You are on page 1of 27

European Journal of Medicinal Chemistry 164 (2019) 214e240

Contents lists available at ScienceDirect

European Journal of Medicinal Chemistry


journal homepage: http://www.elsevier.com/locate/ejmech

Review article

Histone deacetylase 8 (HDAC8) and its inhibitors with selectivity to


other isoforms: An overview
Suvankar Banerjee, Nilanjan Adhikari 1, Sk Abdul Amin, Tarun Jha*
Natural Science Laboratory, Division of Medicinal and Pharmaceutical Chemistry, Department of Pharmaceutical Technology, P. O. Box 17020, Jadavpur
University, Kolkata, 700032, West Bengal, India

a r t i c l e i n f o a b s t r a c t

Article history: The histone deacetylases (HDACs) enzymes provided crucial role in transcriptional regulation of cells
Received 22 September 2018 through deacetylation of nuclear histone proteins. Discoveries related to the HDAC8 enzyme activity
Received in revised form signified the importance of HDAC8 isoform in cell proliferation, tumorigenesis, cancer, neuronal disor-
4 December 2018
ders, parasitic/viral infections and other epigenetic regulations. The pan-HDAC inhibitors can confront
Accepted 16 December 2018
Available online 19 December 2018
these conditions but have chances to affect epigenetic functions of other HDAC isoforms. Designing of
selective HDAC8 inhibitors is a key feature to combat the pathophysiological and diseased conditions
involving the HDAC8 activity. This review is concerned about the structural and positional aspects of
Keywords:
Histone deacetylase
HDAC8 in the HDAC family. It also covers the contributions of HDAC8 in the pathophysiological condi-
HDAC8 tions, a preliminary discussion about the recent scenario of HDAC8 inhibitors. This review might help to
Cancer deliver the structural, functional and computational information in order to identify and design potent
HDAC8 inhibitor and selective HDAC8 inhibitors for target specific treatment of diseases involving HDAC8 enzymatic
Structure-activity relationship (SAR) activity.
Quantitative structure-activity relationship © 2018 Elsevier Masson SAS. All rights reserved.
(QSAR)

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
2. A short trip to HDAC enzymes and its classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
3. FDA approved HDAC inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
3.1. Vorinostat (suberoylanilide hydroxamic acid/SAHA/Zolinza®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
3.2. Romidepsin (Istodax®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.3. Belinostat (Beleodaq®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.4. Panobinostat (Farydak®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.5. Pracinostat (SB939) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
3.6. Chidamide (HBI-8000/Epidaza®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
4. Role of HDAC8 in different diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
5. Structure of HDAC8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
6. An insight into the HDAC8 inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
7. QSAR studies on HDAC8 inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
8. Future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
Conflicts of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236

* Corresponding author.
E-mail address: tjupharm@yahoo.com (T. Jha).
1
Present Address: School of Pharmaceutical Technology, ADAMAS University, Barasat-Barrackpore Road, P. O. Jagannathpur, Kolkata e 700126, West Bengal, India.

https://doi.org/10.1016/j.ejmech.2018.12.039
0223-5234/© 2018 Elsevier Masson SAS. All rights reserved.
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 215

1. Introduction to the other HDAC isoforms of the class. Additionally, the C-ter-
minal protein binding domain of other HDACs is found to be absent
Histone deacetylases (HDACs), called as eraser enzymes, belong in HDAC8 [4]. Moreover, the attachment of some non-histone
to the class of Zn2þ or nicotinamide adenine dinucleotide (NADþ)- proteins like cohesin, cortactin, Estrogen Related Receptor a
dependent proteolytic enzymes. These HDACs are involved in the (ERRa) with HDAC8 has been observed [1,4]. Beside the arguable
transcriptional repression and chromatin condensation mecha- fact about the histone proteins as the major or actual target of
nisms by removing the acetate moiety form acetylated 3-amino HDAC8 enzyme, there are a number of non-histone proteins (such
groups of histone lysine and other non-histone protein [1e3]. This as SMC3, EPR-a and p53) those have been reported as important
process of deletion of the acetate group from the acetylated factors of HDAC8 deacetylation [11,12]. The importance of HDAC8
3 -amino lysine of histone proteins is known as “deacetylation of enzyme activity for the expression of both wild type and mutated
histone proteins” (Fig. 1). Hence, the HDAC enzymes are also known p53 has also been observed [13]. The knockdown of HDAC8 resulted
as “Lysin Deacetylase” (KDAC) enzymes [3]. in the antiproliferative activity for the cells carrying a p53 mutant,
Due to the deacetylation property, HDACs participate in a but the same results have not been seen for the cells with wild type
number of important biological signaling pathways through regu- p53 indicating that HDAC8 inhibition can serve well for tumors
lation and modulation of these pathways [4]. The enzymatic ac- containing mutant p53 as an adjuvant therapy [4,7]. Though the
tivities of HDAC isoforms are responsible for maintenance of several important role of HDAC8 in the patterning of mice skull cranial
normal physiological conditions. These HDAC isoforms are also neural creast cell and skull stability have been seen, the expression
involved in the occurrence and progression of a number of patho- of HDAC8 has been seen in colon, lung and pancreatic cancers,
physiological conditions and diseases. These include neurodegen- acute lymphocytic leukemia (ALL) and acute myeloid leukemia
erative disorders, cancers, inflammation, autoimmune diseases and (AML) including childhood neuroblastoma [4,14,15]. The knock-
metabolic disfunctions [4]. For contributions of the HDAC enzymes down of HDAC8 gene through the RNAi mediated process proved
in the pathophysiological conditions, these HDAC enzymes have the essentiality of HDAC8 enzyme for proliferation in human colon,
became an interesting and important target to combat the diseased lung and cervical cancers, gastric adeno carcinoma and leukemia
conditions and especially cancers and tumors [5,6]. The interrup- [4]. The role of HDAC8 in both sistosomasis and viral infections was
tion in the normal expression of HDAC enzymes or their activity reported. The abnormal enzymatic activity of HDAC8 isoform has
causes abnormal cellular functionality lead to generation of mul- also been reported in Cornelia de Lange Syndrome (CdLS) [16,17].
tiple diseased conditions [4]. The abnormal HDAC expression and HDACs also affect the gelatinase enzymes (MMP-2, MMP-9) in
their abnormal deacetylation activity became a very popular and cancer progration. HDAC enzyme causes up regulation of the
interesting topic for the prevention of the pathophysiological gelatinase enzymes leading to the tumor progression and metas-
conditions including cancer generation and progression [1,4,7,8]. tasis [1e3]. The inhibition of HDAC enzymes including HDAC8
HDAC enzymes are categorized into two groups: the Zn2þ-depen- causes inhibition of the deacetylation activity of HDAC8 enzyme
dent HDACs [namely class I (HDAC1, 2, 3 and 8), class II (HDAC4, 5, 6, and induces acetylation of the histone and non-histone proteins.
7, 9 and 10) and class IV (HDAC11)] and the nicotinamide adenine This inhibition leads to restrict the DNA repairing, cell cycle arrest,
dinucleotide (NADþ)-dependent enzymes [class III sirtuins]. More alteration in genetic expression and induces cellular apoptosis [2].
than 15 year ago, HDAC8 was discovered as a Zn2þ dependant class From studies on HDAC inhibitors, it has been seen that most of the
Ic HDAC enzyme which has a size of 42 kDa comprising 377 amino HDAC inhibitors including the HDAC8 inhibitors contains three
acids [1,4,9,10]. Among the diverse class of HDACs, HDAC8 can be pharmacophoric features in its scaffold known as the Zinc binding
identified easily from the other HDAC enzymes of the class due to group (ZBG), CAP region and a linker moiety presents between the
its unique structural and functional specializations compared to the CAP and ZBG [1e3]. It has been seen that the ZBG causes cheletion
other HDAC isoforms of the class [1,4]. The X-linked structure of with the HDAC8 Zn2þ ion where the hydrophobic CAP region fits
human HDAC8 and its co-factor independent activity is dissimilar and interacts with the HDAC8 pocket whereas the linker helps to

Fig. 1. Deacetylation of lysine residue of histone proteins by histone deacetylase 8 (HDAC8) along with inhibition of HDAC8 by HDAC8 inhibitor.
216 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

maintain the distance between the ZBG and the CAP region while localized into the cytoplasm [10]. The conserved serine residue
binding [1e3]. The linker region also interacts with the amino acid incorporated into the class IIA isoforms that is responsible for
residues of the HDAC8 catalytic domain [18]. Till date, only five phosphorylation [1]. The HDAC10 is composed of an active deace-
HDAC inhibitors namely, Vorinostat (SAHA/Zolinza®), Romidepsin tylase domain along with an incomplete domain while the HDAC6
(Istodax®), Panobinostat (Farydak®), Belinostat (Beleodaq®) and comprises two catalytic domains along with a carboxyl terminus
Pracinostat (SB939) have been approved by USFDA (United States zinc figure motif [1,2].
Food and Drug Administration) and one HDAC inhibitor namely
Chidamide (HBI-8000/Epidaza®) is approved by Food and Drug 3. FDA approved HDAC inhibitors
Administration, China for the treatment of peripheral T-cell lym-
phoma (PTCL) [3]. However, these HDAC inhibitors are pan-HDAC During the long era to discover an effective HDAC inhibitor for
inhibitors. These interact with all classes of HDACs and lead to prevention of HDAC enzymes which contribute in development
adverse effects. Therefore, isoform-specific HDAC8 inhibitors are and progression of cancer, a number of ligands have been synthe-
important. sized with HDAC inhibitory activity [22e26]. Unfortunately, only a
handful of inhibitors among them provided promising results.
2. A short trip to HDAC enzymes and its classification Some of the promising molecules have been tested and are evalu-
ated in the different phases of clinical trials. However, only six
The group of HDAC contains a number of isoforms in its family. A HDAC inhibitors (HDACIs) have successfully passed the clinical trial
total of 18 HDAC isoforms belong to the group comprising either studies till now (Fig. 3) and proved themselves to combat cancer. 5
Zn2þ or NADþ in their structures [8]. On the basis of their structural, out of 6 approved HDAC inhibitors have been approved by the
functional as well as origin, HDACs are categorized into four USFDA (United States Food and Drug Administration).
different classes. Among these four classes of HDACs, the class I and
class II isoforms exhibit structural similarity with the yeast RPD3 3.1. Vorinostat (suberoylanilide hydroxamic acid/SAHA/Zolinza®)
deacetylase and yeast Hda I enzymes respectively [4,12,19]. The 11
HDACs belongs to class I, II and IV are Zn2þ-dependent metal- Vorinostat (suberoylanilide hydroxamic acid/SAHA, FDA-01) is a
loenzymes whereas the class III HDACs (known as sirtuins/SIRTs) potent and orally active HDAC inhibitor. It shows activity against all
are NADþ-dependent with 7 members (SIRT 1-7) [2,9,20]. The class the HDAC enzymes (Table 2, Fig. 3) including HDAC8 (HDAC8
IV HDAC family contains only HDAC11 (Table 1) which differs from IC50 ¼ 1480 nM) [1,2,25]. For haematological malignancies such as
the other three classes of HDACs. The structural difference is due to cutaneous T-cell lymphoma (CTCL) and solid tumors, responses
its sequence with other class of HDACs. However, the catalytic re- were noted for SAHA [27]. A maximum of 400 mg of SAHA can be
gion resembles for both class I and class II HDACs [10,12]. tolerated orally for haematological cancers. However, for solid tu-
The major class of HDACs (class I and class II) are known as the mors, a maximum of 600 mg can be tolerated [28]. Not only that,
classical HDACs. These possess the conserved deacetylase core SAHA was found to be an effective radiosensitizer in human glio-
domain which is responsible for the elimination of N-terminal blastoma in preclinical studies [29].
acetyl group from the lysine residues in presence of Zn2þ ion [21]. SAHA effectively exhibited a potent antiproliferative and
The isoforms of class I HDACs can be further subdivided into three apoptotic activity in type I and type II human endometrial cancers
subclasses namely class IA (HDAC1, HDAC2), class IB (HDAC3) and by modulating the expressions of some genes related to specific
class IC (HDAC8) depending on their phylogenic data. HDAC8 is Insulin-like growth factor-I (IGF-I) receptor signaling pathway [30].
being regulated by protein kinase A (PKA) whereas casein kinase-2 SAHA enhanced the phosphorylation of IGF-IR, upregulated the
is responsible for the regulation of class IA isoforms (HDAC1, expression of PTEN and p21 as well as reduced the levels of cyclin
HDAC2) [4]. The class I HDACs are responsible for the catalysis of a D1 and p53 in the type I human endometrial cancer cell lines
wide range of non-histone and histone substrates including a va- whereas caused the upregulation of IGF-IR, expression of p21 and
riety of transcriptional factors [12]. Similarly, the class II HDACs downregulation of cyclin D1, total AKT and PTEN expressions when
(HDAC4, HDAC5, HDAC6, HDAC7, HDAC10) can also be classified evaluated in the type II human endometrial cancer cell lines.
into two subclasses such as class IIA (HDAC4, HDAC5, HDAC7) and Fascinatingly, SAHA was also found to hyperacetylate the histone
class IIB (HDAC6, HDAC9, HDAC10) [1e3,10,12]. The class I HDACs H3 in the type I and II endometrial cancers [31,32]. In murine and
localizes into the neucleoplasm of cells whereas the class II and human lung cancer cell lines along with the genetically engineered
class IV HDACs are found in both cytoplasm and nucleus (Fig. 2). lung cancer models in mouse, SAHA profoundly decreased cancer
Interestingly, the SIRTs are also localized into the mitochondria. cell proliferation along with expressions of cyclin D1 and cyclin E
HDAC6 is the only enzyme of class II HDAC family which is mostly but increased the expression of p27, apoptosis and histone

Table 1
Classification of HDACs (listed below are the cellular location, non-histone substrates and complexes HDACs).

Class Dependency Sub class Isoform Cellular localization Non-histone substrates Complex

I Zn2þ IA HDAC1 Nucleus SHP,p53,MyoD,E2F1, STAT3, NF-kB,CtIP,AMPK, RB1 Sin3, NURD


HDAC2 Nucleus GCCR,BCL6,STAT3, YY1 Sin3, NURD
IB HDAC3 Nucleus SHP, YY1, GATA1, p65, STAT3, MEF2D NCOR1/NCOR2-GPS2-TBL1X
IC HDAC8 Nucleus SMC3, actin ____
II IIA HDAC4 Nucleus/Cytoplasm GATA1, HP1 NCOR1/NCOR2
HDAC5 Nucleus/Cytoplasm SMAD7, HP1 ____
HDAC7 Nucleus/Cytoplasm PLAG1, PLAG2 Sin3, NCOR2
IIB HDAC6 Mostly Cytoplasm a-tubulin, HSP90, SHP, SMAD ____
HDAC9 Nucleus/Cytoplasm ____ ____
HDAC10 Nucleus/Cytoplasm ____ NCOR2
III NADþ e SIRTs Nucleus/Cytoplasm/Mitochondria Not given Not given
IV Zn2þ e HDAC11 Nucleus/Cytoplasm ____ ____
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 217

Fig. 2. Schematic representation of histone deacetylase (HDAC) super family members.

acetylation [33]. Radiosensitization by SAHA along with capecita- 37 patients having various type of tumors [10]. The FDA approval
bine under hypoxia reduced in vitro colonogenicity and inhibited for romidepsin in the treatment of relapsed or refractory peripheral
in vivo tumor growth in colorectal carcinoma xenograft models T-cell lymphomas (PTCL) was mainly established from the data
[34]. However, SAHA in combination with CHOP (cyclophospha- obtained from the pivotal GPI-06-0002 trial. In the trial of 131
mide, doxorubicin, vincristine and prednisone) exhibited weak patients with relapsed or refractory PTCL, the treatment with
prognosis in the clinical trials while treating patients having un- romidepsin provided better beneficial results [39].
treated peripheral T-cell lumphoma (PTCL) [35]. SAHA was also
found to be a potent anticancer agent effective against gastroin-
3.3. Belinostat (Beleodaq®)
testinal cancer [36]. It was interesting to note that SAHA basically
acts as a pan-HDAC inhibitor due to inhibition of all HDACs (class I,
Belinostat (FDA-03) is a HDAC inhibitor comprising a
II and IV HDAC) but not the NADþ-dependent sirtuins or class III
sulphonamide-based hydroxamate structure. It is chemically (2E)-
HDACs [19,37].
N-hydroxy-3-[3-(phenylsulfamoyl)-phenyl]-prop-2-enamide
which was primarily metabolized by the enzyme UGT1A1 and to
some extent by CYP2A6, CYP2C9 and CYP3A4 as observed in the
3.2. Romidepsin (Istodax®)
in vitro studies (Table 2, Fig. 3). The in vitro studies revealed that
Belinostat also inhibited the enzymatic activity of HDAC8
Romidepsin (FDA-02) (Table 2, Fig. 3), reported in 1994 in the
(IC50 ¼ 3530 nM) along with CYP2C8 and CYP2C9 [26,40].
scientific literature performed by a group of researchers from
Fujisawa Pharmaceutical Company in Tsukuba of Japan. It was
isolated from Chromobacterium violaceum culture obtained from a 3.4. Panobinostat (Farydak®)
soil sample [38]. In the phase I clinical study, the maximum toler-
ated dose (MTD) of romidepsin was identified when carried out on Panobinostat (FDA-04), another hydroxamate HDAC inhibitor,
218 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Fig. 3. Structures of FDA approved HDAC inhibitors (FDA-01 e FDA-06).

Table 2
Histone deacetylase inhibitors (HDACIs) approved by the FDA for the treatment of cancer.

