ArticlePDF Available

Design and Synthesis of a New Series of 3,5-Disubstituted-1,2,4-Oxadiazoles as Potential Colchicine Binding Site Inhibitors: Antiproliferative activity, Molecular docking, and SAR Studies

Authors:

Abstract and Figures

The development of anticancer compounds targeting the colchicine-binding site of tubulin, termed colchicine-binding site inhibitors (CBSIs) is a promising research area for pharmaceutical companies and research institutes. A series of 3,5-disubstituted 1,2,4-oxadiazoles, sharing common structural features with colchicine and combretastatins, was designed and synthesized for screening as antiproliferative CBSIs. All targets were submitted to National Cancer Institute (NCI), USA for screening at 10 mM in full NCI 60 cell panel. In addition, molecular docking studies were performed for the newly synthesized oxadiazole derivatives as CBSIs in β-tubulin against the β-tubulin pocket of colchicine. Also, 5a -as the most active member- was evaluated for its ability to inhibit β-tubulin polymerization against colchicine 6 as a reference standard. The targets showed significant antiproliferative activities. Among the series, target 5a demonstrated the broadest and most potent antiproliferative activity (positive cytotoxic effect (PCE) = 33/60 and the mean growth inhibition (MGI) for the most sensitive cell lines (33) is 24.9 %. In addition, targets (5a-k) showed selective potency towards renal cancer in a particular A498 cell line. The order of potency of the three most potent targets and the standard colchicine (retrieved from NCI database) against A498 is: colchicine (GI = 84.5%) > 5a (GI = 78%) > 5d (GI = 51%) > 5f (32%). Compound 5a achieved superior binding affinity (-8.06 kcal/mol) to that of colchicine 6 itself which achieved (-7.40 kcal/mol). Also, 5a showed a two-fold potent inhibitory activity of tubulin polymerization (IC50: 1.18uM) compared to that of colchicine (IC50: 2.37uM). Thus, target 5a represents a promising selective anticancer CBSI.
Content may be subject to copyright.
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
Cite this: DOI: 10.1039/d1nj02885e
Design and synthesis of a new series of
3,5-disubstituted-1,2,4-oxadiazoles as potential
colchicine binding site inhibitors: antiproliferative
activity, molecular docking, and SAR studies
Rana T. Diab,
a
Zakaria K. Abdel-Sami,
a
Eatedal H. Abdel-Aal,
a
Ahmed A. Al-Karmalawy *
b
and Nader E. Abo-Dya *
ac
The development of anticancer compounds targeting the colchicine-binding site of tubulin, termed
colchicine-binding site inhibitors (CBSIs) is a promising research area for pharmaceutical companies and
research institutes. A series of 3,5-disubstituted 1,2,4-oxadiazoles, sharing common structural features
with colchicine and combretastatins, was designed and synthesized for screening as antiproliferative
CBSIs. All targets were submitted to National Cancer Institute (NCI), USA for screening at 10 mM in full
NCI 60 cell panel. In addition, molecular docking studies were performed for the newly synthesized
oxadiazole derivatives as CBSIs in b-tubulin against the b-tubulin pocket of colchicine. Also, 5a – as the
most active member – was evaluated for its ability to inhibit b-tubulin polymerization against colchicine
6as a reference standard. The targets showed significant antiproliferative activities. Amongst the series,
target 5a demonstrated the broadest and most potent antiproliferative activity (positive cytotoxic effect
(PCE) = 33/60 and the mean growth inhibition (MGI) for the most sensitive cell lines (33) is 24.9%. In
addition, targets (5a–k) showed selective potency towards renal cancer in particular the A498 cell line.
The order of potency of the three most potent targets and the standard colchicine (retrieved from
NCI database) against A498 is: colchicine (GI = 84.5%) 45a (GI = 78%) 45d (GI = 51%) 45f (32%).
Compound 5a achieved superior binding affinity (8.06 kcal mol
1
) compared with that of colchicine 6
itself, which achieved (7.40 kcal mol
1
). Also, 5a showed a two-fold potent inhibitory activity of tubulin
polymerization (IC
50
: 1.18 mM) compared with that of colchicine (IC
50
: 2.37 mM). Thus, target 5a
represents a promising selective anticancer CBSI.
1. Introduction
Cancer is one of the top 10 leading causes of death globally.
It was responsible for about 10 million deaths in 2020.
1
Despite
the presence of multiple treatment strategies for cancer, che-
motherapy remains one of the most significant treatments in
cancer management.
2–4
Unfortunately, most current anticancer
drugs have side effects due to the lack of tumor specificity,
multidrug resistance, toxicity, and reduced bioavailability.
5–8
Thus, chemists are actively searching for new drugs with higher
efficiency and lower side effects.
Microtubule-binding agents (MTAs) inhibit cell proliferation
by halting microtubule assembly/disassembly dynamics during
themitoticphaseofthecellcycle.
9
MTAs are classified into
two major groups, the microtubule-stabilizing agents, and
microtubule-destabilizing agents (Fig. 1). The later ‘‘destabilizing’’
agents prevent tubulin polymerization and promote depolymeri-
zation when present at high concentrations.
10
Most of these
agents bind to either the ‘‘vinca’’ domain near the exchangeable
GTP-binding site (E-site) or the ‘‘colchicine’’ domain at the
intradimer interface between a-andb-tubulin. Examples of vinca
site binders include vinblastine, vincristine, dolastatins, eribulin,
and tasidotin. Colchicine-site binders include colchicine and
its analogs, combretastatins, 2-methoxyestradiol, phenylahistins
(diketopiperazine), steganacins, and curacins.
10
Most of ‘‘micro-
tubule-stabilizing’’ agents bind to the taxoid binding site on
b-tubulin, which is located on the inside surface of the micro-
tubule and enhances its polymerization at high drug concen-
trations, e.g., taxol (paclitaxel, Taxolt), docetaxel (Taxoteret),
epothilones, ixabepilone (Ixemprat), and patupilone.
11
a
Department of Pharmaceutical Organic Chemistry, Faculty of Pharmacy,
Zagazig University, Zagazig 44519, Egypt. E-mail: nader_elmaghry88@yahoo.com
b
Department of Pharmaceutical Medicinal Chemistry, Faculty of Pharmacy, Horus
University-Egypt, New Damietta 34518, Egypt. E-mail: akarmalawy@horus.edu.eg
c
Department of Pharmaceutical Chemistry, Faculty of Pharmacy,
University of Tabuk, Tabuk 71491, Saudi Arabia
Electronic supplementary information (ESI) available. See DOI: 10.1039/
d1nj02885e
Received 12th June 2021,
Accepted 27th October 2021
DOI: 10.1039/d1nj02885e
rsc.li/njc
NJC
PAPER
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
View Journal
New J. Chem. This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021
Tubulin inhibitors that bind to the colchicine binding site
have therapeutic merit over other MTAs, where they inhibit
angiogenesis and usually overcome multidrug resistance
(MDR). Colchicine (Fig. 2) is a very cheap alkaloid isolated
from natural sources. Colchicine has a well-defined binding
site to microtubule (CBS) and effectively arrests mitosis. Unfor-
tunately, its narrow therapeutic index (as it binds to tubulin in
non-cancerous cells) and multiple mechanisms of action led
to systemic toxicities and multiorgan dysfunction in a clinical
trial and thus was not approved by the FDA for cancer
treatment.
12,13
Encouragingly, colchicine served as a template
for the development of advantageous MTAs targeting the CBS,
which are often: (i) effective in P-glycoprotein, MRP1, and
MRP2 overexpression-mediated multidrug resistance, (ii) effec-
tive as a vasculature disrupting agents, and (iii) cost-effective
compared to biologicals, such as monoclonal antibodies.
13–15
Accordingly, the development of novel analogs of colchicine
binding site inhibitors (CBSIs) with lower side effects is a goal
of many former and current research endeavors.
16
Combretastatin A-4 (CA-4) is structurally similar to colchi-
cine and has been extensively studied to improve its activity
and pharmacokinetics properties. Molecules that fall into the
combretastatin family generally share three common structure
features: a trimethoxy A ring, a B-ring containing substituents
often at C30and C40, and an ethene bridge between the two
rings which provides necessary structural rigidity (Fig. 2).
Structure–activity relationship studies found that the cis-
configuration of the double bond and the 3,4,5-trimethoxy
group on ring A is essential for the anti-tubulin activity of
CA-4.
17
However, this is not always the case with its analogs.
The trans-configuration of a-methyl chalcone Iinhibited
cancer cell proliferation in K562 cells at an IC
50
=0.21nM.
18,19
The clinical application of CA-4 as an anticancer CBSI was
hampered by its poor water solubility, short half-life, isomeriza-
tion of the ethene bond, as well as its toxic side effects.
20
To overcome these limitations, many CA-4 analogs have been
developed that retained the biological effects of the parent
molecule with enhanced water solubility and pharmacokinetic
properties (Fig. 2).
The common structural features in several reported potent
colchicine and CA-4 analogs (Fig. 2):
(1) Two aryl groups separated by a 2-3 carbons spacer e.g.:
(a) cis-double bond,
(b) a,b-unsaturated ketone,
(c) a,g-disubstituted five-membered heterocycle such as
imidazole II,
21
thiazole III,
22
and 1,3,4-oxadiazole IV.
23
Fig. 1 Tubulin binding sites.
Fig. 2 Colchicine, combretastatin A-4, examples of reported CBSIs (I,II,andIII), and the novel targets 5a–k.
Paper NJC
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
(2) One of the two aryl groups is always a trimethoxy
phenyl ring.
Targets 5a–k are structurally similar to compounds I,II, and
III, where they contain a a,g-disubstituted five-membered
heterocycle (1,2,4-oxadiazole) and one substituent is 3,4,5-
trimethoxyphenyl group and the other one is aryl or aralkyl
groups (Fig. 2).
1.1. Rational of molecular design
CA-4 binds to the colchicine binding site in tubulin, inhibits
tubulin aggregation in microtubules, alters microtubule dynamics
and this has been confirmed by biochemical experiments and
X-ray crystallography.
24
Docking studies and molecular
dynamics simulations examined the binding models of struc-
turally diverse CBSIs and concluded seven common pharma-
cophoric points: three hydrogen bond acceptors (A1, A2, and
A3), one hydrogen bond donor (D1), two hydrophobic centers
(H1 and H2), and one planar group (R1).
25,26
The seven points
of the pharmacophoric model are partitioned into two planes.
Plane A consists of points A2, A3, and H2, and plane B consists
of points (A1, D1, H1, and R1), Fig. 3.
26,27
In the light of these studies, the molecules that will have
these seven pharmacophoric features will be considered as
promising tubulin inhibitors. Consequently, we could predict
the essential pharmacophoric features for our newly designed
compounds to synthesize promising anticancer agents acting
as CBSIs compared to colchicine as a reference standard
(Fig. 4).
25,28
2. Results and discussion
2.1. Chemistry
3,4,5-Trimethoxybenzamidoxime 2was synthesized by heating
under reflux 3,4,5-trimethoxybenzonitrile 1,withhydroxylamine
HCl in the presence of equivalent sodium carbonate (Scheme 1).
29
N-Acylbenzotriazoles 4a–k were synthesized from commercially
available carboxylic acids 3a–k using two different procedures
Fig. 3 Seven pharmacophoric features: three hydrogen bond acceptors
(A1, A2 and A3), one hydrogen bond donor (D1), two hydrophobic centers
(H1 and H2), and one planar group (R1).
Fig. 4 Schematic representation showing the applied pharmacophoric features for the design of the newly synthesized compounds as CBSIs in respect
to colchicine as a standard reference.
NJC Paper
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
New J. Chem. This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021
depending on the chemical nature of the functional groups
present in the acid. First, N-acylbenzotriazoles 4a and 4e–k were
synthesized from commercially available carboxylic acids 3a and
3e–k, respectively, using the thionyl chloride procedure.
30
On the
other hand, N-acyl benzotriazoles containing free amino group
4b–d were synthesized using the tosyl chloride route.
31
Intermediate
N-acylbenzotriazoles 4a–k were crystallized from CH
2
Cl
2
/hexanes
and their melting points were compared to reported values and
were taken to the next step without further identification.
31–36
Targets 5a–k were characterized using
1
HNMR,
13
CNMR, as
well as IR spectroscopy. The purity of targets was confirmed by
elemental analyses.
2.1.1. Synthesis of 1,2,4-oxadiazoles 5a–k. 1,2,4-Oxadi-
azoles 5a–k were synthesized by the reaction of N-acylbenzo-
triazole 4a–k with 3,4,5-trimethoxybenzamidoxime 2in the
presence of triethylamine under reflux conditions (Scheme 1).
37
Targets 5a–k were characterized using
1
HNMR,
13
CNMR, as well
as IR spectroscopy, and their purity was confirmed by elemental
analyses. The mechanism of the reaction involved initial depro-
tonation of the hydroxyl group of amidoxime by adding triethyl-
amine to give the nitroxide anion of amidoxime. The anion was
acylated by the electron-deficient carbonyl group of N-acylbenzo-
triazoles to form the N-O-acyl amidoxime derivative 7.Intra-
molecular cyclodehydration of intermediate N-O-acyl amidoxime
Scheme 1 Reagents and conditions; (a) NH
2
OHHCl, Na
2
CO
3
, ethanol, reflux; (b) 1-H-benzotriazole, SOCl
2
,CH
2
Cl
2
, stirring for 2–3 h at r.t.;
(c) (1) DMAP, TEA, CH
2
Cl
2
, 0.5 h at 25 1C, (2) 1-H-benzotriazole, 1.5 h; (d) TEA, ethanol, reflux.
Paper NJC
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
derivatives produced the corresponding 1,2,4-oxadiazoles
(Scheme 2).
