopadhyay, D.; Riezman, H. Proteasome-Independent Functions of Ubiquitin in Endocytosis and Signaling. Science 2007, 315, 201–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Hu, M.; Li, P.; Li, M.; Li, W.; Yao, T.; Wu, J.-W.; Gu, W.; Cohen, R.E.; Shi, Y. Crystal Structure of a UBP-Family Deubiquitinating Enzyme in Isolation and in Complex with Ubiquitin Aldehyde. Cell 2002, 111, 1041–1054. [Google Scholar] [CrossRef] [Green Version]
  77. Hu, M.; Li, P.; Song, L.; Jeffrey, P.D.; Chernova, T.A.; Wilkinson, K.D.; Cohen, R.E.; Shi, Y. Structure and mechanisms of the proteasome-associated deubiquitinating enzyme USP14. EMBO J. 2005, 24, 3747–3756. [Google Scholar] [CrossRef]
  78. Avvakumov, G.V.; Walker, J.R.; Xue, S.; Finerty, P.J.; MacKenzie, F.; Newman, E.M.; Dhe-Paganon, S. Amino-terminal Dimerization, NRDP1-Rhodanese Interaction, and Inhibited Catalytic Domain Conformation of the Ubiquitin-specific Protease 8 (USP8). J. Biol. Chem. 2006, 281, 38061–38070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Renatus, M.; Gil Parrado, S.; D’Arcy, A.; Eidhoff, U.; Gerhartz, B.; Hassiepen, U.; Pierrat, B.; Riedl, R.; Vinzenz, D.; Worpenberg, S.; et al. Structural Basis of Ubiquitin Recognition by the Deubiquitinating Protease USP2. Structure 2006, 14, 1293–1302. [Google Scholar] [CrossRef] [Green Version]
  80. Ismail, I.H.; Davidson, R.; Gagné, J.-P.; Xu, Z.Z.; Poirier, G.G.; Hendzel, M.J. Germline Mutations in BAP1 Impair Its Function in DNA Double-Strand Break Repair. Cancer Res. 2014, 74, 4282–4294. [Google Scholar] [CrossRef] [Green Version]
  81. Yu, H.; Pak, H.; Hammond-Martel, I.; Ghram, M.; Rodrigue, A.; Daou, S.; Barbour, H.; Corbeil, L.; Hébert, J.; Drobetsky, E.; et al. Tumor suppressor and deubiquitinase BAP1 promotes DNA double-strand break repair. Proc. Natl. Acad. Sci. USA 2014, 111, 285–290. [Google Scholar] [CrossRef] [Green Version]
  82. Lee, H.-S.; Seo, H.-R.; Lee, S.-A.; Choi, S.; Kang, D.; Kwon, J. BAP1 promotes stalled fork restart and cell survival via INO80 in response to replication stress. Biochem. J. 2019, 476, 3053–3066. [Google Scholar] [CrossRef]
  83. Zhang, Y.; Shi, J.; Liu, X.; Feng, L.; Gong, Z.; Koppula, P.; Sirohi, K.; Li, X.; Wei, Y.; Lee, H.; et al. BAP1 links metabolic regulation of ferroptosis to tumour suppression. Nat. Cell Biol. 2018, 20, 1181–1192. [Google Scholar] [CrossRef]
  84. Wang, L.; Guan, R.; Xie, L.; Liao, X.; Xiong, K.; Rees, T.W.; Chen, Y.; Ji, L.; Chao, H. An ER-Targeting Iridium(III) Complex That Induces Immunogenic Cell Death in Non-Small-Cell Lung Cancer. Angew. Chem. Int. Ed. 2021, 60, 4657–4665. [Google Scholar] [CrossRef] [PubMed]
  85. Van Loosdregt, J.; Coffer, P.J. Post-translational modification networks regulating FOXP3 function. Trends Immunol. 2014, 35, 368–378. [Google Scholar] [CrossRef] [PubMed]
  86. Li, M.; Chen, D.; Shiloh, A.; Luo, J.; Nikolaev, A.Y.; Qin, J.; Gu, W. Deubiquitination of p53 by HAUSP is an important pathway for p53 stabilization. Nature 2002, 416, 648–653. [Google Scholar] [CrossRef] [PubMed]
  87. Zhang, X.-Y.; Varthi, M.; Sykes, S.M.; Phillips, C.; Warzecha, C.; Zhu, W.; Wyce, A.; Thorne, A.W.; Berger, S.L.; McMahon, S.B. The Putative Cancer Stem Cell Marker USP22 Is a Subunit of the Human SAGA Complex Required for Activated Transcription and Cell-Cycle Progression. Mol. Cell 2008, 29, 102–111. [Google Scholar] [CrossRef] [Green Version]
  88. Zhang, H.; Han, B.; Lu, H.; Zhao, Y.; Chen, X.; Meng, Q.; Cao, M.; Cai, L.; Hu, J. USP22 promotes resistance to EGFR-TKIs by preventing ubiquitination-mediated EGFR degradation in EGFR-mutant lung adenocarcinoma. Cancer Lett. 2018, 433, 186–198. [Google Scholar] [CrossRef]
  89. Lin, Z.; Tan, C.; Qiu, Q.; Kong, S.; Yang, H.; Zhao, F.; Liu, Z.; Li, J.; Kong, Q.; Gao, B.; et al. Ubiquitin-specific protease 22 is a deubiquitinase of CCNB1. Cell Discov. 2015, 1, 15028. [Google Scholar] [CrossRef]
  90. Zhou, A.; Lin, K.; Zhang, S.; Chen, Y.; Zhang, N.; Xue, J.; Wang, Z.; Aldape, K.D.; Xie, K.; Woodgett, J.R.; et al. Nuclear GSK3β promotes tumorigenesis by phosphorylating KDM1A and inducing its deubiquitylation by USP22. Nat. Cell Biol. 2016, 18, 954–966. [Google Scholar] [CrossRef] [Green Version]
  91. Wang, W.; Zou, L.; Zhou, D.; Zhou, Z.; Tang, F.; Xu, Z.; Liu, X. Overexpression of ubiquitin carboxyl terminal hydrolase-L1 enhances multidrug resistance and invasion/metastasis in breast cancer by activating the MAPK/Erk signaling pathway. Mol. Carcinog. 2015, 55, 1329–1342. [Google Scholar] [CrossRef]
  92. Singh, A.N.; Oehler, J.; Torrecilla, I.; Kilgas, S.; Li, S.; Vaz, B.; Guérillon, C.; Fielden, J.; Hernandez-Carralero, E.; Cabrera, E.; et al. The p97–Ataxin 3 complex regulates homeostasis of the DNA damage response E3 ubiquitin ligase RNF 8. EMBO J. 2019, 38, e102361. [Google Scholar] [CrossRef]
  93. Shi, Z.; Chen, J.; Zhang, X.; Chu, J.; Han, Z.; Xu, D.; Gan, S.; Pan, X.; Ye, J.; Cui, X. Ataxin-3 promotes testicular cancer cell proliferation by inhibiting anti-oncogene PTEN. Biochem. Biophys. Res. Commun. 2018, 503, 391–396. [Google Scholar] [CrossRef]
  94. Butler, L.R.; Densham, R.M.; Jia, J.; Garvin, A.J.; Stone, H.R.; Shah, V.; Weekes, D.; Festy, F.; Beesley, J.; Morris, J.R. The proteasomal de-ubiquitinating enzyme POH1 promotes the double-strand DNA break response. EMBO J. 2012, 31, 3918–3934. [Google Scholar] [CrossRef]