Sl No. HDAC inhibitor Therapeutic use Trade name Company name Year of approval Reference

(FDA-01) Vorinostat (SAHA) Cutaneous T-cell lymphoma Zolinza Merck 2006 [27,28]
(FDA-02) Romidepsin Cutaneous T-cell lymphoma Istodax Celgene 2009 [10,38]
(FDA-03) Belinostat Relapsed multiple myeloma Beleidaq Spectrum 2014 [40]
(FDA-04) Panobinostat Relapsed multiple myeloma Farydak Novartis 2015 [10]
(FDA-05) Chidamide (HBI-8000) Relapsed or refractory peripheral T-cell Epidaza Shenzhen-core biotechnology limited 2015 [41]
lymphoma and pancreatic cancer
(FDA-06) Pracinostat (SB939) Acute myeloid leukemia ____ Helsinn Group and MEI Pharma 2016 [42]

acts as a non-selective pan HDAC inhibitor (Table 2, Fig. 3). It was Chidamide is the only HDAC inhibitor approved by Chinese Food
approved by USFDA for the treatment of multiple myolema [1,10]. It and Drug Administration (CFDA) unlike other approved HDAC in-
also exhibited activity against haematological and solid cancers in hibitors which are approved by USFDA [43].
phase II clinical trial [10]. It is important to note that all these FDA approved drugs are
nonselective pan-HDAC inhibitors those possess a number of
adverse effects and therefore, their use is limited to particular type
3.5. Pracinostat (SB939)
of cancers. However, HDAC8 has a direct relation in progression of
several types of cancers. Therefore, specific and selective HDAC8
Pracinostat (FDA-05) was found to inhibit class I, II, IV HDACs
inhibitors may be effective to combat these particular cancers with
without inhibiting class III HDACs and HDAC6. It had no effect on
lower adverse effects. Moreover exept these six approved HDAC
other zinc dependent enzymes (Table 2, Fig. 3). It accumulated into
inhibitors (FDA01-FDA06), a number of promising ligands such as
the tumor cells and subsequently exhibited histone deacetylase
Givinostat, Tacedinaline, Entinostat, Resminostat, etc. [10] also have
inhibition followed by acetylation of histones, remodelling of
entered into the preclinical and clinical studies. These compounds
chromatin, transcription of tumor suppressor genes as well as
bear a possibility to become an important weapon in combating
apoptosis in tumor cells [41]. Pracinostat exhibited promising
various cancers and designing newer molecules through modifi-
in vitro activity against HDAC8 isoform (HDAC8 IC50 ¼ 140 nM) [42].
cation of these compounds may provide excellent HDAC8
inhibition.
3.6. Chidamide (HBI-8000/Epidaza®)

Chidamide (FDA-06), a benzamide derived HDAC inhibitor, is 4. Role of HDAC8 in different diseases
also known as HBI-8000 that inhibits class I HDACs (HDAC1,
HDAC2, HDAC3 and HDAC8) as well as Class IIb HDAC (HDAC10). It The overexpression of HDAC8 has been observed in a variety of
is being used for the treatment of pancreatic cancer (Table 2, Fig. 3). cancers (colon, lung, breast, pancreatic cancers and neuroblastoma
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 219

in childhood cancer) (Fig. 4). neuroblastoma models as far as toxicity and efficacy in pre-clinical
The level of HDAC8 expression is almost identical in cancer cells trial was concerned [53e55] but inhibition of the other HDAC iso-
and in healthy tissues with a tendency to increase its expression in forms in neuroblastoma causes apoptosis of cells [56e58]. By
cancer cells [15,44]. The RNA interference (RNAi) mediated mediating the ubiquintin mediated-degeneration of telomerase 1B
knockdown of HDAC8 is helpful to restrain the proliferation of (hEST1B) protein, HDAC8 also affects the human telomerase ac-
human cervical, colon and lung cancer cell lines while the upre- tivity basically through a phosphorylation function-dependant
gulation encourages cell proliferation and also prevents apoptosis process [4]. Moreover, along with inv(16) protein and other co-
in hepatocellular cancer [14,45]. It has also been reported that repressors, HDAC8 activates acute myeloid leukemia-I (AML-I) by
transcription factor SRY-box 4 (SOX 4/sex determining region Y) regulating the genes responsible for the formation of abnormal
activates HDAC8 promoter which is responsible for the expression haematopoetic cell proliferation, similar to leukemia [59,60].
of HDAC8 in cancer cell lines [46e49]. PCI-34051 (9) [Fig. 8] was HDAC8 also reduces the tumor suppressor expression by upregu-
found to be effective for the treatment of T-cell lymphoma and lating the Tap73 expression [61]. HDAC8-mediated downregulation
leukemia. It was also found as specific cytotoxic agent for Jurkat, of suppressor of Cytokine Signaling 1/3 (SOSC1/3) expression cau-
HuT78 and Molt-4 cell lines. PCI-34051 (9) induces calcium- ses the alteration of Janus kinase 2/signal transduction and acti-
mediated caspase-dependant apoptosis [50]. In case of solid state vation of transcription (JAK2/STAT) while inhibition of HDAC8
cancer cell lines like A549 (lung), Rko (colon), U87 (glioma) and provides hindrance in SOS1/3 mediated cell proliferation along
MCF-7 (breast), the same treatment cannot be used to induce with the suppression of haematopoetic cell activity in myelopro-
apoptosis [50]. liferative neoplasms [62,63]. Bcl-2 Modifying Factor (BMF) has a
Upregulation of HDAC8 results in an overexpression of HDAC1 significant role as it executes the methyl selenopyruvate-mediated
and HDAC6 in invasive breast tumor cells to promote invasion. It apoptosis of cells where methyl selenopyruvate works as HDAC8
also causes the overexpression of matrix metalloproteinase-9 antagonist. The HDAC8-STAT3 mediated regulations of BMF have a
(MMP-9) as evidenced in MCF-7 breast cancer cells [51]. The role crucial role in cellular apoptosis and cell proliferation [5,6]. A drop
of HDAC8 in the formation of the crenial neural crest cell of mice in HDAC8 concentration or inhibition of HDAC8 enzymatic activity
skull has also been observed [4]. Deletion of HDAC8 in mice causes reveals its effect on the proliferation of cells having p53 mutation in
perinatal lethality due to instability of the skull [4]. In case of neural them. This phenomenon clearly pointed out the influence of HDAC8
crest-derived neuroblastoma, the expression of HDAC8 can be inhibitors for p53 mutation-carrying tumor adjuvants [4]. HDAC8
correlated with tumor stage [52]. Specific inhibition of HDAC8 over also helps to mould platform, carries the cytokines to the nucleus,
non-selective pan-HDAC inhibition executed promising results in to act on a specific locus of DNA and affecting the cancer sensitive

Fig. 4. Implications of HDAC8 in different Diseases.


220 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

genes expression [4]. forms hydrogen bond with the catalytic Y306 of the human HDAC8
The vital role of HDAC8 enzyme is also found in parasitic dis- [71e74]. The tandem histidine pair namely H142 and H143 amino
eases like schistosomasis (by Schistosoma sp. or Flatworm) [4]. acid residues of HDAC8 is crucial for the stabilization of the
Therefore, HDAC8 can be targeted to combat schistosomasis. The transition-state [75]. H142 serves as the general base catalyst
significance of HDACs was elucidated while testing the Trichostatin whereas H143 serves as both the general acid and the general base
(TSA), a pan-HDAC inhibitor, into the schistosome larvae and adult catalysts [73] (Fig. 5).
parasites. Application of TSA led to the death of larvae and adult The metal-bound water molecule is deprotonated by the histi-
worms [4]. Nevertheless, effects of the deacetylation activity by dine to interact carbonyl function of acetyl-L-lysine. A tetrahedral
HDAC8 were also found in other viral infections like influenza A and intermediate is therefore formed (Fig. 5). To collapse the tetrahe-
Uukuniemi virus [64]. The in vitro studies revealed the HDAC8 dral intermediate, another histidine residue is found to protonate
dependant increased endocytosis, acidification and penetration of the amino-leaving group of the substrate acetyl-L-lysine. Finally,
influenza A and Uukuniemi virus into the host cells but the regu- products of lysine and acetate are formed (Fig. 5) [75]. Recently, the
lation of HDAC1 provides the opposite results [64]. The down- structural and functional importance of the conserved glycine-rich
regulation of HDAC8 and/or HDAC3 using silencing RNA (siRNA) loop G302GGGY in the active site of HDAC8 has been explored [73].
affects the pathway involving conversion of early endosomes to This catalytic loop is crucial since missense and nonsense muta-
lysosomes causing in profitless endocytosis and inadequate acid tions in this HDAC8 loop may lead to Cornelia de Lange Syndrome
transformation for viruses resulting into reduction of viral infection (CdLS). The CdLS, a multiple congenital disease [76], results from
[64]. gene mutations regulated by HDAC8 [16] and other proteins [65]. A
Lower concentration of HDAC8 induces remodelling/refashion- child diagnosed with CdLS where different missense mutations in
ing of microtubular network leading to the centripetal movements. HDAC8 such as G304R has been identified [65,73]. Porter et al. [73]
Hence, this phenomenon reveals the essential role of HDAC8 in suggested that the glycine residues in the G302GGGY segment are
centrosomal designing and endosomal mobility mediated by very crucial for the maintenance of the Y306 residue. Also, all
cohesion. The influence of the centrosomal self-deacetylation of the glycine residues especially G304 and G305 provide flexibility on the
microtubules and/or of regulators to the centrosomal cohesion for G302GGGY loop as suggested by the molecular dynamics (MD)
viral entry into the host body is still a cloudy concept [4]. HDAC8, in simulation study [73]. This results in a transition of Y306 between
congenital malformation, has been spotted as a crucial enzyme for ‘in’ and ‘out’ conformations. This transition of the catalytic Y306
deacetylation of SMC3. This is a cohesin subunit which helps in residue might be modulated by the HDAC8 inhibitors. Residues like
binding two sister chromatids [4,16]. In the early mitotic phase of G151, F152, H180, F208, M274 and Y306 of HDAC8 construct the
cell division, HDAC8-mediated deacetylation of chromatin releases active site tunnel walls [77]. These mentioned amino acid residues
acetylated SMC3 that induces the recycling of cohesin into the are generally hydrophobic and conserved in nature for the HDAC
succeeding cell cycles. However, the absence of HDAC8 activity class I isoforms. These help during the binding of an inhibitor
enhances the amusement of acetylated cohesin leading to the containing bulky hydrophobic linker to the HDAC8 enzyme. The
diminished chromatid affinity. This produces defective manifestation of the movement of F152, away from M274, occurs
transcription-mediated symptoms like CdLS [16]. The relation be- which might open a distinctive HDAC8 specific sub pocket [77].
tween the proximal presence of mutated residues of the enzyme to Although belonging to the class I HDAC family, there are a
the active site and the reduction of HDAC8 activity has also been number of differences present between other HDACs and HDAC8.
proposed. The N-(phenyl carbamothioyl), an activator of HDAC8 Those are following:
specifically, restores the activity of mutants apart from those mu-
tants present near the active site. For the substrate binding, the  Human HDAC8 is an X-linked protein which does not depend on
required stabilization of enzyme conformation is done by the any co-complexes for the activity [78,79].
activator which is prevented due to the proximity of the mutant  The C-terminal protein binding domain of the other HDACs of
and the active site [65]. the class is absent for HDAC8 whereas the L1 loop of the
The expression of HDAC8 has been specifically noticed in several enzyme, closest to the active site is eminently flexible in nature
cancers like colon, lung and breast cancers, hepatocellular carci- and can goes under conformational changes specific to the
noma, pancreatic and gastric cancers, acute lymphocytic leukemia substrate [18].
(ALL) and acute myeloid leukemia (AML) along with childhood  The L1 and the L6 loops of HDAC8 along with its catalytic
neuroblastoma [5,15,62,63,66]. It is observed that RNAi-mediated tyrosine forms a pocket specific to HDAC8 isoform which re-
HDAC8 knockdown inhibits the cellular proliferation in several quires a “L” shaped structure of inhibitors for selective binding
cancers (such as human colon and lung cancer, gastric adenocar- with HDAC8. But for the other HDAC isoforms the presence of
cinoma, leukemia and cervical cancer) [66e68]. Upregulation of the L1 and L6 loops hinders the binding of inhibitors with “L”
HDAC8 is found to inhibit the apoptosis and therefore, induces cell shaped structure [80].
proliferation in hepatocellular and castric carcinoma cell line [15].  HDAC8 can also be differentiated from the other class I HDACs in
The inv(16) fusion protein in association with HDAC8 represses the size and composition of its N-terminal L1 loop which shapes
AML-I regulated gene transcription [69]. Moreover, the knockdown a large portion in one side of the active site and extended to the
of HDAC8 tends to the repression of HoxA5 that decreases the protein surface [4,81].
HoxA5 dependant wild type mutant p53 expression [70]. The  The non-histone proteins like cohesin, ERR-a or cortacin are
ectopic expression of HDAC8 enhances the p53 transcription associated with the HDAC8 enzyme. This provides HDAC8 the
whereas the HDAC8 inactivation may be effective for p53 mutation ability to control processes like microtubular integrity, chro-
containing tumor cells [7]. matid separation, muscle contraction and energy homeostasis
[11,21].
5. Structure of HDAC8  The amino acid forming the HDAC8 pocket is also similar with
other HDACs of class I exept that HDAC8 contains a methionine
In 2004, the first crystal structure of HDAC8 enzyme was whereas the other isoforms contains a leucin in place of meth-
explored [18]. It was observed that the carbonyl group of the sub- eonine [18].
strate acetyl-L-lysine coordinates to the Zn2þ ion and subsequently
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 221

Fig. 5. Schematic representation of the histone deacetylase 8 mechanisms.

 From the structural aspects, HDAC8 consists of a nuclear local- 6. An insight into the HDAC8 inhibitors
ization sequence present between the catalytic domain of the
enzyme and serine binding motif in an end of the catalytic Recently, HDAC8 is emerged as an important druggable target
domain. HDAC8 is the smallest isoform among the HDACs pre- and HDAC8 inhibitors are useful to treat various diseases [1e3].
sent into the class I of the HDAC family. The negative regulation However, no isoform-specific HDAC8 inhibitor is available in
of the HDAC8 catalytic function through the cAMP dependent dosage form. Therefore, designing and discovery of potent HDAC8
PKA phosphorylation of S39 proximal to the catalytic domain inhibitors is a challenging task in the field of anticancer
[11]. therapeutics.
The chemical structures of reported HDAC8 inhibitors generally
Inside the catalytic domain, seven loops are present. The Zn2þ possess three pharmacophoric features. These contain a surface
ion is present between the Loop7 and Loop4 of the domain. The recognition function or a cap group, a hydrophobic cavity-binding
Zn2þ ion is conserved with D(A,V,L,F)Hx~100D sequence where linker motif and a zinc-binding group (ZBG). This hydrophobic
D178, H180 and D267 (D,H and D of the sequence) ligate with the cap group may be able to confer HDAC isoform-specific surface
catalytic Zn2þ ion. It has also been reported that the substitution of recognition at the binding site rim. Hydroxamates are generally
HDAC8 Zn2þ ion with Fe2þ ion resulted enhanced HDAC8 activity. utilized to synthesize HDAC8 inhibitors as this functional group
This indicated to possibility of HDAC8 activities as a ferrous enzyme acts as a strong ZBG [1]. Hydroxamic acids as ZBG express high
for in vivo activity (Fig. 6) [81]. affinity toward the catalytic Zn2þ and therefore, coordinate strongly
According to Somoza et al. [18], HDAC8 active site consists of a to the Zn2þ ion at the enzyme active site. Mostly, compounds
long and narrow tunnel where the walls are comprised of F152, bearing hydroxamate moiety exhibit HDAC8 inhibiton in lower
F208, H180, G151, M274 and Y306 basically hydrophobic in nature. nanomolar concentration. Nevertheless, the strong zinc chelating
In the end of the hydrophobic tunnel, the Zn2þ ion was bound to the ability of hydroxamic acid offers nonselective HDAC inhibition or
carboxylate oxygen atoms of D178 and D267 and H180 heterocyclic interacts with other zinc-dependent metalloenzymes such as
nitrogen atom. The Y306, D183 and D176 have also been seen aminopeptidase N (APN) and matrix metalloproteinases (MMPs).
important in the active site. From SAHA bound human HDAC8 Moreover, these hydroxamate compounds display poor absorbtion
crystal structure (PDB ID: 1T69) of the active site (Fig. 6), it has been property in vivo. In these circumstances, researchers have paid their
seen that the hydroxamate group of the compound forms cheletion attention to design non-hydroxamate (weak Zn2þ chelator in
with the HDAC8 catalytic Zn2þ ion whereas the phenyl cap moiety comparison to hydroxamate) HDAC inhibitors. These non-
has been fit into the HDAC8 pocket. H142 and H143 as well as hydroxamate ZBGs contain different structural motifs such as car-
phenylalanine are forming the wall of the lysine binding channel boxylic acids, boronic acids, thiols, oximes, trifluoromethyl ketones,
while the Y306 and D178 interact with the ZBG of SAHA [81]. A list hydroxy-pyridin-2-thiones and b-lactams (Fig. 7). Compounds
of reported crystal structures of histone deacetylase 8 as per Protein possessing these non-hydroxamate ZBGs also yield effective HDAC8
Data Bank (PDB) [82] record is depicted in Table 3. inhibition.
222 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Fig. 6. Contours of different the ligand binding site of HDAC (HDAC8, HDAC6 and HDAC4) isoforms along with their catalytic Zn2þ and important amino acid residues for ligand
binding. (A) and (B) Ligand (SAHA) bound crystal structure of HDAC8 (PDB: 1T69), (C) Ligand binding site of HDAC4 (PDB: 2VQJ) and (D) Ligand binding site of HDAC6 (PDB: 5W5K).

In this section, hydroxamate and non-hydroxamate ZBGs con- HDAC8 inhibitory activity. A number of hydroxamate and mer-
taining potential HDAC8 inhibitors are discussed and their selec- captoacetamide small molecule HDAC inhibitors were synthesized
tivity and nonselectivity have also been highlighted in details by Kozikowski et al. [86]. The thiol and carbonyl groups of these
(Table 4). compounds were found to contribute properly during Zn2þ chela-
A series of cyclic hydroxamic acid containing peptides (CHAPs) tion. Compared to thiol, the hydroxamate group was proven to be
was reported by Furumai and co-workers [83]. These synthesized better ZBG (3, IC50 ¼ 1,050 nM) (Fig. 8).
CHAP derivatives enhanced the G1 cell population and decreased Several active bis-(aryl)-type HDAC inhibitors were designed
the growth of HeLa cell at S phase. Importantly, the CHAP31 (1) and synthesized by Moradei et al. [87]. The 2-amino-5-aryl anilide-
[Fig. 8] displayed good HDAC inhibition (IC50 ¼ 3.32 nM) obtained based benzamide derivatives exhibited good HDAC1 and HDAC2
from B16/BL6 melanoma. inhibition. Among these compounds, LAQ-824 (4) showed HDAC8
Some cyclic tetrapeptide compounds having pentafluoroethyl inhibitory potency with an IC50 value of 400 nM (Fig. 8). Arts et al.
and trifluoromethyl ketone ZBGs were synthesized by Jose and co- [88] developed some 5-pyrimidinyl hydroxamic acid derivatives.
workers [84]. These compounds were tested against different Among these compounds, R306465 (5) exhibited remarkable
HDACs along with p21 promoter assay. Compound containing the class I HDAC inhibition. R306465 (5) [Fig. 8] showed 16-fold better
pentafluoroethyl ketone moiety (HDAC8 IC50 ¼ 780 nM) exhibited HDAC8 inhibition (IC50 ¼ 23 nM) in vitro than SAHA (FDA-01,
lower HDAC8 inhibitory activity compared to the corresponding IC50 ¼ 370 nM) [Fig. 3]. By inspecting the ligand-bound HDAC8 X-
trifluoromethyl ketone derivative (HDAC8 IC50 ¼ 370 nM). More- ray crystal structures, KrennHrubec et al. [89] designed some new
over, compound possessing the trifluoromethyl ketone function hydroxamates different from SAHA (1). These compounds, lacking
with a thioether linker was found to be the most effective com- the linker function, were a few more folds selective toward HDAC8
pound (2) (HDAC8 IC50 ¼ 230 nM). It may be probably due to the over HDAC1 and HDAC6. The right angle structural conformation of
proper coordination with the catalytic Zn2þ ion (Fig. 8). Compounds the best active compound in series (6, HDAC8 IC50 ¼ 300 nM) may
containing b-lactam function was reported to display better HDAC properly fit into the HDAC8 subpocket resulting in stronger HDAC8
inhibition than sodium butyrate [85]. Such type of compounds also binding (Fig. 8).
triggered NF-kB mediated cytotoxicity during modulation of The triazolylphenyl-containing HDAC inhibitors, synthesized by
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 223

Table 3
List of reported crystal structures of histone deacetylase 8 as per Protein Data Bank (PDB) record.

