Next Article in Journal
Facile Synthesis, Characterization, and Antimicrobial Assessment of a Silver/Montmorillonite Nanocomposite as an Effective Antiseptic against Foodborne Pathogens for Promising Food Protection
Next Article in Special Issue
A First-Class Degrader Candidate Targeting Both KRAS G12D and G12V Mediated by CANDDY Technology Independent of Ubiquitination
Previous Article in Journal
Size–Pore-Dependent Methanol Sequestration from Water–Methanol Mixtures by an Embedded Graphene Slit
Previous Article in Special Issue
PROTACs in Epigenetic Cancer Therapy: Current Status and Future Opportunities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

PROTACs in the Management of Prostate Cancer

by
Poornachandra Yedla
1,*,
Ahmed O. Babalghith
2,*,
Vindhya Vasini Andra
3 and
Riyaz Syed
4
1
Department of Pharmacogenomics, Institute of Translational Research, Asian Healthcare Foundation, Asian Institute of Gastroenterology Hospitals, Gachibowli, Hyderabad 500082, India
2
Department of Medical Genetics, Faculty of Medicine, Umm Al-Qura University, Makkah 21955, Saudi Arabia
3
Department of Medical Oncology, Omega Hospitals, Gachibowli, Hyderabad 500032, India
4
Department of Chemiinformatics, Centella Scientific, JHUB, Jawaharlal Nehru Technological University, Hyderabad 500085, India
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(9), 3698; https://doi.org/10.3390/molecules28093698
Submission received: 8 March 2023 / Revised: 15 April 2023 / Accepted: 17 April 2023 / Published: 25 April 2023

Abstract

:
Cancer treatments with targeted therapy have gained immense interest due to their low levels of toxicity and high selectivity. Proteolysis-Targeting Chimeras (PROTACs) have drawn special attention in the development of cancer therapeutics owing to their unique mechanism of action, their ability to target undruggable proteins, and their focused target engagement. PROTACs selectively degrade the target protein through the ubiquitin–proteasome system, which describes a different mode of action compared to conventional small-molecule inhibitors or even antibodies. Among different cancer types, prostate cancer (PC) is the most prevalent non-cutaneous cancer in men. Genetic alterations and the overexpression of several genes, such as FOXA1, AR, PTEN, RB1, TP53, etc., suppress the immune response, resulting in drug resistance to conventional drugs in prostate cancer. Since the progression of ARV-110 (PROTAC for PC) into clinical phases, the focus of research has quickly shifted to protein degraders targeting prostate cancer. The present review highlights an overview of PROTACs in prostate cancer and their superiority over conventional inhibitors. We also delve into the underlying pathophysiology of the disease and explain the structural design and linkerology strategies for PROTAC molecules. Additionally, we touch on the various targets for PROTAC in prostate cancer, including the androgen receptor (AR) and other critical oncoproteins, and discuss the future prospects and challenges in this field.

1. Introduction

Targeted therapies work to treat cancer by impeding certain proteins that promote tumor growth and metastasis throughout the body. These therapies have made significant strides in the treatment of cancer over the past few decades and are now an effective option for cancer patients [1]. As a result, the development of small-molecule inhibitors (SMIs) enables effective therapeutic approaches to target overexpressed proteins in several cancers. However, due to the poor therapeutic efficacy, genetic alterations/mutations in the target protein, drug resistance, and off-target effects of these SMIs treatments, researchers are currently looking for more effective and precise approaches to target the oncoproteins that are associated with cancer progression [2]. One such approach is Proteolysis-Targeting Chimeras (PROTACs), which overcomes the constraints of conventional SMIs and selectively degrades the target protein. Instead of altering the protein’s function by dwelling in an active site, PROTACs use the cell’s own degradation machinery to degrade the target protein [3]. This innovative approach not only makes it possible to create more effective and targeted therapies for cancer and other intractable diseases, but it also opens up hitherto “undruggable” regions of the proteome [4,5]. PROTACs are now acknowledged as a novel approach within drug development and they have the potential to become the next generation of anticancer therapeutics.
Cancer is one of the most intricate and rapidly propagating diseases of the present century with a high mortality rate. Among several cancers, prostate cancer (PC) is the most prevalent urogenital malignancy in men, characterized by aberrant cell division that causes the prostate gland to develop abnormally [6]. The primary cause of PC mortality is metastatic prostate cancer, where cancer cells have metastasized to other regions of the body such as the bone, pelvis, lumbar vertebra, bladder, rectum, and brain [7]. Approximately half (48%) of all incidents of cancer in males are prostate, lung, bronchus, and colorectal cancers; PC alone accounts for 27% of diagnoses. In 2020, there were 1,414,249 newly diagnosed cases and 375,000 deaths globally because of PC, and in 2022, PC cases were projected to total 268,490 cases (14% of all new cancer cases) and 34,500 deaths alone in the United States [8,9]. Somatic and germline genetic aberrations, increased AR expression, gene amplification, activation of CYP17A1 to increase androgen production, and the emergence of AR splice variants are promising contributing factors for PC progression [10]. The current treatment for early stage prostate cancer is androgen deprivation therapy (ADT), which can be achieved either surgically or medically by castration using anti-androgens or luteinizing hormone-releasing hormone (LHRH) agonists or antagonists. Even though ADT results in remissions that last around two to three years, the condition eventually advances to castration-resistant prostate cancer (CRPC), which has a dismal prognosis and significant treatment complications [11].
With the burgeoning clinical requirement and several limitations of SMIs of PC, there has been considerable development in the discovery of anticancer PROTACs that target the various oncoproteins that are involved in disease progression. Due to its distinctive mode of action, PROTAC technology may compensate for the shortcomings of conventional drug therapy, which accelerates its development. Further, PROTAC is a highly promising technology in terms of its clinical applications, since ARV-110, ARV-766, and AR-LDD, which are some of the AR-targeting PROTACs, are in clinical trials for the treatment of PC [12]. In this review, we introduce the principles and development of prostate cancer-targeted PROTAC technology and summarize the application of PROTACs in targeting the crucial proteins that are involved in tumorigenesis. In addition, we summarized the genetic abnormalities (somatic and germline mutations) of different oncoproteins associated with PC development, which may help in the expansion of precision medicine for PC in the near future.

2. Ubiquitination and Proteolysis-Targeting Mechanism

Protein degradation can take on many different forms and is critical for regulating various cellular and biological processes. The ubiquitin–proteasome system (UPS) is responsible for the degradation of most proteins in eukaryotic cells. This pathway involves the covalent attachment of the highly conserved 76 amino acid protein ubiquitin onto target proteins prior to their recognition and degradation by the 26S proteasome [13,14]. The cellular protein degradation process is necessary for optimal cellular activity, including proliferation, differentiation, and cell death. One such important pathway for post-translational protein regulation is ubiquitin-dependent proteolysis. Ubiquitin-activating enzymes (E1), ubiquitin-conjugating enzymes (E2), and ubiquitin–protein ligases (E3) are the three vital enzymes in the ubiquitination process that catalyze the attachment of ubiquitin to lysine residues in substrate proteins and promote degradation. E3 ligases cause the association of this ubiquitin with the Protein Of Interest (POI), which acts as a signal for proteolytic enzymes to start breaking down the proteins (Figure 1) [15]. On the basis of this ubiquitin–proteasome system, the PROTAC (proteolysis-targeting chimaeras) technology was developed for target protein degradation [16].
PROTACs are heterobifunctional molecules with the following three structural features: (i) a ligand that binds to the target protein, (ii) a ligand that recruits an E3 ligase, and (iii) a suitable linker that connects these two ligands. This structure allows the POI and E3 ligase to become adjacent and accessible for conjugation as the PROTAC molecule becomes sandwiched between these two proteins. Later, the E3 ligase complex ubiquitinates the POI and leads to its identification and degradation by the 26S proteasome, which is a component of the ubiquitin–proteasome system in eukaryotes. A single PROTAC molecule may carry out this catalytic process repeatedly, permitting several turnovers of ubiquitylation events that result in the degradation of the recruited substrate [17].
Three different steps are involved in the ubiquitin–proteasome system. In the initial step, the E1 ubiquitin-activating enzyme uses ATP to activate ubiquitin, resulting in an activated ubiquitin-adenylate. This intermediate is then transformed into a thioester by joining with an E1 enzyme’s catalytic cysteine. In the second step, the activated ubiquitin is transported through a transthioesterification reaction to the catalytic cysteine of an E2 ubiquitin-conjugating enzyme. The carboxy terminus of ubiquitin then forms an isopeptide bond with the side chain of a lysine residue on the surface of the target protein after the formation of a ternary complex between the E2 enzyme, an E3 ubiquitin ligase, and the substrate protein. The ubiquitin protein itself contains seven lysine residues, which provides options for other ubiquitins to bind and initiate polyubiquitination. The fate of the ubiquitinated substrate protein depends on the fashion in which the polyubiquitin chain has formed. Ubiquitin chain forming via linkage between k48/k11 lysine residues of ubiquitin protein usually results in proteolysis. This ubiquitin chain is eventually recognized by the 26s proteasome system, which degrades the substrate protein thereafter [17,18].

2.1. Types of Targeted Protein Degraders

In addition to PROTACs, there are various modalities that can degrade the targeted protein via proteasomal degradation, such as molecular glues [19], double-mechanism degraders [20], selective AR degraders (SARD) [21], hydrophobic tagging technologies A(HyT) [22], the transcription factor PROTAC (TF-PROTAC) [23], the dual-PROTAC [24], and the selective estrogen receptor degrader (SERD) [25]. Apart from proteasomal degradation, TPD via lysosome is also being employed by researchers. This pathway can go via two ways, namely the endosome–lysosome system and the autophagy–lysosome system. The drugs that target the endosome–lysosome system include the lysosome-targeting chimera (LYTAC) [26], bispecific aptamer chimera [27], antibody-based PROTAC (AbTAC) [28], and GlueTAC [29]. Drugs that target the autophagy–lysosome system include the autophagy-targeting chimera (AUTAC) [30], autophagosome tethering compound (ATTEC) [31], autophagy-targeting chimera (AUTOTAC) [32], and chaperone-mediated autophagy-based degrader (CMA-based degrader) [33,34].

2.2. Mechanism of Action of ARV110 in Treating Prostate Cancer

Currently, there are three PROTACs targeting the AR in clinical trials (Table 1). One of them is ARV-110, which was developed by Arvinas. Of all the PROTACs in clinical trials, ARV-110 was the first to enter clinical trial. ARV110 degrades AR by inducing protein degradation via hijacking the above mentioned pathway of UPS. ARV-110 shows its activity by bringing the AR and ubiquitin-recruiting enzyme (CRBN E3 ligase) in to close proximity. This is achieved when the thalidomide warhead and another component of ARV-110 bind to the E3 ligase covalently and AR non-covalently, respectively, forming a ternary complex (E3 ligase-ARV-110-AR). The presence of the E3 ligase and AR makes the process of ubiquitination more efficient. Once the AR is ubiquitinated, it is recognized and broken down by 26s proteasome (Figure 2). The free PROTAC and E3 ligase complex are then ready to form a ternary complex with another AR [3].

3. Advantages and Disadvantages of PROTACs over SMIs

PROTAC technology has recently garnered significant interest among researchers in the field of cancer therapeutics, especially in the area of prostate cancer drug discovery. This is due to the fact that PROTACs possess distinct benefits and variations compared to the traditional small molecule inhibitors. (Table 2) [18].
The development of effective anticancer drugs poses challenges in terms of targeting oncoproteins and achieving effective binding with them. However, PROTACs offer a unique solution as they can target proteins without binding to the active site, making previously undruggable targets now druggable [35]. In contrast to PROTACs, conventional small-molecule inhibitors (SMIs) require larger doses to be effective because they need to bind with many target proteins. The unique aspect of PROTACs is that they do not have to bind to the active site of the target protein, making previously undruggable targets now druggable. The “hook effect” occurs when too many PROTAC molecules bind to either the target protein or E3 ligase, reducing the efficiency of the ternary complex formation and leading to less target degradation. This is why low doses of PROTACs are preferred, as a single molecule can form the ternary complex by binding with both the target protein and the E3 ligase. The catalytic nature of PROTACs allows for longer action and a different PK/PD profile compared to SMIs. PROTACs only need a small amount of interaction with the target to induce noticeable degradation. Their event-driven mechanism of action has a wider biological relevance than SMIs as it can overcome resistance mechanisms such as drug target overexpression and resistance-causing mutations.
Another significant advantage is that proteins that can be degraded using PROTAC do not require an enzyme active site but a small-molecule binding site that an E3 ligase can access. A ligand must typically have a moderate affinity to use these sites, and a recruited E3 ligase must have access to the POI surface close to the binding site. High-affinity ligands are generally not required for this process, which is usually strenuous to design. An ideal E3 ligase ligand would be a powerful binder with a delayed off-rate, even though the ligand affinity for the POI might be modest. The capacity to turn over and function repeatedly is lost by PROTACs that covalently bind to the POI. In contrast, PROTACs that covalently bind to an E3 ligase would decrease the three-body assembly kinetics to a two-body issue and increase catalytic efficiency [17,18,36]. Crew’s lab presented examples of successful degradation of three RTKs, EGFR, HER2, and c-Met, including various mutants of EGFR and c-Met. This demonstrates that PROTACs are capable of causing the degradation of active receptor tyrosine kinases.
Like any other technique, PROTAC technology has both advantages and disadvantages. The main issue with PROTACs is that they do not conform to the usual “rule of 5”, unlike traditional, small-molecule drugs [37]. The characteristics and molecular weights of PROTAC can be altered by the introduction of short linkers, but ‘linker-ology’ is more than just a means of joining the POI and E3 warheads. Another significant challenge is the limited availability of ligands for E3 ligases, with only a small number of E3 ligases such as CRBN, VHL, MDM2, IAPs, DCAF15, DCAF16, RNF114, KEAP1, FEM1B, etc., out of the 600 known E3 ligases being utilized in clinical practice [38]. Although research is being conducted in this area, cell surface proteins are currently not considered optimal targets for PROTAC therapy as the UPS resides inside the cell [17]. The choice of the ideal linker to join these two binding components has proven to be a significant design challenge for PROTACs to date. Most of the time, scientists use synthetically accessible chemistry to experiment with linkers of varying lengths. However, the research has shown that sometimes, having a linker of an arbitrary length can negatively impact the target. In other cases, the cooperation between the target and the E3 ligase is dependent on a protein–protein interface that includes the linker. There is an increasing body of evidence that suggests that a stable ternary complex, formed by protein–protein interactions, is crucial for successful degradation [39].