38
2.2. Biological evaluation
2.2.1. Anti-proliferative activities towards 60 subpanel
tumor cell lines. All the target compounds 5a–k were submitted
to the Developmental Therapeutics Program hosted by the
National Cancer Institute, USA. Their antiproliferative activities
were tested against 60 diverse human cancer cell lines (the
NCI-60) by single-dose assay (concentration 10
5
M). The com-
plete NCI-60 results are represented in the ESI.The percentages
of growth inhibition were significant and the most potent
compounds were 5a–e and 5k (Table 1).
2.2.2. Tubulin polymerization inhibitory activity. Tubulin
polymerization inhibitory activity was also evaluated for the
most active compound (5a) using ELISA assay for unpolymerized
Tubulin Beta (TUBb) to confirm its proposed mechanism of action.
The polymerization was monitored through the fluorescence
intensity change after their incubation. Compound (5a)showed
superior inhibitory activity (IC
50
=1.18mM) against the colchicine
binding site in b-tubulin compared to colchicine as a reference
standard (IC
50
=2.37mM) as depicted in Table 2.
2.3. Molecular docking studies
At first, we studied the colchicine binding site in b-tubulin, the
co-crystallized native inhibitor (N-[(7S)-1,2,3,10-tetramethoxy-9-
oxo-5,7-dihydro-5H-benzo[d]heptalen-7-yl]ethanamide) was found
to be stabilized inside its binding pocket of the CBS receptor
through the formation of one hydrogen bond with Lys350
amino acid.
Molecular docking of the tested oxadiazole derivatives 5a–k,
colchicine 6, and the co-crystallized native inhibitor (N-[(7S)-
1,2,3,10-tetramethoxy-9-oxo-5,7-dihydro-5H-benzo[d]heptalen-7-yl]
ethanamide 7inside the active site of the CBS in b-tubulin was
performed.
All of the newly tested synthesized compounds 5a–k were got
stabilized inside the CBS in b-tubulin by variable scores and
binding interactions with the amino acids of the receptor
pocket. They showed promising binding scores (from 6.41
to 8.06 kcal mol
1
), compared to both colchicine (6,7.40),
and the docked co-crystallized inhibitor (7,8.13). Also, the
RMSD_refine values of the selected poses were within the
acceptable range (from 0.93 to 2.26).
The docked co-crystallized inhibitor (7) fitted inside the CBS
in the b-tubulin pocket through a hydrogen bond formation
with Lys350 amino acid at 2.72 Å. Also, on the other hand,
colchicine (6) – as the first known tubulin destabilizing agent –
got stabilized inside the colchicine binding site by a pi–H bond
formation with Lys350 amino acid at 3.67 Å (Table 3). Accord-
ingly, this indicates the very crucial role of Lys350 amino acid to
stabilize ligands inside the pocket of the CBS in b-tubulin.
It is worth mentioning that all of the newly synthesized
derivatives 5a–k achieved nearly a fingerprint binding mode to
both colchicine (6) and the co-crystallized inhibitor in binding
to the same essential amino acid (Lys350) required for the
inhibitory activity of the CBS in b-tubulin, indicating an
expected affinity and subsequent intrinsic activity to the newly
synthesized derivatives as promising CBSIs.
Surprisingly, the target compound (5a) of the newly synthesized
derivatives achieved superior binding affinity (8.06 kcal mol
1
)to
that of colchicine (6) itself, which achieved (7.40 kcal mol
1
).
Also, their RMSD_refine values were 1.85 and 0.95, respectively,
indicating highly valid selected poses as well. On the other
hand, target derivative (5e) showed a very promising binding
score (7.38 kcal mol
1
) with an RMSD_refine value of 1.60
compared to colchicine (6) as mentioned before. At the same
Scheme 2 Mechanism of the reaction of 3,4,5-trimethoxybenzamidoxime (2)andN-acylbenzotriazoles 4a–k to form oxadiazoles 5a–k.
NJC Paper
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
New J. Chem. This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021
time, the obtained binding scores of compounds 5a and 5e
were very close to those of the co-crystallized inhibitor (7) with a
binding score of 8.13 kcal mol
1
.
Moreover, both 5a and 5e targets as the most two promising
derivatives bound Lys350 amino acid at 3.62 and 3.65 Å,
respectively, by a pi–H bond formation. Also, compound 5a
formed an extra pi-H bond with Leu253 at 3.92 Å and com-
pound 5e bound Met257 by a second H-bond at 3.52 Å (Table 4).
By considering the docking findings of the newly synthe-
sized compounds 5a–k against the binding pocket of the CBS in
b-tubulin compared to its co-crystallized inhibitor 7and colchi-
cine 6as a second promising reference standard, together with
the great matching with the biological cytotoxicity studies per-
formed as well, we can conclude a very promising idea concerning
their recommended mechanism of action as CBSIs.
2.4. Structure–activity relationship (SAR) study
SAR studies based on the biological activities (Tables 1 and 2) of
the newly synthesized series of 3,5 disubstituted 1,2,4-oxadi-
azoles regarding their pharmacophoric features (represented in
Fig. 3 and 4) together with the previously discussed docking
results (Tables 3 and 4), we can conclude the following inter-
esting results:
(a) The most potent designed target (5a), which possesses all
the required pharmacophoric points for binding to the CBS has
the highest antiproliferative activity. It was found to be the only
member containing a benzyl moiety as a planar pharmacophoric
group which may explain its superior anti-proliferative and
CBSI biological activities. Also, it may clarify the reason behind
its attractive fitting and binding score observed through the
Table 1 Percentage growth inhibition (GI%) of compounds 5a–k over 60 subpanel tumor cell lines
Comp.
no.
60 cell lines assay in one dose 10.0 mM concentration: GI%
MGI% MGI-S PCE Most sensitive cell lines
5a 15.5 24.9 33/60 Leukemia (K-562 : 30%, RPMI-8226: 20%, SR:23.%), NSC Lung Cancer (A549/ATCC:11%, EKVX: 23%, HOP-92:
61%, NCI-H226: 15%, NCI-H23: 13%, NCI-H322M:13%, NCI-H522: 22%), Colon Cancer (HCT-15: 19%,
HCT-116: 13%, HT29: 11%), CNS Cancer (SF-268: 14%, SF-295: 11%, SNB-75: 27%), Melanoma (LOX IMVI: 17%,
UACC-62: 35%, SK-MEL-5: 12%), Ovarian Cancer (IGROV1: 34%, OVCAR-4: 23%), Renal Cancer (A498: 78%,
CAKI-1: 43%, UO-31: 46%, 786-0: 15%, ACHN:15%, RXF 393: 12%, SN12C:19%) Prostate Cancer (PC-3: 37%),
Breast Cancer (MCF7: 23%,MDA-MB-231/ATCC: 29%, HS 578T: 26%, T-47D: 34%).
5b 8.1 17.9 23/60 Leukemia (HL-60(TB):21%, RPMI-8226: 23%), NSC Lung Cancer (EKVX: 25%,NCI-H226: 21%,A549/ATCC:11%,
NCI-H23: 12%, NCI-H522: 14%), Colon Cancer (HCT-15: 14%), CNS Cancer (SF-268: 13%, SF-539: 14%),
Melanoma (M14: 12%, UACC-62: 13%), Ovarian Cancer (IGROV1: 15% SK-OV-3: 11%), Renal Cancer (RXF 393:
15%, CAKI-1: 14%, UO-31: 22%), Prostate Cancer (PC-3: 31%), Breast Cancer (MCF7: 22%,MDA-MB-231/ATCC:
20%, T-47D:37%, MDA-MB-468: 23%, HS 578T:10%)
5c 8.1 19.8 16/60 Leukemia (CCRF-CEM:16%, HL-60(TB): 30%, K-562: 17%, RPMI-8226: 14%, MOLT-4: 12.%), NSC Lung Cancer
(A549/ATCC:17%, NCI-H522: 13%), Colon Cancer (HCT-15: 16%, KM12: 16%), Renal Cancer (SN12C:19%, CAKI-
1: 24%, UO-31: 40%), Prostate Cancer (PC-3: 33%), Breast Cancer (MCF7: 12%, BT-549: 12%, T-47D: 27%).
5d 9.4 21.3 21/60 Leukemia (K-562: 17%, CCRF-CEM:11%, SR:27.%, HL-60(TB):40%, MOLT-4: 14%,RPMI-8226: 21%), NSC Lung
Cancer (EKVX: 13%, NCI-H226: 14%, NCI-H522: 19%), Colon Cancer (HCT-15: 12%, KM12: 20%), CNS Cancer
(SNB-75: 32%), Melanoma (LOX IMVI: 14%, UACC-62: 15%), Renal Cancer (A498: 51%, CAKI-1: 24%, UO-31:
33%, SN12C:17%) Prostate Cancer (PC-3: 30%), Breast Cancer (MDA-MB-231/ATCC: 16%, T-47D: 16%)
5e 10.2 19.6 27/60 Leukemia (K-562: 24%, CCRF-CEM:11%, SR:11%, HL-60(TB):17%,RPMI-8226: 23%), NSC Lung Cancer (HOP-62:
19%, NCI-H23: 13%, NCI-H522: 25%), Colon Cancer (HCT-116: 13%), CNS Cancer (SF-539: 15%, SNB-19: 17%,
SNB-75: 59%), Melanoma (LOX IMVI: 13%, UACC-62: 22%, M14: 15%,K-MEL-2: 12%), Ovarian Cancer (OVCAR-
8: 12%), Renal Cancer (CAKI-1: 28%, UO-31: 25%, 786-0: 15%, ACHN:18%, SN12C:16%) Prostate Cancer (PC-3:
16%), Breast Cancer (MCF7: 12%,MDA-MB-231/ATCC: 31%, HS 578T: 34%, T-47D: 15%)
5f 2.2 19.8 9/60 NSC Lung Cancer (HOP-92: 16%, NCI-H226: 11%), CNS Cancer (SNB-75: 26%), Melanoma (UACC-62: 17%),
Renal Cancer (CAKI-1: 19%, UO-31: 31%, A498: 32%) Prostate Cancer (PC-3: 14%), Breast Cancer (T-47D: 13%)
5g 4.9 16.5 10/60 Leukemia (HL-60(TB):33%), NSC Lung Cancer (NCI-H226: 12%), Colon Cancer (HCT-15: 11%) Melanoma
(UACC-62: 19%), Renal Cancer (UO-31: 13%, SN12C:11%) Prostate Cancer (PC-3: 19%), Breast Cancer
(MCF7: 11%,MDA-MB-231/ATCC: 11%, T-47D: 25%)
5h 2.3 20.8 6/60 Leukemia (SR:19%), NSC Lung Cancer (HOP-92: 39%), CNS Cancer (SNB-75: 23%), Renal Cancer (UO-31: 20%)
Prostate Cancer (PC-3: 13%), Breast Cancer (HS 578T: 11%)
5i 2.5 14.4 10/60 Leukemia (K-562: 13%, MOLT-4: 13%), NSC Lung Cancer (NCI-H522: 12%), CNS Cancer (SNB-75: 17%,
SNB-19: 11%), Melanoma (UACC-62: 23%), Renal Cancer (UO-31: 11%, SN12C:12%) Breast Cancer
(MCF7: 15%, T-47D: 17%)
5j 1.9 13.6 7/60 NSC Lung Cancer (HOP-92: 12%), CNS fCancer (SNB-75: 24%, SNB-19: 11%), Melanoma (UACC-62: 12%),
Renal Cancer (UO-31: 11%) Breast Cancer (MCF7: 13%, MDA-MB-468: 12%)
5k 8.1 19.7 18/60 Leukemia (HL-60(TB):29%), NSC Lung Cancer (EKVX: 14%, HOP-92: 35%, NCI-H522: 19%), Colon Cancer (HCT-
15: 19%, HCT-116: 13%, HT29: 11%), CNS Cancer (SNB-19: 14%SNB-75: 30%), Melanoma (UACC-62: 28%),
Ovarian Cancer (IGROV1: 23%), Renal Cancer (CAKI-1: 17%, UO-31: 26%, 786-0: 14%, SN12C:15%) Prostate
Cancer (PC-3: 17%), Breast Cancer (MCF7: 18%,MDA-MB-231/ATCC: 20%, HS 578T: 11%, T-47D: 19%)
PCE: positive cytotoxic effect, which is the ratio between some cell lines with percentage growth inhibition 410% and the total number of cell
lines, MGI%: mean growth inhibition percentage, MGI-S%: mean growth inhibition percentage for the most sensitive cell lines (with percentage
growth inhibition 410%).
Table 2 IC
50
values for (TUBb) inhibition of compound 5a and colchicine
Compound Tubulin binding, IC
50
,uM
5a 1.18 0.037
Colchicine 2.37 0.075
Paper NJC
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
molecular docking study especially its binding to the crucial
amino acid (Lys350) that is mainly responsible for the inhibi-
tory effect of the CBS as discussed before. The positive cytotoxic
effect (PCE) of 5a is 33/60 and the mean growth inhibition
(MGI) for the most sensitive cell lines (33) is 24.9%, while MGI
for the 60 cell lines = 15.5%.
(b) Target (5e) has a second 3,4,5-trimethoxyphenyl group
(a phenyl moiety was added as the planar pharmacophoric
group attached to a trimethoxy group to introduce a second
3,4,5-trimethoxyphenyl group in addition to the first permanent
hydrophobic one). This allowed 5e to be the second most active
target of the whole series, PCE = 27/60 and MGI for 60 cell lines =
10.2%, MGI for 27 cell lines = 19.6. Besides, it is worth mention-
ing that 5e was also able to form a pi-H bond to the crucial
amino acid (Lys350) as described earlier, which may explain its
better anti-proliferative activity as the second most promising
candidate.
(c) The absence of benzyl side chain (planar group: R1)
in (5f) dramatically reduced its PEC to 9/60 and MGI for 60
cell lines to 2.2 and MGI for the nine cell lines to 19.8.
Table 3 3D representations showing the binding interactions and positioning of colchicine (6) and the docked co-crystallized inhibitor (7) into the CBS
in b-tubulin. The red dash represents H-bonds and the black dash represents H–pi interactions
Comp. 3D binding interactions 3D pocket positioning
*Col. (6)
*Inh. (7)
*Col. = colchicine; *Inh. = co-crystallized inhibitor of CBS.
