Invest Clin 63(3): 243 - 261, 2022 https://doi.org/10.54817/IC.v63n3a04
Corresponding author: Gustavo Benaim . Laboratorio de Señalización Celular y Bioquímica de Parásitos, Instituto
de Estudios Avanzados (IDEA) and Instituto de Biología Experimental, Universidad Central de Venezuela (UCV),
Caracas, Venezuela. Tel: +584143060359; E-mail: gbenaim@gmail.com
Effect of tetrahydroquinoline derivatives
on the intracellular Ca2+ homeostasis
in breast cancer cells (MCF-7) and its
relationship with apoptosis.
Semer Maksoud1, Adriana Mayora2, Laura Colman3, Felipe Sojo4,5, Adriana A. Pimentel6,
Vladimir V. Kouznetsov7, Diego R. Merchán-Arenas7, Ángel H. Romero8,
Francisco Arvelo4,5, Juan Bautista De Sanctis9 and Gustavo Benaim10,11
1Department of Neurology and Experimental Therapeutics and Molecular Imaging
Laboratory, Massachusetts General Hospital, MA 02129, USA.
2Instituto de Medicina Experimental. Facultad de Medicina. Universidad Central de
Venezuela.
3Instituto Pasteur, Montevideo, Uruguay.
4Centro de Biociencias, Fundación Instituto de Estudios Avanzados (IDEA). Caracas,
Venezuela.
5Laboratorio de Cultivo de Tejidos y Biología de Tumores, Instituto de Biología
Experimental, Universidad Central de Venezuela (UCV), Caracas, Venezuela.
6Facultad de Farmacia, Universidad Central de Venezuela. Caracas, Venezuela.
7Laboratorio de Química Orgánica y Biomolecular (CMN), Universidad Industrial de
Santander. Parque Tecnológico Guatiguara, Piedecuesta. Colombia.
8Cátedra de Química General, Facultad de Farmacia, Universidad Central de Venezuela.
Caracas, Venezuela.
9Institute of Molecular and Translational Medicine. Palacky Unversity.
Olomouc. Czech Republic.
10Laboratorio de Señalización Celular y Bioquímica de Parásitos. Instituto de Estudios
Avanzados (IDEA). Caracas, Venezuela.
11Instituto de Biología Experimental, Universidad Central de Venezuela (UCV). Caracas.
Venezuela.
Key words: Ca2+; apoptosis; tetrahydroquinolines; mitocondria; SERCA; NF-κB.
Abstract. Tetrahydroquinoline derivatives are interesting structures exhib-
iting a wide range of biological activities, including antitumor effects. In this
investigation, the effect of the synthesized tetrahydroquinolines JS-56 and JS-92
on apoptosis, intracellular Ca2+ concentration ([Ca2+]i), and the sarco(endo)plas-
mic reticulum Ca2+-ATPase (SERCA) activity was determined on MCF-7 breast
cancer cells. Colorimetric assays were used to assess MCF-7 cells viability and
SERCA activity. Fura-2 and rhodamine 123 were used to measure the intracellu-
lar Ca2+ concentration and the mitochondrial electrochemical potential, respec-
244 Maksoud et al.
Investigación Clínica 63(3): 2022
Efecto de derivados de tetrahidroquinolinas sobre la
homeostasis del Ca2+ intracelular en células de cáncer de mama
(MCF-7) y su relación con la apoptosis.
Invest Clin 2022; 63 (3): 243 – 261
Palabras clave: Ca2+, apoptosis, tetrahidroquinolinas, Mitocondria, SERCA, NF-κB.