4. Genetic Alterations Affecting Prostate Cancer

PC is one of the most heritable human cancers with a strong genetic component. An integrated investigation of advanced prostate cancer found that nearly 90% of individuals with metastatic castration-resistant prostate cancer (mCRPC) had clinically actionable germline and somatic alterations. Both somatic and germline genetic aberrations can affect the tumorigenesis as well as the disease progression of PC [40]. The genetic abnormalities of PC that cause sporadic cancer are predominantly somatic mutations. The most common changes in prostate cells are the activation of oncogenes and the loss of function of tumor-suppressor genes. Additionally, structural lesions such as genomic rearrangements, including amplifications, translocations, or deletions, are a leading cause of the development of PC [41]. Copy number variations, such as the gain or loss of segments of genomic DNA, can also lead to the amplification of oncogenes and deletion of tumor-suppressor genes. Another potential outcome of chromosomal rearrangements is the formation of gene fusions that promote oncogenesis. Single nucleotide polymorphisms (SNPs), point mutations such as missense and nonsense mutations, and frameshifts are less prevalent in prostate cancer [42].
The development of prostate cancer is mediated by abnormalities in AR signaling, and androgen deprivation therapy (ADT) is the cornerstone of systemic treatment for individuals with advanced PC. On the other hand, research findings in the biology of PC have revealed that up to 60% of individuals with advanced PC have clinically actionable molecular and genetic alterations in non-AR-related pathways (Figure 3) [43]. Hence, it is essential to focus on both AR- and non-AR-related targets for improvements in PC therapy (Figure 4).

4.1. Non-AR-Related Pathways

A spectrum of genetic alterations was observed in different molecular pathways of PC, including DNA repair, Phosphoinositide 3-Kinase (PI3K), Ras/Raf/MAPK, TGF-β/SMAD4, epigenetic deregulation, and many more pathways [44]. The incidence of germline mutations in DNA repair genes is significantly higher in metastatic PC as compared to localized PC. Genomic studies have identified germline or somatic aberrations in various DNA repair genes such as ATM, BRCA1, BRCA2, CHEK2, FANCA, RAD51D, and PALB2. BRCA2 was the most frequently altered of the major DNA repair genes, followed by ATM. Deletions (BRCA2, ATM, and PALB2), frameshift mutations (BRCA1 and CHEK2), SNPs (RAD51D), and loss of functional mutations (FANCA) are some of the genetic aberrations observed in PC [45]. Phosphatase and Tensin homologue (PTEN) is a tumor-suppressor gene that negatively regulates the PI3K pathway. PTEN appears to be more often mutated in PC metastases, indicating a function for PTEN in the disease progression. The common alterations in the PTEN gene include deletions and frameshift mutations. Furthermore, the Ras/Raf/MAPK pathway is a critical signaling mechanism that plays a critical role in many cancers. Ras and Raf promote MAPK signaling and may increase AR transcriptional activity. The point mutations in RAS oncogenes (HRAS, KRAS, and NRAS) transform the normal cellular genes into abnormally activated oncogenes, which can lead to tumorigenesis in PC.
The E26 transformation-specific (ETS) family of transcription factors, such as ETVs and ERG genes, are involved in the regulation of cell cycles, cell proliferation, and several cellular functions. The development of PC is associated with recurrent gene fusions in ERG, ETV1, ETV4, and ETV5, as well as deletions and translocations in ETS2 [46]. SMAD4 serves as a mutual downstream node of the transforming growth factor-beta (TGF-β) and bone morphogenetic protein (BMP) pathways and controls cell proliferation. Deletions in the SMAD4 gene affects the TGF-β/SMAD4 pathway that leads to the metastatic progression of PC [47]. Annexin A7 (ANXA7) gene encodes Ca2+-activated GTPase, which promotes cell proliferation. ANXA7 expression was shown to be often lost in prostate cancer, particularly in metastasis and local recurrences of hormone-refractory prostate cancer. In PC, the hemizygous deletion of ANXA7 leads to a decreased expression and function that results in cancer progression [48]. AT motif binding factor 1 (ATBF1) gene expression is linked to reduced cell proliferation, increased expression of the CDKN1A (p21) tumor suppressor, and decreased expression of the AFP oncoprotein. Hence, ATBF1 is an equitable candidate tumor-suppressor gene in PC. A deletion of 21 or 24 nucleotides in the coding region has been found not only in sporadic tumors, but also in the germline DNA of PC [49].
The CDKN1B gene provides the instructions for making the p27 protein, which plays a crucial role in controlling cell division. A variation of the CDKN1B gene, which is known as an SNP variant, has been linked to an increased risk of advanced prostate cancer (PC). Additionally, the absence of the CDKN1B gene leads to abnormal cell division [50]. CHEK2, a checkpoint kinase, regulates the p53-mediated DNA damage signaling pathway and has been found to have several alterations, including frameshift mutations, in both hereditary and sporadic prostate tumors [51]. The EPHB2 gene codes for a receptor tyrosine kinase protein and has been found to have both frameshift and nonsense mutations in 10% of sporadic PC cases.
Studies have investigated whether certain genetic alterations in Glutathione S-transferases (GSTs)—GSTP1, GSTM1, and GSTT1—that play a role in carcinogen metabolism and defense against reactive oxygen species are linked to prostate cancer risk. The results showed that specific mutations and SNPs in GSTP1, GSTM1, and GSTT1 have been linked to an increased risk of prostate cancer [52]. The Kruppel-Like Factor 5 (KLF5) is a transcription factor that is involved in cell proliferation, differentiation, and carcinogenesis. The hemizygous and homozygous deletion of KLF5 gene in human PC is rare [53]. Further, KLF6 is a zinc finger transcription factor that has been linked to cell proliferation and differentiation. A germline SNP in KLF6 is significantly associated with an increased relative risk of PC [54].
Macrophage scavenger receptors (MSRs) are trimeric membrane glycoproteins that facilitate the binding, internalization, and processing of a wide array of macromolecules. Genetic aberrations in MSR1, including truncating mutations, have been presented to be associated with the PC risk in both sporadic and hereditary cancers [55]. MYC is a cellular proto-oncogene that is well associated with cell transformation. The overexpression of MYC can immortalize prostatic epithelial cells. Hence, the gain of functional mutations identified in MYC is clearly an oncogenic factor in PC [56]. NK3 Homeobox 1 (NKX3-1) is a prostate-specific gene in humans. The NKX3-1 gene is expressed at higher levels in mature prostate tissues, but its expression is decreased in prostate cancer cells. Due to the gene being haploinsufficient, a hemizygous deletion leads to a reduction in both its expression and function [57]. TP53 is a tumor-suppressor gene whose modifications are implicated in the molecular genetics of several human cancers. Furthermore, the p53 gene is one of the most commonly mutated genes in human cancers. While mutations of p53 are relatively rare in early-stage prostate cancer, they become more frequent in higher-stage, higher-grade tumors and metastases. Deletions and functional mutations of the p53 gene have been found to be prevalent in advanced prostate cancer [58].
The human serum paraoxonase and arylesterase 1 (PON1) gene is involved in the elimination of carcinogenic lipid-soluble radicals. The expression of PON1 varies widely in humans, and specific polymorphisms in the gene have been linked to different levels of PON1 protein in the blood. Single nucleotide polymorphisms (SNPs) in the coding region of PON1 have been associated with a decreased serum paraoxonase activity, which may contribute to the development of cancer [59]. The RNASEL gene encodes a 2′-5′-oligoadenylate (2–5 A)-dependent RNase L protein that plays a vital role in the innate immune response. The variant (Arg462Gln) in RNASEL is allied with fewer enzymatic activities and was significantly associated with PC risk. The frameshift, truncating mutation, as well as the deletions in this gene are linked to an elevated risk of PC [60]. Steroid 5-alpha-reductase 2 (SRD5A2) converts testosterone to di-hydro testosterone, which is essential for the development of the prostate. The SNP variants in SRD5A2 are associated with more aggressive features of PC such as a higher tumor stage, metastasis, and extracapsular disease [61]. In addition, prostate cells express a vitamin D receptor (VDR), which mediates the functions of 1, 25-dihydroxy vitamin D, and the polymorphisms in VDR are connected with an increased risk of more aggressive PC [62].

4.2. AR Pathway

Androgens play a crucial role in the growth and maintenance of the prostate. The relationship between androgen levels and prostate cancer (PC) has been well established. The intracellular AR, a member of the ligand-dependent transcription factor family, is responsible for mediating androgen activity. When testosterone binds with AR, it triggers the transcription of androgen-responsive genes, thus controlling the growth of both the normal prostate gland and PC [63]. The androgen-signaling axis is a primary target for the treatment of metastatic prostate cancer through androgen ablation therapy. However, this treatment method can become ineffective over time as a result of resistance. Recent findings suggest that this resistance is caused not by a lack of androgen signaling but by genetic alterations that lead to abnormal activation of the androgen-signaling axis. It has been observed that many primary and metastatic prostate cancers exhibit genetic changes in the androgen signaling pathway, such as amplifications, mutations, and changes in AR corepressors (NCOR1/2) and coactivators (NCOA1/2), which result in castration resistance [64].
Further, structural rearrangements of AR were identified in one-third of metastatic PCs. In general, FOXA1 suppresses the androgen signaling pathway and promotes prostate tumor growth. Therefore, recurrent mutations in the AR associating factor FOXA1 lead to the development of PC and the alterations in FOXA1 have been identified in 3–4% of both untreated, localized PC and metastatic PC [47].
A significant number of prostate cancer patients have mutations in specific genes, with these alterations being more common in metastatic cases compared to localized ones. Traditional small-molecule inhibitors (SMIs) struggle to target these mutations, as multiple genetic abnormalities lead to resistance in prostate cancer. Each mutation in a particular gene can alter protein signals differently, so it is important to carefully select the correct gene to target. Some genetic abnormalities, such as SNP, deletions, and point mutations, can render the target protein untreatable. To address these mutations, PROTACs (Proteolysis-Targeting Chimera) offer a promising alternative therapeutic option for prostate cancer. Currently, PROTACs are designed to target the AR in prostate cancer. While other genes in addition to AR are altered in prostate cancer, AR remains the primary focus of established PROTACs. These alterations and targets could be valuable for the development of new PROTACs for precision cancer therapy and medicine for prostate cancer.

5. Prominent PROTACs Targeting Prostate Cancer

Over the past several decades, the pathophysiology of prostate cancer has been extensively studied and understood. Inhibiting the AR is widely considered as one of the most effective therapies for prostate cancer, despite its tendency to develop resistance over time. To address this issue, alternative approaches have been proposed, including the use of PROTACs, which target and degrade the AR. The removal of the protein reduces the risk of developing resistance and has gained significant attention in the treatment of prostate cancer, particularly after the development of ARV-110. A significant number of PROTACs (nearly 170) have been reported to target various proteins that are associated with prostate cancer (Table S1). In this article, we focus on the development of various AR and non-AR target degrading PROTACs and their potential to treat prostate cancer, with an emphasis on drug optimization and medicinal chemistry.

5.1. Androgen Receptor Targeting PROTACs

5.1.1. CRBN-Based PROTACs

There are now a significant number of unique AR degraders that have been revealed and are based on the target protein degradation (Proteolysis-Targeting Chimera) principle. Lukas et al. synthesized three series of degraders based on FDA approved AR inhibitor, enzalutamide. They modified the exit vector portion of the enzalutamide with benzamide, phenoxy, and piperidine scaffolds along with linker composition. The subseries with an exit vector of biphenyl/phenoxy type, demonstrated effective AR degradation at a concentration of 1 µM. A polyethylene glycol (PEG) linker containing PROTAC was discovered to be active compared to simple alkyl chain. The final subseries had PROTACs with a phenoxy exit vector with more rigid linkers, which resulted in the best molecules. Furthermore, stereochemistry of the molecules influenced the target degradation, as the geometrical orientation of the ternary complexes was likely altered. PROTAC 35 (compound 1) from final subseries was chosen for additional in-depth investigation since it significantly degraded even at 0.1 mM, resulting in DC50 values of 75 nM and 77 nM in A549 and LNCaP cell lines, respectively. Additionally, they carried out cell specificity tests, which demonstrated that compound 1 was more selective for the PC cell line LNCaP [65].
A series of CRBN-mediated AR PROTACs, which combine a tetramethylcyclobutane-based AR antagonist with the TD-106 CRBN binder, were developed and evaluated by Akshay D. Takwale et al. for the treatment of metastatic castration-resistant prostate cancer. Among the synthesized, TD-802 (compound 2) displayed the most effective AR protein degradation, with a DC50 of 12.5 nM and a maximum degradation of 93% in the LNCaP cell lines. The effectiveness of AR degradation is highly dependent on the linker location within the TD-106 (CRBN binder). In addition, pharmacokinetic studies in mice demonstrated that over 70% of the degraders, including TD-802, exhibited good microsomal stability after 30 min of incubation, a long half-life of over 4 h, and high plasma exposure following intravenous and intraperitoneal injections. Finally, TD-802 induced AR degradation and demonstrated strong antitumor activity in an in vivo xenograft mouse model, indicating that CRBN-based AR degraders may have potential for use in the development of new AR therapies [66].
Jian-Jia Liang et al. developed and tested a series of new PROTACs based on RU59063 derivatives and cereblon ligands for their ability to degrade the AR protein in LNCaP (AR+) cells. Almost all the PROTACs displayed a good binding affinity and efficacy against AR. Among the compounds, A16 (compound 3) was found to be the most promising, with an 85% AR antagonist efficacy compared to enzalutamide (89%). Further, the Western blot analysis revealed that A16 caused only a minor decrease in the amount of AR protein in LNCaP cells. Molecular docking experiments with AR and DDB1-CRBN E3 ubiquitin ligase complex (PDB: 2AXA) indicated that A16 has binding properties similar to other non-steroidal AR ligands, with key interactions [67].
Ga Yeong Kim et al. designed, synthesized, and evaluated a series of PROTACs consisting of PEG chain-linked bicalutamide analogues and thalidomides (E3 ligase recruiters) as novel AR degraders. Among the synthesized, 13b (compound 4) was found to be highly effective in decreasing the mRNA levels of transcriptional genes that are dependent on AR, such as prostate-specific antigens and TMPRSS2. It was also more effective than 13c (compound 5), which had a slightly greater AR inhibitory activity compared to bicalutamide. The author’s findings suggested that replacing the trifluoromethyl and sulfide groups with iodine and ether, respectively, may improve the AR antagonist activity of bicalutamide. Further, the AR antagonist activity was not affected when the hydroxyl group was replaced with fluorine in the terminal phenyl moiety at the solvent exposure site, indicating that the ether linkage may be used as the linker moiety for conjugating the AR ligand with thalidomide. The compounds 13b and 13c (compounds 4 and 5) were significantly degraded by the AR in LNCaP cells in a dose- and time-dependent manner. Interestingly, the AR variant 7 (AR-V) form was more susceptible to compound 13c than the wild type. These results suggest that compound 13c targets and degrades AR through direct binding and does not modulate the mRNA expression [68].
In another study, researchers led by Shaomeng Wang reported a highly potent and effective AR degrader ARD-61 (compound 14) for the treatment of PC. However, it was found to be inactive when administered orally in mice. In order to improve the oral bioavailability of ARD-61, the authors opted to utilize the cereblon as the E3 ligase ligand instead of the VHL ligand, owing to its lower molecular weight and superior drug-like properties. The first series of molecules was synthesized using the AR antagonist of ARD-61 and thalidomide with both flexible and semi-rigid linkers, and it was observed that constrained linkers had almost the same potency as linear and flexible linkers. The next series of molecules was synthesized by changing the linking position from the meta to the ortho position in the phenyl ring of the cereblon ligand, which decreased the activity. The most promising molecule from this series was then tested using modified AR antagonists with the same linker and CRBN ligand, showing that the AR antagonist of ARD-61 is essential for its activity. The DC50 values of the promising compounds were tested on the VCaP cell line, resulting in <0.5 nM activity and <5 nM for DC90. The potencies of these compounds were also evaluated, and the mechanism of action was tested. Further pharmacokinetic properties showed that the selected compounds have an excellent oral bioavailability. The most promising molecule, ARD-2128 (compound 6), was tested further and found to be effective in reducing the levels of AR, PSA, TMPRSS2, and FKBP5. Additionally, it was discovered that ARD-2128 is highly effective at reducing tumors, even at lower doses compared to enzalutamide [69].
Weiguo Xiang et al. reported the development of a potent and orally active compound for AR degradation. The initial series was synthesized using AR antagonists from Pfizer and Arvinas, along with the CRBN ligand and linkers of various lengths. The first series of compounds was synthesized by hooking Pfizer’s molecule to the isoindoline-1,3-dione moiety of thalidomide at the five position through the para-position of its phenyl group, with amine groups on both ends and linear alkyl groups of various lengths. Although all the compounds showed promising results similar to ARV-110, with a DC50 of 1.6 nM and Dmax of 98% in the VCaP cell line, their oral bioavailability was not ideal. Additional compounds were synthesized using semi-rigid and rigid linkers. Rigid linkers resulted in significant improvements in the DC50, Dmax, and kinetic properties. The lead compound was further optimized by modifying the AR antagonist portion of the molecule, specifically by altering the cyclohexane ring and the linker atom connecting the phenyl and cyclohexane rings. The results demonstrated that the cyclohexane ring and tertiary nitrogen (as the linker atom) are highly effective and essential for the activity. The modified compounds using different rigid linkers were identified as highly effective toward AR, with DC50 values ranging from 0.01 to 1.4 nM and Dmax values of 93 to 100%. After subjecting the selected AR degraders in LNCaP cell lines to further analysis, including AR degradation kinetics, pharmacodynamics, and tissue distribution studies, it was determined that the most promising molecule among them was ARD-2585 (compound 7) (Figure 5) [70].