Table 4 3D representations showing the binding interactions and positioning between the most two promising newly synthesized compounds
(5a and 5e) inside the CBS in b-tubulin. The red dash represents H-bonds and the black dash represents H–pi interactions
Comp. 3D binding interactions 3D pocket positioning
5a
5e
NJC Paper
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
New J. Chem. This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021
(d) Replacement of the benzyl group of (5a), with a branched
nonplanar hydrophobic group (isopropyl group) in (5j), gave
the least active target with PCE = 7/60 and the MGI for 60 cell
lines to = 1.9%.
(e) Insertion of a methylene group between the isopropyl
group and the main chain gave a slightly more active compound
(5k) (PCE = 18/60, MGI = 8.07%, and MGI for 18 cell lines =
19.7%).
(f) Replacement of the Cbz–Phe–OH side chain with o-, m-,
and p-aminophenyl groups gave (5b), (5c), and (5d), respec-
tively. The three compounds are of comparable PCEs (16–21/60)
and MGI for 60 cell lines = 8.1–9.4%. The order of activity was
p-(5d)4m-(5c)4o-(5b).
(g) Replacement of the p-amino-group of (5d)byp-methoxy
(5g) reduced the activity, PCE = 10/60 and MGI for 60 cell lines =
4.9%.
(h) Replacement of the aryl group at C5 by n-propyl in (5h)
gave the lowest activity (PCE = 6/60 and MGI for 60 cell lines =
2.3%). This indicates the great importance of the presence of
the second aryl moiety for the receptor fitting and the subse-
quent enhanced anti-proliferative activity of the studied mem-
bers as CBSIs.
(i) Finally, the presence of long-chain unsaturated alkyl
group (5i) did not significantly improve the activity (PCE = 10/
60 and MGI for 60 cell lines = 2.5%).
It is noteworthy that targets (5a–k) showed selective potency
towards renal cancer in a particular A498 cell line, which
represents renal cell carcinoma (RCC). RCC is one of the most
common kidney cancers and is highly resistant to chemo-
therapy.
39
The order of potency of the most potent targets
against A498 is 5a (GI = 78%) 45d (GI = 51%) 45f (32%).
Moreover, targets 5a–f, k inhibited the growth of CAKI-1-
clear cell renal cell carcinoma (ccRCC) cell line – with GI%
ranging from 14 – to 43%. ccRCC Is the most common and
lethal form of urological cancer diagnosed globally.
40
In addi-
tion, all the designed targets (5a–k) showed significant growth
inhibition against UO-31 (GI = 11–46%).
3. Conclusion
In conclusion, we have designed and synthesized eleven 3,5-
disubstituted oxadiazole derivatives (5a–k). The targets were
evaluated for their antiproliferative activity at 10
5
M in full
NCI 60 cell panel and showed moderate antiproliferative
activities. In addition, molecular docking studies revealed that
targets 5a–k showed promising affinities towards the colchicine
binding site in b-tubulin using. Target 5a demonstrated the
most potent antiproliferative activity in terms of positive cyto-
toxic effect (PCE = 33/60) and growth inhibition (MGI for 33 cell
lines = 24.9%) and showed a binding affinity more than that of
the docked Colchicine (6), and very near to that of the co-
crystallized inhibitor (7). Furthermore, 5a inhibited b-tubulin
polymerization in vitro with IC
50
= 1.18 mM, which is twofold the
potency of Colchicine (6), IC
50
= 2.37 mM. In the light of these
findings, 5a is a promising candidate for further investigation
as an anticancer CBSI and as a lead for future development of
more potent anticancer CBSIs to combat multidrug resistance
(MDR) cancers.
4. Experimental section
4.1. Chemistry
4.1.1. General. All reagents were purchased from common
commercial sources and used without further purification.
Melting points (m.p.) were measured on melting point appara-
tus SMP3, Stuart Scientific (UK), and were uncorrected.
1
H NMR
(400 MHz) and
13
C NMR (100 MHz) spectra were determined on
a Bruker 400 MHz FT-NMR spectrometer using DMSO-d
6
. The
chemical shift (d) was reported on ppm relative to the internal
tetramethylsilane (TMS) standard and the coupling constant ( J)
was calculated in Hz (Hertz). IR was determined using Bruker
Alpha ft IR spectrometer. TLC-MASS was determined using
Advion compact mass spectrometer (CMS) NY|USA using elec-
trospray (ES) ionization mode. Reactions were monitored
by TLC (Thin Layer Chromatography) on silica gel 50 GF254
(E-Merck, Germany) and using a UV lamp for visualization at a
wavelength (l) of 254 nm.
4.1.1. Synthesis of 3,4,5-trimethoxybenzamidoxime (2).
3,4,5-Trimethoxy benzonitrile (1) (5 mmol, 950 mg) was reacted
with hydroxylamine hydrochloride (5 mmol, 410 mg) and
sodium carbonate (5 mmol, 540 mg) in ethanol (25 ml). The
mixture was heated under reflux for 5 h. To drive the reaction to
completion, additional hydroxylamine hydrochloride (5 mmol,
410 mg) and sodium carbonate (5 mmol, 540 mg) were added
and the reaction was heated under reflux for another 14 hrs.
The mixture was filtered while hot. The product was then
crystallized from ethanol giving amidoxime as pure white
crystals, 95% yield, m.p. 155–157 1C (as reported).
41
1
H NMR (DMSO-d
6
, 400 MHz, dppm): 9.55 (s, 1H, OH), 5.79
(s, 2H, aromatic H), 5.80 (s, 2H, NH
2
), 3.78 (s, 5H, 3,5
di-methoxy), 3.55 (s, 3H, 4-methoxy).
4.1.2. Synthesis of N-acyl benzotriazoles (4a, e–k). A mix-
ture of benzotriazole (4 mmol, 480 mg) in methylene chloride
and SOCl
2
(1.2 mmol, 140 mg) was stirred for 20–30 m at
(20 1C), then appropriate acid (1 mmol) was added. The stirring
was continued for 2–3 hr at room temperature, then acidic
workup using 5 N hydrochloric acid (3 20 ml) or basic workup
using sodium carbonate (3 20 ml) was performed. The
methylene chloride layer was dried over anhydrous sodium
sulfate. The solid was crystallized from hexane/methylene chloride
to give the corresponding N-acyl benzotriazole (4a, e–k).
Benzyl (S)-(1-(1H-benzo[d][1,2,3]triazol-1-yl)-1-oxo-3-phenylpropan-
2-yl)carbamate (4a). White microcrystals, yield (98%), m.p. 160–
162 1C (ref. 32, 162–163 1C).
(2-Aminophenyl)(1H-benzo[d][1,2,3]triazol-1-yl)methanone (4b).
Pale yellow microcrystals, yield (60%), m.p. 130–132 1C(ref.33,
130–132 1C).
Paper NJC
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
(3-Aminophenyl)(1H-benzo[d][1,2,3]triazol-1-yl)methanone (4c).
Pale yellow microcrystals, yield (75%), m.p. 171–172 1C(ref.31,
171–172 1C).
(4-Aminophenyl)(1H-benzo[d][1,2,3]triazol-1-yl)methanone (4d).
Pale yellow microcrystals, yield (80%), m.p. 178–180 1C(ref.31,
178–180 1C).
(1H-Benzo[d][1,2,3]triazol-1-yl)(3,4,5-trimethoxyphenyl)methanone
(4e). White microcrystals, yield (95%), m.p. 125–127 1C(ref.33,
126–128 1C).
Benzyl (2-(1H-benzo[d][1,2,3]triazol-1-yl)-2-oxoethyl)carbamate
(4f). White microcrystals, yield (93%), m.p. 110–112 1C (ref. 33,
107–109 1C).
(1H-Benzo[d][1,2,3]triazol-1-yl)(4-methoxyphenyl)methanone (4g).
White microcrystals, yield (90%), m.p. 103–105 1C (ref. 33, 102–
104 1C).
1-(1H-Benzo[d][1,2,3]triazol-1-yl)butan-1-one (4h). White
microcrystals, yield (88%), m.p. 57–59 1C (ref. 34, 59–60 1C).
(9Z,12Z)-1-(1H-Benzo[d][1,2,3]triazol-1-yl)octadeca-9,12-dien-1-
one (4i). Colorless oil, yield (90%), (ref. 35 oil).
Benzyl (S)-(1-(1H-benzo[d][1,2,3]triazol-1-yl)-3-methyl-1-oxobutan-
2-yl)carbamate (4j). White microcrystals, yield (87%), m.p. 107–
108 1C (ref. 36, 107–108 1C).
Benzyl (S)-(1-(1H-benzo[d][1,2,3]triazol-1-yl)-4-methyl-1-oxopentan-
2-yl)carbamate (4k). White microcrystals, yield (80%), m.p. 150–
151 1C (ref. 32, 150–152 1C).
4.1.3. Synthesis of N-acyl benzotriazole from m,p-amino-
benzoic acid (4b–d). A mixture of p-toluene sulfonyl chloride
(1 mmol, 190 mg) and DMAP (0.13 mmol, 15 mg) was stirred in
methylene chloride (5 ml) for 10 min. o,m, and p-aminobenzoic
acids (1 mmol, 0.14 gm) were dissolved in methylene chloride
(5 ml) containing TEA (1.5 mmol, 151 mg), then were added
to the reaction mixture. After 20 m, benzotriazole (1.2 mmol,
142 mg) was added and the reaction was stirred for another
1.5 h at 25 1C. When the reaction was complete, methylene
chloride (50 ml) was added and the organic layer was washed
with saturated sodium carbonate (3 10 ml), water (2 10 ml),
and brine (1 10 ml). The methylene chloride layer was dried
over anhydrous sodium sulfate, then hexane (20 ml) was added.
The precipitated solid was filtered and dried under vacuum to
give the target N-acylbenzotriazoles (4b–d).
4.1.4. Synthesis of 3,5-disubstituted-1,2,4-oxadiazoles
(5a–k). 3,4,5-Trimethoxy benzamidoxime (1 mmol, 230 mg)
was reacted with N-acyl benzotriazoles (1 mmol) in ethanol
(20 ml) containing TEA (1 mmol, 100 mg). The mixture was
stirred for 10–20 m, then it was heated under reflux for (2–30 h)
for different acyl benzotriazoles. The product was obtained by
precipitation from ethanol. The reaction mixture was filtered,
then the product was washed on a filter paper with water,
followed by ethanol, and dried under vacuum and recrystallized
from ethyl acetate to obtain pure oxadiazoles. The oily oxadia-
zoles (5i–k) were obtained by evaporation of ethanol under
vacuum, then the residue was dissolved in ethyl acetate (15 ml)
and washed with sodium carbonate (3 5 ml) to remove
benzotriazole, then dried over anhydrous sodium sulfate, ethyl
acetate was evaporated under vacuum to give the product.
4.1.4.1. Benzyl (S)-(2-phenyl-1-(3-(3,4,5-trimethoxyphenyl)-
1,2,4-oxadiazol-5-yl)ethyl) carbamate (5a). White solid. Yield:
90%, m.p. 112–115 1C.
1
H NMR (DMSO-d
6
, 400 MHz, d, ppm):
8.36 (d, 1H, J= 8 Hz, N
H), 7.30–7.25 (m, 12H, C
2
H, C
6
Hof
trimethoxy phenyl ring and 10 Ar–
H of phenyl rings), 5.20 (d,
1H, J= 6 Hz, C
H–NH), 5.00 (s, 2
H, CH
2
–O), 3.86 (s, 6H, 3,5-di
(OC
H
3
) of 3,4,5-trimethoxyphenyl), 3.75 (s, 3H, 4-OC
H
3
of 3,4,5-
trimethoxyphenyl), 3.36 (s, 2H, C
H
2
-phenyl).
13
C NMR (DMSO-
d
6
, 100 MHz, d, ppm): 179.81 (
C
3
-1,2,4-oxadiazole), 167.44 (
C
5
of
1,2,4-oxadiazole), 155.76 (
CQO), 153.40 (
C
3
and
C
5
of 3,4,5-
trimethoxyphenyl), 140.22 (
C
4
of 3,4,5-trimethoxyphenyl),
136.74 (
C
1
of phenyl of Cbz), 136.55 (
C
1
of phenyl of Phe),
129.28 (
C
3
and
C
5
of phenyl of Cbz and
C
3
and
C
5
of phenyl of
Phe), 128.31 (
C
4
of phenyl of Cbz and Phe), 127.83 (
C
2
and
C
6
of phenyl of Phe), 127.57 (
C
2
and
C
6
of phenyl of Cbz), 121.27
(
C
1
of 3,4,5-trimethoxyphenyl), 104.25 (
C
2
and
C
6
of 3,4,5-
trimethoxyphenyl), 65.64 (
CH–O), 60.1 (4-O
CH
3
), 56.05 (3,5-di-
O
CH
3
), 49.91 (
CH
2
–N), 37.63 (
CH
2
– of Phe). Anal. calcd for
C
27
H
27
N
3
O
5
: C, 66.25, H, 5.56, N, 8.58%; found; C, 66.43; H,
5.79; N, 8.81%.
4.1.4.2. 2-(3-(3,4,5-Trimethoxyphenyl)-1,2,4-oxadiazol-5-yl)-
aniline (5b). White solid. Yield: 85%, m.p. 152–155 1C.
1
HNMR
(DMSO-d
6
, 400 MHz, d, ppm): 7.857.83 (d, 1H, J=7.2Hz,C
3
Hof
aniline), 7.39–7.35 (m, 3H, C
5
H of aniline and NH
2
), 6.97–6.93
(m, 3H, C
6
H of aniline and C
2
H, C
6
Hoftrimethoxyphenyl),
6.69 (s, 1H, C
4
H of aniline), 3.90 (s, 6H, 3,5 di (OC
H
3
)of
3,4,5-trimethoxy phenyl), 3.75 (s, 3H, 4-OC
H
3
of 3,4,5-tri-
methoxyphenyl).