PDB Organism Inhibitor Substrate Form Residue Res. (Å) Year

1T64 H. sapiens TSA – WT 754 1.90 2004


1T67 H. sapiens M344 – WT 377 2.31 2004
1T69 H. sapiens SAHA – WT 377 2.91 2004
1VKG H. sapiens CRA 19156 – WT 754 2.20 2004
1W22 H. sapiens PSTB-Hydroxamate – WT 754 2.50 2004
2V5W H. sapiens e Acetylpeptide Y306F 791 2.00 2007
2V5X H. sapiens Hydroxamate – S39D 776 2.25 2007
3EW8 H. sapiens M344 – D101L 388 1.80 2008
3EWF H. sapiens e Acetylpeptide H143A 1572 2.50 2008
3EZP H. sapiens M344 – D101 N 776 2.65 2008
3EZT H. sapiens M344 – D101E 776 2.85 2008
3FO6 H. sapiens M344 – D101A 776 2.55 2008
3F07 H. sapiens APHA – WT 1164 3.30 2008
3F0R H. sapiens TSA – WT 1164 2.54 2008
3MZ3 H. sapiens M344 – WT/Co2þ 778 3.20 2010
3MZ4 H. sapiens M344 – D101/Mn2þ 778 1.85 2010
3MZ6 H. sapiens M344 – D101/Fe2þ 389 2.00 2010
3MZ7 H. sapiens M344 – D101/Co2þ 389 1.90 2010
3SFF H. sapiens Amino acid derivative – WT 378 2.00 2011
3SFH H. sapiens Amino acid derivative – WT 378 2.70 2011
3RQD H. sapiens Largazole thiol – WT 788 2.14 2011
4BZ5 S. mansoni e Acetylpeptide WT apo 1784 1.78 2013
4BZ6 S. mansoni SAHA – WT 1784 2.00 2013
4BZ7 S. mansoni M344 – WT 1784 1.65 2013
4BZ8 S. mansoni J1038 – WT 1784 2.21 2013
4BZ9 S. mansoni J1075 – WT 1784 2.00 2013
4CQF S. mansoni Marcapto acetamide – WT 1784 2.30 2014
4QA0 H. sapiens SAHA – C153F 778 2.24 2014
4QA1 H. sapiens M344 – A188T 1156 1.92 2014
4QA2 H. sapiens SAHA – I243 N 778 2.38 2014
4QA3 H. sapiens TSA – T311 M 778 2.88 2014
4QA4 H. sapiens M344 – H344R 389 1.98 2014
4QA5 H. sapiens e Acetylpeptide A188T/Y306F 788 1.76 2014
4QA6 H. sapiens e Acetylpeptide I243 N/Y306F 788 2.05 2014
4QA7 H. sapiens e Acetylpeptide H334R/Y306F 394 2.31 2014
4RN0 H. sapiens Largazole analog – S39D 778 1.76 2015
4RN1 H. sapiens Largazole analog – S39D 778 2.18 2015
4RN2 H. sapiens Largazole analog – S39D 778 2.39 2015
5D1B H. sapiens TSA – G117E 778 2.90 2015
5D1C H. sapiens e Tetrapeptide D233G/Y306F 790 1.42 2015
5D1D H. sapiens e Tetrapeptide P91L/Y306F 790 2.01 2015
5DC5 H. sapiens M344 – D176 N 778 1.94 2016
5DC6 H. sapiens e Tetrapeptide D176 N/Y306F 790 1.55 2016
5DC7 H. sapiens e Tetrapeptide D176/Y306F 790 2.30 2016
5DC8 H. sapiens e Tetrapeptide H142a/Y306F 790 1.30 2016
5FUE S. mansoni 3- benzamidobenzo hydroxamate – WT 1784 2.19 2016
5BWZ H. sapiens Droxinostat – S39E 778 1.59 2016
5FCW H. sapiens Hydroxamic Acid – WT 766 1.97 2016
5THS H. sapiens M344 – G302A 778 1.90 2016
5THT H. sapiens M344 – G303A 1556 2.40 2016
5THU H. sapiens M344 – G304A 778 1.95 2016
5THV H. sapiens M344 – G304A 778 1.86 2016
6HQY S. mansoni PCI-34051 – WT 1788 2.50 2018
6HRQ S. mansoni NCC-149 – WT 1788 1.85 2018
6HSF S. mansoni PCI-34051 – H292 M 1788 1.90 2018
6HSG S. mansoni NCC-149 – H292 M 1788 1.85 2018
6HSH S. mansoni Quisinostat – WT 1788 1.54 2018
6HSK H. sapiens Quisinostat – Mutant 160 2.10 2018
6HSZ S. mansoni Benzohydroxamate inhibitor 2 – WT 1788 2.37 2018
6HT8 S. mansoni Benzohydroxamate inhibitor 3 – WT 1788 2.50 2018
6HTG S. mansoni Benzohydroxamate inhibitor 4 – WT 1788 1.94 2018
6HTH S. mansoni Benzohydroxamate inhibitor 5 – WT 1788 1.95 2018
6HTI S. mansoni Benzohydroxamate inhibitor 6 – WT 1788 1.69 2018
6HTT S. mansoni Benzohydroxamate inhibitor 7 – WT 1788 1.75 2018
6HTZ S. mansoni Benzohydroxamate inhibitor 8 – WT 1788 1.84 2018
6HU0 S. mansoni Benzohydroxamate inhibitor 9 – WT 1788 1.75 2018
6HU1 S. mansoni Benzohydroxamate inhibitor 10 – WT 1788 2.00 2018
6HU2 S. mansoni Benzohydroxamate inhibitor 11 – WT 1788 1.99 2018
6HU3 S. mansoni Triazole hydroxamate inhibitor – WT 1788 1.65 2018

This table is based on the table reported by Amin et al. [1] and Chakrabarti et al. [4].
Data taken from Ref. [82].
224 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Fig. 7. Different reported zinc binder groups namely hydroxamates, trifluoromethyl ketones, thiols, boronic acids, carboxylic acids, oximes, hydroxy-pyridin-2-thiones and b-
lactams for HDAC8 inhibition.

Chen et al. [90] consisted of an interesting modification in their activity against HDAC6, the structural information obtained from
cap region. Most of these compounds of this series were weaker these chiral inhibitors can be useful to design effective HDAC8 in-
HDAC8 inhibitors among which compound 7 showed an IC50 of hibitors with higher potency (Fig. 9).
406 nM. Though the selectivity profile was not much interesting, Giannini et al. [95] synthesized a number of N-hydroxy-(4-
these compounds exhibited good selectivity towards HDAC6 oxime)-cinnamate-derived cytotoxic HDAC8 inhibitors. Com-
(Fig. 8). Kozikowski et al. [91] synthesized another series of pounds having bulky heteroaryl- or aryl-substituted oxime group
phenylisoxazole-containing hydroxamates as HDAC inhibitors with at the para position of the cinnamic acid moiety provided better
greater selectivity towards HDAC6. These compounds also dis- activity towards HDAC8 over other isoforms. The 4-nitrobenzyl
played some class I and class IIb HDAC inhibitory activity. Experi- group containing compound (13, IC50 ¼ 70 nM) [Fig. 9] showed a
mental evidences suggested that these phenylisoxazole-based minimum of 2-fold HDAC8 selectivity over other HDAC isoforms.
compounds were potential HDAC8 inhibitors. Among these com- Moreover, the 2-morpholinylmethyl analog (14) was also providing
pounds of the series, compound 8 (Fig. 8) yields the highest HDAC8 a good HDAC8 activity (IC50 ¼ 235 nM) [Fig. 9]. Canzoneri et al. [96]
inhibition (IC50 ¼ 938 nM). Balasubramanian et al. [50] reported a discovered some nuclear localization signal peptide (NLS)-derived
compound named PCI-34051 (9, HDAC8 IC50 ¼ 10 nM) which HDAC inhibitors. The lysine-enriched sequence of the NLS might
exhibited highly potent HDAC8 inhibitory activity (Fig. 8). copy the histones core N-terminal Lys-tail. These NLS-derived
Andrianov et al. [92] synthesized some hydroxamate-based HDAC inhibitors helped in nuclear delivery and localization. They
derivatives those had more potency against HDAC1 and HDAC6 synthesized NLS-derived peptide and 1, 2, 3-triazole moiety con-
compared to HDAC8 (10, IC50 ¼ 779 nM) [Fig. 9]. taining hydroxamates. These experiments showed that the length
During in vivo pharmacodynamic analysis conducted on more of the linker was an important component for inhibition of HDAC8
than 100 pyrimidyl hydroxamates, a molecule (JNJ-26481585, 11) enzymatic activity (15, IC50 ¼ 1,243 nM) [Fig. 9].
was recognized by Arts et al. [93]. This molecule showed good Oyelere et al. [97] found out a new class of nonpeptide HDAC
HDAC1 (IC50 ¼ 110 mM) and HDAC2 (IC50 ¼ 330 mM) inhibitions but inhibitors. These compounds were derived from the macrocyclic
highly potent and selective HDAC8 inhibition (11, IC50 ¼ 4.26 nM) backbone of macrolide antibiotics. Compounds containing these
[Fig. 9]. Smil et al. [94] reported some HDAC6 selective chiral scaffolds are not much active against HDAC8 (16, IC50 ¼ 994 nM)
HDAC8 inhibitors containing the 3, 4-dihydroquinoxaline-2-(1H)- but effective against HDAC1 and HDAC2 (Fig. 9). Montero et al. [98]
one and the piperazine-2,5-dione aryl hydroxamate (12, reported a pool of b-amino acid-containing compounds having
IC50 ¼ 210 nM) scaffolds. Though these compounds dedicated their structural similarity with the cyclic tetrapeptide. Changes in the
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 225

Table 4
List of hydroxamate and non-hydroxamate ZBGs containing potential HDAC8 inhibitors.

Sl No. Scaffold Most active isoform Best HDAC8 inhibitory activity References

Cpd No. IC50 (nM)

1 Peptide containing cyclic hydroxamic acid derivative e 1 3.32 [83]


2 Trifluoro methyl and tetrafluoro ethyl ketone based cyclic HDAC1, HDAC8 2 230 [84]
tetrapeptide derivative
3 Mercapto acetamide based hydroxamate HDAC6 3 1050 [86]
4 Amino phenyl benzamide based hydroxamate HDAC1, HDAC2 4 400 [87]
5 5-pyrimidinyl based hydroxamate HDAC1 5 23 [88]
6 4-napthyl-phenyl based linkerless hydroxamate HDAC8 6 300 [89]
7 Triazolyl based hydroxamate HDAC6 7 406 [90]
8 Phenyl iso-oxazole based hydroxamate HDAC6 8 938 [91]
9 Aryl hydroxamate derivative HDAC8 9 10 [50]
10 Amide derivative Pan HDAC 10 779 [92]
11 Pyrimidyl hydroxamate HDAC1, HDAC2 11 4.26 [93]
12 3,4-dihydroquinoxaline-2-(1H) one derivative HDAC6 12 210 [94]
13 Piperazine-2,5-dione aryl hydroxamate HDAC6 12 210 [94]
14 N-hydroxy-(4-oxime)-cinnamate based hydroxamate HDAC8 13 70 [95]
15 2-morpholinyl ethyl hydroxamate HDAC8 14 235 [95]
16 1,2,3-triazole based NLSP hydroxamate HDAC1, HDAC2 15 1243 [96]
17 Non peptide macrocyclic derivative HDAC1, HDAC6 16 994 [97]
18 Cyclic a/b tetra peptide derivative HDAC8 17 120 [98]
19 Benzyloxy phenyl carbamoyl hydroxamate HDAC-8 18 2710 [99]
20 a/b Cyclic tetra peptide HDAC6 17 133 [100]
21 Boronic acid-based derivative HDAC6 19 6600 [101]
22 2-piperazinyl-5- pyrimidyl hydroxamate Pan HDAC 20 e [102]
23 Acetyl urea based hydroxamate HDAC1 21 270 [103]
24 Triazole-4-yl phenyl based hydroxamate Pan HDAC 22 1190 [104]
25 Tricyclic ketolide based hydroxamate HDAC8 23 544.6 [105]
26 Cyclo peptide based hydroxamate HDAC1, HDAC2, HDAC8 24 23,000 [106]
27 Tetrahydro isoquinoline based hydroxamate HDAC8 25 580 [107]
28 Tetrahydro isoquinoline based hydroxamate HDAC6 26 47 [108]
29 Tetrahydro isoquinoline based hydroxamate HDAC6 27 146 [109]
30 Benzohydroxamate derivative HDAC8 28 23 [110]
31 Oxime amide derivative Pan HDAC 29 22,600 [111]
32 Gamma lactym derivative HDAC6 30 119.7 [112]
33 Gamma and del- lactum based hydroxamate HDAC3 31 17 [113]
34 1,4-dithia-7-aza spiro[4.4] nonane-8-carboxylate based derivatives HDAC8 32 21 [114]
35 Short chained(phenyl butyryl) based hydroxamate HDAC1 33 4000 [115]
36 SAHA derivative with osthole derived CAP region HDAC1 34 267.85 [116]
37 Ferrocinyl based JAHA derivative Pan HDAC 35 2 [117]
38 Amino acid derivative HDAC1, HDAC2 36 90 [70]
39 SAHA derivative HDAC6 37 220 [118]
40 2,2,3,3,4,4,5,5,6,6,7,7-dodecafluoro-N-hydroxy octane di-amide derivative HDAC1 38 1,400a [119]
41 Tetrapeptide derivatives without ZBG HDAC8 39 17 [120]
42 Cu(I) catalyzed azide alkyne cyclo addition library of HDAC8 inhibitors HDAC8 40 70 [121]
43 Alkoxamide linked hydroxamate HDAC4, HDAC5, HDAC6 41 893 [122]
44 Phenyl glycin based hydroxamate HDAC8 42 967 [123]
45 Quinazoline-4-one derivative HDAC6 43 420 [124]
46 3-Hydroxypyridin-2-thione derivative HDAC6 44 800 [125]
47 N-Methylpyrrole (Py)-N-methylimidazole (Im) polyamides HDAC1, HDAC2, HDAC8 45 130 [126]
(pips) derivative
48 Amide-linked p-substituted phenyl hydroxamate HDAC6 46 689 [127]
49 Carrbostyril derivative Pan HDAC 47 6 [128]
50 Triazole derivative HDAC8 48 53 [130]
51 Aminotetralin derivative HDAC6, HDAC8 49 30 [132]
52 1,2,4-oxadiazole containing 2-amino benzamide derivative HDAC1, HDAC2 50 200 [133]
53 Benzohydroxamate derivative HDAC6 51 3 [134]
54 Pteroate hydroxamide derivative HDAC6 52 581 [135]
55 Benzothiophene based hydroxamate derivative HDAC6 53 1400 [136]
56 Benzohydroxamate derivative HDAC6 54 376 [137]
57 Suberonyl hydroxamate derivative HDAC8 55 119 [139]
58 Glutamic acid derivative HDAC8, MMP-2 56 2890 [8]
59 Aminopyrrolidinone derivative HDAC6 57 80 [140]
60 1-hydroxypyridine-2-thione derivative HDAC8 58 980 [141]
61 2,5-Disubstituted-1,3,4-oxadiazole derivative HDAC8 59 98 [25]
62 Triazole analog HDAC8 60 0.8 [142]
63 Tetrahydro isoquinoline based hydroxamate HDAC1, HDAC3 61 44 [143]
64 N1-hydroxy tetrapthalamide based derivative HDAC8 62 5500 [26]
65 Propargylamine derivative HDAC6 63 417 [144]
66 Meta-sulfamoyl N-hydroxybenzamide derivative HDAC8 64 50 [145]
67 iso-Combretastatin based derivatives HDAC8 65 60 [146]
a
HDAC8 Kd value.
226 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Fig. 8. Structures of HDAC8 inhibitors (1-9).

Fig. 9. Structures of HDAC8 inhibitors (10-17).


