Next Article in Journal
Cellular and Molecular Mechanisms of Non-Invasive Brain Stimulation Techniques: A Systematic Review on the Implications for the Treatment of Neurological Disorders
Next Article in Special Issue
Enhancing the Nucleoside Analog Response with Translational Therapeutic Approaches to Overcome Resistance
Previous Article in Journal
Macrophage-Derived Exosomal MALAT1 Induced by Hyperglycemia Regulates Vascular Calcification Through miR-143-3p/MGP Axis in Cultured Vascular Smooth Muscle Cells and Diabetic Rat Carotid Artery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Next-Generation HDAC Inhibitors: Advancing Zinc-Binding Group Design for Enhanced Cancer Therapy

Department of Pharmaceutical Chemistry and Technology, Faculty of Pharmacy, An-Najah National University, Nablus 00433, Palestine
Cells 2025, 14(24), 1997; https://doi.org/10.3390/cells14241997
Submission received: 14 November 2025 / Revised: 10 December 2025 / Accepted: 14 December 2025 / Published: 15 December 2025
(This article belongs to the Special Issue Small Molecule Inhibitors: A New Era in Cancer Treatment)

Highlights

What are the main findings?
  • This review systematically analyzes recently developed HDAC inhibitors in recent years, emphasizing structural evolution across CAP, linker, and zinc-binding groups (ZBGs).
  • It highlights the emergence of hydroxamate and non-hydroxamate ZBGs.
What are the implications of the main findings?
  • Understanding SAR trends across diverse scaffolds provides a rational foundation for designing next-generation HDAC inhibitors with improved isoform selectivity, potency, and therapeutic tolerance.
  • The integration of structural insights with updated clinical trial data supports more informed medicinal-chemistry strategies aimed at accelerating the development of clinically viable HDAC-targeted anticancer agents.

Abstract

Histone deacetylases (HDACs) are pivotal epigenetic regulators that control gene expression, cell proliferation, and differentiation, and their dysregulation is closely associated with the onset and progression of multiple cancers. The therapeutic importance of these enzymes is reflected by FDA approval of HDAC inhibitors for oncology indications. Despite this clinical success, most FDA-approved agents employ conventional zinc-binding groups (ZBGs) such as hydroxamic acid and 2-aminoanilide, which are frequently linked to metabolic instability, genotoxicity, and poor pharmacokinetic behavior. These limitations have spurred the development of structurally diverse and safer HDAC inhibitors incorporating alternative ZBGs. This review provides a comprehensive analysis of recently developed HDAC inhibitors reported in the last few years, emphasizing their structure–activity relationships (SARs), chemical scaffolds, and binding features—including cap, linker, and ZBG motifs. Both hydroxamate-based and non-hydroxamate inhibitors, such as benzamides, hydrazides, and thiol-containing analogs, are critically evaluated. Moreover, the potency and selectivity profiles of these inhibitors are summarized across different cancer and normal cell lines, as well as specific HDAC isoforms, providing a clearer understanding of their therapeutic potential. Emerging dual-target HDAC inhibitors, such as HDAC–tubulin, HDAC–PI3K and HDAC–CDK hybrids, are also discussed for their synergistic anticancer effects.

1. Introduction

Although significant advancements in potential anticancer therapies have been made, cancer continues to be a leading global trigger of mortality [1,2]. The efficacy of current FDA-approved drugs is severely hampered by prevalent issues like multiple drug resistance and debilitating side effects, underscoring a critical and immediate need for novel, less toxic therapeutics [3,4]. Extensive ongoing research is dedicated to identifying compounds with improved safety profiles. In pursuit of this goal, drug designers are focusing on synthesizing new chemical structures that specifically inhibit key biological pathways central to cancer progression, such as EGFR, CDKs, Ras, HDACs, and tubulin proteins, which serve as foundational targets for innovative anticancer interventions [5,6,7].
In parallel with the discovery of new HDAC inhibitors, recent advances in HDAC-based therapeutic strategies have expanded the scope of cancer treatment beyond classical enzymatic inhibition [8]. Combination regimens pairing HDAC inhibitors with targeted agents—including kinase inhibitors, DNMT inhibitors, and immunomodulatory drugs—have demonstrated enhanced synergistic efficacy and are increasingly explored in clinical trials [9,10]. Moreover, the emergence of HDAC-directed PROTAC degraders represents a transformative approach aimed at selectively inducing the proteasomal degradation of HDAC isoforms rather than merely blocking their catalytic activity [11,12].
HDACs are essential enzymes that regulate gene expression by removing acetyl groups from histone proteins. This critical deacetylation process directly influences chromatin structure and, consequently, the accessibility of DNA for transcription. The human proteome contains 18 distinct HDAC isoforms, categorized into four main classes (I, II, III, and IV), each defined by specific biological functions and tissue distribution [13]. Dysregulation of HDAC activity is strongly implicated in various pathologies, including cancer, neurodegenerative, and inflammatory disorders [14]. In a cancer context, impaired HDAC function can promote cell proliferation and survival by repressing crucial tumor suppressor genes [15]. Given this significant role in disease pathogenesis, the pharmaceutical discovery and development of therapeutic HDACi have become a major focus of research in recent years [16,17,18].
While all four classes of HDACs have been implicated in various cancers, Class I HDACs (HDAC1, HDAC2, HDAC3, and HDAC8) are generally considered the most directly and frequently associated with tumorigenesis and are often the primary targets in initial cancer therapeutic strategies [19,20].
Over recent years, HDACi have garnered significant attention as promising anticancer therapeutics, demonstrating efficacy against both solid and hematological malignancies [21]. To date, regulatory authorities have approved four pan-HDACi; these approved agents, including the hydroxamate-based compounds Vorinostat (SAHA), Belinostat (PXD101), Panobinostat (LBH589), and the thiol prodrug Romidepsin (FK228), are a testament to the clinical viability of HDAC inhibition [22,23].
Structurally, HDACi are categorized according to their ZBGs, which fundamentally dictate their binding mode, potency, and selectivity. The major ZBG-based classes include hydroxamates, 2-aminobenzamides, cyclic peptides, thiols, and short-chain fatty acids. While hydroxamates and benzamides dominate clinically approved HDAC inhibitors, increasing concerns regarding their metabolic instability, off-target metal chelation, and toxicity have accelerated the search for safer, more pharmacologically favorable alternatives to next-generation HDACi [24,25]. The hydrazide group, which could be considered as next-generation agents, has recently gained significant attention as an emerging ZBG with promising medicinal chemistry attributes. As highlighted in the recent perspective by Raucci et al. [26], hydrazides offer improved chemical stability, reduced side effects, and a well-characterized therapeutic profile, supported by their presence in clinically established drugs such as isoniazid. Beyond their documented antioxidant, antimicrobial, and anti-inflammatory properties, hydrazides have demonstrated selective and potent HDAC1–3 inhibition, first evidenced by the high-throughput discovery of UF010. These growing findings collectively position the hydrazide moiety as a compelling next-generation ZBG, capable of overcoming key limitations associated with traditional HDAC inhibitors [27,28].
The ZBG represents the core pharmacophore responsible for coordinating with the catalytic Zn2+ ion, and even subtle structural modifications can significantly influence binding affinity, isoform selectivity, metabolic stability, and overall therapeutic performance [29]. Newly explored ZBGs (next generation)—such as hydrazides, thiol-based motifs, and other non-classical chelators—provide refined metal-binding geometries that improve stereoelectronic complementarity with individual HDAC isoforms while minimizing off-target metal chelation and toxicity [30,31]. These design innovations are central to next-generation HDACi, as they promote stronger zinc coordination, more favorable cap–linker–ZBG alignment within the catalytic channel, and enhanced pharmacokinetic behavior. Collectively, such optimized ZBG architectures serve as key drivers in the development of safer, more selective, and more potent next-generation HDAC inhibitors capable of delivering superior therapeutic outcomes in cancer therapy.
This review aims to provide a clear and comprehensive overview of the most recent and genuinely novel HDACi, with a specific emphasis on compounds reported in the past few years. In addition to summarizing hydroxamate- and non-hydroxamate-based HDACis, the review highlights structurally innovative scaffolds, detailed SAR features, isoform selectivity, and therapeutic potential. Particular attention is devoted to newly emerging HDACi with promising preclinical profiles that demonstrate clear potential for translation into future clinical applications, and FDA-approved agents. Moreover, a dedicated section discusses the chemical design and recent advances in dual-target HDAC inhibitors, an increasingly important strategy in modern anticancer drug development. By focusing on the most innovative agents and integrating updated insights from the latest literature, this review aims to enhance readability, address gaps highlighted by peer evaluation, and provide meaningful guidance for future medicinal chemistry efforts and the discovery of next-generation epigenetic therapeutics.

2. FDA-Approved and Clinically Approved HDACi

HDACi have emerged as one of the most promising classes of epigenetic modulators in cancer therapy. Based on their chemical scaffolds and ZBGs, HDACis can be broadly classified into hydroxamic acids, cyclic peptides, benzamides, and short-chain fatty acids, each exhibiting distinct selectivity profiles toward HDAC isoforms and diverse clinical indications [15,32].
Hydroxamic acid derivatives represent the most extensively studied class due to their strong chelation of the catalytic zinc ion in HDAC active sites [33]. Among them, Vorinostat and Belinostat were the first to achieve FDA approval (in 2006 and 2014, respectively) (Figure 1) for the treatment of cutaneous and peripheral T-cell lymphomas (CTCL and PTCL) [34,35].
The FDA’s approval of Duvyzat (givinostat) marks a major advancement in the treatment of Duchenne muscular dystrophy (DMD)—a severe genetic disease-causing progressive muscle weakness and degeneration. As a histone deacetylase (HDAC) inhibitor, Duvyzat works by modulating gene expression to enhance dystrophin production, reduce inflammation, and protect muscle tissue. Clinical trials confirmed its safety and efficacy, leading to its regulatory approval. While not a cure, Duvyzat represents a meaningful step forward, offering patients a targeted therapy that addresses key molecular mechanisms of DMD and bringing new hope for improved quality of life and future therapeutic innovation [45,46].
Other analogs, such as Panobinostat and Resminostat, have highlighted clinical efficacy in multiple myeloma (MM) and colorectal or hepatocellular carcinoma (HCC), respectively [9,47]. Resminostat is a hydroxamate-based HDACi with the chemical structure (E)-3-[1-[4-[(dimethylamino)methyl]phenyl]sulfonylpyrrol-3-yl]-N-hydroxyprop-2-enamide. Furthermore, this compound belongs to the class of hydroxamic acid derivatives and exerts its pharmacological activity through the classical hydroxamate–zinc ion chelation mechanism at the HDAC catalytic site [9,48]. The presence of the N-hydroxyacrylamide group serves as the zinc-binding pharmacophore (ZBG), while the pyrrole–sulfonyl–dimethylaminophenyl scaffold provides both hydrophobic and hydrogen-bonding interactions that contribute to HDAC selectivity and favorable pharmacokinetics [49].
Resminostat exhibits broad inhibitory activity against Class I, IIb, and IV HDAC isoforms, with notable selectivity toward HDAC1, HDAC3, and HDAC6. This dual inhibitory action allows it to modulate both nuclear histone acetylation and cytoplasmic deacetylation processes. Mechanistically, Resminostat induces cell cycle arrest and apoptosis by upregulating p21^Waf1/Cip1 and downregulating oncogenic transcription factors such as c-Myc. It also restores differentiation pathways and suppresses epithelial-to-mesenchymal transition (EMT), leading to decreased invasiveness in solid tumors [50,51].
Additionally, Resminostat exemplifies a well-balanced, broad-spectrum HDAC inhibitor with the capacity to modulate several oncogenic signaling pathways simultaneously. Its distinctive chemical scaffold—combining a pyrrole-based CAP group, optimized linker, and a hydroxamic acid ZBG—contributes to a pharmacological profile that supports both potency and tolerability. Importantly, Resminostat’s progression from preclinical models to clinical evaluation underscores its translational relevance, particularly in malignancies characterized by profound epigenetic dysregulation. The most compelling evidence has emerged in hepatocellular carcinoma and multiple myeloma, where Resminostat has demonstrated clinically meaningful activity, reinforcing the therapeutic value of multi-HDAC targeting approaches in complex, treatment-resistant cancers.

