Next Article in Journal
Biological Functions of Hydrogen Sulfide in Plants
Next Article in Special Issue
Irreversible Electroporation in Pancreatic Cancer—An Evolving Experimental and Clinical Method
Previous Article in Journal
Interactions between Major Bioactive Polyphenols of Sugarcane Top: Effects on Human Neural Stem Cell Differentiation and Astrocytic Maturation
Previous Article in Special Issue
Topical Delivery of Hedgehog Inhibitors: Current Status and Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in Anticancer Drug Development Targeting Ubiquitination-Related Factors

College of Life Sciences, Capital Normal University, Beijing 100048, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 15104; https://doi.org/10.3390/ijms232315104
Submission received: 31 August 2022 / Revised: 24 October 2022 / Accepted: 26 October 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Targeted Therapies and Molecular Methods in Cancer)

Abstract

:
Ubiquitination is extensively involved in critical signaling pathways through monitoring protein stability, subcellular localization, and activity. Dysregulation of this process results in severe diseases including malignant cancers. To develop drugs targeting ubiquitination-related factors is a hotspot in research to realize better therapy of human diseases. Ubiquitination comprises three successive reactions mediated by Ub-activating enzyme E1, Ub-conjugating enzyme E2, and Ub ligase E3. As expected, multiple ubiquitination enzymes have been highlighted as targets for anticancer drug development due to their dominant effect on tumorigenesis and cancer progression. In this review, we discuss recent progresses in anticancer drug development targeting enzymatic machinery components.

1. Mechanisms of Ubiquitination

In the process of ubiquitination, ubiquitin (Ub) protein is covalently attached to substrate proteins either as a monomer (monoubiquitination) or as a polymer chain (polyubiquitination). Monoubiquitination simply adds a single Ub to one or multiple lysine (K) residues of substrates. Polyubiquitination modifies proteins with the C-terminal glycine residue of the Ub (Ub-G76), which is linked with one K residue of the proximal Ub via an isopeptide bond. All K residues in Ub, including K6, K11, K27, K29, K33, K48, and K63, can serve as a linkage site. Of note, in some particular cases, Ub can be conjugated to non-K residues, such as methionine-1 (M1), cysteine (C), serine (S), threonine (T), and tyrosine (Y) [1,2,3,4]. A poly-Ub chain is generally homogenous where all the Ub units are linked through the same residue sites. Of note, atypical types of poly-Ub chains are observed in eukaryotic cells [5,6], including mixed and branched poly-Ub chains. In mixed poly-Ub chains, the Ub units employ distinct residues for linkage [7], while branched poly-Ub chains contain a Ub unit providing more than one K residue concurrently for linkage formation (Figure 1) [1,2,3,4]. Upon ubiquitination, protein stability or activity is modified, and the linkage types largely decide the fate of substrates [8]. In general, mono-Ub, K6-, K11-, K29-, and K48-linked poly-Ub chains generally act as a “death label” for proteins and drive them for proteasome or lysosome-mediated degradation [9,10]. K63-linked poly-Ub chains usually regulate the activity and subcellular localization of substrate proteins [11,12,13]. Interestingly, the same Ub linkage could result in distinct outcomes. For instance, the K63-linked poly-Ub chain plays a role of a degradation inducer for octamer-binding transcription factor-4 (Oct4) in mouse embryonic stem cells [14,15,16]. The K11-linked poly-Ub chain does not serve as a “death label” for β-catenin but increases its stability in human colorectal cancer cells [17,18].
Ubiquitination is achieved by three types of enzymes, including the Ub activating enzyme E1, Ub binding enzyme E2, and Ub ligase E3. In this process, E1 produces the Ub–adenosine monophosphate (Ub–AMP) intermediate through ATP-dependent adenylation at Ub-G76. Next, the Ub–AMP attacks the E1 active C site via nucleophilic reaction. A thioester bond forms between the E1 and Ub-G76 accompanying AMP release [19]. Subsequently, E2 mediates a transthiolation reaction to accept Ub at its active C site. At the final step, an E3 ligase simultaneously associates with the E2-Ub and a protein substrate, mediating isopeptide bond formation between the Ub and the substrate [8,20,21]. During this process, E2s largely decide the length and topological structure of Ub chains, while E3s usually specifically recognize substrates through protein–protein interactions [20,22]. In human cells, there are two E1s, 40 E2s, and over 600 E3s [23,24,25]. All E2s are characterized by a ~150 amino acid enzymatic UBC domain with the active site C. Those E2s only possessing the UBC domain in sequence are grouped as class I, while class II and III E2s additionally possess an extended C- and N-terminus, respectively. Class IV E2s contain extension regions at both ends. The extension regions usually regulate subcellular localization of E2s and mediate their interaction with other partners [25,26]. E3s can be divided into three types based on their sequences and Ub transferring mechanisms, including the really interesting new gene (RING)-type, homologous to the E6-AP COOH terminus (HECT)-type and RING between RING (RBR)-type E3s [27]. The RING-type E3s contain a characteristic RING or U-box domain with a similar RING finger fold in the structure. In the process of Ub transfer, the RING-type E3s directly hand over Ub from the E2s to the substrates [27], during which the RING domain, rather than the U-box, displays zinc ions (Zn2+-dependent). To date, more than 500 RING-type E3s have been identified [27]. There are nearly 30 HECT-type E3s encoded in human cells [28]. In addition to the catalytic HECT domain, these E3s usually contain a variable N-terminal domain responsible for substrate recognition [29]. According to the sequences of the N terminus, HECT-type E3s can be further divided into three groups, including the WW domain-containing neuronal precursor cell-expressed developmentally downregulated 4 (Nedd4)/Nedd4-like E3s, HERC (HECT and RCC1-like)- and RLD (RCC1-Like domains)-containing E3s, and the HECT-type E3s without WW and RLD domains [30]. During ubiquitination, HECT E3s achieve Ub transfer through a two-step reaction. Firstly, Ub is transferred from E2 to the catalytic C site of E3 to form a HECT–Ub thioester intermediate. Next, the Ub is ligated to the substrates [27]. In human cells, there are 14 RBR-type E3s encoded, all of which are featured by two conserved Zn2+-binding RING domains, RING1 and RING2 [31]. RBR-type E3s employ similar two-step reactions loading Ubs to substrates despite the absence of HECT domains, which attributes to the catalytic active site of C in RING2 [32].
Ubiquitination is widely involved in various biological processes, such as ribosome function [33], DNA replication [34,35], transcription [36], and inflammatory response [37]. Abnormal ubiquitination of key proteins could cause severe illnesses, such as metabolic disorder, inflammation, degenerative diseases, and malignant cancers [38]. Thus, targeting the ubiquitination process for drug development to achieve improved disease therapy, especially in the anticancer field, has become a research hotspot in recent years.

2. Cancer-Related Ubiquitination Factors

2.1. E1

The ubiquitin proteasome system (UPS) is an important process regulating protein homeostasis. Inhibition of E1 UAE can cause the accumulation of misfolded proteins in the endoplasmic reticulum, thereby destroying the homeostasis of the endoplasmic reticulum and allowing cells to enter the process of apoptosis [39,40]. This strategy appears effective to treat B-cell lymphoma and malignant hematological diseases (Table 1) [39,40]. Furthermore, UAE inhibition can stabilize some tumor suppressors, such as p53, thus delaying the progress of cancer development [41,42]. There is also evidence showing that this inhibition activates the NF-κB signaling pathway so as to suppress malignant tumor growth [43].

2.2. E2-Ubiquitin-Conjugating Enzyme 2C (E2-Ube2c)

Ube2c gene is located on chromosome 20q13.12 encoding a 179 amino acid E2 with a molecular weight of 19 kDa. It cooperates with the anaphase promoting complex/cyclome (APC/C) E3 complex to polyubiquitinate proteins for degradation [44,45]. The activity of Ube2c (Table 1) is involved in cell-cycle regulation and loss of genetic stability [46]. In cancer cells, Ube2c inhibits apoptosis and enhances tumor growth [47]. Ube2c overexpression has been detected in various cancers, such as hepatocellular carcinoma [48], esophageal cancer [49,50], and breast cancers [51,52]. In gastric cancer cells, Ube2c inhibition interrupts cell-cycle progression, while its elevation enhances cancer cell proliferation [47]. In rectal cancers, siRNA-mediated Ube2c depletion induces apoptosis and inhibits cell proliferation, further disturbing cancer growth and invasion [44].

2.3. E3s

2.3.1. Glycoprotein 78 (Gp78)

Gp78 is evolutionarily conserved and encodes a transmembrane glycoprotein possessing the leucine zipper motif and RING domain mediating E3 ligase activity [53,54]. Gp78 is localized on the endoplasmic reticulum membrane and interacts with endoplasmic reticulum-associated degradation (ERAD) E2s via the C terminus of its RING domain [55,56]. Gp78 promotes tumor metastasis and proliferation [57]. In sarcomas, Gp78 interacts with the transmembrane metastasis inhibitor KAI1 (also known as CD82), and inhibition of Gp78 can promote the accumulation of KAI1, lead to apoptosis, and reduce tumor metastasis (Table 1) [58].

2.3.2. Mouse Double Minute 2/Human Homolog of Mouse Double Minute 2 (MDM2/HDM2)

Tumor protein p53 (TP53) (also known as p53) is a tumor suppressor widely involved in critical signaling pathways, including autophagy, aging, differentiation, proliferation, DNA repair, and tumorigenesis [59,60,61,62,63,64]. p53 dysfunction is closely related to human cancers and neurodegenerative diseases [65,66]. Deletion in the p53 gene aggravates deterioration of cancers [67,68,69,70]. RING-type E3 MDM2 binds to p53 through its N terminus to ubiquitinate p53 protein [71,72,73]. MDM2 mediates multiple monoubiquitination of p53 to promote its nuclear export to cytoplasm for proteasomal degradation [74]. MDM2 can also modify p53 via poly-Ub chain for degradation in the nucleus [75,76]. Furthermore, MDM2 binds to the transactivation domain of p53 and inhibits its transcriptional activity [77]. The activity of MDM2 in repressing p53 determines its oncogenetic effect. It has been shown that MDM2 displays an elevated expression in various human cancers, such as osteosarcoma, neuroblastoma, lung cancer, colon cancer, breast cancer, and liver cancer, where it enhances p53 degradation, leading to poor survival and prognosis of patients (Table 1) [78,79,80]. In fact, in addition to MDM2, multiple RING-type E3s are involved in regulating p53 levels, including constitutive photomorphogenesis protein 1 (COP1), p53-induced protein with a RING-H2 domain (Pirh2), and co-chaperone carboxyl terminus Hsp70/90-interacting protein (CHIP). These E3s ubiquitinate p53 to inhibit its function and, thus, promote cancer development [81].