S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 227

positions of b-3-amino acid, Zn2þ chelating amino acid group, the potency (21, IC50 ¼ 270 nM) [Fig. 10]. In the series of the triazole-4-
amino acid chirality and the alkylation of backbone amide nitrogen yl phenyl based hydroxamate compounds synthesized by He et al.
atom around the macrocyclic feature, were made to design selec- [104], it was noticed that these compounds showed poor HDAC8
tive HDAC inhibitors. Among these compounds, cpd 17 (Fig. 9) inhibition. Though these compounds were pan-HDAC inhibitors,
resulted in potent HDAC8 inhibitory activity (IC50 ¼ 120 nM). these were more selective towards HDAC3 and HDAC6. Among
A series of triazole-4-yl substituted hydroxamate molecules these compounds, compound with cyclohexyl function (22)
containing aryl cap group were synthesized and evaluated against showed a weak HDAC8 inhibition with an IC50 of 1190 nM (Fig. 10).
HDAC isoforms by He et al. [99]. Commonly, most of these com- Mwakwari et al. [105] introduced tricyclic ketolide imitating the
pounds were pan-HDAC inhibitors. However, some of them were macrocyclic peptide skeleton. These ketolide-based HDAC in-
active in nanomolar concentration against HDAC isoforms like hibitors showed HDAC1 and HDAC2 selectivity over other HDAC
HDAC1, HDAC3, HDAC6 and HDAC10. Cpd 18 provided poor HDAC8 isoforms. The best active compound was having a moderate HDAC8
inhibition (IC50 ¼ 2,710 nM) [Fig. 10]. inhibition (23, IC50 ¼ 544.6 nM) and might be targeted against the
Olsen and Ghadiri [100] developed some compounds bearing tumor-associated macrophages (Fig. 10).
cyclic tetrapeptide scaffold by keeping the same side chain motif Terracciano and coworkers [106] isolated a compound named
(17) [Fig. 9]. The hydroxamate group showed the most beneficial FR235222 (24, IC50 ¼ 23,000 nM) from the fermented fungus broth
effect while compared to other zinc binding groups. Compounds of Acremonium sp. This cyclopeptide molecule was synthesized by
containing boronic acid function also provided HDAC inhibition. As the same group of workers showed potency against HDAC (Fig. 10).
far as the boronic acid derivatives were concerned, Suzuki et al. In the development process of potent HDAC8 inhibitors, Zhang et al.
[101] pointed out that the (S)-conformation of the compound had a [106] synthesized and evaluated a pool of 1, 2, 3, 4-
significant importance regarding HDAC inhibition. The derivative tetrahydroisoquinoline-3-carboxylic acid compounds as potent
though selective towards HDAC6 possessed very poor HDAC8 in- HDAC8 inhibitors. Among these compounds, 25 (Fig. 10) exhibited
hibition (19, HDAC8 IC50 ¼ 6,600 nM) [Fig. 10]. Angibaud et al. [102] good HDAC8 inhibition (IC50 ¼ 580 nM). Again, they developed a
reported some substituted 2-piperazinyl-5-pyrimidylhydroxamic new group of HDAC8 inhibitors having the tetrahydroisoquinoline
acid derivatives as potent HDAC inhibitors. The biphenyl moiety as a linker [108]. Compound having the 4-methyloxyphenyl
substituted 2-piperazinyl-5-pyrimidylhydroxamate analog (20) group associated with the tetrahydroisoquinoline moiety and the t-
[Fig. 10] showed an excellent HDAC inhibition (IC50 ¼ 0.9 nM) with butoxycarbonyl (Boc) group showed higher contribution towards
a good antiproliferative property against human ovarian cancer HDAC8 inhibition (26, IC50 ¼ 47 nM) (Fig. 11).
A2780 cell line (IC50 ¼ 29 nM). This compound (20) or the modifi- Zhang et al. [109] further optimized the previously derived lead
cation of this compound can be effective against HDAC8. compound to synthesize a compound ZYJ-34c (27) having tetra-
Some acylurea-connected straight-chain hydroxamate de- hydroisoquinoline scaffold (Fig. 11). Cpd 27 was orally active and
rivatives were reported as promising HDAC inhibitors by Wang more potent HDAC8 inhibitor compared to the FDA-approved SAHA
et al. [103]. Compound containing 5-carbon linker function (FDA-01) [Figure-3]. Cpd 27 (ZYJ-34c) also exhibited higher in vivo
exhibited the best HDAC1 inhibition (IC50 ¼ 53 nM) but provided antitumor potency against human breast cancer MDA-MB-231-
poor HDAC8 inhibition (IC50 ¼ 608 nM). However, the compound xenograft model [109]. It also showed efficacy against mouse
with 6-carbon linker function produced better HDAC8 inhibitory hepatoma-H22 pulmonary metastasis as well as human colon

Fig. 10. Structures of HDAC8 inhibitors (18-25).


228 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Fig. 11. Structures of HDAC8 inhibitors (26-37).

cancer HCT116 xenograft model [109]. Tang et al. [110] employed an among these compounds [Fig. 11]. Fass et al. [115] attempted to
unique two step protocol to design selective HDAC8 inhibitors and develop some short-chain hydroxamic acids with potential HDAC
discovered a group of benzohydroxamic acid based compounds as inhibitory activity. Compound containing hydroxamate as the ZBG
potent and selective HDAC8 inhibitors. In an in vitro study of these having phenyl ring as the cap group showed the best HDAC8 in-
compounds against HDAC2, HDAC3 and HDAC8 enzymatic activity hibition (33, IC50 ¼ 4000 nM). However, it was more effective
the compounds actually provided promising results. The com- against HeLa HDAC (IC50 ¼ 600 nM) [Fig. 11]. Compounds contain-
pounds were remarkably selective along with excellent sub nano- ing the aryl cap group was found to be preferable than compounds
molar potency toward HDAC8 enzymatic activity that HDAC2 and with the linear alkyl chain as far as the HDAC8 inhibition was
HDAC3. The most active compound of series (28) [Fig. 11] was able concerned.
to exhibit a good activity (IC50 ¼ 23 nM) against HDA8 and was 15 Huang et al. [116] reported some osthole-based SAHA-linked
fold and 65 fold selective toward HDAC8 than HDAC2 and HDAC3 hydroxamates as effective and selective HDAC8 inhibitors
respectively. compared to HeLa HDACs. Cpd 34 among these compounds exerted
Botta et al. [111] reported effective HDAC inhibitors by modi- the most potent HDAC8 inhibitory activity (IC50 ¼ 267.85 nM)
fying the ZBG with oximes. As per their observations through [Fig. 11]. It is better than SAHA (FDA-01) [Fig. 3]. Spencer et al. [117]
computational analysis, it was noticed that the a-oxime as a ZBG synthesized Jay Amin hydroxamic acid (JAHA, 35, HDAC8
perfectly adapted in the active site for coordinating to the Zn2þ ion. IC50 ¼ 2 nM), an organometallic analog of SAHA with the ferrocenyl
Cpd 29 bearing the a-oxime function exhibited HDAC8 inhibition of group as the phenyl bioisostere. It had promising class I HDAC
IC50 ¼ 22,600 nM (Fig. 11). inhibitory property (Fig. 11). A number of potential chiral a-amino-
A number of g-lactum based compounds were derived and ketone-based HDAC8 inhibitors were reported by Whitehead and
evaluated by Choi and co-workers [112]. These compounds exerted co-workers [70]. One of these compounds comprising dichlor-
promising class I and class II HDAC inhibition but tended to be more ophenyl group showed potent HDAC8 inhibition (36, IC50 ¼ 90 nM)
effective against HDAC6. Cpd 30 (Fig. 11) showed the best HDAC6 with a minimum of 20-fold selectivity over HDAC1, HDAC2 and
inhibition (IC50 ¼ 0.8 nM) with an acceptable HDAC8 inhibition HDAC6 (Fig. 11). Guerrant et al. [118] explored some dual inhibitors
(IC50 ¼ 119.7 nM). Neelarapu et al. [113] synthesized a number of of HDACs and topoisomerase-II by combining two prototype
diazide-containing isoxazole- and pyrazole-derived compounds as chemotherapeutic agents (SAHA and anthracycline analogs). Few of
HDAC inhibitors. Compounds bearing pyrazole scaffold showed them showed medium to poor HDAC8 inhibition. Cpd 37 among
good HDAC3 and HDAC8 inhibition. Among these compounds, cpd these bifunctional compounds was the most potent HDAC8 inhib-
31 (Fig. 11) resulted in a good HDAC8 inhibition (IC50 ¼ 17 nM). itor (IC50 ¼ 220 nM) (Fig. 11). Henkes et al. [119] modified the linker
Zhang et al. [114] designed and synthesized a new class of tri- spacer of SAHA with the perfluorinated alkyl function. They ana-
peptidomimetics having good HDAC8 inhibitory profile. Compound lysed the effect of this spacer on different HDACs. Compound
possessing the 1, 4-dithia-7-azaspiro- [4,4]-nonane-8-carboxylic bearing the anthracene ring (38, HDAC8 Kd value ¼ 1,400 nM) were
acid as the cap group and 6-aminohexanoic acid as the linker more active than SAHA (HDAC8 Kd value ¼ 2,900 nM) against
function was the most potent HDAC8 inhibitor (32, IC50 ¼ 21 nM) HDAC8 (Fig. 12) [119].
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 229

Fig. 12. Structures of HDAC8 inhibitors (38-49).

Vaidya et al. [120] derived some tetrapeptide derivatives and reported some compounds having the amide-linked p-substituted
evaluated them against HDAC isoforms. This study revealed the phenyl hydroxamic acid and evaluated against HDAC2, HDAC4,
importance of hydroxamate ZBG group for imparting the HDAC HDAC6 and HDAC8 enzymes. The ethylene-piperazine function as
inhibitory activity. Cpd 39 showed potent in vitro HDAC8 inhibitory the cap group (HDAC8 IC50 ¼ 21,000 nM) provided poor activity
activity (IC50 ¼ 17 nM) (Fig. 12). Using the click chemistry-based Cu against HDAC8. Further, activity of these compounds was increased
(I)-catalyzed-azide-alkyne cycloaddition (CuAAC), a library of by reducing the chain length from the phenethyl group (HDAC8
HDAC8 inhibitors was synthesized by Suzuki et al. [121]. This study IC50 ¼ 1,150 nM) to the benzyl function (HDAC8 IC50 ¼ 1,030 nM)
resulted in selective and potent HDAC8 inhibitor 40 (IC50 ¼ 70 nM) and also from the benzyl to the phenyl function (46, HDAC8
[Fig. 12]. Marek et al. [122] reported synthesis and HDAC inhibitory IC50 ¼ 689 nM) [Fig. 12]. The activity profile against HDAC8 was
activity of some alkoxamide-linked hydroxamates. These com- decreased in case of the 4-methylamide analog (HDAC8
pounds possessed not much potency against HDAC8 but cpd 40 IC50 ¼ 1950 nM) compared to the 4-phenyl hydroxamic acid.
showed comparatively effective HDAC8 inhibition (41, Tashima et al. [128] designed and synthesized 15 carbostyril de-
IC50 ¼ 893 nM) [Fig. 12]. rivatives with promising HDAC inhibitory activity. The amide linker
Zhang et al. [123] synthesized some arylhydroxamic acids as (1-CONH) and the corresponding reverse amide (1-NHCO) between
potential HDAC8 inhibitors. Compounds containing the ortho and the cap group and the ZBG displayed some important hydrogen-
para-substituted phenyl rings were poor HDAC8 inhibitors. The 3- bonding interactions with HDAC enzyme. The best HDAC8 inhibi-
trifluorophenyl analog showed the most potent HDAC8 inhibition tory potency was shown by cpd 47 (IC50 ¼ 6 nM) [Fig. 12]. Ononye
(42, IC50 ¼ 967 nM) [Fig. 12]. During an attempt to develop selective et al. [129] synthesized some a- and b-tropolones as highly effec-
HDAC inhibitors for the treatment of Alzheimer's disease Yu et al. tive and selective HDAC2 inhibitors over HDAC1, HDAC4-6 and
[124] synthesized some quinazoline-4-one derivated compounds HDAC8. These compounds also showed good HDAC8 inhibition.
as HDAC inhibitors. The compounds were highly selective toward In an attempt to increase the efficacy of NCC-149 (40, HDAC8
HDAC6 but also appeared to possess promising HDAC8 inhibition. IC50 ¼ 70 nM) [Fig. 12], Suzuki et al. [130] modified the aryl linker of
The compound 43 delivered the highest in vitro HDAC8 inhibition the scaffold with different heterocyclic aryl groups such as triazole,
(HDAC8 IC50 ¼ 420 nM) and was almost equipotent to HDAC6 thiophene, etc. They found a compound with the modified triazole
enzymatic activity [Fig. 12]. By means of Suzuki cross-coupling orientation showed slight increase in the HDAC8 inhibitory potency
reactions, Patil et al. [125] reported some new HDAC inhibitors (48, IC50 ¼ 53 nM) [Fig. 12]. A series of biphenylacrylohydroxamic
comprising aromatic and heteroaromatic moiety with 3-hydroxy- acids as HDAC inhibitors (particularly HDAC6) was synthesized by
pyridin-2-thione (3-HPT) group as effective ZBG. Among these Cincinelli et al. [131]. The influence on HDAC8 inhibition was
compounds, the p-methyl biphenyl derivative exhibited the most observed by studying these compounds with the adamantyl or
potent HDAC8 inhibition (44, IC50 ¼ 800 nM) [Fig. 12]. without the adamantyl moiety. Interestingly, incorporation of an
Saha et al. [126] derived some new type of HDAC8 inhibitors adamantyl moiety was not effective for HDAC8 inhibition. Though
with pyrrole-imidazole polyamides (PIP) conjugates. The derivative the hydrophobic functionality was favourable in cap region, an
of JAHA complexed with PIP-d (JAHA-PIP-d, 45) showed excellent unexpected decrease in the HDAC8 and other HDAC isoforms was
HDAC8 inhibition (IC50 ¼ 130 nM) [Fig. 12]. Wagner et al. [127] observed for compounds containing adamantyl cap group.
230 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Tang et al. [132] brought a new class of aminotetralin-linked design newer HDAC6 inhibitors Lee et al. [137] synthesized HPB
dual inhibitors of HDAC6 and HDAC8. Among these compounds, (54) [Fig. 13], a benzohydroxamate derivative for selective HDAC6
cpd 49 showed most potent HDAC inhibitory activity (HDAC8 inhibition. The compound appeared to have higher HDAC6 inhibi-
IC50 ¼ 30 nM) (Fig. 12). Cai et al. [133] synthesized two series of 2- tory activity and was also effectively inhibited cellular growth in
aminobenzamides and hydroxamates which showed significant both tumor and normal cells without inducing apoptosis in normal
in vitro antiproliferative activity against NCI-H661, U937, A549, cells [137,138]. The compound was also provided a nanomolar ac-
MDAMB-231 and HCT-116 cancer cell lines. The hydroxamate de- tivity against HDAC8 (IC50 ¼ 376 nM). Further modification and
rivatives showed an excellent HDAC8 inhibition and selectivity over redesigning of newer molecules from this compound can help to
HDAC1 and HDAC2. These hydroxamates showed the importance of discover potent dual inhibitors of HDAC6 and HDAC8.
the phenyl ring as the cap group for HDAC8 inhibition (50, To design and synthesize some SAHA analogs, Zhang et al. [139]
IC50 ¼ 200 nM) [Fig. 13]. In a successful effort to enhance the po- reported some SAHA derivatives having comparatively better
tency of HDAC6 inhibitors, Shen and his co-workers [134] were able HDAC8 binding affinity (PDB: 1T69) than SAHA (FDA-01, Fig. 3). Cpd
to synthesize a series of bicyclic cap group containing benzohy- 55 (Fig. 13) with the trifluoromethyl group at the third position and
droxamate derivatives with higher affinity toward HDAC6 inhibi- the chlorine atom at the sixth position of the phenyl ring provided
tion. The most active compound (51) [Fig. 13] of this series was the highest HDAC inhibitory potency among these molecules.
exhibited an inhibitory activity of IC50 ¼ 3 nM against HDAC6 and Halder et al. [8] reported some isoglutamine derivatives as dual
with a decent nanomolar activity (IC50 ¼ 176 nm) against HDAC8 inhibitors of MMP-2 and HDAC8. Cpd 56 showed the most potent
enzymatic activity. Two folic acid and pteroic acid based hydrox- HDAC8 inhibiton (IC50 ¼ 2,890 nM) and exhibited anti-migratory
amate derivatives were designed and synthesized by Sodji et al. and anti-invasive properties tested in human lung carcinoma
[135] and were analysed against in vitro HDAC1, HDAC6 and HDAC8 A549 cell line (Fig. 13). In an attempt to design orally bioavailable
enzymatic activity. The observation of this experiment highlighted HDAC6 inhibitors, Lin et al. [140] designed and synthesized some
the folic acid derivative as totally inactive against all three HDAC aminopyrrolidinone derivatives as potent HDAC6 inhibitors. The
isoforms whereas the pteroic hydroxamate derivative (52) [Fig. 13] best active compounds provided an excellent HDAC6 inhibition
was excellently inhibited HDAC6 (IC50 ¼ 17.6 nM) along with along with an excellent HDAC8 inhibition (57, IC50 ¼ 80 nM)
moderate activity against HDAC8 (IC50 ¼ 581 nM) and poorly [Fig. 13]. It also strongly induced the acetylation of tubulin and
inhibited HDAC1 enzymatic activity (IC50 ¼ 2,390 nM). Further minimized the p21 levels. Some compounds containing 1-
modification of the linker motif of the compound was attempted hydroxypyridine-2-thione (1-HPT) group as ZBG was synthesized
but was unable to provide promising activity against the selected and evaluated by Muthyala and co-workers [141]. Molecular
HDAC isoforms. De Vreese et al. [136] synthesized some benzo- docking and molecular dynamic (MD) simulation study strongly
thiophene cap group containing hydroxamate derivatives as potent validated the binding pattern of these compounds with HDAC8.
HDAC inhibitors. The compounds clearly suggested that the 4- These compounds were found to form an octahedral structure
bromo substitution in the fused benzothiophene phenyl ring was while coordinating with the catalytic Zn2þ ion of HDAC8. Com-
unfavourable than its unsubstituted analog. The compound 53 pound 58 exhibited effective HDAC8 inhibitory activity
(Fig. 13) was 1,400 nM active against HDAC8 enzymatic activity but (IC50 ¼ 980 nM) [Fig. 13].
was more selective toward HDAC6 than HDAC8 isoform. In order to Pidugu et al. [25] synthesized some glycine or alanine linked

Fig. 13. Structures of HDAC8 inhibitors (50-61).