13
C NMR (DMSO-d
6
, 100 MHz, d,ppm):
174.64 (
C
5
-1,2,4-oxadiazole), 167.23 (
C
3
-1,2,4-oxadiazole), 153.38
(
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 148.74 (
C
1
of aniline),
140.24 (
C
4
of 3,4,5-trimethoxyphenyl), 134.05 (
C
3
of aniline),
128.49 (
C
5
of aniline), 121.49 (
C
1
of 3,4,5-trimethoxyphenyl),
116.59 (
C
2
of aniline), 115.63(
C
6
of aniline), 104.64 (
C
2
and
C
6
of 3,4,5-trimethoxyphenyl), 103.46 (
C
4
of aniline), 60.18 (3,5 di-
CH
3
O), 56.16 (4-
CH
3
O). Anal. calcd for C
17
H
17
N
3
O
4
:C,62.38;H,
5.23; N, 12.84%; found; C, 62.16; H, 5.37; N, 13.12%.
4.1.4.3. 3-(3-(3,4,5-Trimethoxyphenyl)-1,2,4-oxadiazol-5-yl)-
aniline (5c). White solid. Yield: 40%, m.p. 112–114 1C.
1
HNMR
(DMSO-d
6
, 400 MHz, d, ppm): 7.40–7.27 (m, 5H, C
5
,C
4
and C
2
–H
of aniline, C
2
–H, C
6
–H of trimethoxy phenyl), 6.88 (d, 1H, J=6.8Hz,
C
6
–H of aniline), 5.60 (s, 2H, NH
2
), 3.89 (s, 6H, 3,5-di(OCH
3
)
of 3,4,5-trimethoxy phenyl), 3.75 (s, 3H,4-OCH
3
of 3,4,5-
trimethoxyphenyl).
13
C NMR (DMSO-d
6
, 100 MHz, d,ppm):
176.04 (
C
5
-1,2,4-oxadiazole), 167.99 (
C
3
-1,2,4-oxadiazole), 153.39
(
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 149.65 (
C
1
of aniline),
140.16 (
C
4
of 3,4,5-trimethoxyphenyl), 130.03 (
C
5
of aniline),
123.73 (
C
3
of aniline), 121.57 (
C
1
of 3,4,5-trimethoxyphenyl),
118.46 (
C
6
of aniline), 114.93 (
C
4
of aniline), 112.22 (
C
2
of
aniline), 104.28 (
C
2
and
C
6
of 3,4,5-trimethoxyphenyl), 60.19
NJC Paper
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
New J. Chem. This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021
(3,5 di-
CH
3
O), 56.05 (4-
CH
3
O). Anal. calcd for C
17
H
17
N
3
O
4
:C,
62.38; H, 5.23; N, 12.84%; found; C, 62.50; H, 5.41; N, 13.10%.
4.1.4.4. 4-(3-(3,4,5-Trimethoxyphenyl)-1,2,4-oxadiazol-5-yl)-
aniline (5d). White solid. Yield: 51%, m.p. 155–158 1C. IR (KBr)
cm
1
: 3300 (NH
2
), 2928, 2855 (CH aliphatic), 1559 (CQN), 1512
(CQC), 1258, 1125(C–O).
1
H NMR (DMSO-d
6
, 400 MHz, d, ppm):
7.85 (d, 2H, J= 8.8 Hz, C
3
H, C
5
H of aniline), 7.31 (s, 2H, C
2
H,
C
6
H of trimethoxy phenyl), 6.71 (d, 2H, J= 8.4 Hz, C
2
H, C
6
H
of aniline), 6.18(s, 2H, N
H
2
), 3.88 (s, 6H, 3,5-di (OC
H
3
)of
3,4,5-trimethoxy phenyl), 3.74 (s, 3H, 4-OC
H
3
of 3,4,5-tri-
methoxyphenyl).
13
C NMR (DMSO-d
6
, 100 MHz, d, ppm):
176.55 (
C
5
-1,2,4-oxadiazole), 168.17 (
C
3
-1,2,4-oxadiazole),
153.92 (
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 153.82 (
C
1
of
aniline), 140.45 (
C
4
of 3,4,5-trimethoxyphenyl), 130.32 (
C
3
and
C
5
of aniline), 122.39 (
C
1
of 3,4,5-trimethoxyphenyl), 114.07 (
C
2
and
C
6
of aniline), 110.04 (
C
1
of aniline), 104.78 (
C
2
and
C
6
of
3,4,5-trimethoxyphenyl), 60.81 (3,5 di-
CH
3
O), 56.58 (4-
CH
3
O).
Anal. calcd for C
17
H
17
N
3
O
4
: C, 62.38; H, 5.23; N, 12.84%; found;
C, 62.19; H, 5.40; N, 13.08%.
4.1.4.5. 3,5-Bis(3,4,5-trimethoxyphenyl)-1,2,4-oxadiazole (5e).
White solid. Yield: 52%, m.p. 152–155 1C.
1
H NMR (DMSO-d
6
,
400 MHz, d, ppm): 7.43 (s, 2H, C
2
-
H, C
6
H of trimethoxy phenyl
on C
5
), 7.35 (s, 2H, C
2
H, C
6
H of trimethoxy phenyl on C
3
),
3.92 (s, 12H, 2 [3,5-di(OC
H
3
) of 3,4,5 - trimethoxy phenyl]),
3.78 (s, 6H, 2 [4-O
C
H
3
of 3,4,5-trimethoxyphenyl]).
13
C NMR
(DMSO-d
6
, 100 MHz, d, ppm): 175.19 (
C
5
-1,2,4-oxadiazole),
167.97 (
C
3
-1,2,4-oxadiazole), 153.46 (
C
3
and
C
5
of 3,4,5-
trimethoxyphenyl on C
5
-1,2,4-oxadiazole), 153.38 (
C
3
and
C
5
of 3,4,5-trimethoxyphenyl on C
3
-1,2,4-oxadiazole), 140.16 (
C
4
of
3,4,5-trimethoxyphenyl-1,2,4-oxadiazole), 129.99 (
C
1
of 3,4,5-
trimethoxyphenyl), 121.58 (
C
1
of 3,4,5-trimethoxyphenyl),
104.56 (
C
2
and
C
5
of 3,4,5-trimethoxyphenyl), 104.47 (
C
2
and
C
6
of 3,4,5-trimethoxyphenyl), 60.30 (4-
CH
3
O of 3,4,5-
trimethoxyphenyl), 60.19 (4-
CH
3
O of 3,4,5-trimethoxyphenyl),
56.29 (3,5 di-
CH
3
O of 3,4,5-trimethoxyphenyl), 56.06 (3,5 di-
CH
3
O of 3,4,5-trimethoxyphenyl). Anal. calcd for C
20
H
22
N
2
O
7
:C,
59.70; H, 5.51; N, 6.96%; found; C, 59.89; H, 5.63; N, 7.22%.
4.1.4.6. Benzyl (3-(3,4,5-trimethoxyphenyl)-1,2,4-oxadiazol-5-
yl)methylcarbamate (5f). White solid. Yield: 70%, m.p. 102–
104 1C.
1
H NMR (DMSO-d
6
, 400 MHz, d, ppm): 8.18 (s, 1H,
N
H), 7.35 (m, 5H, C
H of phenyl-CH
2
ring), 7.23 (s,2H, C
2
H,
C
6
H of trimethoxy phenyl), 5.07 (s, 2H, C
H
2
O), 4.58(s, 2H,
CH
2
N
H), 3.84 (s, 6H, 3,5-di(OC
H
3
) of 3,4,5-trimethoxy phenyl),
3.73 (s, 3H, 4-OC
H
3
of 3,4,5-trimethoxyphenyl).
13
C NMR
(DMSO-d
6
, 100 MHz, d, ppm): 177.83 (
C
5
-1,2,4-oxadiazole),
167.54 (
C
3
-1,2,4-oxadiazole), 156.45 (CQO), 153.44 (
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 140.22 (
C
4
of 3,4,5-trimethoxy-
phenyl), 136.77 (
C
1
of phenyl of Cbz), 128.44 (
C
3
and
C
5
of
phenyl of Cbz), 127.87 (
C
2
and
C
4
and
C
6
of phenyl of Cbz),
121.33 (
C
1
of 3,4,5-trimethoxyphenyl), 104.27 (
C
2
and
C
6
of
3,4,5-trimethoxyphenyl), 65.97 (
CH–O), 60.23 (4-
CH
3
Oof
3,4,5-trimethoxyphenyl), 56.09 (3,5 di-
CH
3
O of 3,4,5-trimethoxy-
phenyl), 37.05 (
CH
2
–N). Anal. calcd for C
20
H
21
N
3
O
5
: C, 60.14, H,
5.30, N, 10.52%; found; C, 60.39; H, 5.52; N, 10.79%.
4.1.4.7. 4-(Methoxyphenyl)-3-(3,4,5-trimethoxyphenyl)-1,2,4-oxa-
diazole (5g). White solid. Yield: 40%, m.p. 148–150 1C; IR (KBr)
cm
1
: 2943, 2930 (CH aliphatic), 1583 (CQN), 1502 (CQC), 1230,
1128 (C–O). Anal. calcd for C
20
H
22
N
2
O
7
: C, 63.15; H, 5.30; N,
8.18%; found; 63.38; H, 5.43; N, 8.40%.
4.1.4.8. 5-Propyl-3-(3,4,5-trimethoxyphenyl)-1,2,4-oxadiazole
(5h). White solid. Yield: 50%, m.p. 53–55 1C.
1
H NMR (DMSO-
d
6
, 400 MHz, d, ppm): 7.26 (s, 2H, C
2
H, C
6
Hoftrimethoxy
phenyl), 3.85 (s, 6H,3,5-di (OC
H
3
) of 3,4,5-trimethoxy phenyl),
3.74 (s, 3H, 4-OC
H
3
of 3,4,5-trimethoxyphenyl), 3.00 (t, 2H,
J=7.4Hz,C
1
H of propyl), 1.84–1.80 (m, 2H, C
2
H of propyl),
1.00 (t, 3H, J=7.4Hz,C
3
Hofpropyl).
13
CNMR(DMSO-d
6
,
100 MHz, d, ppm): 180.17 (
C
5
-1,2,4-oxadiazole), 167.34
(
C
3
-1,2,4-oxadiazole), 153.36 (
C
3
and
C
5
of 3,4,5-trimethoxy-
phenyl), 140.03 (
C
4
of 3,4,5-trimethoxyphenyl), 121.5733
(
C
1
of 3,4,5-trimethoxyphenyl), 104.17 (
C
2
and
C
6
of 3,4,5-
trimethoxyphenyl), 60.15 (4-
CH
3
O of 3,4,5-trimethoxyphenyl),
56.02 (3,5 di-
CH
3
O of 3,4,5-trimethoxyphenyl), 27.58 (
C
1
of
propyl), 19.63 (
C
2
of propyl), 13.36 (
C
3
of propyl). Anal. calcd
for C
14
H
18
N
2
O
4
: C, 60.42; H,6.52%; N, 10.07; found; C, 60.71;
H, 6.68; N, 10.28%.
4.1.4.9. 5-((9Z,12Z)-Heptadeca-9,12-dien-1-yl)-3-(3,4,5-trimethoxy-
phenyl)-1,2,4-oxadiazole (5i). Orange oil. Yield: 50%.
1
H NMR
(DMSO-d
6
, 400 MHz, d, ppm): 7.25 (s, 2H, C
2
H, C
6
Hof
trimethoxy phenyl), 5.32–5.30 (m, 4H, 2 (C
HQC
H)), 3.85
(s, 6H, 3,5-di (OC
H
3
) of 3,4,5-trimethoxy phenyl), 3.73 (s, 3H,
4-OC
H
3
of 3,4,5-trimethoxyphenyl), 3.00 (t, 2H, J= 7.6 Hz, C
Hof
(CHQC–CH–C = CH)), 2.73 (t, 2H, J= 6 Hz, C
1
H of hepta-
decadienyl), 2.0 (dd, 4H, J= 6.4, 6 Hz, 2 [C
H
2
of (CH
2
–CQCH)]),
1.78 (q, 2H, C
2
H of heptadecadienyl), 1.32–1.23 (m, 14H, C
3
to
C
7
H and C
15
,C
16
–H of heptadecadienyl), 0.83 (t, 3H,J= 6.8 Hz,
C
17
H of heptadecadienyl).
13
C NMR (DMSO-d
6
, 100 MHz, d,
ppm): 180.38 (
C
5
-1,2,4-oxadiazole), 167.15 (
C
3
-1,2,4-oxadiazole),
153.51 (
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 140.49 (
C
4
of 3,4,5-
trimethoxyphenyl), 130.40 (
C
9
and
C
13
of heptadecadienyl),
128.82 (
C
12
of heptadecadienyl), 127.21 (
C
10
of heptadeca-
dienyl), 121.63 (
C
1
of 3,4,5-trimethoxyphenyl), 104.20 (
C
2
and
C
6
of 3,4,5-trimethoxyphenyl), 60.18 (4-
CH
3
O of 3,4,5-trimeth-
oxyphenyl), 56.03 (3,5 di-
CH
3
O of 3,4,5-trimethoxyphenyl), 28.13
(
C
15
of heptadecadienyl), 25.66 (
C
11
of heptadecadienyl), 22.105
(
C
16
of heptadecadienyl) 13.83 (
C
17
of heptadecadienyl). Anal.
calcd for C
28
H
42
N
2
O
4
: C, 71.46; H, 9.00; N, 5.95%; found; C,
71.32; H, 8.95; N, 6.17%.