2.3.3. SCFSkp2 Complex

The RING-type E3 complex Skp1–Cul1–Skp2 (SCFSkp2) is composed of multiple subunits including F-box protein S-phase kinase-associated protein 2 (Skp2), Cullin 1 (Cul1), Ring box 1 (Rbx1, also known as Roc1), and Skp1. Cul1 serves as a rigid scaffold to allow assembly of Rbx1 and Skp1 [82,83]. Rbx1 recruits E2 through its RING domain and constitutes the catalytic center with Cul1 to allow direct transfer of Ub to substrates [82]. Skp1 provides a docking site for Skp2. Skp2 specifically recognizes substrates via its F-box domain [84]. Skp2 (Table 1) is oncogenetic. Deficiency in Skp2 inhibits the development of breast cancer [85]. Several substrates have been identified to mediate its roles in tumorigenesis and malignancy. For instance, it ubiquitinates serine/threonine kinase Akt with K63-linked poly-Ub chain in the process of tumorigenesis. Cancers can also occur with enhanced degradation of forkhead box O1 (FOXO1) by Skp2 [86,87]. Transcriptional coactivator yes-associated protein (Yap) promotes tumor growth and metastasis [88]. Deletion of Yap can inhibit the development of breast cancer [89]. Skp2 can promote its activity through ubiquitination with K63-linked poly-Ub chain [90]. MutT homolog-1 (MTH1) is a nucleotide pyrophosphatase that oxidizes dNTPs to protect genomic DNA and prevents reactive oxygen species (ROS)-induced cytotoxicity in cancers [91,92]. In melanoma cells, MTH1 can be regulated by Skp2-mediated K63 polyubiquitination, which stabilizes MTH1 protecting DNA integrity [93].

2.3.4. Inhibitor of Apoptosis-Related Proteins (IAPs)

IAPs are RING-type E3s [94]. The human genome encodes five IAPs proteins in total, which are X-linked inhibitor of apoptosis (XIAP), cellular inhibitor of apoptosis proteins 1 and 2 (cIAP1/2), melanoma-inhibitor of apoptosis (ML-IAP), and inhibitor of apoptosis-like protein-2 (ILP2). In addition to catalytic C-terminal RING domains, their characteristic sequence is the baculoviral IAP repeat (BIR) responsible for protein–protein interaction [95,96,97,98]. IAPs are involved in regulating diverse biological processes, including apoptosis, immune response, cell cycle, and migration [99,100]. Many types of human cancers display overexpressed IAPs that enhance tumor growth. For instance, IAPs (Table 1) repress apoptosis of melanoma cells for uncontrollable growth [101]. In contrast, there is evidence showing a putative role of IAPs as a tumor suppressor since cIAP deficiency constitutively activates NF-κB signaling in malignant hematopoietic cancer cells [102].

2.3.5. WW Domain-Containing E3 Ubiquitin Protein Ligase 2 (Wwp2)

Wwp2 (also known as AIP-2, atrophin-1-interacting protein 2) is an HECT-type Nedd4 family E3 [103]. In addition to the HECT domain at the C terminus, Wwp2 contains a C2 domain at the N terminus and a tryptophan–tryptophan (WW) domain in the middle [104]. The WW region specifically recognizes phosphorylated PPxY motif-containing substrates [105]. One of the dominant substrates of Wwp2 is tumor suppressor phosphatase and tensin homolog (PTEN). PTEN is a lipid phosphatase [106]. It can inhibit tumor occurrence by repressing the PI3K/Akt pathway, and it plays an important role in cell survival and apoptosis regulation [107]. Mutations in PTEN can be observed in breast/prostate cancers and other human diseases [108,109,110]. Wwp2 can bind with the phosphatase domain of PTEN to promote its ubiquitination and degradation. In addition, another HECT-type E3 neural precursor cell expressed developmentally downregulated 4-1 (Nedd4-1) is involved in mediating PTEN polyubiquitination for degradation (Table 1) [104,110,111].
Table 1. Ubiquitination factors involved in cancer regulation.
Table 1. Ubiquitination factors involved in cancer regulation.
Ubiquitination Factors Cancer TypesReferences
UAEB-cell lymphoma and malignant hematological diseases[39,40,43]
Ube2cHepatocellular carcinoma, esophageal cancer, breast cancers, and gastric cancer[47,48,49,50,51,52]
Gp78Sarcomas[58]
MDM2/HDM2Osteosarcoma, neuroblastoma, lung cancer, colon cancer, breast cancer, and liver cancer[78,79,80]
SCFSkp2 complexBreast cancer and melanoma cells[85,86,87,88,89]
IAPsMelanoma cells and malignant hematopoietic cancer[101,102]
Wwp2Breast cancer and prostate cancer[108,109,110]

3. Drug Development Targeting Ubiquitination-Related Factors

Since ubiquitination largely determines the stability and activity of proteins, it could be effective to develop drugs specifically targeting this process so as to achieve a better therapeutic effect. This has been become a hotspot in the field of disease therapy. These drugs can be derived from small molecules, peptides/proteins, or oligonucleotides [112]. The strategy to identify candidate targets for drug development requires combining computer-aided drug design methods and a high-throughput dataset of disease-related transcriptomes, drug responses, protein interactomes, and drug–target networks [113,114].

3.1. Proteinase Inhibitors

To date, three protease inhibitors have been approved by the Federal Drug Administration (FDA), bortezomib, carfilzomib, and ixazomib (Table 2). They are used to treat multiple myeloma [115,116,117]. Peptide borate inhibitor bortezomib blocks the chymotrypsin-like activity of the 26S proteasome to treat multiple myeloma patients [118,119,120,121]. Of note, use of bortezomib alone generally results in severe side-effects, which can be minimized through combining other drugs [122,123]. Bortezomib was also approved by FDA to treat mantle cell lymphoma [124]. Currently, use of bortezomib for other cancer types, such as autoimmune hemolysis and colon cancer, is under clinical trials [125,126].
Tetrapeptide epoxyketone carfilzomib selectively binds to the 26S proteasome to inhibit the activities of chymotrypsin-like protein, leading to accumulation of ubiquitinated substrates. In 2012, carfilzomib was approved by the FDA for clinical therapy of multiple myeloma [115,116,127,128,129]. Carfilzomib is currently at the stage of clinical trial for treatments of multiple cancers including neuroendocrine cancer, kidney cancer, lymphoma, acute myeloid, acute lymphoblastic leukemia, amyloidosis, and small-cell lung cancers [130].
Ixazomib (MLN2238, Ninlaro) is a boric acid derivative that inhibits chymotrypsin-like activity of 20S proteasome [130,131,132]. Compared with bortezomib, ixazomib displays a stronger antitumor effect [132]. Strikingly, treatment combining ixazomib and lenalidomide/dexamethasone can significantly increase the survival rate of myeloma patients [133]. Furthermore, ixazomib has completed a stage I clinical trial for glioblastoma, as well as a stage II clinical trial for malignant myeloid and lymphoid hematological cancers [130].
In addition to bortezomib, carfilzomib, and ixazomib, other protease inhibitors, such as marizomib, oprozomib, and delanzomib, are undergoing clinical trials for multiple myeloma and other cancers [115].

3.2. Drug Development Targeting the E1 Enzyme

Mammal genomes encode two E1s: UAE (also known as UBE1) and UBA6. Tak-243 (also known as MLN7243) is a small-molecule inhibitor specifically targeting UAE. Its application can reduce the overall mono- and polyubiquitination levels of cancer cells, leading to cancer cell death through an impaired cell-cycle process and accumulated DNA damage (Table 2) [134,135]. Tak-243 is currently undergoing clinical trials to treat refractory acute myeloid leukemia, refractory myelodysplastic syndrome, and chronic myelomonocytic leukemia [130].

3.3. Development of Drugs Targeting E3 Ligases

Generally, small0molecule drugs are designed using a lock-and-key mechanism to specifically target disease-related proteins. This strategy usually requires target proteins possessing a proper pocket region in structure as a binding site for small molecules. For those targets lacking unique pockets, two key strategies can be employed. First of all, proteolysis-targeting chimera (PROTAC) technology is a useful platform driving target proteins for degradation. The heterogeneous chimera is mainly composed of two important components, including one small-molecule ligand to specifically bind its target protein and one ligand to recruit an E3 to mediate ubiquitination of the captured protein for degradation [136,137,138,139]. Two examples designed using PROTAC with E3 cereblon (CRBN) as a ligand are Arv-110 and Arv-471 (Table 2) that are currently at the clinical trial stage. Arv-110 targets androgen receptor to treat metastatic castration-resistant prostate cancer (mCRPC), while Arv-471 targets estrogen receptor to treat breast cancer [130,138]. Although PROTAC technology is promising to develop drugs for human diseases, the sizes of molecules designed are always big. Moreover, choices of small-molecule ligands with high specificity in capturing E3 ligases are relatively limited. Another attractive strategy is represented by molecular glue degraders. They are small molecules that can modify molecular surface and, thus, enable novel interactions between targeted proteins and E3s, achieving ubiquitination-mediated protein degradation [140]. Compared with PROTAC, molecular glue degraders are smaller in size and easier to optimize their chemical properties. Several molecular glue degraders have been reported. CC-90009 recruits the CUL4–DDB1–CRBN–RBX1 (CRL4CRBN) E3 complex to ubiquitinate G1-to-S phase transition 1 (GSPT1) for proteasomal degradation. It is currently undergoing a phase I/II clinical trial to treat leukemia [130,141]. Serdemetan (JNJ-26854165) serves as an HDM2 E3 ligase antagonist, which mainly prevents degradation of p53 by inhibiting the interaction of HDM2 with p53. Moreover, serdemetan is capable of inhibiting cholesterol transport. It is under clinical investigation for human cell lymphoma and multiple myoma (Table 2) [130,142,143].