S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 231

oxadiazole derivatives having excellent class I HDAC inhibitory cpd 64 provided the best in vitro HDAC8 inhibition (IC50 ¼ 50 nM)
activity. Cpd 59 (Fig. 13) showed the maximum in vitro HDAC8 in- [Fig. 14]. In a recent study on HDAC inhibitors, Lamaa et al. [146]
hibition (IC50 ¼ 98 nM) among these compounds. It also showed designed and synthesized a number of iso-combretastatin based
effective in vitro antiproliferative activity against MDA-MB-231 compounds with inhibitory activity on different HDAC isoforms.
breast cancer cell line (IC50 ¼ 230 nM) better than SAHA The in vitro test results of these compounds against HDAC6, HDAC8
(IC50 ¼ 6,000 nM). Ingham et al. [142] discovered selective HDAC8 and HDAC11 enzymatic activity displayed these compounds more
inhibitors from a large set of small molecule library of Boston selective toward HDAC8 than HDAC6 and HDAC11 isoforms. Cpd 65
University Centre for Molecular Discovery. Initially, they screened a delivered the highest HDAC8 inhibitory activity (IC50 ¼ 60 nM) than
diverse set of 134 esters and subsequently converted these into a the other compounds of the series. Cpd 65 (Fig. 14) was also tested
total of 120 hydroxamic acids and methyl hydroxamates. This study on different tumor cell lines and was appeared to provide a
led to the development of a highly potent HDAC8 inhibitor namely promising antiproliferative activity against HCT 116 (GI50 ¼ 8 nM),
OJI-1 (60, HDAC8 IC50 ¼ 0.8 nM) [Fig. 13]. Cpd 60 was identified as A549 (GI50 ¼ 22 nM), K562 (GI50 ¼ 20 nM), PC3 (GI50 ¼ 28 nM), U87
the most potent and selective HDAC8 inhibitor till now over all (GI50 ¼ 8 nM), MCF-7 (GI50 ¼ 8 nM) and MiaPaca2 (GI50 ¼ 18 nM)
other HDACs (HDAC1 to HDAC7 and HDAC9). cell lines [146].
Taha et al. [143] reported some tetrahydroisoquinoline based
derivatives. These compounds showed the importance of the
7. QSAR studies on HDAC8 inhibitors
number of carbon atoms in the alkyl chain contributing effectively
towards HDAC8 inhibition. Hydroxamate derivative containing 4
Among the possible adaptable approaches for the identification
carbon atoms at the alkyl chain length showed the maximum
and discovery of molecules with optimistic biological potential,
in vitro HDAC8 inhibitory potency (61, IC50 ¼ 44 nM). However,
quantitative structure activity relationship (QSAR) study is one of
most of these compounds were selective towards HDAC1 and
the beneficial methods for identification and designing of potent
HDAC3 (Fig. 13). The N1-hydroxytetrephthalamide-based mole-
and selective inhibitors. For the past few years, different compu-
cules showed promising HDAC inhibition in which the indole cap
tational methodologies and statistically validated QSAR approaches
containing panobinostat (FDA-04, Fig. 3) and quisinostat (JNJ-
have been employed for screening and identification of molecules
26481585, 11) (Fig. 9) are in clinical trials [1]. Wang et al. [26]
with promising HDAC8 inhibitory activity and also for identification
combined both scaffolds and reported the indole cap containing
of structural features beneficial for HDAC8 inhibition. A number of
N1-hydroxytetrephthalamide derivatives. Cpd 62 (Fig. 14) showed
QSAR studies have been conducted for the identification and
promising HDAC inhibitory activity (IC50 ¼ 74 nM) though this
screening of potent HDAC8 inhibitors (Table 5).
molecule was comparatively HDAC8 inhibitor (IC50 ¼ 5,500 nM). A
In 2009, Ortore and co-workers [147] conducted a rational
group of propargylanime derivatives were synthesized by Wünsch
docking-based 3D-QSAR study on a diverse set of 48 hydroxamate
et al. [144] to inhibit Zn2þ containing HDAC isoforms. The com-
derivatives to procure potent HDAC1 and HDAC8 selective in-
pounds were seemed to provide a decent inhibitory activity against
hibitors. This set of molecules also included ligands like TSA, vor-
HDAC8 though it was more selective toward HDAC6 and HDAC1
inostat and MS-344. From the study, authors concluded that the
isoforms. The compound 63 (Fig. 14) was able to provide the highest
activity of these compounds may be dependable on a few things
HDAC8 inhibition (IC50 ¼ 417 nM) among the compounds of this
such as the hydrogen bond formation between the compound and
series.
the enzyme, polar interaction between the compound and D101
Zhao et al. [145] synthesized a group of potent HDAC8 selective
amino acid residue of HDAC8. The aromatic stack interaction
meta-sulfamoyl N-hydroxybenzamide derivatives among which
involving a phenylalanine residue of the enzyme might also cause

Fig. 14. Structures of HDAC8 inhibitors (62-65).


232 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Table 5
QSAR studies recently conducted on HDAC8 inhibitors.

Sl. No. Scaffold QSAR Technique Year Reference

01 Diverse Hyderoxamates Docking based 3D QSAR 2009 [147]


02 Diverse molecules Pharmacophore based 3D QSAR 2010 [148]
03 Diverse molecules 3D QSAR Pharmacophore mapping 2014 [149]
04 Diverse molecules Lazy learning-based classification QSAR 2015 [150]
05 Diverse compounds Pharmacophre-based 3D QSAR 2015 [8]
06 Diverse compounds Regression-based 2D QSAR, Pharmacophore-based 3D QSAR 2015 [151]
07 Diverse molecules SVC and SVM-based 2D QSAR 2016 [152]
08 SAHA derivative MLR based 2D-QSAR, Gaussian kernel based non-linear SVM 2D QSAR 2016 [153]
09 N-hydroxyfurylacrylamides DFT based MLR and ANN 2D QSAR 2016 [154]
10 Aryl valporic acids MLR 2D QSAR, Feed forward ANN 2D QSAR 2017 [155]
11 Diverse hydroxamates Pharmacophore based 3D QSAR 2017 [156]
12 Diverse molecules DFT based pharmacophore dependant 3D QSAR 2018 [157]
13 Diverse compounds Fragment based Bayesian classification QSAR 2018 [3]

HDAC1/HDAC8 selectivity [147]. The 3D-QSAR study also suggested


that these compounds with short indole linker and non-linker
molecules with more than one phenyl or aromatic ring might
provide HDAC8 selectivity.
Another study was conducted by Thangapandian et al. [148]
with the intention of designing novel HDAC8 inhibitors through a
pharmacophore-based 3D-QSAR analysis. They generated a couple
of pharmacophore hypothesis. These models were developed using
features like hydrogen bond acceptor (HBA), hydrogen bond donor
(HBD), metal binding group and hydrophobic features. The results
suggested six pharmacophore features as important components of
these compounds for better protein-ligand interaction between
HDAC8 enzyme and its inhibitors. The model suggested the HBA
and the HBD groups as the zinc binding features (ZBG) of these
compounds whereas the HY (hydrophobic) feature was suggested
to be a better fit as cap group for interaction with the surface res-
idues of HDAC8 pocket. The second pharmacophore model found to Fig. 15. Schematic representation of the pharmacophoric features of compound 66.
contain two hydrophobic features, one acting as the cap group and
the other one as the linker spacer between the cap and the ZBG
[148]. JFD03558 (69), KM00731 (70) and KM02921 (71) have been
Debanath et al. [149] applied a pharmacophore-based 3D-QSAR screened as the lead compounds for HDAC8 inhibition (Fig. 16)
study on a diverse set of 20 molecules with HDAC8 inhibitory ac- [150].
tivity. Authors reported that the best pharmacophore hypothesis Halder et al. [8] conducted some structure-based and ligand-
containing hydrogen bond acceptor (HBA), hydrogen bond donor based 3D-QSAR pharmacophore mapping study on a diverse set
(HBD) and ring aromatic (RA) features provided the regression of molecules in order to design dual inhibitors of HDAC8 and
coefficient (R2) of 0.982 and leave-one-out regression coefficient MMP2. Based on the best pharmacophore hypothesis constructed
(Q2LOO) of 73.14% with 15.34% standard deviation (SD) [149]. The with a hydrophobic, a HBA and two hydrogen bond acceptor lipid
pharmacophoric features of the selected models signify the pres- (HBAL) features, they designed and synthesized some glutamate
ence of such features in these compounds to provide a good derivatives possessed both HDAC8 and MMP-2 inhibitory potency.
protein-ligand interaction for HDAC8. It might be possible that the A regression based 2D-QSAR and pharmacophore-based 3D-
indolyl ethyl moiety acts as the cap group, the hydroxamate group QSAR studies using 16 inhibitors with class I HDAC inhibitory ac-
as ZBG along with the ring aromatic features of the cpd 66 (Fig. 15) tivity including HDAC8 inhibitory potency was carried out by Noor
as the linker moiety of the molecule to deliver the HDAC8 et al. [151]. They reported the best 2D regression based model for
inhibition. HDAC8 inhibition using 5 different descriptors that provided the R2
The constructed 3D-QSAR model yielded a R2 of 0.969 and a of 0.98 with a standard error (SE) of 0.24. The 2D-QSAR model
standard deviation of 0.144 along with a Q2 (R2CV) of 0.712. It also suggested that higher molecular polarizibility of these molecules
identified the presence of HBD group near the indole nitrogen can be detrimental toward their HDAC8 inhibitory activity. The
atom. The hydroxamate group and the linker nitrogen atoms are lipophilicity (logP) of these molecules seemed to be beneficial for
favourable for HDAC8 inhibition. The hydrophobic region predicted their activity. The polar surface area of the molecules (Topological
as favourable near the indolylethyl and phenyl linker moiety. The polar surface area/TPSA) was also positively correlated with their
electron withdrawing groups were also suggested to be favourable HDAC8 inhibitory activity whereas the number of hydrogen bond
for the compounds near the nitrogen atoms and the ZBG group. acceptor groups (HBA) was negatively correlated with the activity.
In order to screen potential molecules for HDAC8 inhibition, Cao Moreover, through the screening process 10 molecules (72-81)
et al. [150] performed a lazy-learning based classification QSAR [Fig. 17] were identified as lead molecules which might possess
study on a diverse set of 75 molecules including 38 inhibitors and potent HDAC inhibitory activity [151].
37 non-inhibitor compounds. These compounds were screened on Cao et al. [152] further implemented the QSAR technique in a
the basis of their inhibitory potency toward HDAC8. From this hybrid strategy for screening and identification of novel HDAC in-
study, five molecules such as AW00409 (67), HTS03338 (68), hibitors along with the molecular dynamics (MD) simulation study.
In this study, they applied a support vector component (SVC) and
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 233

Fig. 16. Structures of the compound 67-71 screened as lead compounds for HDAC8 inhibition.

Fig. 17. Structure of molecule 72-81 screened as lead compounds for HDAC8 inhibition.

support vector machine (SVM) based classification QSAR technique coefficient with 76.67% accuracy and predicted 76% of the accuracy
on a diverse set of 80 HDAC8 inhibitors. The SVC model provided a for the test set. Through the SVM model, 8 different physico-
93.33% accuracy rate with a 100% true positive rate (TPR) and chemical parameters of these molecules were identified which
13.33% of false positive rate (FPR) for the training set along with 84% might have influence on the HDAC8 inhibitory activity of these
accuracy on test set. The SVM model provided 74% regression compounds such as XlogP, TPSA, Span, Inertia Y, H_Don, N_Atoms and
234 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Inertia Z [152]. Using the screening process out of 59,652 molecules QSAR model. In this study, 40 hydroxamate derivatives were used
only two lead molecules were identified as promising hit com- as the training set compounds to develop the 2D-QSAR models
pounds for effective HDAC8 inhibition [152]. whereas 500 aryl valporic acid derivatives, designed by the authors
Praseetha et al. [153] performed a multiple linear regression were used as the test set compounds. These test set compounds
(MLR) based 2D QSAR and Gaussian karnel aided SVM based 2D were designed with a ZBG, a spacer group along with a lipophilic
QSAR models on 1, 2, 4-oxadiazole containing SAHA derivatives. tail in their structure. The docking technique using Autodock 4.2
The constructed MLR model with 5 descriptors resulted in an R2 of software was employed to perform docking of these molecules in
0.945, cross validated R2 (R2CV) of 0.905 with a mean error of 0.067. 3D-structure of HDAC8 (PDB ID-3F07) in order to screen these
The SVM aided model yielded an R2 of 0.990, R2CV of 0.923 with compounds [155]. A pool of 1194 descriptors including 1179 de-
mean error of 0.017 containing 5 different descriptors. The MLR and scriptors calculated by Dragon software, 10 quantum descriptors
SVM model both suggested the importance of 3D-MoRSE de- and 5 molecular descriptors procured from the blind docking was
scriptors (Mor32e and Mor09u) for these compounds for their used to carry out the QSAR study [155]. As a result of this study one
HDAC8 inhibitory activity [153]. The linear model signified the lead aryl valporic acid derivative DAVP042 (82) [Fig. 18(C)] was
importance of Mor32e by suggesting the increase of atomic elec- identified as a hit molecule and tested on in vitro human cancer cell
tronegativity may be beneficial for HDAC8 inhibition of these lines such as A549, MCF-7, HCT-116 and U937. The compound
compounds. Moreover, JGI7 which also indicating toward the (DAVP042) exhibited highest potency against HCT-116 cell line and
reduction of molecular topological charge (lower values of JGI7) was least potent in A549 cell line. Moreover, it provided moderate
may be possible to enhance HDAC8 inhibitory activity of the com- anti proliferative activity in MCF-7 and U937 cell lines [155].
pounds. The descriptor L3s and R5eþ also established negative A pharmacophore based 3D-QSAR model was constructed by
correlation with the HDAC8 inhibitory activity of the compounds Manal et al. [156] on a series of hydroxamic acid derivatives with
[153]. potent HDAC8 inhibitory activity. The best generated pharmaco-
Tassine and Elhallaoui [154] applied the density function theory phoric hypothesis resulted in the R2 of 0.850 with an external cross
(DFT)-based calculations on a dataset of 38 N-hydroxyfurylamide validated R2 of 0.890. The model was consisted with five pharma-
HDAC8 inhibitors. The MLR model provided a correlation coeffi- cophoric features such as a hydrogen bond donor (HBD), a ring
cient value of 0.800 with a standard deviation value of 0.288 aromatic (RA) and three hydrogen bond acceptor (HBA) features
whereas the artificial neural network-based 2D-QSAR model pro- and displayed the importance of these features toward the HDAC8
duced a correlation coefficient of 0.980. The MLR model was con- inhibitory activity of these compounds. The 3D-QSAR model
structed using a total of five independent variables. The equation generated contours of the best and the least active compounds of
indicated the negative correlation with the total energy of mole- that dataset suggested that the HBD feature near the hydroxamate
cules and positively correlated with the hardness. The melting nitrogen atom of the most active compound (83) is favouring its
point of the dataset compounds were also contributed negatively activity whereas the HBD feature of the least active compound (84)
toward the HDAC8 inhibitory activity of these molecules. The suggested the negative influence on its activity [Fig. 18(A)]. The
equation was also suggested that the lesser ionization potential and contours also suggested beneficial effects of bulky group substitu-
higher electron affinity of these molecules were preferable for the tion as the ring aromatic (RA) feature near the thiazole ring of the
higher HDAC8 inhibition [154]. most active compound. The presence of RA feature of the least
A molecular modelling study on aryl valporic acid HDAC8 in- active molecule near its methoxy phenyl moiety suggested it to be
hibitors was conducted by Martinez-pacheco and co-workers [155]. detrimental for its HDAC8 inhibition [156]. For the most active
They used the MLR 2D-QSAR analysis and feed forward artificial compound, the thiazole moiety was also identified to provide
neural network (ANN) technique to develop and optimize the 2D electron withdrawing effect, providing the positive influence

Fig. 18. (A) Schematic representation of the 3D-QSAR study on the most active (83) and the least active compound (84) of the dataset. (B) Structure of two hit molecules (85-86)
screened in the study. (C) Structure of valporic acid derivative DAVP042 (82) identified as lead molecule for HDAC8 inhibition.
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 235

virtual screening of the promising compounds from a diverse set of


molecules. In this study, they generated a number of pharmaco-
phore hypothesis using different combinations of pharmacophoric
features on a training set of 16 compounds with higher to less
HDAC8 inhibitory activity. Among these models, the one
comprising one HBD and two RA features provided the best sta-
tistical outcome which is given in Fig. 19 along with a most active
compound (87) of that training set.
They also used the DFT based approaches along with the phar-
macophoric hypothesis to virtually screen the potential HDAC8
inhibitors from the set of molecules. In this process of screening,
only 193 drug like molecules were screened from 1,10,000 com-
pounds using the 3D-QSAR pharmacophore model and ADMET
study [157]. Then a docking study with the HDAC8 3D structure was
conducted on the 193 screened compounds. As a result only 11 lead
molecules were identified as promising for HDAC inhibition. Thus, a
DFT based HOMO-LOMO study was future conducted on the lead
molecules and finally three hit molecules KM055296 (88),
HTS10917 (89) and SPB02579 (90) were identified as promising
Fig. 19. Structure and 3D-QSAR generated pharmacophore pheactures of compound lead molecules for potent HDAC8 inhibition (Fig. 20) [157].
87. In a very recent study, Amin et al. [3] conducted a fragment
based Bayesian classification QSAR method on 852 HDAC8 in-
hibitors to identify the good and bad molecular fragments
toward its HDAC8 inhibitory activity but the carboxyl group in its responsible for imparting the activity. In this study, a total of 1266
structure with its electron withdrawing effect seemed to be influ- molecules with HDAC8 inhibitory potency was obtained from the
encing negatively towards its activity. binding databases and subjected to initial screening to eliminate
On the basis of pharmacophore based 3D-QSAR model and duplicate molecules as well as by applying Lipinski's rule of five
docking results, two lead molecules DB08732 (85) and 15602 (86) [158] and Veber's rules [159]. Finally, the remaining 852 molecules
[Fig. 18(B)] were screened as hit molecules for HDAC8 inhibition were used to construct a QSAR model based on Naïve Bayes clas-
[156]. The ADMET, density function theory (DFT) and electrostatic sifier modelling and Laplacian-corrected Bayesian classifier
potential calculation study was also conducted on these molecules. modelling. 588 molecules were used as the training set and the rest
The ADMET study disclosed that these compounds (DB08732 and of the 264 molecules were used as the test set compounds to sta-
15602) both posses drug like characteristics whereas the DFT and tistically validate the model. The study was also able to provide an
electrostatic potential calculations exhibited the similarity of these insight into the HDAC8 inhibitors to identify important molecular
molecules with the FDA approved HDAC8 inhibitor vorinostat fragments obtained from ECFP_6 fingerprints and physico-chemical
(SAHA) on the basis of HOMO (Highest occupied molecular orbital) descriptors [3]. The outcome of the study revealed that either
and LOMO (Lowest occupied molecular orbital) energy, band en- hydroxamate or carboxamate group can act as the good ZBG than
ergy gaps and molecular electrostatic potentials of these molecules. the benzamide group for better HDAC8 inhibition. The study also
These finally upheld DB08732 (85) and 15602 (86) as lead proto- provided information about the surface recognition or cap group of
type molecules to design potent HDAC8 inhibitor [Fig. 18(B)] [156]. these molecules. It also indicated about a positive correlation be-
In order to identify novel HDAC8 inhibitors, Kim et al. [157] tween the hydrophobicity of the cap group with the HDAC8
employed the pharmacophore-based 3D-QSAR approach for the inhibitory activity of the molecules [3]. The higher hydrophobic

Fig. 20. Structures of the molecules (88-90) screened by Kim and his co-workers (2018).
236 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

Fig. 21. Schematic representation of the summary of the SAR and QSAR studies conducted on HDAC8 inhibitors.