5.1.2. VHL-Based PROTACs

Xin Han et al. analyzed the correlation between the binding affinity of the VHL ligand component of the VHL protein and the efficacy of the resultant PROTAC degraders. The study revealed that utilizing VHL ligands with a low binding affinity (Ki = 2–3 μM) to the VHL protein can result in the development of highly effective and potent degraders. According to prior research, it was observed that co-crystal structures of VHL ligands in conjunction with VHL protein demonstrated that the 4-methylthiazole group of the ligand binds to the hydrophobic region of the VHL protein. Hence, the region was modified with different groups such as nitrile, bromine, chlorine, fluorine, and hydrogen, and the resulting compounds were tested for their binding affinity to VHL protein. Using ARD-61 (compound 14) as a template, a series of new PROTAC AR degraders was synthesized using VHL ligands with a wide range of binding affinities to VHL. It was found that almost all of the synthesized molecules were effective at reducing AR at all tested concentrations, with three compounds showing excellent activity (greater than 80% degradation at 10 nM), despite having a weak binding affinity to VHL. To further optimize the compound with a very weak VHL binding affinity ARD-266 (compound 8), the linker was modified by reducing the number of rings and atoms within the rings, resulting in a more potent compound (>90% degradation at 10 nM). The biological evaluation of ARD-266 demonstrated that it has DC50 values of 0.5, 1, and 0.2 nM in the LNCaP, VCaP, and 22Rv1 cell lines, respectively. After a 6-h treatment, ARD-266 was able to reduce the AR protein level by more than 95% at concentrations of 30 nM in both the LNCaP and VCaP cell lines and 10 nM in the 22Rv1 cell line. In comparison to conventional AR inhibitors, ARD-266 was found to be more potent, with an IC50 value of 6 nM compared to 182 nM for the AR antagonist [71].
Linrong Chen et al. reported a series of AR degraders, including A031, based on two distinct AR antagonists (azabicyclic and fused bicyclic compounds) and four E3 ligands. They developed PROTACs with a variety of connecting options, such as various warheads, and modified the linker using different heterocyclic and phenyl ring combinations along with an alkyl chain, resulting in a moderate increase in activity. The modifications made to the attachment points on the CRBN ligand did not yield any enhancement in activity. Nevertheless, modifications made to the AR antagonist component led to a reduction in activity. It is noteworthy that substituting the E3 ligase ligand with VHL-2 and shortening the linker length led to the development of improved molecules such as A031 (compound 9). The compound A031 showed promising activity even at lower concentrations, as low as 0.125 µM. Further, the additional biological studies revealed that A031 has comparable efficacy and five times less toxicity as compared to EZLA (the positive control) [72].
Munoz et al. reported a series of niclosamide-based PROTACs for the treatment of AR+ve PC that use VHL-032 as an E3 ligase-binding warhead. The compounds demonstrated moderate IC50 values against the PC-3, LNCaP, and 22Rv1 cell lines. However, it was found that their mechanism of action did not involve AR degradation, as confirmed by Western blotting assays, even at concentrations as high as 1 µM. This is in contrast to the clinical candidate ARV-110, which demonstrated AR degradation. However, Niclo-Click PROTAC 5 (compound 10) exhibited a decent IC50 of 1.04 ± 0.18 µM in the LNCaP cell line [73].
Lee et al. designed and developed a unique PROTAC using 4-(4-phenylthiazol-2-yl)morpholine called MTX-23 (compound 11) which selectively targets and degrades both the full-length AR (AR-FL) and the AR splice variant 7 (AR-V7). In immunoblotting studies, MTX-23 had a DC50 of 0.37 µM for AR-V7 and 2 µM for AR-FL. The compound was found to inhibit proliferation and enhance apoptosis selectively in androgen-responsive PC cells. MTX-23 was tested in cells resistant to four FDA-approved SAT agents (abiraterone, enzalutamide, apalutamide, and darolutamide) and exhibited promising results. It decreased cellular proliferation and reduced tumor growth in both in vitro and in vivo, potentially making it effective against castration-resistant prostate cancer by degrading both AR-V7 and AR-FL [74].
Jemilat Salami et al. synthesized a series of PROTACs based on enzalutamide to target the AR using VHL as an E3 ligase-binding warhead. Among the compounds synthesized, ARCC-4 (compound 12) was the most potent, demonstrating a Dmax of 98% and a DC50 of 5 nM. In cellular models of castration-resistant prostate cancer (CRPC), ARCC-4 was more effective at inducing apoptosis and inhibiting proliferation of AR-amplified prostate cancer cells compared to its parent compound, enzalutamide. Additionally, ARCC-4 retained its ability to degrade AR and inhibit cell proliferation in a high-androgen environment. The mechanism of action for ARCC-4 was determined to be proteasomal degradation through polyubiquitination of AR, as confirmed using the Tandem Ubiquitin-Binding Element (TUBE1) pull-down assay. ARCC-4 was also able to effectively degrade several clinically relevant mutants of AR, including AR-F876L, AR-T877A, AR-L702H, AR-H874Y, and AR-M896V. ARCC-4 was effective at degrading the AR protein in VCaP cells even with increasing concentrations of the synthetic androgen R1881 up to 1 nM [75].
Xin Han et al. conducted a comprehensive optimization process to develop a highly potent and effective AR-degrading PROTAC. They utilized enzalutamide as the AR-binding ligand and VHL as the E3 ligand and synthesized a set of PROTACs through the use of an amide linker with varying chain lengths. The biological studies of synthesized compounds indicated that an optimal linker length of 11–12 atoms resulted in the most effective AR degradation. The introduction of a pyridine group directly connected to the ethynyl group significantly increased the activity and solubility of the compound. The optimization data demonstrated that increasing the rigidity of the linker enhances the activity of the PROTAC. The researchers then modified the stereochemistry of VHL to examine its role in binding to the VHL protein, eventually selecting ARD-69 (compound 13) for further investigation. ARD-69 achieved DC50 values of 0.86 and 0.76 nM in LNCaP and VCaP cell lines, respectively, and induced greater than 95% AR degradation at 10 nM in both cell lines [76].
Steven Kregel et al. reported the development of the AR degrader ARD-61 (compound 14), which demonstrated effectiveness in addressing resistance mechanisms that often arise during PC treatment. From the results, it was identified that ARD-61 caused PARP cleavage in all AR-dependent cell lines, regardless of AR levels, AR splice variant expression, or AR mutational status. ARD-61 was also effective in model systems resistant to enzalutamide and AR splice variants, including AR-V7. Although ARD-61 is unable to bind and degrade AR-V7, it was able to inhibit tumor cell growth in models overexpressing AR-V7. In addition, ARD-61 was degraded to full-length AR in all four CRISPR-induced clones overexpressing AR-V7 but had no effect on the production of AR-V7. When tested in the aggressive, metastatic, castration- and enzalutamide-resistant CWR-R1 EnzR xenograft model, which expresses high levels of AR-V7, ARD-61 was found to be very effective in vivo. These findings suggest that, despite the potential for AR-V7 to enhance AR signaling, full-length AR is necessary for the development and survival of castration- and enzalutamide-resistant prostate cancer (Figure 6) [77].

5.1.3. Miscellaneous

Bohan Ma et al. developed a peptide-based PROTAC that utilizes ultra-small gold nanoparticles as a delivery method. The synthesized degraders target the DNA-binding domain (DBD) of AR and induce the degradation of both AR and AR-V7. The desired PROTAC sequence was synthesized using a solid phase peptide synthesis, with a flexible linker introduced between the AR-DBD- and MDM2-targeting sequences. Using isothermal titration calorimetry, the binding affinity between the AR pep-PROTAC and AR-V7 was determined to be 12.2 nM and 49.6 nM, respectively. The Au-AR pep-PROTAC was found to stimulate the degradation of AR and AR-V7 and reduce the growth of AR- and AR-V7-positive prostate cancer cell lines, including LNCaP, C4-2, and CWR22Rv1, with IC50 values of 230.8 nM, 248.1 nM, and 126.9 nM, respectively. The DC50 values for half-maximal degradation were found to be 175.2 nM, 193.0 nM, and 48.8 nM (AR)/79.2 nM (AR-V7) in LNCaP, C4-2, and CWR22rv1 cells, respectively. Furthermore, it was determined that the Au-AR pep-PROTAC compound elicits degradation of AR in both the cytoplasmic and nuclear compartments [78].
Hang Xie et al. reported the synthesis of two series of selective AR degraders (SARDs) based on the AR agonist RU59063 (4-(4,4-dimethyl-3-(4-hydroxybutyl)-5-oxo-2-thioxo-1-imidazolidinyl)-2-(trifluoromethyl)benzonitrile) with hydrophobic degrons. The analysis of the structure–activity relationship revealed that compounds with alkyl chain linkers had stronger antiproliferative activity compared to those with phenyl groups. The presence of the 1,2,3-triazole group also significantly impacted the compounds’ activities. A9 (compound 15) was found to have a strong inhibitory effect against LNCaP cells and a shared AR-binding affinity with enzalutamide. The molecular docking study revealed that the interaction between the extensive hydrophobic tagging chains of the synthesized compounds and the AR could affect the rate of degradation (Figure 7) [79].

5.2. Non-AR-Targeting PROTACs

There are a variety of proteins that have been identified as potential targets for the treatment of prostate cancer. Apart from the AR, PROTACs have been developed to target some of these proteins. For example, c-Myc and protein kinases, which are frequently overexpressed in PC and have been linked to cancer progression. These non-AR-targeting PROTACs have the potential to be used as therapeutic agents for the treatment of PC.
Kanak Raina et al. reported the use of PROTAC-induced BET protein degradation for the treatment of castration-resistant prostate cancer (CRPC). They demonstrated that the small-molecule pan-BET degrader ARV-771 (compound 16) was more effective than BET inhibition in cellular models of CRPC. In order to optimize the lead compound for in vivo studies, the linker’s length and composition were modified to attain the desired pharmacological properties. In vitro studies on the CRPC cell lines 22Rv1, VCaP, and LnCaP95 showed that ARV-771 potently degrades BRD2/3/4 with a DC50 smaller than 5 nM. The loss of the BET function was determined by measuring the concentration of the c-MYC protein, which decreased to an IC50 smaller than 1 nM. The induction of apoptosis by ARV-771 was demonstrated by the substantial breakage of poly (ADP-ribose) polymerase (PARP) in 22Rv1 cells. ELISA and immunoblotting studies exhibited that ARV-771 dramatically reduced the AR protein levels in VCaP cells, but not JQ1 or OTX015. When the VCaP cells were given a 10 nM dose of ARV-771, the levels of both the FL-AR and AR-V7 mRNA decreased. The in vivo studies of ARV-771, administered at a dose of 10 mg/kg, in male Nu/Nu mice with AR-V7+ 22Rv1 tumor xenografts, demonstrated a significant reduction in the AR-V7 levels in the tumors [80].
Rong Hu et al. designed and synthesized the selective BRD4-PROTAC (WWL0245) using a dual BET/PLK1 inhibitor, WNY0824. Computational analyses revealed that a linker and an E3 ligand, such as lenalidomide and thalidomide, could be introduced to the 1-methylpiperidin-4-yl motif of WNY0824, which is exposed to the solvent region. WWL0245 (compound 17) showed selective cytotoxicity in BETi-sensitive cancer cell lines, including AR-positive PC cell lines. The authors evaluated the effect of WWL0245 and WWL0202 (compound 18) on the expression of BRD2/3/4, PLK1, and c-Myc in VCaP cells. Both compounds significantly reduced the BRD4 levels and inhibited c-Myc at concentrations as low as 10 nM but had no significant impact on protein levels. Due to its exceptional antiproliferative activity, selectivity, and low molecular weight, WWL0202 was selected for further studies. WWL0202 was found to have a DC50 in the sub-nanomolar range, and when tested on 22Rv1 and VCaP cells, it resulted in a reduction in the BRD4 and c-Myc levels in a concentration-dependent manner [81].
Fei Zhou et al. designed multiple cyclin-dependent kinase (CDK)-targeting PROTAC degraders. Using computational studies, they identified the hydrophobic region of two pan-CDK inhibitors (AT-7519 and FN-1501) as the linker attachment points. They designed two series of protein degraders by connecting the two ligands and the CRBN ligand with the same set of linkers. The degraders made with AT-7519 selectively degraded CDK2 at a concentration of 1 µM. After performing a CCK8 assay against PC-3 cell lines, A9 and F3 (compounds 19 and 20) were found to be potent with IC50 values of 0.84 and 0.12 µM, respectively. Mechanistic studies revealed that it could effectively and selectively reduce the levels of CDK2 and CDK9 in a concentration-dependent manner, with DC50 values of 62 and 33 nM, respectively. Further synthesis of molecules based on F3 did not result in any improved activity. The compound F3 was found to significantly extend the S phase by 9.8% at a concentration of 250 nM, and higher concentrations caused cell cycle arrest at the G2/M phase. These findings provide a strong platform for targeting non-AR targets for PC therapy using the TPD (targeted protein degradation) strategy (Figure 7) [82].

