Next Article in Journal
Morphology-Dependent Interactions between α-Synuclein Monomers and Fibrils
Previous Article in Journal
Silica Nanoparticles Promote Apoptosis in Ovarian Granulosa Cells via Autophagy Dysfunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Strategies in Proteolysis-Targeting Chimeras (PROTACs): Highlights from 2022

1
College of Pharmacy, Gachon University, 191 Hambakmoe-ro, Yeonsu-gu, Incheon 21936, Republic of Korea
2
Gachon Pharmaceutical Research Institute, Gachon University, 191 Hambakmoe-ro, Yeonsu-gu, Incheon 21936, Republic of Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(6), 5190; https://doi.org/10.3390/ijms24065190
Submission received: 25 January 2023 / Revised: 27 February 2023 / Accepted: 7 March 2023 / Published: 8 March 2023
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)

Abstract

:
Targeted protein degradation (TPD) is a promising therapeutic modality that has garnered attention in academic, industrial, and pharmaceutical research for treating diseases such as cancer, neurodegenerative disorders, inflammation, and viral infections. In this context, proteolysis-targeting chimeras (PROTACs) present a reliable technology for degrading disease-causing proteins. PROTACs complement small-molecule inhibitors, which primarily rely on direct protein regulation. From concept-to-clinic, PROTACs have evolved from cell impermeable peptide molecules to orally bioavailable drugs. Despite their potential in medicinal chemistry, certain aspects regarding PROTACs remain unclear. The clinical significance of PROTACs is primarily limited owing to their lack of selectivity and drug-like properties. This review focused on recently reported PROTAC strategies, particularly in 2022. It aimed to address and overcome the challenges posed by classical PROTACs by correlating them with emerging approaches with improved selectivity and controllability, cell permeability, linker flexibility, druggability, and PROTAC-based approaches, developed in 2022. Furthermore, recently reported PROTAC-based approaches are discussed, highlighting each of their advantages and limitations. We predict that several improved PROTAC molecules will be accessible for treating patients exhibiting various conditions, including cancer, neurodegenerative disorders, inflammation, and viral infections.

1. Introduction

Classic drug discovery enables the identification of small-molecule compounds that inhibit target proteins by occupying well-defined active or allosteric sites in receptors, enzymes, or ion channels [1,2,3]. However, the lack of binding sites in various proteins renders them undruggable [4]. Approximately 80% of proteins, such as transcription factors and non-enzymatic and scaffolding proteins, are considered undruggable targets [5,6]. Alternative modalities have eventually emerged to address this issue [7]. Among these, proteolysis-targeting chimeras (PROTACs) caused a paradigm-shift in drug discovery with its ability to harness a natural degradation pathway (ubiquitin–protease system) for targeted protein degradation (TPD) [8]. From concept-to-clinic, PROTACs have evolved from being cell impermeable peptide molecules to orally bioavailable drugs. PROTACs are heterobifunctional molecules comprising ligands of the protein of interest (POI) and E3 ligase, connected by a linker. These heterobifunctional degraders bind with a POI on one end and ubiquitin E3 ligase on the other end, connected by a linker to form a ternary complex, further activating the ubiquitin–proteosome system to degrade the target protein (POI) (Figure 1).
A single PROTAC molecule can trigger multiple POI molecules and therefore act catalytically through an event-driven mechanism. PROTACs have garnered considerable attention as they target the “undruggable” [9] targets, thus overcoming drug resistance [10]. Clinical trials have been initiated for more than 10 PROTAC molecules [11]. Although remarkable progress has been made with respect to these chimeras, certain inherent shortcomings are still associated with them: (a) they are less selective and hence degrade other proteins (off-target toxicity) and exhibit non-selective degradation in an undesired tissue (on-target effect); (b) they induce uncontrollable catalytic degradation; (c) their chemical composition lies beyond the Lipinski’s rule of five, thus impeding cell permeability; (d) linker flexibility is essential; and (e) some of the small-molecule degraders exhibit difficulty in targeting hydrophobic pockets owing to the lack of well-defined binding sites [12]. However, these limitations were addressed by a prior novel PROTACs report [13]. The present review is primarily focused on collating information that addresses the above-mentioned challenges associated with classical PROTACs and discusses the strategies developed regarding the same in 2022. Furthermore, the advantages and limitations of newly emerging TPD strategies were highlighted. Specifically, the review is categorized as modalities with the following improvements:
  • Selectivity and controllability;
  • Cell permeability;
  • Linker flexibility;
  • Druggability;
  • PROTAC-based approaches.

2. Discussion

2.1. Modalities with Improved Selectivity and Controllability

Classical small-molecule PROTACs exhibit unfavorable pharmacokinetics and hence lack cell or tissue selectivity, which might cause systemic toxicity owing to nonspecific distribution in normal or undesired tissues [14]. Ligand-modification strategies such as folate–, antibody–, and aptamer–PROTAC conjugates have been developed recently for tumor-targeted delivery [15,16,17]. These ligand–PROTAC conjugates exhibited advantages such as enhanced accumulation in tumors and antitumor potency in vivo and increased cellular uptake. Despite the superior catalytic behavior exhibited by PROTACs, their off-target effect remains a challenge to be addressed. Off-target effects can be reduced by regulating PROTAC activity, thereby improving target specificity and antitumor activity [18]. This can be achieved by spatiotemporally controlling the biological system with light. In the presence of light, PROTAC activity was controlled by utilizing photoactive ligands, viz., cages and photo switches. The caging approach is based on rendering the PROTACs inactive with a photocleavable group that blocks the binding process with either the E3 ligase or POI. Uncaging with light at an appropriate wavelength resulted in the generation of active PROTACs, leading to TPD. These heterobifunctional degraders were also activated in response to internal or external stimuli and called smart activable PROTACs. Based on their activation mechanism, reduction-, enzyme-, and hypoxia-activable PROTACs have been developed. However, limited tumor penetration, low serum stability, and heterogenous expression of receptors in various tumor cells were found to be their shortcomings. In this regard, several new strategies were developed to overcome cell or tissue selectivity, tumor specificity, and controllability, as discussed below (Figure 2).