  95. Byrne, A.; McLaren, R.P.; Mason, P.; Chai, L.; Dufault, M.R.; Huang, Y.; Liang, B.; Gans, J.D.; Zhang, M.; Carter, K. Knockdown of human deubiquitinase PSMD14 induces cell cycle arrest and senescence. Exp. Cell Res. 2010, 316, 258–271. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, B.; Ma, A.; Zhang, L.; Jin, W.-L.; Qiang, X.; Xu, G.; Qiu, B.; Yang, Z.; Liu, Y.; Xia, Q.; et al. POH1 deubiquitylates and stabilizes E2F1 to promote tumour formation. Nat. Commun. 2015, 6, 8704. [Google Scholar] [CrossRef]
  97. Jin, Z.; Li, Y.; Pitti, R.; Lawrence, D.; Pham, V.C.; Lill, J.R.; Ashkenazi, A. Cullin3-Based Polyubiquitination and p62-Dependent Aggregation of Caspase-8 Mediate Extrinsic Apoptosis Signaling. Cell 2009, 137, 721–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Kim, S.-W.; Ramasamy, K.; Bouamar, H.; Lin, A.-P.; Jiang, D.; Aguiar, R.C.T. MicroRNAs miR-125a and miR-125b constitutively activate the NF- B pathway by targeting the tumor necrosis factor alpha-induced protein 3 (TNFAIP3, A20). Proc. Natl. Acad. Sci. USA 2012, 109, 7865–7870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Balkhi, M.Y.; Iwenofu, O.H.; Bakkar, N.; Ladner, K.J.; Chandler, D.S.; Houghton, P.J.; London, C.A.; Kraybill, W.; Perrotti, D.; Croce, C.M.; et al. miR-29 Acts as a Decoy in Sarcomas to Protect the Tumor Suppressor A20 mRNA from Degradation by HuR. Sci. Signal. 2013, 6, ra63. [Google Scholar] [CrossRef] [Green Version]
  100. Jensen, D.E.; Proctor, M.; Marquis, S.T.; Gardner, H.P.; Ha, S.I.; Chodosh, L.A.; Ishov, A.M.; Tommerup, N.; Vissing, H.; Sekido, Y.; et al. BAP1: A novel ubiquitin hydrolase which binds to the BRCA1 RING finger and enhances BRCA1-mediated cell growth suppression. Oncogene 1998, 16, 1097–1112. [Google Scholar] [CrossRef] [Green Version]
  101. Nishikawa, H.; Wu, W.; Koike, A.; Kojima, R.; Gomi, H.; Fukuda, M.; Ohta, T. BRCA1-Associated Protein 1 Interferes with BRCA1/BARD1 RING Heterodimer Activity. Cancer Res. 2009, 69, 111–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Hauri, S.; Comoglio, F.; Seimiya, M.; Gerstung, M.; Glatter, T.; Hansen, K.; Aebersold, R.; Paro, R.; Gstaiger, M.; Beisel, C. A High-Density Map for Navigating the Human Polycomb Complexome. Cell Rep. 2016, 17, 583–595. [Google Scholar] [CrossRef] [Green Version]
  103. Lee, H.-S.; Lee, S.-A.; Hur, S.-K.; Seo, J.-W.; Kwon, J. Stabilization and targeting of INO80 to replication forks by BAP1 during normal DNA synthesis. Nat. Commun. 2014, 5, 5128. [Google Scholar] [CrossRef] [Green Version]
  104. Machida, Y.J.; Machida, Y.; Vashisht, A.A.; Wohlschlegel, J.A.; Dutta, A. The Deubiquitinating Enzyme BAP1 Regulates Cell Growth via Interaction with HCF-1. J. Biol. Chem. 2009, 284, 34179–34188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Yu, H.; Mashtalir, N.; Daou, S.; Hammond-Martel, I.; Ross, J.; Sui, G.; Hart, G.W.; Rauscher, F.J.; Drobetsky, E.; Milot, E.; et al. The Ubiquitin Carboxyl Hydrolase BAP1 Forms a Ternary Complex with YY1 and HCF-1 and Is a Critical Regulator of Gene Expression. Mol. Cell. Biol. 2010, 30, 5071–5085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Giorgi, C.; Marchi, S.; Pinton, P. The machineries, regulation and cellular functions of mitochondrial calcium. Nat. Rev. Mol. Cell Biol. 2018, 19, 713–730. [Google Scholar] [CrossRef]
  107. Baughman, J.M.; Rose, C.M.; Kolumam, G.; Webster, J.D.; Wilkerson, E.M.; Merrill, A.E.; Rhoads, T.W.; Noubade, R.; Katavolos, P.; Lesch, J.; et al. NeuCode Proteomics Reveals Bap1 Regulation of Metabolism. Cell Rep. 2016, 16, 583–595. [Google Scholar] [CrossRef] [Green Version]
  108. Ruan, H.-B.; Han, X.; Li, M.-D.; Singh, J.P.; Qian, K.; Azarhoush, S.; Zhao, L.; Bennett, A.M.; Samuel, V.T.; Wu, J.; et al. O-GlcNAc Transferase/Host Cell Factor C1 Complex Regulates Gluconeogenesis by Modulating PGC-1α Stability. Cell Metab. 2012, 16, 226–237. [Google Scholar] [CrossRef] [Green Version]
  109. Carbone, M.; Adusumilli, P.S.; Alexander, H.R., Jr.; Baas, P.; Bardelli, F.; Bononi, A.; Bueno, R.; Felley-Bosco, E.; Galateau-Salle, F.; Jablons, D.; et al. Mesothelioma: Scientific clues for prevention, diagnosis, and therapy. CA Cancer J. Clin. 2019, 69, 402–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Rai, K.R.; Pilarski, R.; Cebulla, C.M.; Abdelrahman, M.H. Comprehensive review ofBAP1tumor predisposition syndrome with report of two new cases. Clin. Genet. 2016, 89, 285–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Kittaneh, M.; Berkelhammer, C. Detecting germline BAP1 mutations in patients with peritoneal mesothelioma: Benefits to patient and family members. J. Transl. Med. 2018, 16, 194. [Google Scholar] [CrossRef] [Green Version]
  112. Panou, V.; Gadiraju, M.; Wolin, A.; Weipert, C.M.; Skarda, E.; Husain, A.N.; Patel, J.D.; Rose, B.; Zhang, S.R.; Weatherly, M.; et al. Frequency of Germline Mutations in Cancer Susceptibility Genes in Malignant Mesothelioma. J. Clin. Oncol. 2018, 36, 2863–2871. [Google Scholar] [CrossRef] [PubMed]
  113. Kobrinski, D.A.; Yang, H.; Kittaneh, M. BAP1: Role in carcinogenesis and clinical implications. Transl. Lung Cancer Res. 2020, 9 (Suppl. S1), S60–S66. [Google Scholar] [CrossRef]