3.4. MDM2 Inhibitors

E3 MDM2 interacts with tumor suppressor p53 to mediate its ubiquitination and proteasome degradation. Targeting the MDM2–p53 complex is an effective modality for cancer treatment [79,144]. A panel of small-molecule inhibitors of MDM2, including RG7112, RG7388, AMG-232, APG-115, BI-907828, CGM097, sirmadlin, milademetan, SAR405838, MK-8242, PRIMA1, and APR-246, is undergoing clinical trials investigating their therapeutic effects on cancers (Table 2) [145,146,147].

3.5. IAPs Inhibitors

RING-type E3 IAPs are involved in cancer development regulation through inhibiting apoptosis and may serve as potential target for cancer immunotherapy [148,149,150]. GDC-0152, LCL161, AT-406, AEG40826, TL-32711, and APG-1387 are drugs targeting IAPs at the clinical stages (Table 2) [151]. These small-molecule drugs enable the interaction of XIAP and cIAP1 to undergo proteasome dependent degradation. Their design is inspired by the natural compound IAP antagonist Smac/DIABLO domain [152]. GDC-0152 is the first IAP antagonist that has been put into clinical trials [153], which is used to treat solid cancers via intravenous injection. As the first oral Smac mimic inhibitor of apoptosis [154], the effects of Lcl161 in treating solid tumors, including small-cell lung cancer, breast cancer, myelofibrosis, and thrombocytosis, are now under clinical investigation [130].

3.6. SCFSkp2 Inhibitors

As discussed above, Skp2 is overexpressed in various cancers and promotes tumor occurrence and malignancy. Skp2 has been suggested as a promising target for anticancer drug development [155,156,157,158]. Currently, a panel of drugs targeting Skp2 is under clinical trials for treatment of acute lymphoblastic leukemia, prostate cancer, prostate adenocarcinoma, and colon cancer. These promising drugs include curcumin, quercetin, lycopene, silibinin, epigalocatechin-3-gallate (EGCG), and vitamin D3 [130,159,160,161,162,163].

3.7. Nedd4-1 Inhibitors

Nedd4 family E3 Nedd4-1 is highly expressed in non-small-cell lung cancer (NSCLC), colon cancer, gastric cancer, and other cancers to promote cancer development [164,165]. To target Nedd4-1 is a potential avenue for the development of novel anticancer drugs [166,167,168]. Indole-3-carbinol (I3C) is derived from Cruciferae (Table 2). It possesses an indole carbinol structure and acts as an inhibitor against Nedd4-1, and it is currently at the clinical trial stage to treat human melanoma and breast cancer [130,169,170].

3.8. Hoil-1-Interacting Protein (HOIP) Inhibitors

RBR-type E3 HOIP is involved in regulating human immune response [171,172]. Bendamustine (Table 2), as an effective drug inhibiting HOIP, has been approved by the FDA and the European Medicines Agency (EMA) for the treatment of multiple myeloma, chronic leukemia, rituximab/refractory follicular, and low-grade lymphoma [173]. Furthermore, it is undergoing clinical trials for other human cancers, including ovarian cancer, metastatic HER2-negative breast cancer, and different types of lymphoma [130,174,175,176].
Table 2. Drugs targeting ubiquitination factors approved or in clinical trials.
Table 2. Drugs targeting ubiquitination factors approved or in clinical trials.
DrugsMechanism and ApplicationReferences
Arv-110As an inhibitor of PROTACs with E3 CRBN as a ligand, Arv-110 targets the androgen receptor to treat mCRPC.[130,138]
Arv-471As an inhibitor of PROTACs with E3 CRBN as a ligand, Arv-471 targets the estrogen receptor to treat breast cancer.[130,138]
Bortezomib1. Bortezomib blocks the chymotrypsin-like activity of the 26S proteasome to treat multiple myeloma patients.
2. Bortezomib is the first drug approved by the FDA for the treatment of multiple myeloma and myeloma.
[118,119,120,121]
Bendamustine1. Bendamustine inhibits RBR-type E3 HOIP.
2. Bendamustine was approved by the FDA for clinical therapy of EMA for the treatment of multiple myeloma, chronic leukemia, rituximab/refractory follicular, and low-grade lymphoma.
[130,174,175,176]
Carfilzomib1. Carfilzomib selectively binds to the 26S proteasome to inhibit the activities of chymotrypsin-like protein, leading to the accumulation of ubiquitinated substrates.
2. Carfilzomib was approved by the FDA for clinical therapy of multiple myeloma.
[115,116,127,128,129]
CC-90009CC-90009 recruits the CRL4CRBN E3 complex to ubiquitinate GSPT1 for proteasomal degradation.[130,141]
Curcumin, quercetin, lycopene, silibinin, EGCG, and vitaminD3These drugs promote apoptosis by downregulating Skp2 levels.[130,159,160,161,162,163]
GDC-0152, LCL161, AT-406, AEG40826, TL-32711, and APG-1387The drugs bind to the BIR3 domain of the IAP proteins through its N terminus, blocking caspase–IAP complex formation and promoting apoptosis.[150,151,152]
Ixazomib1. Ixazomib belongs to boric acid and inhibits chymotrypsin-like activity of 20S proteasome.
2. Ixazomib was approved by the FDA for clinical therapy of multiple myeloma.
[130,131,132]
I3CI3C binds to the HECT domain of Nedd4-1 and inhibits cancer cell proliferation.[130,169,170]
RG7112, RG7388, AMG-232, APG-115, BI-907828, CGM097, sirmadlin, milademetan, SAR405838, MK-8242, PRIMA1, and APR-246These drugs inhibit the interaction of MDM2 with p53 so as to maintain p53 activity.[79,144,145]
Serdemetan (JNJ-26854165)1. Serdemetan prevents degradation of p53 by inhibiting the interaction of HDM2 with p53.
2. Serdemetan is capable of inhibiting cholesterol transport.
[130,142,143]
Tak-243 (MLN7243)Tak-243 is the first UAE inhibitor based on a ubiquitination mechanism to enter the clinic.[134,135]

4. Concluding Remarks and Perspectives

Ubiquitination-related factors are highly involved in regulating cancer development, providing them as promising “druggable” targets for cancer therapy. In spite of striking expectations, inhibitors or agonists against E1/E2/E3 components might possibly result in adverse effects on normal activities in cells since ubiquitination extensively monitors diverse fundamental biological processes. Thus, achieving high specificity in cancer cell targeting is critical to meet this challenge. Innovative solutions could be derived from multidisciplinary integration. For example, pH-response nanomaterials can serve as vectors mediating unique release of inhibitors/agonists against E1/E2/E3 in the cancer-specific pH environment. Furthermore, it will be of great importance to dissect the detailed mechanisms underlying the ubiquitination-related regulation of cancer onset and development. To date, increasing insights and techniques have been developed that would open a new arena for cancer therapy in the future.

Author Contributions

Q.L., figures, manuscript writing and final approval of manuscript; W.Z., conceptualization and design, financial support, manuscript writing, and final approval of manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Our work was supported by the NSFC fund (81872003).

Institutional Review Board Statement

No applicable.

Informed Consent Statement

No applicable.

Data Availability Statement

No applicable.

Acknowledgments

We apologize to the authors whose work could not be cited in this review due to space constraints. We appreciate the great work about ubiquitination from all past and present lab members.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

UbUbiquitin
Ub-G76C-terminal glycine residue of the Ub
M1Methionine-1
Oct4Octamer-binding transcription factor-4
E1Ub activating enzyme
E2Ub binding enzyme
E3Ub ligase
CCysteine
RINGReally interesting new gene
HECTHomologous to E6-AP COOH terminus
RBRRING between RING
Nedd4Neuronal precursor cell-expressed developmentally downregulated 4
HERCHECT and RLD-containing
RLDRCC1-like domains
UPSUbiquitin proteasome system
Ube2cUbiquitin-conjugating enzyme 2C
Gp78Glycoprotein 78
ERADEndoplasmic reticulum-associated degradation
MDM2/HDM2Mouse double minute 2/human homolog of mouse double minute 2
TP53Tumor protein p53
COP1Constitutive photomorphogenesis protein 1
Pirh2P53-induced protein with a RING-H2 domain
CHIPCo-chaperone carboxyl terminus Hsp70/90-interacting protein
Skp2S-phase kinase-associated protein 2
Cul1Cullin 1
Rbx1Ring box 1
Skp1S-phase kinase-associated protein 1
SCFSkp2 complexThe RING-type E3 complex Skp1–Cul1–Skp2
FOXO1Forkhead box O1
YapYes-associated protein
MTH1MutT homolog-1
ROSOxygen species
IAPsInhibitor of apoptosis-related proteins
XIAPX-linked inhibitor of apoptosis
cIAP1/2Cellular inhibitor of apoptosis proteins 1 and 2
ML-IAPMelanoma inhibitor of apoptosis
ILP2Inhibitor of apoptosis-like protein 2
BIRBaculoviral IAP repeat
Wwp2WW domain-containing E3 Ubiquitin protein ligase 2
AIP-2Atrophin-1-interacting protein 2
PTENPhosphatase and tensin homolog
Nedd4-1Neural precursor cell expressed developmentally downregulated-4-1
FDAFederal Drug Administration
PROTACProteolysis-targeting chimera
CRBNCereblon
mCRPCMetastatic castration resistant prostate cancer
CRL4CRBN E3 complexCUL4–DDB1–CRBN–RBX1 E3 complex
GSPT1G1-to S phase transition 1
EGCGEpigalocatechin-3-gallate
NSCLCNon-small-cell lung cancer
I3CIndole-3-carbinol
HOIPHoil-1-interacting protein
EMAEuropean Medicines Agency