2.1.1. Aptamer-Based PROTAC

Single-stranded DNA or RNA oligonucleotides <100 nucleotides in length are called aptamers [19]. They exhibit properties such as high affinity, high specificity, low toxicity, enhanced tissue penetration, easy chemical synthesis, and excellent stability. Nucleolin is regarded as an important target in cancer therapeutics; it is a multifunctional nucleolar protein localized in the nucleoplasm, cytoplasm, and cell membrane that plays a vital role in performing a wide array of cellular activities [20]. In general, PROTACs are hydrophobic and display low cell-type selectivity. To address this, Tan et al. (2022) reported the first proof-of-concept example by constructing a novel PROTAC-utilizing nucleic acid aptamer, a 26-base guanine-rich ssDNA, as the targeting ligand of nucleolin [21]. The enhanced expression of nucleolin on the cell surface is generally restricted in tumor cells. Therefore, the aptamer was internalized into the tumor cells through nucleolin-dependent macropinocytosis, thus acting as a potential tumor-targeting element and anticancer therapeutic agent [22]. The PROTAC 1 was designed by conjugating an aptamer with a small molecule, VHL, which recruited an E3 ligase ligand via a DBCO-azide click reaction connected by PEG linker (Figure 3). The resultant aptamer-constructed PROTAC induced nucleolin degradation in breast cancer cells, both in vitro and in vivo. Furthermore, this approach demonstrated selective degradation of tumor cell proteins, good water solubility, and excellent serum stability. This is a promising strategy, with its only limitation being the shortage of aptamers that specifically bind to E3 ligases. In contrast to the recently reported aptamer conjugate BET–PROTAC, a targeted delivery tool for specific cell types [23], this novel PROTAC compound promoted the generation of nucleolin–PROTAC–VHL ternary complex by using a nucleolin-binding aptamer as a ligand.
Aptamers are recognized artificially via an iterative in vitro selection strategy called the systematic evolution of ligands by exponential enrichment (SELEX) [24,25,26]. Modification of microenvironmental pH, complementary stands, or metabolic molecules leads to switching of the unique structure of aptamers, allowing for the selective modulation of aptamers and POI interactions [27,28,29]. Recently, Xie et al. (2023) developed apt-PROTAC 2 in which the aptamer directed the oncogenic nucleolin to E3 ligase for degradation [30]. This PROTAC 2 was synthesized via the click reaction between thalidomide-O-amido-propargyl, a cereblon ligand (CRBNL), and azide-modified AS1411 aptamer. A series of thymidines or PEG groups were employed as the linkers. Furthermore, a light-inducible nucleolin degrader, termed opto-PROTAC, was developed by hybridizing the apt-PROTAC with the complimentary photolabile oligonucleotide (CP). Upon irradiation with UV (365 nm), light-triggered breakage of photolabile bonds resulted in the liberation of the active PROTAC degrader (Figure 4). The major advantages of this strategy include (a) apt-PROTACs act as an alternative approach for undruggable TPD owing to their ability to recognize all types of proteins, specifically mutants; (b) it is a convenient mode of synthesis by utilizing nucleotides as linkers; (c) it provides convenient spatiotemporal regulation of the activity of apt-PROTACs by adopting conformational switchable aptamers; and (d) the selective accumulation of apt-PROTACs in cancer cells resulting in reduced on-target toxicity in normal cells.

2.1.2. Pre-PROTACs

To improve therapeutic efficacy and reduce systemic toxicity, drugs are designed to be activated by specific triggers in a predetermined pathway. Reactive oxygen species (ROS), such as hydrogen peroxide, superoxide, and hydroxyl radicals, play an important role in regulating signaling pathways [31]. Low ROS levels are maintained in normal cells, while it is much higher in the tumor microenvironment [32]. Chen et al. (2022) designed ROS-responsive Pre-PROTACs 3 by conjugating an ROS-triggered leaving group, arylboronic acid, to the parent PROTACs [33]. This PROTAC molecule 3 was designed by recruiting a CRBN ligand and BRD3 protein ligand to opposite ends connected by varying linkers (Figure 5). These PROTACs showed excellent TPD of BRD3. Its major limitations are as follows: (a) the occasional generation of unwanted by-products, although the caging strategy is efficient in TPD and (b) the inhibitory activity of Pre-PROTAC is sensitive to the cell line with high ROS levels. Notably, this is the first report of ROS-activated Pre-PROTAC, which was proved to have potential in controlling PROTAC activation through an endogenous ROS-related microenvironment.

2.1.3. PS-Degrons

He et al. (2022) proposed PS-Degrons involving the activation of a TPD platform by adopting a light-triggered pathway [34]. The PS-Degrons 4 were designed by conjugating photosensitizers and target protein ligands (Figure 6). This methodology resulted in the controlled knockdown of the target protein, human estrogen receptor α, with high spatiotemporal precision via light-triggered 1O2. Furthermore, PS-Degrons enable elevated anti-proliferation in MCF7 cells. This modality complements existing technologies by overcoming systemic toxicity.

2.1.4. AP-PROTACs

Despite the rapid progress achieved by PROTACs, the spatiotemporal control of degradation still remained elusive. Methods to attain conditional degradation involved selective targeting or light-driven photo-pharmacology [35]. Traditional examples of light-driven photo-pharmacology include caged and photo-switchable PROTACs, with irreversible and reversible modes of action, respectively. PHOtochemically Targeting Chimeras (PHOTACs) are a small group of photo-switchable PROTACs having an azobenzene photo-switch with an active E (trans) or Z (cis) isomer that induces TPD. Tate et al. (2022) reported a novel arylazopyrazole photo-switch (AP-PROTACs), 5 and 6 the first example of a non-azobenzene PHOTAC technology [36]. Using this, two types of AP-PROTACs were synthesized: (a) AP-PROTACs 5 with pan-bromodomain inhibitor JQ1 as the POI ligand and thalidomide as a CRBN recruiter and (b) AP-PROTACs 6 with a multi-kinase inhibitor as the PROTAC warhead, with lenalidomide as a CRBN-recruiting ligand connected by arylazopyrazole (AAP) linkers (Figure 7). These AP-PROTACs 5 and 6 have the following advantages: (a) AAP linkers acting as a potential “plug and play” photo-switch to control biological outcomes; (b) good PSS isomer abundance; (c) long z-isomer half-life; and (d) stable reversible switching for multiple cycles. Notably, to the best of our knowledge, this is the first example of a multi-target PROTAC that adopts a photo-switchable degradation pathway.

2.1.5. RT-PROTACs

To overcome the limitation of systemic toxicity caused by undesirable off-TPD posed by classical PROTACs, Li et al. (2023) developed a novel radiotherapy-triggered PROTAC 7 (RT-PROTAC) prodrug strategy for spatiotemporally-controlled TPD through X-ray radiation [37]. RT-PROTAC 7 attaches a bromodomain (BRD)-targeting PROTAC to an X-ray inducible phenyl azide cage (Figure 8). This prodrug has been activated by X-ray radiation, both in vitro and in vivo. Thus, activated PROTACs 8 target BRD4 and BRD2 with remarkable antitumor potency in the MCF-7 xenograft model.

2.1.6. NTR-Based PROTACs

The most common characteristic of tumors is their state of low oxygen levels or hypoxia, resulting in uncontrollable metastasis, therapeutic resistance, and poor prognosis [38]. Hypoxia is generally accompanied by high levels of nitroreductase (NTR). The catalytic behavior of PROTACs occasionally results in off-tissue target degradation, leading to potential toxicity. Thus, Zhu et al. (2022) developed the first NTR-responsive PROTAC 9 comprising a VHL E3 ligase ligand (caged by an NTR activating group) and an epidermal growth factor target protein ligand connected by a polyethylene glycol (PEG) linker (Figure 9) [39]. This strategy allowed precise control of the PROTAC 10 for selective TPD with good antitumor efficacy.