  114. Yoshikawa, Y.; Emi, M.; Nakano, T.; Gaudino, G. Mesothelioma developing in carriers of inherited genetic mutations. Transl. Lung Cancer Res. 2020, 9 (Suppl. S1), S67–S76. [Google Scholar] [CrossRef]
  115. Kuznetsov, J.N.; Aguero, T.H.; Owens, D.A.; Kurtenbach, S.; Field, M.G.; Durante, M.A.; Rodriguez, D.A.; King, M.L.; Harbour, J.W. BAP1 regulates epigenetic switch from pluripotency to differentiation in developmental lineages giving rise to BAP1-mutant cancers. Sci. Adv. 2019, 5, eaax1738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Harbour, J.W.; Onken, M.D.; Roberson, E.D.O.; Duan, S.; Cao, L.; Worley, L.A.; Council, M.L.; Matatall, K.A.; Helms, C.; Bowcock, A.M. Frequent Mutation of BAP1 in Metastasizing Uveal Melanomas. Science 2010, 330, 1410–1413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Njauw, C.-N.J.; Kim, I.; Piris, A.; Gabree, M.; Taylor, M.; Lane, A.M.; DeAngelis, M.M.; Gragoudas, E.; Duncan, L.M.; Tsao, H. Germline BAP1 Inactivation Is Preferentially Associated with Metastatic Ocular Melanoma and Cutaneous-Ocular Melanoma Families. PLoS ONE 2012, 7, e35295. [Google Scholar] [CrossRef] [Green Version]
  118. Peña-Llopis, S.; Vega-Rubín-De-Celis, S.; Liao, A.; Leng, N.; Pavía-Jiménez, A.; Wang, S.; Yamasaki, T.; Zhrebker, L.; Sivanand, S.; Spence, P.; et al. BAP1 loss defines a new class of renal cell carcinoma. Nat. Genet. 2012, 44, 751–759. [Google Scholar] [CrossRef]
  119. Popova, T.; Hebert, L.; Jacquemin, V.; Gad, S.; Caux-Moncoutier, V.; Dubois-D’Enghien, C.; Richaudeau, B.; Renaudin, X.; Sellers, J.; Nicolas, A.; et al. Germline BAP1 Mutations Predispose to Renal Cell Carcinomas. Am. J. Hum. Genet. 2013, 92, 974–980. [Google Scholar] [CrossRef] [Green Version]
  120. Lin, M.; Zhang, L.; Hildebrandt, M.A.T.; Huang, M.; Wu, X.; Ye, Y. Common, germline genetic variations in the novel tumor suppressor BAP1 and risk of developing different types of cancer. Oncotarget 2017, 8, 74936–74946. [Google Scholar] [CrossRef] [Green Version]
  121. Carbone, M.; Yang, H.; Pass, H.I.; Krausz, T.; Testa, J.R.; Gaudino, G. BAP1 and cancer. Nat. Rev. Cancer 2013, 13, 153–159. [Google Scholar] [CrossRef]
  122. Testa, J.R.; Cheung, M.; Pei, J.; Below, J.E.; Tan, Y.; Sementino, E.; Cox, N.J.; Dogan, A.U.; Pass, H.I.; Trusa, S.; et al. Germline BAP1 mutations predispose to malignant mesothelioma. Nat. Genet. 2011, 43, 1022–1025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Carbone, M.; Ferris, L.K.; Baumann, F.; Napolitano, A.; Lum, C.A.; Flores, E.G.; Gaudino, G.; Powers, A.; Bryant-Greenwood, P.; Krausz, T.; et al. BAP1 cancer syndrome: Malignant mesothelioma, uveal and cutaneous melanoma, and MBAITs. J. Transl. Med. 2012, 10, 179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Everett, R.D.; Meredith, M.; Orr, A.; Cross, A.; Kathoria, M.; Parkinson, J. A novel ubiquitin-specific protease is dynamically associated with the PML nuclear domain and binds to a herpesvirus regulatory protein. EMBO J. 1997, 16, 1519–1530. [Google Scholar] [CrossRef] [Green Version]
  125. Holowaty, M.N.; Zeghouf, M.; Wu, H.; Tellam, J.; Athanasopoulos, V.; Greenblatt, J.; Frappier, L. Protein Profiling with Epstein-Barr Nuclear Antigen-1 Reveals an Interaction with the Herpesvirus-associated Ubiquitin-specific Protease HAUSP/USP7. J. Biol. Chem. 2003, 278, 29987–29994. [Google Scholar] [CrossRef] [Green Version]
  126. Lu, Y.; Orr, A.; Everett, R.D. Stimulation of the Replication of ICP0-Null Mutant Herpes Simplex Virus 1 and pp71-Deficient Human Cytomegalovirus by Epstein-Barr Virus Tegument Protein BNRF1. J. Virol. 2016, 90, 9664–9673. [Google Scholar] [CrossRef] [Green Version]
  127. Everett, R.D. HSV-1 Biology and Life Cycle. Methods Mol. Biol. 2014, 1144, 1–17. [Google Scholar] [CrossRef] [PubMed]
  128. Laurence, A.; Belkaid, Y.; O’Shea, J.J. A Degrading View of Regulatory T Cells. Immunity 2013, 39, 201–203. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, L.; Guo, C.; Ji, T.; Chen, X. SOX2 Regulates lncRNA CCAT1/MicroRNA-185-3p/FOXP3 Axis to Affect the Proliferation and Self-Renewal of Cervical Cancer Stem Cells. Nanoscale Res. Lett. 2021, 16, 2. [Google Scholar] [CrossRef]
  130. Cai, J.; Chen, H.; Peng, S.; Meng, J.; Wang, Y.; Zhou, Y.; Qian, X.; Sun, X.; Pang, X.; Zhang, Y.; et al. USP7-TRIM27 axis negatively modulates antiviral type I IFN signaling. FASEB J. 2018, 32, 5238–5249. [Google Scholar] [CrossRef] [Green Version]
  131. Oliner, J.D.; Kinzler, K.W.; Meltzer, P.S.; George, D.L.; Vogelstein, B. Amplification of a gene encoding a p53-associated protein in human sarcomas. Nature 1992, 358, 80–83. [Google Scholar] [CrossRef]
  132. Vassilev, L.T.; Vu, B.T.; Graves, B.; Carvajal, D.; Podlaski, F.; Filipovic, Z.; Kong, N.; Kammlott, U.; Lukacs, C.; Klein, C.; et al. In Vivo Activation of the p53 Pathway by Small-Molecule Antagonists of MDM2. Science 2004, 303, 844–848. [Google Scholar] [CrossRef] [Green Version]
  133. Hu, M.; Gu, L.; Li, M.; Jeffrey, P.D.; Gu, W.; Shi, Y. Structural Basis of Competitive Recognition of p53 and MDM2 by HAUSP/USP7: Implications for the Regulation of the p53–MDM2 Pathway. PLoS Biol. 2006, 4, e27. [Google Scholar] [CrossRef] [PubMed]