6. Design and Medicinal Chemistry of PROTACs

Prior to 2008, PROTACs were primarily peptide-based; however, the introduction of the first small-molecule PROTAC by the Crews group, utilizing MDM2, marked a pivotal shift in their synthesis. The crucial part of developing PROTACs is the rational design of all the three components: (i) the E3 ligase warhead (E3 ligand), (ii) the target-binding warhead, and (iii) the linker [83].

6.1. Design of E3 Ligands

To date, researchers have identified more than 600 E3 ubiquitin ligases in the human genome; however, the potential of these ligases remains constrained by the dearth of small molecules exhibiting a strong affinity and specificity for them [84]. When designing a E3 ligase ligand for specific protein degradation, there are a few important factors to consider, such as the expression levels of different types of E3 ligases in the tissue where the target protein is concentrated, and the E3 ligase selectivity for the target protein. However, for any given target protein, VHL and CRBN ligands are a good starting point due to their wide scope of efficacy [85,86]. Based on the existing VHL and CRBN ligands, many other analogues that mimic their activity were reported using modern techniques, such as fragment-based drug design, structure-based drug design, high-throughput screening, DNA-encoded libraries, phage display technology, etc., and have been employed in the design of novel PROTACs [87].

6.2. Design of Linker

The linker plays a critical role in the formation of ternary complexes in PROTAC design, with various elements of the linker involved in specific interactions. As a result, optimizing the linker is a key goal in the development of PROTACs, considering factors such as length, composition, stiffness, and attachment points [88]. The important points to be considered when designing linkers are linker length composition and the linker’s ability to target protein the attachment site (the solvent-exposed site of the ligand). The linker length, if not appropriate, can affect the proximity between E3 ligase and target proteins, which subsequently affects activity. Tertiary structure formation between E3 ligase-PROTAC-target protein, pharmacokinetic properties of PROTAC depends on linker composition [86]. For example, the rigidity of the linker significantly increases the kinetic and dynamic properties of PROTACs [70].

6.3. Design of a Target Protein-Binding Ligand

Most of the ligands used to design this component of PROTACs are knowledge-based. They usually use ligands that are already approved as inhibitors of POI or reported to have high binding affinity with the target protein. It is not necessary to consider the active binding site of the protein when designing a ligand, unlike conventional inhibitors. It can bind to any site of the protein unless it is not inducing degradation [86].

7. Conclusions and Future Recommendations

AR plays a key role in the development and progression of prostate cancer, and genetic alterations in the AR gene have been identified in a significant proportion of PC cases. These genetic changes can result in different variants of the AR protein, which can affect its function and response to treatment. This diversity of AR variants can make it difficult to develop drugs that effectively target the receptor, as a single therapy may not be able to inhibit all of the different variants. This is a major challenge in the development of small-molecule therapeutics for PC to target all the AR aberrations. However, this theory has been repeatedly refuted by targeted protein degradation. It has been 22 years since the advent of the targeted protein degradation strategy utilizing PROTACs as a means of inducing protein degradation [89]. The clinical translation of PROTACs has been demonstrated through the initiation of several clinical trials, indicating their therapeutic potential. Furthermore, a plethora of PROTACs targeting both AR and non-AR proteins have been described in the literature, with a specific focus on PC. The ability of certain PROTACs to effectively degrade these AR variants highlights their potential as a promising strategy for overcoming treatment resistance in prostate cancer and other diseases.
Although most PROTACs developed for PC have focused on the AR, it should be noted that non-AR proteins also contribute significantly to the pathogenesis of PC, comprising 60% of the disease. Despite the significance of non-AR proteins, no PROTACs have been developed to specifically target these proteins. These non-AR proteins often exhibit a variety of genetic alterations, making them difficult to target with small-molecule inhibitors. In the current review article, we summarized the genetic alterations in non-AR proteins and discussed how targeting the same through PROTACs could be an effective strategy in the treatment of PC, as they may overcome resistance mechanisms that are associated with traditional targeted therapies. Further, the use of PROTACs as a precision medicine approach for the treatment of PC is a promising area of research. PROTACs have the potential to selectively degrade specific proteins that drive the disease, thereby offering a more targeted and personalized approach to the treatment of prostate cancer for each individual patient. This method of therapy has the potential to improve the efficacy and specificity of treatment, leading to better patient outcomes. In addition, the use of PROTACs can also lead to the identification of new therapeutic targets that may have been previously undiscovered. The potential of PROTACs as a precision medicine approach for prostate cancer treatment warrants further investigation in preclinical and clinical studies.
The field of PROTAC design and synthesis is challenged by the need for further exploration in the development of ternary complex co-crystal structures using silico methods, the identification of additional E3 ligase ligands, the optimization of synthetic procedures, and the enhancement of PROTAC pharmacokinetic properties. Additionally, there is a need to optimize the binding affinity and specificity of ligands and to gain a deeper understanding of the PROTAC mechanism of action.
Furthermore, in silico approaches such as Rosetta, PRosettaC, PROTAC-Model, and others can also aid in rationalizing PROTAC-mediated ternary complex formation and enhance PROTAC development efficiency. The adoption of these technologies in the design, development, and optimization of PROTACs may prove to be a promising and time-efficient method in the field of drug discovery. In addition, novel strategies, such as molecular glue, LYTAC, AUTAC, ATTEC, and RIBOTAC, which also promote degradation depending on proximity, are broadening the opportunities for drug development and may have therapeutic advantages [19,26,30,31,90].
The role of artificial intelligence in the design and development of target protein degraders (PROTACs) is a rapidly growing area of research that has enormous potential. By incorporating AI techniques, such as machine learning, computer-aided drug design, and molecular dynamics simulations, the discovery of highly specific and potent PROTACs can be accelerated. A recent study published in the Journal of Chemical Information and Modeling utilized deep learning algorithms to identify new E3 ubiquitin ligase-based PROTACs, showing improved binding affinity and specificity compared to traditional designs [91].
Another study published in the Journal of Medicinal Chemistry applied a machine learning approach to optimize the linker component of PROTACs, leading to improved stability and specificity [92]. These findings demonstrate the potential for AI to significantly enhance the efficiency and success rate of PROTAC design and development. It is imperative that further research be conducted in this field to fully leverage the benefits of AI and contribute to the advancement of therapeutic solutions for various diseases. As the use of AI continues to grow in the field of medicinal chemistry, it is expected to play an increasingly important role in the design and development of PROTACs. The integration of AI into PROTAC research has the potential to revolutionize the way we approach drug discovery and development, leading to faster and more effective solutions for various diseases.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28093698/s1, Table S1. PROTACs that target different biological proteins involved in the progression of prostate cancer [93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173].

Author Contributions

Conceptualization, P.Y., A.O.B. and R.S.; writing—original draft preparation, P.Y., V.V.A. and R.S.; writing—review and editing, P.Y., A.O.B. and R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