3. Zinc Chelation Mechanism Across Diverse HDAC Isoforms

The structural architecture of HDACi inhibitors follows a highly conserved pharmacophore model consisting of three essential domains: a surface-interacting cap (CAP), a channel-penetrating linker, and a metal-chelating ZBG [29]. The structural architecture of HDACi is characterized by a conserved pharmacophore model composed of three key elements: a CAP, a channel-penetrating linker, and a metal-chelating ZBG. While hydroxamic acids have traditionally served as the classical ZBG due to their strong affinity for zinc, their clinical translation is often limited by mutagenicity concerns and off-target toxicities [29,52]. Consequently, the design of next-generation HDACi increasingly emphasizes the development of alternative ZBGs capable of maintaining high potency while improving safety profiles. The in silico approaches—including virtual screening, docking, molecular dynamics, and predictive modeling—are becoming indispensable tools in the design of new or next-generation HDAC inhibitors. These methods accelerate the identification of promising scaffolds, clarify zinc-binding mechanisms at the molecular level, and support the optimization of potency, selectivity, and drug-like properties before synthesis. By integrating computational strategies with experimental validation, researchers can more efficiently advance HDACi candidates with enhanced safety and translational potential [53].
As illustrated in Figure 2, the CAP region typically contains an aromatic or heteroaromatic moiety that anchors the inhibitor at the rim of the HDAC active-site pocket, providing isoform-tuning opportunities through steric and electronic modifications. The linker segment, often a flexible aliphatic or semi-rigid aromatic chain, serves as the structural spine that positions the terminal ZBG deep within the catalytic channel [29,54]. Finally, the ZBG—shown in the figure as a classical hydroxamate—coordinates the catalytic Zn2+ ion and dictates both potency and isoform selectivity.
In the context of this review, the pharmacophore model illustrated in Figure 2 does more than depict the classical CAP–linker–ZBG arrangement; it provides a conceptual scaffold that allows for a critical evaluation of how recent HDAC inhibitor designs succeed—or fail—to overcome longstanding limitations in this drug class. By examining newly reported inhibitors through this structural lens, we highlight not only how modifications in each domain affect potency and HDAC isoform preference, but also how certain strategies address persistent challenges such as metabolic instability, off-target toxicity, and poor tumor selectivity. Notably, the growing shift toward non-hydroxamate ZBGs, including hydrazides, benzamides, and emerging heteroatom-based chelators, reflects an intentional move away from the well-known drawbacks of hydroxamic acids. This trend demonstrates a deeper medicinal chemistry rationale: enhancing pharmacokinetic stability and therapeutic tolerability without compromising zinc-binding efficiency. Thus, the figure serves as a visual framework that anchors our critical discussion of recent SAR data and underscores emerging design principles shaping next-generation HDAC inhibitors, particularly those aimed at achieving isoform selectivity, reducing systemic toxicity, and improving clinical translation potential.
The success of early-stage clinical candidates established the hydroxamate functional group as the dominant and most widely utilized ZBG for HDACi. As exemplified by the co-crystal structure of HDAC2 with Vorinostat (SAHA) (PDB: 4LXZ), Figure 3A, the hydroxamate moiety functions as a potent, bidentate ligand capable of chelating the catalytic Zn2+ ion within the HDAC active site with high affinity. This could illustrate the critical coordination chemistry: the two oxygen atoms of the hydroxamate group coordinate directly to the zinc ion, forming a stable five-membered chelate ring. This chelation effectively blocks the enzyme’s hydrolytic deacetylase activity by sequestering the Zn2+ cofactor, which is essential for activating the water molecule required for substrate turnover. This specific, high-affinity zinc chelation mechanism is central to the efficacy of seminal HDACi compounds such as Vorinostat, Belinostat, and Panobinostat [49,55].
The subsequent co-crystal structures of Vorinostat bound to the catalytic sites of other isoforms, such as HDAC6 (Figure 3B) and HDAC8 (Figure 3C), further confirm this conserved inhibitory principle. Despite the distinct cellular localizations and primary functions of these two enzyme classes—HDAC2 is a nuclear enzyme involved in gene silencing, while HDAC6 is predominantly cytoplasmic and targets non-histone substrates like alpha-tubulin, the core interaction remains structurally identical [56]. In both cases, the hydroxamate ZBG of Vorinostat acts as a potent, bidentate ligand that consistently chelates the essential catalytic Zn2+ ion. Furthermore, this chelation efficiently blocks the active site tunnel and prevents the zinc ion from initiating the hydrolysis of the natural acetyl-lysine substrate. The conservation of the deep, narrow active site pocket geometry across different HDAC classes is precisely what allows this structurally simple hydroxamate ZBG to achieve its characteristic broad-spectrum (pan-inhibitory) activity.
Finally, these clinical studies underscore the therapeutic potential and evolving diversity of HDACi. The progress across different chemical scaffolds highlights the ongoing efforts to balance potency, isoform selectivity, and safety—ultimately guiding the next generation of HDAC-targeted anticancer agents.

4. Hydroxamate-Based HDACi

The success of early-stage clinical candidates established the hydroxamate functional group as the dominant and most widely utilized ZBG for HDACi [30,57,58]. Hydroxamates function as highly potent, bidentate zinc-chelating ligands that anchor deeply within the HDAC catalytic pocket, forming a stable coordination complex with the catalytic Zn2+ ion and thereby efficiently suppressing enzymatic deacetylase activity. This robust zinc-binding capability underpins their presence in several first-generation HDAC inhibitors, including Vorinostat (SAHA), Belinostat, and Panobinostat, all of which leverage the strong, dual-oxygen coordination of the hydroxamate motif to achieve broad-spectrum HDAC inhibition. While this interaction confers substantial potency, it is also associated with well-recognized drawbacks—such as metabolic instability, off-target metal chelation, and dose-limiting toxicity—which collectively motivate the search for alternative ZBGs in next-generation HDAC inhibitor design [59,60]. Notably, compounds like Pracinostat and Abexinostat have advanced to Phase II or III clinical trials (Table 1), showing broader activity against solid tumors, leukemia, and lymphomas [61,62].
The clinical development landscape of hydroxamate-based HDAC inhibitors continues to mature, as evidenced by the advancement of multiple agents across early- and mid-stage clinical studies and the clear documentation of their progress through registered ClinicalTrials.gov identifiers (Table 1). Resminostat (4SC AG), for instance, has completed Phase II evaluation in cutaneous lymphomas (NCT02953301) [63,64], while Pracinostat (Helsinn Healthcare SA) has advanced through Phase II/III trials in acute myeloid leukemia (NCT03151408) [62,65,66]. Abexinostat, developed by Xynomic Pharmaceuticals, has likewise reached Phase I/II stages in non-Hodgkin lymphoma (NCT04024696) [35]. Recently developed agents such as Bisthianostat from Shanghai Therion Pharmaceutical (Phase I; NCT03618602) [67,68] and the pan-HDAC inhibitor Ivaltinostat (CrystalGenomics), which has undergone Phase I/II studies in pancreatic adenocarcinoma (NCT05249101), further highlight ongoing innovation within this class [69,70]. Additional inhibitors—including Quisinostat (Janssen; Phase Ib–IVa, NCT01486277) [71,72,73], CUDC-101 and CUDC-907 (Curis Inc.; Phase I, NCT01384799 and NCT01742988, respectively) [74,75,76,77,78,79], CHR-3996 (Chroma Therapeutics; Phase I, NCT00697879) [80,81], and MPT0E028 (Taipei Medical University; Phase I, NCT02350868)—have generated essential pharmacokinetic, tolerability, and preliminary efficacy data supporting their translational relevance [82,83]. More recently, REC-2282 (AR-42) from Recursion Pharmaceuticals has completed Phase II/III clinical evaluation in neurofibromatosis type 2 [84,85], while R-306465 (Johnson & Johnson; Phase I, NCT00677001) has demonstrated activity in advanced solid tumors [86,87]. The integration of these trial identifiers within the review enhances scientific transparency and underscores the growing clinical validation of hydroxamate-based HDAC inhibitors, collectively reflecting a pipeline that is both diversified and steadily advancing toward therapeutic application.
Table 1. Representative Hydroxamate-Based HDACi in Clinical Trials: Structures, Company, Clinical Trials, HDAC Targets, and Cancer Applications.
Table 1. Representative Hydroxamate-Based HDACi in Clinical Trials: Structures, Company, Clinical Trials, HDAC Targets, and Cancer Applications.
Name and StructureCompanyClinical
Trial (ClinicalTrials ID)
Cancer TypeHDAC ClassRef.
Resminostat
Cells 14 01997 i001
4SC AGPhase II completed
(NCT02953301)
Mycosis Fungoides, Sézary Syndrome, lymphomaI & II[63,64]
Pracinostat a
Cells 14 01997 i002
Helsinn Healthcare SAPhase II/III; Ongoing/Completed (NCT03151408)Acute Myeloid LeukemiaI, II,& IV[62,65,66]
Abexinostat
Cells 14 01997 i003
Xynomic PharmaceuticalsPhase I/II; Completed/Ongoing (NCT04024696) non-Hodgkin lymphomaI & II[35]
Bisthianostat
Cells 14 01997 i004
Shanghai Theorion PharmaceuticalPhase I completed
(NCT03618602)
Myelomapan-HDAC[67,68]
Quisinostat
Cells 14 01997 i005
Janssen Research & DevelopmentPhase Ib-IVa
Completed (NCT01486277)
Cutaneous T-cell LymphomaI & II[71,72,73]
Ivaltinostat b
Cells 14 01997 i006
CrystalGenomicsPhase I/II Completed/Ongoing
(NCT05249101)
Pancreatic Adenocarcinomapan-HDAC[69,70]
CUDC-101
Cells 14 01997 i007
Curis, Inc.Phase I completed
(NCT01384799)
Head & Neck CancerI & II[77,78,79]
CUDC-907
Cells 14 01997 i008
Curis, Inc.Phase I completed
(NCT01742988)
LymphomaI & II[74,75,76]
CHR 3996
Cells 14 01997 i009
Chroma TherapeuticsPhase I completed
(NCT00697879)
Solid tumorsI[80,81]
MPT0E028
Cells 14 01997 i010
Taipei Medical UniversityPhase I completed
(NCT02350868)
colorectal cancer and B-cell lymphoma HDAC I, 2 & 6[82,83]
REC-2282 (AR-42)
Cells 14 01997 i011
Recursion PharmaceuticalsPhase II & III completed
(NCT02350868)
Neurofibromatosis Type 2HDAC I and IIb[84,85]
R-306465
Cells 14 01997 i012
Johnson & Johnson PharmaceuticalPhase I completed
(NCT00677001)
Advanced Solid
tumors
HDAC I[86,87]
a in Combination with Azacitidine; b in combination with Capecitabine or Capecitabine.
The hydroxamate-based HDACi included in this review share a conserved pharmacophore architecture consisting of a CAP group, a linker segment, and a ZBG. The hydroxamate moiety (–CONHOH), present across compounds such as Resminostat, Pracinostat, Bisthianostat, Ivaltinostat, and CUDC-series inhibitors, functions as a bidentate ZBG that strongly chelates the catalytic Zn2+ ion, facilitating broad and potent inhibition across HDAC classes I, II, and in some cases IIb and IV. The linker domain—typically a flexible aliphatic chain or an aromatic spacer—ensures proper alignment of the ZBG deep within the catalytic pocket while maintaining structural adaptability across various HDAC isoforms. Differences in linker length and rigidity, observed, for example, between the more extended linkers in CUDC-101 and CUDC-907 and the shorter, semi-rigid linkers in CHR-3996 and R-306465, correlate with differences in isoform preference and cytotoxic performance. The CAP region, which engages the rim of the active site, plays a dominant role in selectivity, cellular uptake, and pharmacokinetic behavior. Bulky aromatic CAP groups in agents such as Resminostat, Quisinostat, and CUDC-907 contribute to enhanced selectivity toward HDAC1/2/6 and improved interaction with surface residues, whereas heterocyclic CAP moieties found in CHR-3996 support tighter HDAC1-focused inhibition with reduced off-target effects. Likewise, the more voluminous CAP structures of pan-HDAC inhibitors such as Bisthianostat and Ivaltinostat enable broader engagement of class I and II enzymes, aligning with their wide-spectrum profiles in clinical evaluation. Overall, SAR trends across these inhibitors underscore the delicate interplay between CAP size and polarity, linker geometry, and hydroxamate ZBG positioning—each component collectively shaping HDAC isoform selectivity, potency, therapeutic window, and clinical translatability.
Many works were carried out by synthesizing compounds containing this scaffold with potent activities on HDAC isoenzymes [88,89,90]. Cursaro et al. designed and synthesized a series of vanillyl-based hydroxamic acid HDACi targeting HDAC6 and HDAC8, which are key epigenetic regulators implicated in aggressive neural tumors such as neuroblastoma and glioblastoma. SAR studies and molecular docking highlighted the influence of the CAP group (vanillyl), linker, and zinc-binding orientation on isoform selectivity. Among these synthesized derivatives, St.1 (Table 2) emerged as a highly potent and selective HDAC6 inhibitor (IC50 = 4.5 nM) [91].
In another study, a series of hydroxamic acid-based HDACi incorporating a tetrahydro-β-carboline core linked by an aliphatic chain was designed and synthesized. Among them, compound St.2 (Table 2) exhibited potent HDAC1 inhibition and strong antiproliferative activity across various tumor cell lines. Molecular docking studies supported its high binding affinity toward HDAC1, suggesting that this structural scaffold represents a promising framework for developing novel HDACi with potential efficacy against solid tumors [92]. In another study, two novel dihydroxamate-based HDAC1 inhibitors were designed and evaluated through molecular docking, molecular dynamics simulations, and enzymatic assays. Compound St.3 (Table 2) highlighted strong interactions with HDAC1, including hydrogen bonding and zinc coordination, and exhibited higher binding affinity (−6.2 kcal/mol), greater complex stability. These findings highlight St.3 as a promising lead for developing selective HDAC1-targeted anticancer agents [93]. In a recent study, a new series of N-hydroxycinnamamide-based hydroxamic acid derivatives was synthesized as potent HDACi targeting THP-1 monocytic leukemia cells. Among 22 compounds (including 20 novel ones), St.4 (Table 2) showed the strongest antiproliferative and pro-apoptotic effects, comparable to vorinostat, and effectively altered cell-cycle progression. Both compounds inhibited HDAC class I and II with similar potency to vorinostat and increased acetylation of histones H2A, H2B, H3, and H4. Molecular modeling confirmed zinc chelation by the hydroxamate group, supporting their potential as dual-class HDACi for leukemia therapy [94].
In another recent study, a novel hydroxamate-based HDAC1-selective inhibitor, compound St.5 (Table 2), was designed and synthesized to combat castration-resistant prostate cancer (CRPC). This compound exhibited remarkable HDAC1 selectivity, leading to increased histone H3 acetylation at Lys9/27 and potent cytotoxicity. Mechanistic studies revealed that St.5 induces G1 phase arrest through Cyclin D1 downregulation and triggers intrinsic apoptosis via caspase-3 activation and an elevated Bax/Bcl-2 ratio. Additionally, in vivo, St.5 significantly reduced tumor growth (by 75%) in PC3 xenograft models without evident toxicity, highlighting it as a promising selective HDAC1 inhibitor for advanced prostate cancer therapy [95]. In another study, a series of (arylidene)aminoxy-based HDACi were synthesized using the SAHA scaffold to explore their potential against uveal melanoma (UM), a highly aggressive and treatment-resistant cancer. The compounds were evaluated for inhibitory activity against HDAC1, 3, 6, and 8, supported by molecular docking to elucidate binding interactions. Among them, the quinoline derivative St.6 (Table 2) emerged as the most potent, showing nanomolar HDAC6 inhibition and notable antiproliferative activity on UM cell lines at micromolar concentrations. VS13 also modulated HDAC target gene expression comparably to SAHA, indicating its promise as a lead compound for developing HDAC6-targeted therapies in uveal melanoma [96]. In another investigation, a series of thiazole-based hydroxamate derivatives was developed as novel HDACi with strong anticancer potential. Among them, compound St.7 was the most potent against HDAC1 and showed significant cytotoxicity against HepG2 cells, accompanied by increased acetylation of histones H3 and H4. These findings identify thiazole-based hydroxamates as promising scaffolds for further development of multimodal HDAC-targeted anticancer agents [97]. In another study work, a series of tetrahydro-β-carboline-based hydroxamate derivatives were synthesized as novel HDACi. Among them, St.8 (Table 2) displayed potent HDAC1 inhibition and strong antiproliferative activity against the A549 lung cancer cell line. Treatment with St.8 led to elevated acetylation of histones H3 and H4, G2/M phase arrest, increased ROS generation, and DNA damage, ultimately triggering apoptosis [98].
In another work, nigranoic acid (NA) and manwuweizic acid (MA), two triterpenoids identified as HDACi via docking-based virtual screening, were chemically modified to generate a series of derivatives with enhanced biological activity. Among them, hydroxamic acid derivatives exhibited improved inhibition of HDAC1/2/4/6 (with the lowest IC50 = 1.14 μM) while showing no effect on HDAC8. Notably, compound St.9 (Table 2) highlighted strong anti-inflammatory activity in J774A.1 macrophages by increasing histone acetylation, inhibiting IL-1β maturation, caspase-1 cleavage, and NLRP3 inflammasome activation, without compromising cell viability. These findings highlight compound St.9 as a promising low-cytotoxic anti-inflammatory HDAC inhibitor and provide a valuable scaffold for probing the link between HDACs and inflammasome regulation [99].
A novel coumarin–hydroxamate hybrid (St.10; Table 2) was designed through a pharmacophore fusion strategy to enhance HDAC1 inhibition and overcome the limitations of current HDACi in breast cancer therapy. Molecular docking confirmed the strong binding affinity of St.10 to HDAC1, while in vitro and in vivo studies highlighted its potent anti-breast cancer activity with no systemic toxicity [100].
A study containing two hydroxamic acid derivatives, St.11 and St.12 Table 2, explored their potential as selective HDAC6 inhibitors for treating liver fibrosis. Chronic liver injury often progresses to fibrosis and cirrhosis due to persistent hepatitis, with limited effective therapies available. In this work, molecular docking, enzyme inhibition, and cellular assays were conducted to assess the specificity and anti-fibrotic effects of DR-3 and FDR2. Both compounds showed strong selectivity for HDAC6 over HDAC1, significantly reducing hepatic stellate cell (HSC) activation, fibrogenic gene expression, and collagen deposition [89].
In another work, a series of 3-substituted quinoline hydroxamic acids was synthesized and tested against A549 and HCT116 cancer cells and HDAC isoforms (1, 2, 6, 8). Substitution at C3 favored HDAC6 selectivity. Compound St.13 showed potent antiproliferative activity and nanomolar HDAC6 inhibition. Mechanistic studies revealed activation of caspase-dependent apoptosis, increased H2AX phosphorylation, and sub-G1 cell-cycle arrest, confirming HDAC6-targeted anticancer potential [101]. Recently, a series of imidazolyl-hydroxamic acid Schiff base derivatives was synthesized and evaluated as HDACi. Among them, St.14 (Table 2) showed the strongest anticancer activity across multiple cell lines, especially A2780 ovarian cancer, by inducing apoptosis, reducing colony formation and migration, and exhibiting antiangiogenic effects comparable to Vorinostat. In vivo, St.14 (50 mg/kg) outperformed Vorinostat in the 4T1 breast cancer mouse model. In silico studies confirmed stable HDAC–ligand interactions, suggesting St.14 as a potent lead for solid tumor therapy [102].
In another study, a novel series of 11 urushiol-based hydroxamic acid HDACi was designed and synthesized. These compounds displayed potent inhibition of class I HDACs (HDAC8, IC50 = 16–41 nM), with minimal activity against HDAC6. Docking studies highlighted key interactions contributing to HDAC8 inhibition, while Western blot analyses confirmed enhanced acetylation of histone H3 and SMC3, supporting selectivity for class I HDACs. St.15 Table 2 was the most active urushiol-based hydroxamates and represents promising scaffolds for further development as anticancer agents [103].
A new series of 2H-1,2,3-triazole-based hydroxamate analogs, inspired by vorinostat, was synthesized and evaluated for anticancer potential. Among 15 designed compounds, St.16 (Table 2) showed potent antiproliferative and cytotoxic effects against MCF-7 breast cancer cells with minimal toxicity toward normal HEK-293 cells. Mechanistic studies revealed apoptosis induction, S and G2/M phase arrest, and ROS generation. Molecular docking, HDAC inhibition assays, and MD simulations confirmed strong binding affinity toward HDAC1 and HDAC6 [104]. In another study, an HDAC6-selective inhibitor, St.17 (Table 2), was developed through a SAR approach for the treatment of T-cell prolymphocytic leukemia (T-PLL), an aggressive and currently incurable hematological malignancy. St.17 highlighted high potency and selectivity for HDAC6, with strong on-target activity and a favorable safety profile in non-transformed cells. In primary T-PLL patient cells, where HDAC6 is overexpressed, St.17 elicited robust biological responses, and combination studies revealed synergistic effects with approved anticancer agents, including bendamustine, idasanutlin, and venetoclax [105]. In another good work, a series of tetrahydro-β-carboline (THβC)-based hydroxamic acids was designed and synthesized as novel selective HDAC6 inhibitors using a scaffold-hopping strategy. Several analogues highlighted sub-nanomolar HDAC6 inhibition (IC50 < 5 nM) with strong selectivity. Molecular docking studies clarified that the SAR of St.18 (Table 2) was the most potent derivative and displayed favorable pharmacokinetic properties following oral administration in mice, highlighting its promise as a potent and selective HDAC6-targeted anticancer lead compound [106].
A series of estratriene-based hydroxamic acid derivatives was designed as HDACi using estrone and estradiol scaffolds linked via alkoxy chains. St.19 (Table 2) showed the best activity, exhibiting potent antiproliferative effects against HeLa. Molecular docking confirmed strong HDAC2 and HDAC6 binding through zinc coordination and favorable cap orientation, supporting these estratriene-based scaffolds as promising candidates for anticancer HDAC inhibitor development [107]. In another study, a series of hydroxamic acid-based HDAC6 inhibitors incorporating adamantane and natural terpene (camphane, fenchane) fragments as CAP groups were synthesized and evaluated through in silico, in vitro, and in vivo approaches. Among the synthesized compounds, St.20 showed the most potent HDAC6 inhibition and anti-β-amyloid aggregation properties, featuring an adamantane-amide-hydrocarbon scaffold, as a promising multitarget neuroprotective candidate for Alzheimer’s disease therapy [108].
Hydroxamate derivatives, including coumarin-, tetrahydro-β-carboline-, quinazolinone-, estratriene-, and urushiol-based scaffolds, have underscored potent and selective inhibition of various HDAC isoforms, especially HDAC1, HDAC6, and HDAC8, leading to cell cycle arrest, induction of apoptosis, and enhanced histone acetylation in multiple cancer models. SAR studies, supported by molecular docking and molecular dynamics simulations, have guided the optimization of zinc-binding groups, linkers, and surface recognition moieties to improve selectivity, potency, and pharmacokinetic profiles. In addition to classical cancer targets, dual- and multi-target inhibitors combining HDAC inhibition with PIM-1, PARP-1, IDO1, or ribonucleotide reductase inhibition have shown synergistic antiproliferative effects and promising in vivo efficacy.