2.1.7. ENCTACs

Hypoxia also affects gene expression and influences epigenetic modifications [38]. Xing et al. (2022) developed enzyme-derived clicking PROTACs 11 (ENCTACs) for stimulating responsive protein degradation in the tumor microenvironment [40]. ENCTACs depend on nitroreductase (NTR) enzyme for selectively forming degraders at the hypoxic site and result in targeted epigenetic BRD4 degradation. The design of NTR-oriented ENCTACs 11 comprises orthogonal cross-linking between the uncaged cysteine residue pre-modified with E3 ligase cereblon (CRBN) recruiting ligand, pomalidomide, and JQ1 (2-cyanobenzothiazole (CBT)-preconjugated BRD4 ligand), connected by a PEG2-NH2 linker (Figure 10). ENCTACs are formed locally within tumors only through the hypoxia-induced activation of the NTR enzyme. Its advantages include (a) precise modulation of hypoxia signaling in living mice, zebrafish, and live cells with solid tumors and (b) controlled cross-linking without prior covalent conjugation, maximizing pharmacological performance due to the large molecular weight. However, limitations of this approach include a lack of selectivity between normoxic and hypoxic cells, possibly resulting in detrimental off-target effects and the potential effect of ENCTACs on the stability of other proteins in the cells.

2.1.8. POLY-PROTACs

Yu et al. (2022) reported a novel polymeric PROTAC 13 (POLY-PROTAC) nanoplatform for the tumor-targeted degradation of the bromodomain and extra-terminal (BET) protein, BRD4 [41]. POLY-PROTAC 13 consists of a VHL ligand and BRD4 protein ligand connected by optimizable linkers (Figure 11). Its advantages include (a) activation through multiple stimuli (extracellular matrix metalloproteinase-2, intracellular acidity, and reductive conditions), resulting in improved tumor accumulation, enhanced protein degradation, and deep tumor penetration; (b) tumor-specific PROTAC delivery being achieved owing to an extracellular tumor acidity-triggered biorthogonal click reaction; and (c) its combination with photodynamic therapy (PDT), suppressing more than 95% of MDA-MB-231 tumor growth in triple-negative breast cancer.

2.1.9. Smart Nano-PROTACs

An interesting modality for cancer-specific protein degradation has been reported by Kanyi et al. (2021) [42]. The smart nano-PROTACs 15 comprise a semiconducting polymer backbone linked to cyclooxygenase1/2 (COX-1/2)-targeting PROTAC peptide (CPP) via a cathepsin B (CatB)-cleavable segment (Figure 12). Through the ubiquitin–proteasome system, COX-1/2 is degraded by CPP activation catalyzed by the tumor-overexpressed CatB. The advantages of this SPNCOX-mediated technology include enhancement of tumor immunogenicity, intra-tumoral degradation of COX-1/2, and reprogramming of the immunosuppressive tumor microenvironment, thus inhibiting tumor growth, metastasis, and recurrence.

2.2. Modalities with Improved Linker Flexibility

Split-PROTACs

An inherent challenge associated with the activity of traditional PROTACs is the high dependency on the length and nature of the linker connecting the target protein and E3 ligase ligands. To address this, Kodadek et al. (2022) created “split PROTACs” 16 with optimizable linker flexibility (Figure 13) [43]. This involved combining the modified E3 ubiquitin ligand and POI ligand with residues suitable for coupling when mixed together in a combinatorial fashion. The aldehyde- and alkoxyamine-functionalized fragments were examined using oxime ligation methodology. This strategy eased the task of synthesizing several PROTACs. In addition, the PROTAC molecule assembled efficiently upon premixing high concentrations of the components and then adding them to the cells. The only limitation of this protocol was that at relatively low concentrations, the POI ligand and E3 ligase ligand displayed inefficient coupling in situ.

2.3. Modalities with Improved Cell Permeability

The molecular weight of PROTACs is considerably higher than that of traditional small-molecule inhibitors, potentially resulting in pharmacokinetic hurdles to their cell permeability. An attempt to address this issue was made by developing in-cell click-formed compounds called CLIPTACs [44]. This novel strategy enabled the intracellular assembly of PROTAC molecules from small fragments via a click reaction; the resulting compounds exhibited enhanced cell permeability. However, the success of CLIPTACs is limited by their low reaction yields, resulting in leftover unreacted fragments that interfered with the formation of ternary complex, a key step in protein degradation.

2.3.1. Hydrophobic Tags

Neurodegenerative disorders are localized to the central nervous system (CNS). Despite the potential chemical knockdown of neurodegenerative disorder-related aggregation-prone proteins by PROTACs, these degraders have poor CNS permeability. Ishikawa et al. (2022) documented the conversion of PROTAC degraders into hydrophobic tags 17 (HyTs) [45], which are heterobifunctional molecules composed of POI ligand linked to a hydrophobic degron, viz., diamantane, that mimics the exposed hydrophobic region of misfolded proteins and is recognized by heat shock protein 70 kDa (HSP70) or C-terminal of HSC70-interacting protein (CHIP) [46]. Hydrophobic tagging technology utilized HSP70, the protein quality control machinery for TPD, via the proteasomal pathway (Figure 14). Consequently, the developed HyTs displayed potentially decreased levels of mutant huntingtin, an aggregation-prone protein. Moreover, the blood–brain barrier penetration potential and brain-to-plasma concentration values were calculated. The findings successfully detected a brain-permeable HyT using immobilized artificial membrane (IAM) chromatography analysis. The major advantages of this conversion were as follows: (a) more potent than the parent PROTAC, (b) a decrease in the molecular weight and number of hydrogen bond donors and acceptors, and (c) improved drug-like properties. Furthermore, this was the first example of a small-molecule hydrophobic tag targeting an aggregation-prone protein.

2.3.2. Stapled PROTACs

In contrast to traditional PROTACs, the recently developed linear peptide-based PROTACs are considered undruggable because of their proteolytic stability and poor cellular uptake. To address this limitation, Hu et al. (2022) developed stapled (SP-)PROTACs 18 by combining peptide stapling with PROTAC [47]. The SP-PROTAC 18 structure comprises an MDM2/MDMX-targeting moiety, viz., PMI-N8A peptide, PEG linker, and VHL E3 ligase recruiting hydroxyproline (Hyp)- and homoleucine (Hle)-containing hexapeptide (LA-Hyp-Y-Hle-P) (Figure 15). The synthesized degraders showed improved proteolytic stability, better cellular permeability, good pharmacokinetic profile, and similar binding to MDM2 and MDMX. Furthermore, SP-PROTACs inhibited tumor progression in orthotopic colorectal cancer xenograft models by the simultaneous atypical degradation of both MDM2 and MDMX and p53 activation. The main advantages of SP-PROTACs are (a) target protein specificity, (b) high binding potency with both the target protein and E3 ligase, (c) ease of synthesis, (d) potent in vitro and in vivo activities, and (e) efficient death of cancer cells. There are certain limitations associated with SP-PROTACs. First, the hydrocarbon-stapled peptide has poor water solubility. Second, they cause deficient TPD owing to the cleavage of peptide-based E3 ligase-recruiting ligands in the cellular environment.