P.C. would like to acknowledge the Asian Healthcare foundation for providing research facilities. R.S. acknowledges the Startup India Seed Fund Scheme, DLabs, Indian School of Business, for the financial assistance.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Baudino, T. Targeted Cancer Therapy: The Next Generation of Cancer Treatment. Curr. Drug Discov. Technol. 2015, 12, 3–20. [Google Scholar] [CrossRef] [PubMed]
  2. Dong, J.; Zhang, Q.; Wang, Z.; Huang, G.; Li, S. Recent Advances in the Development of Indazole-Based Anticancer Agents. ChemMedChem 2018, 13, 1490–1507. [Google Scholar] [CrossRef]
  3. Békés, M.; Langley, D.R.; Crews, C.M. PROTAC Targeted Protein Degraders: The Past Is Prologue. Nat. Rev. Drug Discov. 2022, 21, 181–200. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Z.; Zhang, Y.; Xiang, Y.; Kang, X. Small-Molecule PROTACs for Cancer Immunotherapy. Molecules 2022, 27, 5439. [Google Scholar] [CrossRef] [PubMed]
  5. Prozzillo, Y.; Fattorini, G.; Santopietro, M.V.; Suglia, L.; Ruggiero, A.; Ferreri, D.; Messina, G. Targeted Protein Degradation Tools: Overview and Future Perspectives. Biology 2020, 9, 421. [Google Scholar] [CrossRef]
  6. Denmeade, S.R.; Isaacs, J.T. A History of Prostate Cancer Treatment. Nat. Rev. Cancer 2002, 2, 389–396. [Google Scholar] [CrossRef]
  7. Wang, Z.; Wu, S.; Zhu, C.; Shen, J. The Role of Ferroptosis in Esophageal Cancer. Cancer Cell Int. 2022, 22, 2–6. [Google Scholar] [CrossRef]
  8. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  9. NIH. Available online: https://seer.cancer.gov/statfacts/html/prost.html (accessed on 7 September 2022).
  10. Gandhi, J.; Afridi, A.; Vatsia, S.; Joshi, G.; Joshi, G.; Kaplan, S.A.; Smith, N.L.; Khan, S.A. The Molecular Biology of Prostate Cancer: Current Understanding and Clinical Implications. Prostate Cancer Prostatic. Dis. 2018, 21, 22–36. [Google Scholar] [CrossRef]
  11. Yap, T.A.; Smith, A.D.; Ferraldeschi, R.; Al-Lazikani, B.; Workman, P.; De Bono, J.S. Drug Discovery in Advanced Prostate Cancer: Translating Biology into Therapy. Nat. Rev. Drug Discov. 2016, 15, 699–718. [Google Scholar] [CrossRef]
  12. Mullard, A. Targeted Protein Degraders Crowd into the Clinic. Nat. Rev. Drug Discov. 2021, 20, 247–250. [Google Scholar] [CrossRef]
  13. Nandi, D.; Tahiliani, P.; Kumar, A.; Chandu, D. The Ubiquitin-Proteasome System. J. Biosci. 2006, 31, 137–155. [Google Scholar] [CrossRef]
  14. Livneh, I.; Cohen-Kaplan, V.; Cohen-Rosenzweig, C.; Avni, N.; Ciechanover, A. The Life Cycle of the 26S Proteasome: From Birth, through Regulation and Function, and onto Its Death. Cell Res. 2016, 26, 869–885. [Google Scholar] [CrossRef] [PubMed]
  15. 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]
  16. Kumar, D.; Hassan, M.I. Targeted Protein Degraders March towards the Clinic for Neurodegenerative Diseases. Ageing Res. Rev. 2022, 78, 101616. [Google Scholar] [CrossRef] [PubMed]
  17. Bondeson, D.P.; Mares, A.; Smith, I.E.; Ko, E.; Campos, S.; Miah, A.H.; Mulholland, K.E.; Routly, N.; Buckley, D.L.; Gustafson, J.L.; et al. Catalytic in vivo protein knockdown by small-molecule PROTACs. Nat. Chem Biol. 2015, 11, 611–617. [Google Scholar] [CrossRef]
  18. Li, K.; Crews, C.M. PROTACs: Past, Present and Future. Chem. Soc. Rev. 2022, 51, 5214–5236. [Google Scholar] [CrossRef]
  19. Dong, G.; Ding, Y.; He, S.; Sheng, C. Molecular Glues for Targeted Protein Degradation: From Serendipity to Rational Discovery. J. Med. Chem. 2021, 64, 10606–10620. [Google Scholar] [CrossRef] [PubMed]
  20. Yang, Z.; Sun, Y.; Ni, Z.; Yang, C.; Tong, Y.; Liu, Y.; Li, H.; Rao, Y. Merging PROTAC and molecular glue for degrading BTK and GSPT1 proteins concurrently. Cell Res. 2021, 31, 1315–1318. [Google Scholar] [CrossRef]
  21. Bradbury, R.H.; Acton, D.G.; Broadbent, N.L.; Brooks, A.N.; Carr, G.R.; Hatter, G.; Hayter, B.R.; Hill, K.J.; Howe, N.J.; Jones, R.D.; et al. Discovery of AZD3514, a small-molecule androgen receptor downregulator for treatment of advanced prostate cancer. Bioorg. Med. Chem. Lett. 2013, 23, 1945–1948. [Google Scholar] [CrossRef]
  22. Gustafson, J.L.; Neklesa, T.K.; Cox, C.S.; Roth, A.G.; Buckley, D.L.; Tae, H.S.; Sundberg, T.B.; Stagg, D.B.; Hines, J.; McDonnell, D.P.; et al. Small-Molecule-Mediated Degradation of the Androgen Receptor through Hydrophobic Tagging. Angew. Chem. 2015, 54, 9659–9662. [Google Scholar] [CrossRef]
  23. Liu, J.; Chen, H.; Kaniskan, H.U.; Xie, L.; Chen, X.; Jin, J.; Wei, W. TF-PROTACs Enable Targeted Degradation of Transcription Factors. J. Am. Chem. Soc. 2021, 143, 8902–8910. [Google Scholar] [CrossRef]
  24. Zheng, M.; Huo, J.; Gu, X.; Wang, Y.; Wu, C.; Zhang, Q.; Wang, W.; Liu, Y.; Liu, Y.; Zhou, X.; et al. Rational Design and Synthesis of Novel Dual PROTACs for Simultaneous Degradation of EGFR and PARP. J. Med. Chem. 2021, 64, 7839–7852. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, X.; Wang, Y.; Li, X.; Wu, J.; Zhao, L.; Li, W.; Liu, J. Dynamics-Based Discovery of Novel, Potent Benzoic Acid Derivatives as Orally Bioavailable Selective Estrogen Receptor Degraders for ERalpha+ Breast Cancer. J. Med. Chem. 2021, 64, 7575–7595. [Google Scholar] [CrossRef] [PubMed]
  26. Banik, S.M.; Pedram, K.; Wisnovsky, S.; Ahn, G.; Riley, N.M.; Bertozzi, C.R. Lysosome-targeting chimaeras for degradation of extracellular proteins. Nature 2020, 584, 291–297. [Google Scholar] [CrossRef]
  27. Miao, Y.; Gao, Q.; Mao, M.; Zhang, C.; Yang, L.; Yang, Y.; Han, D. Bispecific aptamer chimeras enable targeted protein degradation on cell membranes. Angew. Chem. 2021, 133, 11367–11371. [Google Scholar] [CrossRef]
  28. Cotton, A.D.; Nguyen, D.P.; Gramespacher, J.A.; Seiple, I.B.; Wells, J.A. Development of Antibody-Based PROTACs for the Degradation of the Cell-Surface Immune Checkpoint Protein PD-L1. J. Am. Chem. Soc. 2021, 143, 593–598. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, H.; Han, Y.; Yang, Y.; Lin, F.; Li, K.; Kong, L.; Liu, H.; Dang, Y.; Lin, J.; Chen, P.R. Covalently Engineered Nanobody Chimeras for Targeted Membrane Protein Degradation. J. Am. Chem. Soc. 2021, 143, 16377–16382. [Google Scholar] [CrossRef] [PubMed]
  30. Takahashi, D.; Moriyama, J.; Nakamura, T.; Miki, E.; Takahashi, E.; Sato, A.; Akaike, T.; Itto-Nakama, K.; Arimoto, H. AUTACs: Cargo-Specific Degraders Using Selective Autophagy. Mol. Cell 2019, 76, 797–810.e10. [Google Scholar] [CrossRef]
  31. Li, Z.; Zhu, C.; Ding, Y.; Fei, Y.; Lu, B. ATTEC: A potential new approach to target proteinopathies. Autophagy 2020, 16, 185–187. [Google Scholar] [CrossRef]
  32. Ji, C.H.; Kim, H.Y.; Lee, M.J.; Heo, A.J.; Park, D.Y.; Lim, S.; Shin, S.; Ganipisetti, S.; Yang, W.S.; Jung, C.A.; et al. The AUTOTAC chemical biology platform for targeted protein degradation via the autophagy-lysosome system. Nat. Commun. 2022, 13, 904. [Google Scholar] [CrossRef] [PubMed]
  33. Fan, X.; Jin, W.Y.; Lu, J.; Wang, J.; Wang, Y.T. Rapid and reversible knockdown of endogenous proteins by peptide-directed lysosomal degradation. Nat. Neurosci. 2014, 17, 471–480. [Google Scholar] [CrossRef]
  34. Zhao, L.; Zhao, J.; Zhong, K.; Tong, A.; Jia, D. Targeted Protein Degradation: Mechanisms, Strategies and Application. Signal Transduct. Target. Ther. 2022, 7, 113. [Google Scholar] [CrossRef] [PubMed]
  35. Pettersson, M.; Crews, C.M. PROteolysis TArgeting Chimeras (PROTACs)—Past, Present and Future. Drug Discov. Today Technol. 2019, 31, 15–27. [Google Scholar] [CrossRef] [PubMed]
  36. Burslem, G.M.; Smith, B.E.; Lai, A.C.; Jaime-Figueroa, S.; McQuaid, D.C.; Bondeson, D.P.; Toure, M.; Dong, H.; Qian, Y.; Wang, J.; et al. The Advantages of Targeted Protein Degradation Over Inhibition: An RTK Case Study. Cell Chem. Biol. 2018, 25, 67–77.e3. [Google Scholar] [CrossRef]
  37. Han, X.; Sun, Y. Review Strategies for the Discovery of Oral PROTAC Degraders Aimed at Cancer Therapy. Cell Rep. Phys. Sci. 2022, 3, 101062. [Google Scholar] [CrossRef]
  38. Clague, M.J.; Heride, C.; Urbé, S. The Demographics of the Ubiquitin System. Trends Cell Biol. 2015, 25, 417–426. [Google Scholar] [CrossRef] [PubMed]
  39. Zaidman, D.; Prilusky, J.; London, N. ProsetTac: Rosetta Based Modeling of PROTAC Mediated Ternary Complexes. J. Chem. Inf. Model. 2020, 60, 4894–4903. [Google Scholar] [CrossRef]
  40. Baca, S.C.; Prandi, D.; Lawrence, M.S.; Mosquera, J.M.; Romanel, A.; Drier, Y.; Park, K.; Kitabayashi, N.; MacDonald, T.Y.; Ghandi, M.; et al. Punctuated Evolution of Prostate Cancer Genomes. Cell 2013, 153, 666–677. [Google Scholar] [CrossRef]
  41. Wang, G.; Zhao, D.; Spring, D.J.; Depinho, R.A. Genetics and Biology of Prostate Cancer. Genes Dev. 2018, 32, 1105–1140. [Google Scholar] [CrossRef]
  42. Dong, J.T. Prevalent Mutations in Prostate Cancer. J. Cell Biochem. 2006, 97, 433–447. [Google Scholar] [CrossRef]
  43. Robinson, D.; van Allen, E.M.; Wu, Y.M.; Schultz, N.; Lonigro, R.J.; Mosquera, J.M.; Montgomery, B.; Taplin, M.E.; Pritchard, C.C.; Attard, G.; et al. Integrative Clinical Genomics of Advanced Prostate Cancer. Cell 2015, 161, 1215–1228. [Google Scholar] [CrossRef]
  44. Eeles, R.; Goh, C.; Castro, E.; Bancroft, E.; Guy, M.; Olama, A.A.; Easton, D.; Kote-Jarai, Z. The Genetic Epidemiology of Prostate Cancer and Its Clinical Implications. Nat. Rev. Urol. 2014, 11, 18–31. [Google Scholar] [CrossRef] [PubMed]
  45. Lozano, R.; Castro, E.; Aragón, I.M.; Cendón, Y.; Cattrini, C.; López-Casas, P.P.; Olmos, D. Genetic Aberrations in DNA Repair Pathways: A Cornerstone of Precision Oncology in Prostate Cancer. Br. J. Cancer 2021, 124, 552–563. [Google Scholar] [CrossRef]
  46. Sizemore, G.M.; Pitarresi, J.R.; Balakrishnan, S.; Ostrowski, M.C. The ETS Family of Oncogenic Transcription Factors in Solid Tumours. Nat. Rev. Cancer 2017, 17, 337–351. [Google Scholar] [CrossRef]
  47. Grasso, C.S.; Wu, Y.M.; Robinson, D.R.; Cao, X.; Dhanasekaran, S.M.; Khan, A.P.; Quist, M.J.; Jing, X.; Lonigro, R.J.; Brenner, J.C.; et al. The Mutational Landscape of Lethal Castration-Resistant Prostate Cancer. Nature 2012, 487, 239–243. [Google Scholar] [CrossRef] [PubMed]
  48. Srivastava, M.; Montagna, C.; Leighton, X.; Glasman, M.; Naga, S.; Eidelman, O.; Ried, T.; Pollard, H.B. Haploinsufficiency of Anx7 Tumor Suppressor Gene and Consequent Genomic Instability Promotes Tumorigenesis in the Anx7(+/−) Mouse. Proc. Natl. Acad. Sci. USA 2003, 100, 14287–14292. [Google Scholar] [CrossRef] [PubMed]
  49. Sun, X.; Frierson, H.F.; Chen, C.; Li, C.; Ran, Q.; Otto, K.B.; Cantarel, B.M.; Vessella, R.L.; Gao, A.C.; Petros, J.; et al. Frequent Somatic Mutations of the Transcription Factor ATBF1 in Human Prostate Cancer. Nat. Genet. 2005, 37, 407–412. [Google Scholar] [CrossRef]
  50. Kibel, A.S.; Suarez, B.K.; Belani, J.; Oh, J.; Webster, R.; Brophy-Ebbers, M.; Guo, C.; Catalona, W.J.; Picus, J.; Goodfellow, P.J. CDKN1A and CDKN1B Polymorphisms and Risk of Advanced Prostate Carcinoma. Cancer Res. 2003, 63, 2033–2036. [Google Scholar]
  51. Dong, X.; Wang, L.; Taniguchi, K.; Wang, X.; Cunningham, J.M.; McDonnell, S.K.; Qian, C.; Marks, A.F.; Slager, S.L.; Peterson, B.J.; et al. Mutations in CHEK2 Associated with Prostate Cancer Risk. Am. J. Hum. Genet. 2003, 72, 270–280. [Google Scholar] [CrossRef]
  52. Ntais, C.; Polycarpou, A.; Ioannidis, J.P.A. Association of GSTM1, GSTT1, and GSTP1 Gene Polymorphisms with the Risk of Prostate Cancer: A Meta-Analysis. Cancer Epidemiol. Biomark Prev. 2005, 14, 176–181. [Google Scholar] [CrossRef]
  53. Chen, C.; Bhalala, H.v.; Vessella, R.L.; Dong, J.T. KLF5 Is Frequently Deleted and Down-Regulated but Rarely Mutated in Prostate Cancer. Prostate 2003, 55, 81–88. [Google Scholar] [CrossRef] [PubMed]
  54. Narla, G.; Heath, K.E.; Reeves, H.L.; Li, D.; Giono, L.E.; Kimmelman, A.C.; Glucksman, M.J.; Narla, J.; Eng, F.J.; Chan, A.M.; et al. KLF6, a Candidate Tumor Suppressor Gene Mutated in Prostate Cancer. Science 2001, 294, 2563–2566. [Google Scholar] [CrossRef] [PubMed]
  55. Xu, J.; Lilly Zheng, S.; Komiya, A.; Mychaleckyj, J.C.; Isaacs, S.D.; Chang, B.; Turner, A.R.; Ewing, C.M.; Wiley, K.E.; Hawkins, G.A.; et al. Common Sequence Variants of the Macrophage Scavenger Receptor 1 Gene Are Associated with Prostate Cancer Risk. Am. J. Hum. Genet. 2003, 72, 208–212. [Google Scholar] [CrossRef] [PubMed]
  56. Hubbard, G.K.; Mutton, L.N.; Khalili, M.; McMullin, R.P.; Hicks, J.L.; Bianchi-Frias, D.; Horn, L.A.; Kulac, I.; Moubarek, M.S.; Nelson, P.S.; et al. Combined MYC Activation and Pten Loss Are Sufficient to Create Genomic Instability and Lethal Metastatic Prostate Cancer. Cancer Res. 2016, 76, 283–292. [Google Scholar] [CrossRef]
  57. Abdulkadir, S.A.; Magee, J.A.; Peters, T.J.; Kaleem, Z.; Naughton, C.K.; Humphrey, P.A.; Milbrandt, J. Conditional Loss of Nkx3.1 in Adult Mice Induces Prostatic Intraepithelial Neoplasia. Mol. Cell Biol. 2002, 22, 1495–1503. [Google Scholar] [CrossRef] [PubMed]