  134. Yang, X.; Zang, H.; Luo, Y.; Wu, J.; Fang, Z.; Zhu, W.; Li, Y. High expression of USP22 predicts poor prognosis and advanced clinicopathological features in solid tumors: A meta-analysis. OncoTargets Ther. 2018, 11, 3035–3046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Day, I.N.; Thompson, R.J. UCHL1 (PGP 9.5): Neuronal biomarker and ubiquitin system protein. Prog. Neurobiol. 2010, 90, 327–362. [Google Scholar] [CrossRef] [PubMed]
  136. Tezel, E.; Hibi, K.; Nagasaka, T.; Nakao, A. PGP9.5 as a prognostic factor in pancreatic cancer. Clin. Cancer Res. 2000, 6, 4764–4767. [Google Scholar] [PubMed]
  137. Otsuki, T.; Yata, K.; Takata-Tomokuni, A.; Hyodoh, F.; Miura, Y.; Sakaguchi, H.; Hatayama, T.; Hatada, S.; Tsujioka, T.; Sato, Y.; et al. Expression of protein gene product 9·5 (PGP9·5)/ubiquitin-C-terminal hydrolase 1 (UCHL-1) in human myeloma cells. Br. J. Haematol. 2004, 127, 292–298. [Google Scholar] [CrossRef]
  138. Leiblich, A.; Cross, S.S.; Catto, J.W.; Pesce, G.; Hamdy, F.C.; Rehman, I. Human prostate cancer cells express neuroendocrine cell markers PGP 9.5 and chromogranin A. Prostate 2007, 67, 1761–1769. [Google Scholar] [CrossRef]
  139. Liu, X.; Zeng, B.; Ma, J.; Wan, C. Comparative Proteomic Analysis of Osteosarcoma Cell and Human Primary Cultured Osteoblastic Cell. Cancer Investig. 2009, 27, 345–352. [Google Scholar] [CrossRef]
  140. Hibi, K.; Westra, W.H.; Borges, M.; Goodman, S.; Sidransky, D.; Jen, J. PGP9.5 As a Candidate Tumor Marker for Non-Small-Cell Lung Cancer. Am. J. Pathol. 1999, 155, 711–715. [Google Scholar] [CrossRef] [Green Version]
  141. Kim, H.J.; Magesh, V.; Lee, J.-J.; Kim, S.; Knaus, U.G.; Lee, K.-J. Ubiquitin C-terminal hydrolase-L1 increases cancer cell invasion by modulating hydrogen peroxide generated via NADPH oxidase 4. Oncotarget 2015, 6, 16287–16303. [Google Scholar] [CrossRef]
  142. Brichory, F.; Beer, D.; Le Naour, F.; Giordano, T.; Hanash, S. Proteomics-based identification of protein gene product 9.5 as a tumor antigen that induces a humoral immune response in lung cancer. Cancer Res. 2001, 61, 7908–7912. [Google Scholar] [PubMed]
  143. Nicastro, G.; Menon, R.P.; Masino, L.; Knowles, P.P.; McDonald, N.Q.; Pastore, A. The solution structure of the Josephin domain of ataxin-3: Structural determinants for molecular recognition. Proc. Natl. Acad. Sci. USA 2005, 102, 10493–10498. [Google Scholar] [CrossRef] [Green Version]
  144. Mao, Y.; Senic-Matuglia, F.; Di Fiore, P.P.; Polo, S.; Hodsdon, M.E.; De Camilli, P. Deubiquitinating function of ataxin-3: Insights from the solution structure of the Josephin domain. Proc. Natl. Acad. Sci. USA 2005, 102, 12700–12705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Burnett, B.; Li, F.; Pittman, R.N. The polyglutamine neurodegenerative protein ataxin-3 binds polyubiquitylated proteins and has ubiquitin protease activity. Hum. Mol. Genet. 2003, 12, 3195–3205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Kakarougkas, A.; Ismail, A.; Katsuki, Y.; Freire, R.; Shibata, A.; Jeggo, P.A. Co-operation of BRCA1 and POH1 relieves the barriers posed by 53BP1 and RAP80 to resection. Nucleic Acids Res. 2013, 41, 10298–10311. [Google Scholar] [CrossRef]
  147. Stitzel, M.L.; Durso, R.; Reese, J.C. The proteasome regulates the UV-induced activation of the AP-1-like transcription factor Gcn4. Genes Dev. 2001, 15, 128–133. [Google Scholar] [CrossRef] [Green Version]
  148. Opipari, A.W.; Hu, H.M.; Yabkowitz, R.; Dixit, V.M. The A20 zinc finger protein protects cells from tumor necrosis factor cytotoxicity. J. Biol. Chem. 1992, 267, 12424–12427. [Google Scholar] [CrossRef] [PubMed]
  149. He, K.-L.; Ting, A.T. A20 Inhibits Tumor Necrosis Factor (TNF) Alpha-Induced Apoptosis by Disrupting Recruitment of TRADD and RIP to the TNF Receptor 1 Complex in Jurkat T Cells. Mol. Cell. Biol. 2002, 22, 6034–6045. [Google Scholar] [CrossRef] [Green Version]
  150. Chanudet, E.; Huang, Y.; Ichimura, K.; Dong, G.; Hamoudi, R.A.; Radford, J.; Wotherspoon, A.C.; Isaacson, P.G.; Ferry, J.; Du, M.-Q. A20 is targeted by promoter methylation, deletion and inactivating mutation in MALT lymphoma. Leukemia 2010, 24, 483–487. [Google Scholar] [CrossRef]
  151. Novak, U.; Rinaldi, A.; Kwee, I.; Nandula, S.V.; Rancoita, P.M.V.; Compagno, M.; Cerri, M.; Rossi, D.; Murty, V.V.; Zucca, E.; et al. The NF-κB negative regulator TNFAIP3 (A20) is inactivated by somatic mutations and genomic deletions in marginal zone lymphomas. Blood 2009, 113, 4918–4921. [Google Scholar] [CrossRef]
  152. Hochstrasser, M. Origin and function of ubiquitin-like proteins. Nature 2009, 458, 422–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Schulman, B.A.; Harper, J.W. Ubiquitin-like protein activation by E1 enzymes: The apex for downstream signalling pathways. Nat. Rev. Mol. Cell Biol. 2009, 10, 319–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Barghout, S.H.; Schimmer, A.D. E1 Enzymes as Therapeutic Targets in Cancer. Pharmacol. Rev. 2021, 73, 1–56. [Google Scholar] [CrossRef] [PubMed]