2.4. Modalities with Improved Druggability

In drug discovery and development, small-molecule inhibitors primarily target either the active site or an allosteric site of an enzyme to inhibit the function of the POI (occupancy-driven mode) [48]. In contrast, PROTACs are event-driven therapeutics, wherein one degrader molecule efficiently triggers the degradation of many POI molecules. PROTACs have been widely investigated for degrading previously undruggable targets. The characteristics of such undruggable targets include the lack of well-defined binding sites, non-catalytic protein–protein interaction functional modes, and ambiguous 3D structures. PROTACs negate the need for an active site owing to their event-driven pharmacology. Recent emerging approaches for targeting undruggable targets include covalent drugs, protein-based drugs, peptide-based drugs, bifunctional molecules, and RNA therapeutics.

2.4.1. Bridged PROTACs

Classical PROTACs were successful in targeting druggable and difficult-to-drug proteins. Recently, degraders such as transcription factor TF-PROTACs [49] and RNA-PROTACs [50] have also been developed to target undruggable TFs and RNA-binding proteins, respectively, with small-molecule binding sites. Jin et al. (2022) introduced a novel strategy, the bridged PROTAC 19, which specifically targets undruggable proteins by utilizing a small-molecule binder from the binding fragment of the POI [51]. This novel compound 19 successfully degraded the undruggable protein cyclin D1 via a unique binding mode. The bridged PROTAC 19 is composed of VHL E3 ligase, CDK4/6 binder connected to VHL E3 ligase by a linker, CDK4/6 as the bridged protein, and cyclin D1 as the target protein (POI) (Figure 16). Compared with CDK4/6 degraders and inhibitors, this novel platform showed remarkable efficiency in suppressing cancer cell proliferation. Moreover, bridged PROTACs formed quaternary complexes, such as cyclin D1-CDK6-PROTAC-VHL, and provided a proof-of-concept for degrading undruggable proteins lacking small-molecule binders.

2.4.2. TF-DUBTACs

The targeted degradation of tumor suppressor proteins remains less explored. To address this, Wei et al. (2022) demonstrated a novel TF-DUBTAC 20 platform [52]. It comprises a DNA oligonucleotide linked to the covalent ligand of deubiquitinase OTUB1 (Figure 17). A series of TF-DUBTACs 20, including FOXO-DUBTAC, p53-DUBTAC, and IRF-DUBTAC, have been developed via a copper-free strain-promoted azide-alkyne cycloaddition (SPAAC) reaction. These compounds stabilized cellular FOXO3A, p53, and IRF3 in an OTUB1-dependent manner. TF-DUBTACs suppressed tumorigenesis via the selective stabilization of tumor suppressor transcription factors.

2.4.3. Nanobody-Based Degrader

Many undruggable proteins have surfaces to which other large biomolecules can bind and act as ligands for TPD. Antibodies display exceptional specificity and significant affinity for their antigens by exploiting such protein surfaces. Nanobodies, fragments of single variable heavy chain domains, retain antigen specificity and have been employed as POI ligands in PROTAC synthesis. In this context, the yeast surface display platform, a cost-effective method, was developed to generate nanobody ligands for proteins [53]. Hemoglobin disorders are treated by the reactivation of fetal hemoglobin (HbF, α2γ2) [54]. The low survival rates of patients with sickle cell disease are because of the high-level production of HbF [55]. BCL11A subjugates HbF expression, thereby switching from fetal to adult hemoglobin expression during erythropoiesis [56]. Dassama et al. (2022) developed a cell-permeant nanobody fused to a miniature protein and an E3 adaptor to form a degrader 21 (Figure 18) [57]. This nanobody degrader 21 efficiently depleted cellular BCL11A in differentiated primary erythroid precursor cells. Furthermore, it significantly induced fetal hemoglobin expression. This strategy has emerged to compliment traditionally “undruggable” targeted degradation.

2.5. PROTAC-Based Approaches

2.5.1. BacPROTACs

ClpC: ClpP (ClpCP) protease is a key quasi-proteasomal particle crucial for microbial protein quality control, stress tolerance, and pathogenicity. ClpCP was also observed in Gram-positive bacteria and mycobacteria. Similar to eukaryotes, only a few bacterial species undergo TPD. PROTAC technology is primarily focused on the ubiquitin tagging of eukaryotes and needs to be extended to bacterial degradation. Phosphorylated arginine residues (pArg) function as a degradation signal, which is recognized by ClpCP in the form of a simple phosphate group attached to the arginine residue of the target protein [58]. Clausen et al. (2022) developed a novel small-molecule degrader, the BacPROTAC 22. These pArg-containing BacPROTACs 22 recruit POI to the ClpC domain, resulting in their degradation by the ClpCP protease (Figure 19) [59]. These small-molecule adapters redirect ClpCP protease for specifically targeting neo-substrates. Moreover, degradation and drug susceptibility assays using this method were performed in mycobacteria, revealing in vivo activity by selectively targeting endogenous proteins through fusion with an established degron 23.
The significant benefits of BacPROTACs include (a) potent in vivo activity, (b) selective and efficient degradation of POI in bacteria, (c) cell permeability, (d) wide applicability to bacterial proteins owing to the efficacy of reprogramming ClpCP protease, and (e) the feasibility of repurposing various protein ligands. BacPROTACs have two major limitations. First, the in vivo degradation efficacy is lower than that in vitro. Second, when deregulating the ClpC1P1P2 protease, cyclomarin (Cym)-based degraders might interfere with the cellular levels of unrelated proteins.

2.5.2. DUB-Based Degrader

Deubiquitinating enzymes (DUBs) have emerged as potential targets for treating various human cancers and other diseases. Steinebach et al. (2022) reported the first DUB degrader 24 and revealed a new class of drug targets for TPD [60]. DUB-degraders are PROTACs targeting a ubiquitin-specific protease 7. The design of the degrader adopted glutarimide-based CRBN binder and USP7 inhibitors, which were connected by variable linkers (Figure 20). These small-molecule degraders 24 efficiently induce apoptosis in USP7-dependent cancer cells.

2.5.3. HEMTACs

Heat shock protein 90 (HSP90), an ATP-dependent molecular chaperone, is widely expressed in all eukaryotic cells and displays tumor-selective pharmacokinetics [61,62,63]. HSP90 plays an important role in regulating cytoprotection, cell cycle progression, and cell signaling via its increased production against stress response. Traditional E3 ligase VHL- and CRBN-based small-molecule PROTAC degraders cause off-target effects and resistance mechanism in cells [64,65,66]. To address this, Li et al. (2023) developed heat shock protein 90-mediated targeting chimeras 25 (HEMTACs) for intracellular protein degradation [67]. HSP90 binds to the target protein and facilitates the downregulation of proteins. HEMTACs 25 comprise a POI binding ligand, linker, and HSP90-binding ligand (Figure 21). Cyclin-dependent kinases regulate eukaryotic cell division, proliferation, and gene expression [68]. CDK4/6 is usually overexpressed and hyperactivated in cancer cells and plays a key role in initiating the cell cycle [69,70,71]. By adopting HEMTACs, CDK4/6 was successfully degraded, thus expanding the potential for TPD.