References

  1. Hochstrasser, M. Ubiquitin-dependent protein degradation. Annu. Rev. Genet. 1996, 30, 405–439. [Google Scholar] [CrossRef] [PubMed]
  2. Peng, J.; Schwartz, D.; Elias, J.E.; Thoreen, C.C.; Cheng, D.; Marsischky, G.; Roelofs, J.; Finley, D.; Gygi, S.P. A proteomics approach to understanding protein ubiquitination. Nat. Biotechnol. 2003, 21, 921–926. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, P.; Peng, J. Characterization of polyubiquitin chain structure by middle-down mass spectrometry. Anal. Chem. 2008, 80, 3438–3444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Yau, R.; Rape, M. The increasing complexity of the ubiquitin code. Nat. Cell Biol. 2016, 18, 579–586. [Google Scholar] [CrossRef] [PubMed]
  5. Wagner, S.A.; Beli, P.; Weinert, B.T.; Nielsen, M.L.; Cox, J.; Mann, M.; Choudhary, C. A proteome-wide, quantitative survey of in vivo ubiquitylation sites reveals widespread regulatory roles. Mol. Cell Proteom. 2011, 10, M111.013284. [Google Scholar] [CrossRef] [Green Version]
  6. Xu, P.; Duong, D.M.; Seyfried, N.T.; Cheng, D.; Xie, Y.; Robert, J.; Rush, J.; Hochstrasser, M.; Finley, D.; Peng, J. Quantitative proteomics reveals the function of unconventional ubiquitin chains in proteasomal degradation. Cell 2009, 137, 133–145. [Google Scholar] [CrossRef] [Green Version]
  7. Nakasone, M.A.; Livnat-Levanon, N.; Glickman, M.H.; Cohen, R.E.; Fushman, D. Mixed-linkage ubiquitin chains send mixed messages. Structure 2013, 21, 727–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Pickart, C.M. Mechanisms underlying ubiquitination. Annu. Rev. Biochem. 2001, 70, 503–533. [Google Scholar] [CrossRef] [PubMed]
  9. Pickart, C.M.; Fushman, D. Polyubiquitin chains: Polymeric protein signals. Curr. Opin. Chem. Biol. 2004, 8, 610–616. [Google Scholar] [CrossRef]
  10. Livneh, I.; Kravtsova-Ivantsiv, Y.; Braten, O.; Kwon, Y.T.; Ciechanover, A. Monoubiquitination joins polyubiquitination as an esteemed proteasomal targeting signal. Bioessays 2017, 39, 1700027. [Google Scholar] [CrossRef]
  11. Ea, C.K.; Deng, L.; Xia, Z.P.; Pineda, G.; Chen, Z.J. Activation of IKK by TNFalpha requires site-specific ubiquitination of RIP1 and polyubiquitin binding by NEMO. Mol. Cell 2006, 22, 245–257. [Google Scholar] [CrossRef] [PubMed]
  12. Li, L.; Xin, H.; Xu, X.; Huang, M.; Zhang, X.; Chen, Y.; Zhang, S.; Fu, X.Y.; Chang, Z. CHIP mediates degradation of Smad proteins and potentially regulates Smad-induced transcription. Mol. Cell Biol. 2004, 24, 856–864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kanayama, A.; Seth, R.B.; Sun, L.; Ea, C.K.; Hong, M.; Shaito, A.; Chiu, Y.H.; Deng, L.; Chen, Z.J. TAB2 and TAB3 activate the NF-kappaB pathway through binding to polyubiquitin chains. Mol. Cell 2004, 15, 535–548. [Google Scholar] [CrossRef]
  14. Skaug, B.; Jiang, X.; Chen, Z.J. The role of ubiquitin in NF-kappaB regulatory pathways. Annu. Rev. Biochem. 2009, 78, 769–796. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, S.; Ulrich, H.D. Distinct consequences of posttranslational modification by linear versus K63-linked polyubiquitin chains. Proc. Natl. Acad. Sci. USA 2010, 107, 7704–7709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Liao, B.; Jin, Y. Wwp2 mediates Oct4 ubiquitination and its own auto-ubiquitination in a dosage-dependent manner. Cell Res. 2010, 20, 332–344. [Google Scholar] [CrossRef]
  17. Wang, J.; Zhang, Y.; Hou, J.; Qian, X.; Zhang, H.; Zhang, Z.; Li, M.; Wang, R.; Liao, K.; Wang, Y.; et al. Ube2s regulates Sox2 stability and mouse ES cell maintenance. Cell Death Differ. 2016, 23, 393–404. [Google Scholar] [CrossRef] [Green Version]
  18. Li, Z.; Wang, Y.; Li, Y.; Yin, W.; Mo, L.; Qian, X.; Zhang, Y.; Wang, G.; Bu, F.; Zhang, Z.; et al. Ube2s stabilizes β-Catenin through K11-linked polyubiquitination to promote mesendoderm specification and colorectal cancer development. Cell Death Dis. 2018, 9, 456. [Google Scholar] [CrossRef] [Green Version]
  19. Hann, Z.S.; Ji, C.; Olsen, S.K.; Lu, X.; Lux, M.C.; Tan, D.S.; Lima, C.D. Structural basis for adenylation and thioester bond formation in the ubiquitin E1. Proc. Natl. Acad. Sci. USA 2019, 116, 15475–15484. [Google Scholar] [CrossRef] [Green Version]
  20. Nandi, D.; Tahiliani, P.; Kumar, A.; Chandu, D. The ubiquitin-proteasome system. J. Biosci. 2006, 31, 137–155. [Google Scholar] [CrossRef]
  21. Ciechanover, A.; Heller, H.; Katz-Etzion, R.; Hershko, A. Activation of the heat-stable polypeptide of the ATP-dependent proteolytic system. Proc. Natl. Acad. Sci. USA 1981, 78, 761–765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ye, Y.; Rape, M. Building ubiquitin chains: E2 enzymes at work. Nat. Rev. Mol. Cell Biol. 2009, 10, 755–764. [Google Scholar] [CrossRef] [Green Version]
  23. Varshavsky, A. The ubiquitin system, an immense realm. Annu. Rev. Biochem. 2012, 81, 167–176. [Google Scholar] [CrossRef] [PubMed]
  24. Metzger, M.B.; Hristova, V.A.; Weissman, A.M. HECT and RING finger families of E3 ubiquitin ligases at a glance. J. Cell Sci. 2012, 125 Pt 3, 531–537. [Google Scholar] [CrossRef] [Green Version]
  25. 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]
  26. Plafker, S.M.; Plafker, K.S.; Weissman, A.M.; Macara, I.G. Ubiquitin charging of human class III ubiquitin-conjugating enzymes triggers their nuclear import. J. Cell Biol. 2004, 167, 649–659. [Google Scholar] [CrossRef]
  27. Morreale, F.E.; Walden, H. Types of Ubiquitin Ligases. Cell 2016, 165, 248–248.e1. [Google Scholar] [CrossRef]
  28. Scheffner, M.; Staub, O. HECT E3s and human disease. BMC Biochem. 2007, 8 (Suppl. S1), S6. [Google Scholar] [CrossRef] [Green Version]
  29. Deshaies, R.J.; Joazeiro, C.A. RING domain E3 ubiquitin ligases. Annu. Rev. Biochem. 2009, 78, 399–434. [Google Scholar] [CrossRef] [PubMed]
  30. Rotin, D.; Kumar, S. Physiological functions of the HECT family of ubiquitin ligases. Nat. Rev. Mol. Cell Biol. 2009, 10, 398–409. [Google Scholar] [CrossRef]
  31. Marín, I.; Lucas, J.I.; Gradilla, A.C.; Ferrús, A. Parkin and relatives: The RBR family of ubiquitin ligases. Physiol. Genom. 2004, 17, 253–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Dove, K.K.; Klevit, R.E. RING-Between-RING E3 Ligases: Emerging Themes amid the Variations. J. Mol. Biol. 2017, 429, 3363–3375. [Google Scholar] [CrossRef] [PubMed]
  33. Spence, J.; Gali, R.R.; Dittmar, G.; Sherman, F.; Karin, M.; Finley, D. Cell cycle-regulated modification of the ribosome by a variant multiubiquitin chain. Cell 2000, 102, 67–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Spence, J.; Sadis, S.; Haas, A.L.; Finley, D. A ubiquitin mutant with specific defects in DNA repair and multiubiquitination. Mol. Cell Biol. 1995, 15, 1265–1273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hofmann, R.M.; Pickart, C.M. Noncanonical MMS2-encoded ubiquitin-conjugating enzyme functions in assembly of novel polyubiquitin chains for DNA repair. Cell 1999, 96, 645–653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kaiser, P.; Flick, K.; Wittenberg, C.; Reed, S.I. Regulation of transcription by ubiquitination without proteolysis: Cdc34/SCF(Met30)-mediated inactivation of the transcription factor Met4. Cell 2000, 102, 303–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Deng, L.; Wang, C.; Spencer, E.; Yang, L.; Braun, A.; You, J.; Slaughter, C.; Pickart, C.; Chen, Z.J. Activation of the IkappaB kinase complex by TRAF6 requires a dimeric ubiquitin-conjugating enzyme complex and a unique polyubiquitin chain. Cell 2000, 103, 351–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Popovic, D.; Vucic, D.; Dikic, I. Ubiquitination in disease pathogenesis and treatment. Nat. Med. 2014, 20, 1242–1253. [Google Scholar] [CrossRef]
  39. Best, S.; Hashiguchi, T.; Kittai, A.; Bruss, N.; Paiva, C.; Okada, C.; Liu, T.; Berger, A.; Danilov, A.V. Targeting ubiquitin-activating enzyme induces ER stress-mediated apoptosis in B-cell lymphoma cells. Blood Adv. 2019, 3, 51–62. [Google Scholar] [CrossRef] [Green Version]
  40. Xu, G.W.; Ali, M.; Wood, T.E.; Wong, D.; Maclean, N.; Wang, X.; Gronda, M.; Skrtic, M.; Li, X.; Hurren, R.; et al. The ubiquitin-activating enzyme E1 as a therapeutic target for the treatment of leukemia and multiple myeloma. Blood 2010, 115, 2251–2259. [Google Scholar] [CrossRef]