  58. Beltran, H.; Prandi, D.; Mosquera, J.M.; Benelli, M.; Puca, L.; Cyrta, J.; Marotz, C.; Giannopoulou, E.; Chakravarthi, B.V.S.K.; Varambally, S.; et al. Divergent Clonal Evolution of Castration-Resistant Neuroendocrine Prostate Cancer. Nat. Med. 2016, 22, 298–305. [Google Scholar] [CrossRef] [PubMed]
  59. Marchesani, M.; Hakkarainen, A.; Tuomainen, T.P.; Kaikkonen, J.; Pukkala, E.; Uimari, P.; Seppälä, E.; Matikainen, M.; Kallioniemi, O.P.; Schleutker, J.; et al. New Paraoxonase 1 Polymorphism I102V and the Risk of Prostate Cancer in Finnish Men. J. Natl. Cancer Inst. 2003, 95, 812–818. [Google Scholar] [CrossRef]
  60. Carpten, J.; Nupponen, N.; Isaacs, S.; Sood, R.; Robbins, C.; Xu, J.; Faruque, M.; Moses, T.; Ewing, C.; Gillanders, E.; et al. Germline Mutations in the Ribonuclease L Gene in Families Showing Linkage with HPC1. Nat. Genet. 2002, 30, 181–184. [Google Scholar] [CrossRef]
  61. Cicek, M.S.; Conti, D.v.; Curran, A.; Neville, P.J.; Paris, P.L.; Casey, G.; Witte, J.S. Association of Prostate Cancer Risk and Aggressiveness to Androgen Pathway Genes: SRD5A2, CYP17, and the AR. Prostate 2004, 59, 69–76. [Google Scholar] [CrossRef]
  62. Medeiros, R.; Morais, A.; Vasconcelos, A.; Costa, S.; Pinto, D.; Oliveira, J.; Lopes, C. The Role of Vitamin D Receptor Gene Polymorphisms in the Susceptibility to Prostate Cancer of a Southern European Population. J. Hum. Genet. 2002, 47, 413–418. [Google Scholar] [CrossRef]
  63. Newmark, J.R.; Hardy, D.; Tonb, D.C.; Carter, B.S.; Epsteint, J.I.; Isaacs, W.B.; Browns, T.R.; Barrack, E.R. Androgen Receptor Gene Mutations in Human Prostate Cancer (Polymerase Chain Reaction/Denaturing Gradient Gel Electrophoresis/Somatic Mutation). Proc. Natl. Acad. Sci. USA 1992, 89, 6319–6323. [Google Scholar] [CrossRef] [PubMed]
  64. Giovannucci, E.; Stampfer, M.J.; Krithivas, K.; Brown, M.; Brufsky, A.; Talcott, J.; Hennekens, C.H.; Kantoff, P.W. The CAG Repeat within the Androgen Receptor Gene and Its Relationship to Prostate Cancer. Proc. Natl. Acad. Sci. USA 1997, 94, 3320–3323. Available online: www.pnas.org (accessed on 1 April 1997). [CrossRef] [PubMed]
  65. Gockel, L.M.; Pfeifer, V.; Baltes, F.; Bachmaier, R.D.; Wagner, K.G.; Bendas, G.; Gütschow, M.; Sosič, I.; Steinebach, C. Design, Synthesis, and Characterization of PROTACs Targeting the Androgen Receptor in Prostate and Lung Cancer Models. Arch. Pharm. 2022, 355, 156. [Google Scholar] [CrossRef]
  66. Takwale, A.D.; Jo, S.H.; Jeon, Y.U.; Kim, H.S.; Shin, C.H.; Lee, H.K.; Ahn, S.; Lee, C.O.; du Ha, J.; Kim, J.H.; et al. Design and Characterization of Cereblon-Mediated Androgen Receptor Proteolysis-Targeting Chimeras. Eur. J. Med. Chem. 2020, 208, 112769. [Google Scholar] [CrossRef]
  67. Liang, J.J.; Xie, H.; Yang, R.H.; Wang, N.; Zheng, Z.J.; Zhou, C.; Wang, Y.L.; Wang, Z.J.; Liu, H.M.; Shan, L.H.; et al. Designed, Synthesized and Biological Evaluation of Proteolysis Targeting Chimeras (PROTACs) as AR Degraders for Prostate Cancer Treatment. Bioorg. Med. Chem. 2021, 45, 16331. [Google Scholar] [CrossRef]
  68. Kim, G.Y.; Song, C.W.; Yang, Y.S.; Lee, N.R.; Yoo, H.S.; Son, S.H.; Lee, S.J.; Park, J.S.; Lee, J.K.; Inn, K.S. Chemical Degradation of Androgen Receptor (Ar) Using Bicalutamide Analog–Thalidomide Protacs. Molecules 2021, 26, 2525. [Google Scholar] [CrossRef] [PubMed]
  69. Han, X.; Zhao, L.; Xiang, W.; Qin, C.; Miao, B.; McEachern, D.; Wang, Y.; Metwally, H.; Wang, L.; Matvekas, A.; et al. Strategies toward Discovery of Potent and Orally Bioavailable Proteolysis Targeting Chimera Degraders of Androgen Receptor for the Treatment of Prostate Cancer. J. Med. Chem. 2021, 64, 12831–12854. [Google Scholar] [CrossRef] [PubMed]
  70. Xiang, W.; Zhao, L.; Han, X.; Qin, C.; Miao, B.; McEachern, D.; Wang, Y.; Metwally, H.; Kirchhoff, P.D.; Wang, L.; et al. Discovery of ARD-2585 as an Exceptionally Potent and Orally Active PROTAC Degrader of Androgen Receptor for the Treatment of Advanced Prostate Cancer. J. Med. Chem. 2021, 64, 13487–13509. [Google Scholar] [CrossRef]
  71. Han, X.; Zhao, L.; Xiang, W.; Qin, C.; Miao, B.; Xu, T.; Wang, M.; Yang, C.Y.; Chinnaswamy, K.; Stuckey, J.; et al. Discovery of Highly Potent and Efficient PROTAC Degraders of Androgen Receptor (AR) by Employing Weak Binding Affinity VHL E3 Ligase Ligands. J. Med. Chem. 2019, 62, 11218–11231. [Google Scholar] [CrossRef]
  72. Chen, L.; Han, L.; Mao, S.; Xu, P.; Xu, X.; Zhao, R.; Wu, Z.; Zhong, K.; Yu, G.; Wang, X. Discovery of A031 as Effective Proteolysis Targeting Chimera (PROTAC) Androgen Receptor (AR) Degrader for the Treatment of Prostate Cancer. Eur. J. Med. Chem. 2021, 216, 113307. [Google Scholar] [CrossRef] [PubMed]
  73. Munoz, E.; Chen, G.; Hossain, A.; Wu, S.; Oceguera Nava, E.; Hang, J.; Lee, T.; Zhang, Q.; Wang, G.; Chen, Q.-H. Synthesis and Biological Evaluation of Niclosamide PROTACs. Bioorg. Med. Chem. Lett. 2022, 72, 128870. [Google Scholar] [CrossRef]
  74. Lee, G.T.; Nagaya, N.; Desantis, J.; Madura, K.; Sabaawy, H.E.; Kim, W.J.; Vaz, R.J.; Cruciani, G.; Kim, I.Y. Effects of MTX-23, a Novel PROTAC of Androgen Receptor Splice Variant-7 and Androgen Receptor, on CRPC Resistant to Second-Line Antiandrogen Therapy. Mol. Cancer Ther. 2021, 20, 490–499. [Google Scholar] [CrossRef] [PubMed]
  75. Salami, J.; Alabi, S.; Willard, R.R.; Vitale, N.J.; Wang, J.; Dong, H.; Jin, M.; McDonnell, D.P.; Crew, A.P.; Neklesa, T.K.; et al. Androgen Receptor Degradation by the Proteolysis-Targeting Chimera ARCC-4 Outperforms Enzalutamide in Cellular Models of Prostate Cancer Drug Resistance. Commun. Biol. 2018, 1, 100. [Google Scholar] [CrossRef] [PubMed]
  76. Han, X.; Wang, C.; Qin, C.; Xiang, W.; Fernandez-Salas, E.; Yang, C.Y.; Wang, M.; Zhao, L.; Xu, T.; Chinnaswamy, K.; et al. Discovery of ARD-69 as a Highly Potent Proteolysis Targeting Chimera (PROTAC) Degrader of Androgen Receptor (AR) for the Treatment of Prostate Cancer. J. Med. Chem. 2019, 62, 941–964. [Google Scholar] [CrossRef]
  77. Kregel, S.; Wang, C.; Han, X.; Xiao, L.; Fernandez-Salas, E.; Bawa, P.; McCollum, B.L.; Wilder-Romans, K.; Apel, I.J.; Cao, X.; et al. Androgen Receptor Degraders Overcome Common Resistance Mechanisms Developed during Prostate Cancer Treatment. Neoplasia 2020, 22, 111–119. [Google Scholar] [CrossRef]
  78. Ma, B.; Fan, Y.; Zhang, D.; Wei, Y.; Jian, Y.; Liu, D.; Wang, Z.; Gao, Y.; Ma, J.; Chen, Y.; et al. De Novo Design of an Androgen Receptor DNA Binding Domain-Targeted Peptide PROTAC for Prostate Cancer Therapy. Adv. Sci. 2022, 9, 2201859. [Google Scholar] [CrossRef]
  79. Xie, H.; Liang, J.J.; Wang, Y.L.; Hu, T.X.; Wang, J.Y.; Yang, R.H.; Yan, J.K.; Zhang, Q.R.; Xu, X.; Liu, H.M.; et al. The Design, Synthesis and Anti-Tumor Mechanism Study of New Androgen Receptor Degrader. Eur. J. Med. Chem. 2020, 204, 112512. [Google Scholar] [CrossRef]
  80. Raina, K.; Lu, J.; Qian, Y.; Altieri, M.; Gordon, D.; Rossi, A.M.K.; Wang, J.; Chen, X.; Dong, H.; Siu, K.; et al. PROTAC-Induced BET Protein Degradation as a Therapy for Castration-Resistant Prostate Cancer. Proc. Natl. Acad. Sci. USA 2016, 113, 7124–7129. [Google Scholar] [CrossRef]
  81. Hu, R.; Wang, W.L.; Yang, Y.Y.; Hu, X.T.; Wang, Q.W.; Zuo, W.Q.; Xu, Y.; Feng, Q.; Wang, N.Y. Identification of a Selective BRD4 PROTAC with Potent Antiproliferative Effects in AR-Positive Prostate Cancer Based on a Dual BET/PLK1 Inhibitor. Eur. J. Med. Chem. 2022, 227, 113922. [Google Scholar] [CrossRef]
  82. Zhou, F.; Chen, L.; Cao, C.; Yu, J.; Luo, X.; Zhou, P.; Zhao, L.; Du, W.; Cheng, J.; Xie, Y.; et al. Development of Selective Mono or Dual PROTAC Degrader Probe of CDK Isoforms. Eur. J. Med. Chem. 2020, 187, 111952. [Google Scholar] [CrossRef]
  83. Schneekloth, A.R.; Pucheault, M.; Tae, H.S.; Crews, C.M. Targeted Intracellular Protein Degradation Induced by a Small Molecule: En Route to Chemical Proteomics. Bioorg. Med. Chem. Lett. 2008, 18, 5904–5908. [Google Scholar] [CrossRef]
  84. Li, W.; Bengtson, M.H.; Ulbrich, A.; Matsuda, A.; Reddy, V.A.; Orth, A.; Chanda, S.K.; Batalov, S.; Joazeiro, C.A.P. Genome-Wide and Functional Annotation of Human E3 Ubiquitin Ligases Identifies MULAN, a Mitochondrial E3 That Regulates the Organelle’s Dynamics and Signaling. PLoS ONE 2008, 3, e1487. [Google Scholar] [CrossRef]
  85. Guenette, R.G.; Yang, S.W.; Min, J.; Pei, B.; Potts, P.R. Target and Tissue Selectivity of PROTAC Degraders. Chem. Soc. Rev. 2022, 51, 5740–5756. [Google Scholar] [CrossRef] [PubMed]
  86. Cao, C.; He, M.; Wang, L.; He, Y.; Rao, Y. Chemistries of Bifunctional PROTAC Degraders. Chem. Soc. Rev. 2022, 51, 7066–7114. [Google Scholar] [CrossRef]
  87. Ishida, T.; Ciulli, A. E3 Ligase Ligands for PROTACs: How They Were Found and How to Discover New Ones. SLAS Discov. 2021, 26, 484–502. [Google Scholar] [CrossRef] [PubMed]
  88. Troup, R.I.; Fallan, C.; Baud, M.G.J. Current Strategies for the Design of PROTAC Linkers: A Critical Review. Explor. Target Antitumor. Ther. 2020, 1, 273–312. [Google Scholar] [CrossRef]
  89. Hughes, S.J.; Testa, A.; Thompson, N.; Churcher, I. The rise and rise of protein degradation: Opportunities and challenges ahead. Drug Discov. Today 2021, 26, 2889–2897. [Google Scholar] [CrossRef]
  90. Dey, S.K.; Jaffrey, S.R. RIBOTACs: Small Molecules Target RNA for Degradation. Cell Chem. Biol. 2019, 26, 1047–1049. [Google Scholar] [CrossRef] [PubMed]
  91. Li, F.; Hu, Q.; Zhang, X.; Sun, R.; Liu, Z.; Wu, S.; Tian, S.; Ma, X.; Dai, Z.; Yang, X.; et al. DeepPROTACs is a deep learning-based targeted degradation predictor for PROTACs. Nat. Commun. 2022, 13, 7133. [Google Scholar] [CrossRef]
  92. Bemis, T.A.; La Clair, J.J.; Burkart, M.D. Unraveling the role of linker design in proteolysis targeting chimeras: Miniperspective. J. Med. Chem. 2021, 64, 8042–8052. [Google Scholar] [CrossRef] [PubMed]
  93. Cao, C.; Yang, J.; Chen, Y.; Zhou, P.; Wang, Y.; Du, W.; Zhao, L.; Chen, Y. Discovery of SK-575 as a Highly Potent and Efficacious ProteolysisTargeting Chimera Degrader of PARP1 for Treating Cancers. J. Med. Chem. 2020, 63, 11012–11033. [Google Scholar] [CrossRef]
  94. Qu, X.; Liu, H.; Song, X.; Sun, N.; Zhong, H.; Qiu, X.; Yang, X.; Jiang, B. Effective degradation of EGFRL858R+T790M mutant proteins by CRBN-based PROTACs through both proteosome and autophagy/lysosome degradation systems. Eur. J. Med. Chem. 2021, 218, 113328. [Google Scholar] [CrossRef] [PubMed]
  95. Sohrabi, M.; Saeedi, M.; Larijani, B.; Mahdavi, M. Recent advances in biological activities of rhodium complexes: Their applications in drug discovery research. Eur. J. Med. Chem. 2021, 216, 113308. [Google Scholar] [CrossRef]
  96. Jansa, J.; Jorda, R.; Skerlova, J.; Pachl, P.; Perina, M.; Reznickova, E.; Heger, T.; Gucky, T.; Rezacova, P.; Lycka, A.; et al. Imidazo[1,2c]pyrimidin-5(6H)-one inhibitors of CDK2: Synthesis, kinase inhibition and co-crystal structure. Eur. J. Med. Chem. 2021, 216, 113309. [Google Scholar] [CrossRef] [PubMed]
  97. Sang, Z.; Song, Q.; Cao, Z.; Deng, Y.; Tan, Z.; Zhang, L. Design, synthesis and evaluation of novel dimethylamino chalcone-O-alkylamines derivatives as potential multifunctional agents against Alzheimer’s disease. Eur. J. Med. Chem. 2021, 216, 113310. [Google Scholar] [CrossRef]
  98. Liu, Y.Y.; Guo, Z.; Wang, J.Y.; Wang, H.M.; Da Qi, J.; Ma, J.; Piao, H.R.; Jin, C.H.; Jin, X. Synthesis and evaluation of the epithelial-to- mesenchymal inhibitory activity of indazole-derived imidazoles as dual ALK5/p38alpha MAP inhibitors. Eur. J. Med. Chem. 2021, 216, 113311. [Google Scholar] [CrossRef]
  99. Singh, S.S.; Mattheolabakis, G.; Gu, X.; Withers, S.; Dahal, A.; Jois, S. A grafted peptidomimetic for EGFR heterodimerization inhibition: Implications in NSCLC models. Eur. J. Med. Chem. 2021, 216, 113312. [Google Scholar] [CrossRef]
  100. Lu, R.; Wang, Y.; Liu, C.; Zhang, Z.; Li, B.; Meng, Z.; Jiang, C.; Hu, Q. Design, synthesis and evaluation of 3-amide-5-aryl benzoic acid derivatives as novel P2Y(14)R antagonists with potential high efficiency against acute gouty arthritis. Eur. J. Med. Chem. 2021, 216, 113313. [Google Scholar] [CrossRef]
  101. Zhang, Z.; Xing, X.; Guan, P.; Song, S.; You, G.; Xia, C.; Liu, T. Recent progress in agents targeting polo-like kinases: Promising therapeutic strategies. Eur. J. Med. Chem. 2021, 217, 113314. [Google Scholar] [CrossRef]
  102. Villard, A.L.; Aubertin, A.M.; Peyrottes, S.; Perigaud, C. An original pronucleotide strategy for the simultaneous delivery of two bioactive drugs. Eur. J. Med. Chem. 2021, 216, 113315. [Google Scholar] [CrossRef] [PubMed]