5. Non-Hydroxamate-Based HDAC Inhibitors

HDACs have emerged as critical epigenetic targets in cancer therapy, neurodegenerative diseases, and immune-related disorders, prompting extensive research into both hydroxamate- and non-hydroxamate-based inhibitors. Non-hydroxamate HDACi, such as 2-aminobenzamides [109,110], cyclic peptides [111,112,113], thiols [114], coumarin-sulfonamide [115], and hydrazides [116], and other novel scaffolds, further expand the therapeutic potential by offering alternative mechanisms, reduced toxicity, and isoform-selectivity. Collectively, these studies highlight the versatility of HDAC-targeted small molecules as valuable frameworks for developing next-generation therapeutics with enhanced efficacy and safety profiles across diverse disease models [117]. Here, in this section, various works that do not contain a hydroxamate scaffold will be included accordingly.
Benzamide derivatives, including Entinostat, Chidamide, Tacedinaline, and Mocetinostat (Table 3), represent a newer generation of selective class I and class IIb HDACi [25,118]. These compounds generally possess improved safety and selectivity profiles, making them attractive for combination therapy. For instance, Entinostat and Chidamide have reached Phase III clinical trials for solid tumors and hematological malignancies; all clinically successful agents were listed in Table 3 with their chemical structures and targeted tumors. Structurally, these inhibitors share the canonical pharmacophore triad of CAP group, linker, and ZBG, which was presented before in Figure 2, but their variations in aromatic substituents and linker design strongly influence isoform selectivity and biological potency. All four compounds possess an ortho-aminobenzamide moiety as the ZBG, which coordinates with the catalytic zinc ion in a bidentate fashion while providing hydrogen bonding interactions within the enzyme pocket. This benzamide group binds less aggressively than hydroxamates, leading to enhanced selectivity toward HDAC1–3 and reduced off-target effects.
The CAP and linker regions are key determinants of activity and selectivity. Entinostat features a pyridine cap connected via a flexible carbamate linker, allowing optimal orientation at the rim of the catalytic site, while chidamide introduces fluorophenyl and pyridyl rings connected by a conjugated amide linker, improving rigidity, lipophilicity, and membrane permeability. Tacedinaline adopts a more compact diaryl amide framework, providing strong class I inhibition through a short, rigid linker that enhances zinc-pocket alignment. In contrast, mocetinostat incorporates heteroaromatic scaffolds in both the CAP and linker regions, improving hydrogen bonding and π-stacking interactions, thereby broadening its selectivity toward Class I and IV HDACs. Collectively, these structural variations illustrate the SAR principles governing benzamide HDACi—fine-tuning of cap polarity, linker rigidity, and ZBG geometry enables precise modulation of isoform selectivity, potency, and clinical performance.
A novel series of class I-selective HDACi containing 2-aminobenzamide zinc-binding groups linked to piperazinyl-pyrazine or piperazinyl-pyrimidine cores was synthesized and evaluated. Several compounds, including St.21 (Table 4), exhibited high selectivity for HDAC1, 2, and 3 over other HDAC isoforms and showed superior in vitro activity compared to clinically tested inhibitors such as Entinostat. Molecular docking and dynamics studies supported the observed structure–activity relationships [128]. In another amino-benzamide structure, the impact of incorporating a β-carboline cap into HDACi containing cinnamic acid linkers and benzamide zinc-binding groups was evaluated. A series of β-carboline–cinnamide conjugates was synthesized and evaluated for HDAC inhibition and antiproliferative activity against various human cancer cells. Most compounds exhibited superior HDAC inhibitory activity compared to the standard drug Entinostat, and St.22 (Table 4), showing notable potency against HCT-15 cells [129]. In a recent study, a series of novel benzamide derivatives with modified linker groups was synthesized as selective HDAC3 inhibitors. These compounds displayed potent antiproliferative activity against multiple cancer cell lines, with minimal toxicity toward normal human cells. Among them, St.23 (Table 4) showed significant HDAC3 selectivity (~47-fold over HDAC2), induced G2/M cell cycle arrest and apoptosis in 4T1 breast cancer cells, and underscored a favorable in vivo pharmacokinetic profile [130].
Selective HDAC6 inhibitors were designed using thiol as the ZBG to overcome the limitations of hydroxamate-based inhibitors, such as poor pharmacokinetics and potential genotoxicity. A series of thiol-based HDAC6 inhibitors was synthesized, and their SAR was analyzed via molecular docking. Notably, St.24 exhibited high selectivity (29-fold) while St.25 (Table 4) showed the highest activity against HDAC6. The use of a pyrimidine linker in these thiol-based inhibitors represents a novel scaffold, potentially offering improved pharmacokinetic properties and reduced genotoxicity compared to traditional hydroxamate derivatives [131]. In this study, selective Class I HDAC inhibition was explored using a cyclic peptide approach inspired by natural depsipeptides. Building on Largazole, a fluorinated analog of Romidepsin St.26 (Table 4) was synthesized in 12 steps. This analog showed potent inhibitory activity against Class I HDACs with negligible HDAC6 inhibition, confirming its high selectivity. This structure significantly inhibited the growth of NCI-H1975 and HT29 cancer cells while showing reduced cytotoxicity toward normal cell lines (WRL-68 and HEK293). Mechanistic studies underscored that St.26 induced cell-cycle arrest and apoptosis [132]. In recent work, 33 chrysin derivatives were synthesized and characterized to assess their HDAC inhibitory and anticancer activities. Among them, compound St.27 showed the strongest HDAC inhibition and was most selective toward HDAC8 [133].
In another work, a series of sulfur-based selective HDACi was developed by modifying ajoene, a natural compound from garlic, to overcome the toxicity issues associated with hydroxamic acid-based inhibitors. Structure–activity and docking studies revealed potent and highly selective HDAC8 inhibitors with a novel zinc-binding group. Among them, St.28 (Table 4) underscored strong antiproliferative activity against neuroblastoma cell lines and showed significant in vivo efficacy in a BE(2)-C xenograft mouse model, highlighting its promise as a new HDAC8-targeted anticancer agent [134].
A growing body of research has focused on developing non-hydroxamic HDACi to overcome the genotoxicity and poor pharmacokinetic limitations of hydroxamate-based agents [49]. Several alternative ZBGs have been explored, including thiol, carboxylate, sulfur-based, and aminobenzamide scaffolds, each offering enhanced selectivity and safety profiles. Thiol-based HDAC6 inhibitors, such as compound St.24, underscored notable potency and selectivity with potentially improved pharmacokinetics, while sulfur-containing derivatives like compound St.28 showed HDAC8 selectivity and in vivo antitumor efficacy in neuroblastoma models. Similarly, carboxyl-containing chrysin derivatives St.27 exhibited selective HDAC8 inhibition and antiproliferative activity against colon cancer cells. Among aminobenzamide analogs, several agents achieved class I or HDAC3 selectivity, strong apoptosis induction, and tumor growth suppression in vivo. Collectively, these studies highlight the promising therapeutic potential of non-hydroxamic HDACi with diverse ZBGs as safer, selective, and potent anticancer candidates.