  155. Groen, E.J.; Gillingwater, T.H. UBA1: At the Crossroads of Ubiquitin Homeostasis and Neurodegeneration. Trends Mol. Med. 2015, 21, 622–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Tsherniak, A.; Vazquez, F.; Montgomery, P.G.; Weir, B.A.; Kryukov, G.; Cowley, G.S.; Gill, S.; Harrington, W.F.; Pantel, S.; Krill-Burger, J.M.; et al. Defining a Cancer Dependency Map. Cell 2017, 170, 564–576. [Google Scholar] [CrossRef] [Green Version]
  157. Luo, J.; Solimini, N.L.; Elledge, S.J. Principles of Cancer Therapy: Oncogene and Non-oncogene Addiction. Cell 2009, 136, 823–837. [Google Scholar] [CrossRef] [Green Version]
  158. Dobbelstein, M.; Moll, U. Targeting tumour-supportive cellular machineries in anticancer drug development. Nat. Rev. Drug Discov. 2014, 13, 179–196. [Google Scholar] [CrossRef]
  159. Soucy, T.A.; Smith, P.G.; Milhollen, M.A.; Berger, A.J.; Gavin, J.M.; Adhikari, S.; Brownell, J.E.; Burke, K.E.; Cardin, D.P.; Critchley, S.; et al. An inhibitor of NEDD8-activating enzyme as a new approach to treat cancer. Nat. Cell Biol. 2009, 458, 732–736. [Google Scholar] [CrossRef]
  160. Hyer, M.L.; Milhollen, M.A.; Ciavarri, J.; Fleming, P.; Traore, T.; Sappal, D.; Huck, J.; Shi, J.; Gavin, J.; Brownell, J.; et al. A small-molecule inhibitor of the ubiquitin activating enzyme for cancer treatment. Nat. Med. 2018, 24, 186–193. [Google Scholar] [CrossRef]
  161. Duda, D.M.; Borg, L.A.; Scott, D.C.; Hunt, H.W.; Hammel, M.; Schulman, B.A. Structural Insights into NEDD8 Activation of Cullin-RING Ligases: Conformational Control of Conjugation. Cell 2008, 134, 995–1006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Osaka, F.; Saeki, M.; Katayama, S.; Aida, N.; Toh-E, A.; Kominami, K.; Toda, T.; Suzuki, T.; Chiba, T.; Tanaka, K.; et al. Covalent modifier NEDD8 is essential for SCF ubiquitin-ligase in fission yeast. EMBO J. 2000, 19, 3475–3484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Toth, J.I.; Yang, L.; Dahl, R.; Petroski, M.D. A Gatekeeper Residue for NEDD8-Activating Enzyme Inhibition by MLN4924. Cell Rep. 2012, 1, 309–316. [Google Scholar] [CrossRef] [Green Version]
  164. Tisdale, M.J. Cachexia in cancer patients. Nat. Rev. Cancer 2002, 2, 862–871. [Google Scholar] [CrossRef]
  165. Adams, J. The development of proteasome inhibitors as anticancer drugs. Cancer Cell 2004, 5, 417–421. [Google Scholar] [CrossRef] [Green Version]
  166. Hideshima, T.; Richardson, P.; Chauhan, D.; Palombella, V.J.; Elliott, P.J.; Adams, J.; Anderson, K.C. The proteasome inhibitor PS-341 inhibits growth, induces apoptosis, and overcomes drug resistance in human multiple myeloma cells. Cancer Res. 2001, 61, 3071–3076. [Google Scholar] [PubMed]
  167. Chauhan, D.; Hideshima, T.; Mitsiades, C.; Richardson, P.; Anderson, K.C. Proteasome inhibitor therapy in multiple myeloma. Mol. Cancer Ther. 2005, 4, 686–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Hideshima, T.; Mitsiades, C.; Akiyama, M.; Hayashi, T.; Chauhan, D.; Richardson, P.; Schlossman, R.; Podar, K.; Munshi, N.C.; Mitsiades, N.; et al. Molecular mechanisms mediating antimyeloma activity of proteasome inhibitor PS-341. Blood 2003, 101, 1530–1534. [Google Scholar] [CrossRef] [Green Version]
  169. Liu, F.-T.; Agrawal, S.G.; Gribben, J.G.; Ye, H.; Du, M.-Q.; Newland, A.C.; Jia, L. Bortezomib blocks Bax degradation in malignant B cells during treatment with TRAIL. Blood 2008, 111, 2797–2805. [Google Scholar] [CrossRef] [Green Version]
  170. Pei, X.-Y.; Dai, Y.; Grant, S. Synergistic Induction of Oxidative Injury and Apoptosis in Human Multiple Myeloma Cells by the Proteasome Inhibitor Bortezomib and Histone Deacetylase Inhibitors. Clin. Cancer Res. 2004, 10, 3839–3852. [Google Scholar] [CrossRef] [Green Version]
  171. Mata-Cantero, L.; Lobato-Gil, S.; Aillet, F.; Lang, V.; Rodriguez, M.S. The Ubiquitin-Proteasome System (UPS) as a Cancer Drug Target: Emerging Mechanisms and Therapeutics. In Stress Response Pathways in Cancer; Wondrak, G., Ed.; Springer: Dordrecht, The Netherlands, 2015; pp. 225–264. [Google Scholar]
  172. Zhang, X.; Linder, S.; Bazzaro, M. Drug Development Targeting the Ubiquitin–Proteasome System (UPS) for the Treatment of Human Cancers. Cancers 2020, 12, 902. [Google Scholar] [CrossRef] [Green Version]
  173. Kortuem, K.M.; Stewart, A.K. Carfilzomib. Blood 2013, 121, 893–897. [Google Scholar] [CrossRef] [PubMed]
  174. Meng, L.; Mohan, R.; Kwok, B.H.B.; Elofsson, M.; Sin, N.; Crews, C.M. Epoxomicin, a potent and selective proteasome inhibitor, exhibits in vivo antiinflammatory activity. Proc. Nat. Acad. Sci. USA 1999, 96, 10403–10408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Demo, S.D.; Kirk, C.J.; Aujay, M.A.; Buchholz, T.J.; Dajee, M.; Ho, M.N.; Jiang, J.; Laidig, G.J.; Lewis, E.R.; Parlati, F.; et al. Antitumor Activity of PR-171, a Novel Irreversible Inhibitor of the Proteasome. Cancer Res. 2007, 67, 6383–6391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Arastu-Kapur, S.; Anderl, J.L.; Kraus, M.; Parlati, F.; Shenk, K.D.; Lee, S.J.; Muchamuel, T.; Bennett, M.K.; Driessen, C.; Ball, A.J.; et al. Nonproteasomal Targets of the Proteasome Inhibitors Bortezomib and Carfilzomib: A Link to Clinical Adverse Events. Clin. Cancer Res. 2011, 17, 2734–2743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Stewart, A.K.; Rajkumar, S.V.; Dimopoulos, M.A.; Masszi, T.; Špička, I.; Oriol, A.; Hájek, R.; Rosiñol, L.; Siegel, D.S.; Mihaylov, G.G.; et al. Carfilzomib, Lenalidomide, and Dexamethasone for Relapsed Multiple Myeloma. N. Engl. J. Med. 2015, 372, 142–152. [Google Scholar] [CrossRef] [PubMed]