groups such as biphenyl and phenyl imidazole derivated groups Hence, apart from these, theoretical and hypothetical in-
were suggested to be preferable as cap group than less bulky phenyl terventions along with lots of experimental studies have to be
derivatives for higher HDAC8 inhibition of these molecules. performed simultaneously to gather knowledge about the detail
enzyme-drug interaction mechanisms. Last but not the least, the
8. Future perspectives ligand-bound X-ray crystal structures of different HDAC8 inhibitors
also provides valuable insights about the binding mode of in-
Design and development of potential and selective HDAC8 in- teractions and the related chemical structures involved into these
hibitors is an important and challenging task. Based on the three mechanisms. These crucial ligand-binding interactions and the SAR
crucial structural and pharmacophoric requirements (cap function, as well as molecular modelling studies in combination with
linker feature and the zinc binding group), newer effective HDAC8 rigorous approaches, in a nutshell, should be effective to design
inhibitors may be designed. In a recent study, Amin and his co- potential target-specific HDAC8 inhibitors as lead candidates in
workers proposed a modified fish-like structural orchestration of future those may be used as effective HDAC8 inhibitors.
the pharmacophoric characters that probably enhance the HDAC8
inhibitory potential [1]. In some cases, spacially for the benzohy- Conflicts of interest
droxamate derivatives, the compounds were seemed to possess
promising affinity toward both HDAC6 and HDAC8 isoforms. So The authors have no conflict of interests.
designing of potent HDAC6/HDAC8 dual inhibitors is also plausible.
Apart from the hydroxamate ZBG, there is other non-hydroxamate Acknowledgments
ZBGs (such as thiol, carboxylic acid, boronic acid, trifluoromethyl
ketone, oxime, hydroxypyridine thione, tropolone, b-lactam and SB is grateful to the All India Council for Technical Education
other kind of heterocyclic ZBGs) those may also coordinate with the (AICTE) New Delhi for awarding a fellowship. NA is thankful to
zinc ion to execute HDAC8 inhibition [160]. Therefore, work is University Grants Commission (UGC), New Delhi, India for
continuing for designing potential HDAC8 inhibitors comprising providing Rajiv Gandhi National Fellowship (RGNF). SAA sincerely
such type of non-hydroxamate ZBGs [160]. Being a zinc-dependent acknowledge Jadavpur University, Kolkata for awarding Junior
metalloenzyme, targeting the ZBG of HDAC8 inhibitors must be Research Fellowship under State Government Fellowship Scheme
optimized as an initial strategy for designing higher active in- of Jadavpur University, Kolkata, India. TJ is thankful for the financial
hibitors that have already been proposed. In the current scenario, support from UPE Phase II Program of UGC, New Delhi to Jadavpur
rational drug designing (RDD) and molecular modelling strategies University, Kolkata, India. We also thank the support from
appeared to be effective tools for optimization as well as identifi- Department of Pharmaceutical Technology, Jadavpur University,
cation of lead compounds for HDAC8 inhibitors. These various Kolkata, India for providing the research facilities.
computational approaches can also be productive to design and/or
identify lead molecules for dual inhibition of HDAC isoforms. An References
overall summarized outcome of SAR and QSAR studies of all these
different HDAC8 inhibitors is also shown in Fig. 21. [1] S.A. Amin, N. Adhikari, T. Jha, Structureeactivity relationships of
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 237

hydroxamate-based histone deacetylase-8 inhibitors: reality behind anti- summary: vorinostat for treatment of advanced primary cutaneous t-cell
cancer drug discovery, Future Med. Chem. 9 (2017) 2211e2237. lymphoma, The Oncologist 12 (2007) 1247e1252.
[2] S.A. Amin, N. Adhikari, T. Jha, Is dual inhibition of metalloenzymes HDAC-8 [28] W.K. Kelly, O.A. O'Connor, M.L. Krug, J.H. Chiao, M. Heaney, T. Curley,
and MMP-2 a potential pharmacological target to combat hematological B. MacGregore-Cortelli, W. Tong, J.P. Secrist, L. Schwartz, S. Richardson,
malignancies? Pharmacol. Res. 122 (2017) 8e19. E. Chu, S. Olgac, P.A. Marks, H. Scher, V.M. Richon, Phase I study of an oral
[3] S.A. Amin, N. Adhikari, T. Jha, Diverse classes of HDAC8 inhibitors: in search histone deacetylase inhibitor, suberoylanilide hydroxamic acid, in patients
of molecular fingerprints that regulate activity, Future Med. Chem. 10 (2018) with advanced cancer, J. Clin. Oncol. 23 (2005) 3923e3931.
1589e1602. [29] W. Shi, Y.R. Lawrence, H. Choy, M. Werner-Wasik, D.W. Andrews, J.J. Evans,
[4] A. Chakrabarti, I. Oehme, O. Witt, G. Oliveira, W. Sippl, C. Romier, R.J. Pierce, K.D. Judy, C.J. Farrell, Y. Moshel, A.C. Berger, V. Bar-Ad, A.P. Dicker, Vorinostat
M. Jung, HDAC8: a multifaceted target for therapeutic interventions, Trends as a radiosensitizer for brain metastasis: a phase i clinical trial, J. Neuro
Pharmacol. Sci. 36 (2015) 481e492. Oncol. 118 (2014) 313e319.
[5] N. Adhikari, S.A. Amin, P. Trivedi, T. Jha, B. Ghosh, HDAC3 is a potential [30] R. Sarfstein, I. Bruchim, A. Fishman, H. Werner, The mechanism of action of
validated target for cancer: an overview on the benzamide-based selective the histone deacetylase inhibitor vorinostat involves interaction with the
HDAC3 inhibitors through comparative SAR/QSAR/QAAR approaches, Eur. J. insulin-like growth factor signalling pathway, PLoS One 6 (2011) e24468.
Med. Chem. 157 (2018) 1127e1142. [31] J.V. Bokhman, Two pathogenetic types of endometrial carcinoma, Gynecol.
[6] H. Nian, W.H. Bisson, W.-M. Dashwood, J.T. Pinto, R.H. Dashwood, Alpha-keto Oncol. 15 (1983) 10e17.
acid metabolites of organoselenium compounds inhibit histone deacetylase [32] S.F. Lax, Molecular genetic pathways in various types of endometrial carci-
activity in human colon cancer cells, Carcinogenesis 30 (2009) 1416e1423. noma: from a phenotypical to a molecular-based classification, Virchows
[7] A. Chakrabarti, J. Melesina, F.R. Kolbinger, I. Oehme, J. Senger, O. Witt, Arch. 444 (2004) 213e223.
W. Sippl, M. Jung, Targeting histone deacetylase 8 as a therapeutic approach [33] T. Ma, F. Galimberti, C.P. Erkmen, V. Memoli, F. Chinyengetere, L. Sempere,
to cancer and neurodegenerative diseases, Future Med. Chem. 8 (2016) J.H. Beumer, B.N. Anyang, W. Nugent, D. Johnstone, G.J. Tsongalis, J.M. Kurie,
1609e1634. H. Li, J. DiRenzo, Y. Guo, S.J. Freemantle, K.H. Dragnev, E. Dmitrovsky,
[8] A.K. Halder, S. Mallick, D. Shikha, A. Saha, K.D. Sahab, T. Jha, Design of dual Comparing histone deacetylase inhibitor responses in genetically engineered
MMP-2/HDAC8 inhibitors by pharmacophore mapping, molecular docking, mouse lung cancer models and a window of opportunity trial in patients
synthesis and biological activity, RSC Adv. 5 (2015) 72373e72386. with lung cancer, Mol. Canc. Therapeut. 12 (2013) 1545e1555.
[9] I.V. Gregoretti, Y.M. Lee, H.V. Goodson, Molecular evolution of the histone [34] M.G. Saelen, A.H. Ree, A. Kristian, K.G. Fleten, T. Furre, H.H. Hektoen,
deacetylase family: functional implications of phylogenetic analysis, J. Mol. K. Flatmark, Radiosensitization by the histone deacetylase inhibitor vorino-
Biol. 338 (2004) 17e31. stat under hypoxia and with capecitabine in experimental colorectal carci-
[10] M. Mottamal, S. Zheng, T.L. Huang, G. Wang, Histone deacetylase inhibitors noma, Radiat. Oncol. 7 (2012) 165.
in clinical studies as templates for new anticancer agents, Molecules 20 [35] Y. Oki, A. Younes, A. Copeland, F. Hagemeister, L.E. Fayad, P. McLaughlin,
(2015) 3898e3941. J. Shah, N. Fowler, J. Romaguera, Phase I study of vorinostat in combination
[11] B.J. Wilson, A.M. Tremblay, G. Deblois, G. Sylvain-Drolet, V. Gigue re, An with standard chop in patients with newly diagnosed peripheral T-cell
acetylation switch modulates the transcriptional activity of estrogen-related lymphoma, Br. J. Haematol. 162 (2013) 138e141.
receptor alpha, Mol. Endocrinol. 24 (2010) 1349e1358. [36] T. Doi, T. HamaguchiKuniaki, S. ChinKiyohiko, H. Noguchi, T. Otsuki,
[12] P. Bertrand, Inside HDAC with HDAC inhibitors, Eur. J. Med. Chem. 45 (2010) A. Mehta, A. Ohtsu, Evaluation of safety, pharmacokinetics, and efficacy of
2095e2116. vorinostat, a histone deacetylase inhibitor, in the treatment of gastrointes-
[13] W. Yan, S. Liu, E. Xu, J. Zhang, Y. Zhang, X. Chen, X. Chen, Histone deacetylase tinal (gi) cancer in a phase i clinical trial, Int. J. Clin. Oncol. 18 (2013) 87e95.
inhibitors suppress mutant p53 transcription via histone deacetylase 8, [37] A.A. Lane, B.A. Chabner, Histone deacetylase inhibitors in cancer therapy,
Oncogene 32 (2013) 599e609. J. Clin. Oncol. 27 (2009) 5459e5468.
[14] M. Haberland, M.H. Mokalled, R.L. Montgomery, E.N. Olson, Epigenetic con- [38] H. Ueda, H. Nakajima, Y. Hori, T. Fujita, M. Nishimura, T. Goto, M. Okuhara,
trol of skull morphogenesis by histone deacetylase 8, Genes Dev. 23 (2009) FR901228, a novel antitumor bicyclic depsipeptide produced by Chromo-
1625e1630. bacterium violaceum No. 968. I. Taxonomy, fermentation, isolation, physico-
[15] M. Nakagawa, Y. Oda, T. Eguchi, S. Aishima, T. Yao, F. Hosoi, Y. Basaki, M. Ono, chemical and biological properties, and antitumor activity, J. Antibiot. 47
M. Kuwano, M. Tanaka, M. Tsuneyoshi, Expression profile of class I histone (1994) 301e310.
deacetylases in human cancer tissues, Oncol. Rep. 18 (2007) 769e774. [39] V. Sandor, S. Bakke, R.W. Robey, M.H. Kang, M.V. Blagosklonny, J. Bender,
[16] M.A. Deardorff, M. Bando, R. Nakato, E. Watrin, T. Itoh, M. Minamino, R. Brooks, R.L. Piekarz, E. Tucker, W.D. Figg, K.K. Chan, B. Goldspiel, A.T. Fojo,
K. Saitoh, M. Komata, Y. Katou, D. Clark, K.E. Cole, E. De Baere, C. Decroos, S.P. Balcerzak, S.E. Bates, Phase I trial of the histone deacetylase inhibitor,
N. Di Donato, S. Ernst, L.J. Francey, Y. Gyftodimou, K. Hirashima, M. Hullings, depsipeptide (FR901228, NSC 630176), in patients with refractory neo-
Y. Ishikawa, C. Jaulin, M. Kaur, T. Kiyono, P.M. Lombardi, L. Magnaghi-Jaulin, plasms, Clin. Canc. Res. 8 (2002) 718e728.
G.R. Mortier, N. Nozaki, M.B. Petersen, H. Seimiya, V.M. Siu, Y. Suzuki, [40] B. Coiffier, B. Pro, H.M. Prince, F. Foss, L. Sokol, M. Greenwood, D. Caballero,
K. Takagaki, J.J. Wilde, P.J. Willems, C. Prigent, G. Gillessen-Kaesbach, P. Borchmann, F. Morschhauser, M. Wilhelm, L. Pinter-Brown,
D.W. Christianson, F.J. Kaiser, L.G. Jackson, T. Hirota, I.D. Krantz, K. Shirahige, S. Padmanabhan, A. Shustov, J. Nichols, S. Carroll, J. Balser, B. Balser,
HDAC8 mutations in Cornelia de Lange syndrome affect the cohesin acety- S. Horwitz, Results from a pivotal, open-label, phase II study of romidepsin in
lation cycle, Nature 489 (2012) 313e317. relapsed or refractory peripheral T-cell lymphoma after prior systemic
[17] L. Jackson, A.D. Kline, M.A. Barr, S. Koch, de Lange syndrome: a clinical review therapy, J. Clin. Oncol. 30 (2012) 631e636.
of 310 individuals, Am. J. Med. Genet. 47 (1993) 940e946. [41] V. Novotny-Diermayr, S. Hart, K.C. Goh, A. Cheong, L.-C. Ong, H. Hentze,
[18] J.R. Somoza, R.J. Skene, B.A. Katz, C. Mol, J.D. Ho, A.J. Jennings, C. Luong, M.K. Pasha, R. Jayaraman, K. Ethirajulu, J.M. Wood, The oral HDAC inhibitor
A. Arvai, J.J. Buggy, E. Chi, J. Tang, B.C. Sang, E. Verner, R. Wynands, pracinostat (SB939) is efficacious and synergistic with the JAK2 inhibitor
E.M. Leahy, D.R. Dougan, G. Snell, M. Navre, M.W. Knuth, R.V. Swanson, pacritinib (SB1518) in preclinical models of AML, Blood Canc. J. 2 (2012) e69.
D.E. McRee, L.W. Tari, Structural snapshots of human HDAC8 provide insights [42] V. Novotny-Diermayr, K. Sangthongpitag, C.Y. Hu, X. Wu, N. Sausgruber,
into the class I histone deacetylases, Structure 12 (2004) 1325e1334. P. Yeo, G. Greicius, S. Pettersson, A.L. Liang, Y.K. Loh, Z. Bonday, K.C. Goh,
[19] W.S. Xu, R.B. Parmigiani, P.A. Marks, Histone deacetylase inhibitors: molec- H. Hentze, S. Hart, H. Wang, K. Ethirajulu, J.M. Wood, SB939, a novel potent
ular mechanisms of action, Oncogene 26 (2007) 5541e5552. and orally active histone deacetylase inhibitor with high tumor exposure
[20] P. Trivedi, N. Adhikari, S.A. Amin, T. Jha, B. Ghosh, Design, synthesis and and efficacy in mouse models of colorectal cancer, Mol. Canc. Therapeut. 9
biological screening of 2-aminobenzamides as selective HDAC3 inhibitors (2010) 642e652.
with promising anticancer effects, Eur. J. Pharm. Sci. 124 (2018) 165e181. [43] S. Ren Qiao, W. Li, X. Wang, M. He, Y. Guo, L. Sun, Y. He, Y. Ge, Q. Yu, Chi-
[21] Z. Li, W.-G. Zhu, Targeting histone deacetylases for cancer therapy: from damide, a novel histone deacetylase inhibitor, synergistically enhances
molecular mechanisms to clinical implications, Int. J. Biol. Sci. 10 (2014) gemcitabine cytotoxicity in pancreatic cancer cells, Biochem. Biophys. Res.
757e770. Commun. 434 (2013) 95e101.
[22] C. Monneret, Histone deacetylase inhibitors, Eur. J. Med. Chem. 40 (2005) [44] J. Wu, C. Du, Z. Lv, C. Ding, J. Cheng, H. Xie, L. Zhou, S. Zheng, The up-
1e13. regulation of histone deacetylase 8 promotes proliferation and inhibits
[23] J. Roche, P. Bertrand, Inside HDACs with more selective HDAC inhibitors, Eur. apoptosis in hepatocellular carcinoma, Dig. Dis. Sci. 58 (2013) 3545e3553.
J. Med. Chem. 121 (2016) 451e483. [45] Z.A. Gurard-Levin, J. Kim, M. Mrksich, Combining mass spectrometry and
[24] R. Sangwan, R. Rajan, P.K. Mandal, HDAC as onco target: reviewing the peptide arrays to profile the specificities of histone deacetylases, Chem-
synthetic approaches with SAR study of their inhibitors, Eur. J. Med. Chem. biochem 10 (2009) 2159e2161.
158 (2018) 620e706. [46] M.W. Schilham, M.A. Oosterwegel, P. Moerer, J. Ya, P.A.J. de Boer, M. van de
[25] V.R. Pidugu, N.S. Yarla, S.R. Pedada, A.M. Kalle, A. KrishnaSatyab, Design and Wetering, S. Verbeek, W.H. Lamers, A.M. Kruisbeek, A. Cumano, H. Clevers,
synthesis of novel HDAC8 inhibitory 2, 5-disubstituted-1, 3, 4-oxadiazoles Defects in cardiac outflow tract formation and pro-B-lymphocyte expansion
containing glycine and alanine hybrids with anticancer activity, Bioorg. in mice lacking Sox-4, Nature 380 (1996) 711e714.
Med. Chem. 24 (2016) 5611e5617. [47] M.R. Potzner, K. Tsarovina, E. Binder, A. Penzo-Me ndez, V. Lefebvre,
[26] X. Wang, X. Li, J. Li, J. Hou, Y. Qu, C. Yu, F. He, W. Xu, J. Wu, Design, synthesis, H. Rohrer, M. Wegner, E. Sock, Sequential requirement of Sox4 and Sox11
and preliminary bioactivity evaluation of N1 -hydroxyterephthalamide de- during development of the sympathetic nervous system, Development 137
rivatives with indole cap as novel histone deacetylase inhibitors, Chem. Biol. (2010) 775e784.
Drug Des. 89 (2017) 38e46. [48] M. Bergsland, M. Werme, M. Malewicz, T. Perlmann, J. Muhr, The estab-
[27] B.S. Mann, J.R. Johnson, M.H. Cohen, R. Justice, R. Pazdur, FDA approval lishment of neuronal properties is controlled by Sox4 and Sox11, Genes Dev.
238 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