6. HDAC-Based Dual-Target Inhibitors

Over the past five years, increasing attention from both academia and industry has been directed toward the development of isoform-selective HDACi and HDAC-based dual inhibitors (hydroximate and non-hydroximate-based scaffolds) that simultaneously modulate HDACs and complementary oncogenic pathways. This dual-targeting paradigm offers the potential to enhance therapeutic efficacy, reduce off-target effects, and overcome resistance mechanisms, leading to substantial progress and innovative designs in this rapidly evolving field [135,136]. A series of 3,4,5-trimethoxyphenyl-based hybrids was designed and synthesized by Mohamed et al. as dual EGFR/HDACi by combining pharmacophoric features of both targets. The derivatives containing hydroxamic acid showed the most potent anticancer activity against HepG2, MCF-7, HCT116, and A549 cell lines Table 5, and among them, compound St.29 exhibited remarkable EGFR inhibition (IC50 = 0.063 µM), comparable to staurosporine, and selective HDAC6 inhibition. Mechanistic studies confirmed its pro-apoptotic effect through increased Bax, caspase-3, and caspase-8 levels, along with reduced Bcl-2 expression and cell cycle arrest at G1/S phase [137]. In another work, a series of phenylurea hydroxamic acids was developed by combining the pharmacophores of HDAC and VEGFR-2 inhibitors. Among them, St.30 (Table 5) effectively inhibited HDAC and showed modest VEGFR-2 activity. Molecular docking revealed that the hydroxamic acid moiety maintains zinc coordination in HDAC while interacting with key residues in the VEGFR-2 active site [138].
In a related study, a new series of phthalazinone derivatives was designed and synthesized to target PARP-1 alone or both PARP-1 and HDAC1 simultaneously. Several derivatives displayed remarkable enzyme inhibition, with DLC-1 showing exceptional PARP-1 potency and strong antiproliferative effects across breast cancer cell lines, inducing G1 phase arrest and apoptosis. Among the dual inhibitors, St.31 (Table 5) exhibited potent activity against both PARP-1 and HDAC1. These results highlight phthalazinone-based hybrids as promising scaffolds for developing dual PARP-1/HDAC1 inhibitors with strong anticancer potential [139]. A series of dual PIM-1/HDACi was designed and synthesized based on a 3-cyanopyridine-hydroximate scaffold (zinc-binding groups). Several compounds exhibited broad-spectrum anticancer activity across the NCI-60 cancer cell line panel. Notably, hydroxamic acid St.32 (Table 5) showed potent antiproliferative effects and displayed strong activity (GI50 ≤ 3 μM) and inducing G2/M cell cycle arrest and pre-G1 apoptosis. Particularly, St.32, a promising dual PIM-1/HDACi with potential for further in vivo development in anticancer therapy [140].
In numerous studies, researchers have sought to design and synthesize diverse series of compounds exhibiting dual inhibitory activity against both HDAC and tubulin [6,141,142]. In another work on dual inhibitors, researchers developed a new class of dual tubulin/HDAC inhibitors derived from the natural product millepachine, a known tubulin polymerization inhibitor. Their study reported the synthesis and evaluation of several analogs, among which compound St.33 (Table 5) demonstrated exceptional potency against PC-3 prostate cancer cells, and effectively inhibited both microtubule polymerization and HDAC activity, leading to G2/M cell cycle arrest and pronounced apoptosis. In a PC-3 xenograft model, St.33 achieved a remarkable tumor inhibition rate of 90.07% at 20 mg/kg, significantly outperforming the reference compound CA-4. The compound also displayed favorable in vivo drug metabolism characteristics [141].
A series of phthalazino [1,2-b]-quinazolinone-based hybrids bearing ortho-aminoanilide or hydroxamic acid groups was developed as a multi-target HDACi for solid tumor therapy. Among them, St.34 showed nanomolar inhibitory potency against cancer cells and HDAC subtypes, surpassing SAHA (vorinostat). Mechanistic studies revealed that St.34 enhanced histone H3 and α-tubulin acetylation and activated the p53 signaling pathway, leading to HepG2 cell growth inhibition [143]. A 1,10-phenanthroline-based hydroxamate derivative, St.35 (Table 5), was synthesized accordingly and identified as a dual HDAC/RR inhibitor with notable anticancer potential. Dual inhibition induced ROS-mediated apoptosis, confirming synergistic action. Docking and molecular dynamics revealed strong binding to HDAC isoforms, especially HDAC7 (–9.63 kcal/mol), surpassing SAHA. These findings highlight St.35 as a promising dual-target anticancer scaffold for further development [144].
Table 5. The structures of HDAC-based dual-target inhibitors and their IC50 values against a panel of cancer cell lines, HDAC enzymes, and other targets for the most active agents.
Table 5. The structures of HDAC-based dual-target inhibitors and their IC50 values against a panel of cancer cell lines, HDAC enzymes, and other targets for the most active agents.
CodeStructuresEvaluated Cancer/Normal Cell LinesEvaluated TargetRef.
Cell LinesIC50/%VHDACIC50
St.29Cells 14 01997 i046MCF-7
HepG2
HCT116
A549
0.621 µM
0.536 µM
1.206 µM
0.797 µM
HDAC1
HDAC2
HDAC4
HDAC6
HDAC8
EGFR
0.148 µM
0.168 µM
5.852 µM
0.060 µM
2.257 µM
63 nM
[137]
St.30Cells 14 01997 i047EPC1.0 µMHDAC6
EPCs
166 nM
1.0 µM
[138]
St.31Cells 14 01997 i048MDA-MB-436
MDA-MB-231
MCF-7
0.30 µM
2.70 µM
2.41 µM
HDAC1
PARP-1
31 nM
<0.2 nM
[139]
St.32Cells 14 01997 i049HeLa65.94 nMHDAC1
HDAC6
PIM-1
63.65 nM
62.39 nM
343.87 nM
[140]
St.33Cells 14 01997 i050PC-316 nMHDAC2
Tubulin
0.43 µM
4.82 µM
[141]
St.34Cells 14 01997 i051HepG20.58 µMHDAC1
HDAC6
13.37 nM
42.74 nM
[143]
St.35Cells 14 01997 i052SiHa
HepG2
MCF-7
Cal27
16.43 µM
>100 µM
>100 µM
50.98 µM
HDACs
RR
10.80 µM
9.34 µM
[144]
St.36Cells 14 01997 i053SU-DHL-61.20 µMHDAC1
HDAC4
HDAC6
HDAC11
EZH2wt
0.19 µM
>10 µM
0.03 µM
>10 µM
0.59 nM
[145]
St.37Cells 14 01997 i054HCT-116
SW480
MDA-MB-231
MCF-7
HK2
16.42 µM
23.43 µM
43.74 µM
36.46 µM
>80 µM
HDAC1
HDAC6
IDO1
1.078 µM 58.23 nM
86 nM
[146]
St.38Cells 14 01997 i055MDA-MB-231
HeLa
HepG2
2.47 µM
1.51 µM
4.52 µM
HDAC1
HDAC2
HDAC3
CDK9
1.73 µM
>50 µM
1.11 µM
0.17 µM
[147]
St.39Cells 14 01997 i056JEKO-10.9 µMHDAC1
HDAC2
HDAC3
PI3K α
PI3K β
75.5 nM
70.9 nM
1.9 nM
2.5 nM
10.0 nM
[148]
In a recent study, researchers developed a novel class of tazemetostat-derived HDAC/EZH2 dual inhibitors to exploit the therapeutic synergy of co-targeting epigenetic regulators in hematological malignancies. Among the synthesized compounds, St.36 (Table 5) showed outstanding dual-inhibitory potency, displaying strong antiproliferative activity against EZH2-mutant DLBCL and multiple AML cell lines [145]. In a recent study, a series of imidazothiazole-based hydroxamic acid derivatives was developed as dual IDO1/HDAC6 inhibitors using a pharmacophore fusion strategy. Among them, St.37 (Table 5) emerged as the most potent, exhibiting strong IDO1 inhibition and high HDAC6 selectivity over other isoforms, validated by Western blot. Docking studies suggested favorable binding to both IDO1 and HDAC6. Biologically, St.37 induced G2/M cell cycle arrest in HCT-116 cells and showed significant in vivo antitumor activity in CT26 tumor-bearing mice with minimal toxicity [146].
Several recent studies have focused on the development of dual HDAC/CDK inhibitory agents [147,149]. In this study, a series of dual CDK9/HDACi based on N-(2-aminophenyl)-5-(4-aryl-pyrimidin-2-yl)amino-1H-indole-2-carboxamide scaffolds was designed and synthesized to target transcriptional dysregulation in cancer. Among them, St.38 (Table 5) exhibited strong antiproliferative activity across multiple cancer cell lines and potently inhibited CDK9, HDAC1, and HDAC3. It induced mitochondria-mediated apoptosis and G2/M arrest, outperforming reference inhibitors AZD-5438 and Mocetinostat. In vivo, St.38 significantly reduced MDA-MB-231 tumor growth by 76.8%, establishing it as a promising dual CDK9/HDAC inhibitor with potent anticancer potential [147]. In another study, a series of dual PI3K/HDACi was designed by integrating the morpholino-triazine core of the PI3K inhibitor with the hydrazide moiety of an HDAC1–3 selective inhibitor. Among them, St.39 (Table 5) underscored potent inhibition of both PI3K isoforms (IC50 = 2.5–80.5 nM) and HDAC1–3 (IC50 = 1.9–75.5 nM), exhibiting strong antiproliferative effects across multiple cancer cell lines. In mantle cell lymphoma (Jeko-1) cells, and induced markedly better apoptosis than either single-target inhibitor [148].

7. Conclusions

HDACi remain one of the most dynamic and rapidly evolving classes of epigenetic anticancer agents, with several FDA-approved drugs establishing HDACs as clinically validated therapeutic targets. Nevertheless, the majority of clinically advanced HDACi continue to rely on classical hydroxamate or non-hydroximate ZBGs, which, despite their strong zinc-chelating potency, are limited by metabolic instability, suboptimal pharmacokinetics, and dose-limiting toxicities. This review systematically evaluated HDACi reported in recent years, highlighting key innovations across the three essential pharmacophore components—CAP region engineering, linker optimization, and ZBG diversification. Comparative assessment of biological activities across cancer models and HDAC isoforms reveals clear medicinal chemistry trends that are redefining the design of next-generation HDACi. Structural analyses show that tailoring the CAP group and fine-tuning linker rigidity can substantially enhance potency and selectivity, even when employing weaker but safer ZBGs. Growing evidence indicates that strategic ZBG diversification, rather than simply maximizing Zn2+ affinity, represents the most promising pathway toward generating HDACi with improved drug-like properties, reduced off-target effects, and greater clinical feasibility.
Importantly, recent advances demonstrate that computational and in silico methodologies—including ligand-based virtual screening, docking studies, molecular dynamics simulations, and predictive pharmacokinetic modeling—offer powerful tools to accelerate HDACi discovery. These approaches provide deep insights into binding interactions, isoform selectivity, ZBG behavior within zinc-containing catalytic pockets, and the overall structural determinants of activity. By integrating these computational techniques with experimental validation, researchers can rapidly prioritize promising scaffolds, refine molecular designs, and more efficiently advance candidates with genuine translational potential.
As the field progresses, the synergy between innovative ZBG chemotypes, isoform-selective CAP motifs, optimized linker architectures, and advanced computational design strategies is poised to deliver the next generation of HDAC inhibitors. These emerging compounds are expected to be not only potent but also safer, more selective, and more clinically viable—marking a significant step forward in the development of durable epigenetic therapies for cancer and other diseases.

Funding

This research received no external funding.

Data Availability Statement

This is a Review article, and there are no new data in this manuscript.

Acknowledgments

The author wishes to thank An-Najah National University (www.najah.edu) for their support in carrying out this work.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
HDACsHistone deacetylases
HDACiHistone deacetylase inhibitors
FDAThe Food and Drug Administration of the United States
ZBGzinc-binding group
SARstructure–activity relationships
CDKcyclin-dependent kinase
PI3KPhosphoinositide 3-kinase
EGFRMultidisciplinary Digital Publishing Institute
DNADeoxyribonucleic Acid
DMDDuchenne muscular dystrophy
HCChepatocellular carcinoma
EMTEpithelial-to-mesenchymal transition
PDBProtein Data Bank
HLHodgkin Lymphoma
HepG2Liver Cancer cell lines
MCF-7Breast Cancer cell lines
HCT116Colon Cancer cell lines
A549Lung cancer cell lines
St.Structure
IC50inhibitory concentration
GI50half-maximum growth inhibition
µMMicro molar
nMNano molar
SK-MEL-2Male malignant melanoma
HS-5Fibroblast morphology Cell
NANot applicable
MDA-MB-436Human breast adenocarcinoma Cells
MDA-MB-231Human breast adenocarcinoma Cells
PC3Prostate Cancer Cell line
PDXOPatient-Derived Xenograft Organoids
UMUveal Melanoma
PC9human pulmonary adenocarcinoma cell lines
J774ASarcoma Cell lines
4T1metastatic mouse mammary carcinoma cell line
LX-2Human Hepatic Stellate Cell Line
HeLaCervical Cancer cell lines
EPCEndothelial Progenitor Cells
HT-29Colon cancer cell lines
PIM-1Proto-oncogene, serine/threonine kinase
VEGFRvascular endothelial growth factor receptor
IDO1Indoleamine 2,3-dioxygenase
SH-SY5YNeuroblastoma cell line
EZH2wtwild-type of the Enhancer of Zeste Homolog 2 protein