  41. Kitagaki, J.; Yang, Y.; Saavedra, J.E.; Colburn, N.H.; Keefer, L.K.; Perantoni, A.O. Nitric oxide prodrug JS-K inhibits ubiquitin E1 and kills tumor cells retaining wild-type p53. Oncogene 2009, 28, 619–624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Henningsen, K.M.; Manzini, V.; Magerhans, A.; Gerber, S.; Dobbelstein, M. MDM2-Driven Ubiquitination Rapidly Removes p53 from Its Cognate Promoters. Biomolecules 2021, 12, 22. [Google Scholar] [CrossRef] [PubMed]
  43. 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]
  44. Zhang, Y.; Tian, S.; Li, X.; Ji, Y.; Wang, Z.; Liu, C. UBE2C promotes rectal carcinoma via miR-381. Cancer Biol. Ther. 2018, 19, 230–238. [Google Scholar] [CrossRef] [Green Version]
  45. Townsley, F.M.; Aristarkhov, A.; Beck, S.; Hershko, A.; Ruderman, J.V. Dominant-negative cyclin-selective ubiquitin carrier protein E2-C/UbcH10 blocks cells in metaphase. Proc. Natl. Acad. Sci. USA 1997, 94, 2362–2367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Nicolau-Neto, P.; Palumbo, A.; De Martino, M.; Esposito, F.; de Almeida Simão, T.; Fusco, A.; Nasciutti, L.E.; Meireles Da Costa, N.; Ribeiro Pinto, L.F. UBE2C Is a Transcriptional Target of the Cell Cycle Regulator FOXM1. Genes 2018, 9, 188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. 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] [Green Version]
  48. Ieta, K.; Ojima, E.; Tanaka, F.; Nakamura, Y.; Haraguchi, N.; Mimori, K.; Inoue, H.; Kuwano, H.; Mori, M. Identification of overexpressed genes in hepatocellular carcinoma, with special reference to ubiquitin-conjugating enzyme E2C gene expression. Int. J. Cancer 2007, 121, 33–38. [Google Scholar] [CrossRef]
  49. Matsumoto, A.; Ishibashi, Y.; Urashima, M.; Omura, N.; Nakada, K.; Nishikawa, K.; Shida, A.; Takada, K.; Kashiwagi, H.; Yanaga, K. High UBCH10 protein expression as a marker of poor prognosis in esophageal squamous cell carcinoma. Anticancer Res. 2014, 34, 955–961. [Google Scholar] [PubMed]
  50. Li, L.; Li, X.; Wang, W.; Gao, T.; Shi, Z. UBE2C is involved in the functions of ECRG4 on esophageal squamous cell carcinoma. Biomed. Pharmacother. 2018, 98, 201–206. [Google Scholar] [CrossRef]
  51. Qin, T.; Huang, G.; Chi, L.; Sui, S.; Song, C.; Li, N.; Sun, S.; Li, N.; Zhang, M.; Zhao, Z.; et al. Exceptionally high UBE2C expression is a unique phenomenon in basal-like type breast cancer and is regulated by BRCA1. Biomed. Pharmacother. 2017, 95, 649–655. [Google Scholar] [CrossRef] [PubMed]
  52. Mo, C.H.; Gao, L.; Zhu, X.F.; Wei, K.L.; Zeng, J.J.; Chen, G.; Feng, Z.B. The clinicopathological significance of UBE2C in breast cancer: A study based on immunohistochemistry, microarray and RNA-sequencing data. Cancer Cell Int. 2017, 17, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Nabi, I.R.; Watanabe, H.; Raz, A. Identification of B16-F1 melanoma autocrine motility-like factor receptor. Cancer Res. 1990, 50, 409–414. [Google Scholar] [PubMed]
  54. Shimizu, K.; Tani, M.; Watanabe, H.; Nagamachi, Y.; Niinaka, Y.; Shiroishi, T.; Ohwada, S.; Raz, A.; Yokota, J. The autocrine motility factor receptor gene encodes a novel type of seven transmembrane protein. FEBS Lett. 1999, 456, 295–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Fang, S.; Ferrone, M.; Yang, C.; Jensen, J.P.; Tiwari, S.; Weissman, A.M. The tumor autocrine motility factor receptor, gp78, is a ubiquitin protein ligase implicated in degradation from the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 2001, 98, 14422–14427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zhong, X.; Shen, Y.; Ballar, P.; Apostolou, A.; Agami, R.; Fang, S. AAA ATPase p97/valosin-containing protein interacts with gp78, a ubiquitin ligase for endoplasmic reticulum-associated degradation. J. Biol. Chem. 2004, 279, 45676–45684. [Google Scholar] [CrossRef] [Green Version]
  57. Chiu, C.G.; St-Pierre, P.; Nabi, I.R.; Wiseman, S.M. Autocrine motility factor receptor: A clinical review. Expert Rev. Anticancer Ther. 2008, 8, 207–217. [Google Scholar] [CrossRef]
  58. Tsai, Y.C.; Mendoza, A.; Mariano, J.M.; Zhou, M.; Kostova, Z.; Chen, B.; Veenstra, T.; Hewitt, S.M.; Helman, L.J.; Khanna, C.; et al. The ubiquitin ligase gp78 promotes sarcoma metastasis by targeting KAI1 for degradation. Nat. Med. 2007, 13, 1504–1509. [Google Scholar] [CrossRef]
  59. Suzuki, N.; Johmura, Y.; Wang, T.W.; Migita, T.; Wu, W.; Noguchi, R.; Yamaguchi, K.; Furukawa, Y.; Nakamura, S.; Miyoshi, I.; et al. TP53/p53-FBXO22-TFEB controls basal autophagy to govern hormesis. Autophagy 2021, 17, 3776–3793. [Google Scholar] [CrossRef]
  60. Beck, J.; Turnquist, C.; Horikawa, I.; Harris, C. Targeting cellular senescence in cancer and aging: Roles of p53 and its isoforms. Carcinogenesis 2020, 41, 1017–1029. [Google Scholar] [CrossRef]
  61. Synoradzki, K.J.; Bartnik, E.; Czarnecka, A.M.; Fiedorowicz, M.; Firlej, W.; Brodziak, A.; Stasinska, A.; Rutkowski, P.; Grieb, P. TP53 in Biology and Treatment of Osteosarcoma. Cancers 2021, 13, 4284. [Google Scholar] [CrossRef] [PubMed]
  62. Ren, Y.A.; Mullany, L.K.; Liu, Z.; Herron, A.J.; Wong, K.K.; Richards, J.S. Mutant p53 Promotes Epithelial Ovarian Cancer by Regulating Tumor Differentiation, Metastasis, and Responsiveness to Steroid Hormones. Cancer Res. 2016, 76, 2206–2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cho, S.; Kim, M.J.; Choi, Y.Y.; Yoo, S.S.; Lee, W.K.; Lee, E.J.; Jang, E.J.; Bae, E.Y.; Jin, G.; Jeon, H.S.; et al. Associations between polymorphisms in DNA repair genes and TP53 mutations in non-small cell lung cancer. Lung Cancer 2011, 73, 25–31. [Google Scholar] [CrossRef]
  64. Stein, Y.; Rotter, V.; Aloni-Grinstein, R. Gain-of-Function Mutant p53: All the Roads Lead to Tumorigenesis. Int. J. Mol. Sci. 2019, 20, 6197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Baugh, E.H.; Ke, H.; Levine, A.J.; Bonneau, R.A.; Chan, C.S. Why are there hotspot mutations in the TP53 gene in human cancers? Cell Death Differ. 2018, 25, 154–160. [Google Scholar] [CrossRef] [PubMed]
  66. Kanapathipillai, M. Treating p53 Mutant Aggregation-Associated Cancer. Cancers 2018, 10, 154. [Google Scholar] [CrossRef] [Green Version]
  67. Yue, Z.; Zhou, Y.; Zhao, P.; Chen, Y.; Yuan, Y.; Jing, Y.; Wang, X. p53 Deletion promotes myeloma cells invasion by upregulating miR19a/CXCR5. Leuk. Res. 2017, 60, 115–122. [Google Scholar] [CrossRef]
  68. Seifert, H.; Mohr, B.; Thiede, C.; Oelschlägel, U.; Schäkel, U.; Illmer, T.; Soucek, S.; Ehninger, G.; Schaich, M. The prognostic impact of 17p (p53) deletion in 2272 adults with acute myeloid leukemia. Leukemia 2009, 23, 656–663. [Google Scholar] [CrossRef] [Green Version]
  69. Kadioglu, O.; Saeed, M.; Mahmoud, N.; Azawi, S.; Mrasek, K.; Liehr, T.; Efferth, T. Identification of potential novel drug resistance mechanisms by genomic and transcriptomic profiling of colon cancer cells with p53 deletion. Arch. Toxicol. 2021, 95, 959–974. [Google Scholar] [CrossRef]
  70. Jia, Z.; He, J.; Cen, L.; Han, W.; Jiang, N.; Yang, J.; Zhou, M. P53 deletion is independently associated with increased age and decreased survival in a cohort of Chinese patients with diffuse large B-cell lymphoma. Leuk. Lymphoma 2012, 53, 2182–2185. [Google Scholar] [CrossRef]
  71. Maki, C.G.; Huibregtse, J.M.; Howley, P.M. In vivo ubiquitination and proteasome-mediated degradation of p53(1). Cancer Res. 1996, 56, 2649–2654. [Google Scholar] [PubMed]
  72. Haupt, Y.; Maya, R.; Kazaz, A.; Oren, M. Mdm2 promotes the rapid degradation of p53. Nature 1997, 387, 296–299. [Google Scholar] [CrossRef] [PubMed]
  73. Honda, R.; Tanaka, H.; Yasuda, H. Oncoprotein MDM2 is a ubiquitin ligase E3 for tumor suppressor p53. FEBS Lett. 1997, 420, 25–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Lai, Z.; Ferry, K.V.; Diamond, M.A.; Wee, K.E.; Kim, Y.B.; Ma, J.; Yang, T.; Benfield, P.A.; Copeland, R.A.; Auger, K.R. Human mdm2 mediates multiple mono-ubiquitination of p53 by a mechanism requiring enzyme isomerization. J. Biol. Chem. 2001, 276, 31357–31367. [Google Scholar] [CrossRef] [Green Version]