20 (2006) 3475e3486. [72] A. Vannini, C. Volpari, P. Gallinari, P. Jones, M. Mattu, A. Carfí, R. De Francesco,
[49] T. Higuchi, T. Nakayama, T. Arao, K. Nishio, O. Yoshie, SOX4 is a direct target C. Steinkühler, S. Di Marco, Substrate binding to histone deacetylases as
gene of FRA-2 and induces expression of HDAC8 in adult T-cell leukemia/ shown by the crystal structure of the HDAC8-substrate complex, EMBO Rep.
lymphoma, Blood 121 (2013) 3640e3649. 8 (2007) 879e884.
[50] S. Balasubramanian, J. Ramos, W. Luo, M. Sirisawad, E. Verner, J.J. Buggy, [73] N.J. Porter, N.H. Christianson, C. Decroos, D.W. Christianson, Structural and
A novel histone deacetylase 8 (HDAC8)-specific inhibitor PCI-34051 induces functional influence of the glycine-rich loop G302GGGY on the catalytic
apoptosis in T-cell lymphomas, Leukemia 22 (2008) 1026e1034. tyrosine of histone deacetylase 8, Biochemistry 55 (2016) 6718e6729.
[51] S.Y. Park, J.A. Jun, K.J. Jeong, H.J. Heo, J.S. Sohn, H.Y. Lee, C.G. Park, J. Kang, [74] G. Estiu, N. West, R. Mazitschek, E. Greenberg, J.E. Bradner, O. Wiest, On the
Histone deacetylases 1, 6 and 8 are critical for invasion in breast cancer, inhibition of histone deacetylase 8, Bioorg. Med. Chem. 18 (2010)
Oncol. Rep. 25 (2011) 1677e1681. 4103e4110.
[52] I. Oehme, H.E. Deubzer, D. Wegener, D. Pickert, J.-P. Linke, B. Hero, A. Kopp- [75] S.M. Gantt, C. Decroos, M.S. Lee, L.E. Gullett, C.M. Bowman, D.W. Christianson,
Schneider, F. Westermann, S.M. Ulrich, A. von Deimling, M. Fischer, O. Witt, C.A. Fierke, General baseegeneral acid catalysis in human histone deacety-
Histone deacetylase 8 in neuroblastoma tumorigenesis, Clin. Canc. Res. 15 lase 8, Biochemistry 55 (2016) 820e832.
(2009) 91e99. [76] J. Liu, I.D. Krantz, Cornelia de Lange syndrome, cohesin, and beyond, Clin.
[53] I. Oehme, H.E. Deubzer, M. Lodrini, T. Milde, O. Witt, Targeting of HDAC8 and Genet. 76 (2009) 303e314.
investigational inhibitors in neuroblastoma, Expert Opin. Investig. Drugs 18 [77] A.A. Tabackman, R. Frankson, E.S. Marsan, K. Perry, K.E. Cole, Structure of
(2009) 1605e1617. ‘linkerless’ hydroxamic acid inhibitor-HDAC8 complex confirms the forma-
[54] O. Witt, H.E. Deubzer, T. Milde, I. Oehme, HDAC family: what are the cancer tion of an isoform-specific subpocket, J. Struct. Biol. 195 (2016) 373e378.
relevant targets? Cancer Lett. 277 (2009) 8e21. [78] J.J. Buggy, M.L. Sideris, P. Mak, D.D. Lorimer, B. McIntosh, J.M. Clark, Cloning
[55] E. Koeneke Rettig, F. Trippel, W.C. Mueller, J. Burhenne, A. Kopp-Schneider, and characterization of a novel human histone deacetylase, HDAC8, Bio-
J. Fabian, A. Schober, U. Fernekorn, A.V. Deimling, H.E. Deubzer, T. Milde, chem. J. 350 (2000) 199e205.
O. Witt, I. Oehme, Selective inhibition of HDAC8 decreases neuroblastoma [79] I. Van den Wyngaert, W. de Vries, A. Kremer, J. Neefs, P. Verhasselt,
growth in vitro and in vivo and enhances retinoic acid-mediated differen- W.H. Luyten, S.U. Kass, Cloning and characterization of human histone
tiation, Cell Death Dis. 6 (2015) e1657. deacetylase 8, FEBS Lett. 478 (2000) 77e83.
[56] M. Lodrini, I. Oehme, C. Schroeder, T. Milde, M.C. Schier, A. Kopp-Schneider, [80] M. Marek, T.B. Shaik, T. Heimburg, A. Chakrabarti, J. Lancelot, E. Ramos-
J.H. Schulte, M. Fischer, K.D. Preter, F. Pattyn, M. Castoldi, M.U. Muckenthaler, Morales, C. Da Veiga, D. Kalinin, J. Melesina, D. Robaa, K. Schmidtkunz,
A.E. Kulozik, F. Westermann, O. Witt, H.E. Deubzer, MYCN and HDAC2 T. Suzuki, R. Holl, E. Ennifar, R.J. Pierce, M. Jung, W. Sippl, C. Romier, Char-
cooperate to repress miR-183 signaling in neuroblastoma, Nucleic Acids Res. acterization of histone deacetylase 8 (HDAC8) selective inhibition reveals
41 (2013) 6018e6033. specific active site structural and functional determinants, J. Med. Chem. 61
[57] I. Oehme, J.P. Linke, B.C. Bo € ck, T. Milde, M. Lodrini, B. Hartenstein, I. Wiegand, (2018) 10000e10016.
C. Eckert, W. Roth, M. Kool, S. Kaden, H.J. Gro €ne, J.H. Schulte, S. Lindner, [81] P.M. Lombardi, K.E. Cole, D.P. Dowling, D.W. Christianson, Structure, mech-
A. Hamacher-Brady, N.R. Brady, H.E. Deubzer, O. Witt, Histone deacetylase 10 anism, and inhibition of histone deacetylases and related metalloenzymes,
promotes autophagy-mediated cell survival, Proc. Natl. Acad. Sci. U.S.A. 110 Curr. Opin. Struct. Biol. 21 (2011) 735e743.
(2013) E2592eE2601. [82] Research Collaboratory for Structural Bioinformatics (RCSB) Protein Data
[58] O. Witt, H.E. Deubzer, M. Lodrini, T. Milde, I. Oehme, Targeting histone Bank (PDB), 2018. www.rcsb.org. (Accessed 14 November 2018).
deacetylases in neuroblastoma, Curr. Pharmaceut. Des. 15 (2009) 436e447. [83] R. Furumai, A. Matsuyama, N. Kobashi, K.H. Lee, M. Nishiyama, H. Nakajima,
[59] K.L. Durst, B. Lutterbach, T. Kummalue, A.D. Friedman, S.W. Hiebert, The inv A. Tanaka, Y. Komatsu, N. Nishino, M. Yoshida, S. Horinouchi, FK228 (dep-
(16) fusion protein associates with corepressors via a smooth muscle myosin sipeptide) as a natural prodrug that inhibits class I histone deacetylases,
heavy-chain domain, Mol. Cell Biol. 23 (2003) 607e619. Cancer Res. 62 (2002) 4916e4921.
[60] B. Lutterbach, S.W. Hiebert, Role of the transcription factor AML-1 in acute [84] B. Jose, Y. Oniki, T. Kato, N. Nishino, Y. Sumida, M. Yoshida, Novel histone
leukemia and hematopoietic differentiation, Gene 245 (2000) 223e235. deacetylase inhibitors: cyclic tetrapeptide with trifluoromethyl and penta-
[61] M. Marek, S. Kannan, A.-T. Hauser, M.M. Mour~ ao, S. Caby, V. Cura, D.A. Stolfa, fluoroethyl ketones, Bioorg. Med. Chem. Lett 14 (2004) 5343e5346.
K. Schmidtkunz, J. Lancelot, L. Andrade, J.-P. Renaud, G. Oliveira, W. Sippl, [85] S. Oh, J.-C. Jungb, M.A. Avery, Synthesis of new b-lactam analogs and eval-
M. Jung, J. Cavarelli, R.J. Pierce, C. Romier, Structural basis for the inhibition of uation of their histone deacetylase (HDAC) activity, Z. Naturforsch. 62b
histone deacetylase 8 (HDAC8), a key epigenetic player in the blood fluke (2007) 1459e1464.
Schistosoma mansoni, PLoS Pathog. 9 (2013) e1003645. [86] A.P. Kozikowski, Y. Chen, A. Gaysin, B. Chen, M.A. D'Annibale, C.M. Suto,
[62] C.Q. Chen, K. Yu, Q.X. Yan, C.Y. Xing, Y. Chen, Z. Yan, Y.F. Shi, K.W. Zhao, B.C. Langley, Functional differences in epigenetic modulators superiority of
S.M. Gao, Pure curcumin increases the expression of SOCS1 and SOCS3 in mercaptoacetamide-based histone deacetylase inhibitors relative to
myeloproliferative neoplasms through suppressing class I histone deacety- hydroxamates in cortical neuron neuroprotection Studies, J. Med. Chem. 50
lases, Carcinogenesis 34 (2013) 1442e1449. (2007) 3054e3061.
[63] S.M. Gao, C.Q. Chen, L.Y. Wang, L.L. Hong, J.B. Wu, P.H. Dong, F.J. Yu, Histone [87] O.M. Moradei, T.C. Mallais, S. Frechette, I. Paquin, P.E. Tessier, S.M. Leit,
deacetylases inhibitor sodium butyrate inhibits JAK2/STAT signaling through M. Fournel, C. Bonfils, M.-C. Trachy-Bourget, J. Liu, T.P. Yan, A.-H. Lu, J. Rahil,
upregulation of SOCS1 and SOCS3 mediated by HDAC8 inhibition in J. Wang, S. Lefebvre, Z. Li, A.F. Vaisburg, J.M. Besterman, Novel aminophenyl
myeloproliferative neoplasms, Exp. Hematol. 41 (2013) 261e270. benzamide-type histone deacetylase inhibitors with enhanced potency and
[64] Y. Yamauchi, H. Boukari, I. Banerjee, I.F. Sbalzarini, P. Horvath, A. Helenius, selectivity, J. Med. Chem. 50 (2007) 5543e5546.
Histone deacetylase 8 is required for centrosome cohesion and influenza A [88] J. Arts, P. Angibaud, A. Marie €n, W. Floren, B. Janssens, P. King, J. van Dun,
virus entry, PLoS Pathog. 7 (2011) e1002316. L. Janssen, T. Geerts, R.W. Tuman, D.L. Johnson, L. Andries, M. Jung, M. Janicot,
[65] C. Decroos, C.M. Bowman, J.A. Moser, K.E. Christianson, M.A. Deardorff, K. van Emelen, R306465 is a novel potent inhibitor of class I histone
D.W. Christianson, Compromised structure and function of HDAC8 mutants deacetylases with broad-spectrum antitumoral activity against solid and
identified in Cornelia de Lange Syndrome spectrum disorders, ACS Chem. haematological malignancies, BJC (Br. J. Cancer) 97 (2007) 1344e1353.
Biol. 9 (2014) 2157e2164. [89] K. Krennhrubec, B.L. Marshall, M. Hedglin, E. Verdin, S.M. Ulrich, Design and
[66] J. Gao, B. Siddoway, Q. Huang, H. Xia, Inactivation of CREB mediated gene evaluation of ‘Linkerless’ hydroxamic acids as selective HDAC8 inhibitors,
transcription by HDAC8 bound protein phosphatase, Biochem. Biophys. Res. Bioorg. Med. Chem. Lett 17 (2007) 2874e2878.
Commun. 379 (2009) 1e5. [90] Y. Chen, M. Lopez-Sanchez, D.N. Savoy, D.D. Billadeau, G.S. Dow,
[67] B. Gryseels, K. Polman, J. Clerinx, L. Kestens, Human schistosomiasis, Lancet A.P. Kozikowski, A series of potent and selective, triazolylphenyl-based
368 (2006) 1106e1118. histone deacetylases inhibitors with activity against pancreatic cancer cells
[68] D. Van der Kleij, A. Van Remoortere, J.H.N. Schuitemaker, M.L. Kapsenberg, and Plasmodium falciparum, J. Med. Chem. 51 (2008) 3437e3448.
A.M. Deelder, A.G.M. Tielens, C.H. Hokke, Maria Yazdanbakhsh, Triggering of [91] A.P. Kozikowski, S. Tapadar, D.N. Luchini, K.H. Kim, D.D. Billadeau, Use of the
innate immune responses by schistosome egg glycolipids and their carbo- nitrile oxide cycloaddition (NOC) reaction for molecular probe generation: a
hydrate epitope GalNAc beta 1-4(Fuc alpha 1-2Fuc alpha 1-3) GlcNAc, new class of enzyme selective histone deacetylase inhibitors (HDACIs)
J. Infect. Dis. 185 (2002) 531e539. showing picomolar activity at HDAC6, J. Med. Chem. 51 (2008) 4370e4373.
[69] A. Vannini, C. Volpari, G. Filocamo, E.C. Casavola, M. Brunetti, D. Renzoni, [92] V. Andrianov, V. Gailite, D. Lola, E. Loza, V. Semenikhina, I. Kalvinsh, P. Finn,
P. Chakravarty, C. Paolini, R. De Francesco, P. Gallinari, C. Steinkühler, S. Di K.D. Petersen, J.W. Ritchie, N. Khan, A. Tumber, L.S. Collins, S.M. Vadlamudi,
Marco, Crystal structure of a eukaryotic zinc-dependent histone deacetylase, F. Bjo€ rkling, M. Sehested, Novel amide derivatives as inhibitors of histone
human HDAC8, complexed with a hydroxamic acid inhibitor, Proc. Natl. deacetylase: design, synthesis and SAR, Eur. J. Med. Chem. 44 (2009)
Acad. Sci. U.S.A. 101 (2004) 15064e15069. 1067e1085.
[70] L. Whitehead, M.R. Dobler, B. Radetich, Y. Zhu, P.W. Atadja, T. Claiborne, [93] J. Arts, P. King, A. Marie€n, W. Floren, A. Belie€n, L. Janssen, I. Pilatte, B. Roux,
J.E. Grob, A. McRiner, M.R. Pancost, A. Patnaik, W. Shao, M. Shultz, L. Decrane, R. Gilissen, I. Hickson, V. Vreys, E. Cox, K. Bol, W. Talloen, I. Goris,
R. Tichkule, R.A. Tommasi, B. Vash, P. Wang, T. Stams, Human HDAC isoform L. Andries, M.D. Jardin, M. Janicot, M. Page, K. van Emelen, P. Angibaud, JNJ-
selectivity achieved via exploitation of the acetate release channel with 26481585, a novel ‘second-generation’ oral histone deacetylase inhibitor,
structurally unique small molecule inhibitors, Bioorg. Med. Chem. 19 (2011) shows broad-spectrum preclinical antitumoral activity, Cancer Ther. Clin.
4626e4634. Can. Res. 15 (2009) 6841e6851.
[71] D.P. Dowling, S.L. Gantt, S.G. Gattis, C.A. Fierke, D.W. Christianson, Structural [94] D.V. Smil, S. Manku, Y.A. Chantigny, S. Leit, A. Wahhab, T.P. Yan, M. Fournel,
studies of human histone deacetylase 8 and its site-specific variants com- C. Maroun, Z. Li, A.M. Lemieux, A. Nicolescu, J. Rahil, S. Lefebvre, A. Panetta,
plexed with substrate and inhibitors, Biochemistry 47 (2008) 13554e13563. J.M. Besterman, R. De ziel, Novel HDAC6 isoform selective chiral small
S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240 239