References

  1. Sharma, S.; Kumar, C.; Kushwaha, H.; Jha, S.K.; Chawla, S.; Sharma, A.; Midha, T.; Huddar, V. Advancing anticancer drug development: Overcoming challenges and exploring new therapeutic strategies. Ayush J. Integr. Oncol. 2025, 2, 8–27. [Google Scholar] [CrossRef]
  2. Zafar, A.; Khatoon, S.; Khan, M.J.; Abu, J.; Naeem, A. Advancements and limitations in traditional anti-cancer therapies: A comprehensive review of surgery, chemotherapy, radiation therapy, and hormonal therapy. Discov. Oncol. 2025, 16, 607. [Google Scholar] [CrossRef] [PubMed]
  3. Abavisani, M.; Khoshrou, A.; Eshaghian, S.; Karav, S.; Sahebkar, A. Overcoming antibiotic resistance: The potential and pitfalls of drug repurposing. J. Drug Target. 2025, 33, 341–367. [Google Scholar] [CrossRef]
  4. Khan, N.; Raza, U.; Zaidi, S.A.A.; Nuer, M.; Abudurousuli, K.; Paerhati, Y.; Aikebaier, A.; Zhou, W. Drugging the ‘undruggable’KRAS: Breakthroughs, challenges, and opportunities in pancreatic cancer. Cancer Biol. Med. 2025, 22, 762–788. [Google Scholar]
  5. Patrick, G.L. An Introduction to Medicinal Chemistry, 7th ed.; Oxford University Press: Oxford, UK, 2023. [Google Scholar]
  6. Tran, C.; Hamze, A. Recent Advancements in the Development of HDAC/Tubulin Dual-Targeting Inhibitors. Pharmaceuticals 2025, 18, 341. [Google Scholar] [CrossRef]
  7. Hawash, M. Highlights on Specific Biological Targets; Cyclin-Dependent Kinases, Epidermal Growth Factor Receptors, Ras Protein, and Cancer Stem Cells in Anticancer Drug Development. Drug Res. 2019, 69, 471–478. [Google Scholar] [CrossRef] [PubMed]
  8. Cheshmazar, N.; Hamzeh-Mivehroud, M.; Charoudeh, H.N.; Hemmati, S.; Melesina, J.; Dastmalchi, S. Current trends in development of HDAC-based chemotherapeutics. Life Sci. 2022, 308, 120946. [Google Scholar] [CrossRef] [PubMed]
  9. Talom, A.; Barhoi, A.; Jirpu, T.; Dawn, B.; Ghosh, A. Clinical progress and functional modalities of HDAC inhibitor-based combination therapies in cancer treatment. Clin. Transl. Oncol. 2025, 1–15. [Google Scholar] [CrossRef]
  10. Shirbhate, E.; Singh, V.; Kore, R.; Koch, B.; Veerasamy, R.; Tiwari, A.K.; Rajak, H. Synergistic strategies: Histone deacetylase inhibitors and platinum-based drugs in cancer therapy. Expert Rev. Anticancer Ther. 2025, 25, 121–141. [Google Scholar] [CrossRef]
  11. Márquez-Cantudo, L.; Ramos, A.; Coderch, C.; de Pascual-Teresa, B. Proteasomal degradation of Zn-dependent Hdacs: The E3-ligases implicated and the designed protacs that enable degradation. Molecules 2021, 26, 5606. [Google Scholar] [CrossRef]
  12. Peterson, J.J.; Lewis, C.A.; Burgos, S.D.; Manickam, A.; Xu, Y.; Rowley, A.A.; Clutton, G.; Richardson, B.; Zou, F.; Simon, J.M. A histone deacetylase network regulates epigenetic reprogramming and viral silencing in HIV-infected cells. Cell Chem. Biol. 2023, 30, 1617–1633.E9. [Google Scholar] [CrossRef]
  13. Mohite, R.; Doshi, G. Elucidation of the role of the epigenetic regulatory mechanisms of PI3K/Akt/mTOR signaling pathway in human malignancies. Curr. Cancer Drug Targets 2024, 24, 231–244. [Google Scholar] [CrossRef]
  14. Tokarz, P.; Kaarniranta, K.; Blasiak, J. Role of the cell cycle re-initiation in DNA damage response of post-mitotic cells and its implication in the pathogenesis of neurodegenerative diseases. Rejuvenation Res. 2016, 19, 131–139. [Google Scholar] [CrossRef]
  15. Pu, J. The Role of Histone Deacetylase (HDAC) Inhibitors and Cytokine-Induced Killer Cell (CIK) in Multiple Myeloma. Ph.D. Thesis, Universitäts-und Landesbibliothek Bonn, Bonn, Germany, 2025. [Google Scholar]
  16. Su, M.; Gong, X.; Liu, F. An update on the emerging approaches for histone deacetylase (HDAC) inhibitor drug discovery and future perspectives. Expert Opin. Drug Discov. 2021, 16, 745–761. [Google Scholar] [CrossRef] [PubMed]
  17. Theodoropoulou, M.A.; Mantzourani, C.; Kokotos, G. Histone Deacetylase (HDAC) Inhibitors as a Novel Therapeutic Option Against Fibrotic and Inflammatory Diseases. Biomolecules 2024, 14, 1605. [Google Scholar] [CrossRef]
  18. Tarawneh, A.H.; Al-Trawneh, S.A.; Yesiloglu, T.Z.; Zessin, M.; Robaa, D.; Barinka, C.; Schutkowski, M.; Sippl, W.; Ross, S.A. Novel selective indole based histone deacetylase 10 inhibitors as anticancer therapeutics. Sci. Rep. 2025, 15, 33307. [Google Scholar] [CrossRef]
  19. Reichert, N.; Choukrallah, M.-A.; Matthias, P. Multiple roles of class I HDACs in proliferation, differentiation, and development. Cell. Mol. Life Sci. 2012, 69, 2173–2187. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Wang, H.; Zhan, Z.; Gan, L.; Bai, O. Mechanisms of HDACs in cancer development. Front. Immunol. 2025, 16, 1529239. [Google Scholar] [CrossRef] [PubMed]
  21. Cappellacci, L.; Perinelli, D.R.; Maggi, F.; Grifantini, M.; Petrelli, R. Recent progress in histone deacetylase inhibitors as anticancer agents. Curr. Med. Chem. 2020, 27, 2449–2493. [Google Scholar] [CrossRef] [PubMed]
  22. Bakrim, S.; Atifi, F.; Omari, N.E.; Zaid, Y.; Aanniz, T.; Lee, L.H.; Zengin, G.; Bouyahya, A. Clinical Applications of HDAC Inhibitors as Anticancer Agents in Prostate, Breast, Ovarian, and Cervical Cancers. ChemistrySelect 2025, 10, e202405484. [Google Scholar] [CrossRef]
  23. Ruzic, D.; Djoković, N.; Srdić-Rajić, T.; Echeverria, C.; Nikolic, K.; Santibanez, J.F. Targeting histone deacetylases: Opportunities for cancer treatment and chemoprevention. Pharmaceutics 2022, 14, 209. [Google Scholar] [CrossRef] [PubMed]
  24. Abdelsalam, M.; Zmyslia, M.; Schmidtkunz, K.; Vecchio, A.; Hilscher, S.; Ibrahim, H.S.; Schutkowski, M.; Jung, M.; Jessen-Trefzer, C.; Sippl, W. Design and synthesis of bioreductive prodrugs of class I histone deacetylase inhibitors and their biological evaluation in virally transfected acute myeloid leukemia cells. Arch. Pharm. 2024, 357, 2300536. [Google Scholar] [CrossRef]
  25. Abdelraheem, A.M.; Aly, O.M.; Mohassab, A.M.; Mohamed, M.F. HDAC as a cancer target: Various scaffolds based on HDAC inhibitors and an overview of their advances. Octahedron Drug Res. 2024, 5, 22–50. [Google Scholar] [CrossRef]
  26. Raucci, A.; Zwergel, C.; Valente, S.; Mai, A. Advancements in Hydrazide-Based HDAC Inhibitors: A Review of Recent Developments and Therapeutic Potential. J. Med. Chem. 2025, 68, 14171–14194. [Google Scholar] [CrossRef]
  27. Zhao, C.; Chai, Q.; Zhang, Y. Recent progress of small molecular HDAC modulators for cancer therapy and beyond. Pharmacol. Discov. 2025, 5, 8. [Google Scholar] [CrossRef]
  28. Zhang, L.; Yang, Y.; Li, Y.; Wang, C.; Bian, C.; Wang, H.; Wang, F. Epigenetic regulation of histone modifications in glioblastoma: Recent advances and therapeutic insights. Biomark. Res. 2025, 13, 80. [Google Scholar] [CrossRef]
  29. Morales-Herrejón, G.; García-Vázquez, J.B.; Fernández-Pomares, C.; Bakalara, N.; Correa-Basurto, J.; Mendoza-Figueroa, H.L. Computationally Guided Design, Synthesis, and Evaluation of Novel Non-Hydroxamic Histone Deacetylase Inhibitors, Based on N-Trifluoroacetamide as a Zinc-Binding Group, Against Breast Cancer. Pharmaceuticals 2025, 18, 351. [Google Scholar] [CrossRef]
  30. Geurs, S.; Clarisse, D.; De Bosscher, K.; D’hooghe, M. The zinc-binding group effect: Lessons from non-hydroxamic acid vorinostat analogs. J. Med. Chem. 2023, 66, 7698–7729. [Google Scholar] [CrossRef]
  31. Li, Y.; Wang, F.; Chen, X.; Wang, J.; Zhao, Y.; Li, Y.; He, B. Zinc-dependent deacetylase (HDAC) inhibitors with different zinc binding groups. Curr. Top. Med. Chem. 2019, 19, 223–241. [Google Scholar] [CrossRef] [PubMed]
  32. Begum, R.; Parsons, J.L.; Jones, A.M. Adverse drug reaction profiles of histone deacetylase inhibitors. Sci. Rep. 2025, 15, 35880. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, L.; Zhang, J.; Jiang, Q.; Zhang, L.; Song, W. Zinc binding groups for histone deacetylase inhibitors. J. Enzym. Inhib. Med. Chem. 2018, 33, 714–721. [Google Scholar] [CrossRef]
  34. Inoue, Y.; Yasunaga, J.-I. Impact of Novel Agents on Allogeneic Hematopoietic Cell Transplantation in Patients with T-Cell Lymphomas. Cells 2025, 14, 1306. [Google Scholar] [CrossRef] [PubMed]
  35. Gui, L.; Xie, Z.; Qin, Y.; Liu, P.; Yang, J.; Chen, X.; Li, Z.; Tao, R.; Shi, Y. Safety, pharmacokinetics, and efficacy of abexinostat, an novel histone deacetylase inhibitor, in Chinese patients with relapsed/refractory B cell non-Hodgkin lymphoma: A Phase 1 study. BMC Cancer 2025, 25, 967. [Google Scholar] [CrossRef] [PubMed]
  36. Afifi, S.; Michael, A.; Azimi, M.; Rodriguez, M.; Lendvai, N.; Landgren, O. Role of Histone Deacetylase Inhibitors in Relapsed Refractory Multiple Myeloma: A Focus on Vorinostat and Panobinostat. Pharmacotherapy 2015, 35, 1173–1188. [Google Scholar] [CrossRef] [PubMed]
  37. Piper, W.L.; Waddell, J.A.; Solimando, D.A., Jr. Drug monographs: Belinostat and idelalisib. Hosp. Pharm. 2014, 49, 1009–1013. [Google Scholar] [CrossRef] [PubMed]
  38. Yadav, R.; Mishra, P.; Yadav, D. Histone deacetylase inhibitors: A prospect in drug discovery. Turk. J. Pharm. Sci. 2018, 16, 101. [Google Scholar] [CrossRef]
  39. Calvo, E.; Reddy, G.; Boni, V.; García-Cañamaque, L.; Song, T.; Tjornelund, J.; Choi, M.R.; Allen, L.F. Pharmacokinetics, metabolism, and excretion of 14C-labeled belinostat in patients with recurrent or progressive malignancies. Investig. New Drugs 2016, 34, 193–201. [Google Scholar] [CrossRef]
  40. Imai, Y.; Maru, Y.; Tanaka, J. Action mechanisms of histone deacetylase inhibitors in the treatment of hematological malignancies. Cancer Sci. 2016, 107, 1543–1549. [Google Scholar] [CrossRef]
  41. Van Veggel, M.; Westerman, E.; Hamberg, P. Clinical Pharmacokinetics and Pharmacodynamics of Panobinostat. Clin. Pharmacokinet. 2018, 57, 21–29. [Google Scholar] [CrossRef]
  42. Aartsma-Rus, A. Histone deacetylase inhibition with givinostat: A multi-targeted mode of action with the potential to halt the pathological cascade of Duchenne muscular dystrophy. Front. Cell Dev. Biol. 2025, 12, 1514898. [Google Scholar] [CrossRef]
  43. Lamb, Y.N. Givinostat: First approval. Drugs 2024, 84, 849–856. [Google Scholar] [CrossRef]
  44. Fiorentini, F.; Germani, M.; Del Bene, F.; Pellizzoni, C.; Cazzaniga, S.; Rocchetti, M.; Bettica, P. Population pharmacokinetic–pharmacodynamic analysis of givinostat. Expert Opin. Drug Metab. Toxicol. 2023, 19, 229–238. [Google Scholar] [CrossRef]
  45. Anjum, A.F.; Anjum, M.B.; ur Rehman, R. Unleashing the Potential of Givinostat: A Novel Therapy for Duchenne Muscular Dystrophy. Curr. Ther. Res. 2025, 102, 100787. [Google Scholar] [CrossRef]
  46. Kumar, D.; Kumar, R.; Singh, H.R.; Tanwar, R.; Gupta, V. Duvyzat (Givinostat) in Duchenne Muscular Dystrophy: Mechanisms, Clinical Impact, and Future Directions. Curr. Pharm. Res. 2025, 1, 308–322. [Google Scholar] [CrossRef]
  47. Galluzzi, L.; Spada, S. Epigenetic Regulation of Cancer-Part D; Academic Press: Cambridge, MA, USA, 2025; Volume 390. [Google Scholar]
  48. Hao, W.; Zhang, Q.; Ma, Y.; Ding, Y.; Zhao, C.; Tian, C. Mechanism and application of HDAC inhibitors in the treatment of hepatocellular carcinoma. J. Mol. Med. 2025, 103, 469–484. [Google Scholar] [CrossRef]
  49. Pires, G.S.; Tolomeu, H.V.; Rodrigues, D.A.; Lima, L.M.; Fraga, C.A.M.; Pinheiro, P.d.S.M. Drug Discovery for Histone Deacetylase Inhibition: Past, Present and Future of Zinc-Binding Groups. Pharmaceuticals 2025, 18, 577. [Google Scholar] [CrossRef]