3. Conclusions

Drug discovery and development have witnessed rapidly successful protein-degradation technologies in the form of PROTACs and several newly emergent modalities. This is evident from the initiation of a clinical trial for >10 PROTAC molecules. The journey of PROTACs from concept-to-clinic took approximately two decades. Despite the remarkable progress achieved, certain inherent limitations remain to be addressed. These involve issues associated with selectivity, permeability, role of linkers, and druggability. Several new modalities have been developed to address these challenges. However, these emerging modalities are still immature and require further elucidation for obtaining successful PROTAC-based drug candidates. We hypothesize that several PROTAC molecules will be accessible to patients exhibiting various conditions, including cancer, neurodegenerative disorders, inflammation, and viral infections, in the future.

Author Contributions

Conceptualization, R.T. and D.S.; methodology, R.T. and D.S.; writing—original draft preparation, R.T.; writing—review and editing, R.T. and D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of Korea (NRF-2020R1A6A1A3043708 and NRF-2021R1A2C1012280).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that there are no conflicts of interest.

References

  1. Blanco, M.J.; Gardinier, K.M. New Chemical Modalities and Strategic Thinking in Early Drug Discovery. ACS Med. Chem. Lett. 2020, 11, 228–231. [Google Scholar] [CrossRef]
  2. Cohen, P.; Cross, D.; Janne, P.A. Kinase drug discovery 20 years after imatinib: Progress and future directions. Nat. Rev. Drug Discov. 2021, 20, 551–569. [Google Scholar] [CrossRef] [PubMed]
  3. Burslem, G.M.; Crews, C.M. Small-Molecule Modulation of Protein Homeostasis. Chem. Rev. 2017, 117, 11269–11301. [Google Scholar] [CrossRef]
  4. Dang, C.V.; Reddy, E.P.; Shokat, K.M.; Soucek, L. Drugging the ‘undruggable’ cancer targets. Nature Reviews Cancer 2017, 17, 502–508. [Google Scholar] [CrossRef] [PubMed]
  5. Overington, J.P.; Al-Lazikani, B.; Hopkins, A.L. How Many Drug Targets are There? Nat. Rev. Drug Discov. 2006, 5, 993–996. [Google Scholar] [CrossRef]
  6. Crews, C.M. Targeting the Undruggable Proteome: The Small Molecules of my Dreams. Chem. Biol. 2010, 17, 551–555. [Google Scholar] [CrossRef] [Green Version]
  7. Gurevich, E.; Gurevich, V. Beyond Traditional Pharmacology: New Tools and Approaches. Br. J. Pharmacol. 2015, 172, 3229–3241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Glickman, M.H.; Ciechanover, A. The Ubiquitin-Proteasome Proteolytic Pathway: Destruction for the Sake of Construction. Physiol. Rev. 2002, 82, 373–428. [Google Scholar] [CrossRef]
  9. Samarasinghe, K.T.G.; Crews, C.M. Targeted protein degradation: A promise for undruggable proteins. Cell Chem. Biol. 2021, 28, 934–951. [Google Scholar] [CrossRef]
  10. Sun, X.; Rao, Y. PROTACs as Potential Therapeutic Agents for Cancer Drug Resistance. Biochemistry 2020, 59, 240–249. [Google Scholar] [CrossRef]
  11. Bekes, 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]
  12. Garber, K. The PROTAC GoldRush. Nat. Biotechnol. 2022, 40, 12–16. [Google Scholar] [CrossRef]
  13. Zhong, Y.; Chi, F.; Wu, H.; Liu, Y.; Xie, Z.; Huang, W.; Shi, W.; Qian, H. Emerging Targeted Protein Degradation Tools For Innovative Drug Discovery: From Classical Protacs To The Novel And Beyond. Eur. J. Med. Chem. 2022, 231, 114142–114165. [Google Scholar] [CrossRef] [PubMed]
  14. Kevin, M.; Muireann, C.; Andrew, X.Z.; Fiona, P.; Castaldi, M.P.; Goran, D.; Helen, B.; Clay, S.; Pete, N. Proteolysis-targeting chimeras in drug development: A safety perspective. Br. J. Pharm. 2020, 177, 1709–1718. [Google Scholar]
  15. Cotton, A.D.; Nguyen, D.P.; Gramespacher, J.A.; Seiple IBWells, 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]
  16. Thomas, H.P.; Pragya, A.; Robert, A.B.; Jinhua, C.; Geoffrey, D.R.; Gauri, D.; Isabel, F.; Karen, E.G.; Amrita, V.K.; Susan, K.; et al. Antibody conjugation of a chimeric BET degrader enables in vivo activity. ChemMedChem 2020, 15, 17–25. [Google Scholar]
  17. Jing, L.; He, C.; Yi, L.; Yudao, S.; Fanye, M.; Kaniskan, H.U.; Jian, J.; Wenyi, W. Cancer selective target degradation by folate-caged PROTACs. J. Am. Chem. Soc. 2021, 143, 7380–7387. [Google Scholar]
  18. Nalawansha, D.A.; Crews, C.M. PROTACs: An Emerging Therapeutic Modality in Precision Medicine. Cell Chem. Biol. 2020, 27, 998–1014. [Google Scholar] [CrossRef]
  19. Nimjee, S.M.; White, R.R.; Becker, R.C.; Sullenger, B.A. Aptamers as therapeutics. Annu. Rev. Pharmacol. Toxicol. 2017, 57, 61–79. [Google Scholar] [CrossRef]
  20. Koutsioumpa, M.; Papadimitriou, E. Cell Surface Nucleolin as a target for Anti-cancer Therapies. Recent Pat. Anti-Cancer Drug Discov. 2014, 9, 137–152. [Google Scholar] [CrossRef] [PubMed]
  21. Lin, Z.; Ling, L.; Xia, W.; Huimin, L.; Yibin, Z.; Tiantian, X.; Hui, Z.; Xiaodong, L.; Tianhuan, P.; Xing, S.; et al. Development of a novel PROTAC using the nucleic acid aptamer as a targeting ligand for tumor selective degradation of nucleolin. Mol. Ther. Nucleic Acids 2022, 30, 66–79. [Google Scholar]
  22. Reyes-Reyes, E.M.; Teng, Y.; Bates, P.J. A new paradigm for aptamer therapeutic AS1411 action: Uptake by macropinocytosis and its stimulation by a nucleolin-dependent mechanism. Cancer Res. 2010, 70, 8617–8629. [Google Scholar] [CrossRef] [Green Version]
  23. He, S.; Gao, F.; Ma, J.; Ma, H.; Dong, G.; Sheng, C. Aptamer-PROTAC conjugates (APCs) for tumor-specific targeting in breast cancer. Angew Chem. Int. Ed. Engl. 2021, 60, 23299–23305. [Google Scholar] [CrossRef]