  103. Zhu, H.; Li, W.; Shuai, W.; Liu, Y.; Yang, L.; Tan, Y.; Zheng, T.; Yao, H.; Xu, J.; Zhu, Z.; et al. Discovery of novel Nbenzylbenzamide derivatives as tubulin polymerization inhibitors with potent antitumor activities. Eur. J. Med. Chem. 2021, 216, 113316. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, C.; Gao, M.; Liu, S.; Zou, Z.; Ren, R.; Zhang, C.; Xie, H.; Sun, J.; Qi, Y.; Qu, Q.; et al. Pyxinol bearing amino acid residues: Easily achievable and promising modulators of P-glycoprotein-mediated multidrug resistance. Eur. J. Med. Chem. 2021, 216, 113317. [Google Scholar] [CrossRef] [PubMed]
  105. Lee, S.; Kim, J.; Jo, J.; Chang, J.W.; Sim, J.; Yun, H. Recent advances in development of hetero-bivalent kinase inhibitors. Eur. J. Med. Chem. 2021, 216, 113318. [Google Scholar] [CrossRef]
  106. Li, Z.; Xin, W.; Wang, Q.; Zhu, M.; Zhou, H. Design and synthesis of N-(3-sulfamoylphenyl)amides as Trypanosoma brucei leucyl-tRNA synthetase inhibitors. Eur. J. Med. Chem. 2021, 217, 113319. [Google Scholar] [CrossRef]
  107. Srivastava, S.; Ahmad, R.; Khare, S.K. Alzheimer’s disease and its treatment by different approaches: A review. Eur. J. Med. Chem. 2021, 216, 113320. [Google Scholar] [CrossRef]
  108. Aaghaz, S.; Sharma, K.; Jain, R.; Kamal, A. beta-Carbolines as potential anticancer agents. Eur. J. Med. Chem. 2021, 216, 113321. [Google Scholar] [CrossRef]
  109. Li, Y.S.; He, M.; Zhou, T.S.; Wang, Q.; He, L.; Wang, S.J.; Hu, B.; Wei, B.; Wang, H.; Cui, Z.N. 2,5-Disubstituted furan derivatives containing 1,3,4-thiadiazole moiety as potent alpha-glucosidase and E. coli beta-glucuronidase inhibitors. Eur. J. Med. Chem. 2021, 216, 113322. [Google Scholar] [CrossRef]
  110. Jo, J.; Lee, D.; Park, Y.H.; Choi, H.; Han, J.; Park, D.H.; Choi, Y.K.; Kwak, J.; Yang, M.K.; Yoo, J.W.; et al. Discovery and optimization of novel 3-benzyl-N-phenyl-1H-pyrazole-5-carboxamides as bifunctional antidiabetic agents stimulating both insulin secretion and glucose uptake. Eur. J. Med. Chem. 2021, 217, 113325. [Google Scholar] [CrossRef]
  111. Zhou, W.; Xu, C.; Dong, G.; Qiao, H.; Yang, J.; Liu, H.; Ding, L.; Sun, K.; Zhao, W. Development of phenyltriazole thiol-based derivatives as highly potent inhibitors of DCN1-UBC12 interaction. Eur. J. Med. Chem. 2021, 217, 113326. [Google Scholar] [CrossRef]
  112. Wu, S.; Xu, C.; Xia, K.; Lin, Y.; Tian, S.; Ma, H.; Ji, Y.; Zhu, F.; He, S.; Zhang, X. Ring closure strategy leads to potent RIPK3 inhibitors. Eur. J. Med. Chem. 2021, 217, 113327. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, J.; Chen, C.; Wang, D.; Zhang, J.; Zhang, T. Emerging small-molecule inhibitors of the Bruton’s tyrosine kinase (BTK): Current development. Eur. J. Med. Chem. 2021, 217, 113329. [Google Scholar] [CrossRef] [PubMed]
  114. Kayamba, F.; Malimabe, T.; Ademola, I.K.; Pooe, O.J.; Kushwaha, N.D.; Mahlalela, M.; van Zyl, R.L.; Gordon, M.; Mudau, P.T.; Zininga, T.; et al. Design and synthesis of quinoline-pyrimidine inspired hybrids as potential plasmodial inhibitors. Eur. J. Med. Chem. 2021, 217, 113330. [Google Scholar] [CrossRef] [PubMed]
  115. Grieco, I.; Bissaro, M.; Tiz, D.B.; Perez, D.I.; Perez, C.; Martinez, A.; Redenti, S.; Mariotto, E.; Bortolozzi, R.; Viola, G.; et al. Developing novel classes of protein kinase CK1delta inhibitors by fusing [1,2,4]triazole with different bicyclic heteroaromatic systems. Eur. J. Med. Chem. 2021, 216, 113331. [Google Scholar] [CrossRef]
  116. Gediya, P.; Parikh, P.K.; Vyas, V.K.; Ghate, M.D. Histone deacetylase 2: A potential therapeutic target for cancer and neurodegenerative disorders. Eur. J. Med. Chem. 2021, 216, 113332. [Google Scholar] [CrossRef]
  117. Gao, D.; Jin, N.; Fu, Y.; Zhu, Y.; Wang, Y.; Wang, T.; Chen, Y.; Zhang, M.; Xiao, Q.; Huang, M.; et al. Rational drug design of benzothiazole-based derivatives as potent signal transducer and activator of transcription 3 (STAT3) signaling pathway inhibitors. Eur. J. Med. Chem. 2021, 216, 113333. [Google Scholar] [CrossRef]
  118. Dhokne, P.; Sakla, A.P.; Shankaraiah, N. Structural insights of oxindole based kinase inhibitors as anticancer agents: Recent advances. Eur. J. Med. Chem. 2021, 216, 113334. [Google Scholar] [CrossRef]
  119. Ren, C.; Sun, N.; Kong, Y.; Qu, X.; Liu, H.; Zhong, H.; Song, X.; Yang, X.; Jiang, B. Structure-based discovery of SIAIS001 as an oral bioavailability ALK degrader constructed from Alectinib. Eur. J. Med. Chem. 2021, 217, 113335. [Google Scholar] [CrossRef]
  120. Shim, J.H. Inhibitory Effects of Cycloheterophyllin on Melanin Synthesis. Molecules 2021, 26, 2526. [Google Scholar] [CrossRef]
  121. Azzouz, Z.; Bettache, A.; Boucherba, N.; Prieto, A.; Martinez, M.J.; Benallaoua, S.; de Eugenio, L.I. Optimization of beta-1,4-Endoxylanase Production by an Aspergillus niger Strain Growing on Wheat Straw and Application in Xylooligosaccharides Production. Molecules 2021, 26, 2527. [Google Scholar] [CrossRef] [PubMed]
  122. Rubino, F.M. The Redox Potential of the beta-(93)-Cysteine Thiol Group in Human Hemoglobin Estimated from In Vitro Oxidant Challenge Experiments. Molecules 2021, 26, 2528. [Google Scholar] [CrossRef] [PubMed]
  123. Kim, H.; Yang, W.S.; Htwe, K.M.; Lee, M.N.; Kim, Y.D.; Yoon, K.D.; Lee, B.H.; Lee, S.; Cho, J.Y. Dipterocarpus tuberculatus Roxb. Ethanol Extract Has Anti-Inflammatory and Hepatoprotective Effects In Vitro and In Vivo by Targeting the IRAK1/AP-1 Pathway. Molecules 2021, 26, 2529. [Google Scholar] [CrossRef] [PubMed]
  124. Budeev, A.; Kantin, G.; Dar’in, D.; Krasavin, M. Diazocarbonyl and Related Compounds in the Synthesis of Azoles. Molecules 2021, 26, 2530. [Google Scholar] [CrossRef]
  125. Dao, T.B.; Nguyen, T.M.; Nguyen, V.Q.; Tran, T.M.; Tran, N.M.; Nguyen, C.H.; Nguyen, T.H.; Nguyen, H.H.; Sichaem, J.; Tran, C.L.; et al. Flavones from Combretum quadrangulare Growing in Vietnam and Their Alpha-Glucosidase Inhibitory Activity. Molecules 2021, 26, 2531. [Google Scholar] [CrossRef] [PubMed]
  126. Zhen, L.; Lange, H.; Crestini, C. An Analytical Toolbox for Fast and Straightforward Structural Characterisation of Commercially Available Tannins. Molecules 2021, 26, 2532. [Google Scholar] [CrossRef] [PubMed]
  127. Kato, I.; Ukai, Y.; Kondo, N.; Nozu, K.; Kimura, C.; Hashimoto, K.; Mizusawa, E.; Maki, H.; Naito, A.; Kawai, M. Identification of Thiazoyl Guanidine Derivatives as Novel Antifungal Agents Inhibiting Ergosterol Biosynthesis for Treatment of Invasive Fungal Infections. J. Med. Chem. 2021, 64, 10482–10496. [Google Scholar] [CrossRef]
  128. Jimenez, T.; Botero, J.; Otaegui, D.; Calvo, J.; Hernandez, F.J.; San Sebastian, E. Rational Design and Experimental Analysis of Short-Oligonucleotide Substrate Specificity for Targeting Bacterial Nucleases. J. Med. Chem. 2021, 64, 12855–12864. [Google Scholar] [CrossRef]
  129. Rossino, G.; Rui, M.; Linciano, P.; Rossi, D.; Boiocchi, M.; Peviani, M.; Poggio, E.; Curti, D.; Schepmann, D.; Wunsch, B.; et al. Bitopic Sigma 1 Receptor Modulators to Shed Light on Molecular Mechanisms Underpinning Ligand Binding and Receptor Oligomerization. J. Med. Chem. 2021, 64, 14997–15016. [Google Scholar] [CrossRef]
  130. Plesselova, S.; Garcia-Cerezo, P.; Blanco, V.; Reche-Perez, F.J.; Hernandez-Mateo, F.; Santoyo-Gonzalez, F.; Giron-Gonzalez, M.D.; Salto-Gonzalez, R. Polyethylenimine-Bisphosphonate-Cyclodextrin Ternary Conjugates: Supramolecular Systems for the Delivery of Antineoplastic Drugs. J. Med. Chem. 2021, 64, 12245–12260. [Google Scholar] [CrossRef]
  131. Ryan, M.D.; Parkes, A.L.; Corbett, D.; Dickie, A.P.; Southey, M.; Andersen, O.A.; Stein, D.B.; Barbeau, O.R.; Sanzone, A.; Thommes, P.; et al. Discovery of Novel UDP-N-Acetylglucosamine Acyltransferase (LpxA) Inhibitors with Activity against Pseudomonas aeruginosa. J. Med. Chem. 2021, 64, 14377–14425. [Google Scholar] [CrossRef]
  132. Mann, M.K.; Zepeda-Velazquez, C.A.; Gonzalez-Alvarez, H.; Dong, A.; Kiyota, T.; Aman, A.M.; Loppnau, P.; Li, Y.; Wilson, B.; Arrowsmith, C.H.; et al. Structure-Activity Relationship of USP5 Inhibitors. J. Med. Chem. 2021, 64, 1501715036. [Google Scholar] [CrossRef] [PubMed]
  133. Zhang, Z.; Ghosh, A.; Connolly, P.J.; King, P.; Wilde, T.; Wang, J.; Dong, Y.; Li, X.; Liao, D.; Chen, H.; et al. GutRestricted Selective Cyclooxygenase-2 (COX-2) Inhibitors for Chemoprevention of Colorectal Cancer. J. Med. Chem. 2021, 64, 11570–11596. [Google Scholar] [CrossRef]
  134. Singh, H.; Thirupathi, A.; Das, B.; Janni, M.; Kumari, R.; Singh, S.; Rashid, M.; Wahajuddin, M.; Balamurali, M.M.; Jagavelu, K.; et al. 2,3-Difunctionalized Benzo[b]thiophene Scaffolds Possessing Potent Antiangiogenic Properties. J. Med. Chem. 2022, 65, 120134. [Google Scholar] [CrossRef] [PubMed]
  135. Facchini, F.A.; Minotti, A.; Luraghi, A.; Romerio, A.; Gotri, N.; Matamoros-Recio, A.; Iannucci, A.; Palmer, C.; Wang, G.; Ingram, R.; et al. Synthetic Glycolipids as Molecular Vaccine Adjuvants: Mechanism of Action in Human Cells and In Vivo Activity. J. Med. Chem. 2021, 64, 12261–12272. [Google Scholar] [CrossRef]
  136. Chen, H.; Liu, J.; Kaniskan, H.U.; Wei, W.; Jin, J. Folate-Guided Protein Degradation by Immunomodulatory Imide Drug-Based Molecular Glues and Proteolysis Targeting Chimeras. J. Med. Chem. 2021, 64, 12273–12285. [Google Scholar] [CrossRef] [PubMed]
  137. Bowden, G.D.; Stotz, S.; Kinzler, J.; Geibel, C.; Lammerhofer, M.; Pichler, B.J.; Maurer, A. DoE Optimization Empowers the Automated Preparation of Enantiomerically Pure [(18)F]Talazoparib and its In Vivo Evaluation as a PARP Radiotracer. J. Med. Chem. 2021, 64, 1569015701. [Google Scholar] [CrossRef]
  138. Bursavich, M.G.; Harrison, B.A.; Acharya, R.; Costa, D.E.; Freeman, E.A.; Hrdlicka, L.A.; Jin, H.; Kapadnis, S.; Moffit, J.S.; Murphy, D.; et al. Discovery of the Oxadiazine FRM-024: A Potent CNSPenetrant Gamma Secretase Modulator. J. Med. Chem. 2021, 64, 14426–14447. [Google Scholar] [CrossRef]
  139. Heightman, T.D.; Berdini, V.; Bevan, L.; Buck, I.M.; Carr, M.G.; Courtin, A.; Coyle, J.E.; Day, J.E.H.; East, C.; Fazal, L.; et al. Discovery of ASTX029, A Clinical Candidate Which Modulates the Phosphorylation and Catalytic Activity of ERK1/2. J. Med. Chem. 2021, 64, 12286–12303. [Google Scholar] [CrossRef]
  140. Xiong, B.; Wang, Y.; Chen, Y.; Xing, S.; Liao, Q.; Chen, Y.; Li, Q.; Li, W.; Sun, H. Strategies for Structural Modification of Small Molecules to Improve Blood-Brain Barrier Penetration: A Recent Perspective. J. Med. Chem. 2021, 64, 13152–13173. [Google Scholar] [CrossRef]
  141. Deng, X.; Salgado-Polo, F.; Shao, T.; Xiao, Z.; Van, R.; Chen, J.; Rong, J.; Haider, A.; Shao, Y.; Josephson, L.; et al. Imaging Autotaxin In Vivo with (18)F-Labeled Positron Emission Tomography Ligands. J. Med. Chem. 2021, 64, 15053–15068. [Google Scholar] [CrossRef]
  142. Fallica, A.N.; Barbaraci, C.; Amata, E.; Pasquinucci, L.; Turnaturi, R.; Dichiara, M.; Intagliata, S.; Gariboldi, M.B.; Marras, E.; Orlandi, V.T.; et al. Nitric Oxide Photo-Donor Hybrids of Ciprofloxacin and Norfloxacin: A Shift in Activity from Antimicrobial to Anticancer Agents. J. Med. Chem. 2021, 64, 11597–11613. [Google Scholar] [CrossRef]
  143. Monsarrat, C.; Compain, G.; Andre, C.; Engilberge, S.; Martiel, I.; Olieric, V.; Wolff, P.; Brillet, K.; Landolfo, M.; Silva da Veiga, C.; et al. Iterative Structure-Based Optimization of Short Peptides Targeting the Bacterial Sliding Clamp. J. Med. Chem. 2021, 64, 17063–17078. [Google Scholar] [CrossRef] [PubMed]
  144. McCoull, W.; Boyd, S.; Brown, M.R.; Coen, M.; Collingwood, O.; Davies, N.L.; Doherty, A.; Fairley, G.; Goldberg, K.; Hardaker, E.; et al. Optimization of an Imidazo[1,2-a]pyridine Series to Afford Highly Selective Type I1/2 Dual Mer/Axl Kinase Inhibitors with In Vivo Efficacy. J. Med. Chem. 2021, 64, 13524–13539. [Google Scholar] [CrossRef] [PubMed]
  145. Cristina Mendonca Nogueira, T.; Vinicius Nora de Souza, M. New FDA oncology small molecule drugs approvals in 2020: Mechanism of action and clinical applications. Bioorganic Med. Chem. 2021, 46, 116340. [Google Scholar] [CrossRef]
  146. Sun, J.; Wang, J.; Wang, X.; Hu, X.; Cao, H.; Bai, J.; Li, D.; Hua, H. Design and synthesis of beta-carboline derivatives with nitrogen mustard moieties against breast cancer. Bioorganic Med. Chem. 2021, 45, 116341. [Google Scholar] [CrossRef]
  147. Jiang, Y.; Liu, S.; Tian, G.; Cheung, H.J.H.; Li, X.; Li, X.D. Concise solid-phase synthesis enables derivatisation of YEATS domain cyclopeptide inhibitors for improved cellular uptake. Bioorganic Med. Chem. 2021, 45, 116342. [Google Scholar] [CrossRef]
  148. Sato, J.; Kusano, H.; Aoki, T.; Shibuya, S.; Yokoo, K.; Komano, K.; Oguma, T.; Matsumoto, S.; Sato, T.; Yasuo, K.; et al. A novel tricyclic beta-lactam exhibiting potent antibacterial activities against carbapenem-resistant Enterobacterales: Synthesis and structure-activity-relationships. Bioorganic Med. Chem. 2021, 46, 116343. [Google Scholar] [CrossRef]