  178. Chauhan, D.; Tian, Z.; Zhou, B.; Kuhn, D.; Orlowski, R.; Raje, N.; Richardson, P.; Anderson, K.C. In Vitro and In Vivo Selective Antitumor Activity of a Novel Orally Bioavailable Proteasome Inhibitor MLN9708 against Multiple Myeloma Cells. Clin. Cancer Res. 2011, 17, 5311–5321. [Google Scholar] [CrossRef] [Green Version]
  179. Luo, Z.; Pan, Y.; Jeong, L.S.; Liu, J.; Jia, L. Inactivation of the Cullin (CUL)-RING E3 ligase by the NEDD8-activating enzyme inhibitor MLN4924 triggers protective autophagy in cancer cells. Autophagy 2012, 8, 1677–1679. [Google Scholar] [CrossRef]
  180. Aubry, A.; Yu, T.; Bremner, R. Preclinical studies reveal MLN4924 is a promising new retinoblastoma therapy. Cell Death Discov. 2020, 6, 1–12. [Google Scholar] [CrossRef] [Green Version]
  181. Swords, R.T.; Kelly, K.R.; Smith, P.G.; Garnsey, J.J.; Mahalingam, D.; Medina, E.; Oberheu, K.; Padmanabhan, S.; O’Dwyer, M.; Nawrocki, S.T.; et al. Inhibition of NEDD8-activating enzyme: A novel approach for the treatment of acute myeloid leukemia. Blood 2010, 115, 3796–3800. [Google Scholar] [CrossRef] [Green Version]
  182. Yao, W.-T.; Wu, J.-F.; Yu, G.-Y.; Wang, R.; Wang, K.; Li, L.-H.; Chen, P.; Jiang, Y.-N.; Cheng, H.; Lee, H.W.; et al. Suppression of tumor angiogenesis by targeting the protein neddylation pathway. Cell Death Dis. 2014, 5, e1059. [Google Scholar] [CrossRef] [Green Version]
  183. Pulvino, M.; Liang, Y.; Oleksyn, D.; DeRan, M.; Van Pelt, E.; Shapiro, J.; Sanz, I.; Chen, L.; Zhao, J. Inhibition of proliferation and sur- vival of diffuse large B-cell lymphoma cells by a small-molecule inhib-itor of the ubiquitin-conjugating enzyme Ubc13-Uev1A. Blood 2012, 120, 1668–1677. [Google Scholar] [CrossRef]
  184. Bauer, S.; Demetri, G.; Jeay, S.; Dummer, R.; Guerreiro, N.; Tan, D.; Kumar, A.; Meille, C.; Van Bree, L.; Halilovic, E.; et al. A phase I, open-label, multi-center, dose escalation study of oral NVP-CGM097, a p53/HDM2-protein-protein interaction inhibitor, in adult patients with selected advanced solid tumors. Ann. Oncol. 2016, 27 (Suppl. S6), vi114–vi135. [Google Scholar] [CrossRef] [Green Version]
  185. Holzer, P.; Masuya, K.; Furet, P.; Kallen, J.; Valat-Stachyra, T.; Ferretti, S.; Berghausen, J.; Bouisset-Leonard, M.; Buschmann, N.; Pissot-Soldermann, C.; et al. Discovery of a Dihydroisoquinolinone Derivative (NVP-CGM097): A Highly Potent and Selective MDM2 Inhibitor Undergoing Phase 1 Clinical Trials in p53wt Tumors. J. Med. Chem. 2015, 58, 6348–6358. [Google Scholar] [CrossRef]
  186. Bauer, T.; Hong, D.; Somaiah, N.; Cai, C.; Song, S.; Kumar, P.; Gajee, R.; Rosen, M.; Kochan, J.; Chen, S.; et al. Abstract B27: A phase I dose escalation study of the MDM2 inhibitor DS-3032b in patients with advanced solid tumors and lymphomas. Mol. Cancer Ther. 2015, 14 (Suppl. S2), Abstract nr B27. [Google Scholar] [CrossRef]
  187. Gounder, M.M.; Bauer, T.M.; Schwartz, G.K.; Masters, T.; Carvajal, R.D.; Song, S.; Kumar, P.; Gajee, R.; Zernovak, O.; Rosen, M.M.; et al. A phase 1 study of the MDM2 inhibitor DS-3032b in patients (pts) with advanced solid tumors and lymphomas. J. Clin. Oncol. 2016, 34 (Suppl. S15), 2581. [Google Scholar] [CrossRef]
  188. DiPersio, J.F.; Erba, H.P.; Larson, R.A.; Luger, S.M.; Tallman, M.S.; Brill, J.M.; Vuagniaux, G.; Rouits, E.; Sorensen, J.M.; Zanna, C. Oral Debio1143 (AT406), an Antagonist of Inhibitor of Apoptosis Proteins, Combined With Daunorubicin and Cytarabine in Patients With Poor-Risk Acute Myeloid Leukemia—Results of a Phase I Dose-Escalation Study. Clin. Lymphoma Myeloma Leuk. 2015, 15, 443–449. [Google Scholar] [CrossRef] [Green Version]
  189. Infante, J.R.; Dees, E.C.; Olszanski, A.J.; Dhuria, S.V.; Sen, S.; Cameron, S.; Cohen, R.B. Phase I Dose-Escalation Study of LCL161, an Oral Inhibitor of Apoptosis Proteins Inhibitor, in Patients With Advanced Solid Tumors. J. Clin. Oncol. 2014, 32, 3103–3110. [Google Scholar] [CrossRef] [PubMed]
  190. Schilder, R.J.; Albertella, M.; Strauss, J.; Sydvander, M.; Nair, S.M.; Urakpo, K.; Norin, S.; Öhd, J. A phase 1/2 study with birinapant in combination with pembrolizumab. J. Clin. Oncol. 2020, 36 (Suppl. S15), TPS3131. [Google Scholar] [CrossRef]
  191. Chen, J.; Dexheimer, T.S.; Ai, Y.; Liang, Q.; Villamil, M.A.; Inglese, J.; Maloney, D.J.; Jadhav, A.; Simeonov, A.; Zhuang, Z. Selective and Cell-Active Inhibitors of the USP1/ UAF1 Deubiquitinase Complex Reverse Cisplatin Resistance in Non-small Cell Lung Cancer Cells. Chem. Biol. 2011, 18, 1390–1400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Rath, B.H.; Camphausen, K.; Tofilon, P.J. Glioblastoma radiosensitization by pimozide. Transl. Cancer Res. 2016, 5 (Suppl. S6), S1029–S1032. [Google Scholar] [CrossRef] [PubMed]
  193. Green, A.K.; Corty, R.W.; Wood, W.A.; Meeneghan, M.; Reeder-Hayes, K.E.; Basch, E.; Milowsky, M.I.; Dusetzina, S.B. Comparative Effectiveness of Mitoxantrone Plus Prednisone Versus Prednisone Alone in Metastatic Castrate-Resistant Prostate Cancer After Docetaxel Failure. Oncolologist 2015, 20, 516–522. [Google Scholar] [CrossRef] [Green Version]