  50. Shukla, Y.K.; Vandana; Mandal, V.; Asati, V.; Keservani, R.K.; Bharti, S.K. An updated patent review on rational combinations of HDAC inhibitors for cancer chemotherapy (2020–present): Part 2-patent published. Expert Opin. Ther. Pat. 2025, 35, 875–900. [Google Scholar] [CrossRef] [PubMed]
  51. Rizwan, A.; Aqeel, A.; Farooqi, H. Decoding HDACs and its inhibitors-artificial intelligence assisted smart software based super computational modelling technology in targeting cancer and neurological disorders of the brain. Netw. Model. Anal. Health Inform. Bioinform. 2025, 14, 104. [Google Scholar] [CrossRef]
  52. Al Rahim, N.A.A.; Mahmood, A.A.R.; Tahtamouni, L.H.; Bawadi, R.M.; Almasri, A.Y.; Hamad, M.A.; Hussein, N.A.; Yasin, S.R.; Saleh, A.M. Synthesis, In Silico, and Biological Evaluation of Non-Hydroxamate Benzoic Acid–Based Derivatives as Potential Histone Deacetylase Inhibitors (HDACi). Chem. Biodivers. 2025, e01492. [Google Scholar] [CrossRef]
  53. Baselious, F.; Hilscher, S.; Handke, L.; Barinka, C.; Schutkowski, M.; Sippl, W. In silico screening of a designed focused chemical space identifies novel alkyl hydrazides as potent HDAC11 inhibitors. Comput. Biol. Med. 2025, 196, 110695. [Google Scholar] [CrossRef] [PubMed]
  54. Olaoye, O.O.; Erdogan, F.; Gracia-Hernandez, M.; Garcha, H.K.; Sedighi, A.; Ashraf, Q.F.; Nawar, N.; Geletu, M.; Seo, H.-S.; Abdallah, D.I. Improved Pharmacokinetic Profiles of HDAC6 Inhibitors via Cap Group Modifications. J. Med. Chem. 2025, 68, 18216–18229. [Google Scholar] [CrossRef]
  55. Ganesan, A. Targeting the zinc-dependent histone deacetylases (HDACs) for drug discovery. In Chemical Epigenetics; Springer: Cham, Switzerland, 2020; pp. 1–27. [Google Scholar]
  56. Spallotta, F.; Illi, B. The role of HDAC6 in glioblastoma multiforme: A new avenue to therapeutic interventions? Biomedicines 2024, 12, 2631. [Google Scholar] [CrossRef] [PubMed]
  57. Liu, L.; Dong, L.; Bourguet, E.; Fairlie, D.P. Targeting class IIa HDACs: Insights from phenotypes and inhibitors. Curr. Med. Chem. 2021, 28, 8628–8672. [Google Scholar] [CrossRef]
  58. Yousefian, M.; Hashemi, M.; Eskandarpour, V.; Zarghi, A.; Hadizadeh, F.; Ghodsi, R. New indolin-2-ones, possessing sunitinib scaffold as HDAC inhibitors and anti-cancer agents with potential VEGFR inhibition activity; design, synthesis and biological evaluation. Bioorg. Chem. 2025, 156, 108231. [Google Scholar] [CrossRef]
  59. Barone, S.; Bello, I.; Guadagni, A.; Cerchia, C.; Filocamo, G.; Cassese, E.; Alfano, A.I.; Esposito, C.; Morel, Á.J.F.; Brunetti, M. Challenging triple negative breast cancer through HDAC6 selective inhibition: Novel cap-group identification, structure-activity relationships, computational and biological studies. Eur. J. Med. Chem. 2025, 292, 117634. [Google Scholar] [CrossRef]
  60. Patel, V.K.; Shirbhate, E.; Tiwari, P.; Kore, R.; Veerasamy, R.; Mishra, A.; Rajak, H. Multi-targeted HDAC inhibitors as anticancer agents: Current status and future prospective. Curr. Med. Chem. 2023, 30, 2762–2795. [Google Scholar] [CrossRef]
  61. Mensah, A.A.; Spriano, F.; Sartori, G.; Priebe, V.; Cascione, L.; Gaudio, E.; Tarantelli, C.; Civanelli, E.; Aresu, L.; Rinaldi, A. Study of the antilymphoma activity of pracinostat reveals different sensitivities of DLBCL cells to HDAC inhibitors. Blood Adv. 2021, 5, 2467–2480. [Google Scholar] [CrossRef] [PubMed]
  62. Garcia-Manero, G.; Kazmierczak, M.; Wierzbowska, A.; Fong, C.Y.; Keng, M.K.; Ballinari, G.; Scarci, F.; Adès, L. Pracinostat combined with azacitidine in newly diagnosed adult acute myeloid leukemia (AML) patients unfit for standard induction chemotherapy: PRIMULA phase III study. Leuk. Res. 2024, 140, 107480. [Google Scholar] [CrossRef] [PubMed]
  63. Brunetto, A.T.; Ang, J.E.; Lal, R.; Olmos, D.; Molife, L.R.; Kristeleit, R.; Parker, A.; Casamayor, I.; Olaleye, M.; Mais, A. First-in-human, pharmacokinetic and pharmacodynamic phase I study of Resminostat, an oral histone deacetylase inhibitor, in patients with advanced solid tumors. Clin. Cancer Res. 2013, 19, 5494–5504. [Google Scholar] [CrossRef]
  64. Clinicaltrials.gov. NCT02953301. Available online: https://clinicaltrials.gov/study/NCT02953301#publications (accessed on 26 November 2025).
  65. Hou, K.; Dong, X.; Niu, W. Novel small-molecule therapies for myelodysplastic syndromes with IPSS-R ≥ 3.5 in patients aged 60 or older: Current landscape and challenges. Ther. Adv. Hematol. 2025, 16, 20406207251371298. [Google Scholar] [CrossRef]
  66. Clinicaltrials.gov. NCT03151408. Available online: https://clinicaltrials.gov/study/NCT03151408?intr=pracinostat&term=NCT03151408&rank=1 (accessed on 26 November 2025).
  67. Tian, J.; Han, M.; Song, F.; Liu, Y.; Shen, Y.; Zhong, J. Advances of HDAC inhibitors in tumor therapy: Potential applications through immune modulation. Front. Oncol. 2025, 15, 1576781. [Google Scholar] [CrossRef]
  68. Clinicaltrials.gov. NCT03618602. Available online: https://clinicaltrials.gov/study/NCT03618602?intr=Bisthianostat&rank=1 (accessed on 28 November 2025).
  69. Kim, B.; Huh, K.Y.; Yu, K.S.; Lee, S. Pharmacokinetics, pharmacodynamics and safety of oral formulation (CG-750) of ivaltinostat, a histone deacetylase inhibitor, compared to IV formulation (CG-745). Br. J. Clin. Pharmacol. 2024, 90, 1103–1114. [Google Scholar] [CrossRef] [PubMed]
  70. Clinicaltrials.gov. NCT05249101. Available online: https://clinicaltrials.gov/study/NCT05249101?intr=Ivaltinostat&rank=1 (accessed on 28 November 2025).
  71. Venugopal, B.; Baird, R.; Kristeleit, R.S.; Plummer, R.; Cowan, R.; Stewart, A.; Fourneau, N.; Hellemans, P.; Elsayed, Y.; Mcclue, S.; et al. A phase I study of quisinostat (JNJ-26481585), an oral hydroxamate histone deacetylase inhibitor with evidence of target modulation and antitumor activity, in patients with advanced solid tumors. Clin. Cancer Res. 2013, 19, 4262–4272. [Google Scholar] [CrossRef]
  72. Asad, M. Exploiting Epigenetic Pathway to Target Colorectal Cancer Stemness. Ph.D. Thesis, Oklahoma State University, Stillwater, OK, USA, 2025. [Google Scholar]
  73. Clinicaltrials.gov. NCT01486277. Available online: https://clinicaltrials.gov/study/NCT01486277?intr=Quisinostat&term=NCT01486277&rank=1 (accessed on 27 November 2025).
  74. Gou, P.; Fang, C.; Xu, M.; Zhang, D.; Wu, X.; Zhang, L.; Li, X.; Li, M.; Gan, L.; Luo, J. The dual HDAC/PI3K inhibitor CUDC-907 inhibits the growth and proliferation of MYC-driven Group 3 medulloblastoma. Cell Death Discov. 2025, 11, 172. [Google Scholar] [CrossRef]
  75. Yang, C.; Ren, X.; Cui, Y.; Jiang, H.; Li, M.; Yu, K.; Shen, S.; Li, M.; Zhang, X.; Zhao, X. New hopes and challenges in targeted therapy and immunotherapy for primary central nervous system lymphoma. Front. Immunol. 2025, 16, 1438001. [Google Scholar] [CrossRef]
  76. Clinicaltrials.gov. NCT01742988. Available online: https://clinicaltrials.gov/study/NCT01742988?intr=CUDC-907&term=NCT01742988&rank=1 (accessed on 27 November 2025).
  77. Ravikumar, B.; Cichońska, A.; Sahni, N.; Aittokallio, T.; Rahman, R. Advancements in Rational Multi-Targeted Drug Discovery: Improving the Efficacy-Safety Balance of Small Molecule Cancer Therapeutics. In Polypharmacology: Strategies for Multi-Target Drug Discovery; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2025; pp. 109–125. [Google Scholar]
  78. Gonçalves, M.T.; Lavareze, L.; Egal, E.S.A.; Altemani, A.; Mariano, F.V. Cell culture in salivary gland tumor research: Molecular insights of pathogenic targets and personalized medicine. Cytotechnology 2025, 77, 70. [Google Scholar] [CrossRef]
  79. Clinicaltrials.gov. NCT01384799. Available online: https://clinicaltrials.gov/study/NCT01384799?intr=cudc%20101&rank=4 (accessed on 27 November 2025).
  80. Guha, S.; Jagadeesan, Y.; Pandey, M.M.; Mittal, A.; Chitkara, D. Targeting the epigenome with advanced delivery strategies for epigenetic modulators. Bioeng. Transl. Med. 2025, 10, e10710. [Google Scholar] [CrossRef] [PubMed]
  81. Clinicaltrials.gov. NCT00697879. Available online: https://clinicaltrials.gov/study/NCT00697879?intr=CHR%203996&rank=2 (accessed on 27 November 2025).
  82. Rembiałkowska, N.; Rekiel, K.; Urbanowicz, P.; Mamala, M.; Marczuk, K.; Wojtaszek, M.; Żywica, M.; Radzevičiūtė-Valčiukė, E.; Novickij, V.; Kulbacka, J. Epigenetic dysregulation in cancer: Implications for gene expression and DNA repair-associated pathways. Int. J. Mol. Sci. 2025, 26, 6531. [Google Scholar] [CrossRef] [PubMed]
  83. Lin, C.-C.; Chen, T.W.-W.; Shiah, H.-S.; Tan, K.T.; Lin, C.-T.; Hsu, T.-A.; Lin, M.-C.; Liou, J.-P.; Pan, S.-L.; Huang, H.-L. Phase I first-in-human trial of ABT-301, an oral pan-HDAC inhibitor, in patients with advanced solid tumors. Am. Soc. Clin. Oncol. 2023, 41, e15137. [Google Scholar] [CrossRef]
  84. Harris, A.W.; Scott, R.C.; Butchbach, M.E. The effect of coadministration of D156844 and AR42 (REC-2282) on the survival and motor phenotype of mice with spinal muscular atrophy. Sci. Rep. 2025, 15, 28866. [Google Scholar] [CrossRef] [PubMed]
  85. Clinicaltrials.gov. NCT05130866. Available online: https://clinicaltrials.gov/study/NCT05130866?intr=REC-2282%20&rank=1 (accessed on 27 November 2025).
  86. Banerjee, S.; Adhikari, N.; Amin, S.A.; Jha, T. Histone deacetylase 8 (HDAC8) and its inhibitors with selectivity to other isoforms: An overview. Eur. J. Med. Chem. 2019, 164, 214–240. [Google Scholar] [CrossRef]
  87. Clinicaltrials.gov. NCT00677001. Available online: https://clinicaltrials.gov/study/NCT00677001?intr=R-306465&rank=1 (accessed on 28 November 2025).
  88. Mak, J.Y.; Wu, K.-C.; Gupta, P.K.; Barbero, S.; McLaughlin, M.G.; Lucke, A.J.; Tng, J.; Lim, J.; Loh, Z.; Sweet, M.J. HDAC7 inhibition by phenacetyl and phenylbenzoyl hydroxamates. J. Med. Chem. 2021, 64, 2186–2204. [Google Scholar] [CrossRef]
  89. Borrello, M.T.; Ruzic, D.; Paish, H.; Graham, E.; Collins, A.L.; Scott, R.; Higginbotham, S.; Radovic, B.; Nelson, G.; Bulmer, D. Pharmacological manipulation of liver fibrosis progression using novel HDAC 6 inhibitors. FEBS J. 2025, 292, 3397–3411. [Google Scholar] [CrossRef]
  90. Ali, K.H.; Sharma, C.; Oh, Y.J.; Yu, J.H.; Kim, S.-H.; Lee, H.; Seo, Y.H. Design and evaluation of highly selective HDAC6 inhibitors derived from the natural product tryptoline. Mol. Pharmacol. 2025, 107, 100062. [Google Scholar] [CrossRef]
  91. Cursaro, I.; Frattaruolo, L.; Scalvini, L.; Contri, C.; Bichicchi, A.; Tardiolo, N.; Tudino, V.; Rossi, S.; Manti, E.N.; Cappello, M.; et al. New Vanillyl-capped HDAC inhibitors exhibit anti-tumor efficacy in neuroblastoma and glioblastoma cells. Bioorg. Chem. 2025, 166, 109085. [Google Scholar] [CrossRef] [PubMed]
  92. Shi, J.; Wang, J.; Wang, X.; Qu, C.; Ye, C.; Li, X.; Chen, X.; Xu, Z. Design, synthesis and antiproliferative evaluation of tetrahydro-β-carboline histone deacetylase inhibitors bearing an aliphatic chain linker. RSC Adv. 2024, 14, 12762–12771. [Google Scholar] [CrossRef]
  93. Saul, J.G.; Huckleby, A.E.; Gugello, M.C.; Urbanczyk, J.; Desmarais, S.; Shin, H.; Bokka, A.; Jeon, J.; Tripathy, J.N.; Kim, S.-K. Inhibitory Effect of Novel Dihydroxamate Derivatives for Histone Deacetylase 1. Front. Biosci. Sch. 2025, 17, 38998. [Google Scholar] [CrossRef]