  149. Yao, H.; Guo, Q.; Wang, M.; Wang, R.; Xu, Z. Discovery of pyrazole N-aryl sulfonate: A novel and highly potent cyclooxygenase-2 (COX-2) selective inhibitors. Bioorganic Med. Chem. 2021, 46, 116344. [Google Scholar] [CrossRef] [PubMed]
  150. Chia, J.Y.; Miki, T.; Mihara, H.; Tsutsumi, H. Biofunctional supramolecular hydrogels fabricated from a short self-assembling peptide modified with bioactive sequences for the 3D culture of breast cancer MCF-7 cells. Bioorganic Med. Chem. 2021, 46, 116345. [Google Scholar] [CrossRef]
  151. Yang, H.; Li, Q.; Su, M.; Luo, F.; Liu, Y.; Wang, D.; Fan, Y. Design, synthesis, and biological evaluation of novel 6-(pyridin-3-yl) quinazolin-4(3H)one derivatives as potential anticancer agents via PI3K inhibition. Bioorganic Med. Chem. 2021, 46, 116346. [Google Scholar] [CrossRef] [PubMed]
  152. Tsai, W.C.; Aleem, A.M.; Tena, J.; Rivera-Velazquez, M.; Brah, H.S.; Tripathi, S.; D’Silva, M.; Nadler, J.L.; Kalyanaraman, C.; Jacobson, M.P.; et al. Docking and mutagenesis studies lead to improved inhibitor development of ML355 for human platelet 12-lipoxygenase. Bioorganic Med. Chem. 2021, 46, 116347. [Google Scholar] [CrossRef] [PubMed]
  153. Toviwek, B.; Phuangsawai, O.; Konsue, A.; Hannongbua, S.; Riley, J.; Mutter, N.; Anderson, M.; Webster, L.; Hallyburton, I.; Read, K.D.; et al. Preparation, biological & cheminformatics-based assessment of N(2),N(4)-diphenylpyrimidine-2,4-diamine as potential Kinase-targeted antimalarials. Bioorganic Med. Chem. 2021, 46, 116348. [Google Scholar]
  154. Tsai, W.C.; Gilbert, N.C.; Ohler, A.; Armstrong, M.; Perry, S.; Kalyanaraman, C.; Yasgar, A.; Rai, G.; Simeonov, A.; Jadhav, A.; et al. Kinetic and structural investigations of novel inhibitors of human epithelial 15-lipoxygenase-2. Bioorganic Med. Chem. 2021, 46, 116349. [Google Scholar] [CrossRef]
  155. Wang, B.; Feng, W.; Wang, J.; Dong, Y.; Liu, Y.; Yao, Y.; Zhang, J.; Shi, W.; Liu, L.; Zhang, H.; et al. Discovery of potent and selective Bcl-2 inhibitors with acyl sulfonamide skeleton. Bioorganic Med. Chem. 2021, 47, 116350. [Google Scholar] [CrossRef] [PubMed]
  156. Biteau, N.G.; Roy, V.; Lambry, J.C.; Becker, H.F.; Myllykallio, H.; Agrofoglio, L.A. Synthesis of acyclic nucleoside phosphonates targeting flavindependent thymidylate synthase in Mycobacterium tuberculosis. Bioorganic Med. Chem. 2021, 46, 116351. [Google Scholar] [CrossRef]
  157. Ma, Z.; Jiang, L.; Li, B.; Liang, D.; Feng, Y.; Liu, L.; Jiang, C. Discovery of benzimidazole derivatives as potent and selective aldehyde dehydrogenase 1A1 (ALDH1A1) inhibitors with glucose consumption improving activity. Bioorganic Med. Chem. 2021, 46, 116352. [Google Scholar] [CrossRef]
  158. Chang, T.C.; Tanaka, K. In vivo organic synthesis by metal catalysts. Bioorganic Med. Chem. 2021, 46, 116353. [Google Scholar] [CrossRef]
  159. Kumar, S.; Mittal, A.; Mittal, A. A review upon medicinal perspective and designing rationale of DPP-4 inhibitors. Bioorganic Med. Chem. 2021, 46, 116354. [Google Scholar] [CrossRef]
  160. Schwarthoff, S.; Tischer, N.; Sager, H.; Schatz, B.; Rohrbach, M.M.; Raztsou, I.; Robaa, D.; Gaube, F.; Arndt, H.D.; Winckler, T. Evaluation of gamma-carboline-phenothiazine conjugates as simultaneous NMDA receptor blockers and cholinesterase inhibitors. Bioorganic Med. Chem. 2021, 46, 116355. [Google Scholar] [CrossRef]
  161. Mahajan, S.; Choudhary, S.; Kumar, P.; Tomar, S. Antiviral strategies targeting host factors and mechanisms obliging +ssRNA viral pathogens. Bioorganic Med. Chem. 2021, 46, 116356. [Google Scholar] [CrossRef]
  162. Yudi Utomo, R.; Asawa, Y.; Okada, S.; Ban, H.S.; Yoshimori, A.; Bajorath, J.; Nakamura, H. Development of curcumin-based amyloid beta aggregation inhibitors for Alzheimer’s disease using the SAR matrix approach. Bioorganic Med. Chem. 2021, 46, 116357. [Google Scholar] [CrossRef]
  163. Liu, X.J.; Xu, L.; Pang, X.J.; Ying Yuan, X.; Yu, G.X.; Li, Y.R.; Guan, Y.F.; Zhang, Y.B.; Song, J.; Zhang, Q.R.; et al. Progress in the development of small molecular inhibitors of the Bruton’s tyrosine kinase (BTK) as a promising cancer therapy. Bioorganic Med. Chem. 2021, 47, 116358. [Google Scholar] [CrossRef] [PubMed]
  164. Hopkins, B.T.; Bame, E.; Bajrami, B.; Black, C.; Bohnert, T.; Boiselle, C.; Burdette, D.; Burns, J.C.; Delva, L.; Donaldson, D.; et al. Discovery and Preclinical Characterization of BIIB091, a Reversible, Selective BTK Inhibitor for the Treatment of Multiple Sclerosis. J. Med. Chem. 2022, 65, 1206–1224. [Google Scholar] [CrossRef] [PubMed]
  165. Tong, X.; Liu, X.; Tan, X.; Li, X.; Jiang, J.; Xiong, Z.; Xu, T.; Jiang, H.; Qiao, N.; Zheng, M. Generative Models for De Novo Drug Design. J. Med. Chem. 2021, 64, 14011–14027. [Google Scholar] [CrossRef] [PubMed]
  166. Zheng, L.; Ren, R.; Sun, X.; Zou, Y.; Shi, Y.; Di, B.; Niu, M.M. Discovery of a Dual Tubulin and Poly(ADP-Ribose) Polymerase-1 Inhibitor by Structure-Based Pharmacophore Modeling, Virtual Screening, Molecular Docking, and Biological Evaluation. J. Med. Chem. 2021, 64, 15702–15715. [Google Scholar] [CrossRef]
  167. Cui, H.; Carlson, A.S.; Schleiff, M.A.; Divakaran, A.; Johnson, J.A.; Buchholz, C.R.; Zahid, H.; Vail, N.R.; Shi, K.; Aihara, H.; et al. 4-Methyl-1,2,3-Triazoles as N-Acetyl-Lysine Mimics Afford Potent BET Bromodomain Inhibitors with Improved Selectivity. J. Med. Chem. 2021, 64, 10497–10511. [Google Scholar] [CrossRef]
  168. Luo, W.; Huang, Z.; Xu, D.; Yang, M.; Zhu, Y.; Shen, L.; Chen, S.; Tao, X.; Bin, W.; Hu, Y.; et al. Discovery and preclinical evaluations of JBD0131, a novel nitrodihydro-imidazooxazole anti-tuberculosis agent. Bioorganic Med. Chem. Lett. 2022, 72, 128871. [Google Scholar] [CrossRef]
  169. Mykhailenko, O.; Bezruk, I.; Ivanauskas, L.; Georgiyants, V. Comparative analysis of apocarotenoids and phenolic constituents of Crocus sativus stigmas from 11 countries: Ecological impact. Arch. Der Pharm. 2022, 355, e2100468. [Google Scholar] [CrossRef]
  170. Wang, R. Current perspectives on naturally occurring saponins as anticancer agents. Arch. Der Pharm. 2022, 355, e2100469. [Google Scholar] [CrossRef]
  171. Maher, M.; Zaher, A.F.; Mahmoud, Z.; Kassab, A.E. Recent green approaches for the synthesis of pyrazolo[3,4-d]pyrimidines: A mini review. Arch. Der Pharm. 2022, 355, e2100470. [Google Scholar] [CrossRef]
  172. Aggul, A.G.; Uzun, N.; Kuzu, M.; Taslimi, P.; Gulcin, I. Some phenolic natural compounds as carbonic anhydrase inhibitors: An in vitro and in silico study. Arch. Der Pharm. 2022, 355, e2100476. [Google Scholar] [CrossRef] [PubMed]
  173. Solangi, M.; Khan, K.M.; Chigurupati, S.; Saleem, F.; Qureshi, U.; Ul-Haq, Z.; Jabeen, A.; Felemban, S.G.; Zafar, F.; Perveen, S.; et al. Isatin thiazoles as antidiabetic: Synthesis, in vitro enzyme inhibitory activities, kinetics, and in silico studies. Arch. Der Pharm. 2022, 355, e2100481. [Google Scholar] [CrossRef] [PubMed]
Figure 1. General mechanism of PROTAC-mediated ubiquitination and proteasomal degradation of target protein. The heterobifunctional PROTAC is composed of a ligand that binds to the target protein and a ligand that binds to the E3 ubiquitin ligase by a linker.
Figure 1. General mechanism of PROTAC-mediated ubiquitination and proteasomal degradation of target protein. The heterobifunctional PROTAC is composed of a ligand that binds to the target protein and a ligand that binds to the E3 ubiquitin ligase by a linker.
Molecules 28 03698 g001
Figure 2. Mechanism of ARV110 in the degradation of AR protein. Initially, the ubiquitin protein binds to an E1 ubiquitin-activating enzyme and then transfers it to an E2 ubiquitin-conjugating enzyme. ARV110 binds with cereblon E3 ligase, AR, and forms a ternary complex. Finally, the cereblon E3 ligase delivers ubiquitin to the AR and induces poly-ubiquitination. The ubiquitinated AR proteins are recognized and degraded by proteasome.
Figure 2. Mechanism of ARV110 in the degradation of AR protein. Initially, the ubiquitin protein binds to an E1 ubiquitin-activating enzyme and then transfers it to an E2 ubiquitin-conjugating enzyme. ARV110 binds with cereblon E3 ligase, AR, and forms a ternary complex. Finally, the cereblon E3 ligase delivers ubiquitin to the AR and induces poly-ubiquitination. The ubiquitinated AR proteins are recognized and degraded by proteasome.
Molecules 28 03698 g002
Figure 3. Overview of genetic aberrations in different genes involved in the progression of prostate cancer. Genes and their respective chromosomal location were depicted in outer and inner circles, respectively. Each genetic variation is denoted with a single letter along with a distinct color.
Figure 3. Overview of genetic aberrations in different genes involved in the progression of prostate cancer. Genes and their respective chromosomal location were depicted in outer and inner circles, respectively. Each genetic variation is denoted with a single letter along with a distinct color.
Molecules 28 03698 g003
Figure 4. The intricate signaling pathways, including both androgen receptor (AR) and non-AR pathways, and their respective genes have been implicated in the progression of prostate cancer. The relevant genes have been highlighted in red.
Figure 4. The intricate signaling pathways, including both androgen receptor (AR) and non-AR pathways, and their respective genes have been implicated in the progression of prostate cancer. The relevant genes have been highlighted in red.
Molecules 28 03698 g004
Figure 5. The chemical structures of CRBN-based PROTACs that target the androgen receptor typically display the POI ligand in blue, the E3 ligase ligand in red, and the linker region in black.
Figure 5. The chemical structures of CRBN-based PROTACs that target the androgen receptor typically display the POI ligand in blue, the E3 ligase ligand in red, and the linker region in black.
Molecules 28 03698 g005
Figure 6. The chemical structures of VHL-based PROTACs that target the androgen receptor typically display the POI ligand in blue, the E3 ligase ligand in red, and the linker region in black.
Figure 6. The chemical structures of VHL-based PROTACs that target the androgen receptor typically display the POI ligand in blue, the E3 ligase ligand in red, and the linker region in black.
Molecules 28 03698 g006
Figure 7. Chemical structures of miscellaneous E3 ligase ligand-based PROTAC (compound 15)-targeting androgen receptor and PROTACs targeting non-AR targets (compound 16–20). The POI ligand is shown in blue, the E3 ligase ligand is in red, and the linker is represented by black.
Figure 7. Chemical structures of miscellaneous E3 ligase ligand-based PROTAC (compound 15)-targeting androgen receptor and PROTACs targeting non-AR targets (compound 16–20). The POI ligand is shown in blue, the E3 ligase ligand is in red, and the linker is represented by black.
Molecules 28 03698 g007
Table 1. PROTAC degraders in clinical trials for prostate cancer [3].
Table 1. PROTAC degraders in clinical trials for prostate cancer [3].
Name of the PROTACClinical Trial No.Date of Entry into Clinical TrialsHighest Clinical PhaseTarget ProteinE3 Ligase Used
ARV-110NCT03888612March, 2019Phase IIARCereblon
CC-94676NCT04428788June, 2020Phase IAR_
ARV-766NCT05067140October, 2021Phase IARCereblon
Table 2. Notable differences between SMIs and PROTACs.
Table 2. Notable differences between SMIs and PROTACs.
ParameterSMIsPROTACs
DruggabilityOnly limited targets are druggableExtended scope for druggable targets
SelectivityLack of selectivityHighly selective
DosageRequire high dose; thus they are less potentMinimal dose and highly protein
MOACompetitive active site inhibitionProteasomal degradation of target protein
Drug–target interactions
  • Usually they are non-covalent and require numerous interactions for good activity
  • Require active site binding
  • POI requires minimal interactions and E3 ligase requires strong interactions for good activity
  • No need for active site binding
Lipinski RO5Most of them obeys RO5Most of the PROTACs do not obey Lipinski rules
Synthetic feasibilityUsually easy to synthesizeUsually strenuous to synthesize
MOA—mechanism of action; POI—protein of interest; Ro5—rule of 5.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yedla, P.; Babalghith, A.O.; Andra, V.V.; Syed, R. PROTACs in the Management of Prostate Cancer. Molecules 2023, 28, 3698. https://doi.org/10.3390/molecules28093698

AMA Style

Yedla P, Babalghith AO, Andra VV, Syed R. PROTACs in the Management of Prostate Cancer. Molecules. 2023; 28(9):3698. https://doi.org/10.3390/molecules28093698

Chicago/Turabian Style

Yedla, Poornachandra, Ahmed O. Babalghith, Vindhya Vasini Andra, and Riyaz Syed. 2023. "PROTACs in the Management of Prostate Cancer" Molecules 28, no. 9: 3698. https://doi.org/10.3390/molecules28093698

Article Metrics

Back to TopTop