  75. Lohrum, M.A.; Woods, D.B.; Ludwig, R.L.; Bálint, E.; Vousden, K.H. C-terminal ubiquitination of p53 contributes to nuclear export. Mol. Cell Biol. 2001, 21, 8521–8532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Li, M.; Brooks, C.L.; Wu-Baer, F.; Chen, D.; Baer, R.; Gu, W. Mono- versus polyubiquitination: Differential control of p53 fate by Mdm2. Science 2003, 302, 1972–1975. [Google Scholar] [CrossRef] [Green Version]
  77. Kussie, P.H.; Gorina, S.; Marechal, V.; Elenbaas, B.; Moreau, J.; Levine, A.J.; Pavletich, N.P. Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain. Science 1996, 274, 948–953. [Google Scholar] [CrossRef]
  78. Oliner, J.D.; Saiki, A.Y.; Caenepeel, S. The Role of MDM2 Amplification and Overexpression in Tumorigenesis. Cold Spring Harb. Perspect. Med. 2016, 6, a026336. [Google Scholar] [CrossRef] [Green Version]
  79. Zafar, A.; Wang, W.; Liu, G.; Xian, W.; McKeon, F.; Zhou, J.; Zhang, R. Targeting the p53-MDM2 pathway for neuroblastoma therapy: Rays of hope. Cancer Lett. 2021, 496, 16–29. [Google Scholar] [CrossRef]
  80. Hou, H.; Sun, D.; Zhang, X. The role of MDM2 amplification and overexpression in therapeutic resistance of malignant tumors. Cancer Cell Int. 2019, 19, 216. [Google Scholar] [CrossRef]
  81. Sane, S.; Rezvani, K. Essential Roles of E3 Ubiquitin Ligases in p53 Regulation. Int. J. Mol. Sci. 2017, 18, 442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Mason, B.; Laman, H. The FBXL family of F-box proteins: Variations on a theme. Open Biol. 2020, 10, 200319. [Google Scholar] [CrossRef]
  83. Cardozo, T.; Pagano, M. The SCF ubiquitin ligase: Insights into a molecular machine. Nat. Rev. Mol. Cell Biol. 2004, 5, 739–751. [Google Scholar] [CrossRef]
  84. Zheng, N.; Schulman, B.A.; Song, L.; Miller, J.J.; Jeffrey, P.D.; Wang, P.; Chu, C.; Koepp, D.M.; Elledge, S.J.; Pagano, M.; et al. Structure of the Cul1-Rbx1-Skp1-F boxSkp2 SCF ubiquitin ligase complex. Nature 2002, 416, 703–709. [Google Scholar] [CrossRef] [PubMed]
  85. Chan, C.H.; Li, C.F.; Yang, W.L.; Gao, Y.; Lee, S.W.; Feng, Z.; Huang, H.Y.; Tsai, K.K.; Flores, L.G.; Shao, Y.; et al. The Skp2-SCF E3 ligase regulates Akt ubiquitination, glycolysis, herceptin sensitivity, and tumorigenesis. Cell 2012, 149, 1098–1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Huang, H.; Regan, K.M.; Wang, F.; Wang, D.; Smith, D.I.; van Deursen, J.M.; Tindall, D.J. Skp2 inhibits FOXO1 in tumor suppression through ubiquitin-mediated degradation. Proc. Natl. Acad. Sci. USA 2005, 102, 1649–1654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Cai, Z.; Moten, A.; Peng, D.; Hsu, C.C.; Pan, B.S.; Manne, R.; Li, H.Y.; Lin, H.K. The Skp2 Pathway: A Critical Target for Cancer Therapy. Semin. Cancer Biol. 2020, 67 (Pt 2), 16–33. [Google Scholar] [CrossRef] [PubMed]
  88. Lamar, J.M.; Stern, P.; Liu, H.; Schindler, J.W.; Jiang, Z.G.; Hynes, R.O. The Hippo pathway target, YAP, promotes metastasis through its TEAD-interaction domain. Proc. Natl. Acad. Sci. USA 2012, 109, E2441–E2450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Chen, Q.; Zhang, N.; Gray, R.S.; Li, H.; Ewald, A.J.; Zahnow, C.A.; Pan, D. A temporal requirement for Hippo signaling in mammary gland differentiation, growth, and tumorigenesis. Genes Dev. 2014, 28, 432–437. [Google Scholar] [CrossRef] [Green Version]
  90. Yao, F.; Zhou, Z.; Kim, J.; Hang, Q.; Xiao, Z.; Ton, B.N.; Chang, L.; Liu, N.; Zeng, L.; Wang, W.; et al. SKP2- and OTUD1-regulated non-proteolytic ubiquitination of YAP promotes YAP nuclear localization and activity. Nat. Commun. 2018, 9, 2269. [Google Scholar] [CrossRef]
  91. Patel, A.; Burton, D.G.; Halvorsen, K.; Balkan, W.; Reiner, T.; Perez-Stable, C.; Cohen, A.; Munoz, A.; Giribaldi, M.G.; Singh, S.; et al. MutT Homolog 1 (MTH1) maintains multiple KRAS-driven pro-malignant pathways. Oncogene 2015, 34, 2586–2596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Sakumi, K.; Furuichi, M.; Tsuzuki, T.; Kakuma, T.; Kawabata, S.; Maki, H.; Sekiguchi, M. Cloning and expression of cDNA for a human enzyme that hydrolyzes 8-oxo-dGTP, a mutagenic substrate for DNA synthesis. J. Biol. Chem. 1993, 268, 23524–23530. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, J.Y.; Liu, G.Z.; Wilmott, J.S.; La, T.; Feng, Y.C.; Yari, H.; Yan, X.G.; Thorne, R.F.; Scolyer, R.A.; Zhang, X.D.; et al. Skp2-Mediated Stabilization of MTH1 Promotes Survival of Melanoma Cells upon Oxidative Stress. Cancer Res. 2017, 77, 6226–6239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Vaux, D.L.; Silke, J. IAPs, RINGs and ubiquitylation. Nat. Rev. Mol. Cell Biol. 2005, 6, 287–297. [Google Scholar] [CrossRef] [PubMed]
  95. Deveraux, Q.L.; Reed, J.C. IAP family proteins—Suppressors of apoptosis. Genes Dev. 1999, 13, 239–252. [Google Scholar] [CrossRef]
  96. Hinds, M.G.; Norton, R.S.; Vaux, D.L.; Day, C.L. Solution structure of a baculoviral inhibitor of apoptosis (IAP) repeat. Nat. Struct. Biol. 1999, 6, 648–651. [Google Scholar]
  97. Eckelman, B.P.; Drag, M.; Snipas, S.J.; Salvesen, G.S. The mechanism of peptide-binding specificity of IAP BIR domains. Cell Death Differ. 2008, 15, 920–928. [Google Scholar] [CrossRef] [Green Version]
  98. Dubrez, L.; Rajalingam, K. IAPs and cell migration. Semin. Cell Dev. Biol. 2015, 39, 124–131. [Google Scholar] [CrossRef]
  99. Estornes, Y.; Bertrand, M.J. IAPs, regulators of innate immunity and inflammation. Semin. Cell Dev. Biol. 2015, 39, 106–114. [Google Scholar] [CrossRef]
  100. Oberoi-Khanuja, T.K.; Murali, A.; Rajalingam, K. IAPs on the move: Role of inhibitors of apoptosis proteins in cell migration. Cell Death Dis. 2013, 4, e784. [Google Scholar] [CrossRef] [Green Version]
  101. Liang, J.; Zhao, W.; Tong, P.; Li, P.; Zhao, Y.; Li, H.; Liang, J. Comprehensive molecular characterization of inhibitors of apoptosis proteins (IAPs) for therapeutic targeting in cancer. BMC Med. Genom. 2020, 13, 7. [Google Scholar] [CrossRef] [PubMed]
  102. Lau, R.; Pratt, M.A. The opposing roles of cellular inhibitor of apoptosis proteins in cancer. ISRN Oncol. 2012, 2012, 928120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Zhang, R.; Zhang, J.; Luo, W.; Luo, Z.; Shi, S. WWP2 Is One Promising Novel Oncogene. Pathol. Oncol. Res. 2019, 25, 443–446. [Google Scholar] [CrossRef] [PubMed]
  104. Bernassola, F.; Karin, M.; Ciechanover, A.; Melino, G. The HECT family of E3 ubiquitin ligases: Multiple players in cancer development. Cancer Cell 2008, 14, 10–21. [Google Scholar] [CrossRef] [PubMed]
  105. Zou, W.; Chen, X.; Shim, J.H.; Huang, Z.; Brady, N.; Hu, D.; Drapp, R.; Sigrist, K.; Glimcher, L.H.; Jones, D. The E3 ubiquitin ligase Wwp2 regulates craniofacial development through mono-ubiquitylation of Goosecoid. Nat. Cell Biol. 2011, 13, 59–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Myers, M.P.; Pass, I.; Batty, I.H.; Van der Kaay, J.; Stolarov, J.P.; Hemmings, B.A.; Wigler, M.H.; Downes, C.P.; Tonks, N.K. The lipid phosphatase activity of PTEN is critical for its tumor supressor function. Proc. Natl. Acad. Sci. USA 1998, 95, 13513–13518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Carnero, A.; Blanco-Aparicio, C.; Renner, O.; Link, W.; Leal, J.F. The PTEN/PI3K/AKT signalling pathway in cancer, therapeutic implications. Curr. Cancer Drug Targets 2008, 8, 187–198. [Google Scholar] [CrossRef]
  108. Li, J.; Yen, C.; Liaw, D.; Podsypanina, K.; Bose, S.; Wang, S.I.; Puc, J.; Miliaresis, C.; Rodgers, L.; McCombie, R.; et al. PTEN, a putative protein tyrosine phosphatase gene mutated in human brain, breast, and prostate cancer. Science 1997, 275, 1943–1947. [Google Scholar] [CrossRef]
  109. Steck, P.A.; Pershouse, M.A.; Jasser, S.A.; Yung, W.K.; Lin, H.; Ligon, A.H.; Langford, L.A.; Baumgard, M.L.; Hattier, T.; Davis, T.; et al. Identification of a candidate tumour suppressor gene, MMAC1, at chromosome 10q23.3 that is mutated in multiple advanced cancers. Nat. Genet. 1997, 15, 356–362. [Google Scholar] [CrossRef]
  110. Maddika, S.; Kavela, S.; Rani, N.; Palicharla, V.R.; Pokorny, J.L.; Sarkaria, J.N.; Chen, J. WWP2 is an E3 ubiquitin ligase for PTEN. Nat. Cell Biol. 2011, 13, 728–733. [Google Scholar] [CrossRef] [Green Version]