molecule histone deacetylase inhibitors, Bioorg. Med. Chem. Lett 19 (2009) neuronal chromatin, ACS Med. Chem. Lett. 2 (2011) 39e42.
688e692. [116] W.J. Huang, C.C. Chen, S.W. Chao, C.C. Yu, C.Y. Yang, J.H. Guh, Y.C. Lin, C.I. Kuo,
[95] G. Giannini, M. Marzi, R. Pezzi, T. Brunetti, G. Battistuzzi, M.D. Marzo, P. Yang, C.I. Chang, Synthesis and evaluation of aliphatic-chain hydrox-
W. Cabri, L. Vesci, C. Pisano, N-Hydroxy-(4-oxime)-cinnamide: a versatile amates capped with osthole derivatives as histone deacetylase inhibitors,
scaffold for the synthesis of novel histone deacetylase (HDAC) inhibitors, Eur. J. Med. Chem. 46 (2011) 4042e4049.
Bioorg. Med. Chem. Lett 19 (2009) 2346e2349. [117] J. Spencer, J. Amin, M. Wang, G. Packham, S.S.S. Alwi, G.J. Tizzard, S.J. Coles,
[96] J.C. Canzoneri, P.C. Chen, A.K. Oyelere, Design and synthesis of novel histone R.M. Paranal, J.E. Bradner, T.D. Heightman, Synthesis and biological evalua-
deacetylase inhibitor derived from nuclear localization signal peptide, Bio- tion of JAHAs: ferrocene-based histone deacetylase inhibitors, ACS Med.
org. Med. Chem. Lett 19 (2009) 6588e6590. Chem. Lett. 2 (2011) 358e362.
[97] A.K. Oyelere, P.C. Chen, W. Guerrant, S.C. Mwakwari, R. Hood, Y. Zhang, [118] W. Guerrant, V. Patil, J.C. Canzoneri, A.K. Oyelere, Dual targeting of histone
Y. Fan, Non-peptide macrocyclic histone deacetylase inhibitors, J. Med. deacetylase and topoisomerase II with novel bifunctional inhibitors, J. Med.
Chem. 52 (2009) 456e468. Chem. 55 (2012) 1465e1477.
[98] A. Montero, J.M. Beierle, C.A. Olsen, M.R. Ghadiri, Design, synthesis, biological [119] L.M. Henkes, P. Haus, F. J€ ager, J. Ludwig, F.J. Meyer-Almes, Synthesis and
evaluation, and structural characterization of potent Histone Deacetylase biochemical analysis of 2,2,3,3,4,4,5,5,6,6,7,7-dodecafluoro-N-hydroxy-
inhibitors based on cyclic alpha/beta-tetrapeptide architectures, J. Am. Chem. octanediamides as inhibitors of human histone deacetylases, Bioorg. Med.
Soc. 131 (2009) 3033e3041. Chem. 20 (2012) 985e995.
[99] B. He, S. Velaparthi, G. Pieffet, C. Pennington, A. Mahesh, D.L. Holzle, [120] A.S. Vaidya, R. Neelarapu, A. Madriaga, H. Bai, E. Mendonca, H. Abdelkarim,
M. Brunsteiner, R. van Breemen, Y. Sylvie, P. Blond, A. Petukhov, Binding R.B. van Breemen, S.Y. Blond, P.A. Petukhov, Novel histone deacetylase 8
ensemble profiling with photoaffinity labeling (BEProFL) approach: mapping ligands without a zinc chelating group: exploring an ‘upside-down’ binding
the binding poses of HDAC8 inhibitors, J. Med. Chem. 52 (2009) 7003e7013. pose, Bioorg. Med. Chem. Lett 22 (2012) 6621e6627.
[100] C.A. Olsen, M.R. Ghadiri, Discovery of potent and selective histone deacety- [121] T. Suzuki, Y. Ota, M. Ri, M. Bando, A. Gotoh, Y. Itoh, H. Tsumoto, P.R. Tatum,
lase inhibitors via focused combinatorial libraries of cyclic alpha3 b-tetra- T. Mizukami, H. Nakagawa, S. Iida, R. Ueda, K. Shirahige, N. Miyata, Rapid
peptides, J. Med. Chem. 52 (2009) 7836e7846. discovery of highly potent and selective inhibitors of histone deacetylase 8
[101] N. Suzuki, T. Suzuki, Y. Ota, T. Nakano, M. Kurihara, H. Okuda, T. Yamori, using click chemistry to generate candidate libraries, J. Med. Chem. 55 (2012)
H. Tsumoto, H. Nakagawa, N. Miyata, Design, synthesis, and biological ac- 9562e9575.
tivity of boronic acid-based histone deacetylase inhibitors, J. Med. Chem. 52 [122] L. Marek, A. Hamacher, F.K. Hansen, K. Kuna, H. Gohlke, M.U. Kassack,
(2009) 2909e2922. T. Kurz, Histone deacetylase (HDAC) inhibitors with a novel connecting unit
[102] P. Angibaud, K. Van Emelen, L. Decrane, S. van Brandt, P. ten Holte, I. Pilatte, linker region reveal a selectivity profile for HDAC4 and HDAC5 with
B. Roux, V. Poncelet, D. Speybrouck, L. Queguiner, S. Gaurrand, A. Marie €n, improved activity against chemo resistant cancer cells, J. Med. Chem. 56
W. Floren, L. Janssen, M. Verdonck, J. van Dun, J. van Gompel, R. Gilissen, (2013) 427e436.
C. Mackie, M.D. Jardin, J. Peeters, M. Noppe, L. Van Hijfte, E. Freyne, M. Page, [123] L. Zhang, X. Wang, X. Li, L. Zhang, W. Xu, Discovery of a series of hydroxamic
M. Janicot, J. Arts, Identification of a series of substituted 2-piperazinyl-5- acid derivatives as potent histone deacetylase inhibitors, J. Enzym. Inhib.
pyrimidylhydroxamic acids as potent histone deacetylase inhibitors, Bio- Med. Chem. 29 (2014) 582e589.
org. Med. Chem. Lett 20 (2010) 294e298. [124] C.-W. Yu, P.-T. Chang, L.-W. Hsin, J.-W. Chern, Quinazolin-4-one derivatives
[103] H. Wang, Z.-Y. Lim, Y. Zhou, M. Ng, T. Lu, K. Lee, K. Sangthongpitag, K.C. Goh, as selective histone deacetylase-6 inhibitors for the treatment of alzheimer's
X. Wang, X. Wu, H.H. Khng, S.K. Goh, W.C. Ong, Z. Bonday, E.T. Sun, Acylurea disease, J. Med. Chem. 56 (2013) 6775e6791.
connected straight chain hydroxamates as novel histone deacetylase in- [125] V. Patil, Q.H. Sodji, J.R. Kornacki, M. Mrksich, A.K. Oyelere, 3-hydroxypyridin-
hibitors: synthesis, SAR, and in vivo antitumor activity, Bioorg. Med. Chem. 2-thione as novel zinc binding group for selective histone deacetylase in-
Lett 20 (2010) 3314e3321. hibition, J. Med. Chem. 56 (2013) 3492e3506.
[104] R. He, Y. Chen, Y. Chen, A.V. Ougolkov, J.-S. Zhang, D.N. Savoy, D.D. Billadeau, [126] A. Saha, G.N. Pandian, S. Sato, J. Taniguchi, K. Hashiya, T. Bando, H. Sugiyama,
A.P. Kozikowski, Synthesis and biological evaluation of triazol-4-ylphenyl- Synthesis and biological evaluation of a targeted DNA-binding transcrip-
bearing histone deacetylase inhibitors as anticancer agents, J. Med. Chem. 53 tional activator with HDAC8 inhibitory activity, Bioorg. Med. Chem. 21
(2010) 1347e1356. (2013) 4201e4209.
[105] S.C. Mwakwari, W. Guerrant, V. Patil, S.I. Khan, B.L. Tekwani, Z.A. Gurard- [127] F.F. Wagner, D.E. Olson, J.P. Gale, T. Kaya, M. Weïwer, N. Aidoud, M. Thomas,
Levin, M. Mrksich, A.K. Oyelere, Non-peptide macrocyclic histone deacety- E.L. Davoine, B.C. Lemercier, Y.-L. Zhang, E.B. Holson, Potent and selective
lase inhibitors derived from tricyclic ketolide skeleton, J. Med. Chem. 53 inhibition of histone deacetylase 6 (HDAC6) does not require a surface-
(2010) 6100e6111. binding motif, J. Med. Chem. 56 (2013) 1772e1776.
[106] S. Terracciano, S. Di Micco, G. Bifulco, P. Gallinari, R. Riccio, I. Bruno, Synthesis [128] T. Tashima, H. Murata, H. Kodama, Design and synthesis of novel and highly-
and biological activity of cyclotetrapeptide analogues of the natural HDAC active pan-histone deacetylase (pan-HDAC) inhibitors, Bioorg. Med. Chem.
inhibitor FR235222, Bioorg. Med. Chem. 18 (2010) 3252e3260. 22 (2014) 3720e3731.
[107] Y. Zhang, J. Feng, C. Liu, L. Zhang, J. Jiao, H. Fang, L. Su, X. Zhang, J. Zhang, [129] S.N. Ononye, M.D. VanHeyst, E.Z. Oblak, W. Zhou, M. Ammar, A.C. Anderson,
M. Li, B. Wang, W. Xu, Design, synthesis and preliminary activity assay of 1, D.L. Wright, Tropolones as lead-like natural products: the development of
2, 3, 4-tetrahydroisoquinoline-3-carboxylic acid derivatives as novel Histone potent and selective histone deacetylase inhibitors, ACS Med. Chem. Lett. 4
deacetylases (HDACs) inhibitors, Bioorg. Med. Chem. 18 (2010) 1761e1772. (2013) 757e761.
[108] Y. Zhang, J. Feng, Y. Jia, X. Wang, L. Zhang, C. Liu, H. Fang, W. Xu, Develop- [130] T. Suzuki, N. Muto, M. Bando, Y. Itoh, A. Masaki, M. Ri, Y. Ota, H. Nakagawa,
ment of tetrahydroisoquinoline-based hydroxamic acid derivatives: potent S. Iida, K. Shirahige, N. Miyata, Design, synthesis, and biological activity of
histone deacetylase inhibitors with marked in vitro and in vivo antitumor NCC149 derivatives as histone deacetylase 8-selective inhibitors, Chem-
activities, J. Med. Chem. 54 (2011) 2823e2838. MedChem 9 (2014) 657e664.
[109] Y. Zhang, H. Fang, J. Feng, Y. Jia, X. Wang, W. Xu, Discovery of a [131] R. Cincinelli, L. Musso, G. Giannini, V. Zuco, M.D. Cesare, F. Zunino,
tetrahydroisoquinoline-based hydroxamic acid derivative (ZYJ-34c) as his- S. Dallavalle, Influence of the adamantyl moiety on the activity of bipheny-
tone deacetylase inhibitor with potent oral antitumor activities, J. Med. lacrylohydroxamic acid-based HDAC inhibitors, Eur. J. Med. Chem. 79 (2014)
Chem. 54 (2011) 5532e5539. 251e259.
[110] W. Tang, T. Luo, E.F. Greenberg, J.E. Bradner, S.L. Schreiber, Discovery of [132] G. Tang, J.C. Wong, W. Zhang, Z. Wang, N. Zhang, Z. Peng, Z. Zhang, Y. Rong,
histone deacetylase 8 selective inhibitors, Bioorg. Med. Chem. Lett 21 (2011) S. Li, M. Zhang, L. Yu, T. Feng, X. Zhang, X. Wu, J.Z. Wu, L. Chen, Identification
2601e2605. of a novel aminotetralin class of HDAC6 and HDAC8 selective inhibitors,
[111] C.B. Botta, W. Cabri, E. Cini, L. De Cesare, C. Fattorusso, G. Giannini, J. Med. Chem. 57 (2014) 8026e8034.
M. Persico, A. Petrella, F. Rondinelli, M. Rodriquez, A. Russo, M. Taddei, [133] J. Cai, H. Wei, K.H. Hong, X. Wu, M. Cao, X. Zong, L. Li, C. Sun, J. Chen, M. Ji,
Oxime amides as a novel zinc binding group in histone deacetylase in- Discovery and preliminary evaluation of 2-aminobenzamide and hydrox-
hibitors: synthesis, biological activity and computational evaluation, J. Med. amate derivatives containing 1, 2, 4-oxadiazole moiety as potent histone
Chem. 54 (2011) 2165e2182. deacetylase inhibitors, Eur. J. Med. Chem. 96 (2015) 1e13.
[112] L.E. Choi, M. Cho, B. Lee, S.J. Oh, S.K. Park, K. Lee, H.M. Kim, G. Han, Structure [134] S. Shen, V. Benoy, J.A. Bergman, J.H. Kalin, M. Frojuello, G. Vistoli, W. Haeck,
and property based design, synthesis and biological evaluation of gamma- L. Van Den Bosch, A.P. Kozikowski, Bicyclic-capped histone deacetylase 6
lactam based HDAC inhibitors, Bioorg. Med. Chem. Lett 21 (2011) inhibitors with improved activity in a model of axonal charcotmarietooth
1218e1221. disease, ACS Chem. Neurosci. 7 (2016) 240e258.
[113] R. Neelarapu, D.L. Holzle, S. Velaparthi, H. Bai, M. Brunsteiner, S.Y. Blond, [135] Q.H. Sodji, J.R. Kornacki, J.F. McDonald, M. Mrksich, A.K. Oyelere, Design and
P.A. Petukhov, Design, synthesis, docking, and biological evaluation of novel structure activity relationship of tumor-homing histone deacetylase in-
diazide-containing isoxazole and pyrazole-based histone deacetylase probes, hibitors conjugated to folic and pteroic acids, Eur. J. Med. Chem. 96 (2015)
J. Med. Chem. 54 (2011) 4350e4364. 340e359.
[114] Y. Zhang, J. Feng, Y. Jia, Y. Xu, C. Liu, H. Fang, W. Xu, Design, synthesis and [136] R. De Vreese, N. Van Steen, T. Verhaeghe, T. Desmet, N. Bougarne, K. De
primary activity assay of tripeptidomimetics as histone deacetylase in- Bosscher, V. Benoy, W. Haeck, L. Van Den Bosch, M. D’hooghe, Synthesis of
hibitors with linear linker and branched cap group, Eur. J. Med. Chem. 46 benzothiophene-based hydroxamic acids as potent and selective HDAC6
(2011) 5387e5397. inhibitors, Chem. Commun. 51 (2015) 9868e9871.
[115] D.M. Fass, R. Shah, B. Ghosh, K. Hennig, S. Norton, W.-N. Zhao, S.A. Reis, [137] J.-H. Lee, Y. Yao, A. Mahendran, L. Ngo, G. Venta-Perez, M.L. Choy, R. Breslow,
P.S. Klein, R. Mazitschek, R.L. Maglathlin, T.A. Lewis, S.J. Haggarty, Short- P.A. Marks, Creation of a histone deacetylase 6 inhibitor and its biological
chain HDAC inhibitors differentially affect vertebrate development and effects, Proc. Natl. Acad. Sci. Unit. States Am. 112 (2015) 12005e12010.
240 S. Banerjee et al. / European Journal of Medicinal Chemistry 164 (2019) 214e240

[138] R. De Vreese, M. D'hooghe, Synthesis and applications of benzohydroxamic pharmacophore based virtual screening, 3D QSAR and molecular docking
acid-based histone deacetylase inhibitors, Eur. J. Med. Chem. 135 (2017) approach, Am. J. PharmTech Res. 4 (2014) 253e267.
174e195. [150] G.P. Cao, M. Arooj, S. Thangapandian, C. Park, V. Arulalapperumal, Y. Kim,
[139] S. Zhang, W. Huang, X. Li, Z. Yang, B. Feng, Synthesis, biological evaluation, Y.J. Kwon, H.H. Kim, J.K. Suh, K.W. Lee, A lazy learning-based QSAR classifi-
and computer-aided drug designing of new derivatives of hyperactive sub- cation study for screening potential histone deacetylase 8 (HDAC8) in-
eroylanilide hydroxamic acid histone deacetylase inhibitors, Chem. Biol. hibitors, SAR QSAR Environ. Res. 26 (2015) 397e420.
Drug Des. 86 (2015) 795e804. [151] Z. Noor, N. Afzal, S. Rashid, Exploration of novel inhibitors for class I histone
[140] X. Lin, W. Chen, Z. Qiu, L. Guo, W. Zhu, W. Li, Z. Wang, W. Zhang, Z. Zhang, deacetylase isoforms by QSAR modeling and molecular dynamics simulation
Y. Rong, M. Zhang, L. Yu, S. Zhong, R. Zhao, X. Wu, J.C. Wong, G. Tang, Design assays, PLoS One 10 (2015) e0139588.
and synthesis of orally bioavailable aminopyrrolidinone histone deacetylase [152] G.P. Cao, S. Thangapandian, M. Son, R. Kumar, Y.-J. Choi, Y. Kim, Y.J. Kwon, H.-
6 inhibitors, J. Med. Chem. 58 (2015) 2809e2820. H. Kim, J.-K. Suh, K.W. Lee, QSAR modeling to design selective histone
[141] R. Muthyala, W.S. Shin, J. Xie, Y.Y. Sham, Discovery of 1-hydroxypyridine-2- deacetylase 8 (HDAC8) inhibitors, Arch Pharm. Res. (Seoul) 39 (2016)
thiones as selective histone deacetylase inhibitors and their potential 1356e1369.
application for treating leukemia, Bioorg. Med. Chem. Lett 25 (2015) [153] S. Praseetha, S. Bandaru, M. Yadav, A. Nayarisseri, S.S. kumar, Common SAR
4320e4324. derived from multiple QSAR models on vorinostat derivatives targeting
[142] O.J. Ingham, R.M. Paranal, W.B. Smith, R.A. Escobar, H. Yueh, T. Snyder, HDACs in tumor treatment, Curr. Pharmaceut. Des. 22 (2016) 5072e5078.
J.A. Porco Jr., J.E. Bradner, A.B. Beeler, Development of a potent and selective [154] F.Z. Tassine, M. Elhallaoui, DFT-based QSAR study of N-hydroxyfur-
HDAC8 inhibitor, ACS Med. Chem. Lett. 7 (2016) 929e932. ylacrylamide and its derivatives, principal inhibitors of histone deacetylase,
[143] T.Y. Taha, S.M. Aboukhatwa, R.C. Knopp, N. Ikegaki, H. Abdelkarim, J. Neerasa, J. Mater. Environ. Sci. 7 (2016) 2252e2258.
Y. Lu, R. Neelarapu, T.W. Hanigan, G.R.J. Thatcher, P.A. Petukhov, Design, [155] H. Martínez-Pacheco, G. Ramírez-Galicia, M. Vergara-Arias, J. Gertsch,
synthesis, and biological evaluation of tetrahydroisoquinoline-based histone M.J. Fragoso-Va zquez, D. Me ndez-Luna, A.L. Abujamra, C.-P.L. Cristina, R.-
deacetylase 8 selective inhibitors, ACS Med. Chem. Lett. 8 (2017) 824e829. H.M. Cecilia, I. Mendoza-Lujambio, J. Correa-Basurto, Docking and QSAR
[144] M. Wünsch, J. Senger, P. Schultheisz, S. Schwarzbich, K. Schmidtkunz, studies of aryl-valproic acid derivatives to identify anti-proliferative agents
C. Michalek, M. Klaß, S. Goskowitz, P. Borchert, L. Praetorius, W. Sippl, targeting the HDAC8, Anti Cancer Agents Med. Chem. 17 (7) (2017)
M. Jung, N. Sewald, Structureeactivity relationship of propargylamine-based 927e940.
HDAC inhibitors, ChemMedChem 12 (2017) 2044e2053. [156] M. Manal, K. Manish, D. Sanal, A. Selvaraj, V. Devadasan, M.J.N. Chandrasekar,
[145] C. Zhao, J. Zang, Q. Ding, E.S. Inks, W. Xu, C.J. Chou, Y. Zhang, Discovery of Novel HDAC8 inhibitors: a multi-computational approach, SAR QSAR Envi-
meta-sulfamoyl N-hydroxybenzamides as HDAC8 selective inhibitors, Eur. J. ron. Res. 28 (2017) 707e733.
Med. Chem. 150 (2018) 282e291. [157] S. Kim, Y. Lee, S. Kim, S.J. Lee, P.K. Heo, S. Kim, Y.J. Kwon, K.W. Lee, Identi-
[146] D. Lamaa, H.-P. Lin, L. Zig, C. Bauvais, G. Bollot, J. Bignon, H. Levaique, fication of novel human HDAC8 inhibitors by pharmacophore-based virtual
O. Pamlard, J. Dubois, M. Ouaissi, M. Souce, A. Kasselouri, F. Saller, D. Borgel, screening and density functional theory approaches, Bull. Korean Chem. Soc.
C. Jayat-Vignoles, H. Al-Mouhammad, J. Feuillard, K. Benihoud, M. Alami, 39 (2018) 197e206.
A. Hamze, Design and synthesis of tubulin and histone deacetylase inhibitor [158] C.A. Lipinski, F. Lombardo, B.W. Dominy, P.J. Feeney, Experimental and
based on iso-combretastatin A-4, J. Med. Chem. 61 (2018) 6574e6591. computational approaches to estimate solubility and permeability in drug
[147] G. Ortore, F.D. Colo, A. Martinelli, Docking of hydroxamic acids into HDAC1 discovery and development settings, Adv. Drug Deliv. Rev. 46 (2001) 3e26.
and HDAC8: a rationalization of activity trends and selectivities, J. Chem. Inf. [159] D.F. Veber, S.R. Johnson, H.Y. Cheng, B.R. Smith, K.W. Ward, K.D. Kopple,
Model. 49 (2009) 2774e2785. Molecular properties that influence the oral bioavailability of drug candi-
[148] S. Thangapandian, S. John, S. Sakkiah, K.W. Lee, Ligand and structure based dates, J. Med. Chem. 45 (2002) 2615e2623.
pharmacophore modeling to facilitate novel histone deacetylase 8 inhibitor [160] S.A. Amin, N. Adhikari, T. Jha, Structure-activity relationships of HDAC8 in-
design, Eur. J. Med. Chem. 45 (2010) 4409e4417. hibitors: non-hydroxamates as anticancer agents, Pharmacol. Res. 131
[149] S. Debnath, P. Nath, R.K. Nath, Identification of novel HDAC8 inhibitors using (2018) 128e142.

You might also like