4.1.4.10. Benzyl(S)-(2-methyl-1-(3-(3,4,5-trimethoxyphenyl)-1,2,4-
oxadiazol-5-yl)propyl)carbamate (5j). Colorlessoil,yield:55%;IR
(KBr) cm
1
: 3301 (NH), 2954, 2938 (CH aliphatic), 1702 (CQO),
1570 (CQN), 1525 (CQC), 1231, 1125 (C–O);
1
HNMR(DMSO-d
6
,
400 MHz, d, ppm): 8.30 (d, 1H, J=8Hz,N
H), 7.35–7.25 (m, 7H,
C
2
H, C
6
H of trimethoxy phenyl and 5
HofphenylofCbz),5.07
(s, 2H, C
H
2
O), 4.80 (t,1H, J=7.6Hz,C
H-(CH(CH
3
)
2
)), 3.86 (s, 6H,
3,5-di (OC
H
3
) of 3,4,5-trimethoxy phenyl), 3.74 (s, 3H, 4-OC
H
3
of
3,4,5-trimethoxyphenyl), 2.28 (m, 1H, C
H–(CH
3
)
2
), 0.89 (d, 6H, J=
6.8 Hz, CH-(C
H
3
)
2
).
13
CNMR(DMSO-d
6
,100MHz,d, ppm):
179.80 (
C
5
-1,2,4-oxadiazole), 167.44 (
C
3
-1,2,4-oxadiazole), 156.24
Paper NJC
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
(
CQO), 153.40 (
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 140.22
(
C
4
of 3,4,5-trimethoxyphenyl), 136.74 55 (
C
1
of phenyl of Cbz),
128.37 (
C
3
and
C
5
of phenyl of Cbz), 127.93 (
C
2
,
C
4
and
C
6
of
phenyl of Cbz), 121.33 (
C
1
of 3,4,5-trimethoxyphenyl), 104.46
(
C
2
and
C
6
of 3,4,5-trimethoxyphenyl), 65.84 (
CH
2
-O), 60.19
(4-
CH
3
O of 3,4,5-trimethoxyphenyl), 56.05 (3,5 di-
CH
3
O of 3,4,5-
trimethoxyphenyl), 31.07 (
CH-(CH
3
)
2
), 18.21 (CH-(
CH
3
)
2
). Anal.
calcd for C
23
H
27
N
3
O
5
: C, 62.57; H, 6.16; N, 9.52; found; C, 62.75;
H, 6.42; N, 9.73%.
4.1.4.11. Benzyl (S)-(3-methyl-1-(3-(3,4,5-trimethoxyphenyl)-
1,2,4-oxadiazol-5-yl)butyl)carbamate (5k). Colorless oil, yield:
55%.
1
H NMR (DMSO-d
6
, 400 MHz, d, ppm): 8.26 (d, 1H, J=
7.6 Hz, N
H), 7.36–7.30 (m, 5H, 5
H of phenyl of Cbz), 7.24 (s, 2H,
C
2
–H, C
6
–H of trimethoxy phenyl ring) 5.07 (s, 2H, C
H
2
O),
4.96 (app s,1H, CH–N
H), 3.85 (s, 6H, 3,5-di (OC
H
3
) of 3,4,5-
trimethoxy phenyl), 3.74 (s, 3H, 4-OC
H
3
of 3,4,5-trimethoxy-
phenyl), 1.73 (d, 2H, J= 8.4 Hz, C
H
2
–CH(CH
3
)
2
), 1.70–1.67
(m, 1H, C
H(CH
3
)
2
), 0.93 (d, 6H, J= 6.4 Hz, CH(C
H
3
)
2
).
13
C
NMR (DMSO-d
6
100 MHz, d, ppm): 180.60 (
C
5
-1,2,4-oxadiazole,
167.52 (
C
3
-1,2,4-oxadiazole), 156.01 (
CQO), 153.42 (
C
3
and
C
5
of 3,4,5-trimethoxyphenyl), 140.23 (
C
4
of 3,4,5-trimeth-
oxyphenyl), 136.78 (
C
1
of phenyl of Cbz), 128.41 (
C
3
and
C
5
of phenyl of Cbz), 127.72 (
C
2
,
C
4
and
C
6
of phenyl of Cbz),
121.33 (
C
1
of 3,4,5-trimethoxyphenyl), 104.30 (
C
2
and
C
6
of 3,4,5-trimethoxyphenyl), 65.86 (
CH
2
–O), 60.22 (4-
CH
3
Oof
3,4,5-trimethoxyphenyl), 56.09 (3,5 di-
CH
3
O of 3,4,5-trimeth-
oxyphenyl), 46.89 (
CH
2
–NH), 24.18 (
CH(CH
3
)
2
, 22.66 (CH(
CH
3
)
2
.
Anal. calcd for C
24
H
29
N
3
O
6
: C, 63.28 H, 6.42; N, 9.23%,; found; C,
63.39; H, 6.48; N, 9.47%.
4.2. Biological evaluation
4.2.1. Antiproliferative screening. Antiproliferative effects
of compounds 5a–k were screened over about 60 human tumor
cell lines derived from nine different cancer types, according to
the protocol of the Developmental Therapeutics Program,
National Cancer Institute, Bethesda, MD.
42
4.2.2. Evaluation of in vitro tubulin polymerization inhibi-
tion. The effect of compound 5a – as the most promising
candidate on tubulin polymerization was determined using
an ELISA kit for Tubulin b(TUBb), SEB870Hu (Cloud-Clone
Corp., TX, USA). The microtiter plate provided in kits was pre-
coated with a biotin-conjugated antibody specific to Tubulin
Beta (TUBb). Standards or samples were then added to the
appropriate microtiter plate wells with a biotin-conjugated
antibody. Avidin protein conjugated to horseradish peroxidase
(HRP) enzyme was added to microplates to bind the biotin-
labeled antibody and incubated. After the addition of TMB
(3,30,5,500 -tetramethylbenzidine) substrate solution, only wells
that contained the complex gave a characteristic color change.
This color was measured spectrophotometrically after the addi-
tion of the sulfuric acid solution to stop the enzyme–substrate
reaction. The decrease of color intensity was used as an
indicator for tubulin inhibition.
4.2.3. Tubulin polymerization assay. The assay was made
using a highly purified porcine tubulin. The kit used was
(cytoskeleton, cat. BK011P). At first, 95 well plates were warmed
in the fluorimeter at 37 1C for 10 min. After the preparation of
the tested compound in solution 5a, 1.5 mL of the buffer 1 and
20 mL of GTP stock were defrosted and put in ice. Tubulin
glycerol buffer was then removed from 4 1C and placed on the
ice. 88 ml of tubulin was defrosted in a room temperature water
bath, then instantly placed on the ice. Assay components were
immediately mixed, then 5 ml of the control buffer was pipetted
into the first duplicate wells A1 and B1 as well as, 5 mL of
colchicine, followed by 5 mL of the tested compound 5a. The
plate was placed back into a warm plate reader for 1 min, then
50 mL of the tubulin reaction mixture was added into each of the
eight wells and the plate reader was turned on. The intensity of
fluorescence was determined every 50 s for 90 min.
43
After polymerization of tubulin, an increase in fluorescence
emission was measured at 410–450 nm over a 50 min period at
37 1C to the first duplicate wells. 50 mL of the tubulin reaction
was added into each of the eight wells using a single channel
pipette. The fluorescence intensity was recorded every 50 s for
90 min in a multifunctional microplate reader. The area under
the curve was used to determine the concentration that inhi-
bited tubulin polymerization by 50% (IC
50
) (Table 2).
4.3. Molecular docking
Molecular docking studies using MOE 2019.012 suite
44,45
were
performed to propose the expected mechanism of the action for
the tested compounds 5a–k as CBSIs by evaluating their bind-
ing scores and interaction modes at the b-tubulin active site of
the CBS. The co-crystallized native inhibitor (N-[(7S)-1,2,3,10-
tetramethoxy-9-oxo-5,7-dihydro-5H-benzo[d]heptalen-7-yl]ethan-
amide, 7) was isolated from the protein complex and inserted in
the database together with colchicine (colchicine, 6) as the two
reference standards.
4.3.1. Preparation of the newly synthesized oxadiazole
tested compounds (5a–k) and colchicine (6). Each derivative
of the aforementioned newly synthesized compounds (5a–k)
and colchicine (6) as well was built using the MOE builder, and
then prepared for the docking process as described earlier.
46–48
Moreover, all of the prepared synthesized compounds (5a–k)
and the prepared colchicine molecules (6) were imported in
the same database together with the co-crystallized native
inhibitor ((N-[(7S)-1,2,3,10-tetramethoxy-9-oxo-5,7-dihydro-5H-
benzo[d]heptalen-7-yl]ethanamide (7) and saved as an MDB file
to be suitable for importing in the docking process.
4.3.2. Preparation of the active site of colchicine binding
site in b-tubulin. The Protein Data Bank website (https://www.
rcsb.org/) was used to download the X-ray structure of the CBS
receptor in b-tubulin (PDB ID: 5XIW, Resolution: 2.90 Å).
49
Then, the CBS complex was subjected to the default preparation
steps described before in detail.
50–52
4.3.3. Validation of the applied docking process using
the MOE program. A validation process for the MOE docking
procedure was performed at first and before the docking
process by redocking the co-crystallized native inhibitor (N-[(7S)-
1,2,3,10-tetramethoxy-9-oxo-5,7-dihydro-5H-benzo[d]heptalen-7-yl]-
ethanamide) inside its active site of the colchicine binding site
NJC Paper
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
New J. Chem. This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021
in b-tubulin. It is worth mentioning that a very small difference
was observed between both the co-crystallized (native) and the
docked inhibitor molecules (RMSD = 0.54) as represented in the
ESI,indicating a highly valid performance of the applied MOE
program.
53–55
4.3.4. Docking of the database to the binding pocket of
CBS in b-tubulin. A general docking process was performed
using the aforementioned prepared database, which was
inserted as an MDB file at the ligand site. The default specifica-
tions of the applied docking process were adjusted as repre-
sented before in detail.
56–58
Furthermore, we selected the best
pose for each studied compound according to its binding
score, RMSD value, and similar binding interactions with the
co-crystallized ligand for more considerations.
Conflicts of interest
The authors declare no conflict of interest.
References
1 H. Sung, et al., Global cancer statistics 2020: GLOBOCAN
estimates of incidence and mortality worldwide for 36
cancers in 185 countries, CA: A Cancer J. Clin., 2021, 71(3),
209–249.
2 M. R. Harrison, K. D. Holen and G. Liu, Beyond taxanes:
A review of novel agents that target mitotic tubulin and
microtubules, kinases, and kinesins, Clin. Adv. Hematol.
Oncol., 2009, 7(1), 54.
3 M. H. El-Shershaby, et al., From triazolophthalazines to
triazoloquinazolines: A bioisosterism-guided approach
toward the identification of novel PCAF inhibitors with
potential anticancer activity, Bioorg. Med. Chem., 2021, 42,
116266.
4 M. F. Elshal, Concanavalin-A shows synergistic cytotoxicity
with tamoxifen via inducing apoptosis in estrogen receptor-
positive breast cancer, Pharm. Sci., 2021, DOI: 10.34172/
PS.2021.22.
5 S. Eckhardt, Recent progress in the development of anti-
cancer agents, Curr. Med. Chemistry-Anti-Cancer Agents,
2002, 2(3), 419–439.
6 A. Al-Karmalawy, et al., Design and synthesis of new quin-
oxaline derivatives as potential histone deacetylase inhibi-
tors targeting hepatocellular carcinoma: In silico,in vitro,
and SAR studies, Front. Chem., 2021, 648.
7 M. S. Alesawy, et al., Design and discovery of new 1,2,4-
triazolo-[4,3-c]quinazolines as potential DNA intercalators
and topoisomerase II inhibitors, Arch. Pharm., 2020,
e2000237.
8 A. A. Gaber, et al., Pharmacophore-linked pyrazolo[3,4-
d]pyrimidines as EGFR-TK inhibitors: Synthesis, anticancer
evaluation, pharmacokinetics, and in silico mechanistic
studies, Arch. Pharm., 2021, e2100258.
9 C. Dominguez-Brauer, et al., Targeting mitosis in cancer:
Emerging strategies, Mol. Cell, 2015, 60(4), 524–536.
10 C. Dumontet and M. A. Jordan, Microtubule-binding agents:
a dynamic field of cancer therapeutics, Nat. Rev. Drug
Discovery, 2010, 9(10), 790–803.
11 A. E. Prota, et al., A new tubulin-binding site and pharma-
cophore for microtubule-destabilizing anticancer drugs,
Proc. Natl. Acad. Sci. U. S. A., 2014, 111(38), 13817–13821.
12 Y. Finkelstein, et al., Colchicine poisoning: The dark side of
an ancient drug, Clin. Toxicol., 2010, 48(5), 407–414.
13 Q. Wang, et al., Structure-guided design, synthesis, and
biological evaluation of (2-(1H-Indol-3-yl)-1H-Imidazol-4-yl)
(3,4,5-trimethoxyphenyl) methanone (ABI-231) analogues
targeting the colchicine binding site in tubulin, J. Med.
Chem., 2019, 62(14), 6734–6750.
14 M. Khattab and A. A. Al-Karmalawy, Computational repur-
posing of benzimidazole anthelmintic drugs as potential
colchicine binding site inhibitors, Future Med. Chem., 2021,
DOI: 10.4155/fmc-2020-0273.
15 E. C. McLoughlin and N. M. O’Boyle, Colchicine-binding
site inhibitors from chemistry to clinic: A review, Pharma-
ceuticals, 2020, 13(1), 8.
16 K. Sun, et al., Recent advances in research of colchicine
binding site inhibitors and their interaction modes with
tubulin, Future Med. Chem., 2021, 13(9), 839–858.
17 K. Ohsumi, et al., Syntheses and antitumor activity of cis-
restricted combretastatins: 5-membered heterocyclic analo-
gues, Bioorg. Med. Chem. Lett., 1998, 8(22), 3153–3158.
18 S. Ducki, The development of chalcones as promising
anticancer agents, IDrugs, 2007, 10(1), 42.
19 S. Ducki, et al., Combretastatin-like chalcones as inhibitors
of microtubule polymerization. Part 1: Synthesis and biolo-
gical evaluation of antivascular activity, Bioorg. Med. Chem.,
2009, 17(22), 7698–7710.
20 L. Lee, et al., Design, synthesis, and biological evaluations
of 2,5-diaryl-2,3-dihydro-1,3,4-oxadiazoline analogs of
combretastatin-A4, J. Med. Chem., 2010, 53(1), 325–334.
21 J. Chen, et al., Synthesis and antiproliferative activity of
novel 2-aryl-4-benzoyl-imidazole derivatives targeting
tubulin polymerization, Bioorg. Med. Chem., 2011, 19(16),
4782–4795.
22 Y. Lu, et al., Design, synthesis, and SAR studies of 4-substituted
methoxylbenzoyl-aryl-thiazoles analogues as potent and orally
bioavailable anticancer agents, J. Med. Chem., 2011, 54(13),
4678–4693.
23 M. P. Tantak, et al., Indolyl-a-keto-1, 3, 4-oxadiazoles: Synth-
esis, anti-cell proliferation activity, and inhibition of tubulin
polymerization, Bioorg. Med. Chem. Lett., 2021, 37, 127842.
24 R. Gaspari, et al.,Structuralbasisofcis-and trans-
combretastatin binding to tubulin, Chem,2017, 2(1), 102–113.
25 A. M. El-Naggar, et al., Design, eco-friendly synthesis,
molecular modeling and anticancer evaluation of thiazol-
5 (4H)-ones as potential tubulin polymerization inhibitors
targeting the colchicine binding site, RSC Adv.,2020,10(5),
2791–2811.
26 T. L. Nguyen, et al., A common pharmacophore for a diverse
set of colchicine site inhibitors using a structure-based
approach, J. Med. Chem., 2005, 48(19), 6107–6116.
Paper NJC
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2021 New J. Chem.
27 A. A. Al-Karmalawy and M. J. Khattab, Molecular modelling of
mebendazole polymorphs as a potential colchicine binding
site inhibitor, New J. Chem.,2020,44(33), 13990–13996.
28 M. Khattab and A. A. Al-Karmalawy, Revisiting activity
of some nocodazole analogues as a potential anticancer
drugs using molecular docking and DFT calculations, Front.
Chem., 2021, 9, 92.
29 D. Kumar, et al., Synthesis of novel 1,2,4-oxadiazoles and
analogues as potential anticancer agents, Eur. J. Med. Chem.,
2011, 46(7), 3085–3092.
30 A. R. Katritzky, Y. Zhang and S. K. Singh, Efficient conver-
sion of carboxylic acids into N-acylbenzotriazoles, Synthesis,
2003, 2795–2798.
31 K. A. Agha, et al., Efficient synthesis of N-acylbenzotriazoles
using tosyl chloride: En route to suberoylanilide hydroxamic
acid (SAHA), ARKIVOC, 2016, 3, 161–170.
32 A. R. Katritzky, et al., Efficient and selective syntheses of
S-acyl and N-acyl glutathiones, Synlett, 2010, 1337–1340.
33 K. A. Agha, et al.,N-Acylbenzotriazole: Convenient approach
for protecting group-free monoacylation of symmetric dia-
mines, Mon. Chem.-Chem. Mon., 2020, 151(4), 589–598.
34 C. A. Brandenburg, et al., Synthesis of deoxyradicinin, an
inhibitor of Xylella fastidiosa and Liberibacter crescens,a
culturable surrogate for Candidatus Liberibacter asiaticus,
J. Nat. Prod., 2020, 83(6), 1810–1816.
35 M. Zampagni, et al., Novel S-acyl glutathione derivatives
prevent amyloid oxidative stress and cholinergic dysfunc-
tion in Alzheimer disease models, Free Radic. Biol. Med.,
2012, 52(8), 1362–1371.
36 A. R. Katritzky, et al., Efficient syntheses of thiadiazole
peptides, J. Org. Chem., 2010, 75(17), 6009–6011.
37 A. R. Katritzky, A. A. Shestopalov and K. Suzuki, A convenient
synthesis of chiral 1, 2, 4-oxadiazoles from N-protected
(a-aminoacyl) benzotriazoles, ARKIVOC, 2005, 7, 36–55.
38 S. Periyasmay and S. Subbiah, Modifications to the process
for the synthesis of new 1, 2, 4-oxadiazole analogues and the
short review on synthetic methodologies of 1, 2, 4-
oxadiazoles, Int. J. Life Sci. Res., 2019, 7(2), 87–104.
39 P. Makhov, et al., Resistance to systemic therapies in clear
cell renal cell carcinoma: Mechanisms and management
strategies, Mol. Cancer Ther., 2018, 17(7), 1355–1364.
40 M. Pasha, et al., Metformin induces different responses in
clear cell renal cell carcinoma caki cell lines, Biomolecules,
2019, 9(3), 113.
41 S. K. Pathak, et al., The effect of regioisomerism on the
mesomorphic and photophysical behavior of oxadiazole-
based tris (N-salicylideneaniline)s: Synthesis and charac-
terization, New J. Chem., 2017, 41(18), 9908–9917.
42 R. H. Shoemaker, The NCI60 human tumour cell line anti-
cancer drug screen, Nat. Rev. Cancer, 2006, 6(10), 813–823.
43 J. F. Diaz and J. M. Andreu, Assembly of purified GDP-tubulin
into microtubules induced by taxol and taxotere: Reversibility,
ligand stoichiometry, and competition, Biochemistry, 1993,
32(11), 2747–2755.
44 Inc., C.C.G., Molecular operating environment (MOE). 2016,
Chemical Computing Group Inc 1010 Sherbooke St. West,
Suite# 910, Montreal.
45 R. F. Taher, et al., Two new flavonoids and anticancer
activity of Hymenosporum flavum:In vitro and molecular
docking studies, J. Herbmed Pharmacol., 2021, 10(4), DOI:
10.34172/jhp.2021.52.
46 A. A. Zaki, et al., Calendulaglycoside a showing potential
activity against SARS-CoV-2 main protease: Molecular
docking, molecular dynamics, and SAR studies, J. Traditon.
Complem. Med., 2021, DOI: 10.1016/j.jtcme.2021.05.001.
47 R. Soltane, et al., Strong inhibitory activity and action
modes of synthetic maslinic acid derivative on highly patho-
genic coronaviruses: COVID-19 drug candidate, Pathogens,
2021, 10(5), 623.
48 H. A. El Gizawy, et al.,Pimenta dioica (L.) Merr. bioactive
constituents exert anti-SARS-CoV-2 and anti-inflammatory
activities: Molecular docking and dynamics, in vitro, and
in vivo studies, Molecules, 2021, 26(19), 5844.
49 J. Yang, et al., The compound millepachine and its deriva-
tives inhibit tubulin polymerization by irreversibly binding
to the colchicine-binding site in b-tubulin, J. Biol. Chem.,
2018, 293(24), 9461–9472.
50 R. M. Samra, et al., Bioassay-guided isolation of a new
cytotoxic ceramide from Cyperus rotundus L, S. Afr. J. Bot.,
2021, 139, 210–216.
51 M. A. Soltan, et al.,In silico prediction of a multitope vaccine
against Moraxella catarrhalis: Reverse vaccinology and
immunoinformatics, Vaccines, 2021, 9(6), 669.
52 A. Abo Elmaaty, et al., Computational insights on the
potential of some NSAIDs for treating COVID-19: Priority
set and lead optimization, Molecules, 2021, 26(12),
3772.
53 X.-Y. Meng, et al., Molecular docking: A powerful approach
for structure-based drug discovery, Curr. Comput.-Aided
Drug Des., 2011, 7(2), 146–157.
54 A. Kandeil, et al., Bioactive polyphenolic compounds showing
strong antiviral activities against severe acute respiratory
syndrome coronavirus 2, Pathogens, 2021, 10(6), 758.
55 A. E. Abdallah, et al., Design and synthesis of new 4-(2-
nitrophenoxy)benzamide derivatives as potential antiviral
agents: Molecular modeling and in vitro antiviral screening,
New J. Chem., 2021, DOI: 10.1039/D1NJ02710G.
56 A. A. Al-Karmalawy, et al., Molecular docking and dynamics
simulation revealed the potential inhibitory activity of ACEIs
against SARS-CoV-2 targeting the hACE2 receptor, Front.
Chem., 2021, 9(227), DOI: 10.3389/fchem.2021.661230.
57 D. Elebeedy, et al., Anti-SARS-CoV-2 activities of tanshinone
IIA, carnosic acid, rosmarinic acid, salvianolic acid, baica-
lein, and glycyrrhetinic acid between computational and
in vitro insights, RSC Adv., 2021, 11(47), 29267–29286.
58 A. Mahmoud, et al., Telaprevir is a potential drug for repur-
posing against SARS-CoV-2: Computational and in vitro studies,
Heliyon, 2021, 7(9), DOI: 10.1016/j.heliyon.2021.e07962.
NJC Paper
Published on 27 October 2021. Downloaded on 11/11/2021 7:14:17 PM.
View Article Online
... Microtubule-binding agents can be categorized into two primary groups: microtubule-stabilizing agents and microtubule-destabilizing agents. The destabilizing agents could impede tubulin polymerization and encourage depolymerization, particularly at elevated concentrations (Diab et al., 2021;Dumontet & Jordan, 2010;Khattab & Al-Karmalawy, 2021). The majority of these agents attach themselves to either the "vinca" domain adjacent to the exchangeable GTP-binding site (E-site) or the "colchicine" domain found at the interface between alpha and beta tubulin within the dimer, as shown in Figure 1 (Dumontet & Jordan, 2010 (Dumontet & Jordan, 2010). ...
... They effectively impede angiogenesis and often surmount multidrug resistance (Al-Karmalawy & Khattab, 2020;Diab et al., 2021). ...
Article
Although various approaches exist for treating cancer, chemotherapy continues to hold a prominent role in the management of this disease. Besides, microtubules serve as a vital component of the cellular skeleton, playing a pivotal role in the process of cell division making it an attractive target for cancer treatment. Hence, the scope of this work was adapted to design and synthesize new anti-tubulin tetrabromophthalimide hybrids (3-17) with colchicine binding site (CBS) inhibitory potential. The conducted in vitro studies showed that compound 16 displayed the lowest IC 50 values (11.46 µM) at the FaDu cancer cell lines, whereas compound 17 exhibited the lowest IC 50 value (13.62 µM) at the PC3 cancer cell line. However, compound 7b exhibited the lowest IC 50 value (11.45 µM) at the MDA-MB-468 cancer cell line. Moreover, compound 17 was observed to be the superior antitumor candidate against all three tested cancer cell lines (MDA-MB-468, PC3, and FaDu) with IC 50 values of 17.22, 13.15, and 13.62 µM, respectively. In addition, compound 17 showed a well-established upregulation of apoptotic markers (Caspases 3, 7, 8, and 9, Bax, and P53). Moreover, compound 17 induced downregulation of the antiapoptotic markers (MMP2, MMP9, and BCL-2). Furthermore, the colchicine binding site inhibition assay showed that compounds 15a and 17 exhibited particularly significant inhibitory potentials, with IC 50 values of 23.07 and 4.25 µM, respectively, compared to colchicine, which had an IC 50 value of 3.89 µM. Additionally, cell cycle analysis was conducted, showing that compound 17 could prompt cell cycle arrest at both the G0-G1 and G2-M phases. On the other hand, a molecular docking approach was applied to investigate the binding interactions of the examined candidates compared to colchicine towards CBS of the β-tubulin subunit. Thus, the synthesized tetrabromophthalimide hybrids can be regarded as outstanding anticancer candidates with significant apoptotic activity.
... The microtubule cytoskeleton has been identified as a promising target for anticancer agents, where numerous studies have been conducted on novel compounds of both natural and synthetic origin to investigate their effectiveness against cancer. Colchicine binding site inhibitors (CBSIs) and other microtubule-targeting drugs are thought to induce G2/M cell cycle arrest and apoptosis providing further evidence for their anticancer activity (Diab et al., 2021;Hammouda et al., 2022;Kumar et al., 2014). ...
... The protein data bank was employed to download the X-ray structure of the CBS of tubulin with PDB entry: 5XIW (Diab et al., 2021). Hence, to be ready for the docking process, the target protein was protonated, corrected, and energetically minimized, as discussed previously in detail (Elebeedy et al., 2021;Mahmoud et al., 2021). ...
Article
Nowadays, conventional anticancer therapy suffers many pitfalls, including drastic side effects and limited therapeutic efficacy resulting from diminished oral bioavailability. So, in an attempt to enhance their poor solubility and oral bioavailability along with the cytotoxic activity, the developed lead compounds (C1 and C2) were loaded in D-α-tocopherol polyethylene glycol 1000 succinate (TPGS) modified vesicles adopting thin film hydration technique. The formulations of the aforementioned candidates (F1 and F2, respectively) were elected as the optimum formula with desirability values of 0.701 and 0.618, respectively. Furthermore, an outstanding enhancement in the drug's cytotoxic activity against different cancer cell lines (MCF-7, HepG-2, MDA-MB-321, A375, and MGC-803) after being included in the nano-TPGS-modified optimum formula was noticed relative to the unformulated compounds. The formula F1 showed the best cytotoxic activities against HepG-2 with an IC50 = 3 µM. Furthermore, regarding MCF-7, F1 was shown to be the most potent and protective among all the tested formulations with an IC50 = 6 µM. Besides, F1 exerted the best caspase 3/7 activity stimulation (around a 5-folds increase) compared to control in the MCF-7 cell line. Notably, it was disclosed that both C1 and C2 induced cell cycle arrest at the resting S growth phase. Moreover, C1 and C2 decreased tubulin concentrations by approximately 2-folds and 6-folds, respectively. Meanwhile, the conducted molecular docking studies ensure the eligible binding affinities of the assessed compounds. Besides, MD simulations were performed for 1000 ns to confirm the docking results and study the exact behavior of the target candidates (C1 and C2) toward the CBS.
... The nine isolated compounds (1-9) from S. tomentosa were sketched using the ChemDraw program. Each compound was introduced individually into the MOE program window and prepared for docking as discussed before [67,68]. Then, the nine prepared isolates (1-9) were imported into two different databases in order to perform two separate docking processes. ...
... The X-ray structures of both the S and M pr• receptors of SARS-CoV-2 (IDs: 7FCD [53] and 6Y2G [54], respectively) were downloaded from the Protein Data Bank (PDB). Each protein was prepared as described earlier in detail [68,69] Each database was uploaded in a separate general docking process according to the previously discussed methodology [70,71]. Moreover, the best pose for each tested compound was selected according to the binding mode, score, and RMSD as well [72,73]. ...
Article
Full-text available
The anti-MERS-CoV activities of three medicinal plants (Azadirachta indica, Artemisia judaica, and Sophora tomentosa) were evaluated. The highest viral inhibition percentage (96%) was recorded for S. tomentosa. Moreover, the mode of action for both S. tomentosa and A. judaica showed 99.5% and 92% inhibition, respectively, with virucidal as the main mode of action. Furthermore, the anti-MERS-CoV and anti-SARS-CoV-2 activities of S. tomentosa were measured. Notably, the anti-SARS-CoV-2 activity of S. tomentosa was very high (100%) and anti-MERS-CoV inhibition was slightly lower (96%). Therefore, the phytochemical investigation of the very promising S. tomentosa L. led to the isolation and structural identification of nine compounds (1–9). Then, both the CC50 and IC50 values for the isolated compounds against SARS-CoV-2 were measured. Compound 4 (genistein 4’-methyl ether) achieved superior anti-SARS-CoV-2 activity with an IC50 value of 2.13 µm. Interestingly, the mode of action of S. tomentosa against SARS-CoV-2 showed that both virucidal and adsorption mechanisms were very effective. Additionally, the IC50 values of S. tomentosa against SARS-CoV-2 and MERS-CoV were found to be 1.01 and 3.11 µg/mL, respectively. In addition, all the isolated compounds were subjected to two separate molecular docking studies against the spike (S) and main protease (Mpr°) receptors of SARS-CoV-2.
... Two different docking processes were performed for the same database prepared previously according to the methodology illustrated earlier 59,60 . One pose for each compound was selected to be further investigated based on the obtained scores and RMSD values 61,62 . ...
Article
Full-text available
Agri-food wastes, produced following industrial food processing, are mostly discarded, leading to environmental hazards and losing the nutritional and medicinal values associated with their bioactive constituents. In this study, we performed a comprehensive analytical and biological evaluation of selected vegetable by-products (potato, onion, and garlic peels). The phytochemical analysis included UHPLC-ESI-qTOF-MS/MS in combination with molecular networking and determination of the total flavonoid and phenolic contents. Further, the antimicrobial, anti-osteoarthritis and wound healing potentials were also evaluated. In total, 47 compounds were identified, belonging to phenolic acids, flavonoids, saponins, and alkaloids as representative chemical classes. Onion peel extract (OPE) showed the higher polyphenolic contents, the promising antioxidant activity, the potential anti-osteoarthritis activity, and promising antimicrobial activity, especially against methicillin-resistant Staphylococcus aureus (MRSA). Furthermore, OPE revealed to have promising in vivo wound healing activity, restoring tissue physiology and integrity, mainly through the activation of AP-1 signaling pathway. Lastly, when OPE was loaded with nanocapsule based hydrogel, the nano-formulation revealed enhanced cellular viability. The affinities of the OPE major metabolites were evaluated against both p65 and ATF-2 targets using two different molecular docking processes revealing quercetin-3,4′-O-diglucoside, alliospiroside C, and alliospiroside D as the most promising entities with superior binding scores. These results demonstrate that vegetable by-products, particularly, those derived from onion peels can be incorporated as natural by-product for future evaluation against wounds and osteoarthritis.
... Each protein was opened using the MOE, corrected, 3D hydrogenated, and energy-minimized, as discussed before. 88 Three general docking processes with the default program specifications 89 were performed by inserting the appropriate database for each target receptor. The superior compounds with the best docking scores and root mean square deviation (rmsd) values 90 were selected. ...
Article
Full-text available
Paracetamol overdose is the leading cause of drug-induced hepatotoxicity worldwide. Because of N-acetyl cysteine's limited therapeutic efficacy and safety, searching for alternative therapeutic substitutes is necessary. This study investigated four citrus juices: Citrus sinensis L. Osbeck var. Pineapple (pineapple sweet orange), Citrus reticulata Blanco × Citrus sinensis L. Osbeck (Murcott mandarin), Citrus paradisi Macfadyen var. Ruby Red (red grapefruit), and Fortunella margarita Swingle (oval kumquat) to improve the herbal therapy against paracetamol-induced liver toxicity. UHPLC-QTOF-MS/MS profiling of the investigated samples resulted in the identification of about 40 metabolites belonging to different phytochemical classes. Phenolic compounds were the most abundant, with the total content ranked from 609.18 to 1093.26 μg gallic acid equivalent (GAE)/mL juice. The multivariate data analysis revealed that phloretin 3′,5′-di-C-glucoside, narirutin, naringin, hesperidin, 2-O-rhamnosyl-swertisin, fortunellin (acacetin-7-O-neohesperidoside), sinensetin, nobiletin, and tangeretin represented the crucial discriminatory metabolites that segregated the analyzed samples. Nevertheless, the antioxidant activity of the samples was 1135.91−2913.92 μM Trolox eq/mL juice, 718.95−3749.47 μM Trolox eq/mL juice, and 2304.74− 4390.32 μM Trolox eq/mL juice, as revealed from 2,2′-azino-bis-3-ethylbenzthiazoline-6-sulfonic acid, ferric-reducing antioxidant power, and oxygen radical absorbance capacity, respectively. The in vivo paracetamol-induced hepatotoxicity model in rats was established and assessed by measuring the levels of hepatic enzymes and antioxidant biomarkers. Interestingly, the concomitant administration of citrus juices with a toxic dose of paracetamol effectively recovered the liver injury, as confirmed by normal sections of hepatocytes. This action could be due to the interactions between the major identified metabolites (hesperidin, hesperetin, phloretin 3′,5′-di-C-glucoside, fortunellin, poncirin, nobiletin, apigenin-6,8-digalactoside, 6′,7′-dihydroxybergamottin, naringenin, and naringin) and cytochrome P450 isoforms (CYP3A4, CYP2E1, and CYP1A2), as revealed from the molecular docking study. The most promising compounds in the three docking processes were hesperidin, fortunellin, poncirin, and naringin. Finally, a desirable food−drug interaction was achieved in our research to overcome paracetamol overdose-induced hepatotoxicity.
... Also, two different validation processes were carried out by redocking the co-crystallized inhibitor of NF-kB and caspase 3 targets within its binding pocket [35][36][37]. The MOE validity was confirmed by obtaining low root mean square deviations (< 2 Å) and similar binding modes to the co-crystallized inhibitor in both cases [38][39][40]. ...
Article
Aims: Oxidative stress and inflammation have been linked to doxorubicin (DOX)-induced cardiotoxicity, while the exact molecular processes are currently under investigation. The goal of this study is to investigate Metformin's preventive role in cardiotoxicity induced by DOX. Materials and methods: Male albino mice were divided randomly into four groups. Metformin (Met) 200 mg/kg orally (p.o.) was given either alone or when combined with a single DOX (15 mg/kg; i.p.). A control group of 5 mice was also provided. Met was initiated7 days before DOX, lasting for 14 days. Besides, docking studies of Met towards HMGB1, NF-kB, and caspase 3 were performed. Key findings: Heart weight, cardiac troponin T (cTnT), creatine kinase Myocardial Band (CK-MB) levels, malondialdehyde (MDA), and nitric oxide (NO) contents all increased significantly when comparing the DOX group to the control normal group. Conversely, there was a substantial decline in superoxide dismutase (SOD) and glutathione peroxidase (GSH). DOX group depicts a high expression of TLR4, HMGB1, and caspase 3. Immunohistochemical staining revealed an increase in NLRP3 inflammasome and NF-κB expressions alongside histopathological modifications. Additionally, Met dramatically decreased cardiac weight, CK-MB, and cTnT while maintaining the tissues' histological integrity. Inflammatory biomarkers, including HMGB1, TLR4, NF-κB, inflammasome, and caspase 3 were reduced after Met therapy. Furthermore, molecular docking studies suggested the antagonistic activity of Met towards HMGB1, NF-κB, and caspase 3 target receptors. Significance: According to recent evidence, Met is a desirable strategy for improving cardiac toxicity produced by DOX by inhibiting the HMGB1/NF-κB inflammatory pathway, thus preserving heart function.
... One of the most important and recent approaches to investigate the activity of a drug is computer-aided drug design (CADD). [39][40][41][42][43][44][45][46][47][48][49] In this study, the interaction between the designed compounds and the binding site was predicted using molecular docking to calculate the binding affinities. 50 Moreover, to confirm the results of molecular docking, molecular dynamics (MD) simulations for 100 ns were carried out on the best-docked inhibitor-protein complexes. ...
... The previously built database was inserted in place of the ligand during the process of general docking using the ligand site as the docking site. All steps of the docking process methodology were followed as described before in detail [63,64]. The best-docked complexes based on their score values, RMSD, and binding modes were selected for further investigations. ...
Article
Full-text available
Alzheimer’s disease remains a global health challenge and an unmet need requiring innovative approaches to discover new drugs. The current study aimed to investigate the inhibitory activity of Albizia lucidior and Albizia procera leaves against acetylcholinesterase enzyme in vitro and explore their chemical compositions. Metabolic profiling of the bioactive plant, A. lucidior, via UHPLC/MS/MS-based Molecular Networking highlighted the richness of its ethanolic extract with budmunchiamine alkaloids, fourteen budmunchiamine alkaloids as well as four new putative ones were tentatively identified for the first time in A. lucidior. Pursuing these alkaloids in the fractions of A. lucidior extract via molecular networking revealed that alkaloids were mainly concentrated in the ethyl acetate fraction. In agreement, the alkaloid-rich fraction showed the most promising anticholinesterase activity (IC50 5.26 µg/mL) versus the ethanolic extract and ethyl acetate fraction of A. lucidior (IC50 24.89 and 6.90 µg/mL, respectively), compared to donepezil (IC50 3.90 µg/mL). Furthermore, deep in silico studies of tentatively identified alkaloids of A. lucidior were performed. Notably, normethyl budmunchiamine K revealed superior stability and receptor binding affinity compared to the two used references: donepezil and the co-crystallized inhibitor (MF2 700). This was concluded based on molecular docking, molecular dynamics simulations and molecular mechanics generalized born/solvent accessibility (MM–GBSA) calculations.
Article
The combination of antibacterial and antiviral agents is becoming a very important aspect of dealing with resistant bacterial and viral infections. The N ‐phenylthiazole scaffold was found to possess significant anti‐MRSA, antifungal, and anti‐COVID‐19 activities as previously published; hence, a slight refinement was proposed to attach various alkyne lipophilic tails to this promising scaffold, to investigate their effects on the antimicrobial activity of the newly synthesized compounds and to provide a valuable structure–activity relationship. Phenylthiazole 4 m exhibited the most potent anti‐MRSA activity with 8 μg/mL MIC value. Compounds 4 k and 4 m demonstrated potent activity against Clostridium difficile with MIC values of 2 μg/mL and moderate activity against Candida albicans with MIC value of 4 μg/mL. When analyzed for their anti‐COVID‐19 inhibitory effect, compound 4 b emerged with IC 50 =1269 nM and the highest selectivity of 138.86 and this was supported by its binding score of −5.21 kcal mol ⁻¹ when docked against SARS‐CoV‐2 M pro . Two H‐bonds were formed, one with His164 and the other with Met49 stabilizing phenylthiazole derivative 4 b , inside the binding pocket. Additionally, it created two arene‐H bonds with Asn142 and Glu166, through the phenylthiazole scaffold and one arene‐H bond with Leu141 via the phenyl ring of the lipophilic tail.
Article
Full-text available
Introduction: Hymenosporum flavum (Hook.) F. Muell. is the sole species within the genus Hymenosporum known for its antimicrobial activity. The current study aims to examine the prospective activity of H. flavum as a safe supporter of sorafenib (as a reference standard) against hepatocellular carcinoma (HCC). Methods: Isolation and identification of compounds were made by chromatographic and spectroscopic methods. A fingerprint for the plant extract was done using HPLC-MS/MS spectrometric analysis. The total plant extract was examined in vitro for HCC activity. The isolated flavonoids were examined for their cytotoxic activities using molecular docking studies against both RAF-1 and ERK-2, and the promising compounds were further examined in vitro using quantitative reverse transcription polymerase chain reaction (qRT-PCR). Results: Two new flavonols were isolated from the leaf extract of H. flavum (Hook.) F. Muell., quercetin-3-O-(glucopyranosyl 1→2 ribopyranoside) (1) and kaempferol-3-O-(glucopyranosyl 1→2 ribopyranoside) (2), accompanying other six known flavonoids (3-8), and identified via spectroscopic analysis. Moreover, HPLC- PDA/MS/MS spectrometric analysis revealed the presence of seventy phenolic metabolites. The cytotoxic activity of the plant extract confirmed its potential action on HepG2 cells indicated by the production level of lactate dehydrogenase (LDH) upon treatment compared with the normal cells. The isolated flavonoids were examined for their cytotoxic activity using molecular docking studies against both RAF-1 and ERK-2 as proposed mechanisms of their anticancer activities. Furthermore, the compounds 1 and 3, which showed the best in silico results, were further examined in vitro using qRT-PCR. They exhibited promising inhibitory activities against both RAF-1 and ERK-2 gene expression. Moreover, they showed promising cytotoxic activities indicated by the MTT assay. Also, both of them improved the efficiency of sorafenib in targeting both RAF-1 and ERK-2 pathways suggesting synergistic combinations. Conclusion: Our findings showed the potential cytotoxic activity of H. flavum extract on HepG2 cells. Some isolated compounds (1 & 3) exhibited promising inhibitory activities against both RAF-1 and ERK-2 gene expression giving a lead future study for these compounds to be used in pharmaceutical preparations either alone or in combination with sorafenib.
Article
Full-text available
In response to the urgent need to control Coronavirus disease 19 (COVID-19), this study aims to explore potential anti-SARS-CoV-2 agents from natural sources. Moreover, cytokine immunological responses to the viral infection could lead to acute respiratory distress which is considered a critical and life-threatening complication associated with the infection. Therefore, the anti-viral and anti-inflammatory agents can be key to the management of patients with COVID-19. Four bioactive compounds, namely ferulic acid 1, rutin 2, gallic acid 3, and chlorogenic acid 4 were isolated from the leaves of Pimenta dioica (L.) Merr (ethyl acetate extract) and identified using spectroscopic evidence. Furthermore, molecular docking and dynamics simulations were performed for the isolated and identified compounds (1–4) against SARS-CoV-2 main protease (Mpro) as a proposed mechanism of action. Furthermore, all compounds were tested for their half-maximal cytotoxicity (CC50) and SARS-CoV-2 inhibitory concentrations (IC50). Additionally, lung toxicity was induced in rats by mercuric chloride and the effects of treatment with P. dioca aqueous extract, ferulic acid 1, rutin 2, gallic acid 3, and chlorogenic acid 4 were recorded through measuring TNF-α, IL-1β, IL-2, IL-10, G-CSF, and genetic expression of miRNA 21-3P and miRNA-155 levels to assess their anti-inflammatory effects essential for COVID-19 patients. Interestingly, rutin 2, gallic acid 3, and chlorogenic acid 4 showed remarkable anti-SARS-CoV-2 activities with IC50 values of 31 µg/mL, 108 μg/mL, and 360 µg/mL, respectively. Moreover, the anti-inflammatory effects were found to be better in ferulic acid 1 and rutin 2 treatments. Our results could be promising for more advanced preclinical and clinical studies especially on rutin 2 either alone or in combination with other isolates for COVID-19 management.
Article
Full-text available
Drug repurposing is an important approach to the assignment of already approved drugs for new indications. This technique bypasses some steps in the traditional drug approval system, which saves time and lives in the case of pandemics. Direct acting antivirals (DAAs) have repeatedly repurposed from treating one virus to another. In this study, 16 FDA-approved hepatitis C virus (HCV) DAA drugs were studied to explore their activities against severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) human and viral targets. Among the 16 HCV DAA drugs, telaprevir has shown the best in silico evidence to work on both indirect human targets (cathepsin L [CTSL] and human angiotensin-converting enzyme 2 [hACE2] receptor) and direct viral targets (main protease [Mpro]). Moreover, the docked poses of telaprevir inside both hACE2 and Mpro were subjected to additional molecular dynamics simulations monitored by calculating the binding free energy using MM-GBSA. In vitro analysis of telaprevir showed inhibition of SARS-CoV-2 replication in cell culture (IC50 = 11.552 μM, CC50 = 60.865 μM, and selectivity index = 5.27). Accordingly, based on the in silico studies and supported by the presented in vitro analysis, we suggest that telaprevir may be considered for therapeutic development against SARS-CoV-2.
Article
Full-text available
Moraxella catarrhalis (M. catarrhalis) is a Gram-negative bacterium that can cause serious respiratory tract infections and middle ear infections in children and adults. M. catarrhalis has demonstrated an increasing rate of antibiotic resistance in the last few years, thus development of an effective vaccine is a major health priority. We report here a novel designed multitope vaccine based on the mapped epitopes of the vaccine candidates filtered out of the whole proteome of M. catarrhalis. After analysis of 1615 proteins using a reverse vaccinology approach, only two proteins (outer membrane protein assembly factor BamA and LPS assembly protein LptD) were nominated as potential vaccine candidates. These proteins were found to be essential, outer membrane, virulent and non-human homologs with appropriate molecular weight and high antigenicity score. For each protein, cytotoxic T lymphocyte (CTL), helper T lymphocyte (HTL) and B cell lymphocyte (BCL) epitopes were predicted and confirmed to be highly antigenic and cover conserved regions of the proteins. The mapped epitopes constituted the base of the designed multitope vaccine where suitable linkers were added to conjugate them. Additionally, beta defensin adjuvant and pan-HLA DR-binding epitope (PADRE) peptide were also incorporated into the construct to improve the stimulated immune response. The constructed multitope vaccine was analyzed for its physicochemical, structural and immunological characteristics and it was found to be antigenic, soluble, stable, non-allergenic and have a high affinity to its target receptor. Although the in silico analysis of the current study revealed that the designed multitope vaccine has the ability to trigger a specific immune response against M. catarrhalis, additional translational research is required to confirm the effectiveness of the designed vaccine. View Full-Text Keywords: Moraxella catarrhalis; vaccinomics; reverse vaccinology; immunoinformatics; epitope mapping; multitope vaccine
Article
Full-text available
Six compounds namely, tanshinone IIA (1), carnosic acid (2), rosmarinic acid (3), salvianolic acid B (4), baicalein (5), and glycyrrhetinic acid (6) were screened for their anti-SARS-CoV-2 activities against both the spike (S) and main protease (Mpro) receptors using molecular docking studies. Molecular docking recommended the superior affinities of both salvianolic acid B (4) and glycyrrhetinic acid (6) as the common results from the previously published computational articles. On the other hand, their actual anti-SARS-CoV-2 activities were tested in vitro using plaque reduction assay to calculate their IC 50 values after measuring their CC 50 values using MTT assay on Vero E6 cells. Surprisingly, tanshinone IIA (1) was the most promising member with IC 50 equals 4.08 ng ml À1. Also, both carnosic acid (2) and rosmarinic acid (3) showed promising IC 50 values of 15.37 and 25.47 ng ml À1 , respectively. However, salvianolic acid (4) showed a weak anti-SARS-CoV-2 activity with an IC 50 value equals 58.29 ng ml À1. Furthermore, molecular dynamics simulations for 100 ns were performed for the most active compound from the computational point of view (salvianolic acid 4), besides, the most active one biologically (tanshinone IIA 1) on both the S and Mpro complexes of them (four different molecular dynamics processes) to confirm the docking results and give more insights regarding the stability of both compounds inside the SARS-CoV-2 mentioned receptors, respectively. Also, to understand the mechanism of action for the tested compounds towards SARS-CoV-2 inhibition it was necessary to examine the mode of action for the most two promising compounds, tanshinone IIA (1) and carnosic acid (2). Both compounds (1 and 2) showed very promising virucidal activity with a most prominent inhibitory effect on viral adsorption rather than its replication. This recommended the predicted activity of the two compounds against the S protein of SARS-CoV-2 rather than its Mpro protein. Our results could be very promising to rearrange the previously mentioned compounds based on their actual inhibitory activities towards SARS-CoV-2 and to search for the reasons behind the great differences between their in silico and in vitro results against SARS-CoV-2. Finally, we recommend further advanced preclinical and clinical studies especially for tanshinone IIA (1) to be rapidly applied in COVID-19 management either alone or in combination with carnosic acid (2), rosmarinic acid (3), and/or salvianolic acid (4).
Article
Full-text available
Targeting the epidermal growth factor receptors (EGFRs) with small inhibitor molecules has been validated as a potential therapeutic strategy in cancer therapy. Pyrazolo[3,4-d]pyrimidine is a versatile scaffold that has been exploited for developing potential anticancer agents. On the basis of fragment-based drug discovery, considering the essential pharmacophoric features of potent EGFR tyrosine kinase (TK) inhibitors, herein, we report the design and synthesis of new hybrid molecules of the pyrazolo[3,4-d]pyrimidine scaffold linked with diverse pharmacophoric fragments with reported anticancer potential. These fragments include hydrazone, indoline-2-one, phthalimide, thiourea, oxadiazole, pyrazole, and dihydropyrazole. The synthesized molecules were evaluated for their anticancer activity against the human breast cancer cell line, MCF-7. The obtained results revealed comparable antitumor activity with that of the reference drugs doxorubicin and toceranib. Docking studies were performed along with EGFR-TK and ADMET profiling studies. The results of the docking studies showed the ability of the designed compounds to interact with key residues of the EGFR-TK through a number of covalent and noncovalent interactions. The obtained activity of compound 25 (IC 50 = 2.89 µM) suggested that it may serve as a lead for further optimization and drug development. K E Y W O R D S antitumor, docking, EGFR, pyrazolo[3,4-d]pyrimidine, SAR, Schiff bases
Article
Full-text available
Regarding the crucial role of deubiquitinase (DUB) enzymes in many viruses, in particular, Adenovirus, HSV-1, coxsackievirus, and SARS-CoV-2, DUB inhibition was reported as an effective new approach to find new effective antiviral agents. In the present study, a new wave of 4-(2-nitrophenoxy)benzamide derivatives was designed and synthesized to fulfill the basic pharmacophoric features of DUB inhibitors. The molecular docking of the designed compounds against deubiquitinase enzymes of the aforementioned viruses was carried out. Significant molecular docking results directed us to conduct in vitro antiviral screening against the aforementioned viruses. The biological data showed very strong to strong antiviral activities with IC 50 values ranging from 10.22 to 44.68 mM against Adenovirus, HSV-1, and coxsackievirus. Compounds 8 c , 8 d , 10 b , and 8 a were found to be the most potent against Adenovirus, HSV-1, coxsackievirus, and SAR-CoV-2, respectively. Also, the CC 50 values of the examined compounds ranged from 72.93 to 120.50 mM. Finally, the in silico ADMET and toxicity studies demonstrated that the tested members have a good profile of drug-like properties. Furthermore, we concluded the structure-activity relationship (SAR) of the newly designed and synthesized compounds regarding their in vitro results, which may help medicinal chemists in further optimization to obtain more potential antiviral candidates in the near future as well.
Article
Full-text available
The discovery of drugs capable of inhibiting SARS-CoV-2 is a priority for human beings due to the severity of the global health pandemic caused by COVID-19. To this end, repurposing of FDA-approved drugs such as NSAIDs against COVID-19 can provide therapeutic alternatives that could be utilized as an effective safe treatment for COVID-19. The anti-inflammatory activity of NSAIDs is also advantageous in the treatment of COVID-19, as it was found that SARS-CoV-2 is responsible for provoking inflammatory cytokine storms resulting in lung damage. In this study, 40 FDA-approved NSAIDs were evaluated through molecular docking against the main protease of SARS-CoV-2. Among the tested compounds, sulfinpyrazone 2, indomethacin 3, and auranofin 4 were proposed as potential antagonists of COVID-19 main protease. Molecular dynamics simulations were also carried out for the most promising members of the screened NSAID candidates (2, 3, and 4) to unravel the dynamic properties of NSAIDs at the target receptor. The conducted quantum mechanical study revealed that the hybrid functional B3PW91 provides a good description of the spatial parameters of auranofin 4. Interestingly, a promising structure-activity relationship (SAR) was concluded from our study that could help in the future design of potential SARS-CoV-2 main protease inhibitors with expected anti-inflammatory effects as well. NSAIDs may be used by medicinal chemists as lead compounds for the development of potent SARS-CoV-2 (M pro) inhibitors. In addition, some NSAIDs can be selectively designated for treatment of inflammation resulting from COVID-19.
Article
Full-text available
Moraxella catarrhalis (M. catarrhalis) is a Gram-negative bacterium that can cause serious respiratory tract infections and middle ear infections in children and adults. M. catarrhalis has demonstrated an increasing rate of antibiotic resistance in the last few years, thus development of an effective vaccine is a major health priority. We report here a novel designed multitope vaccine based on the mapped epitopes of the vaccine candidates filtered out of the whole proteome of M. catarrhalis. After analysis of 1615 proteins using a reverse vaccinology approach, only two proteins (outer membrane protein assembly factor BamA and LPS assembly protein LptD) were nominated as potential vaccine candidates. These proteins were found to be essential, outer membrane, virulent and non-human homologs with appropriate molecular weight and high antigenicity score. For each protein, cytotoxic T lymphocyte (CTL), helper T lymphocyte (HTL) and B cell lymphocyte (BCL) epitopes were predicted and confirmed to be highly antigenic and cover conserved regions of the proteins. The mapped epitopes constituted the base of the designed multitope vaccine where suitable linkers were added to conjugate them. Additionally, beta defensin adjuvant and pan-HLA DR-binding epitope (PADRE) peptide were also incorporated into the construct to improve the stimulated immune response. The constructed multitope vaccine was analyzed for its physicochemical, structural and immunological characteristics and it was found to be antigenic, soluble, stable, non-allergenic and have a high affinity to its target receptor. Although the in silico analysis of the current study revealed that the designed multitope vaccine has the ability to trigger a specific immune response against M. catarrhalis, additional translational research is required to confirm the effectiveness of the designed vaccine.
Article
Background: Although some benzimidazole-based anthelmintic drugs are found to possess anticancer activity, their modes of binding interactions have not been reported. Methodology: Therefore, in this study, we aimed to investigate the binding interactions and electronic configurations of nine benzimidazole-based anthelmintics against one of the well-known cancer targets (tubulin protein). Results: Binding affinities of docked benzimidazole drugs into colchicine-binding site were calculated where flubendazole > oxfendazole > nocodazole > mebendazole. Flubendazole was found to bind more efficiently with tubulin protein than other drugs. Quantum mechanics studies revealed that the electron density of HOMO of flubendazole and mebendazole together with their molecular electrostatic potential map are closely similar to that of nocodazole. Conclusion: Our study has ramifications for considering repurposing of flubendazole as a promising anticancer candidate.