  111. 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]
  112. Ghidini, A.; Cléry, A.; Halloy, F.; Allain, F.H.T.; Hall, J. RNA-PROTACs: Degraders of RNA-Binding Proteins. Angew Chem. Int. Ed. Engl. 2021, 60, 3163–3169. [Google Scholar] [CrossRef] [PubMed]
  113. Li, K.; Du, Y.; Li, L.; Wei, D.Q. Bioinformatics Approaches for Anti-cancer Drug Discovery. Curr. Drug Targets 2020, 21, 3–17. [Google Scholar] [CrossRef]
  114. Berdigaliyev, N.; Aljofan, M. An overview of drug discovery and development. Future Med. Chem. 2020, 12, 939–947. [Google Scholar] [CrossRef] [PubMed]
  115. Fricker, L.D. Proteasome Inhibitor Drugs. Annu. Rev. Pharmacol. Toxicol. 2020, 60, 457–476. [Google Scholar] [CrossRef] [Green Version]
  116. Narayanan, S.; Cai, C.Y.; Assaraf, Y.G.; Guo, H.Q.; Cui, Q.; Wei, L.; Huang, J.J.; Ashby, C.R., Jr.; Chen, Z.S. Targeting the ubiquitin-proteasome pathway to overcome anti-cancer drug resistance. Drug Resist. Updates 2020, 48, 100663. [Google Scholar] [CrossRef]
  117. Jayaweera, S.P.E.; Wanigasinghe Kanakanamge, S.P.; Rajalingam, D.; Silva, G.N. Carfilzomib: A Promising Proteasome Inhibitor for the Treatment of Relapsed and Refractory Multiple Myeloma. Front. Oncol. 2021, 11, 740796. [Google Scholar] [CrossRef]
  118. Kane, R.C.; Bross, P.F.; Farrell, A.T.; Pazdur, R. Velcade: U.S. FDA approval for the treatment of multiple myeloma progressing on prior therapy. Oncologist 2003, 8, 508–513. [Google Scholar] [CrossRef]
  119. Cohen, P.; Tcherpakov, M. Will the ubiquitin system furnish as many drug targets as protein kinases? Cell 2010, 143, 686–693. [Google Scholar] [CrossRef] [Green Version]
  120. Kisselev, A.F.; Callard, A.; Goldberg, A.L. Importance of the different proteolytic sites of the proteasome and the efficacy of inhibitors varies with the protein substrate. J. Biol. Chem. 2006, 281, 8582–8590. [Google Scholar] [CrossRef] [Green Version]
  121. Kane, R.C.; Farrell, A.T.; Sridhara, R.; Pazdur, R. United States Food and Drug Administration approval summary: Bortezomib for the treatment of progressive multiple myeloma after one prior therapy. Clin. Cancer Res. 2006, 12, 2955–2960. [Google Scholar] [CrossRef] [PubMed]
  122. Dimopoulos, M.; Weisel, K.; Moreau, P.; Anderson, L.D., Jr.; White, D.; San-Miguel, J.; Sonneveld, P.; Engelhardt, M.; Jenner, M.; Corso, A.; et al. Pomalidomide, bortezomib, and dexamethasone for multiple myeloma previously treated with lenalidomide (OPTIMISMM): Outcomes by prior treatment at first relapse. Leukemia 2021, 35, 1722–1731. [Google Scholar] [CrossRef] [PubMed]
  123. Yamamoto, S.; Egashira, N. Pathological Mechanisms of Bortezomib-Induced Peripheral Neuropathy. Int. J. Mol. Sci. 2021, 22, 888. [Google Scholar] [CrossRef] [PubMed]
  124. Cheah, C.Y.; Seymour, J.F.; Wang, M.L. Mantle Cell Lymphoma. J. Clin. Oncol. 2016, 34, 1256–1269. [Google Scholar] [CrossRef]
  125. Pasquale, R.; Giannotta, J.A.; Barcellini, W.; Fattizzo, B. Bortezomib in autoimmune hemolytic anemia and beyond. Ther. Adv. Hematol. 2021, 12, 20406207211046428. [Google Scholar] [CrossRef]
  126. Härtel, H.; Theiß, J.; Abdelaziz, M.O.; Raftery, M.J.; Pecher, G.; Bogner, E. HCMV-Mediated Interference of Bortezomib-Induced Apoptosis in Colon Carcinoma Cell Line Caco-2. Viruses 2021, 13, 83. [Google Scholar] [CrossRef]
  127. Kortuem, K.M.; Stewart, A.K. Carfilzomib. Blood 2013, 121, 893–897. [Google Scholar] [CrossRef]
  128. 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]
  129. US FDA. Available online: https://www.accessdata.fda.gov/scripts/cder/daf/index.cfm (accessed on 25 October 2022).
  130. Available online: https://clinicaltrials.gov (accessed on 25 October 2022).
  131. Augello, G.; Modica, M.; Azzolina, A.; Puleio, R.; Cassata, G.; Emma, M.R.; Di Sano, C.; Cusimano, A.; Montalto, G.; Cervello, M. Preclinical evaluation of antitumor activity of the proteasome inhibitor MLN2238 (ixazomib) in hepatocellular carcinoma cells. Cell Death Dis. 2018, 9, 28. [Google Scholar] [CrossRef] [Green Version]
  132. Kupperman, E.; Lee, E.C.; Cao, Y.; Bannerman, B.; Fitzgerald, M.; Berger, A.; Yu, J.; Yang, Y.; Hales, P.; Bruzzese, F.; et al. Evaluation of the proteasome inhibitor MLN9708 in preclinical models of human cancer. Cancer Res. 2010, 70, 1970–1980. [Google Scholar] [CrossRef] [Green Version]
  133. Moreau, P.; Masszi, T.; Grzasko, N.; Bahlis, N.J.; Hansson, M.; Pour, L.; Sandhu, I.; Ganly, P.; Baker, B.W.; Jackson, S.R.; et al. Oral Ixazomib, Lenalidomide, and Dexamethasone for Multiple Myeloma. N. Engl. J. Med. 2016, 374, 1621–1634. [Google Scholar] [CrossRef] [PubMed]
  134. Jin, J.; Li, X.; Gygi, S.P.; Harper, J.W. Dual E1 activation systems for ubiquitin differentially regulate E2 enzyme charging. Nature 2007, 447, 1135–1138. [Google Scholar] [CrossRef] [PubMed]
  135. 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] [PubMed]
  136. Schneekloth, J.S., Jr.; Fonseca, F.N.; Koldobskiy, M.; Mandal, A.; Deshaies, R.; Sakamoto, K.; Crews, C.M. Chemical genetic control of protein levels: Selective in vivo targeted degradation. J. Am. Chem. Soc. 2004, 126, 3748–3754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Weng, G.; Shen, C.; Cao, D.; Gao, J.; Dong, X.; He, Q.; Yang, B.; Li, D.; Wu, J.; Hou, T. PROTAC-DB: An online database of PROTACs. Nucleic Acids Res. 2021, 49, D1381–D1387. [Google Scholar] [CrossRef]
  138. Nalawansha, D.A.; Crews, C.M. PROTACs: An Emerging Therapeutic Modality in Precision Medicine. Cell Chem. Biol. 2020, 27, 998–1014. [Google Scholar] [CrossRef]
  139. Sakamoto, K.M.; Kim, K.B.; Kumagai, A.; Mercurio, F.; Crews, C.M.; Deshaies, R.J. Protacs: Chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proc. Natl. Acad. Sci. USA 2001, 98, 8554–8559. [Google Scholar] [CrossRef] [Green Version]
  140. Schapira, M.; Calabrese, M.F.; Bullock, A.N.; Crews, C.M. Targeted protein degradation: Expanding the toolbox. Nat. Rev. Drug Discov. 2019, 18, 949–963. [Google Scholar] [CrossRef]
  141. Surka, C.; Jin, L.; Mbong, N.; Lu, C.C.; Jang, I.S.; Rychak, E.; Mendy, D.; Clayton, T.; Tindall, E.; Hsu, C.; et al. CC-90009, a novel cereblon E3 ligase modulator, targets acute myeloid leukemia blasts and leukemia stem cells. Blood 2021, 137, 661–677. [Google Scholar] [CrossRef]
  142. Jones, R.J.; Gu, D.; Bjorklund, C.C.; Kuiatse, I.; Remaley, A.T.; Bashir, T.; Vreys, V.; Orlowski, R.Z. The novel anticancer agent JNJ-26854165 induces cell death through inhibition of cholesterol transport and degradation of ABCA1. J. Pharmacol. Exp. Ther. 2013, 346, 381–392. [Google Scholar] [CrossRef] [Green Version]
  143. Chargari, C.; Leteur, C.; Angevin, E.; Bashir, T.; Schoentjes, B.; Arts, J.; Janicot, M.; Bourhis, J.; Deutsch, E. Preclinical assessment of JNJ-26854165 (Serdemetan), a novel tryptamine compound with radiosensitizing activity in vitro and in tumor xenografts. Cancer Lett. 2011, 312, 209–218. [Google Scholar] [CrossRef]
  144. Klein, A.M.; de Queiroz, R.M.; Venkatesh, D.; Prives, C. The roles and regulation of MDM2 and MDMX: It is not just about p53. Genes Dev. 2021, 35, 575–601. [Google Scholar] [CrossRef] [PubMed]
  145. Konopleva, M.; Martinelli, G.; Daver, N.; Papayannidis, C.; Wei, A.; Higgins, B.; Ott, M.; Mascarenhas, J.; Andreeff, M. MDM2 inhibition: An important step forward in cancer therapy. Leukemia 2020, 34, 2858–2874. [Google Scholar] [CrossRef] [PubMed]
  146. Fang, Y.; Liao, G.; Yu, B. Small-molecule MDM2/X inhibitors and PROTAC degraders for cancer therapy: Advances and perspectives. Acta Pharm. Sin. B 2020, 10, 1253–1278. [Google Scholar] [CrossRef] [PubMed]
  147. LaPlante, G.; Zhang, W. Targeting the Ubiquitin-Proteasome System for Cancer Therapeutics by Small-Molecule Inhibitors. Cancers 2021, 13, 3079. [Google Scholar] [CrossRef]
  148. Kumar, S.; Fairmichael, C.; Longley, D.B.; Turkington, R.C. The Multiple Roles of the IAP Super-family in cancer. Pharmacol. Ther. 2020, 214, 107610. [Google Scholar] [CrossRef]
  149. Feltham, R.; Khan, N.; Silke, J. IAPS and ubiquitylation. IUBMB Life 2012, 64, 411–418. [Google Scholar] [CrossRef]