  94. Onuscakova, M.; Kauerova, T.; Fialova, E.; Pizova, H.; Garaj, V.; Kemka, M.; Frecer, V.; Kollar, P.; Bobal, P. New potent N-hydroxycinnamamide-based histone deacetylase inhibitors suppress proliferation and trigger apoptosis in THP-1 leukaemia cells. Arch. Pharm. 2025, 358, e2400889. [Google Scholar] [CrossRef] [PubMed]
  95. Bao, Y.; Li, J.; Ni, Y.; Bi, M.; Chen, J. The novel selective HDAC1 inhibitor ZJH-1 exhibits potent antitumor activity in castration-resistant prostate cancer, potentially involving HSP90AA1. Chem.-Biol. Interact. 2025, 421, 111777. [Google Scholar] [CrossRef] [PubMed]
  96. Nencetti, S.; Cuffaro, D.; Nuti, E.; Ciccone, L.; Rossello, A.; Fabbi, M.; Ballante, F.; Ortore, G.; Carbotti, G.; Campelli, F. Identification of histone deacetylase inhibitors with (arylidene) aminoxy scaffold active in uveal melanoma cell lines. J. Enzym. Inhib. Med. Chem. 2021, 36, 34–47. [Google Scholar] [CrossRef]
  97. Li, Z.; Qiu, H.; Lu, W.; Duan, N.; Fan, S.; Zhou, R.; Li, X.; Zhang, H.; Liu, N.; Yang, F. Design and synthesis of thiazole-based hydroxamate histone deacetylase inhibitors with potent antitumor efficacy by inducing apoptosis, pyroptosis and cell cycle arrest. Sci. Rep. 2025, 15, 24589. [Google Scholar] [CrossRef]
  98. Fan, S.; Wan, Z.; Qu, Y.; Lu, W.; Li, X.; Yang, F.; Zhang, H. Design and optimization of novel Tetrahydro-β-carboline-based HDAC inhibitors with potent activities against tumor cell growth and metastasis. Bioorg. Med. Chem. Lett. 2024, 114, 129986. [Google Scholar] [CrossRef]
  99. Ni, D.-X.; Wang, Q.; Li, Y.-M.; Cui, Y.-M.; Shen, T.-Z.; Li, X.-L.; Sun, H.-D.; Zhang, X.-J.; Zhang, R.; Xiao, W.-L. Synthesis of nigranoic acid and manwuweizic acid derivatives as HDAC inhibitors and anti-inflammatory agents. Bioorg. Chem. 2021, 109, 104728. [Google Scholar] [CrossRef]
  100. Zhu, S.; Zhu, W.; Zhao, K.; Yu, J.; Lu, W.; Zhou, R.; Fan, S.; Kong, W.; Yang, F.; Shan, P. Discovery of a novel hybrid coumarin-hydroxamate conjugate targeting the HDAC1-Sp1-FOSL2 signaling axis for breast cancer therapy. Cell Commun. Signal. 2024, 22, 361. [Google Scholar] [CrossRef]
  101. Mehndiratta, S.; Chen, M.C.; Chao, Y.H.; Lee, C.H.; Liou, J.P.; Lai, M.J.; Lee, H.Y. Effect of 3-subsitution of quinolinehydroxamic acids on selectivity of histone deacetylase isoforms. J. Enzym. Inhib. Med. Chem. 2021, 36, 74–84. [Google Scholar] [CrossRef]
  102. Jamshidi, Z.; Abnous, K.; Ghodsi, R.; Taghdisi, S.M.; Hadizadeh, F. Novel imidazolyl-hydroxamic acid Schiff base scaffold derivatives as histone deacetylase inhibitors: Design, synthesis, and biological assessment. Bioorg. Chem. 2025, 164, 108879. [Google Scholar] [CrossRef] [PubMed]
  103. Zhou, H.; Qi, Z.; Liu, D.; Xue, X.; Wang, C. Design, Synthesis, and Biological Evaluation of New Urushiol Derivatives as Potent Class I Histone Deacetylase Inhibitors. Chembiochem A Eur. J. Chem. Biol. 2023, 24, e202300238. [Google Scholar] [CrossRef] [PubMed]
  104. Shirbhate, E.; Koch, B.; Singh, V.; Dubey, A.; Yasin, H.K.A.; Rajak, H. Heteroaryl-Capped Hydroxamic Acid Derivatives with Varied Linkers: Synthesis and Anticancer Evaluation with Various Apoptosis Analyses in Breast Cancer Cells, Including Docking, Simulation, DFT, and ADMET Studies. Pharmaceuticals 2025, 18, 1148. [Google Scholar] [CrossRef] [PubMed]
  105. Toutah, K.; Nawar, N.; Timonen, S.; Sorger, H.; Raouf, Y.S.; Bukhari, S.; von Jan, J.; Ianevski, A.; Gawel, J.M.; Olaoye, O.O.; et al. Development of HDAC Inhibitors Exhibiting Therapeutic Potential in T-Cell Prolymphocytic Leukemia. J. Med. Chem. 2021, 64, 8486–8509. [Google Scholar] [CrossRef] [PubMed]
  106. Chen, X.; Wang, J.; Zhao, P.; Dang, B.; Liang, T.; Steimbach, R.R.; Miller, A.K.; Liu, J.; Wang, X.; Zhang, T.; et al. Tetrahydro-β-carboline derivatives as potent histone deacetylase 6 inhibitors with broad-spectrum antiproliferative activity. Eur. J. Med. Chem. 2023, 260, 115776. [Google Scholar] [CrossRef]
  107. Chen, H.; Li, Y.; Liang, Z.; Zhong, Z.; Huang, Y.; Liu, Z.; Gu, Y.; Jiang, L.; Gan, B.; Gan, C. Design, synthesis, and biological evaluation of estratriene-based hydroxamic acid derivatives as histone deacetylase inhibitors. J. Steroid Biochem. Mol. Biol. 2025, 255, 106867. [Google Scholar] [CrossRef] [PubMed]
  108. Neganova, M.; Aleksandrova, Y.; Suslov, E.; Mozhaitsev, E.; Munkuev, A.; Tsypyshev, D.; Chicheva, M.; Rogachev, A.; Sukocheva, O.; Volcho, K.; et al. Novel Multitarget Hydroxamic Acids with a Natural Origin CAP Group against Alzheimer’s Disease: Synthesis, Docking and Biological Evaluation. Pharmaceutics 2021, 13, 1893. [Google Scholar] [CrossRef] [PubMed]
  109. Bülbül, E.F.; Melesina, J.; Ibrahim, H.S.; Abdelsalam, M.; Vecchio, A.; Robaa, D.; Zessin, M.; Schutkowski, M.; Sippl, W. Docking, Binding Free Energy Calculations and In Vitro Characterization of Pyrazine Linked 2-Aminobenzamides as Novel Class I Histone Deacetylase (HDAC) Inhibitors. Molecules 2022, 27, 2526. [Google Scholar] [CrossRef]
  110. Jiang, X.; Jiang, Y.; Li, Q.; Li, L.; Wang, D. Synthesis, Molecular Docking, Dynamics Analysis Studies, and Cytotoxicityactivity Evaluation of Novel Berberine Derivative Bearing Oxadiazole and O-Diaminobenzene Moieties. Química Nova 2025, 48, e-20250121. [Google Scholar] [CrossRef]
  111. Ghosh, R.; Biswas, S.; Bagchi, A.; Chattopadhyay, S.K. Synthesis and Evaluation of 9-epi-Koshidacin B as Selective Inhibitor of Histone Deacetylase 1. J. Nat. Prod. 2024, 87, 2757–2767. [Google Scholar] [CrossRef]
  112. Hosseinzadeh, P.; Watson, P.R.; Craven, T.W.; Li, X.; Rettie, S.; Pardo-Avila, F.; Bera, A.K.; Mulligan, V.K.; Lu, P.; Ford, A.S.; et al. Anchor extension: A structure-guided approach to design cyclic peptides targeting enzyme active sites. Nat. Commun. 2021, 12, 3384. [Google Scholar] [CrossRef]
  113. Patra, A.; Ghosh, S.S.; Saini, G.K. Exploring potential molecular targets and therapeutic efficacy of beauvericin in triple-negative breast cancer cells. Comput. Biol. Chem. 2024, 112, 108154. [Google Scholar] [CrossRef]
  114. Sarkar, K.; Debnath, S.; Sen, D.; Kar, S.; Sil, S.K. Crucial Structural Understanding for Selective HDAC8 Inhibition: Common Pharmacophores, Molecular Docking, Molecular Dynamics, and Zinc Binder Analysis of Selective HDAC8 Inhibitors. Med. Chem. 2025, 21, 597–618. [Google Scholar] [CrossRef]
  115. Alibeg, A.A.A.; Mohammed, M.H. Molecular docking, synthesis, characteristics and preliminary cytotoxic study of new coumarin-sulfonamide derivatives as histone deacetylase inhibitors. Pol. Med. Bibliogr. 2024, 77, 514–525. [Google Scholar] [CrossRef]
  116. Carreiras, M.d.C.; Marco-Contelles, J. Hydrazides as inhibitors of histone deacetylases. J. Med. Chem. 2024, 67, 13512–13533. [Google Scholar] [CrossRef] [PubMed]
  117. Alwash, A.H.; Yaseen, Y.S.; Najumuldeen, Z.; Ibrahim, N.K. In Silico Profiling of New 1,2,3,4-Tetrahydropyrimidine Derivatives Linked to Hydroxamate Moiety by Various Aromatic Linkers as HDACs Inhibitors. Adv. J. Chem. Sect. A 2025, 8, 1201–1223. [Google Scholar]
  118. Tao, L.; Zhou, Y.; Luo, Y.; Qiu, J.; Xiao, Y.; Zou, J.; Zhang, Y.; Liu, X.; Yang, X.; Gou, K. Epigenetic regulation in cancer therapy: From mechanisms to clinical advances. MedComm–Oncology 2024, 3, e59. [Google Scholar] [CrossRef]
  119. Tahghighi, A.; Seyedhashemi, E.; Mohammadi, J.; Moradi, A.; Esmaeili, A.; Pornour, M.; Jafarifar, K.; Ganji, S.M. Epigenetic marvels: Exploring the landscape of colorectal cancer treatment through cutting-edge epigenetic-based drug strategies. Clin. Epigenetics 2025, 17, 34. [Google Scholar] [CrossRef]
  120. Villegas-Vazquez, E.Y.; Marín-Carrasco, F.P.; Reyes-Hernández, O.D.; Báez-González, A.S.; Bustamante-Montes, L.P.; Padilla-Benavides, T.; Quintas-Granados, L.I.; Figueroa-González, G. Revolutionizing ovarian cancer therapy by drug repositioning for accelerated and cost-effective treatments. Front. Oncol. 2025, 14, 1514120. [Google Scholar] [CrossRef]
  121. Wu, C.; Chen, S.; Wu, Z.; Xue, J.; Zhang, W.; Wang, S.; Zhao, X.; Wu, S. Chidamide and orelabrutinib synergistically induce cell cycle arrest and apoptosis in diffuse large B-cell lymphoma by regulating the PI3K/AKT/mTOR pathway. J. Cancer Res. Clin. Oncol. 2024, 150, 98. [Google Scholar] [CrossRef] [PubMed]
  122. Li, L.; Yang, W.; Pan, Y.; Ye, R.; Wang, Y.; Li, S.; Jiang, H.; Zhang, Q.; Wang, X.; Yan, J. Chidamide enhances T-cell-mediated anti-tumor immune function by inhibiting NOTCH1/NFATC1 signaling pathway in ABC-type diffuse large B-cell lymphoma. Leuk. Lymphoma 2024, 65, 895–910. [Google Scholar] [CrossRef] [PubMed]
  123. Kawakubo, K.; Castillo, C.F.-d.; Liss, A.S. Epigenetic regulation of pancreatic adenocarcinoma in the era of cancer immunotherapy. J. Gastroenterol. 2022, 57, 819–826. [Google Scholar] [CrossRef]
  124. Suraweera, A.; O’Byrne, K.J.; Richard, D.J. Epigenetic drugs in cancer therapy. Cancer Metastasis Rev. 2025, 44, 37. [Google Scholar] [CrossRef]
  125. Ho, T.; Coleman, C.; Shah, P.; Yazbeck, V. Advances in Hodgkin’s lymphoma pharmacotherapy: A focus on histone deacetylase inhibitors. Expert Opin. Pharmacother. 2023, 24, 1427–1438. [Google Scholar] [CrossRef] [PubMed]
  126. Liu, G.; Barczak, W.; Lee, L.N.; Shrestha, A.; Provine, N.M.; Albayrak, G.; Zhu, H.; Hutchings, C.; Klenerman, P.; La Thangue, N.B. The HDAC inhibitor zabadinostat is a systemic regulator of adaptive immunity. Commun. Biol. 2023, 6, 102. [Google Scholar] [CrossRef]
  127. Clinicaltrials.gov. NCT01977638. Available online: https://clinicaltrials.gov/study/NCT01977638?intr=CXD101&rank=1#publications (accessed on 28 November 2025).
  128. Ibrahim, H.S.; Abdelsalam, M.; Zeyn, Y.; Zessin, M.; Mustafa, A.M.; Fischer, M.A.; Zeyen, P.; Sun, P.; Bülbül, E.F.; Vecchio, A.; et al. Synthesis, Molecular Docking and Biological Characterization of Pyrazine Linked 2-Aminobenzamides as New Class I Selective Histone Deacetylase (HDAC) Inhibitors with Anti-Leukemic Activity. Int. J. Mol. Sci. 2021, 23, 369. [Google Scholar] [CrossRef] [PubMed]
  129. Namballa, H.K.; Anchi, P.; Lakshmi Manasa, K.; Soni, J.P.; Godugu, C.; Shankaraiah, N.; Kamal, A. β-Carboline tethered cinnamoyl 2-aminobenzamides as class I selective HDAC inhibitors: Design, synthesis, biological activities and modelling studies. Bioorg. Chem. 2021, 117, 105461. [Google Scholar] [CrossRef]
  130. Himaja, A.; Routholla, G.; Patel, T.; Banerjee, S.; Begum, D.; Regula, S.; Pulya, S.; Biswas, S.; Adhikari, N.; Ghosh, B. Design and synthesis of pyridine-based benzamides as potent HDAC3 inhibitors as an armament against breast cancer with in vivo validation. Eur. J. Med. Chem. 2025, 291, 117636. [Google Scholar] [CrossRef]
  131. Guo, Z.; Zhang, Y.; Bao, Y.; Huang, Z.; Gu, X.; Wang, G.; Li, J. Synthesis and structure-activity relationship of thiol-based histone deacetylase 6 inhibitors. Chem. Biol. Drug Des. 2022, 100, 90–107. [Google Scholar] [CrossRef]
  132. Song, X.; Wang, H.; Gao, Y.; Zhang, W.; Lei, X. Synthesis and biological evaluation of the Fluoro analog of Romidepsin with improved selectivity for class I histone deacetylases (HDACs). Bioorg. Chem. 2025, 159, 108348. [Google Scholar] [CrossRef]
  133. Kamloon, T.; Senawong, T.; Senawong, G.; Namwan, N.; Kumboonma, P.; Somsakeesit, L.-O.; Ritchumpon, P.; Nontakitticharoen, M.; Nasomjai, P.; Phaosiri, C. Exploring putative histone deacetylase inhibitors with antiproliferative activity of chrysin derivatives. Med. Chem. Res. 2025, 34, 1308–1320. [Google Scholar] [CrossRef]