  194. Abbi, K.K.; Rybka, W.; Ehmann, W.C.; Claxton, D.F. Phase I/II Study of Clofarabine, Etoposide, and Mitoxantrone in Patients with Refractory or Relapsed Acute Leukemia. Clin. Lymphoma Myeloma Leuk. 2015, 15, 41–46. [Google Scholar] [CrossRef]
  195. Gill, H.; Yim, R.; Pang, H.H.; Lee, P.; Chan, T.S.Y.; Hwang, Y.; Leung, G.M.K.; Ip, H.; Leung, R.Y.Y.; Yip, S.; et al. Clofarabine, cytarabine, and mitoxantrone in refractory/relapsed acute myeloid leukemia: High response rates and effective bridge to allogeneic hematopoietic stem cell transplantation. Cancer Med. 2020, 9, 3371–3382. [Google Scholar] [CrossRef] [Green Version]
  196. Cornbleet, M.A.; Stuart-Harris, R.C.; Smith, I.E.; Coleman, R.E.; Rubens, R.D.; McDonald, M.; Mouridsen, H.T.; Rainer, H.; Van Oosterom, A.T.; Smyth, J.F. Mitoxantrone for the treatment of advanced breast cancer: Single-agent therapy in previously untreated patients. Eur. J. Cancer Clin. Oncol. 1984, 20, 1141–1146. [Google Scholar] [CrossRef]
  197. Neidhart, J.A.; Gochnour, D.; Roach, R.; Hoth, D.; Young, D. A comparison of mitoxantrone and doxorubicin in breast cancer. J. Clin. Oncol. 1986, 4, 672–677. [Google Scholar] [CrossRef]
  198. Cristofanilli, M.; Holmes, F.A.; Esparza, L.; Valero, V.; Buzdar, A.U.; Neidhart, J.A.; Hortobagyi, G.N. Phase I/II trial of high dose mitoxantrone in metastatic breast cancer: The M.D. Anderson Cancer Center experience. Breast Cancer Res. Treat. 1999, 54, 225–233. [Google Scholar] [CrossRef]
  199. Silver, R.T.; Case, D.C.; Wheeler, R.H.; Miller, T.P.; Stein, R.S.; Stuart, J.J.; Peterson, B.A.; Rivkin, S.E.; Golomb, H.M.; Costanzi. Multicenter clinical trial of mitoxantrone in non-Hodgkin’s lymphoma and Hodgkin’s disease. J. Clin. Oncol. 1991, 9, 754–761. [Google Scholar] [CrossRef]
  200. Bajetta, E.; Buzzoni, R.; Valagussa, P.; Bonadonna, G. Mitoxantrone: An active agent in refractory non-Hodgkin’s lymphomas. Am. J. Clin. Oncol. 1988, 11, 100–103. [Google Scholar] [CrossRef] [PubMed]
  201. Davis, R.B.; Van Echo, D.A.; Leone, L.A.; Henderson, E.S. Phase II trial of mitoxantrone in advanced primary liver cancer: A Cancer and Leukemia Group B Study. Cancer Treat. Rep. 1986, 70, 1125–1126. [Google Scholar] [PubMed]
  202. De Weger, V.A.; De Jonge, M.; Langenberg, M.H.G.; Schellens, J.H.M.; Lolkema, M.; Varga, A.; Demers, B.; Thomas, K.; Hsu, K.; Tuffal, G.; et al. A phase I study of the HDM2 antagonist SAR405838 combined with the MEK inhibitor pimasertib in patients with advanced solid tumours. Br. J. Cancer 2019, 120, 286–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Feltham, R.; Bettjeman, B.; Budhidarmo, R.; Mace, P.D.; Shirley, S.; Condon, S.M.; Chunduru, S.K.; McKinlay, M.A.; Vaux, D.L.; Silke, J.; et al. Smac Mimetics Activate the E3 Ligase Activity of cIAP1 Protein by Promoting RING Domain Dimerization. J. Biol. Chem. 2011, 286, 17015–17028. [Google Scholar] [CrossRef] [Green Version]
  204. Dubrez, L.; Berthelet, J.; Glorian, V. IAP proteins as targets for drug development in oncology. Onco Targets Ther. 2013, 9, 1285–1304. [Google Scholar] [CrossRef] [Green Version]
  205. Allensworth, J.L.; Sauer, S.J.; Lyerly, H.K.; Morse, M.A.; Devi, G.R. Smac mimetic Birinapant induces apoptosis and enhances TRAIL potency in inflammatory breast cancer cells in an IAP-dependent and TNF-α-independent mechanism. Breast Cancer Res. Treat. 2013, 137, 359–371. [Google Scholar] [CrossRef]
  206. Cai, Q.; Sun, H.; Peng, Y.; Lu, J.; Nikolovska-Coleska, Z.; McEachern, D.; Liu, L.; Qiu, S.; Yang, C.-Y.; Miller, R.; et al. A Potent and Orally Active Antagonist (SM-406/AT-406) of Multiple Inhibitor of Apoptosis Proteins (IAPs) in Clinical Development for Cancer Treatment. J. Med. Chem. 2011, 54, 2714–2726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Flygare, J.A.; Beresini, M.; Budha, N.; Chan, H.; Chan, I.T.; Cheeti, S.; Cohen, F.; Deshayes, K.; Doerner, K.; Eckhardt, S.G.; et al. Discovery of a Potent Small-Molecule Antagonist of Inhibitor of Apoptosis (IAP) Proteins and Clinical Candidate for the Treatment of Cancer (GDC-0152). J. Med. Chem. 2012, 55, 4101–4113. [Google Scholar] [CrossRef]
  208. Varfolomeev, E.; Blankenship, J.W.; Wayson, S.M.; Fedorova, A.V.; Kayagaki, N.; Garg, P.; Zobel, K.; Dynek, J.N.; Elliott, L.O.; Wallweber, H.J.; et al. IAP Antagonists Induce Autoubiquitination of c-IAPs, NF-κB Activation, and TNFα-Dependent Apoptosis. Cell 2007, 131, 669–681. [Google Scholar] [CrossRef] [Green Version]
  209. Murai, J.; Yang, K.; Dejsuphong, D.; Hirota, K.; Takeda, S.; D’Andrea, A.D. The USP1/UAF1 Complex Promotes Double-Strand Break Repair through Homologous Recombination. Mol. Cell. Biol. 2011, 31, 2462–2469. [Google Scholar] [CrossRef] [Green Version]
  210. Burkhart, R.A.; Peng, Y.; Norris, Z.A.; Tholey, R.M.; Talbott, V.A.; Liang, Q.; Ai, Y.; Miller, K.; Lal, S.; Cozzitorto, J.A.; et al. Mitoxantrone targets human ubiquitin-specific peptidase 11 (USP11) and is a potent inhibitor of pan-creatic cancer cell survival. Mol. Cancer Res. 2013, 11, 901–911. [Google Scholar] [CrossRef] [Green Version]