  150. Michie, J.; Kearney, C.J.; Hawkins, E.D.; Silke, J.; Oliaro, J. The Immuno-Modulatory Effects of Inhibitor of Apoptosis Protein Antagonists in Cancer Immunotherapy. Cells 2020, 9, 207. [Google Scholar] [CrossRef] [Green Version]
  151. Ye, P.; Chi, X.; Cha, J.H.; Luo, S.; Yang, G.; Yan, X.; Yang, W.H. Potential of E3 Ubiquitin Ligases in Cancer Immunity: Opportunities and Challenges. Cells 2021, 10, 3309. [Google Scholar] [CrossRef]
  152. Cossu, F.; Milani, M.; Mastrangelo, E.; Lecis, D. Targeting the BIR Domains of Inhibitor of Apoptosis (IAP) Proteins in Cancer Treatment. Comput. Struct. Biotechnol. J. 2019, 17, 142–150. [Google Scholar] [CrossRef]
  153. 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] [PubMed] [Green Version]
  154. 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]
  155. Wang, Z.; Fukushima, H.; Inuzuka, H.; Wan, L.; Liu, P.; Gao, D.; Sarkar, F.H.; Wei, W. Skp2 is a promising therapeutic target in breast cancer. Front. Oncol. 2012, 1, 57. [Google Scholar] [CrossRef] [Green Version]
  156. Kitagawa, M.; Lee, S.H.; McCormick, F. Skp2 suppresses p53-dependent apoptosis by inhibiting p300. Mol. Cell 2008, 29, 217–231. [Google Scholar] [CrossRef] [PubMed]
  157. Chan, C.H.; Morrow, J.K.; Li, C.F.; Gao, Y.; Jin, G.; Moten, A.; Stagg, L.J.; Ladbury, J.E.; Cai, Z.; Xu, D.; et al. Pharmacological inactivation of Skp2 SCF ubiquitin ligase restricts cancer stem cell traits and cancer progression. Cell 2013, 154, 556–568. [Google Scholar] [CrossRef] [Green Version]
  158. Asmamaw, M.D.; Liu, Y.; Zheng, Y.C.; Shi, X.J.; Liu, H.M. Skp2 in the ubiquitin-proteasome system: A comprehensive review. Med. Res. Rev. 2020, 40, 1920–1949. [Google Scholar] [CrossRef]
  159. Su, J.; Zhou, X.; Wang, L.; Yin, X.; Wang, Z. Curcumin inhibits cell growth and invasion and induces apoptosis through down-regulation of Skp2 in pancreatic cancer cells. Am. J. Cancer Res. 2016, 6, 1949–1962. [Google Scholar]
  160. Huang, H.C.; Lin, C.L.; Lin, J.K. 1,2,3,4,6-penta-O-galloyl-β-D-glucose, quercetin, curcumin and lycopene induce cell-cycle arrest in MDA-MB-231 and BT474 cells through downregulation of Skp2 protein. J. Agric. Food Chem. 2011, 59, 6765–6775. [Google Scholar] [CrossRef]
  161. Roy, S.; Kaur, M.; Agarwal, C.; Tecklenburg, M.; Sclafani, R.A.; Agarwal, R. p21 and p27 induction by silibinin is essential for its cell cycle arrest effect in prostate carcinoma cells. Mol. Cancer Ther. 2007, 6, 2696–2707. [Google Scholar] [CrossRef] [Green Version]
  162. Huang, H.C.; Way, T.D.; Lin, C.L.; Lin, J.K. EGCG stabilizes p27kip1 in E2-stimulated MCF-7 cells through down-regulation of the Skp2 protein. Endocrinology 2008, 149, 5972–5983. [Google Scholar] [CrossRef]
  163. Liu, W.; Asa, S.L.; Fantus, I.G.; Walfish, P.G.; Ezzat, S. Vitamin D arrests thyroid carcinoma cell growth and induces p27 dephosphorylation and accumulation through PTEN/akt-dependent and -independent pathways. Am. J. Pathol. 2002, 160, 511–519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Amodio, N.; Scrima, M.; Palaia, L.; Salman, A.N.; Quintiero, A.; Franco, R.; Botti, G.; Pirozzi, P.; Rocco, G.; De Rosa, N.; et al. Oncogenic role of the E3 ubiquitin ligase NEDD4-1, a PTEN negative regulator, in non-small-cell lung carcinomas. Am. J. Pathol. 2010, 177, 2622–2634. [Google Scholar] [CrossRef]
  165. Eide, P.W.; Cekaite, L.; Danielsen, S.A.; Eilertsen, I.A.; Kjenseth, A.; Fykerud, T.A.; Ågesen, T.H.; Bruun, J.; Rivedal, E.; Lothe, R.A.; et al. NEDD4 is overexpressed in colorectal cancer and promotes colonic cell growth independently of the PI3K/PTEN/AKT pathway. Cell Signal 2013, 25, 12–18. [Google Scholar] [CrossRef] [PubMed]
  166. Zeng, T.; Wang, Q.; Fu, J.; Lin, Q.; Bi, J.; Ding, W.; Qiao, Y.; Zhang, S.; Zhao, W.; Lin, H.; et al. Impeded Nedd4-1-mediated Ras degradation underlies Ras-driven tumorigenesis. Cell Rep. 2014, 7, 871–882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Lu, C.; Thoeni, C.; Connor, A.; Kawabe, H.; Gallinger, S.; Rotin, D. Intestinal knockout of Nedd4 enhances growth of Apc(min) tumors. Oncogene 2016, 35, 5839–5849. [Google Scholar] [CrossRef] [PubMed]
  168. Ye, X.; Wang, L.; Shang, B.; Wang, Z.; Wei, W. NEDD4: A promising target for cancer therapy. Curr. Cancer Drug Targets 2014, 14, 549–556. [Google Scholar] [CrossRef] [PubMed]
  169. Kundu, A.; Quirit, J.G.; Khouri, M.G.; Firestone, G.L. Inhibition of oncogenic BRAF activity by indole-3-carbinol disrupts microphthalmia-associated transcription factor expression and arrests melanoma cell proliferation. Mol. Carcinog. 2017, 56, 49–61. [Google Scholar] [CrossRef] [Green Version]
  170. Quirit, J.G.; Lavrenov, S.N.; Poindexter, K.; Xu, J.; Kyauk, C.; Durkin, K.A.; Aronchik, I.; Tomasiak, T.; Solomatin, Y.A.; Preobrazhenskaya, M.N.; et al. Indole-3-carbinol (I3C) analogues are potent small molecule inhibitors of NEDD4-1 ubiquitin ligase activity that disrupt proliferation of human melanoma cells. Biochem. Pharmacol. 2017, 127, 13–27. [Google Scholar] [CrossRef]
  171. Smit, J.J.; Monteferrario, D.; Noordermeer, S.M.; van Dijk, W.J.; van der Reijden, B.A.; Sixma, T.K. The E3 ligase HOIP specifies linear ubiquitin chain assembly through its RING-IBR-RING domain and the unique LDD extension. Embo. J. 2012, 31, 3833–3844. [Google Scholar] [CrossRef] [Green Version]
  172. Iwai, K.; Fujita, H.; Sasaki, Y. Linear ubiquitin chains: NF-κB signalling, cell death and beyond. Nat. Rev. Mol. Cell Biol. 2014, 15, 503–508. [Google Scholar] [CrossRef] [PubMed]
  173. Derenzini, E.; Zinzani, P.L.; Cheson, B.D. Bendamustine: Role and evidence in lymphoma therapy, an overview. Leuk. Lymphoma 2014, 55, 1471–1478. [Google Scholar] [CrossRef] [PubMed]
  174. Damaj, G.; Gressin, R.; Bouabdallah, K.; Cartron, G.; Choufi, B.; Gyan, E.; Banos, A.; Jaccard, A.; Park, S.; Tournilhac, O.; et al. Results from a prospective, open-label, phase II trial of bendamustine in refractory or relapsed T-cell lymphomas: The BENTLY trial. J. Clin. Oncol. 2013, 31, 104–110. [Google Scholar] [CrossRef] [PubMed]
  175. Santoro, A.; Mazza, R.; Pulsoni, A.; Re, A.; Bonfichi, M.; Zilioli, V.R.; Salvi, F.; Merli, F.; Anastasia, A.; Luminari, S.; et al. Bendamustine in Combination with Gemcitabine and Vinorelbine Is an Effective Regimen As Induction Chemotherapy before Autologous Stem-Cell Transplantation for Relapsed or Refractory Hodgkin Lymphoma: Final Results of a Multicenter Phase II Study. J. Clin. Oncol. 2016, 34, 3293–3299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Weidmann, E.; Kim, S.Z.; Rost, A.; Schuppert, H.; Seipelt, G.; Hoelzer, D.; Mitrou, P.S. Bendamustine is effective in relapsed or refractory aggressive non-Hodgkin’s lymphoma. Ann. Oncol. 2002, 13, 1285–1289. [Google Scholar] [CrossRef]
Figure 1. The ubiquitination machinery. Ubiquitination is achieved through the sequential activity of E1, E2, and E3. RING-type E3s can mediate a direct ligation of Ub from E2s to substrates, whereas HECT-/RBR-type E3s require two-step reactions. Typical types of Ub linkages include MonoUb, multi-MonoUb, poly-Ub, and branched poly-Ub.
Figure 1. The ubiquitination machinery. Ubiquitination is achieved through the sequential activity of E1, E2, and E3. RING-type E3s can mediate a direct ligation of Ub from E2s to substrates, whereas HECT-/RBR-type E3s require two-step reactions. Typical types of Ub linkages include MonoUb, multi-MonoUb, poly-Ub, and branched poly-Ub.
Ijms 23 15104 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

Li, Q.; Zhang, W. Progress in Anticancer Drug Development Targeting Ubiquitination-Related Factors. Int. J. Mol. Sci. 2022, 23, 15104. https://doi.org/10.3390/ijms232315104

AMA Style

Li Q, Zhang W. Progress in Anticancer Drug Development Targeting Ubiquitination-Related Factors. International Journal of Molecular Sciences. 2022; 23(23):15104. https://doi.org/10.3390/ijms232315104

Chicago/Turabian Style

Li, Qianqian, and Weiwei Zhang. 2022. "Progress in Anticancer Drug Development Targeting Ubiquitination-Related Factors" International Journal of Molecular Sciences 23, no. 23: 15104. https://doi.org/10.3390/ijms232315104

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