  134. Cho, H.; Lee, E.; Kim, J.; Shin, S.; Kim, Y.-J.; Lee, H.; Yu, J.H.; Jeon, Y.H.; Lee, S.W.; Lee, S.Y. Discovery of organosulfur-based selective HDAC8 inhibitors with anti-neuroblastoma activity. Eur. J. Pharm. Sci. 2024, 203, 106921. [Google Scholar] [CrossRef]
  135. Roy, R.; Ria, T.; RoyMahaPatra, D.; Sk, U.H. Single inhibitors versus dual inhibitors: Role of HDAC in cancer. ACS Omega 2023, 8, 16532–16544. [Google Scholar] [CrossRef] [PubMed]
  136. Huang, Y.-H.; Huang, Y.-M.; Huang, W.-J.; Yu, M.-C.; Chuang, C.-H.; Hsu, Y.-F.; Chen, H.-C.; Chen, L.-C.; Huang, S.-W.; Hsu, M.-J. The hydroxamate based HDAC inhibitor WMJ-J-09 induces colorectal cancer cell death by targeting tubulin and downregulating survivin. Sci. Rep. 2025, 15, 19590. [Google Scholar] [CrossRef]
  137. Ibrahim, T.S.; Malebari, A.M.; Mohamed, M.F. Design, synthesis, in vitro anticancer evaluation and molecular modelling studies of 3, 4, 5-trimethoxyphenyl-based derivatives as dual EGFR/HDAC hybrid inhibitors. Pharmaceuticals 2021, 14, 1177. [Google Scholar] [CrossRef] [PubMed]
  138. Lee, S.; Wang, S.W.; Yu, C.L.; Tai, H.C.; Yen, J.Y.; Tuan, Y.L.; Wang, H.H.; Liu, Y.T.; Chen, S.S.; Lee, H.Y. Effect of phenylurea hydroxamic acids on histone deacetylase and VEGFR-2. Bioorg. Med. Chem. 2021, 50, 116454. [Google Scholar] [CrossRef]
  139. Wu, J.; Wang, X.; Yao, Y.; Du, N.; Duan, L.; Gong, P. Design, synthesis and antitumor activities of phthalazinone derivatives as PARP-1 inhibitors and PARP-1/HDAC-1 inhibitors. Bioorg. Chem. 2024, 151, 107556. [Google Scholar] [CrossRef]
  140. Bass, A.K.A.; Nageeb, E.M.; El-Zoghbi, M.S.; Mohamed, M.F.A.; Badr, M.; Abuo-Rahma, G.E.A. Utilization of cyanopyridine in design and synthesis of first-in-class anticancer dual acting PIM-1 kinase/HDAC inhibitors. Bioorg. Chem. 2022, 119, 105564. [Google Scholar] [CrossRef]
  141. Xie, S.; Leng, J.; Zhao, S.; Zhu, L.; Zhang, M.; Ning, M.; Zhao, B.; Kong, L.; Yin, Y. Design and biological evaluation of dual tubulin/HDAC inhibitors based on millepachine for treatment of prostate cancer. Eur. J. Med. Chem. 2024, 268, 116301. [Google Scholar] [CrossRef] [PubMed]
  142. Abdulwahab, H.G.; Mansour, R.E.-S.; Farghaly, T.A.; El-Sehrawi, H.M. Discovery of novel benzimidazole derivatives as potent HDACs inhibitors against leukemia with (Thio) Hydantoin as zinc-binding moiety: Design, synthesis, enzyme inhibition, and cellular mechanistic study. Bioorg. Chem. 2024, 146, 107284. [Google Scholar] [CrossRef]
  143. Liu, Q.; Zhang, B.; Wang, Y.; Wang, X.; Gou, S. Discovery of phthalazino[1,2-b]-quinazolinone derivatives as multi-target HDAC inhibitors for the treatment of hepatocellular carcinoma via activating the p53 signal pathway. Eur. J. Med. Chem. 2022, 229, 114058. [Google Scholar] [CrossRef]
  144. Shetty, M.G.; Pai, P.; Dey, B.; Satyamoorthy, K.; Shil, S.; Nayak, U.Y.; T, A.; Sundara, B.K. Evaluation of 1,10-phenanthroline-based hydroxamate derivative as dual histone deacetylases/ribonucleotide reductase inhibitor with antitumor activities. Daru J. Pharm. 2024, 32, 263–278. [Google Scholar] [CrossRef]
  145. Zhang, J.; Yang, M.; Liu, Q.; Xue, X.; Tian, S.; Hu, X.; Li, M.; Li, J.; Chai, Q.; Liu, F. Discovery of epigenetic modulators targeting HDACs and EZH2 simultaneously for the treatment of hematological malignancies. Bioorg. Chem. 2024, 153, 107964. [Google Scholar] [CrossRef] [PubMed]
  146. Zhang, S.; Wang, Y.F.; Lu, H.R.; Yang, X.Q.; Zhang, Y.; Ma, X.L.; Huang, R.Z. Discovery of Novel Imidazothiazole-Based Hydroxamic Acid Derivatives as Potent Indoleamine 2,3-Dioxygenase 1 and Histone Deacetylase 6 Dual Inhibitors. Molecules 2025, 30, 2508. [Google Scholar] [CrossRef]
  147. Chen, X.; Li, R.; Qiu, Y.; Lin, F.; Chen, S.; Li, X.; Sun, H.; Jiang, G.; Fang, H.; Qin, J. Design, synthesis, and biological evaluation of N-(2-amino-phenyl)-5-(4-aryl-pyrimidin-2-yl) amino)-1H-indole-2-carboxamide derivatives as novel inhibitors of CDK9 and class I HDACs for cancer treatment. Bioorg. Chem. 2025, 162, 108577. [Google Scholar] [CrossRef] [PubMed]
  148. Hou, B.; Jia, G.; Li, Z.; Jiang, Y.; Chen, Y.; Li, X. Discovery of hydrazide-based PI3K/HDAC dual inhibitors with enhanced pro-apoptotic activity in lymphoma cells. Eur. J. Med. Chem. 2025, 292, 117658. [Google Scholar] [CrossRef] [PubMed]
  149. Kim, B.; Kang, W.; Neerasa, J.; Chung, H. Design and synthesis of novel cyclin-dependent kinase 4/6 (CDK4/6) and histone deacetylase (HDAC) dual inhibitors: In vitro and in vivo anticancer activity. Eur. J. Med. Chem. 2025, 301, 118192. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The 3D chemical structure, Pharmacokinetics, Pharmacodynamics, Mechanism of action, Tumor Types, and specificity of HDAC inhibitor approved drugs with hydroximate scaffold [36,37,38,39,40,41,42,43,44].
Figure 1. The 3D chemical structure, Pharmacokinetics, Pharmacodynamics, Mechanism of action, Tumor Types, and specificity of HDAC inhibitor approved drugs with hydroximate scaffold [36,37,38,39,40,41,42,43,44].
Cells 14 01997 g001aCells 14 01997 g001b
Figure 2. General pharmacophore architecture of HDACi, highlighting the three essential structural domains (blue for Oxygen, red for nitrogen, white for hydrogen and grey for carbon atoms).
Figure 2. General pharmacophore architecture of HDACi, highlighting the three essential structural domains (blue for Oxygen, red for nitrogen, white for hydrogen and grey for carbon atoms).
Cells 14 01997 g002
Figure 3. Binding pose of Vorinostat at (A) HDAC2 zinc catalytic site based on the experimental crystal structure (PDB: 4LXZ). (B) HDAC6 zinc catalytic site based on the experimental crystal structure (PDB: 5EEI). (C) HDAC8 zinc catalytic site based on the experimental crystal structure (PDB: 4BZ6). Dashed lines indicate Zn2+–ligand coordination interactions.
Figure 3. Binding pose of Vorinostat at (A) HDAC2 zinc catalytic site based on the experimental crystal structure (PDB: 4LXZ). (B) HDAC6 zinc catalytic site based on the experimental crystal structure (PDB: 5EEI). (C) HDAC8 zinc catalytic site based on the experimental crystal structure (PDB: 4BZ6). Dashed lines indicate Zn2+–ligand coordination interactions.
Cells 14 01997 g003aCells 14 01997 g003b
Table 2. The structures and the IC50 values against a panel of cancer cell lines and HDAC enzymes for the most active agents, which contain hydroximate scaffolds.
Table 2. The structures and the IC50 values against a panel of cancer cell lines and HDAC enzymes for the most active agents, which contain hydroximate scaffolds.
CodeStructuresEvaluated Cancer/Normal Cell LinesEvaluated TargetRef.
Cell LinesIC50/%VHDACIC50
St.1Cells 14 01997 i013U87-MG
T98G
U251-MG
51.31 µM
42.60 µM
2.37 µM
HDAC1
HDAC2
HDAC3
HDAC5
HDAC6
HDAC10
403 nM
537 nM
1.278 µM
4.455 µM
4.5 nM
202 nM
[91]
St.2Cells 14 01997 i014HCT116
SK-MEL-2
HS-5
1.54 µM
0.48 µM
>50 µM
HDAC1
HDAC2
HDAC3
HDAC6
8.73 nM
23.5 nM
32.1 nM
>50 nM
[92]
St.3Cells 14 01997 i015NANAHDAC12.96 µM[93]
St.4Cells 14 01997 i016THP-11.60 µMHDAC I/II157.0 nM[94]
St.5Cells 14 01997 i017PC3
PDXO
65 nM
345 nM
HDAC1increased histone H3 acetylation[95]
St.6Cells 14 01997 i018UMNAHDAC1
HDAC3
HDAC6
HDAC8
0.137 µM
0.040 µM
0.010 µM
9.29 µM
[96]
St.7Cells 14 01997 i019HepG2
PC9
HCT116
MCF-7
0.19 µM
0.46 µM
1.19 µM
3.31 µM
HDAC1
HDAC6
0.8 nM
2.5 nM
[97]
St.8Cells 14 01997 i020A5491.09 µMHDAC14.5 nM[98]
St.9Cells 14 01997 i021J774A
(A) LDH
(B) IL-1β
9.98 µM
5.50 µM
HDAC1
HDAC2
HDAC4
HDAC6
HDAC8
1.14 µM
10.56 µM
19.39 µM
2.23 µM
>50 µM
[99]
St.10Cells 14 01997 i0224T1
MDAMB231
21.4%
25.4%
HDAC10.99 µM[100]
St.11Cells 14 01997 i023LX-2NAHDAC1
HDAC6
5.9 nM
78.1 nM
[89]
St.12Cells 14 01997 i024LX-2NAHDAC1
HDAC6
8.8 nM
50.8 nM
St.13Cells 14 01997 i025A549
HCT116
1.29 µM
1.61 µM
HDAC1 HDAC2
HDAC6 HDAC8
2.62 µM
1.31 µM
4.75 nM
1.80 µM
[101]
St.14Cells 14 01997 i026A27808.10 µMPan-HDAC12.58 µM[102]
St.15Cells 14 01997 i027HT-29
MDA-MB-231
4.02 µM
2.31 µM
HDAC816.11 nM[103]
St.16Cells 14 01997 i028A549
MCF-7
<10 µg/mlHDAC1 HDAC63.06 µM
4.08 µM
[104]
St.17Cells 14 01997 i029MV4-11
MRC-9
0.42 µM
>20 µM
HDAC6
HDAC8
8.50 nM
0.334 µM
[105]
St.18Cells 14 01997 i030NANAHDAC62.68 nM[106]
St.19Cells 14 01997 i031HeLa6.74 µMHDAC6.23 µM[107]
St.20Cells 14 01997 i032NANAHDAC60.96 µM[108]
NA: Not applicable (no cell lines were evaluated in these works).
Table 3. Representative Benzamide-Based HDACi in Clinical Trials: Structures, Targets, and Cancer Applications.
Table 3. Representative Benzamide-Based HDACi in Clinical Trials: Structures, Targets, and Cancer Applications.
Name or CodeStructureClinical
Trial
Cancer TypeHDAC ClassRef.
EntinostatCells 14 01997 i033I, II & IIISolid tumors, Chronic Myeloid Leukemia, & Acute Myeloid LeukemiaI[119,120]
ChidamideCells 14 01997 i034Ib/II/IIIT-cell Lymphoma, Angioimmunoblastic T-cell Lymphoma, lymphoma & Breast
cancer
HDAC 1,2,3,10[121,122]
TacedinalineCells 14 01997 i035II/IIIAdvanced pancreatic cancer & Multiple MyelomaI[123,124]
MocetinostatCells 14 01997 i036IIclassical Hodgkin Lymphoma & Relapsed/Refractory LymphomaI & IV[125]
Zabadinostat
(CXD-101)
Cells 14 01997 i037Isolid tumours, lymphoma, and myelomaI[126,127]
Table 4. The structures of Non-Hydroxamate-Based HDACi and the IC50 values against a panel of cancer cell lines and HDAC enzymes for the most active agents.
Table 4. The structures of Non-Hydroxamate-Based HDACi and the IC50 values against a panel of cancer cell lines and HDAC enzymes for the most active agents.
CodeStructuresEvaluated Cancer/Normal Cell LinesEvaluated TargetRef.
Cell LinesIC50HDACIC50
St.21Cells 14 01997 i038MV4-11
MOLM-13
194 nM
318 nM
HDAC1
HDAC2
HDAC3
0.13 µM
0.28 µM
0.31 µM
[128]
St.22Cells 14 01997 i039HCT-15
HT-29
L-132
0.70 µM
0.94 µM
14.50 µM
HDAC I0.97 µM[129]
St.23Cells 14 01997 i040MDA-MB-23134.3 µMHDAC30.69 µM[130]
St.24Cells 14 01997 i041NANAHDAC1
HDAC6
2.124 µM
73 nM
[131]
St.25Cells 14 01997 i042NANAHDAC1
HDAC6
13.0 nM
4.0 nM
[131]
St.26Cells 14 01997 i043HT-29<150 nMHDAC1
HDAC2
HDAC3
HDAC8
HDAC6
0.948 nM
0.856 nM
1.06 nM
4.24 nM
>1000 nM
[132]
St.27Cells 14 01997 i044HCT-11613.04 µMHDAC875.37 µM[133]
St.28Cells 14 01997 i045BE(2)-C
IMR-32
SH-SY5Y
2.52 µM
1.50 µM
1.72 µM
HDAC6
HDAC8
10.78 µM
0.81 µM
[134]
NA: Not applicable (no cell lines were evaluated in these works).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hawash, M. Next-Generation HDAC Inhibitors: Advancing Zinc-Binding Group Design for Enhanced Cancer Therapy. Cells 2025, 14, 1997. https://doi.org/10.3390/cells14241997

AMA Style

Hawash M. Next-Generation HDAC Inhibitors: Advancing Zinc-Binding Group Design for Enhanced Cancer Therapy. Cells. 2025; 14(24):1997. https://doi.org/10.3390/cells14241997

Chicago/Turabian Style

Hawash, Mohammed. 2025. "Next-Generation HDAC Inhibitors: Advancing Zinc-Binding Group Design for Enhanced Cancer Therapy" Cells 14, no. 24: 1997. https://doi.org/10.3390/cells14241997

APA Style

Hawash, M. (2025). Next-Generation HDAC Inhibitors: Advancing Zinc-Binding Group Design for Enhanced Cancer Therapy. Cells, 14(24), 1997. https://doi.org/10.3390/cells14241997

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

Article Metrics

Back to TopTop