  24. Ellington, A.D.; Szostak, J.W. In vitro selection of RNA molecules that bind specific ligands. Nature 1990, 346, 818–822. [Google Scholar] [CrossRef] [PubMed]
  25. Irvine, D.; Tuerk, C.; Gold, L. SELEXION. Systematic evolution of ligands by exponential enrichment with integrated optimization by non-linear analysis. J. Mol. Biol. 1991, 222, 739–761. [Google Scholar] [CrossRef] [PubMed]
  26. Taylor, A.I.; Arangundy-Franklin, S.; Holliger, P. Towards applications of synthetic genetic polymers in diagnosis and therapy. Curr. Opin. Chem. Biol. 2014, 22, 79–84. [Google Scholar] [CrossRef]
  27. Li, L.; Jiang, Y.; Cui, C.; Yang, Y.; Zhang, P.; Stewart, K.; Pan, X.; Li, X.; Yang, L.; Qiu, L.; et al. Modulating aptamer specificity with pH-responsive DNA bonds. J. Am. Chem. Soc. 2018, 140, 13335–13339. [Google Scholar] [CrossRef] [PubMed]
  28. Biniuri, Y.; Luo, G.F.; Fadeev, M.; Wulf, V.; Willner, I. Redox-switchable binding properties of the ATP-aptamer. J. Am. Chem. Soc. 2019, 141, 15567–15576. [Google Scholar] [CrossRef]
  29. Zhao, D.; Yang, G.; Liu, Q.; Liu, W.; Weng, Y.; Zhao, Y.; Qu, F.; Li, L.; Huang, Y. A photo-triggerable aptamer nanoswitch for spatiotemporal controllable siRNA delivery. Nanoscale 2020, 12, 10939–10943. [Google Scholar] [CrossRef]
  30. Miao, C.; Ping, Z.; Yun, K.; Jingrui, L.; Yan, L.; Yao, Z.; Jie, R.; Jun, Z.; Yan, C.; Songbo, X. Inducible Degradation of Oncogenic Nucleolin Using an Aptamer-Based PROTAC. J. Med. Chem. 2023, 66, 1339–1348. [Google Scholar]
  31. Cheung, E.C.; Vousden, K.H. The role of ROS in tumor development and progression. Nat. Rev. Cancer 2022, 22, 280–297. [Google Scholar] [CrossRef]
  32. Winterbourn, C.C. Hydrogen peroxide reactivity and specificity in thiol-based cell signalling. Biochem. Soc. Trans. 2020, 48, 745–754. [Google Scholar] [CrossRef]
  33. Haixia, L.; Chaowei, R.; Renhong, S.; Huihui, W.; Yuexiong, Z.; Xiaobao, Y.; Biao, J.; Hongli, C. Reactive oxygen species-responsive Pre-PROTAC for tumor-specific protein degradation. Chem. Commun. 2022, 58, 10072–10075. [Google Scholar]
  34. Silong, Z.; Yuanyuan, L.; Tao, L.; Yu, Z.; Haimei, L.; Zhengzai, C.; Na, P.; Yi, L.; Juan, X.; Huan, H. Activable Targeted Protein Degradation Platform Based on Light-triggered Singlet Oxygen. J. Med. Chem. 2022, 65, 3632–3643. [Google Scholar]
  35. Reynders, M.T.D. Optical control of targeted protein degradation. Cell Chem. Biol. 2021, 28, 969–986. [Google Scholar] [CrossRef] [PubMed]
  36. Qisi, Z.; Cyrille, S.K.; Milon, M.; Jake, L.G.; Jennifer, R.B.; Sergei, K.; Mikhail, I.; Christopher, P.T.; Leran, Z.; Daniel, C.; et al. Light-mediated multi-target protein degradation using arylazopyrazolephotoswitchable PROTACs (AP-PROTACs). Chem. Commun. 2022, 58, 10933–10936. [Google Scholar]
  37. Chunrong, Y.; Yuchen, Y.; Yujie, L.; Qiankun, N.; Jinghong, L. Radiotherapy-Triggered Proteolysis Targeting Chimera Prodrug Activation in Tumors. J. Am. Chem. Soc. 2023, 145, 385–391. [Google Scholar]
  38. Singleton, D.C.; Macann, A.; Wilson, W.R. Therapeutic targeting of the hypoxic tumor microenvironment. Nat. Rev. Clin. Oncol. 2021, 18, 751–772. [Google Scholar] [CrossRef]
  39. Shi, S.; Yu, D.; Yi, Z.; Jing, N.; Zeyu, C.; Xiaonan, W.; Feihuang, Q.; Yi, D.; Gengchen, Y.; Yunze, W.; et al. Rational Design for Nitroreductase (NTR)-Responsive Proteolysis Targeting Chimeras (PROTACs) Selectively Targeting Tumor Tissues. J. Med. Chem. 2022, 65, 5057–5071. [Google Scholar] [CrossRef]
  40. Thang, C.D.; Jun, W.L.; Caixia, S.; Songhan, L.; Khoa, T.K.; Seok, T.L.; Yu, Y.O.; Yuet, P.K.; Jia, J.M.; Yuguang, M.; et al. Hypoxia deactivates epigenetic feedbacks via enzyme derived clicking proteolysis-targeting chimeras. Sci. Adv. 2022, 8, eabq2216. [Google Scholar]
  41. Jing, G.; Bo, H.; Qiwen, Z.; Lei, Y.; Xingyu, J.; Zhifeng, Z.; Xutong, L.; Tianfeng, X.; Mingyue, Z.; Yi-Hung, C.; et al. Engineered bioorthogonal POLY-PROTAC nanoparticles for tumor-specific protein degradation and precise cancer therapy. Nat. Comm. 2022, 13, 4318. [Google Scholar]
  42. Chi, Z.; Shasha, H.; Ziling, Z.; Penghui, C.; Kanyi, P. Smart Nano-PROTACs Reprogram Tumor Microenvironment for ActivatablePhoto-metabolic Cancer Immunotherapy. Angew. Chemie. 2021, 61, e202114957. [Google Scholar]
  43. Weijun, G.; Thomas, K. Applications and Limitations of Oxime-Linked “Split PROTACs”. ChemBioChem. 2022, 23, e202200275. [Google Scholar]
  44. Lebraud, H.; Wright, D.J.; Johnson, C.N.; Heightman, T.D. Protein Degradation by In-Cell Self-Assembly of Proteolysis Targeting Chimeras. ACS Cent. Sci. 2016, 2, 927–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Keigo, H.; Hiroko, Y.; Shusuke, T.; Yugo, M.; Tatsuya, N.; Kenji, O.; Mayumi, I.; Kayoko, K.; Yui, I.; Shinsaku, N.; et al. Conversion of a PROTAC Mutant Huntingtin Degrader into Small Molecule Hydrophobic Tags Focusing on Drug-like Properties. ACS Med. Chem. Lett. 2022, 13, 396–402. [Google Scholar]
  46. 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. Int. Ed. 2015, 54, 9659–9662. [Google Scholar] [CrossRef]
  47. Si, C.; Xiang, L.; Yinghua, L.; Xing, Y.; Chenchen, G.; Songyan, G.; Jinyang, L.; Bohan, M.; Zhe, W.; Wuyuan, L.; et al. Design of stapled peptide-based PROTACs for MDM2/MDMX atypical degradation and tumor Suppression. Theranostics 2022, 12, 6665–6681. [Google Scholar]
  48. Luh, L.M.; Scheib, U.; Juenemann, K.; Wortmann, L.; Brands, M.; Cromm, P.M. Prey for the Proteasome: Targeted Protein Degradation-A Medicinal Chemist’s Perspective. Angew. Chem. Int. Ed. Engl. 2020, 59, 15448–15466. [Google Scholar] [CrossRef]
  49. Liu, J.; Chen, H.; Kaniskan, H.Ü.; Xie, L.; Chen, X.; Jin, J.; Wenyi, W. TFPROTACs Enable Targeted Degradation of Transcription Factors. J. Am. Chem. Soc. 2021, 143, 8902–8910. [Google Scholar] [CrossRef]
  50. Ghidini, A.; Cléry, A.; Halloy, F.; Allain, F.H.T.; Hall, J. RNAPROTACs: Degraders of RNA-Binding Proteins. Angew. Chem. Int. Ed. 2021, 60, 3163–3169. [Google Scholar] [CrossRef] [PubMed]
  51. Yan, X.; Yue, Z.; Hyerin, Y.; Xiaobao, Y.; Kwang-Su, P.; Ling, X.; Poulikos, I.P.; Xiaoran, H.; Yue, X.; Xian, C.; et al. Bridged Proteolysis Targeting Chimera (PROTAC) Enables Degradation of Undruggable Targets. J. Am. Chem. Soc. 2022, 144, 22622–22632. [Google Scholar]
  52. Jing, L.; Xufen, Y.; He, C.; Kaniskan, H.U.; Ling, X.; Xian, C.; Jian, J.; Wenyi, W. TF-DUBTACs Stabilize Tumor Suppressor Transcription Factors. J. Am. Chem. Soc. 2022, 144, 12934–12941. [Google Scholar]
  53. McMahon, C.; Baier, A.S.; Pascolutti, R.; Wegrecki, M.; Zheng, S.; Ong, J.X.; Erlandson, S.C.; Hilger, D.; Rasmussen, S.G.F.; Ring, A.M.; et al. Yeast surface display platform for rapid discovery of conformationally selective nanobodies. Nat. Struct. Mol. Biol. 2018, 25, 289–296. [Google Scholar] [CrossRef] [Green Version]
  54. Orkin, S.H. Molecular Medicine: Found in Translation. Med 2021, 2, 122–136. [Google Scholar] [CrossRef] [PubMed]
  55. Platt, O.S.; Brambilla, D.J.; Rosse, W.F.; Milner, P.F.; Castro, O.; Steinberg, M.H.; Klug, P.P. Mortality in Sickle Cell Disease—Life Expectancy and Risk Factors for Early Death. New Engl. J. Med. 1994, 330, 1639–1644. [Google Scholar] [CrossRef]
  56. Liu, N.; Hargreaves, V.V.; Zhu, Q.; Kurland, J.V.; Hong, J.; Kim, W.; Sher, F.; Macias-Trevino, C.; Rogers, J.M.; Kurita, R.; et al. Direct Promoter Repression by BCL11A Controls the Fetal to Adult Hemoglobin Switch. Cell 2018, 173, 430–442. [Google Scholar] [CrossRef] [Green Version]
  57. Shen, F.; Zheng, G.; Setegne, M.; Tenglin, K.; Izadi, M.; Xie, H.; Zhai, L.; Orkin, S.H.; Dassama, L.M.K. A Cell-Permeant Nanobody-Based Degrader That Induces Fetal Hemoglobin. ACS Cent. Sci. 2022, 8, 1695–1703. [Google Scholar] [CrossRef]
  58. Trentini, D.B.; Suskiewicz, M.J.; Heuck, A.; Kurzbauer, R.; Deszcz, L.; Mechtler, K.; Clausen, T. Arginine phosphorylation marks proteins for degradation by a Clp protease. Nature 2016, 539, 48–53. [Google Scholar] [CrossRef]
  59. Francesca, E.M.; Stefan, K.; Julia, L.; Sabryna, J.; David, M.H.; Stepan, O.; Anastasia, O.; Juliane, K.; Robert, K.; Lukas, J.; et al. BacPROTACs mediate targeted protein degradation in bacteria. Cell 2022, 185, 2338–2353. [Google Scholar]
  60. Arunima, M.; Izidor, S.; Martina, G.; Patricia, L.; Matic, P.; Sophie, W.; Rabea, V.; Michael, G.; Jan, K.; Christian, S. Targeting the deubiquitinase USP7 for degradation with PROTACs. Chem. Commun. 2022, 58, 8858–8861. [Google Scholar]
  61. Verba, K.A.; Wang, R.Y.-R.; Arakawa, A.; Liu, Y.; Shirouzu, M.; Yokoyama, S.; Agard, D.A. STRUCTURAL BIOLOGY Atomic structure of Hsp90-Cdc37-Cdk4 reveals that Hsp90 traps and stabilizes an unfolded kinase. Science 2016, 352, 1542–1547. [Google Scholar] [CrossRef] [Green Version]
  62. Verba, K.A.; Agard, D.A. How Hsp90 and Cdc37 Lubricate Kinase Molecular Switches. Trends Biochem. Sci. 2017, 42, 799–811. [Google Scholar] [CrossRef]
  63. Sun, Y.; Huang, Y.-H.; Huang, F.-Y.; Mei, W.-L.; Liu, Q.; Wang, C.-C.; Lin, Y.-Y.; Huang, C.; Li, Y.-N.; Dai, H.-F.; et al. 3′-epi-12 beta-hydroxyfroside, a new cardenolide, induces cytoprotective autophagy via blocking the Hsp90/Akt/mTOR axis in lung cancer cells. Theranostics 2018, 8, 2044–2060. [Google Scholar] [CrossRef]
  64. Matyskiela, M.E.; Lu, G.; Ito, T.; Pagarigan, B.; Lu, C.-C.; Miller, K.; Fang, W.; Wang, N.-Y.; Nguyen, D.; Houston, J.; et al. A novel cereblon modulator recruits GSPT1 to the CRL4(CRBN) ubiquitin ligase. Nature 2016, 535, 252. [Google Scholar] [CrossRef]
  65. Churcher, I. Protac-Induced Protein Degradation in Drug Discovery: Breaking the Rules or Just Making New Ones? J. Med. Chem. 2018, 61, 444–452. [Google Scholar] [CrossRef]
  66. Zhang, L.; Riley-Gillis, B.; Vijay, P.; Shen, Y. Acquired Resistance to BET-PROTACs (Proteolysis-Targeting Chimeras) Caused by Genomic Alterations in Core Components of E3 Ligase Complexes. Mol. Cancer Ther. 2019, 18, 1302–1311. [Google Scholar] [CrossRef] [Green Version]
  67. Li, Z.; Ma, S.; Zhang, S.; Ma, Z.; Du, L.; Li, M. Targeted Protein Degradation Induced by HEMTACs Based on HSP90. J. Med. Chem. 2023, 66, 733–751. [Google Scholar] [CrossRef]
  68. Asghar, U.; Witkiewicz, A.K.; Turner, N.C.; Knudsen, E.S. The history and future of targeting cyclin-dependent kinases in cancer therapy. Nat. Rev. Drug Discov. 2015, 14, 130–146. [Google Scholar] [CrossRef] [Green Version]
  69. Classon, M.; Harlow, E. The retinoblastoma tumor suppressor in development and cancer. Nat. Rev. Cancer 2002, 2, 910–991. [Google Scholar] [CrossRef]
  70. Malumbres, M.; Sotillo, R.; Santamaría, D.; Galán, J.; Cerezo, A.; Ortega, S.; Dubus, P.; Barbacid, M. Mammalian cells cycle without the D-type cyclin-elependent kinases Cdk4 and Cdk6. Cell 2004, 118, 493–504. [Google Scholar] [CrossRef] [Green Version]
  71. Fassl, A.; Geng, Y.; Sicinski, P. CDK4 and CDK6 kinases: From basic science to cancer therapy. Science 2022, 375, eabc1495. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic of the mechanism of proteolysis-targeting chimeras (PROTACs).
Figure 1. Schematic of the mechanism of proteolysis-targeting chimeras (PROTACs).
Ijms 24 05190 g001
Figure 2. Mode of action—Aptamer–PROTACs: upon recognition by the cell membrane receptor (nucleolin), aptamer–PROTACs enter the cells, and endocytosis followed by linker cleavage result in the liberation of active PROTAC; Pre-PROTACs: inserting reactive oxygen species (ROS)-triggered leaving group to the parent PROTAC results in the release of the active PROTAC in the presence of ROS of the tumor microenvironment (TME); Hydrophobic tags: hydrophobic degron mimics the exposed hydrophobic region of misfolded proteins and is recognized by the heat shock protein for proteasomal degradation; RT-PROTACs: radiotherapy-triggered PROTAC prodrug incorporated with phenyl azide caging group releases the active PROTAC molecule upon X-ray irradiation; NTR-responsive PROTACs: caging of the nitroreductase (NTR)-responsive group releases the active PROTAC in the presence of hypoxia microenvironment.
Figure 2. Mode of action—Aptamer–PROTACs: upon recognition by the cell membrane receptor (nucleolin), aptamer–PROTACs enter the cells, and endocytosis followed by linker cleavage result in the liberation of active PROTAC; Pre-PROTACs: inserting reactive oxygen species (ROS)-triggered leaving group to the parent PROTAC results in the release of the active PROTAC in the presence of ROS of the tumor microenvironment (TME); Hydrophobic tags: hydrophobic degron mimics the exposed hydrophobic region of misfolded proteins and is recognized by the heat shock protein for proteasomal degradation; RT-PROTACs: radiotherapy-triggered PROTAC prodrug incorporated with phenyl azide caging group releases the active PROTAC molecule upon X-ray irradiation; NTR-responsive PROTACs: caging of the nitroreductase (NTR)-responsive group releases the active PROTAC in the presence of hypoxia microenvironment.
Ijms 24 05190 g002
Figure 3. Aptamer-constructed PROTAC: (A) mode of action; (B) chemical structure.
Figure 3. Aptamer-constructed PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g003
Figure 4. (A) Mode of action in opto-PROTAC; (B) chemical structure of apt-PROTAC.
Figure 4. (A) Mode of action in opto-PROTAC; (B) chemical structure of apt-PROTAC.
Ijms 24 05190 g004
Figure 5. Pre-PROTAC: (A) mode of action; (B) chemical structure.
Figure 5. Pre-PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g005
Figure 6. PS-Degrons: (A) mode of action; (B) chemical structure.
Figure 6. PS-Degrons: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g006
Figure 7. AP-PROTAC: (A) mode of action; (B) chemical structure.
Figure 7. AP-PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g007
Figure 8. RT-PROTAC: (A) mode of action; (B) chemical structure.
Figure 8. RT-PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g008
Figure 9. NTR-based PROTAC: (A) mode of action; (B) chemical structure.
Figure 9. NTR-based PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g009
Figure 10. ENCTAC: (A) mode of action; (B) chemical structure.
Figure 10. ENCTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g010
Figure 11. Poly-PROTAC: (A) mode of action; (B) chemical structure.
Figure 11. Poly-PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g011
Figure 12. Chemical structure of smart nano-PROTAC.
Figure 12. Chemical structure of smart nano-PROTAC.
Ijms 24 05190 g012
Figure 13. Split-PROTAC: (A) mode of action; (B) Chemical structure.
Figure 13. Split-PROTAC: (A) mode of action; (B) Chemical structure.
Ijms 24 05190 g013
Figure 14. Hydrophobic tag: (A) mode of action; (B) chemical structure.
Figure 14. Hydrophobic tag: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g014
Figure 15. Stapled PROTAC: (A) mode of action; (B) chemical structure.
Figure 15. Stapled PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g015
Figure 16. Bridged PROTAC: (A) mode of action; (B) chemical structure.
Figure 16. Bridged PROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g016
Figure 17. TF–DUBTAC: (A) mode of action; (B) chemical structure.
Figure 17. TF–DUBTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g017
Figure 18. Mode of action by nanobody-based degrader.
Figure 18. Mode of action by nanobody-based degrader.
Ijms 24 05190 g018
Figure 19. BacPROTAC: (A) mode of action; (B) chemical structure.
Figure 19. BacPROTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g019
Figure 20. DUB–based degrader: (A) mode of action; (B) chemical structure.
Figure 20. DUB–based degrader: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g020
Figure 21. HEMTAC: (A) mode of action; (B) chemical structure.
Figure 21. HEMTAC: (A) mode of action; (B) chemical structure.
Ijms 24 05190 g021
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

Tamatam, R.; Shin, D. Emerging Strategies in Proteolysis-Targeting Chimeras (PROTACs): Highlights from 2022. Int. J. Mol. Sci. 2023, 24, 5190. https://doi.org/10.3390/ijms24065190

AMA Style

Tamatam R, Shin D. Emerging Strategies in Proteolysis-Targeting Chimeras (PROTACs): Highlights from 2022. International Journal of Molecular Sciences. 2023; 24(6):5190. https://doi.org/10.3390/ijms24065190

Chicago/Turabian Style

Tamatam, Rekha, and Dongyun Shin. 2023. "Emerging Strategies in Proteolysis-Targeting Chimeras (PROTACs): Highlights from 2022" International Journal of Molecular Sciences 24, no. 6: 5190. https://doi.org/10.3390/ijms24065190

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

Article Metrics

Back to TopTop