  211. Schoenfeld, A.R.; Apgar, S.; Dolios, G.; Wang, R.; Aaronson, S.A. BRCA2 Is Ubiquitinated In Vivo and Interacts with USP11, a Deubiquitinating Enzyme That Exhibits Prosurvival Function in the Cellular Response to DNA Damage. Mol. Cell. Biol. 2004, 24, 7444–7455. [Google Scholar] [CrossRef] [Green Version]
  212. Jacko, A.M.; Nan, L.; Li, S.; Tan, J.; Zhao, J.; Kass, D.J.; Zhao, Y. De-ubiquitinating enzyme, USP11, promotes transforming growth factor β-1 signaling through stabilization of transforming growth factor β receptor II. Cell Death Dis. 2016, 7, e2474. [Google Scholar] [CrossRef] [Green Version]
  213. Chauhan, D.; Tian, Z.; Nicholson, B.; Kumar, G.S.; Zhou, B.; Carrasco, R.; McDermott, J.L.; Leach, C.A.; Fulcinniti, M.; Kodrasov, M.P.; et al. A small molecule inhibitor of ubiquitin-specific protease-7 induces apoptosis in multiple myeloma cells and overcomes bortezomib resistance. Cancer Cell 2012, 22, 345–358. [Google Scholar] [CrossRef] [Green Version]
  214. D’Arcy, P.; Brnjic, S.; Olofsson, M.H.; Fryknäs, M.; Lindsten, K.; De Cesare, M.; Perego, P.; Sadeghi, B.; Hassan, M.; Larsson, R.; et al. Inhibition of proteasome deubiquitinating activity as a new cancer therapy. Nat. Med. 2011, 17, 1636–1640. [Google Scholar] [CrossRef]
  215. Khaminets, A.; Behl, C.; Dikic, I. Ubiquitin-Dependent and Independent Signals In Selective Autophagy. Trends Cell Biol. 2016, 26, 6–16. [Google Scholar] [CrossRef] [PubMed]
  216. Duncan, K.; Schäfer, G.; Vava, A.; Parker, M.I.; Zerbini, L.F. Targeting neddylation in cancer therapy. Future Oncol. 2012, 8, 1461–1470. [Google Scholar] [CrossRef]
  217. Fouladkou, F.; Landry, T.; Kawabe, H.; Neeb, A.; Lu, C.; Brose, N.; Stambolic, V.; Rotin, D. The ubiquitin ligase Nedd4-1 is dispensable for the regulation of PTEN stability and localization. Proc. Natl. Acad. Sci. USA 2008, 105, 8585–8590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Wang, X.; Trotman, L.C.; Koppie, T.; Alimonti, A.; Chen, Z.; Gao, Z.; Wang, J.; Erdjument-Bromage, H.; Tempst, P.; Cordon-Cardo, C.; et al. NEDD4-1 Is a Proto-Oncogenic Ubiquitin Ligase for PTEN. Cell 2007, 128, 129–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Trotman, L.C.; Wang, X.; Alimonti, A.; Chen, Z.; Teruya-Feldstein, J.; Yang, H.; Pavletich, N.P.; Carver, B.S.; Cordon-Cardo, C.; Erdjument-Bromage, H.; et al. Ubiquitination Regulates PTEN Nuclear Import and Tumor Suppression. Cell 2007, 128, 141–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Overview of the ubiquitin proteasome system (UPS). The UPS cascade. Substrate protein is ubiquitinated through the sequential action of three enzymes. E1 binds to activated ubiquitin and is transferred to the ubiquitin-conjugating enzyme (E2). The E2 carries the activated ubiquitin to ubiquitin ligase (E3), which then facilitates the transfer of ubiquitin from E2 to a lysine residue in the target protein. Proteins can be modified with a single mono-ubiquitin molecule, or with ubiquitin chains of different lengths and linkage types. Substrate proteins modified with specific chains are recognized and subsequently degraded by the 26S proteasome. Deubiquitinating enzymes (DUBs) remove ubiquitin from substrate proteins by removing mono-ubiquitination or by trimming or removing the ubiquitin chain. Typically, poly-ubiquitination has been associated with protein clearance through proteasomal degradation while mono-ubiquitination which involves the addition of a single ubiquitin moiety to the substrate protein affects cellular processes.
Figure 1. Overview of the ubiquitin proteasome system (UPS). The UPS cascade. Substrate protein is ubiquitinated through the sequential action of three enzymes. E1 binds to activated ubiquitin and is transferred to the ubiquitin-conjugating enzyme (E2). The E2 carries the activated ubiquitin to ubiquitin ligase (E3), which then facilitates the transfer of ubiquitin from E2 to a lysine residue in the target protein. Proteins can be modified with a single mono-ubiquitin molecule, or with ubiquitin chains of different lengths and linkage types. Substrate proteins modified with specific chains are recognized and subsequently degraded by the 26S proteasome. Deubiquitinating enzymes (DUBs) remove ubiquitin from substrate proteins by removing mono-ubiquitination or by trimming or removing the ubiquitin chain. Typically, poly-ubiquitination has been associated with protein clearance through proteasomal degradation while mono-ubiquitination which involves the addition of a single ubiquitin moiety to the substrate protein affects cellular processes.
Cancers 13 01513 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fhu, C.W.; Ali, A. Dysregulation of the Ubiquitin Proteasome System in Human Malignancies: A Window for Therapeutic Intervention. Cancers 2021, 13, 1513. https://doi.org/10.3390/cancers13071513

AMA Style

Fhu CW, Ali A. Dysregulation of the Ubiquitin Proteasome System in Human Malignancies: A Window for Therapeutic Intervention. Cancers. 2021; 13(7):1513. https://doi.org/10.3390/cancers13071513

Chicago/Turabian Style

Fhu, Chee Wai, and Azhar Ali. 2021. "Dysregulation of the Ubiquitin Proteasome System in Human Malignancies: A Window for Therapeutic Intervention" Cancers 13, no. 7: 1513. https://doi.org/10.3390/cancers13071513

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop