Acessibilidade / Reportar erro

The Thin Line between Promiscuous and Privileged Structures in Medicinal Chemistry

Abstract

The concept of privileged structures in medicinal chemistry refers to commonly found substructures in approved drugs and lead small molecules, presenting a broad profile of pharmacological action, i.e., participating in the recognition by several classes of pharmacological targets or by modulating physicochemical properties. Privileged structures are also related to nontoxic effects, which make their use in the design of new drug candidates very attractive. In contrast, another concept also refers to structures or substructures capable of presenting a pleiotropic profile for several pharmacological targets, referred to as promiscuous compounds. Worth mentioning, more recently, a great majority of promiscuous compounds have been classified as Pan Assay Interference Compounds (PAINS). In its great majority, PAINS are electrophilic in nature, capable of covalently reacting indiscriminately with several pharmacological targets, and are associated with specific substructures, which are, in turn, used as an exclusion filter during screening campaigns. This work aims to critically discuss the thin line that separates the two concepts, clarifying their differences from a molecular and pharmacological point of view. Moreover, special considerations regarding PAINS and exclusion filters will be made.

Keywords:
privileged structures; privileged scaffolds; frameworks; chemical promiscuity; PAINS


1. Introduction

Privileged structures and promiscuous compounds represent somewhat intriguing and contradictory phenomena in the field of medicinal chemistry, drawing substantial attention due to their significant impact on drug discovery and development. The concept of privileged structures, originally described in the seminal works of Evans et al.,11 Evans, B. E.; Rittle, K. E.; Bock, M. G.; DiPardo, R. M.; Freidinger, R. M.; Whitter, W. L.; Lundell, G. F.; Veber, D. F.; Anderson, P. S.; Chang, R. S. L.; Lotti, V. J.; Cerino, D. J.; Chen, T. B.; Kling, P. J.; Kunkel, K. A.; Springer, J. P.; Hirshfield, J.; J. Med. Chem. 1988, 31, 2235. [Crossref]
Crossref...
,22 Evans, B. E.; Bock, M. G.; Rittle, K. E.; DiPardo, R. M.; Whitter, W. L.; Veber, D. F.; Anderson, P. S.; Freidinger, R. M.; Proc. Natl. Acad. Sci. U.S.A. 1986, 83, 4918. [Crossref]
Crossref...
refers to specific molecular frameworks that are present in the structures of several bioactive small molecules and, consequently, possess a remarkable ability to exhibit diverse biological activities by interacting with multiple pharmacological targets. These privileged structures serve as valuable templates for medicinal chemists, offering a starting point for the design and synthesis of novel bioactive compounds or drug candidates.33 Duarte, C.; Barreiro, E.; Fraga, C.; Mini-Rev. Med. Chem. 2007, 7, 1108. [Crossref]
Crossref...
,44 Skoreński, M.; Sieńczyk, M.; Pharmaceuticals 2021, 14, 1164. [Crossref]
Crossref...
,55 Kim, J.; Kim, H.; Park, S. B.; J. Am. Chem. Soc. 2014, 136, 14629. [Crossref]
Crossref...

Concurrently and controversially, the notion of promiscuous compounds has garnered attention for its potential implications in drug development. Compounds that are promiscuous, similar to what is observed for privileged structures, have the unique trait of interacting with various biological targets. This extends their biological effects beyond the initially intended receptors. However, unlike privileged structures, promiscuous compounds are frequently associated with negative side effects and/or toxicity.66 Scheiber, J.; Jenkins, J. L.; Sukuru, S. C. K.; Bender, A.; Mikhailov, D.; Milik, M.; Azzaoui, K.; Whitebread, S.; Hamon, J.; Urban, L.; Glick, M.; Davies, J. W.; J. Med. Chem. 2009, 52, 3103. [Crossref]
Crossref...
,77 Espinoza-Fonseca, L. M.; Bioorg. Med. Chem. 2006, 14, 896. [Crossref]
Crossref...
,88 Hu, Y.; Bajorath, J.; Eur. J. Med. Chem. 2013, 69, 393. [Crossref]
Crossref...

Despite the intricate relationship between privileged structures and promiscuous compounds it is absolutely necessary to emphasize the nuances of their concepts and understand their roles in drug design and discovery processes.

In addition, it is important to highlight that the concept of promiscuity in medicinal chemistry has been shaped throughout the 21st century and its “definition” is also currently related to the concept of Pan Assay Interference Compounds (PAINS).99 Baell, J.; Walters, M. A.; Nature 2014, 513, 481. [Crossref]
Crossref...

The exploration of privileged structures and promiscuous compounds represents a dynamic and evolving frontier in medicinal chemistry. Through rigorous scientific inquiry and innovative approaches, researchers from both academia and industry strive to capitalize on the unique attributes of these molecular entities to advance the development of safer and more effective therapeutic innovations.1010 David, L.; Walsh, J.; Sturm, N.; Feierberg, I.; Nissink, J. W. M.; Chen, H.; Bajorath, J.; Engkvist, O.; ChemMedChem. 2019, 14, 1795. [Crossref]
Crossref...
,1111 Blaschke, T.; Miljković, F.; Bajorath, J.; ACS Omega 2019, 4, 6883. [Crossref]
Crossref...
,1212 Mencher, S. K.; Wang, L. G.; BMC Clin. Pharmacol. 2005, 5, 3. [Crossref]
Crossref...
,1313 Hu, Y.; Bajorath, J.; J. Chem. Inf. Model. 2010, 50, 2112. [Crossref]
Crossref...

This comprehensive review focuses on discussing and defining important concepts related to privileged structures and promiscuous compounds, bringing to light the fine line that separates the two. We hope that this work will assist in the critical understanding of these concepts in medicinal chemistry, avoiding their mistaken use.

2. Privileged Structures in Medicinal Chemistry

2.1. Classic privileged structures

In this section, classic privileged structures will be highlighted and discussed. These type of privileged structures and scaffolds are understood as those standard motifs deeply exploited and discussed by the medicinal chemistry community.

The classical privileged structures, as elucidated in prior studies, are inherently linked to well-established findings that highlight their notable biological activities. These structures or substructures serve as valuable starting points in the quest for the discovery of novel drugs. Their selection is grounded in the precise understanding of their interaction capabilities with biological targets, thereby enhancing the efficiency of drug discovery and development processes.33 Duarte, C.; Barreiro, E.; Fraga, C.; Mini-Rev. Med. Chem. 2007, 7, 1108. [Crossref]
Crossref...
,1414 Barreiro, E. J. In Privileged Scaffolds in Medicinal Chemistry: Design, Synthesis, Evaluation; Bräse, S., ed.; The Royal Society of Chemistry: Cambridge, UK, 2015, ch. 1, p. 1-15. [Crossref]
Crossref...

The 7-azaindole scaffold is considered a privileged structure in medicinal chemistry and has garnered significant attention due to its special physicochemical and pharmacological properties.1515 Han, Y.; Dong, W.; Guo, Q.; Li, X.; Huang, L.; Eur. J. Med. Chem. 2020, 203, 112506. [Crossref]
Crossref...
The ability of nitrogen atoms in the 7-azaindole core to form hydrogen bonds with the hinge region in the adenosine triphosphate (ATP) binding site makes 7-azaindole derivatives crucial sources for kinase inhibitor design.1616 Fang, G.; Chen, H.; Cheng, Z.; Tang, Z.; Wan, Y.; Eur. J. Med. Chem. 2023, 115621. [Crossref]
Crossref...

To date, some 7-azaindole derivatives have been marketed or undergone clinical trials for the treatment of kinase-associated diseases. Vemurafenib (1), a serine/ threonine-protein kinase B-raf (BRAF) inhibitor was the first U.S. Food and Drug Administration (FDA)-approved drug based on 7-azaindole for the treatment of BRAF-mutated metastatic melanoma (Figure 1).1717 Kim, A.; Cohen, M. S.; Expert. Opin. Drug. Discovery 2016, 11, 907. [Crossref]
Crossref...
Another example is pexidartinib (2), a tyrosine kinase inhibitor selectively targeting colony-stimulating factor 1 receptor (CSF1R), FDA approved in 2019 and available in the market for treating giant cell tumors of the tendon sheath in adult patients (Figure 1).1818 Monestime, S.; Lazaridis, D.; Drugs R&D 2020, 20, 189. [Crossref]
Crossref...
Additionally, decernotinib (3), a Janus Kinase 3 (JACK3) inhibitor, has been the subject of clinical trials for the treatment of rheumatoid arthritis (Figure 1).1919 Gadina, M.; Schwartz, D. M.; O’Shea, J. J.; Arthritis Rheumatol. 2016, 68, 31. [Crossref]
Crossref...

Figure 1
7-Azaindole-based kinase inhibitors in clinical use or under clinical trial studies.

Most kinase inhibitors are designed to interact with the hinge region, acting as competitive inhibitors with respect to adenosine triphosphate (ATP).2020 Liao, J. J.-L.; J. Med. Chem. 2007, 50, 409. [Crossref]
Crossref...
Among azaindole-derived kinase inhibitors, three different binding modes with kinases can be observed. The “normal” mode, where the azaindole scaffold binds to the hinge; the “inverted” mode, in which the azaindole binds in the opposite direction at a 180° angle compared to the normal mode; and finally, the “non-hinge” mode, where the azaindole binds to a distinct site away from the hinge region (Figure 2). In the “normal” and “inverted” binding modes, the azaindole subunit can form two hydrogen bonds with amino acid residues in the hinge region. In the “normal” mode, azaindole interacts with GK+1 (adjacent to the gatekeeper amino acid residue) and GK+3 residues through bidentate hydrogen bonds (Figure 2). On the other hand, in the “inverted” binding mode, bidentate hydrogen bonds occur between azaindole and the GK+3 residue (Figure 2).1616 Fang, G.; Chen, H.; Cheng, Z.; Tang, Z.; Wan, Y.; Eur. J. Med. Chem. 2023, 115621. [Crossref]
Crossref...
,2121 Irie, T.; Sawa, M.; Chem. Pharm. Bull. 2018, 66, 29. [Crossref]
Crossref...

Figure 2
Different interaction modes of privileged azaindole subunit with kinase active binding site.

The imidazoles hold a distinctive position in the field of heterocyclic chemistry. Due to their versatile properties in the realms of chemistry and pharmacology, their derivatives have been the focus of interest, as they play a significant role in both biological and pharmaceutical contexts.2222 Verma, A.; Joshi, S.; Singh, D.; J. Chem. 2013, 2013, ID 329412. [Crossref]
Crossref...

In biological systems, imidazole constitutes the main structure of various naturally occurring compounds, such as the amino acid histidine (4), the nitrogenous bases guanine (5) and adenosine in nucleic acids (6), the neurotransmitter histamine (7) (Figure 3).2222 Verma, A.; Joshi, S.; Singh, D.; J. Chem. 2013, 2013, ID 329412. [Crossref]
Crossref...
,2323 Alghamdi, S. S.; Suliman, R. S.; Almutairi, K.; Kahtani, K.; Aljatli, D.; Drug Des., Dev. Ther. 2021, 2021, 3289. [Crossref]
Crossref...

Figure 3
Presence of imidazole in different bioactive natural compounds.

The presence of the imidazole nucleus can be observed in a wide variety of chemical compounds of pharmaceutical interest, whether they are of natural or synthetic origin. It presents a classic tautomeric behavior between nitrogen atoms N1 and N3 of the heterocyclic ring that favors its molecular recognition by different receptors. For this reason, imidazole derivatives exhibit a broad spectrum of biological activities, including antibacterial, anticancer, anti-inflammatory, antifungal, antimalarial, antitubercular, and many others.2222 Verma, A.; Joshi, S.; Singh, D.; J. Chem. 2013, 2013, ID 329412. [Crossref]
Crossref...
,2323 Alghamdi, S. S.; Suliman, R. S.; Almutairi, K.; Kahtani, K.; Aljatli, D.; Drug Des., Dev. Ther. 2021, 2021, 3289. [Crossref]
Crossref...
,2424 Tolomeu, H. V.; Fraga, C. A. M.; Molecules 2023, 28, 838. [Crossref]
Crossref...
Among natural compounds, one example is topsentin (8), which exhibits antiviral, antifungal, anticancer, and anti-inflammatory activities (Figure 4).2525 Hwang, J.; Kim, D.; Park, J. S.; Park, H. J.; Shin, J.; Lee, S. K.; Mar. Drugs 2020, 18, 87. [Crossref]
Crossref...
Isonaamine A (9) demonstrates anticancer activity by acting as an inhibitor of the epidermal growth factor receptor (EGFR) (Figure 4).2626 Gibbons, J. B.; Salvant, J. M.; Vaden, R. M.; Kwon, K.-H.; Welm, B. E.; Looper, R. E.; J. Org. Chem. 2015, 80, 10076. [Crossref]
Crossref...
Additionally, pilocarpine (10) is utilized in the management and treatment of xerostomia and glaucoma (Figure 4).2727 Kapourani, A.; Kontogiannopoulos, K. N.; Barmpalexis, P.; Pharmaceuticals 2022, 15, 762. [Crossref]
Crossref...

Figure 4
Examples of natural products containing the privileged imidazole ring.

Additionally, there are several drugs containing the imidazole nucleus currently available in the market, such as losartan (11) an angiotensin II receptor agonist that is used as an antihypertensive agent.2828 Goa, K. L.; Wagstaff, A. J.; Drugs 1996, 51, 820. [Crossref]
Crossref...
In the structure of losartan (11, Figure 5), various pharmacophoric groups play crucial roles in interacting with the angiotensin II (Ang II) receptor. The imidazole heterocycle and the N-butyl side chain mimic the His6 and Ile5 side chains of the C-terminal region of angiotensin II. Additionally, the tetrazole ring in its structure interacts with the Asp1 and Tyr4 residues of the N-terminal region of Ang II, contributing to its pharmacological activity.2929 Xu, F.; Mao, C.; Liu, Y.; Wu, L.; Xu, Z.; Zhang, L.; Curr. Med. Chem. 2009, 16, 3701. [Crossref]
Crossref...

Figure 5
Structure of drugs on market containing privileged imidazole ring.

Ketoconazole (12, Figure 5), an antifungal drug, competitively inhibits sterol 14-α-demethylase (CYP51), a key enzyme in catalyzing the process of oxidative removal of 14-α-methyl during the biosynthesis of ergosterol, the main sterol in fungal membranes.3030 Sohn, C. A.; Clin. Pharm. 1982, 1, 217. [PubMed]
PubMed...
Ketoconazole, like other ergosterol biosynthesis inhibitors in the azole class, has four interaction sites with the CYP51 enzyme. The nitrogen atom in the imidazole ring of ketoconazole at site 1 complexes with the iron of the heme group present in the target enzyme, blocking binding with molecular oxygen. At site 2, oxygen acts as a hydrogen bond acceptor, and the structures at sites 3 and 4 interact with the hydrophobic part of the CYP51 enzyme.3030 Sohn, C. A.; Clin. Pharm. 1982, 1, 217. [PubMed]
PubMed...

Cimetidine (13, Figure 5) is an H2 receptor antagonist that completely blocks the stimulation of H2 receptors located in gastric parietal cells by histamine (7). Cimetidine’s structure is based on the histamine (7) prototype, the natural agonist. Due to the imidazole’s nature, tautomerism can be observed, influencing molecular recognition by different histamine receptor subtypes (H1 and H2). The imidazole ring in cimetidine (13) has a methyl group at C-5, contributing to the tautomeric form essential for the desired selectivity for H2 receptors. The presence of the thioether group on the side chain of its structure provides desirable hydrophobic properties and contributes to inhibiting the tautomeric balance of the heterocyclic ring.3131 Finkelstein, W.; Isselbacher, K. J.; N. Engl. J. Med. 1978, 299, 992. [Crossref]
Crossref...

Other interesting examples of drugs containing the privileged imidazole ring include flumizole (14, Figure 5)3232 McIlhenny, H. M.; Bettis, J. W.; Wiseman, E. H.; J. Pharm. Sci. 1975, 64, 1469. [Crossref]
Crossref...
a non-steroidal anti-inflammatory that inhibits the enzyme cyclooxygenase (COX), azomycin (15, Figure 5),3333 Maeda, K.; Osato, T.; Umezawa, H.; J. Antibiot. 1953, 6, 182. [PubMed]
PubMed...
which has antibiotic activity against Gram-positive bacteria, dacarbazine (16, Figure 5),3434 Al-Badr, A. A.; Alodhaib, M. M.; Profiles Drug Subst., Excipients, Relat. Methodol. 2016, 41, 323. [Crossref]
Crossref...
an anticancer agent whose mechanism of action is not very well established, but the main hypothesis is that it inhibits deoxyribonucleic acid (DNA) synthesis as an analog of purines, nafimidone (17, Figure 5),3535 Hempel, A.; Camerman, N.; Camerman, A.; Mastropaolo, D.; Acta Crystallogr., Sect. E: Struct. Rep. Online 2005, 61, o1387. [Crossref]
Crossref...
an anticonvulsant which, like its metabolite, is capable of inhibiting the main metabolism pathways of important antiepileptic drugs such as phenytoin and carbamazepine and others.

Another heterocycle of extreme importance for medicinal chemists is benzimidazole. This privileged scaffold represents fundamental structures found in diverse libraries of therapeutically relevant agents used in medical research and drug discovery campaigns. Benzimidazole stands out among heterocyclic compounds due to its remarkable biological actions and synthetic applications in medicinal chemistry, serving as an essential building block for the synthesis of a wide range of bioactive compounds.3636 Ali, A. M.; Tawfk, S. S.; Mostafa, A. S.; Massoud, M. A. M.; Chem. Biol. Drug Des. 2022, 100, 656. [Crossref]
Crossref...
,3737 Guo, Y.; Hou, X.; Fang, H.; Mini-Rev. Med. Chem. 2021, 21, 1367. [Crossref]
Crossref...

The core benzimidazole nucleus is present in several well-known drugs in current pharmacotherapy for both humans and animals. Among the benzimidazole-containing compounds are omeprazole (18), which has anti-inflammatory activity, and interrupts gastric acid secretion by selectively inhibiting the H+/K+ ATPase enzyme system; albendazole (19), an anthelmintic that acts by inhibiting tubulin polymerization, it also impairs glucose utilization and reduces the parasite’s glycogen reserves, which results in decreased ATP production, leading to energy depletion and the subsequent death of the parasite. There is also rabeprazole (20), an anti-ulcer drug that also acts by inhibiting the proton pump in gastric parietal cells (H+/K+- ATPase), resulting in inhibition of gastric acid production; telmisartan (21) an antihypertensive agent; capable of blocking the vasoconstrictive and aldosterone-secreting effects of angiotensin II by selectively blocking the binding of angiotensin II to AT1; benomyl (22), an antifungal agent that promotes the depolymerization of microtubules; and bendamustine (23), a drug with anticancer activity that activates the stress response by damaging DNA, induces apoptosis, inhibits mitotic checkpoints and induces mitotic catastrophe (Figure 6).3838 Ajani, O. O.; Aderohunmu, D. V; Ikpo, C. O.; Adedapo, A. E.; Olanrewaju, I. O.; Arch. Pharm. 2016, 349, 475. [Crossref]
Crossref...
,3939 Welsch, M. E.; Snyder, S. A.; Stockwell, B. R.; Curr. Opin. Chem. Biol. 2010, 14, 347. [Crossref]
Crossref...

Figure 6
Examples of benzimidazole-containing drugs.

N-Acylhydrazone (NAH) frameworks are considered privileged structures in medicinal chemistry. This versatile peptidemimetic subunit presents remarkable chemical stability against hydrolysis, the ability to perform putative interactions with different receptors, and synthetic accessibility, facilitating rational modifications during the structural optimization of a lead compound. Such modifications enable the generation of compounds that exhibit affinity and selectivity toward specific pharmacological targets.33 Duarte, C.; Barreiro, E.; Fraga, C.; Mini-Rev. Med. Chem. 2007, 7, 1108. [Crossref]
Crossref...

Despite significant advances in the last decade in the discovery of small-sized NAH-based drugs, only a few NAHs have received clinical approval. Among these drugs, nitrofurazone (24) and nitrofurantoin (25), belonging to the chemical class associated with NAHs, namely semicarbazones, were the first compounds approved for clinical use (Figure 7). Both derivatives were approved for the treatment of bacterial infections. Nitrofurazone (24) is used topically and its mechanism of action is the inactivation of ribosomal proteins, which inhibits the synthesis of proteins, DNA, ribonucleic acids (RNA), and cell wall synthesis, blocking the aerobic metabolism of bacterial cells.4040 Martini, C. A. N.; Scapini, J. G. S.; Collaço, L. M.; Matsubara A.; Veiga Jr., V. F.; Rev. Col. Bras. Cir. 2016, 43, 445. [Crossref]
Crossref...
Nitrofurantoin (25) is administered orally for treating genitourinary tract infections and is reduced by bacterial flavoproteins which generate reactive intermediates that inhibit protein synthesis, aerobic energy metabolism, DNA, RNA, and cell wall synthesis.4141 Guay, D. R.; Drugs 2001, 61, 353. [Crossref]
Crossref...
Another approved drug associated with semicarbazones, carbazochrome (26), is used as a hemostatic agent and it is indicated for capillary and parenchymal bleeding. It increases platelet aggregation and forms a platelet buffer by interacting with α-adrenergic receptors on the surface of platelets (Figure 7). The sodium salt of dantrolene (27) and its azumolene (28) sodium salt analog are also NAH-derived drugs, both approved for treating malignant hyperthermia (MH) and act directly on the contractile mechanism of skeletal muscle, reducing the force of contraction, without showing any effects on neural pathways, the neuromuscular junction or the excitable properties of muscle fiber membranes (Figure 7).4242 Krause, T.; Gerbershagen, M. U.; Fiege, M.; Weisshorn, R.; Wappler, F.; Anaesthesia 2004, 59, 364. [Crossref]
Crossref...
,4343 Corrêa, J. C. R.; Hiene, M. A. C.; Salgado, H. R. N.; J. Anal. Bioanal. Tech. 2013, 177. [Crossref]
Crossref...
,4444 Pinder, R. M.; Brogden, R. N.; Speight, T. M.; Avery, G. S.; Drugs 1977, 13, 3. [Crossref]
Crossref...

Figure 7
Structure of some approved drugs presenting the NAH privileged framework.

Despite only a few of the drugs and drug candidates containing NAH showed significant advances in the pre-clinical stage in recent years, promising lead compounds derived from NAH are documented in the literature. Among these, it is worth highlighting the compound LASSBio-294 (29, Figure 8), which targets the molecular adenosine A2A receptor and has demonstrated selective binding. This NAH derivative was designed for the treatment of heart failure, exhibiting in vitro and in vivo positive inotropic activity and moderate vasodilation.4545 da Silva, J. S.; Pereira, S. L.; do Couto Maia, R.; Landgraf, S. S.; Caruso-Neves, C.; Kümmerle, A. E.; Fraga, C. A. M.; Barreiro, E. J.; Sudo, R. T.; Zapata-Sudo, G.; Life Sci. 2014, 94, 30. [Crossref]
Crossref...
,4646 Silva, C. L. M.; Noël, F.; Barreiro, E. J.; B r. J. Pharmacol. 2002, 135, 293. [Crossref]
Crossref...

Figure 8
Lead compounds containing NAH framework.

From this compound (29), other NAH analogs have been obtained. Its thiophene regioisomer (30) and the N-methylated analog (31) exhibited a more significant vasodilatory effect. The pharmacological properties of compound 30 appear to be associated with the agonism of muscarinic M3 receptors, while the N-methylated analog (31) acts as a blocker of Ca2+ channel types.4747 Zapata-Sudo, G.; Pereira, S. L.; Beiral, H. J. V.; Kummerle, A. E.; Raimundo, J. M.; Antunes, F.; Sudo, R. T.; Barreiro, E. J.; Fraga, C. A. M.; Am. J. Hypertens. 2010, 23, 135. [Crossref]
Crossref...
,4848 Pereira, S. L.; Kummerle, A. E.; Fraga, C. A. M.; Barreiro, E. J.; de Novaes Rocha, N.; Ferraz, E. B.; do Nascimento, J. H. M.; Sudo, R. T.; Zapata-Sudo, G.; Eur. J. Pharmacol. 2013, 702, 316. [Crossref]
Crossref...
,4949 Pereira, S. L.; Kummerle, A. E.; Fraga, C. A. M.; Barreiro, E. J.; Sudo, R. T.; Zapata-Sudo, G.; Fundam. Clin. Pharmacol. 2014, 28, 29. [Crossref]
Crossref...

Compounds derived from quinolines possess various intriguing properties and find applications in diverse fields, including medicinal chemistry. In this domain, quinolines are deemed privileged structures due to their particular capability to be recognized by distinct biotargets, rendering them pertinent in drug discovery process.5050 Mao, Y.; Soni, K.; Sangani, C.; Yao, Y.; Curr. Top. Med. Chem. 2020, 20, 2599. [Crossref]
Crossref...
,5151 Chu, X.-M.; Wang, C.; Liu, W.; Liang, L.-L.; Gong, K.-K.; Zhao, C.-Y.; Sun, K.-L.; Eur. J. Med. Chem. 2019, 161, 101. [Crossref]
Crossref...

The presence of quinoline as a structural constituent can be observed in various compounds, both natural and synthetic, exhibiting a wide range of biological activity. These compounds display diverse pharmacological properties such as analgesic, antibacterial, anti-inflammatory, anticancer, anticonvulsant, and others.5252 Yadav, V.; Reang, J.; Sharma, V.; Majeed, J.; Sharma, P. C.; Sharma, K.; Giri, N.; Kumar, A.; Tonk, R. K.; Chem. Biol. Drug Des. 2022, 100, 389. [Crossref]
Crossref...

It is known that drugs containing the quinoline ring find broad therapeutic applications in the pharmaceutical domain. Chloroquine (32, Figure 9), a commercially available antimalarial, features this subunit. This drug interferes with the parasite’s digestion of hemoglobin, so chloroquine accumulates in the digestive vacuole of the parasite, compromising hemoglobin digestion and leading to its death. The quinoline ring is important for antimalarial activity and enables chloroquine to inhibit the polymerization of the heme group, a toxic by-product of the parasite’s digestion of hemoglobin. By preventing the crystallization of hemozoin from the heme group, chloroquine contributes to the death of the parasite.5353 da Silva, T. H. Á.; de Oliveira, M. T.; dos Santos, H. F.; de Oliveira, A. B.; de Almeida, W. B.; Quim. Nova 2005, 28, 244. [Crossref]
Crossref...
Besides this drug, there are others such as the antiviral saquinavir (33), which acts by blocking a protease enzyme that is necessary for the reproduction of human immunodeficiency virus (HIV). Saquinavir is a competitive peptidomimetic inhibitor that mimics the natural substrate of the viral protease. The quinoline ring present in its structure almost entirely occupies the protease enzyme pocket, helping to inhibit HIV-1 protease activity.5454 Agnello, S.; Brand, M.; Chellat, M. F.; Gazzola, S.; Riedl, R.; Angew. Chem., Int. Ed. 2019, 58, 3300. [Crossref]
Crossref...
The antiprotozoal clioquinol (34) can inhibit the activity of enzymes essential for the metabolism of the protozoan resulting in metabolic dysfunction and cell death. The 8-hydroxyquinoline group present in the clioquinol (34) structure has both antiprotozoal and antibacterial actions. The presence of the nitrogen atom in the quinoline ring allows this compound to have the chelating properties necessary for inhibiting bacterial biofilms.5555 Repac Antić, D.; Parčina, M.; Gobin, I.; Petković Didović, M.; Antibiotics 2022, 11, 1105. [Crossref]
Crossref...
The antiasthmatic montelukast (35) is a selective leukotriene receptor antagonist that inhibits the CysLT1 (cysteine leukotriene) receptor. The quinoline group of montelukast interacts with the hydrophobic region of the CysLT1 receptor binding site and is therefore essential for its activity as it prevents leukotrienes from binding to and activating the receptor, thus blocking the inflammatory and bronchoconstrictive effects mediated by leukotrienes.5656 Zwaagstra, M. E.; Schoenmakers, S. H. H. F.; Nederkoorn, P. H. J.; Gelens, E.; Timmerman, H.; Zhang, M.-Q.; J. Med. Chem. 1998, 41, 1439. [Crossref]
Crossref...
The antibacterial ciprofloxacin (36) which inhibits bacterial topoisomerase type II and topoisomerase IV, which are essential for replication, transcription, repair, and recombination of bacterial DNA. Quinolones, derived from the quinoline nucleus present in ciprofloxacin, bind non-covalently to the enzyme-DNA interface at the active cleavage-binding site, increasing the concentration of enzyme-DNA cleavage complexes.5757 Aldred, K. J.; Kerns, R. J.; Osheroff, N.; Biochemistry 2014, 53, 1565. [Crossref]
Crossref...
The anticancer drug irinotecan (37) also interacts with topoisomerase I, an important enzyme in the cell multiplication process (Figure 9).5252 Yadav, V.; Reang, J.; Sharma, V.; Majeed, J.; Sharma, P. C.; Sharma, K.; Giri, N.; Kumar, A.; Tonk, R. K.; Chem. Biol. Drug Des. 2022, 100, 389. [Crossref]
Crossref...

Figure 9
Examples of drugs containing the privileged quinoline ring.

The basic structure of benzoxazine can be modified to generate various compounds, each with specific properties. This versatility makes benzoxazine intriguing for chemical synthesis, particularly in the production of compounds with potential biological or pharmacological properties, underscoring its significance in the research and development of new drugs.5858 Tang, Z.; Tan, Y.; Chen, H.; Wan, Y.; Curr. Med. Chem. 2023, 30, 372. [Crossref]
Crossref...

In this context, there is a variety of bioactive compounds containing benzoxazine that exhibit diverse pharmacological activities such as anticancer activity, antibacterial effects, antifungal properties, and antituberculosis potential.5858 Tang, Z.; Tan, Y.; Chen, H.; Wan, Y.; Curr. Med. Chem. 2023, 30, 372. [Crossref]
Crossref...
Among the drugs containing benzoxazine are apararenone (38),5959 Kintscher, U.; Edelmann, F.; Cardiovasc. Diabetol. 2023, 22, 162. [Crossref]
Crossref...
a non-steroidal mineralocorticoid receptor antagonist under development for the treatment of diabetic nephropathies and non-alcoholic steatohepatitis. In its structure, the sulfonamide group acts as a hydrogen donor/acceptor with Asn770, the fluorophenyl group is directed towards a hydrophobic pocket interacting with Met852 while the heterocyclic benzoxazine ring acts as a linker thus allowing important interactions in the mineralocorticoid receptor (MR) binding pocket.6060 Iijima, T.; Katoh, M.; Takedomi, K.; Yamamoto, Y.; Akatsuka, H.; Shirata, N.; Nishi, A.; Takakuwa, M.; Watanabe, Y.; Munakata, H.; Koyama, N.; Ikeda, T.; Iguchi, T.; Kato, H.; Kikkawa, K.; Kawaguchi, T.; J. Med. Chem. 2022, 65, 8127. [Crossref]
Crossref...
Elbasvir (39), an approved drug for the treatment of hepatitis C that inhibits NS5A, a protein essential for the replication of the hepatitis C virus (HCV) (Figure 10),6161 Keating, G. M.; Drugs 2016, 76, 617. [Crossref]
Crossref...
and etifoxine (40), a non-benzodiazepine anxiolytic agent indicated for short-term treatment of adjustment disorder. This drug produces its anxiolytic effects by activating the channels containing the β2 and β3 subunits of the GABA-A receptor complex.6262 Choi, Y. M.; Kim, K. H.; Korean J. Pain 2015, 28, 4. [Crossref]
Crossref...

Figure 10
Bioactive compounds featuring the privileged benzoxazine scaffold.

2.2. Emerging privileged structures

It is important to note that several emerging privileged scaffolds have been highlighted in scientific literature and reviewed elsewhere. These include benzoxazine derivatives,5858 Tang, Z.; Tan, Y.; Chen, H.; Wan, Y.; Curr. Med. Chem. 2023, 30, 372. [Crossref]
Crossref...
chromones,6363 Patil, V. M.; Masand, N.; Verma, S.; Masand, V.; Chem. Biol. Drug Des. 2021, 98, 943. [Crossref]
Crossref...
aporphines,6464 Zhu, R.; Jiang, G.; Tang, W.; Zhao, X.; Chen, F.; Zhang, X.; Ye, N.; Eur. J. Med. Chem. 2023, 256, 115414. [Crossref]
Crossref...
indol-3-ylglyoxylamides,6565 Barresi, E.; Robello, M.; Baglini, E.; Poggetti, V.; Viviano, M.; Salerno, S.; Da Settimo, F.; Taliani, S.; Pharmaceuticals 2023, 16, 997. [Crossref]
Crossref...
thiazolidinediones,6666 Sun, J.; Chen, Y.-H.; Liu, H.-Y.; Hondo, E.; Zhou, Y.; Wu, Y.-F.; Curr. Top. Med. Chem. 2021, 21, 2529. [Crossref]
Crossref...
1,3-diazepines,6767 Malki, Y.; Martinez, J.; Masurier, N.; Med. Res. Rev. 2021, 41, 2247. [Crossref]
Crossref...
and benzopyrans.6868 Nazeri, M. T.; Nasiriani, T.; Torabi, S.; Shaabani, A.; Org. Biomol. Chem. 2024, 22, 1102. [Crossref]
Crossref...
Additionally, some works focus on specific therapeutic classes instead of a single privileged structure, such as those targeting bromodomain,6969 Wang, R.; Wang, Y.-A.; Xu, Y.-G.; Shi, L.; Curr. Top. Med. Chem. 2022, 22, 600. [Crossref]
Crossref...
antiviral agents,44 Skoreński, M.; Sieńczyk, M.; Pharmaceuticals 2021, 14, 1164. [Crossref]
Crossref...
and others. This section will focus on emerging privileged structures that the authors believe deserve more detailed and special attention due to their increasing applicability in drug discovery: covalent warheads and privileged structures for the design of Proteolysis Targeting Chimeras (PROTACs).

2.2.1. Emerging privileged covalent warheads

The discovery and development of covalent warheads is a hot topic in medicinal chemistry, especially regarding kinase inhibition. In this context, certain covalent warheads have recently (from 2010 to 2023) emerged as privileged structures, resulting in the discovery of small molecule kinase inhibitors with improved efficacy.

From a historical perspective, kinases emerged in the 1980s as highly coveted targets for pharmacological interventions. During that time, the feasibility of developing competitive inhibitors targeting the ATP binding site, considering selectivity among different kinases, sparked substantial debates in the scientific community.7070 Shenkar, R.; Shi, C.; Austin, C.; Moore, T.; Lightle, R.; Cao, Y.; Zhang, L.; Wu, M.; Zeineddine, H. A.; Girard, R.; McDonald, D. A.; Rorrer, A.; Gallione, C.; Pytel, P.; Liao, J. K.; Marchuk, D. A.; Awad, I. A.; Stroke 2017, 48, 187. [Crossref]
Crossref...
This approach was considered a challenging barrier, given the remarkable conservation of the ATP binding site among the kinome. The year 1995 witnessed a significant milestone with the approval of fasudil (41) in Japan (Figure 11), representing the pioneering small-molecule kinase inhibitor aimed at ROCK1 and ROCK2 to alleviate cerebral vasospasm.7070 Shenkar, R.; Shi, C.; Austin, C.; Moore, T.; Lightle, R.; Cao, Y.; Zhang, L.; Wu, M.; Zeineddine, H. A.; Girard, R.; McDonald, D. A.; Rorrer, A.; Gallione, C.; Pytel, P.; Liao, J. K.; Marchuk, D. A.; Awad, I. A.; Stroke 2017, 48, 187. [Crossref]
Crossref...
Four years later, sirolimus (rapamycin) (42), a natural product, secured its position as the first kinase inhibitor to receive FDA approval.7171 Kahan, B. D.; Expert Opin. Pharmacother. 2001, 2, 1903. [Crossref]
Crossref...
Employed in organ transplant rejection prevention, sirolimus (42) (Figure 11) targets its mammalian target, mTOR,7272 Sehgal, S. N.; Transplant. Proc. 2003, 35, S7. [Crossref]
Crossref...
gaining prominence for its implication in various disease stages.7373 Li, J.; Kim, S. G.; Blenis, J.; Cell Metab. 2014, 19, 373. [Crossref]
Crossref...
In 2001, imatinib (43) (Figure 11) made history as the first synthetic kinase inhibitor to receive FDA approval.7474 Cohen, M. H.; Farrell, A.; Justice, R.; Pazdur, R.; Oncologist 2009, 14, 174. [Crossref]
Crossref...
These advancements signify remarkable achievements in understanding and manipulating kinases, paving the way for the development of more effective therapies across various medical conditions. These three non-covalent drugs represent historical landmarks that have underpinned and propelled the discovery of new protein kinase inhibitors (PKI), culminating in the current pursuit of covalent inhibitors.

Figure 11
Structures of the first approved kinase inhibitors.

Kinase inhibitors are classified into several categories (type 1, 1.5, 2, 3, 4, 5 and 6).7575 Attwood, M. M.; Fabbro, D.; Sokolov, A. V.; Knapp, S.; Schiöth, H. B.; Nat. Rev. Drug Discovery 2021, 20, 798. [Crossref]
Crossref...
,7676 Faridoon; Ng, R.; Zhang, G.; Li, J. J.; Med. Chem. Res. 2023, 32, 1039. [Crossref]
Crossref...
Kinases with the Asp-Phe-Gly (DFG)-in motif and the αC-in conformation (active state) are inhibited by type 1 inhibitors at the ATP site. At the ATP site, type 2 inhibitors inhibit kinases with the DFG-out conformation and αC-in (inactive state). Type 1.5 inhibitors are a subtype of the type-1 inhibitors, binding to an inactive kinase conformation with DFG-in conformation, typical of an active kinase, but with the αC-helix-out. Protein kinase’s small and large lobes are separated by a region where the gatekeeper residue is located. The αC helix typically adopts an “out” shape when type 3 inhibitors inhibit kinases at the allosteric site, which is situated near the ATP site. Kinases are inhibited by type 4 inhibitors at a distant allosteric location from the ATP site. Antibodies and their derivatives that target the extracellular domains of receptor tyrosine kinases (RTKs) are also classified as type 4 inhibitors.7575 Attwood, M. M.; Fabbro, D.; Sokolov, A. V.; Knapp, S.; Schiöth, H. B.; Nat. Rev. Drug Discovery 2021, 20, 798. [Crossref]
Crossref...
,7676 Faridoon; Ng, R.; Zhang, G.; Li, J. J.; Med. Chem. Res. 2023, 32, 1039. [Crossref]
Crossref...
Type 5 bivalent inhibitors target kinases at two distinct binding sites (or domains) at the same time. Type 6 inhibitors form a covalent bond with nucleophilic amino acid residues, with the option of the bond being either irreversible or reversible. In addition, employing the ubiquitin (Ub) breakdown pathway, degraders and molecular glues form another type of kinase inhibitor.7575 Attwood, M. M.; Fabbro, D.; Sokolov, A. V.; Knapp, S.; Schiöth, H. B.; Nat. Rev. Drug Discovery 2021, 20, 798. [Crossref]
Crossref...
,7676 Faridoon; Ng, R.; Zhang, G.; Li, J. J.; Med. Chem. Res. 2023, 32, 1039. [Crossref]
Crossref...

Type 6 covalent inhibitors have experienced remarkable prominence in recent years. Afatinib (44) stood out as the pioneer of this class, obtaining regulatory approval in 2013. Subsequently, ibrutinib (45) and other subsequent drugs followed the same path, as evidenced in Figure 12.7777 Zhao, Z.; Bourne, P. E.; Pharmaceuticals 2022, 15, 1322. [Crossref]
Crossref...
,7878 Roskoski Jr., R.; Pharmacol. Res. 2023, 187, 106552. [Crossref]
Crossref...
,7979 Roskoski Jr., R.; Pharmacol. Res. 2021, 165, 105422. [Crossref]
Crossref...

Figure 12
Structures of approved kinase covalent inhibitors (type 6) with their respective targets. Each covalent warhead is highlighted in blue.

Currently, when considering the most targeted molecular entities in terms of covalent inhibition, we can identify a top three ranking based on a literature survey, including scientific articles and patents from 2001 to 2023. It is observed that the protein tyrosine kinase, Bruton’s Tyrosine Kinase (BTK) leads the list, occupying the first position with approximately 300 publications, followed by the epidermal growth factor receptor HER2, which has about 210 publications, and finally, the kinase protein Akt, with approximately 150 publications.

Beyond assessing the potency and efficacy of type 6 covalent kinase protein inhibitors (CKPIs), the issue of promiscuity of these agents becomes a point of great relevance in drug discovery. This aspect gains special importance in the face of the considerable number of potential CKPIs currently available, being crucial for optimizing the selectivity and effectiveness of these therapies.8080 Xerxa, E.; Miljković, F.; Bajorath, J.; J. Med. Chem. 2023, 66, 7657. [Crossref]
Crossref...

There has been a growing interest in the development of reversible covalent inhibitors,8181 Serafimova, I. M.; Pufall, M. A.; Krishnan, S.; Duda, K.; Cohen, M. S.; Maglathlin, R. L.; McFarland, J. M.; Miller, R. M.; Frödin, M.; Taunton, J.; Nat. Chem. Biol. 2012, 8, 471. [Crossref]
Crossref...
,8282 Gabizon, R.; Shraga, A.; Gehrtz, P.; Livnah, E.; Shorer, Y.; Gurwicz, N.; Avram, L.; Unger, T.; Aharoni, H.; Albeck, S.; Brandis, A.; Shulman, Z.; Katz, B.-Z.; Herishanu, Y.; London, N.; J. Am. Chem. Soc. 2020, 142, 11734. [Crossref]
Crossref...
,8383 Fairhurst, R. A.; Knoepfel, T.; Buschmann, N.; Leblanc, C.; Mah, R.; Todorov, M.; Nimsgern, P.; Ripoche, S.; Niklaus, M.; Warin, N.; Luu, V. H.; Madoerin, M.; Wirth, J.; Graus-Porta, D.; Weiss, A.; Kiffe, M.; Wartmann, M.; Kinyamu-Akunda, J.; Sterker, D.; Stamm, C.; Adler, F.; Buhles, A.; Schadt, H.; Couttet, P.; Blank, J.; Galuba, I.; Trappe, J.; Voshol, J.; Ostermann, N.; Zou, C.; Berghausen, J.; Del Rio Espinola, A.; Jahnke, W.; Furet, P.; J. Med. Chem. 2020, 63, 12542. [Crossref]
Crossref...
,8484 Guo, W.-H.; Qi, X.; Yu, X.; Liu, Y.; Chung, C.-I.; Bai, F.; Lin, X.; Lu, D.; Wang, L.; Chen, J.; Su, L. H.; Nomie, K. J.; Li, F.; Wang, M. C.; Shu, X.; Onuchic, J. N.; Woyach, J. A.; Wang, M. L.; Wang, J.; Nat. Commun. 2020, 11, 4268. [Crossref]
Crossref...
,8585 Bandyopadhyay, A.; Gao, J.; Curr. Opin. Chem. Biol. 2016, 34, 110. [Crossref]
Crossref...
as reversibility offers the possibility of minimizing off-target reactions and avoiding permanent modification of target proteins.8585 Bandyopadhyay, A.; Gao, J.; Curr. Opin. Chem. Biol. 2016, 34, 110. [Crossref]
Crossref...
Conceptually, a reversible covalent inhibitor first establishes non-covalent bonds with its target, and subsequently, the warhead forms a covalent bond, intensifying the interaction and inhibition.8686 Reja, R. M.; Wang, W.; Lyu, Y.; Haeffner, F.; Gao, J.; J. Am. Chem. Soc. 2022, 144, 1152. [Crossref]
Crossref...
To minimize nonspecific reactions, it is desirable for the warhead to have a moderate activation rate and a fast deactivation rate (high Koff), allowing for the rapid reversal of off-target conjugations. However, such a combination would result in reduced affinity (KD, where KD = Koff/Kon). In other words, a low deactivation rate is sought to maximize the thermodynamic benefit of the warhead (Figure 13).8686 Reja, R. M.; Wang, W.; Lyu, Y.; Haeffner, F.; Gao, J.; J. Am. Chem. Soc. 2022, 144, 1152. [Crossref]
Crossref...
Furthermore, a low dissociation rate is desirable to achieve lasting inhibition, represented by a longer residence time. These paradoxical considerations highlight the need to understand and optimize the kinetics of reversible covalent warheads. Notable examples of efforts in this direction can be found in the literature, demonstrating that slow-dissociating warheads can provide highly potent and long-acting inhibitors in relevant kinases.8181 Serafimova, I. M.; Pufall, M. A.; Krishnan, S.; Duda, K.; Cohen, M. S.; Maglathlin, R. L.; McFarland, J. M.; Miller, R. M.; Frödin, M.; Taunton, J.; Nat. Chem. Biol. 2012, 8, 471. [Crossref]
Crossref...
,8787 Bradshaw, J. M.; McFarland, J. M.; Paavilainen, V. O.; Bisconte, A.; Tam, D.; Phan, V. T.; Romanov, S.; Finkle, D.; Shu, J.; Patel, V.; Ton, T.; Li, X.; Loughhead, D. G.; Nunn, P. A.; Karr, D. E.; Gerritsen, M. E.; Funk, J. O.; Owens, T. D.; Verner, E.; Brameld, K. A.; Hill, R. J.; Goldstein, D. M.; Taunton, J.; Nat. Chem. Biol. 2015, 11, 525. [Crossref]
Crossref...
,8888 Krishnan, S.; Miller, R. M.; Tian, B.; Mullins, R. D.; Jacobson, M. P.; Taunton, J.; J. Am. Chem. Soc. 2014, 136, 12624. [Crossref]
Crossref...

Figure 13
Ideal kinetics for a reversible covalent warhead, where a lower Koff compared to Kon is observed, providing a longer residence time.

In the context of covalent warheads, both reversible and irreversible, notable examples stand out, considering the interaction of these fragments with specific amino acid residues. Figure 14 illustrates some examples of warheads currently employed in the approach of covalent inhibitors, with irreversible inhibitors highlighted in the red box and the reversible ones in the green box.8989 Shindo, N.; Ojida, A.; Bioorg. Med. Chem. 2021, 47, 116386. [Crossref]
Crossref...

Figure 14
Most used warheads in covalent inhibition approach. In red, warheads classified as irreversible, and in green, warheads classified as reversible.

As described in the study by Zhao et al.,7777 Zhao, Z.; Bourne, P. E.; Pharmaceuticals 2022, 15, 1322. [Crossref]
Crossref...
ten privileged warheads were identified and highlighted, as shown in Figure 15. A relevant correlation can be established with the research conducted by Xerxa et al.,9090 Xerxa, E.; Laufkötter, O.; Bajorath, J.; Molecules 2023, 28, 5805. [Crossref]
Crossref...
where an analysis of the frequency of these warheads in covalent kinase inhibitors was undertaken. Notably, the top five prominent warheads, totaling 13,949 occurrences, were identified as acrylamide (9,861), terminal alkyne (547), acrylate (535), cyanoacrylamide (481), and aldehyde (479).9090 Xerxa, E.; Laufkötter, O.; Bajorath, J.; Molecules 2023, 28, 5805. [Crossref]
Crossref...

Figure 15
Representation of the ten warheads classified as privileged, highlighted in the study by Xerxa et al.9090 Xerxa, E.; Laufkötter, O.; Bajorath, J.; Molecules 2023, 28, 5805. [Crossref]
Crossref...

In the study conducted by Xerxa et al.,9090 Xerxa, E.; Laufkötter, O.; Bajorath, J.; Molecules 2023, 28, 5805. [Crossref]
Crossref...
a meticulous analysis was undertaken regarding various warheads, aiming to quantify the presence of these scaffolds in the scientific literature. Among the most prevalent warheads, which showed occurrences in the range of 450-550 compounds, a subset was selected for a more rigorous evaluation. To these warheads, a promiscuity degree (PD) was assigned. The more targets protein kinases (PKs) that are reported as being inhibited by a given CPKI, the higher the PD value for that CPKI. This is because the PD value is a ratio of the number of PKs that are affected by that particular CPKI. For the compounds in question, a non-promiscuity indicative value (%) was assigned.9090 Xerxa, E.; Laufkötter, O.; Bajorath, J.; Molecules 2023, 28, 5805. [Crossref]
Crossref...

As observed, the most selective compounds encompass the cyanoacrylamide (88.8%), aldehydes (86%), acrylate (75.9%), and terminal alkynes (66%) warheads. Remarkably, among the four selected, two of the most selective warheads correspond to the reversible ones, namely aldehyde and cyanoacrylamide, corroborating the indication of reduced off-target impact for reversible warheads. The chemical basis underlying reversibility is elucidated in Scheme 1, exemplified in the case of cyanoacrylamides, where the presence of the withdrawing group (nitrile) allows for a reversible thiol-Michael addition reaction (retro-Michael).9191 Abdeldayem, A.; Raouf, Y. S.; Constantinescu, S. N.; Moriggl, R.; Gunning, P. T.; Chem. Soc. Rev. 2020, 49, 2617. [Crossref]
Crossref...

Scheme 1
Mechanism of reversibility of cyanoacrylamides.

Rilzabrutinib (92, Figure 16) (PRN1008) represents a notable example of a reversible covalent inhibitor incorporating the cyanoacrylamide subunit. This compound is currently undergoing phase III clinical evaluation,9292 Kuter, D. J.; Bussel, J. B.; Ghanima, W.; Cooper, N.; Gernsheimer, T.; Lambert, M. P.; Liebman, H. A.; Tarantino, M. D.; Lee, M.; Guo, H.; Daak, A.; Ther. Adv. Hematol. 2023, 14. [Crossref]
Crossref...
as registered at the National Library of Medicine9393 The National Library of Medicine, ClinicalTrials.gov, accessed in May 2024. under the identifier NCT04562766 and on EudraCT9494 European Union Drug Regulating Authorities Clinical Trials Database, https://eudract.ema.europa.eu/, accessed in May 2024.
https://eudract.ema.europa.eu/...
under the number 2020-002063-60. Rilzabrutinib (92, Figure 16) is targeted for the treatment of immune thrombocytopenia, exerting its inhibitory activity on the BTK enzyme.8989 Shindo, N.; Ojida, A.; Bioorg. Med. Chem. 2021, 47, 116386. [Crossref]
Crossref...

Figure 16
Chemical structure of rilzabrutinib (92) (PRN1008) with the cyanoacrylamide subunit highlighted in blue.

Another example highlighting the reversible covalent effect can be observed for α-chlorofluoroacetamides (CFA).9595 Shindo, N.; Fuchida, H.; Sato, M.; Watari, K.; Shibata, T.; Kuwata, K.; Miura, C.; Okamoto, K.; Hatsuyama, Y.; Tokunaga, K.; Sakamoto, S.; Morimoto, S.; Abe, Y.; Shiroishi, M.; Caaveiro, J. M. M.; Ueda, T.; Tamura, T.; Matsunaga, N.; Nakao, T.; Koyanagi, S.; Ohdo, S.; Yamaguchi, Y.; Hamachi, I.; Ono, M.; Ojida, A.; Nat. Chem. Biol. 2019, 15, 250. [Crossref]
Crossref...
It has been demonstrated that the CFA-thiol adduct is gradually hydrolyzed under neutral aqueous conditions, reverting reversibly to an unchanged free thiol (Scheme 2). This reversible hydrolysis contributes to the effective elimination of off-target reactions with solvent-accessible cysteines. This environmentally sensitive hydrolysis may play a crucial role in achieving the high target selectivity of CFA-based covalent inhibitors.9595 Shindo, N.; Fuchida, H.; Sato, M.; Watari, K.; Shibata, T.; Kuwata, K.; Miura, C.; Okamoto, K.; Hatsuyama, Y.; Tokunaga, K.; Sakamoto, S.; Morimoto, S.; Abe, Y.; Shiroishi, M.; Caaveiro, J. M. M.; Ueda, T.; Tamura, T.; Matsunaga, N.; Nakao, T.; Koyanagi, S.; Ohdo, S.; Yamaguchi, Y.; Hamachi, I.; Ono, M.; Ojida, A.; Nat. Chem. Biol. 2019, 15, 250. [Crossref]
Crossref...
It is important to note that the reversible modification of CFA cysteine has a distinct mechanism from that of cyanoacrylamide, which is highly reactive and regenerates the warhead through retro-Michael reaction. In contrast, the initially weakly reactive CFA gains reversible reactivity through the hydrolysis of the cysteine adduct, transforming into the non-reactive glyoxamide hydrated (95, Scheme 2).8989 Shindo, N.; Ojida, A.; Bioorg. Med. Chem. 2021, 47, 116386. [Crossref]
Crossref...

Scheme 2
Mechanism of reversibility of α-chlorofluoroacetamides (CFA).

Evidence indicates that several covalent inhibitors based on CFA have demonstrated superior selectivity towards target proteins when compared to analogous inhibitors based on acrylamide.8989 Shindo, N.; Ojida, A.; Bioorg. Med. Chem. 2021, 47, 116386. [Crossref]
Crossref...

The highly promising research published by Reja et al.,8686 Reja, R. M.; Wang, W.; Lyu, Y.; Haeffner, F.; Gao, J.; J. Am. Chem. Soc. 2022, 144, 1152. [Crossref]
Crossref...
provides another illustration of the reversible covalent inhibition mechanism in action. The study makes a contribution by presenting a novel lysine conjugation chemistry in which a β-hydroxy diazaborine (101) is produced by the reversible conjugation of an RMR1 (100) warhead with lysine residues (Figure 17). The lysine conjugation caused by RMR1 (100) exhibits a substantially longer reverse reaction, with dissociation taking place over the course of hours, in comparison to iminoboronate chemistry (99), a known lysine conjugate with quick dissociation kinetics. They demonstrate that RMR1 (100) may be grafted onto a peptide structure to form strong reversible covalent inhibitors, whose effectiveness is shown against the recombinant protein as well as in living bacterial cells, using staphylococcal sortase (SrtA) as a model system. Significantly, an inhibitor containing RMR1 (100) permits SrtA to remain inhibited for hours after the inhibitor is removed, marking the first instance of the kinetic advantage of lysine-targeted reversible covalent inhibitors.8686 Reja, R. M.; Wang, W.; Lyu, Y.; Haeffner, F.; Gao, J.; J. Am. Chem. Soc. 2022, 144, 1152. [Crossref]
Crossref...
They believe that the RMR1 (100) warhead’s repertoire for producing reversible covalent inhibitors that target lysine will be significantly increased by tuning it to provide a broad range of kinetic profiles for reversible lysine conjugation.8686 Reja, R. M.; Wang, W.; Lyu, Y.; Haeffner, F.; Gao, J.; J. Am. Chem. Soc. 2022, 144, 1152. [Crossref]
Crossref...
Currently, efforts are being made in this direction.

Figure 17
Lysine-targeting covalent warheads.

There are several marketed drugs that are reversible covalent inhibitors (Figure 18). Saxagliptin (102) is a reversible covalent drug that functions as an electrophilic warhead with a nitrile group to target the Ser630 residue of dipeptidyl peptidase-4 (DPP-4).9696 Augeri, D. J.; Robl, J. A.; Betebenner, D. A.; Magnin, D. R.; Khanna, A.; Robertson, J. G.; Wang, A.; Simpkins, L. M.; Taunk, P.; Huang, Q.; Han, S.-P.; Abboa-Offei, B.; Cap, M.; Xin, L.; Tao, L.; Tozzo, E.; Welzel, G. E.; Egan, D. M.; Marcinkeviciene, J.; Chang, S. Y.; Biller, S. A.; Kirby, M. S.; Parker, R. A.; Hamann, L. G.; J. Med. Chem. 2005, 48, 5025. [Crossref]
Crossref...
The FDA gave its approval in 2009 to treat Type 2 diabetes. The FDA granted licenses for telaprevir (103) and boceprevir (104), two reversible covalent α-ketoamide-based drugs that interact with the catalytic serine residue of the hepatitis C NS3 serine protease, in 2011.9797 O’Meara, J. A.; Lemke, C. T.; Godbout, C.; Kukolj, G.; Lagacé, L.; Moreau, B.; Thibeault, D.; White, P. W.; Llinàs-Brunet, M.; J. Biol. Chem. 2013, 288, 5673. [Crossref]
Crossref...
Narlaprevir (105) is a potent second-generation inhibitor of the hepatitis C virus (HCV) NS3 protease, with a Ki of 7 nM.9898 Arasappan, A.; Bennett, F.; Bogen, S. L.; Venkatraman, S.; Blackman, M.; Chen, K. X.; Hendrata, S.; Huang, Y.; Huelgas, R. M.; Nair, L.; Padilla, A. I.; Pan, W.; Pike, R.; Pinto, P.; Ruan, S.; Sannigrahi, M.; Velazquez, F.; Vibulbhan, B.; Wu, W.; Yang, W.; Saksena, A. K.; Girijavallabhan, V.; Shih, N.-Y.; Kong, J.; Meng, T.; Jin, Y.; Wong, J.; McNamara, P.; Prongay, A.; Madison, V.; Piwinski, J. J.; Cheng, K.-C.; Morrison, R.; Malcolm, B.; Tong, X.; Ralston, R.; Njoroge, F. G.; ACS Med. Chem. Lett. 2010, 1, 64. [Crossref]
Crossref...
In 2016, it was created a reversible covalent drug based on α-ketoamide, which was authorized in Russia for the treatment of hepatitis C. The first inhibitor of the 26S proteasome in its class, bortezomib (106), targets the N-terminal threonine of the 26S proteasome with an electrophilic warhead consisting of boronic acid. The FDA gave its approval in 2003 to treat multiple myeloma.9999 Greener, B. S.; Millan, D. S. In Modern Drug Synthesis; Wiley: New Jersey, 2010, p. 99-110. [Crossref]
Crossref...
Ixazomib (107), an additional 26S proteasome inhibitor, is an orally accessible, reversible covalent inhibitor that attaches to the 20S proteasome’s β5 subunit. It was approved by the FDA in 2015 for use in combination with dexamethasone and lenalidomide to treat individuals with multiple myeloma.100100 Shirley, M.; Drugs 2016, 76, 405. [Crossref]
Crossref...
In 2019, the FDA authorized voxelotor (108) as a treatment for sickle cell disease. By forming an imine, its aldehyde warhead attaches to hemoglobin’s N-terminal valine in a reversible manner.101101 Oksenberg, D.; Dufu, K.; Patel, M. P.; Chuang, C.; Li, Z.; Xu, Q.; Silva-Garcia, A.; Zhou, C.; Hutchaleelaha, A.; Patskovska, L.; Patskovsky, Y.; Almo, S. C.; Sinha, U.; Metcalf, B. W.; Archer, D. R.; Br. J. Haematol. 2016, 175, 141. [Crossref]
Crossref...
Another reversible covalent inhibitor that targets a cysteine in the major protease (Mpro) of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is nitmatrelvir (109), which has a nitrile warhead. In 2022, the FDA authorized it for the treatment of coronavirus disease (COVID-19).102102 Greasley, S. E.; Noell, S.; Plotnikova, O.; Ferre, R.; Liu, W.; Bolanos, B.; Fennell, K.; Nicki, J.; Craig, T.; Zhu, Y.; Stewart, A. E.; Steppan, C. M.; J. Biol. Chem. 2022, 298, 101972. [Crossref]
Crossref...

Figure 18
Examples of reversible covalent drugs on the market.

PRN473 (110) (Figure 19) was studied in a clinical investigation of canine pemphigus foliaceous. In toxicological and canine clinical investigations, PRN473 (110) is effectively absorbed in dogs, whereas it is poorly absorbed orally in humans.103103 Smith, P. F.; Krishnarajah, J.; Nunn, P. A.; Hill, R. J.; Karr, D.; Tam, D.; Masjedizadeh, M.; Funk, J. O.; Gourlay, S. G.; Br J. Clin. Pharmacol. 2017, 83, 2367. [Crossref]
Crossref...
As a topical drug, PRN473 (110) has successfully finished a phase 1 clinical trial. Two selective reversible covalent inhibitors of JAK3, (111) and (112) (Figure 19), with half maximal inhibitory concentration (IC50) values of 127 and 154 pM, respectively, were discovered by Forster et al.104104 Forster, M.; Chaikuad, A.; Bauer, S. M.; Holstein, J.; Robers, M. B.; Corona, C. R.; Gehringer, M.; Pfaffenrot, E.; Ghoreschi, K.; Knapp, S.; Laufer, S. A.; Cell Chem. Biol. 2016, 23, 1335. [Crossref]
Crossref...
Compound 112 showed selectivity over JAK1 (416-fold), JAK2 (1753-fold), and tyrosine kinase 2 (TYK2) (5831-fold), whereas compound (111) demonstrated excellent selectivity of 409-, 2724-, and 3614-fold over JAK1, JAK2, and TYK2, respectively. The use of a cyanamide warhead in reversible covalent inhibitor PF-303 (113) (Figure 19) as a chemical probe to study the phenotypic of BTK inhibition in mice was also described by Benson et al.,105105 Benson, M. J.; Rodriguez, V.; von Schack, D.; Keegan, S.; Cook, T. A.; Edmonds, J.; Benoit, S.; Seth, N.; Du, S.; Messing, D.; Nickerson-Nutter, C. L.; Dunussi-Joannopoulos, K.; Rankin, A. L.; Ruzek, M.; Schnute, M. E.; Douhan, J.; J. Immunol. 2014, 193, 185. [Crossref]
Crossref...
PF-303 (113) is an oral bioavailable and strong inhibitor of BTK (IC50 = 0.64 nM). The 2-formyl tetrahydronaphthyridine urea series underwent further, thorough optimization, which produced roblitinib (114) (Figure 19), a therapeutic candidate currently undergoing phase III clinical studies.106106 Weiss, A.; Adler, F.; Buhles, A.; Stamm, C.; Fairhurst, R. A.; Kiffe, M.; Sterker, D.; Centeleghe, M.; Wartmann, M.; Kinyamu-Akunda, J.; Schadt, H. S.; Couttet, P.; Wolf, A.; Wang, Y.; Barzaghi-Rinaudo, P.; Murakami, M.; Kauffmann, A.; Knoepfel, T.; Buschmann, N.; Leblanc, C.; Mah, R.; Furet, P.; Blank, J.; Hofmann, F.; Sellers, W. R.; Graus Porta, D.; Mol. Cancer Ther. 2019, 18, 2194. [Crossref]
Crossref...
Roblitinib (114) showed promising pharmacokinetic properties and signs of fibroblast growth factor receptor 4 (FGFR4) inhibition in a phase 1-2 research. Hepatocellular carcinoma patients showed clinical effectiveness. NS-062 (115) demonstrated stronger target selectivity for EGFR than the corresponding Michael acceptors in a wide range of concentrations (0.1-10 μM) in cells. Oral treatment of NS-062 (115) greatly inhibited tumor growth in a mouse xenograft model (Figure 19).9595 Shindo, N.; Fuchida, H.; Sato, M.; Watari, K.; Shibata, T.; Kuwata, K.; Miura, C.; Okamoto, K.; Hatsuyama, Y.; Tokunaga, K.; Sakamoto, S.; Morimoto, S.; Abe, Y.; Shiroishi, M.; Caaveiro, J. M. M.; Ueda, T.; Tamura, T.; Matsunaga, N.; Nakao, T.; Koyanagi, S.; Ohdo, S.; Yamaguchi, Y.; Hamachi, I.; Ono, M.; Ojida, A.; Nat. Chem. Biol. 2019, 15, 250. [Crossref]
Crossref...

Figure 19
Examples of reversible covalent kinase inhibitors under development.

2.2.2. Emerging privileged structures for the design of proteolysis targeting chimeras (PROTACs)

The pioneering work published by Sakamoto et al.107107 Sakamoto, K. M.; Kim, K. B.; Kumagai, A.; Mercurio, F.; Crews, C. M.; Deshaies, R. J.; Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 8554. [Crossref]
Crossref...
reported the first study on proteolysis targeting chimeras (PROTACs) in 2001. PROTACs are heterobifunctional molecules that have three components: the protein-of-interest (POI) binding moiety, a linker, and E3 ubiquitin ligase binding moiety (Figure 20).108108 Schneekloth, A. R.; Pucheault, M.; Tae, H. S.; Crews, C. M.; Bioorg. Med. Chem. Lett. 2008, 18, 5904. [Crossref]
Crossref...
,109109 Yao, T.; Xiao, H.; Wang, H.; Xu, X.; Int. J. Mol. Sci. 2022, 23, 10328. [Crossref]
Crossref...
,110110 Liu, Z.; Hu, M.; Yang, Y.; Du, C.; Zhou, H.; Liu, C.; Chen, Y.; Fan, L.; Ma, H.; Gong, Y.; Xie, Y.; Mol. Biomed. 2022, 3, 46. [Crossref]
Crossref...
A POI-PROTAC-E3 ligase ternary complex can be formed when the PROTAC molecule simultaneously bind to both the target protein and E3 ligase.111111 Buckley, D. L.; Crews, C. M.; Angew. Chem., Int. Ed. 2014, 53, 2312. [Crossref]
Crossref...
,112112 Gadd, M. S.; Testa, A.; Lucas, X.; Chan, K.-H.; Chen, W.; Lamont, D. J.; Zengerle, M.; Ciulli, A.; Nat. Chem. Biol. 2017, 13, 514. [Crossref]
Crossref...
Protein degradation by proteases occurs once the target protein is polyubiquitinated due to the hijacking of the ubiquitin-protease system (UPS). In eukaryotic cells, the UPS is the principal mechanism for maintaining protein homeostasis eliminating faulty and damaged proteins.113113 Finley, D.; Annu. Rev. Biochem. 2009, 78, 477. [Crossref]
Crossref...
,114114 Hipp, M. S.; Kasturi, P.; Hartl, F. U.; Nat. Rev. Mol. Cell Biol. 2019, 20, 421. [Crossref]
Crossref...
Proteins are broken down by the UPS system by substrate-specific ubiquitination and recognition. Three enzymes are involved in the continuous process of ubiquitination: substrate-specific ligases (E3), ubiquitin-activating enzymes (E1), and ubiquitin-conjugating enzymes (E2).115115 Kliza, K.; Husnjak, K.; Front. Mol. Biosci. 2020, 7, 21. [Crossref]
Crossref...
,116116 Konstantinidou, M.; Li, J.; Zhang, B.; Wang, Z.; Shaabani, S.; Ter Brake, F.; Essa, K.; Dömling, A.; Expert Opin. Drug Discovery 2019, 14, 1255. [Crossref]
Crossref...
,117117 Cromm, P. M.; Crews, C. M.; Cell Chem. Biol. 2017, 24, 1181. [Crossref]
Crossref...
,118118 Hu, Z.; Crews, C. M.; ChemBioChem. 2022, 23. [Crossref]
Crossref...
By creating a ubiquitin-E1 thioester link, E1 binds free ubiquitin (Ub) in an ATP-dependent manner. Later, E1 uses trans-thioesterification to move the activated Ub to E2.119119 Komander, D.; Rape, M.; Annu. Rev. Biochem. 2012, 81, 203. [Crossref]
Crossref...
Lastly, E3 ligase recruits the Ub-tagged E2 and target protein to help with ubiquitin labeling on target proteins.120120 Chen, Y.; Jin, J.; Acta Biochim. Biophys. Sin. 2020, 52, 776. [Crossref]
Crossref...
These ubiquitination processes can be repeated to produce target proteins with poly-ubiquitin chains attached, which point the designated protein toward the 26S proteasome for breakdown.121121 Burslem, G. M.; Crews, C. M.; Cell 2020, 181, 102. [Crossref]
Crossref...
POI and E3 ligase are recruited by PROTACs concurrently, resulting in their spatial closeness. To eliminate POIs from cells, PROTACs mimic the precise substrate identification of E3 ligase and take use of the intracellular protein degradation pathway.122122 Salami, J.; Crews, C. M.; Science 2017, 355, 1163. [Crossref]
Crossref...

Figure 20
PROTAC-induced degradation of target proteins.

The most relevant E3 enzymes for the PROTAC approach are outlined, namely Creblon (CRBN), von Hippel-Lindau (VHL), inhibitors of apoptosis proteins (IAP), and mouse double minute 2 (MDM2), with a special focus on the extensively studied CRBN. CRBN is a 442-amino acid protein that forms a Cullin-4-RING E3 ubiquitin ligase (CRL4) complex and interacts with the adaptor protein damaged DNA-binding protein 1 (DDB1).123123 Ito, T.; Ando, H.; Handa, H.; Cell. Mol. Life Sci. 2011, 68, 1569. [Crossref]
Crossref...
,124124 Chamberlain, P. P.; Lopez-Girona, A.; Miller, K.; Carmel, G.; Pagarigan, B.; Chie-Leon, B.; Rychak, E.; Corral, L. G.; Ren, Y. J.; Wang, M.; Riley, M.; Delker, S. L.; Ito, T.; Ando, H.; Mori, T.; Hirano, Y.; Handa, H.; Hakoshima, T.; Daniel, T. O.; Cathers, B. E.; Nat. Struct. Mol. Biol. 2014, 21, 803. [Crossref]
Crossref...
Within the CRL4 complex, CRBN acts as a substrate-specificity receptor.124124 Chamberlain, P. P.; Lopez-Girona, A.; Miller, K.; Carmel, G.; Pagarigan, B.; Chie-Leon, B.; Rychak, E.; Corral, L. G.; Ren, Y. J.; Wang, M.; Riley, M.; Delker, S. L.; Ito, T.; Ando, H.; Mori, T.; Hirano, Y.; Handa, H.; Hakoshima, T.; Daniel, T. O.; Cathers, B. E.; Nat. Struct. Mol. Biol. 2014, 21, 803. [Crossref]
Crossref...
Among the known ligands for CRBN, thalidomide (116, Figure 21) as an immunomodulatory drug (IMiD) and other IMiD-derived immunomodulatory drugs stand out. Upon binding of IMiDs to CRBN, the E3 ubiquitin ligase activity of CRBN is reconfigured.124124 Chamberlain, P. P.; Lopez-Girona, A.; Miller, K.; Carmel, G.; Pagarigan, B.; Chie-Leon, B.; Rychak, E.; Corral, L. G.; Ren, Y. J.; Wang, M.; Riley, M.; Delker, S. L.; Ito, T.; Ando, H.; Mori, T.; Hirano, Y.; Handa, H.; Hakoshima, T.; Daniel, T. O.; Cathers, B. E.; Nat. Struct. Mol. Biol. 2014, 21, 803. [Crossref]
Crossref...
,125125 Zhu, Y. X.; Braggio, E.; Shi, C.-X.; Bruins, L. A.; Schmidt, J. E.; Van Wier, S.; Chang, X.-B.; Bjorklund, C. C.; Fonseca, R.; Bergsagel, P. L.; Orlowski, R. Z.; Stewart, A. K.; Blood 2011, 118, 4771. [Crossref]
Crossref...
,126126 Krönke, J.; Udeshi, N. D.; Narla, A.; Grauman, P.; Hurst, S. N.; McConkey, M.; Svinkina, T.; Heckl, D.; Comer, E.; Li, X.; Ciarlo, C.; Hartman, E.; Munshi, N.; Schenone, M.; Schreiber, S. L.; Carr, S. A.; Ebert, B. L.; Science 2014, 343, 301. [Crossref]
Crossref...
,127127 Krönke, J.; Fink, E. C.; Hollenbach, P. W.; MacBeth, K. J.; Hurst, S. N.; Udeshi, N. D.; Chamberlain, P. P.; Mani, D. R.; Man, H. W.; Gandhi, A. K.; Svinkina, T.; Schneider, R. K.; McConkey, M.; Järås, M.; Griffiths, E.; Wetzler, M.; Bullinger, L.; Cathers, B. E.; Carr, S. A.; Chopra, R.; Ebert, B. L.; Nature 2015, 523, 183. [Crossref]
Crossref...
As a result, there is an increase in the recruitment of the transcription factors Ikaros (IKZF1) and Aiolos (IKZF3), leading to their subsequent ubiquitination and proteasomal degradation. This interaction and its consequences are responsible for the observed antiproliferative effects of thalidomide (116, Figure 21), pomalidomide (125, Figure 21), and lenalidomide (126, Figure 21) in multiple myeloma.124124 Chamberlain, P. P.; Lopez-Girona, A.; Miller, K.; Carmel, G.; Pagarigan, B.; Chie-Leon, B.; Rychak, E.; Corral, L. G.; Ren, Y. J.; Wang, M.; Riley, M.; Delker, S. L.; Ito, T.; Ando, H.; Mori, T.; Hirano, Y.; Handa, H.; Hakoshima, T.; Daniel, T. O.; Cathers, B. E.; Nat. Struct. Mol. Biol. 2014, 21, 803. [Crossref]
Crossref...
,126126 Krönke, J.; Udeshi, N. D.; Narla, A.; Grauman, P.; Hurst, S. N.; McConkey, M.; Svinkina, T.; Heckl, D.; Comer, E.; Li, X.; Ciarlo, C.; Hartman, E.; Munshi, N.; Schenone, M.; Schreiber, S. L.; Carr, S. A.; Ebert, B. L.; Science 2014, 343, 301. [Crossref]
Crossref...
,127127 Krönke, J.; Fink, E. C.; Hollenbach, P. W.; MacBeth, K. J.; Hurst, S. N.; Udeshi, N. D.; Chamberlain, P. P.; Mani, D. R.; Man, H. W.; Gandhi, A. K.; Svinkina, T.; Schneider, R. K.; McConkey, M.; Järås, M.; Griffiths, E.; Wetzler, M.; Bullinger, L.; Cathers, B. E.; Carr, S. A.; Chopra, R.; Ebert, B. L.; Nature 2015, 523, 183. [Crossref]
Crossref...

Figure 21
Commonly utilized talidomide-derived CRBN ligands and possible linker attachment styles. (117-118) Pomalidomide derivatives; (119-120) 4-hydroxythalidomide derivatives; (121) alkyl type attachment to thalidomide; (122-124) lenalidomide derivatives.

So far, CRBN has demonstrated success as the E3 ligase in PROTACs targeting over 30 distinct proteins, spanning those involved in various cancers128128 Sun, X.; Gao, H.; Yang, Y.; He, M.; Wu, Y.; Song, Y.; Tong, Y.; Rao, Y.; Signal Transduction Targeted Ther. 2019, 4, 64. [Crossref]
Crossref...
and immunological disorders,129129 Bassi, Z. I.; Fillmore, M. C.; Miah, A. H.; Chapman, T. D.; Maller, C.; Roberts, E. J.; Davis, L. C.; Lewis, D. E.; Galwey, N. W.; Waddington, K. E.; Parravicini, V.; Macmillan-Jones, A. L.; Gongora, C.; Humphreys, P. G.; Churcher, I.; Prinjha, R. K.; Tough, D. F.; ACS Chem. Biol. 2018, 13, 2862. [Crossref]
Crossref...
to the protein associated with the neurodegenerative disease Tau130130 Silva, M. C.; Ferguson, F. M.; Cai, Q.; Donovan, K. A.; Nandi, G.; Patnaik, D.; Zhang, T.; Huang, H.-T.; Lucente, D. E.; Dickerson, B. C.; Mitchison, T. J.; Fischer, E. S.; Gray, N. S.; Haggarty, S. J.; eLife 2019, 8, e45457. [Crossref]
Crossref...
and even the NS3 protein of the hepatitis C virus.131131 de Wispelaere, M.; Du, G.; Donovan, K. A.; Zhang, T.; Eleuteri, N. A.; Yuan, J. C.; Kalabathula, J.; Nowak, R. P.; Fischer, E. S.; Gray, N. S.; Yang, P. L.; Nat. Commun. 2019, 10, 3468. [Crossref]
Crossref...

The collection of CRBN ligands with different linker attachment options is presented in Figure 21. The majority of CRBN-targeting PROTACs employ derivatives of pomalidomide (Figure 21, 117, 118), 4-hydroxythalidomide (Figure 21, 119, 120), alkyl-connected thalidomide derivatives (Figure 21, 121), or lenalidomide (Figure 21, 122-124). However, alternatives are possible, including examples with substitution at position 5 of the phthalimide fragment.132132 Bricelj, A.; Steinebach, C.; Kuchta, R.; Gütschow, M.; Sosič, I.; Front. Chem. 2021, 9, 707317. [Crossref]
Crossref...

Based on data extracted from PROTAC-DB in the work of Weng et al.,133133 Weng, G.; Shen, C.; Cao, D.; Gao, J.; Dong, X.; He, Q.; Yang, B.; Li, D.; Wu, J.; Hou, T.; Nucleic. Acids. Res. 2021, 49, D1381. [Crossref]
Crossref...
a statistical analysis was conducted to assess the frequency of various CRBN ligands and linker attachment options used in PROTAC compounds (Figure 22). An overwhelming majority of PROTACs incorporated N-alkylated pomalidomide (133) as the E3 ligase ligand, while acylated pomalidomide (129) was represented with a comparable frequency to 4-hydroxythalidomide derivatives (127). Interestingly, the 5-amino derivative (132) was employed in approximately 5% of PROTACs. Lenalidomide analogs, especially 4-acylated (130) derivatives and alkyl-connected lenalidomide derivatives (131), showed similar frequencies at 7 and 6%, respectively.132132 Bricelj, A.; Steinebach, C.; Kuchta, R.; Gütschow, M.; Sosič, I.; Front. Chem. 2021, 9, 707317. [Crossref]
Crossref...

Figure 22
Frequency of CRBN ligands used in PROTAC compounds.

Protein chemical degraders, exemplified by PROTACs have emerged as a robust strategy to disrupt protein activity in recent years.134134 Chamberlain, P. P.; Hamann, L. G.; Nat. Chem. Biol. 2019, 15, 937. [Crossref]
Crossref...
Notable efforts have been dedicated to the use of covalent inhibitors based on acrylamide as the protein-target binding subunit in PROTACs.135135 Gabizon, R.; London, N.; Curr. Opin. Chem. Biol. 2021, 62, 24. [Crossref]
Crossref...
Additionally, investigations have been conducted on the application of BTK inhibitors based on cyanacrylamide for PROTAC, as depicted in Figure 23.8282 Gabizon, R.; Shraga, A.; Gehrtz, P.; Livnah, E.; Shorer, Y.; Gurwicz, N.; Avram, L.; Unger, T.; Aharoni, H.; Albeck, S.; Brandis, A.; Shulman, Z.; Katz, B.-Z.; Herishanu, Y.; London, N.; J. Am. Chem. Soc. 2020, 142, 11734. [Crossref]
Crossref...
,8484 Guo, W.-H.; Qi, X.; Yu, X.; Liu, Y.; Chung, C.-I.; Bai, F.; Lin, X.; Lu, D.; Wang, L.; Chen, J.; Su, L. H.; Nomie, K. J.; Li, F.; Wang, M. C.; Shu, X.; Onuchic, J. N.; Woyach, J. A.; Wang, M. L.; Wang, J.; Nat. Commun. 2020, 11, 4268. [Crossref]
Crossref...

Figure 23
(a) Structure of the compound RC-1 (134). (b) Structures from the study by Gabizon et al.,8282 Gabizon, R.; Shraga, A.; Gehrtz, P.; Livnah, E.; Shorer, Y.; Gurwicz, N.; Avram, L.; Unger, T.; Aharoni, H.; Albeck, S.; Brandis, A.; Shulman, Z.; Katz, B.-Z.; Herishanu, Y.; London, N.; J. Am. Chem. Soc. 2020, 142, 11734. [Crossref]
Crossref...
with emphasis on the compound RC-3 (135). (c) Mechanism of action of a reversible covalent PROTAC.

The work by Guo et al.,8484 Guo, W.-H.; Qi, X.; Yu, X.; Liu, Y.; Chung, C.-I.; Bai, F.; Lin, X.; Lu, D.; Wang, L.; Chen, J.; Su, L. H.; Nomie, K. J.; Li, F.; Wang, M. C.; Shu, X.; Onuchic, J. N.; Woyach, J. A.; Wang, M. L.; Wang, J.; Nat. Commun. 2020, 11, 4268. [Crossref]
Crossref...
reports that reversible covalent chemistry based on cyanacrylamide can substantially enhance intracellular accumulation and target engagement in PROTACs. The compound RC-1 (134) (Figure 23a) was developed, with an IC50 of 33 nM for BTK and 250 nM for CRBN, providing a degradation rate of BTK of 81% at 200 nM. RC-1 (134) acts as a reversible covalent inhibitor of BTK in PROTACs, exhibiting high target occupancy and functioning as a kinase inhibitor. This dual functionality, combining inhibitory and degradative properties, constitutes a distinct mechanism of action for PROTACs. The relevance of this reversible covalent strategy is emphasized, as it proves to be generalizable for optimizing other PROTACs,8484 Guo, W.-H.; Qi, X.; Yu, X.; Liu, Y.; Chung, C.-I.; Bai, F.; Lin, X.; Lu, D.; Wang, L.; Chen, J.; Su, L. H.; Nomie, K. J.; Li, F.; Wang, M. C.; Shu, X.; Onuchic, J. N.; Woyach, J. A.; Wang, M. L.; Wang, J.; Nat. Commun. 2020, 11, 4268. [Crossref]
Crossref...
thus opening a pathway to enhance the efficacy of these protein degraders.

In the study conducted by Gabizon et al.,8282 Gabizon, R.; Shraga, A.; Gehrtz, P.; Livnah, E.; Shorer, Y.; Gurwicz, N.; Avram, L.; Unger, T.; Aharoni, H.; Albeck, S.; Brandis, A.; Shulman, Z.; Katz, B.-Z.; Herishanu, Y.; London, N.; J. Am. Chem. Soc. 2020, 142, 11734. [Crossref]
Crossref...
the presented data indicate that a significant portion of the degradation induced by irreversible covalent PROTACs is driven by reversible binding before the formation of the covalent bond, while reversible covalent PROTACs predominantly lead to degradation through covalent engagement. These PROTACs demonstrated more pronounced inhibition of B-cell activation compared to ibrutinib (45) and exhibited effective degradation of BTK in primary chronic lymphocytic leukemia cells derived from patients. The most potent reversible covalent PROTAC, RC-3 (135), showed greater selectivity towards BTK compared to covalent, non-covalent, and irreversible PROTACs. These results suggest the potential for developing covalent PROTACs for a wide range of challenging targets.

3. Promiscuous Structures

3.1. Pan Assay Interference Compounds (PAINS)

Pan Assay Interference Compounds (PAINS) are a major challenge for virtual (VS) and high-throughput screening (HTS) in drug discovery and development. PAINS are chemical compounds that often give false positive results in screening campaigns, because they interfere with the assay rather than binding specifically to the target of interest, resulting in the identification of frequent hitters, i.e., promiscuous compounds.99 Baell, J.; Walters, M. A.; Nature 2014, 513, 481. [Crossref]
Crossref...
,1313 Hu, Y.; Bajorath, J.; J. Chem. Inf. Model. 2010, 50, 2112. [Crossref]
Crossref...
The mechanism behind this promiscuity can be related to aggregate formation136136 McGovern, S. L.; Caselli, E.; Grigorieff, N.; Shoichet, B. K.; J. Med. Chem. 2002, 45, 1712. [Crossref]
Crossref...
,137137 Ryan, A. J.; Gray, N. M.; Lowe, P. N.; Chung, C.; J. Med. Chem. 2003, 46, 3448. [Crossref]
Crossref...
chemical instability that results in indiscriminately reactivity with proteins138138 Jöst, C.; Nitsche, C.; Scholz, T.; Roux, L.; Klein, C. D.; J. Med. Chem. 2014, 57, 7590. [Crossref]
Crossref...
,139139 Metz, J. T.; Huth, J. R.; Hajduk, P. J.; J. Comput.-Aided Mol. Des. 2007, 21, 139. [Crossref]
Crossref...
and those that interfere with assay signaling, for example, by being fluorescent, strongly colored or sequester metal ions that are essential to protein’s function.99 Baell, J.; Walters, M. A.; Nature 2014, 513, 481. [Crossref]
Crossref...
,140140 Baell, J. B.; Holloway, G. A.; J. Med. Chem. 2010, 53, 2719. [Crossref]
Crossref...
Despite three possible mechanisms, most PAINS are covalently reactive compounds.99 Baell, J.; Walters, M. A.; Nature 2014, 513, 481. [Crossref]
Crossref...
PAINS comprise approximately 400 structural classes. However, a noteworthy observation reveals that over half of the PAINS present in a typical library can be categorized into a mere 16 easily recognizable classes.99 Baell, J.; Walters, M. A.; Nature 2014, 513, 481. [Crossref]
Crossref...
,140140 Baell, J. B.; Holloway, G. A.; J. Med. Chem. 2010, 53, 2719. [Crossref]
Crossref...
This concentration of PAINS within a limited set of distinct categories underscores the importance of vigilant screening and identification to mitigate potential assay interference in scientific research and drug discovery endeavors.

Some examples of PAINS are toxoflavin, isothiazolones, hydroxyphenyl hydrazones, curcumin, phenol-sulfonamides, rhodanines, enones, quinones, and catechols.1010 David, L.; Walsh, J.; Sturm, N.; Feierberg, I.; Nissink, J. W. M.; Chen, H.; Bajorath, J.; Engkvist, O.; ChemMedChem. 2019, 14, 1795. [Crossref]
Crossref...
,140140 Baell, J. B.; Holloway, G. A.; J. Med. Chem. 2010, 53, 2719. [Crossref]
Crossref...
To identify and remove PAINS from screening library, researchers can use methods, such as: applying substructure filters that flag compounds with known PAINS functional groups140140 Baell, J. B.; Holloway, G. A.; J. Med. Chem. 2010, 53, 2719. [Crossref]
Crossref...
and perform a validation of the identified hits.141141 Campos-Olivas, R.; Curr. Top. Med. Chem. 2011, 11, 43. [Crossref]
Crossref...

By identifying and removing PAINS from screening library, researchers can avoid wasting time and resources on compounds that cannot be developed into drugs and focus on those with genuine biological activity.140140 Baell, J. B.; Holloway, G. A.; J. Med. Chem. 2010, 53, 2719. [Crossref]
Crossref...
Despite that, early filtering of screen libraries, without a deeper investigation of the chemical and biochemical effects of certain frameworks in the structure of a drug candidate, has also been questioned.142142 Capuzzi, S. J.; Muratov, E. N.; Tropsha, A.; J. Chem. Inf. Model. 2017, 57, 417. [Crossref]
Crossref...
,143143 Senger, M. R.; Fraga, C. A. M.; Dantas, R. F.; Silva, F. P.; Drug Discovery Today 2016, 21, 868. [Crossref]
Crossref...

3.2. The thin line between promiscuity and covalent inhibition

Promiscuity and covalent inhibition in medicinal chemistry are two important concepts that relate to the selectivity and efficacy of covalent drugs. Covalent drugs are small molecules that form a covalent bond with their target proteins, resulting in irreversible or pseudo-irreversible inhibition and, consequently, leading to increased residence time.144144 Copeland, R. A.; Nat. Rev. Drug Discovery 2016, 15, 87. [Crossref]
Crossref...
,145145 Bernetti, M.; Masetti, M.; Rocchia, W.; Cavalli, A.; Annu. Rev. Phys. Chem. 2019, 70, 143. [Crossref]
Crossref...
,146146 Lee, K. S. S.; Yang, J.; Niu, J.; Ng, C. J.; Wagner, K. M.; Dong, H.; Kodani, S. D.; Wan, D.; Morisseau, C.; Hammock, B. D.; ACS Cent. Sci. 2019, 5, 1614. [Crossref]
Crossref...
Covalent drugs have several advantages over conventional reversible inhibitors, such as increased duration of action, reduced pharmacokinetic sensitivity, and the potential to target shallow binding sites.147147 Sutanto, F.; Konstantinidou, M.; Dömling, A.; RSC Med. Chem. 2020, 11, 876. [Crossref]
Crossref...
However, covalent drugs also pose challenges in terms of off-target toxicity and immunogenicity risks, which depend on the compound selectivity and reactivity of the electrophilic warheads that mediate the covalent bond formation. In this scenario, promiscuity refers to the tendency of a covalent drug to indiscriminately react with multiple proteins, which may lead to undesirable side effects. Off-target toxicity may cause idiosyncratic reactions, allergic responses, or organ damage. Therefore, careful design and optimization of covalent drugs are required to balance promiscuity and selectivity, and to minimize off-target toxicity. As previously discussed, some examples of successful covalent inhibitors are afatinib (44),148148 Dai, Y.; Hartandi, K.; Ji, Z.; Ahmed, A. A.; Albert, D. H.; Bauch, J. L.; Bouska, J. J.; Bousquet, P. F.; Cunha, G. A.; Glaser, K. B.; Harris, C. M.; Hickman, D.; Guo, J.; Li, J.; Marcotte, P. A.; Marsh, K. C.; Moskey, M. D.; Martin, R. L.; Olson, A. M.; Osterling, D. J.; Pease, L. J.; Soni, N. B.; Stewart, K. D.; Stoll, V. S.; Tapang, P.; Reuter, D. R.; Davidsen, S. K.; Michaelides, M. R.; J. Med. Chem. 2007, 50, 1584. [Crossref]
Crossref...
ibrutinib (45),149149 Woyach, J. A.; Furman, R. R.; Liu, T.-M.; Ozer, H. G.; Zapatka, M.; Ruppert, A. S.; Xue, L.; Li, D. H.-H.; Steggerda, S. M.; Versele, M.; Dave, S. S.; Zhang, J.; Yilmaz, A. S.; Jaglowski, S. M.; Blum, K. A.; Lozanski, A.; Lozanski, G.; James, D. F.; Barrientos, J. C.; Lichter, P.; Stilgenbauer, S.; Buggy, J. J.; Chang, B. Y.; Johnson, A. J.; Byrd, J. C.; New Engl. J. Med. 2014, 370, 2286. [Crossref]
Crossref...
and osimertinib (46)150150 Butterworth, S.; Cross, D. A. E.; Finlay, M. R. V.; Ward, R. A.; Waring, M. J.; Med. Chem. Comm. 2017, 8, 820. [Crossref]
Crossref...
(Figure 12). These covalent inhibitors have shown clinical benefits for patients with cancer by irreversibly inhibiting their target kinases.151151 Boike, L.; Henning, N. J.; Nomura, D. K.; Nat. Rev. Drug Discovery 2022, 21, 881. [Crossref]
Crossref...

In essence, the implementation of optimization steps is crucial for transitioning a given structure from the PAINS condition, primarily associated with compounds exhibiting fragment-like features, to more intricate structures that demand effective recognition by their respective molecular targets. This involves strategically positioning the covalent reactive subunit towards the specific amino acid within the binding site. Moreover, exploring reversible covalent warheads opens up vast opportunities for the development of safe covalent drugs.

3.3. The thin line between promiscuity and multitarget actions

While promiscuity may pose challenges in terms of selectivity and off-target effects, which are linked to toxicity, it also opens new avenues for drug repurposing and the exploration of polypharmacology. Polypharmacology refers to the phenomenon of one drug, with multitarget profile, or the combination of different drugs that results in multiple biological effects by modulating more than one pharmacological target, which may enhance its therapeutic efficacy or reduce its adverse effects.152152 Reddy, A. S.; Zhang, S.; Expert. Rev. Clin. Pharmacol. 2013, 6, 41. [Crossref]
Crossref...
The difference between polypharmacology and promiscuity is that polypharmacology is usually intentional and/or desirable, while promiscuity is usually accidental and undesirable.

Promiscuity can have both positive and negative implications for drug development and therapeutic applications.1111 Blaschke, T.; Miljković, F.; Bajorath, J.; ACS Omega 2019, 4, 6883. [Crossref]
Crossref...
,153153 Zhang, W.; Bai, Y.; Wang, Y.; Xiao, W.; Curr. Pharm. Des. 2016, 22, 3171. [Crossref]
Crossref...
On one hand, promiscuity can enhance drug efficacy by modulating multiple pathways involved in complex diseases, such as cancer, inflammation, or neurodegeneration. For example, aspirin is a drug that has anti-inflammatory, analgesic, antipyretic, and antiplatelet effects by inhibiting cyclooxygenases and has been characterized as a promiscuous drug by modulating other 23 putative targets,1212 Mencher, S. K.; Wang, L. G.; BMC Clin. Pharmacol. 2005, 5, 3. [Crossref]
Crossref...
,154154 Dai, S.-X.; Li, W.-X.; Li, G.-H.; Huang, J.-F.; Peer. J. 2016, 4, e1791. [Crossref]
Crossref...
which opened avenues for drug repositioning for cancer prevention.155155 Alfonso, L.; Ai, G.; Spitale, R. C.; Bhat, G. J.; Br. J. Cancer 2014, 111, 61. [Crossref]
Crossref...
On the other hand, promiscuity can cause adverse drug reactions by binding off-targets that are unrelated to the desired pharmacological effect.66 Scheiber, J.; Jenkins, J. L.; Sukuru, S. C. K.; Bender, A.; Mikhailov, D.; Milik, M.; Azzaoui, K.; Whitebread, S.; Hamon, J.; Urban, L.; Glick, M.; Davies, J. W.; J. Med. Chem. 2009, 52, 3103. [Crossref]
Crossref...
For example, thalidomide (116) is a promiscuous drug that binds to various proteins and has teratogenic effects in addition to its immunomodulatory and antiangiogenic activities.156156 Ito, T.; Ando, H.; Suzuki, T.; Ogura, T.; Hotta, K.; Imamura, Y.; Yamaguchi, Y.; Handa, H.; Science 2010, 327, 1345. [Crossref]
Crossref...
On the other hand, delving into the promiscuity of thalidomide (116) and its analogs has paved the way for the development and discovery of PROTACs, as previously discussed.

The degree of promiscuity of a compound depends on several factors, such as the chemical structure, the binding mode, the target similarity, and the assay conditions. Promiscuity can be quantified by various metrics, such as the number of targets, the target diversity, or the promiscuity index (PI).157157 Hu, Y.; Bajorath, J.; AAPS J. 2013, 15, 104. [Crossref]
Crossref...
The PI is a valuable metric in drug discovery used to quantify the promiscuity of a compound by measuring its interactions with multiple targets relative to the total number of targets tested. A higher PI indicates greater promiscuity, meaning the compound has the potential to interact with a wide range of targets. The PI is calculated using the formula: PI = n/N, where: n is the number of targets a compound interacts with and N is the total number of targets tested. For example, if a compound interacts with 5 out of 50 tested targets, its PI would be 0.1 (5/50). The PI provides valuable insights into the selectivity and potential off-target effects of a compound. Compounds with low PI values are more selective and have a narrower range of interactions, making them potentially safer and more suitable for further development. On the other hand, compounds with high PI values may have a higher risk of off-target effects but could also offer broader therapeutic potential. Studies have utilized the PI to assess and compare the promiscuity of compounds in drug discovery.158158 Jasial, S.; Hu, Y.; Bajorath, J.; PLoS One 2016, 11, e0153873. [Crossref]
Crossref...
For example, a study by Lounkine et al.,159159 Lounkine, E.; Keiser, M. J.; Whitebread, S.; Mikhailov, D.; Hamon, J.; Jenkins, J. L.; Lavan, P.; Weber, E.; Doak, A. K.; Côté, S.; Shoichet, B. K.; Urban, L.; Nature 2012, 486, 361. [Crossref]
Crossref...
used the PI to analyze the promiscuity of approved drugs and experimental compounds, providing valuable insights into their potential off-target effects. Another study by Hu et al.,157157 Hu, Y.; Bajorath, J.; AAPS J. 2013, 15, 104. [Crossref]
Crossref...
used the PI to evaluate the promiscuity of kinase inhibitors, highlighting the importance of considering off-target effects in drug design. Overall, the PI is a valuable tool in drug discovery for assessing compound promiscuity and guiding the design of safer and more selective drugs.

Promiscuity can also be visualized by network representations, such as bipartite graphs or heat maps.160160 Yıldırım, M. A.; Goh, K.-I.; Cusick, M. E.; Barabási, A.-L.; Vidal, M.; Nat. Biotechnol. 2007, 25, 1119. [Crossref]
Crossref...
Promiscuity can be predicted by computational methods, such as ligand-based or target-based approaches, that exploit structural or biological information of compounds and targets.161161 Ezzat, A.; Wu, M.; Li, X.-L.; Kwoh, C.-K.; Briefings Bioinf. 2019, 20, 1337. [Crossref]
Crossref...

Promiscuity is an emerging concept in drug discovery that challenges the traditional one drug-one target paradigm and offers new challenges and opportunities for medicinal chemistry and pharmacology. Understanding and exploiting promiscuity can facilitate the design of multitarget drugs, the discovery of novel drug-target interactions, and the optimization of drug safety profiles.

4. Conclusion and Perspectives

The intricate relationship between promiscuity and privileged structures in medicinal chemistry presents both challenges and opportunities. Promiscuity, the ability of a molecule to interact with multiple targets, has been viewed traditionally as a hurdle due to potential off-target effects and lack of selectivity. However, it also offers a gateway to exploring polypharmacology, where a single compound can affect multiple disease pathways, potentially leading to innovative therapeutic interventions. Privileged structures, molecular frameworks with the propensity to interact with diverse biological targets, have been instrumental in drug discovery by serving as scaffolds for synthesizing compounds with desired pharmacological properties. Their inherent promiscuity provides a starting point for designing multi-target drugs and exploring new therapeutic avenues.

Additionally, the ethical considerations surrounding promiscuity and privileged structures necessitate a careful balance between therapeutic innovation and potential risks. Striking a balance between exploiting promiscuity for therapeutic benefits while minimizing adverse effects remains a crucial challenge.

In essence, the intersection of promiscuity and privileged structures in medicinal chemistry represents a dynamic area of research that holds promise for revolutionizing drug discovery and development. Continued exploration and refinement in this field offer immense potential for the creation of safer, more efficient, and targeted therapeutic interventions for various diseases.

Acknowledgments

The authors would like also to thank INCT-INOFAR (BR, grant number 465.249/2014-0), Fundação Carlos Chagas Filho de Amparo à Pesquisa do Estado do Rio de Janeiro (FAPERJ, grant numbers E-26/010.001273/2016 and SEI-260003/003613/2022, fellowship to CAMF), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, grant number 304.811/2022-0), and Departamento de Ciência e Tecnologia, Ministério da Saúde (DECIT-MS).

  • The authors would like to express their sincere appreciation and gratitude to Prof Eliezer Barreiro for his outstanding and invaluable contributions to the field of Medicinal Chemistry in Brazil. Professor Barreiro’s passion for science, coupled with his mentorship and collaborative spirit, has inspired countless researchers and students alike.

References

  • 1
    Evans, B. E.; Rittle, K. E.; Bock, M. G.; DiPardo, R. M.; Freidinger, R. M.; Whitter, W. L.; Lundell, G. F.; Veber, D. F.; Anderson, P. S.; Chang, R. S. L.; Lotti, V. J.; Cerino, D. J.; Chen, T. B.; Kling, P. J.; Kunkel, K. A.; Springer, J. P.; Hirshfield, J.; J. Med. Chem. 1988, 31, 2235. [Crossref]
    » Crossref
  • 2
    Evans, B. E.; Bock, M. G.; Rittle, K. E.; DiPardo, R. M.; Whitter, W. L.; Veber, D. F.; Anderson, P. S.; Freidinger, R. M.; Proc. Natl. Acad. Sci. U.S.A. 1986, 83, 4918. [Crossref]
    » Crossref
  • 3
    Duarte, C.; Barreiro, E.; Fraga, C.; Mini-Rev. Med. Chem. 2007, 7, 1108. [Crossref]
    » Crossref
  • 4
    Skoreński, M.; Sieńczyk, M.; Pharmaceuticals 2021, 14, 1164. [Crossref]
    » Crossref
  • 5
    Kim, J.; Kim, H.; Park, S. B.; J. Am. Chem. Soc. 2014, 136, 14629. [Crossref]
    » Crossref
  • 6
    Scheiber, J.; Jenkins, J. L.; Sukuru, S. C. K.; Bender, A.; Mikhailov, D.; Milik, M.; Azzaoui, K.; Whitebread, S.; Hamon, J.; Urban, L.; Glick, M.; Davies, J. W.; J. Med. Chem. 2009, 52, 3103. [Crossref]
    » Crossref
  • 7
    Espinoza-Fonseca, L. M.; Bioorg. Med. Chem. 2006, 14, 896. [Crossref]
    » Crossref
  • 8
    Hu, Y.; Bajorath, J.; Eur. J. Med. Chem. 2013, 69, 393. [Crossref]
    » Crossref
  • 9
    Baell, J.; Walters, M. A.; Nature 2014, 513, 481. [Crossref]
    » Crossref
  • 10
    David, L.; Walsh, J.; Sturm, N.; Feierberg, I.; Nissink, J. W. M.; Chen, H.; Bajorath, J.; Engkvist, O.; ChemMedChem. 2019, 14, 1795. [Crossref]
    » Crossref
  • 11
    Blaschke, T.; Miljković, F.; Bajorath, J.; ACS Omega 2019, 4, 6883. [Crossref]
    » Crossref
  • 12
    Mencher, S. K.; Wang, L. G.; BMC Clin. Pharmacol. 2005, 5, 3. [Crossref]
    » Crossref
  • 13
    Hu, Y.; Bajorath, J.; J. Chem. Inf. Model. 2010, 50, 2112. [Crossref]
    » Crossref
  • 14
    Barreiro, E. J. In Privileged Scaffolds in Medicinal Chemistry: Design, Synthesis, Evaluation; Bräse, S., ed.; The Royal Society of Chemistry: Cambridge, UK, 2015, ch. 1, p. 1-15. [Crossref]
    » Crossref
  • 15
    Han, Y.; Dong, W.; Guo, Q.; Li, X.; Huang, L.; Eur. J. Med. Chem. 2020, 203, 112506. [Crossref]
    » Crossref
  • 16
    Fang, G.; Chen, H.; Cheng, Z.; Tang, Z.; Wan, Y.; Eur. J. Med. Chem. 2023, 115621. [Crossref]
    » Crossref
  • 17
    Kim, A.; Cohen, M. S.; Expert. Opin. Drug. Discovery 2016, 11, 907. [Crossref]
    » Crossref
  • 18
    Monestime, S.; Lazaridis, D.; Drugs R&D 2020, 20, 189. [Crossref]
    » Crossref
  • 19
    Gadina, M.; Schwartz, D. M.; O’Shea, J. J.; Arthritis Rheumatol. 2016, 68, 31. [Crossref]
    » Crossref
  • 20
    Liao, J. J.-L.; J. Med. Chem. 2007, 50, 409. [Crossref]
    » Crossref
  • 21
    Irie, T.; Sawa, M.; Chem. Pharm. Bull. 2018, 66, 29. [Crossref]
    » Crossref
  • 22
    Verma, A.; Joshi, S.; Singh, D.; J. Chem. 2013, 2013, ID 329412. [Crossref]
    » Crossref
  • 23
    Alghamdi, S. S.; Suliman, R. S.; Almutairi, K.; Kahtani, K.; Aljatli, D.; Drug Des., Dev. Ther. 2021, 2021, 3289. [Crossref]
    » Crossref
  • 24
    Tolomeu, H. V.; Fraga, C. A. M.; Molecules 2023, 28, 838. [Crossref]
    » Crossref
  • 25
    Hwang, J.; Kim, D.; Park, J. S.; Park, H. J.; Shin, J.; Lee, S. K.; Mar. Drugs 2020, 18, 87. [Crossref]
    » Crossref
  • 26
    Gibbons, J. B.; Salvant, J. M.; Vaden, R. M.; Kwon, K.-H.; Welm, B. E.; Looper, R. E.; J. Org. Chem. 2015, 80, 10076. [Crossref]
    » Crossref
  • 27
    Kapourani, A.; Kontogiannopoulos, K. N.; Barmpalexis, P.; Pharmaceuticals 2022, 15, 762. [Crossref]
    » Crossref
  • 28
    Goa, K. L.; Wagstaff, A. J.; Drugs 1996, 51, 820. [Crossref]
    » Crossref
  • 29
    Xu, F.; Mao, C.; Liu, Y.; Wu, L.; Xu, Z.; Zhang, L.; Curr. Med. Chem. 2009, 16, 3701. [Crossref]
    » Crossref
  • 30
    Sohn, C. A.; Clin. Pharm. 1982, 1, 217. [PubMed]
    » PubMed
  • 31
    Finkelstein, W.; Isselbacher, K. J.; N. Engl. J. Med. 1978, 299, 992. [Crossref]
    » Crossref
  • 32
    McIlhenny, H. M.; Bettis, J. W.; Wiseman, E. H.; J. Pharm. Sci. 1975, 64, 1469. [Crossref]
    » Crossref
  • 33
    Maeda, K.; Osato, T.; Umezawa, H.; J. Antibiot. 1953, 6, 182. [PubMed]
    » PubMed
  • 34
    Al-Badr, A. A.; Alodhaib, M. M.; Profiles Drug Subst., Excipients, Relat. Methodol. 2016, 41, 323. [Crossref]
    » Crossref
  • 35
    Hempel, A.; Camerman, N.; Camerman, A.; Mastropaolo, D.; Acta Crystallogr., Sect. E: Struct. Rep. Online 2005, 61, o1387. [Crossref]
    » Crossref
  • 36
    Ali, A. M.; Tawfk, S. S.; Mostafa, A. S.; Massoud, M. A. M.; Chem. Biol. Drug Des. 2022, 100, 656. [Crossref]
    » Crossref
  • 37
    Guo, Y.; Hou, X.; Fang, H.; Mini-Rev. Med. Chem. 2021, 21, 1367. [Crossref]
    » Crossref
  • 38
    Ajani, O. O.; Aderohunmu, D. V; Ikpo, C. O.; Adedapo, A. E.; Olanrewaju, I. O.; Arch. Pharm. 2016, 349, 475. [Crossref]
    » Crossref
  • 39
    Welsch, M. E.; Snyder, S. A.; Stockwell, B. R.; Curr. Opin. Chem. Biol. 2010, 14, 347. [Crossref]
    » Crossref
  • 40
    Martini, C. A. N.; Scapini, J. G. S.; Collaço, L. M.; Matsubara A.; Veiga Jr., V. F.; Rev. Col. Bras. Cir. 2016, 43, 445. [Crossref]
    » Crossref
  • 41
    Guay, D. R.; Drugs 2001, 61, 353. [Crossref]
    » Crossref
  • 42
    Krause, T.; Gerbershagen, M. U.; Fiege, M.; Weisshorn, R.; Wappler, F.; Anaesthesia 2004, 59, 364. [Crossref]
    » Crossref
  • 43
    Corrêa, J. C. R.; Hiene, M. A. C.; Salgado, H. R. N.; J. Anal. Bioanal. Tech. 2013, 177. [Crossref]
    » Crossref
  • 44
    Pinder, R. M.; Brogden, R. N.; Speight, T. M.; Avery, G. S.; Drugs 1977, 13, 3. [Crossref]
    » Crossref
  • 45
    da Silva, J. S.; Pereira, S. L.; do Couto Maia, R.; Landgraf, S. S.; Caruso-Neves, C.; Kümmerle, A. E.; Fraga, C. A. M.; Barreiro, E. J.; Sudo, R. T.; Zapata-Sudo, G.; Life Sci. 2014, 94, 30. [Crossref]
    » Crossref
  • 46
    Silva, C. L. M.; Noël, F.; Barreiro, E. J.; B r. J. Pharmacol. 2002, 135, 293. [Crossref]
    » Crossref
  • 47
    Zapata-Sudo, G.; Pereira, S. L.; Beiral, H. J. V.; Kummerle, A. E.; Raimundo, J. M.; Antunes, F.; Sudo, R. T.; Barreiro, E. J.; Fraga, C. A. M.; Am. J. Hypertens. 2010, 23, 135. [Crossref]
    » Crossref
  • 48
    Pereira, S. L.; Kummerle, A. E.; Fraga, C. A. M.; Barreiro, E. J.; de Novaes Rocha, N.; Ferraz, E. B.; do Nascimento, J. H. M.; Sudo, R. T.; Zapata-Sudo, G.; Eur. J. Pharmacol. 2013, 702, 316. [Crossref]
    » Crossref
  • 49
    Pereira, S. L.; Kummerle, A. E.; Fraga, C. A. M.; Barreiro, E. J.; Sudo, R. T.; Zapata-Sudo, G.; Fundam. Clin. Pharmacol. 2014, 28, 29. [Crossref]
    » Crossref
  • 50
    Mao, Y.; Soni, K.; Sangani, C.; Yao, Y.; Curr. Top. Med. Chem. 2020, 20, 2599. [Crossref]
    » Crossref
  • 51
    Chu, X.-M.; Wang, C.; Liu, W.; Liang, L.-L.; Gong, K.-K.; Zhao, C.-Y.; Sun, K.-L.; Eur. J. Med. Chem. 2019, 161, 101. [Crossref]
    » Crossref
  • 52
    Yadav, V.; Reang, J.; Sharma, V.; Majeed, J.; Sharma, P. C.; Sharma, K.; Giri, N.; Kumar, A.; Tonk, R. K.; Chem. Biol. Drug Des. 2022, 100, 389. [Crossref]
    » Crossref
  • 53
    da Silva, T. H. Á.; de Oliveira, M. T.; dos Santos, H. F.; de Oliveira, A. B.; de Almeida, W. B.; Quim. Nova 2005, 28, 244. [Crossref]
    » Crossref
  • 54
    Agnello, S.; Brand, M.; Chellat, M. F.; Gazzola, S.; Riedl, R.; Angew. Chem., Int. Ed. 2019, 58, 3300. [Crossref]
    » Crossref
  • 55
    Repac Antić, D.; Parčina, M.; Gobin, I.; Petković Didović, M.; Antibiotics 2022, 11, 1105. [Crossref]
    » Crossref
  • 56
    Zwaagstra, M. E.; Schoenmakers, S. H. H. F.; Nederkoorn, P. H. J.; Gelens, E.; Timmerman, H.; Zhang, M.-Q.; J. Med. Chem. 1998, 41, 1439. [Crossref]
    » Crossref
  • 57
    Aldred, K. J.; Kerns, R. J.; Osheroff, N.; Biochemistry 2014, 53, 1565. [Crossref]
    » Crossref
  • 58
    Tang, Z.; Tan, Y.; Chen, H.; Wan, Y.; Curr. Med. Chem. 2023, 30, 372. [Crossref]
    » Crossref
  • 59
    Kintscher, U.; Edelmann, F.; Cardiovasc. Diabetol. 2023, 22, 162. [Crossref]
    » Crossref
  • 60
    Iijima, T.; Katoh, M.; Takedomi, K.; Yamamoto, Y.; Akatsuka, H.; Shirata, N.; Nishi, A.; Takakuwa, M.; Watanabe, Y.; Munakata, H.; Koyama, N.; Ikeda, T.; Iguchi, T.; Kato, H.; Kikkawa, K.; Kawaguchi, T.; J. Med. Chem. 2022, 65, 8127. [Crossref]
    » Crossref
  • 61
    Keating, G. M.; Drugs 2016, 76, 617. [Crossref]
    » Crossref
  • 62
    Choi, Y. M.; Kim, K. H.; Korean J. Pain 2015, 28, 4. [Crossref]
    » Crossref
  • 63
    Patil, V. M.; Masand, N.; Verma, S.; Masand, V.; Chem. Biol. Drug Des. 2021, 98, 943. [Crossref]
    » Crossref
  • 64
    Zhu, R.; Jiang, G.; Tang, W.; Zhao, X.; Chen, F.; Zhang, X.; Ye, N.; Eur. J. Med. Chem. 2023, 256, 115414. [Crossref]
    » Crossref
  • 65
    Barresi, E.; Robello, M.; Baglini, E.; Poggetti, V.; Viviano, M.; Salerno, S.; Da Settimo, F.; Taliani, S.; Pharmaceuticals 2023, 16, 997. [Crossref]
    » Crossref
  • 66
    Sun, J.; Chen, Y.-H.; Liu, H.-Y.; Hondo, E.; Zhou, Y.; Wu, Y.-F.; Curr. Top. Med. Chem. 2021, 21, 2529. [Crossref]
    » Crossref
  • 67
    Malki, Y.; Martinez, J.; Masurier, N.; Med. Res. Rev. 2021, 41, 2247. [Crossref]
    » Crossref
  • 68
    Nazeri, M. T.; Nasiriani, T.; Torabi, S.; Shaabani, A.; Org. Biomol. Chem. 2024, 22, 1102. [Crossref]
    » Crossref
  • 69
    Wang, R.; Wang, Y.-A.; Xu, Y.-G.; Shi, L.; Curr. Top. Med. Chem. 2022, 22, 600. [Crossref]
    » Crossref
  • 70
    Shenkar, R.; Shi, C.; Austin, C.; Moore, T.; Lightle, R.; Cao, Y.; Zhang, L.; Wu, M.; Zeineddine, H. A.; Girard, R.; McDonald, D. A.; Rorrer, A.; Gallione, C.; Pytel, P.; Liao, J. K.; Marchuk, D. A.; Awad, I. A.; Stroke 2017, 48, 187. [Crossref]
    » Crossref
  • 71
    Kahan, B. D.; Expert Opin. Pharmacother. 2001, 2, 1903. [Crossref]
    » Crossref
  • 72
    Sehgal, S. N.; Transplant. Proc. 2003, 35, S7. [Crossref]
    » Crossref
  • 73
    Li, J.; Kim, S. G.; Blenis, J.; Cell Metab. 2014, 19, 373. [Crossref]
    » Crossref
  • 74
    Cohen, M. H.; Farrell, A.; Justice, R.; Pazdur, R.; Oncologist 2009, 14, 174. [Crossref]
    » Crossref
  • 75
    Attwood, M. M.; Fabbro, D.; Sokolov, A. V.; Knapp, S.; Schiöth, H. B.; Nat. Rev. Drug Discovery 2021, 20, 798. [Crossref]
    » Crossref
  • 76
    Faridoon; Ng, R.; Zhang, G.; Li, J. J.; Med. Chem. Res. 2023, 32, 1039. [Crossref]
    » Crossref
  • 77
    Zhao, Z.; Bourne, P. E.; Pharmaceuticals 2022, 15, 1322. [Crossref]
    » Crossref
  • 78
    Roskoski Jr., R.; Pharmacol. Res. 2023, 187, 106552. [Crossref]
    » Crossref
  • 79
    Roskoski Jr., R.; Pharmacol. Res. 2021, 165, 105422. [Crossref]
    » Crossref
  • 80
    Xerxa, E.; Miljković, F.; Bajorath, J.; J. Med. Chem. 2023, 66, 7657. [Crossref]
    » Crossref
  • 81
    Serafimova, I. M.; Pufall, M. A.; Krishnan, S.; Duda, K.; Cohen, M. S.; Maglathlin, R. L.; McFarland, J. M.; Miller, R. M.; Frödin, M.; Taunton, J.; Nat. Chem. Biol. 2012, 8, 471. [Crossref]
    » Crossref
  • 82
    Gabizon, R.; Shraga, A.; Gehrtz, P.; Livnah, E.; Shorer, Y.; Gurwicz, N.; Avram, L.; Unger, T.; Aharoni, H.; Albeck, S.; Brandis, A.; Shulman, Z.; Katz, B.-Z.; Herishanu, Y.; London, N.; J. Am. Chem. Soc. 2020, 142, 11734. [Crossref]
    » Crossref
  • 83
    Fairhurst, R. A.; Knoepfel, T.; Buschmann, N.; Leblanc, C.; Mah, R.; Todorov, M.; Nimsgern, P.; Ripoche, S.; Niklaus, M.; Warin, N.; Luu, V. H.; Madoerin, M.; Wirth, J.; Graus-Porta, D.; Weiss, A.; Kiffe, M.; Wartmann, M.; Kinyamu-Akunda, J.; Sterker, D.; Stamm, C.; Adler, F.; Buhles, A.; Schadt, H.; Couttet, P.; Blank, J.; Galuba, I.; Trappe, J.; Voshol, J.; Ostermann, N.; Zou, C.; Berghausen, J.; Del Rio Espinola, A.; Jahnke, W.; Furet, P.; J. Med. Chem. 2020, 63, 12542. [Crossref]
    » Crossref
  • 84
    Guo, W.-H.; Qi, X.; Yu, X.; Liu, Y.; Chung, C.-I.; Bai, F.; Lin, X.; Lu, D.; Wang, L.; Chen, J.; Su, L. H.; Nomie, K. J.; Li, F.; Wang, M. C.; Shu, X.; Onuchic, J. N.; Woyach, J. A.; Wang, M. L.; Wang, J.; Nat. Commun. 2020, 11, 4268. [Crossref]
    » Crossref
  • 85
    Bandyopadhyay, A.; Gao, J.; Curr. Opin. Chem. Biol. 2016, 34, 110. [Crossref]
    » Crossref
  • 86
    Reja, R. M.; Wang, W.; Lyu, Y.; Haeffner, F.; Gao, J.; J. Am. Chem. Soc. 2022, 144, 1152. [Crossref]
    » Crossref
  • 87
    Bradshaw, J. M.; McFarland, J. M.; Paavilainen, V. O.; Bisconte, A.; Tam, D.; Phan, V. T.; Romanov, S.; Finkle, D.; Shu, J.; Patel, V.; Ton, T.; Li, X.; Loughhead, D. G.; Nunn, P. A.; Karr, D. E.; Gerritsen, M. E.; Funk, J. O.; Owens, T. D.; Verner, E.; Brameld, K. A.; Hill, R. J.; Goldstein, D. M.; Taunton, J.; Nat. Chem. Biol. 2015, 11, 525. [Crossref]
    » Crossref
  • 88
    Krishnan, S.; Miller, R. M.; Tian, B.; Mullins, R. D.; Jacobson, M. P.; Taunton, J.; J. Am. Chem. Soc. 2014, 136, 12624. [Crossref]
    » Crossref
  • 89
    Shindo, N.; Ojida, A.; Bioorg. Med. Chem. 2021, 47, 116386. [Crossref]
    » Crossref
  • 90
    Xerxa, E.; Laufkötter, O.; Bajorath, J.; Molecules 2023, 28, 5805. [Crossref]
    » Crossref
  • 91
    Abdeldayem, A.; Raouf, Y. S.; Constantinescu, S. N.; Moriggl, R.; Gunning, P. T.; Chem. Soc. Rev. 2020, 49, 2617. [Crossref]
    » Crossref
  • 92
    Kuter, D. J.; Bussel, J. B.; Ghanima, W.; Cooper, N.; Gernsheimer, T.; Lambert, M. P.; Liebman, H. A.; Tarantino, M. D.; Lee, M.; Guo, H.; Daak, A.; Ther. Adv. Hematol. 2023, 14. [Crossref]
    » Crossref
  • 93
    The National Library of Medicine, ClinicalTrials.gov, accessed in May 2024.
  • 94
    European Union Drug Regulating Authorities Clinical Trials Database, https://eudract.ema.europa.eu/, accessed in May 2024.
    » https://eudract.ema.europa.eu/
  • 95
    Shindo, N.; Fuchida, H.; Sato, M.; Watari, K.; Shibata, T.; Kuwata, K.; Miura, C.; Okamoto, K.; Hatsuyama, Y.; Tokunaga, K.; Sakamoto, S.; Morimoto, S.; Abe, Y.; Shiroishi, M.; Caaveiro, J. M. M.; Ueda, T.; Tamura, T.; Matsunaga, N.; Nakao, T.; Koyanagi, S.; Ohdo, S.; Yamaguchi, Y.; Hamachi, I.; Ono, M.; Ojida, A.; Nat. Chem. Biol. 2019, 15, 250. [Crossref]
    » Crossref
  • 96
    Augeri, D. J.; Robl, J. A.; Betebenner, D. A.; Magnin, D. R.; Khanna, A.; Robertson, J. G.; Wang, A.; Simpkins, L. M.; Taunk, P.; Huang, Q.; Han, S.-P.; Abboa-Offei, B.; Cap, M.; Xin, L.; Tao, L.; Tozzo, E.; Welzel, G. E.; Egan, D. M.; Marcinkeviciene, J.; Chang, S. Y.; Biller, S. A.; Kirby, M. S.; Parker, R. A.; Hamann, L. G.; J. Med. Chem. 2005, 48, 5025. [Crossref]
    » Crossref
  • 97
    O’Meara, J. A.; Lemke, C. T.; Godbout, C.; Kukolj, G.; Lagacé, L.; Moreau, B.; Thibeault, D.; White, P. W.; Llinàs-Brunet, M.; J. Biol. Chem. 2013, 288, 5673. [Crossref]
    » Crossref
  • 98
    Arasappan, A.; Bennett, F.; Bogen, S. L.; Venkatraman, S.; Blackman, M.; Chen, K. X.; Hendrata, S.; Huang, Y.; Huelgas, R. M.; Nair, L.; Padilla, A. I.; Pan, W.; Pike, R.; Pinto, P.; Ruan, S.; Sannigrahi, M.; Velazquez, F.; Vibulbhan, B.; Wu, W.; Yang, W.; Saksena, A. K.; Girijavallabhan, V.; Shih, N.-Y.; Kong, J.; Meng, T.; Jin, Y.; Wong, J.; McNamara, P.; Prongay, A.; Madison, V.; Piwinski, J. J.; Cheng, K.-C.; Morrison, R.; Malcolm, B.; Tong, X.; Ralston, R.; Njoroge, F. G.; ACS Med. Chem. Lett. 2010, 1, 64. [Crossref]
    » Crossref
  • 99
    Greener, B. S.; Millan, D. S. In Modern Drug Synthesis; Wiley: New Jersey, 2010, p. 99-110. [Crossref]
    » Crossref
  • 100
    Shirley, M.; Drugs 2016, 76, 405. [Crossref]
    » Crossref
  • 101
    Oksenberg, D.; Dufu, K.; Patel, M. P.; Chuang, C.; Li, Z.; Xu, Q.; Silva-Garcia, A.; Zhou, C.; Hutchaleelaha, A.; Patskovska, L.; Patskovsky, Y.; Almo, S. C.; Sinha, U.; Metcalf, B. W.; Archer, D. R.; Br. J. Haematol. 2016, 175, 141. [Crossref]
    » Crossref
  • 102
    Greasley, S. E.; Noell, S.; Plotnikova, O.; Ferre, R.; Liu, W.; Bolanos, B.; Fennell, K.; Nicki, J.; Craig, T.; Zhu, Y.; Stewart, A. E.; Steppan, C. M.; J. Biol. Chem. 2022, 298, 101972. [Crossref]
    » Crossref
  • 103
    Smith, P. F.; Krishnarajah, J.; Nunn, P. A.; Hill, R. J.; Karr, D.; Tam, D.; Masjedizadeh, M.; Funk, J. O.; Gourlay, S. G.; Br J. Clin. Pharmacol. 2017, 83, 2367. [Crossref]
    » Crossref
  • 104
    Forster, M.; Chaikuad, A.; Bauer, S. M.; Holstein, J.; Robers, M. B.; Corona, C. R.; Gehringer, M.; Pfaffenrot, E.; Ghoreschi, K.; Knapp, S.; Laufer, S. A.; Cell Chem. Biol. 2016, 23, 1335. [Crossref]
    » Crossref
  • 105
    Benson, M. J.; Rodriguez, V.; von Schack, D.; Keegan, S.; Cook, T. A.; Edmonds, J.; Benoit, S.; Seth, N.; Du, S.; Messing, D.; Nickerson-Nutter, C. L.; Dunussi-Joannopoulos, K.; Rankin, A. L.; Ruzek, M.; Schnute, M. E.; Douhan, J.; J. Immunol. 2014, 193, 185. [Crossref]
    » Crossref
  • 106
    Weiss, A.; Adler, F.; Buhles, A.; Stamm, C.; Fairhurst, R. A.; Kiffe, M.; Sterker, D.; Centeleghe, M.; Wartmann, M.; Kinyamu-Akunda, J.; Schadt, H. S.; Couttet, P.; Wolf, A.; Wang, Y.; Barzaghi-Rinaudo, P.; Murakami, M.; Kauffmann, A.; Knoepfel, T.; Buschmann, N.; Leblanc, C.; Mah, R.; Furet, P.; Blank, J.; Hofmann, F.; Sellers, W. R.; Graus Porta, D.; Mol. Cancer Ther. 2019, 18, 2194. [Crossref]
    » Crossref
  • 107
    Sakamoto, K. M.; Kim, K. B.; Kumagai, A.; Mercurio, F.; Crews, C. M.; Deshaies, R. J.; Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 8554. [Crossref]
    » Crossref
  • 108
    Schneekloth, A. R.; Pucheault, M.; Tae, H. S.; Crews, C. M.; Bioorg. Med. Chem. Lett. 2008, 18, 5904. [Crossref]
    » Crossref
  • 109
    Yao, T.; Xiao, H.; Wang, H.; Xu, X.; Int. J. Mol. Sci. 2022, 23, 10328. [Crossref]
    » Crossref
  • 110
    Liu, Z.; Hu, M.; Yang, Y.; Du, C.; Zhou, H.; Liu, C.; Chen, Y.; Fan, L.; Ma, H.; Gong, Y.; Xie, Y.; Mol. Biomed. 2022, 3, 46. [Crossref]
    » Crossref
  • 111
    Buckley, D. L.; Crews, C. M.; Angew. Chem., Int. Ed. 2014, 53, 2312. [Crossref]
    » Crossref
  • 112
    Gadd, M. S.; Testa, A.; Lucas, X.; Chan, K.-H.; Chen, W.; Lamont, D. J.; Zengerle, M.; Ciulli, A.; Nat. Chem. Biol. 2017, 13, 514. [Crossref]
    » Crossref
  • 113
    Finley, D.; Annu. Rev. Biochem. 2009, 78, 477. [Crossref]
    » Crossref
  • 114
    Hipp, M. S.; Kasturi, P.; Hartl, F. U.; Nat. Rev. Mol. Cell Biol. 2019, 20, 421. [Crossref]
    » Crossref
  • 115
    Kliza, K.; Husnjak, K.; Front. Mol. Biosci. 2020, 7, 21. [Crossref]
    » Crossref
  • 116
    Konstantinidou, M.; Li, J.; Zhang, B.; Wang, Z.; Shaabani, S.; Ter Brake, F.; Essa, K.; Dömling, A.; Expert Opin. Drug Discovery 2019, 14, 1255. [Crossref]
    » Crossref
  • 117
    Cromm, P. M.; Crews, C. M.; Cell Chem. Biol. 2017, 24, 1181. [Crossref]
    » Crossref
  • 118
    Hu, Z.; Crews, C. M.; ChemBioChem. 2022, 23. [Crossref]
    » Crossref
  • 119
    Komander, D.; Rape, M.; Annu. Rev. Biochem. 2012, 81, 203. [Crossref]
    » Crossref
  • 120
    Chen, Y.; Jin, J.; Acta Biochim. Biophys. Sin. 2020, 52, 776. [Crossref]
    » Crossref
  • 121
    Burslem, G. M.; Crews, C. M.; Cell 2020, 181, 102. [Crossref]
    » Crossref
  • 122
    Salami, J.; Crews, C. M.; Science 2017, 355, 1163. [Crossref]
    » Crossref
  • 123
    Ito, T.; Ando, H.; Handa, H.; Cell. Mol. Life Sci. 2011, 68, 1569. [Crossref]
    » Crossref
  • 124
    Chamberlain, P. P.; Lopez-Girona, A.; Miller, K.; Carmel, G.; Pagarigan, B.; Chie-Leon, B.; Rychak, E.; Corral, L. G.; Ren, Y. J.; Wang, M.; Riley, M.; Delker, S. L.; Ito, T.; Ando, H.; Mori, T.; Hirano, Y.; Handa, H.; Hakoshima, T.; Daniel, T. O.; Cathers, B. E.; Nat. Struct. Mol. Biol. 2014, 21, 803. [Crossref]
    » Crossref
  • 125
    Zhu, Y. X.; Braggio, E.; Shi, C.-X.; Bruins, L. A.; Schmidt, J. E.; Van Wier, S.; Chang, X.-B.; Bjorklund, C. C.; Fonseca, R.; Bergsagel, P. L.; Orlowski, R. Z.; Stewart, A. K.; Blood 2011, 118, 4771. [Crossref]
    » Crossref
  • 126
    Krönke, J.; Udeshi, N. D.; Narla, A.; Grauman, P.; Hurst, S. N.; McConkey, M.; Svinkina, T.; Heckl, D.; Comer, E.; Li, X.; Ciarlo, C.; Hartman, E.; Munshi, N.; Schenone, M.; Schreiber, S. L.; Carr, S. A.; Ebert, B. L.; Science 2014, 343, 301. [Crossref]
    » Crossref
  • 127
    Krönke, J.; Fink, E. C.; Hollenbach, P. W.; MacBeth, K. J.; Hurst, S. N.; Udeshi, N. D.; Chamberlain, P. P.; Mani, D. R.; Man, H. W.; Gandhi, A. K.; Svinkina, T.; Schneider, R. K.; McConkey, M.; Järås, M.; Griffiths, E.; Wetzler, M.; Bullinger, L.; Cathers, B. E.; Carr, S. A.; Chopra, R.; Ebert, B. L.; Nature 2015, 523, 183. [Crossref]
    » Crossref
  • 128
    Sun, X.; Gao, H.; Yang, Y.; He, M.; Wu, Y.; Song, Y.; Tong, Y.; Rao, Y.; Signal Transduction Targeted Ther. 2019, 4, 64. [Crossref]
    » Crossref
  • 129
    Bassi, Z. I.; Fillmore, M. C.; Miah, A. H.; Chapman, T. D.; Maller, C.; Roberts, E. J.; Davis, L. C.; Lewis, D. E.; Galwey, N. W.; Waddington, K. E.; Parravicini, V.; Macmillan-Jones, A. L.; Gongora, C.; Humphreys, P. G.; Churcher, I.; Prinjha, R. K.; Tough, D. F.; ACS Chem. Biol. 2018, 13, 2862. [Crossref]
    » Crossref
  • 130
    Silva, M. C.; Ferguson, F. M.; Cai, Q.; Donovan, K. A.; Nandi, G.; Patnaik, D.; Zhang, T.; Huang, H.-T.; Lucente, D. E.; Dickerson, B. C.; Mitchison, T. J.; Fischer, E. S.; Gray, N. S.; Haggarty, S. J.; eLife 2019, 8, e45457. [Crossref]
    » Crossref
  • 131
    de Wispelaere, M.; Du, G.; Donovan, K. A.; Zhang, T.; Eleuteri, N. A.; Yuan, J. C.; Kalabathula, J.; Nowak, R. P.; Fischer, E. S.; Gray, N. S.; Yang, P. L.; Nat. Commun. 2019, 10, 3468. [Crossref]
    » Crossref
  • 132
    Bricelj, A.; Steinebach, C.; Kuchta, R.; Gütschow, M.; Sosič, I.; Front. Chem. 2021, 9, 707317. [Crossref]
    » Crossref
  • 133
    Weng, G.; Shen, C.; Cao, D.; Gao, J.; Dong, X.; He, Q.; Yang, B.; Li, D.; Wu, J.; Hou, T.; Nucleic. Acids. Res. 2021, 49, D1381. [Crossref]
    » Crossref
  • 134
    Chamberlain, P. P.; Hamann, L. G.; Nat. Chem. Biol. 2019, 15, 937. [Crossref]
    » Crossref
  • 135
    Gabizon, R.; London, N.; Curr. Opin. Chem. Biol. 2021, 62, 24. [Crossref]
    » Crossref
  • 136
    McGovern, S. L.; Caselli, E.; Grigorieff, N.; Shoichet, B. K.; J. Med. Chem. 2002, 45, 1712. [Crossref]
    » Crossref
  • 137
    Ryan, A. J.; Gray, N. M.; Lowe, P. N.; Chung, C.; J. Med. Chem. 2003, 46, 3448. [Crossref]
    » Crossref
  • 138
    Jöst, C.; Nitsche, C.; Scholz, T.; Roux, L.; Klein, C. D.; J. Med. Chem. 2014, 57, 7590. [Crossref]
    » Crossref
  • 139
    Metz, J. T.; Huth, J. R.; Hajduk, P. J.; J. Comput.-Aided Mol. Des. 2007, 21, 139. [Crossref]
    » Crossref
  • 140
    Baell, J. B.; Holloway, G. A.; J. Med. Chem. 2010, 53, 2719. [Crossref]
    » Crossref
  • 141
    Campos-Olivas, R.; Curr. Top. Med. Chem. 2011, 11, 43. [Crossref]
    » Crossref
  • 142
    Capuzzi, S. J.; Muratov, E. N.; Tropsha, A.; J. Chem. Inf. Model. 2017, 57, 417. [Crossref]
    » Crossref
  • 143
    Senger, M. R.; Fraga, C. A. M.; Dantas, R. F.; Silva, F. P.; Drug Discovery Today 2016, 21, 868. [Crossref]
    » Crossref
  • 144
    Copeland, R. A.; Nat. Rev. Drug Discovery 2016, 15, 87. [Crossref]
    » Crossref
  • 145
    Bernetti, M.; Masetti, M.; Rocchia, W.; Cavalli, A.; Annu. Rev. Phys. Chem. 2019, 70, 143. [Crossref]
    » Crossref
  • 146
    Lee, K. S. S.; Yang, J.; Niu, J.; Ng, C. J.; Wagner, K. M.; Dong, H.; Kodani, S. D.; Wan, D.; Morisseau, C.; Hammock, B. D.; ACS Cent. Sci. 2019, 5, 1614. [Crossref]
    » Crossref
  • 147
    Sutanto, F.; Konstantinidou, M.; Dömling, A.; RSC Med. Chem. 2020, 11, 876. [Crossref]
    » Crossref
  • 148
    Dai, Y.; Hartandi, K.; Ji, Z.; Ahmed, A. A.; Albert, D. H.; Bauch, J. L.; Bouska, J. J.; Bousquet, P. F.; Cunha, G. A.; Glaser, K. B.; Harris, C. M.; Hickman, D.; Guo, J.; Li, J.; Marcotte, P. A.; Marsh, K. C.; Moskey, M. D.; Martin, R. L.; Olson, A. M.; Osterling, D. J.; Pease, L. J.; Soni, N. B.; Stewart, K. D.; Stoll, V. S.; Tapang, P.; Reuter, D. R.; Davidsen, S. K.; Michaelides, M. R.; J. Med. Chem. 2007, 50, 1584. [Crossref]
    » Crossref
  • 149
    Woyach, J. A.; Furman, R. R.; Liu, T.-M.; Ozer, H. G.; Zapatka, M.; Ruppert, A. S.; Xue, L.; Li, D. H.-H.; Steggerda, S. M.; Versele, M.; Dave, S. S.; Zhang, J.; Yilmaz, A. S.; Jaglowski, S. M.; Blum, K. A.; Lozanski, A.; Lozanski, G.; James, D. F.; Barrientos, J. C.; Lichter, P.; Stilgenbauer, S.; Buggy, J. J.; Chang, B. Y.; Johnson, A. J.; Byrd, J. C.; New Engl. J. Med. 2014, 370, 2286. [Crossref]
    » Crossref
  • 150
    Butterworth, S.; Cross, D. A. E.; Finlay, M. R. V.; Ward, R. A.; Waring, M. J.; Med. Chem. Comm. 2017, 8, 820. [Crossref]
    » Crossref
  • 151
    Boike, L.; Henning, N. J.; Nomura, D. K.; Nat. Rev. Drug Discovery 2022, 21, 881. [Crossref]
    » Crossref
  • 152
    Reddy, A. S.; Zhang, S.; Expert. Rev. Clin. Pharmacol. 2013, 6, 41. [Crossref]
    » Crossref
  • 153
    Zhang, W.; Bai, Y.; Wang, Y.; Xiao, W.; Curr. Pharm. Des. 2016, 22, 3171. [Crossref]
    » Crossref
  • 154
    Dai, S.-X.; Li, W.-X.; Li, G.-H.; Huang, J.-F.; Peer. J. 2016, 4, e1791. [Crossref]
    » Crossref
  • 155
    Alfonso, L.; Ai, G.; Spitale, R. C.; Bhat, G. J.; Br. J. Cancer 2014, 111, 61. [Crossref]
    » Crossref
  • 156
    Ito, T.; Ando, H.; Suzuki, T.; Ogura, T.; Hotta, K.; Imamura, Y.; Yamaguchi, Y.; Handa, H.; Science 2010, 327, 1345. [Crossref]
    » Crossref
  • 157
    Hu, Y.; Bajorath, J.; AAPS J. 2013, 15, 104. [Crossref]
    » Crossref
  • 158
    Jasial, S.; Hu, Y.; Bajorath, J.; PLoS One 2016, 11, e0153873. [Crossref]
    » Crossref
  • 159
    Lounkine, E.; Keiser, M. J.; Whitebread, S.; Mikhailov, D.; Hamon, J.; Jenkins, J. L.; Lavan, P.; Weber, E.; Doak, A. K.; Côté, S.; Shoichet, B. K.; Urban, L.; Nature 2012, 486, 361. [Crossref]
    » Crossref
  • 160
    Yıldırım, M. A.; Goh, K.-I.; Cusick, M. E.; Barabási, A.-L.; Vidal, M.; Nat. Biotechnol. 2007, 25, 1119. [Crossref]
    » Crossref
  • 161
    Ezzat, A.; Wu, M.; Li, X.-L.; Kwoh, C.-K.; Briefings Bioinf. 2019, 20, 1337. [Crossref]
    » Crossref

Edited by

Editor handled this article: Albertina Moglioni (Associate)

Publication Dates

  • Publication in this collection
    28 June 2024
  • Date of issue
    2024

History

  • Received
    30 Jan 2024
  • Published
    29 May 2024
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br