Resumen. Los derivados de tetrahidroquinolina son estructuras interesan-
tes que exhiben una amplia gama de actividades biológicas, incluyendo efectos
antitumorales. Se determinó el efecto de las tetrahidroquinolinas sintetizadas
JS-56 y JS-92 sobre la apoptosis, concentración intracelular de Ca2+ ([Ca2+]i) y
la actividad Ca2+-ATPasa del retículo sarco(endo)plásmico (SERCA) en células de
cáncer de mama MCF-7. Se usaron ensayos colorimétricos para evaluar la viabili-
dad de las células MCF-7 y la actividad SERCA. Se emplearon Fura-2 y rodamina
123 para medir la concentración de Ca2+ intracelular y el potencial electroquí-
mico mitocondrial, respectivamente. El ensayo TUNEL se utilizó para analizar la
fragmentación del ADN, mientras que la actividad de caspasas y la expresión géni-
ca dependiente de NF-κB se evaluaron mediante luminiscencia. Modelos in silico
permitieron el análisis del acoplamiento molecular. Estos compuestos aumentan
la concentración de Ca2+ intracelular; la principal contribución es la entrada de
Ca2+ desde el medio extracelular. Tanto JS-56 como JS-92 inhiben la actividad de
SERCA y disipan el potencial electroquímico mitocondrial a través de procesos
dependientes e independientes de la captación de Ca2+ por este orgánulo. Ade-
más, JS-56 y JS-92 generan citotoxicidad en células MCF-7. El efecto de JS-92 es
mayor que JS-56. Ambos compuestos activan las caspasas 7 y 9, provocan la frag-
mentación del ADN y potencian el efecto del 12-miristato-13-acetato de forbol en
la expresión génica dependiente de NF-κB. El análisis de acoplamiento molecular
sugiere que ambos compuestos tienen una alta interacción con SERCA, similar
a la tapsigargina. Ambos derivados de tetrahidroquinolina indujeron la muerte
celular a través de una combinación de eventos apoptóticos, aumento de [Ca2+]i
e inhibición de la actividad SERCA por interacción directa.
Received: 06-04-2022 Accepted: 10-06-2022
tively. TUNEL assay was used to analyze DNA fragmentation, while caspase activi-
ty and NF-κB-dependent gene expression were assessed by luminescence. In silico
models were used for molecular docking analysis. These compounds increase
intracellular Ca2+ concentration; the main contribution is the Ca2+ entry from
the extracellular milieu. Both JS-56 and JS-92 inhibit the activity of SERCA and
dissipate the mitochondrial electrochemical potential through processes depen-
dent and independent of the Ca2+ uptake by this organelle. Furthermore, JS-56
and JS-92 generate cytotoxicity in MCF-7 cells. The effect of JS-92 is higher than
JS-56. Both compounds activate caspases 7 and 9, cause DNA fragmentation,
and potentiate the effect of phorbol 12-myristate-13-acetate on NF-κB-dependent
gene expression. Molecular docking analysis suggests that both compounds have
a high interaction for SERCA, similar to thapsigargin. Both tetrahydroquinoline
derivatives induced cell death through a combination of apoptotic events, in-
crease [Ca2+]i, and inhibit SERCA activity by direct interaction.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 245
Vol. 63(3): 243 - 261, 2022
INTRODUCTION
Breast cancer is the most frequently
diagnosed cancer and the leading cause of
cancer-related death in women worldwide,
with an estimated 2.3 million new cases per
year and approximately more than 685,000
deaths by 2020 1. Among treatments against
this disease is the use of tamoxifen (Nolva-
dex®) for cancer cells expressing estrogen
receptors or trastuzumab (Herceptin®) for
HER-2+ mammary tumors 2. Of particular in-
terest, numerous investigations evaluate the
potential of natural or synthetic compounds
that stimulate apoptosis in breast cancer
cells by disturbing intracellular Ca2+ homeo-
stasis 3-9.
Ca2+ possesses an essential regulatory
role in differentiation, secretion, contrac-
tion, transcription, phosphorylation, and
apoptosis processes. A piece of large cell ma-
chinery composed of different proteins and
organelles contributes to the regulation of
intracellular Ca2+ concentration ([Ca2+]i) 10.
Among them is the sarco(endo)plasmic re-
ticulum Ca2+-ATPase (SERCA), which allows
active transport and a considerable accu-
mulation of this cation at the endoplasmic
reticulum (ER). Active transport maintains
[Ca2+] in the ER. This concentration is ap-
proximately three orders of magnitude high-
er (millimolar range) than cytoplasm [Ca2+]
(~100 nanomolar), and similar to the ex-
tracellular milieu 11. Apoptosis is directly
related to Ca2+ transfer from ER to mito-
chondria. Ca2+ depletion in ER can generate
a phenomenon called “ER stress” in which
the reduction of protein folding capacity,
and consequent accumulation of these mis-
folded proteins, initiates different processes
promoting cell death 12,13. The electropho-
retic uniporter (MCU) induces Calcium ac-
cumulation in the mitochondria. This [Ca2+]
increase generates an important change in
electrochemical potential (Δψm) 14 which
may enable membrane permeabilization
with consequent release of cytochrome c,
caspase-9 activation and cell death 15.
In this investigation, the effect of tetra-
hydroquinolines JS-56 and JS-92 (Fig. 1) was
evaluated on Ca2+ homeostasis and apopto-
sis induction in breast cancer cells MCF-7.
The tested compounds were prepared from
inexpensive and commercially available isat-
in, 2-aminobenzonitrile (or 4-bromoaniline)
and trans-isoeugenol, a major constituent of
clove oil using a previously published two-
step procedure 16,17. These tetrahydroquino-
lines have been shown to exert antitumor
activity against various cancer cell lines, in-
cluding MCF-7, SKBR3, PC3 and Hela cells
16,17; the aim of the study is to ascertain the
mechanism of apoptosis induction by the
compounds. Both compounds induced an in-
Fig. 1. Chemical structure of the tetrahydroquinolines JS-56 and JS-92.
246 Maksoud et al.
Investigación Clínica 63(3): 2022
crease in [Ca2+]i caused in part through SER-
CA inhibition, combined with the dissipation
of mitochondrial Δψm. Molecular docking
analysis showed that these compounds have
similar ligand-protein interactions with SER-
CA when compared to thapsigargin (Tg).
Furthermore, these tetrahydroquinolines
generated apoptosis via caspase activation
and DNA fragmentation. Both compounds
enhanced the effect of phorbol 12-myristate-
13-acetate (PMA) on the NF-κB-dependent
gene expression.
MATERIALS AND METHODS
Chemicals
The preparation of 8’-ciano-4’-(4-hydroxy-
3-methoxyphenyl)-3’-methyl-3’,4’-dihydro-1’H-
spiro[indoline-3,2’-quinolin]-2-one (JS-56) and
6’-bromo-4’-(4-hydroxy-3-methoxyphenyl)-3’-
methyl-3’,4’-dihydro-1’H-spiro[indoline-3,2’-
quinolin]-2-one (JS-92) was performed by fol-
lowing our previous reports which include the
reaction of isatin and 2-aminobenzonitrile or
4-bromoaniline to give the respective ketimine-
isatin derivatives and the subsequent acid-
catalyzed cycloaddition reaction with trans-
isoeugenol 16,17. Both compounds were purified
by silica gel column chromatography and ob-
tained as stable solids. Their structures were
confirmed by various spectroscopic techniques
such as EI-MS, 1H NMR, 13C NMR, and IR, test-
ing each sample as a pure chemical substance
in the following biological experiments.
Cell culture
The human breast cancer cell line MCF-
7 was grown in Dulbecco’s Modified Eagle
Medium (DMEM), supplemented with peni-
cillin (100 U/mL), streptomycin (10,000
μg/mL), 1% GlutaMAX™ and 10% fetal bo-
vine serum (FBS). These cells were kept in
a humidified incubator at 37°C and in an at-
mosphere of 5% CO2.
Cytotoxicity assays
To evaluate the cytotoxicity in MCF-
7 cells, the colorimetric reduction assay of
3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltet-
razolium bromide (MTT) was performed 18.
Ten thousand cells were seeded in 96-well
plates for 48 h allowing confluence. Cells
were exposed to JS-56 and JS-92 for 24 h,
at concentrations ranging from 1 to 50 μM.
In all cases, both tetrahydroquinolines were
dissolved in dimethyl sulfoxide (DMSO). The
final concentration of this solvent was equal
to or less than 0.5%.
All appropriate controls were developed
with 0.5% DMSO. After treatment, cells were
incubated with 0.4 mg/mL of MTT for 2 h at
37°C. Then, the supernatant was removed, and
the formazan crystals were dissolved by adding
50 μL of DMSO. Absorbance was measured in a
multi-detection microplate reader (Sinergy HT,
Bio-Tek) at 570 nm. The IC50 value was defined
as the tetrahydroquinoline concentration caus-
ing a 50% reduction in absorbance compared
to the control of untreated cells, using the
Graphpad Prism 5.0 software.
Measurement of intracellular Ca2+
concentration
Measurements of [Ca2+]i in MCF-7 cells
were made using the fluorescent indicator
Fura-2, essentially as previously described 19.
Briefly, a suspension of 1x106 cells/mL was
prepared in a medium containing 138 mM
NaCl, 5 mM KCl, 1 mM MgCl2, 1.5 mM CaCl2,
5 mM glucose, 10 mM HEPES and 0.1% al-
bumin (extracellular medium), loading cells
with 200 μM sulfinpyrazone, 0.025% plu-
ronic acid, 0.1% albumin, and 2 μM Fura 2
acetoxymethyl ester (Fura 2-AM) for 30 min
at 25°C, in darkness and constant agitation.
After loading, cells were incubated for 15
minutes in the extracellular medium and
washed again using the same solution. When
indicated, these experiments were also
performed using the same wash buffer but
without CaCl2 (by addition of 2 mM EGTA).
Fura-2 fluorescence on loaded cells suspen-
sion was monitored through a Perkin Elmer
LS 55 spectrofluorimeter with a chopper
device. The excitation wavelengths were set
at 340 and 380 nm, while the emission was
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 247
Vol. 63(3): 243 - 261, 2022
measured at 510 nm. All experiments were
carried out at 30°C, with constant agitation
in a stirred cuvette.
Intracellular Ca2+ was calculated ac-
cording to Grynkiewicz et al. 20, using the
following equation: [Ca2+]I = Kd x (R-Rmin/
Rmax-R) x Fmin(380)/Fmax(380) where Kd is the dis-
sociation constant of Fura-2 (224 nM). R is
the ratio of emission fluorescence intensities
to excitation lengths of 340/380 nm. Rmin
and Rmax are ratios at 0 Ca2+ (6 μM digitonin
plus 4 mM EGTA) and saturating Ca2+ (2 mM
CaCl2), respectively. Fmin(380)/Fmax(380) are fluo-
rescent values of cells exposed to digitonin
plus 4 mM EGTA and 2 mM CaCl2, represent-
ing minimal and maximal fluorescent values.
Purification of Ca2+-ATPase from the
sarcoplasmic reticulum
The sarcoplasmic reticulum was ob-
tained from the white skeletal muscle of
the hind legs of rabbits, according to the
method described by Eletr and Inesi 21. The
vesicles derived from the sarcoplasmic retic-
ulum are highly enriched in Ca2+-ATPase (ap-
proximately 90%). To avoid Ca2+ retention
in the lumen of the vesicles, the ionophore
A-23187 (1 μM) was added.
Evaluation of SERCA activity
The purified enzyme (2 μg/mL) was in-
cubated for 45 min at 37°C and continuous
agitation in a final volume of 250 μL of a buf-
fer containing 100 mM KCl, 20 mM MOPS-
Tris (pH 7.0), 1 mM EGTA, 10 mM MgCl2,
4 mM ATP, 1 μM A23817 and 1 mM CaCl2,
obtaining 10 μM of Ca2+ in the medium and
reaching an optimal activity of the enzyme.
Inorganic phosphate production was quanti-
fied using the Fiske and Subbarow 22 colo-
rimetric assay, modified by applying ferrous
sulfate as a reducing agent 23.
Determination of the mitochondrial
electrochemical potential (∆ψm)
Mitochondrial membrane potential
changes were measured using the fluores-
cent marker rhodamine 123, as previously
described 24, with the modifications intro-
duced by Pimentel et al. 6. Rhodamine 123
is a cationic fluorescent dye that allows the
evaluation of the electrochemical potential of
this organelle when distributed according to
the Δψm. Rhodamine has a maximum excita-
tion peak of 488 nm and a maximum emis-
sion peak of 530 nm. For these experiments
3 x 105 cells/mL were resuspended in a me-
dium containing 138 mM NaCl, 5 mM KCl,
1 mM MgCl2, 1.5 mM CaCl2, 5 mM glucose,
10 mM HEPES-KOH, pH 7.4 and then loaded
with rhodamine 123 (20 μg/mL) for 45 min
at 30°C. All measurements were made on a
HITACHI-L-7000 spectrofluorimeter at 30˚C
with continuous stirring.
Determination of caspases 7 and 9 activity
We used Caspase-Glo 3/7 and Caspase-
Glo 9 kits (Promega, Madison, WI) and de-
veloped the respective tests following the
manufacturer’s instructions. Since MCF-7
cells do not possess caspase 3 25, the results
shown correspond to the effect of the dif-
ferent compounds on caspase 7 activity. In
a 96-well plate, 10,000 cells/well were seed-
ed and incubated for 48 h at 37°C in a 5%
CO2 atmosphere. Then, cells were exposed
to different treatments for 24 h. Then, the
cells were lysed, and the luminescence was
measured with a multidetector microplate
reader (Synergy HT, Bio-Tek). Staurosporine
(Stau, 1 μM) was applied as an apoptosis ac-
tivator (positive control), and 0.5% DMSO
was used as a negative control.
Determination of DNA fragmentation
DNA fragmentation was analyzed through
the “dead-end fluorimetric TUNEL (TdT-medi-
ated dUTP Nick-End labeling) system” (Prome-
ga) kit. Cells were double-labeled with Fluores-
cein-12-dUTP (a specific marker of fragmented
DNA) and propidium iodide (DNA marker). Ac-
cording to manufacturing instructions, chang-
es in fluorescence intensity were detected by
flow cytometry (Epics XL Beckman Coulter,
FL). Cells incubated with staurosporine (1 μM)
were used as a positive control.
248 Maksoud et al.
Investigación Clínica 63(3): 2022
Analysis of NF-κB-dependent gene expression
In this assay, HeLa tumor cells trans-
fected with a luciferase reporter gene whose
expression is under control of a modified
IL-6 promoter, which only contains a critical
binding site for NF-κB, were used. The lucif-
erase activity was determined in a luminom-
eter. Five thousand cells/well were seeded in
96-well plates and incubated for 48 h. Sub-
sequently, the cells were exposed to all treat-
ments and corresponding controls for 24 h.
Then, the cells were lysed, the luciferase sub-
strate was added, and the product was quan-
tified by luminescence. Incubation with cul-
ture medium represents the negative control
while phorbol 12-myristate 13-acetate (100
nM) was used as a positive control.
Molecular docking procedure
A 3D crystal structure from SERCA in
complex with BHQ (2,5-ditert-butylbenzene-1,4-
diol) and Tg was downloaded from Protein
Data Bank under PDB code 2AGV. The pro-
tein corresponds to sarco(endo)plasmic re-
ticulum Ca2+-ATPase from Oryctolagus cu-
niculus species, and it was crystallized with a
resolution of 2.4 Å 26. The protein presented
six co-crystallized chemical ligands: two mol-
ecules of BHQ, two molecules of Tg, and two
sodium atoms. Four active sites (two for BHQ
and the other two for Tg) were generated on
the protein: two on α-chain and the other two
on β-chain. The 3D structure was imported
into the Swiss-PdbViewer v4.0.1 software 27.
Hydrogen atoms were added to the protein
structure, and minimization calculations
were carried out on protein structure via AM-
BER force field, using the conjugate gradient
method with an RMA gradient of 0.01 kcal/
Åmol on Swiss-PdbViewer v4.0.1 software.
The prepared protein was exported to an Ar-
gusLab v.4.0.1 program package and saved
as an Agl document. Four active sites were
constructed on the SERCA protein derived
from BHQ and Tg inhibitors. It is well docu-
mented that both compounds, BHQ and Tg,
act as SERCA inhibitors. Specifically, BHQ is
an L-type Ca2+ channel inhibitor, blocks Ca2+
transport by inducing superoxide anion pro-
duction 28. On the contrary, Tg inhibits all
SERCA isozymes at similar concentrations29.
Both inhibitors are bound to the transmem-
brane domain of the enzyme close to the
membrane/cytosolic interface.
Once the protein was prepared, the two
tested tetrahydroquinolines JS56 and JS92
were built using ArgusLab v4.0 30, optimiz-
ing their respective geometries employing
B3LYP/6-31G(d,p) 31. Then, molecular dock-
ing of the two compounds and crystallized
inhibitors (BHQ and TG) over the four active
SERCA sites was performed employing the
ArgusLab (v4.0.1) package program under
Windows 7.0 environment, applying AMBER
force field. The docking was achieved through
their respective grid map dimensions and a
grid point spacing of 0.375 Å. A flexible li-
gand model was used in the docking and
subsequent optimization scheme. Different
orientations of the ligands were scanned and
ranked from their energy scores. Reproduc-
ibility of the calculated affinity and the mini-
mum energy pose was evaluated through 10
replicates for each quinoline and crystallized
inhibitor (BHQ or TS) 32. The ChemScore
scoring function was selected for the evalu-
ation of ligand binding modes. Settings for
genetic algorithm runs were kept at their de-
fault values; that is, the population size was
100, the selection pressure 1.1, and the num-
ber of operations was 100,000. Affinity energy
is reported as the mean of the 10 replicates.
The lowest energy poses (strongest-docking)
for each ligand in each target protein are
summarized in Table 1.
RESULTS
Effect of tetrahydroquinolines JS-56 and
JS-92 on cytotoxicity in MCF-7 cells
First, the possible cytotoxicity generated
by JS-56 and JS-92 on MCF-7 breast cancer
cells was evaluated through an MTT assay af-
ter being treated for 24 h with different con-
centrations of these tetrahydroquinolines. It
can be observed (Fig. 2) that both compounds
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 249
Vol. 63(3): 243 - 261, 2022
Table 1
Binding energies (kcal/mol), H-interactions, and hydrophobic interactions from molecular
docking on the four studied active sites.
Binding
site
Molecule Binding energy
(kcal/mol)
H-interactions (Å) Intermolecular interactions
TG1 JS-56 -13.1366 Two hydrogen bonding
(2.8596 and 2.1804 Å)
between the phenolic
hydrogen of the ligand
and Ile-829 residue.
a) H-π interaction with
Phe-834 and Phe-256.
b) Hydrophobic interactions
with Met-838, Tyr-837, Phe-834,
Ile-829, Leu-828, Pro-827,
Leu-828, Ile-765, Gln-259,
Leu-260 and Val-263.
JS-92 -12.5189 Hydrogen bonding
(2.8988 Å) between the
phenolic hydrogen of
the ligand and Phe-834
residue
a) H-π interaction with
Phe-236 and Phe-834.
b) Hydrophobic interactions
with Met838, Ile829, Pro827,
Leu828, and 765Ile.
Tg -14.1626 No hydrogen bonding a) No π-π interactions.
b) Hydrophobic interactions
with Met-838, Phe-834, Ile-829,
Leu-828, Pro-827, Phe-776,
Val-773, Val-772, Val-769, Asn-768,
Ile-765, Ile-761, Pro-308, Ile-267,
Val-263, Leu-260, Gln-259,
Phe-256, Leu-253,
Leu-249 residues.
Fig. 2. Effect of tetrahydroquinolines JS-56 and JS-92 on MCF-7 cell viability. MCF-7 cells were treated with
different concentrations of these compounds. Each point represents the mean ± SD of three indepen-
dent experiments. IC50 values obtained for JS-56 and JS-92 were 9.74 μM and 5.03 μM, respectively.
250 Maksoud et al.
Investigación Clínica 63(3): 2022
induced cytotoxicity in MCF-7 cells in a dose-
dependent manner, with an IC50 of 9.74 μM and
5.03 μM for JS-56 and JS-92, respectively. At a
concentration of 50 μM, JS-56 decreased cell
viability by approximately 88%. At the same
time, JS-92 exerted a more substantial effect
than JS-56 on this cell line, being able to re-
duce almost 100% of MCF-7 cell viability at the
same concentration.
Effect of tetrahydroquinolines JS-56 and
JS-92 on intracellular Ca2+ mobilization
To study the effect of these tetrahydro-
quinolines on the [Ca2+]i, cells were loaded
with the Ca2+ fluorescent indicator Fura-2. The
addition of JS-56 induced a rapid increase in
[Ca2+]i, reaching a peak followed by a continu-
ous decrease (Fig. 3A), as a consequence of
[Ca2+]i regulation carried out by the cellular
homeostatic machinery until it reached a new
steady-state Ca2+ level, which is greater than
the original basal level. This response is charac-
teristic of compounds capable of inducing the
release of Ca2+ from the ER, with the conse-
quent opening of plasma membrane Ca2+ chan-
nels (SOCE) activated by emptying this intra-
cellular reservoir 19,33. The effect observed upon
the addition of thapsigargin (Tg) (Fig. 3B), a
potent SERCA inhibitor, was similar to that ob-
served with JS-56. However, the maximum Ca2+
peak achieved with Tg was higher than JS-56.
Additionally, when comparing Figures 3A and
3B, the addition of Tg after JS-56 (Fig. 3A) still
induced a slight increase in [Ca2+]i. However,
the tetrahydroquinoline did not exert any dis-
cernible effect after Tg addition (Fig. 3B).
The same experiments were performed
in the absence of extracellular Ca2+ to deter-
mine the origin of the cation increase. Under
these experimental conditions, the maximum
Ca2+ peaks obtained with JS-56 (Fig. 3C) or
Tg (Fig. 3D) were considerably lower than
Fig. 3. Effect of tetrahydroquinoline JS-56 and thapsigargin (Tg) on [Ca2+]i in MCF-7 cells. The arrows in-
dicate the moment in which JS-56 (20 μM) and Tg (2 μM) were added, in the presence (Fig. A and
B) or absence (EGTA, Fig. C and D) of 2 mM CaCl2. The curves are representative of at least three
independent experiments.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 251
Vol. 63(3): 243 - 261, 2022
those reached in the presence of extracellu-
lar Ca2+ (Figs. 3A and 3B). Furthermore, the
Ca2+ levels obtained, after the maximum peak
and the subsequent decrease, were equal to
the basal levels (Figs. 3C and 3D). Moreover,
when Tg was added after JS-56, a minimal rise
in [Ca2+]i was detected, indicating again that,
different from Tg, the tetrahydroquinoline
was not able to release all the Ca2+ from the
ER (Fig. 3D). Likewise, the effect of the tet-
rahydroquinoline JS-92 was evaluated under
the same experimental conditions, obtaining
essentially the same overall results, as can be
observed in Fig. 4.
Effect of tetrahydroquinolines JS-56
and JS-92 on SERCA activity
To identify a possible mechanism of ac-
tion of these compounds, which could explain
their effect on the increase in [Ca2+]i, the
ability to inhibit SERCA activity was exam-
ined, considering that the results obtained
were very similar to those found by the use
of Tg. Fig. 5 shows the effect of increasing
concentrations of JS-56 and JS-92 on ATPase
activity. Both tetrahydroquinolines inhibited
SERCA activity in a dose-dependent man-
ner. The maximum inhibitory effect of JS-
56 and JS-92 on SERCA activity was 56 and
51%, respectively. Albeit none of the tetrahy-
droquinolines could completely inhibit the
SERCA activity, these results demonstrate
that part of the effect of these compounds
on [Ca2+]i elevation was undoubtedly due to
SERCA activity inhibition.
Effect of tetrahydroquinolines JS-56
and JS-92 on the mitochondrial
electrochemical potential
Mitochondria use their electrochemical
membrane potential ψm) to accumulate
Ca2+ through the electrophoretic Ca2+ uni-
porter (MCU). The effects of the tetrahydro-
quinolines and Tg on the mitochondrial Δψ
were studied by using the fluorescent mark-
er rhodamine 123. This fluorescent indica-
Fig. 4. Effect of tetrahydroquinoline JS-92 and thapsigargin (Tg) on [Ca2+]i in MCF-7 cells. The arrows indi-
cate the addition of JS-92 (20 μM) and Tg (2 μM), in the presence (Fig. A and B) or absence (EGTA,
Fig. C and D) of 2 mM CaCl2. The curves are representative of at least three independent experiments.
252 Maksoud et al.
Investigación Clínica 63(3): 2022
tor accumulates in the mitochondrial matrix
according to the electrochemical gradient.
Thus, an increase in fluorescence can be ob-
served due to the mitochondrial H+ gradient
dissipation, concomitantly with the release
of rhodamine 123 from the mitochondria to
the cytoplasm and then to the extracellular
milieu. Results from Fig. 6 demonstrate that
both JS-56 and JS-92 (20 μM) and Tg (2 μM)
produced a rapid and pronounced elevation
in fluorescence due to rhodamine 123 release
from mitochondria. However, the increment
in rhodamine 123 fluorescence caused by Tg
was more significant than those produced by
the tetrahydroquinolines (Figs. 6B and 6D).
It can also be noticed that JS-92 (Fig. 6C)
generated an elevation of rhodamine 123
fluorescence higher than JS-56 (Fig. 6A).
The effect of Tg after the addition
of the tetrahydroquinolines was less pro-
nounced (Figs. 6A and 6C). It is worthwhile
to mention that this fluorescence elevation
with both Tg and tetrahydroquinolines was
expected since it was due to the partial dis-
sipation of mitochondrial Δψ, generated by
the entry of Ca2+ released from the ER6. On
the other hand, after Tg application, the
addition of JS-56 and JS-92 led to an addi-
tional boost in fluorescence (Figs. 6B and
6D), indicating a superior Δψ dissipation,
which cannot be merely explained by the en-
trance of Ca2+ released from the ER. Thus,
these results suggest that both tetrahydro-
quinolines can directly affect the mitochon-
dria by Ca2+-independent mechanisms. As
expected, the positive control, carbonyl cy-
anide-4-(trifluoromethoxy)phenylhydrazone
(FCCP), generated a complete collapse of
the electrochemical gradient.
Effect of tetrahydroquinolines JS-56 and
JS-92 on the activity of caspases 7 and 9
We then evaluated the possible effect
of these tetrahydroquinolines on the induc-
tion of apoptosis in MCF-7 cells. Accordingly,
the activation of caspase 9, a known caspase
that initiates apoptosis, was assessed. The
1/2 IC50, IC50, and 2 IC50 obtained in the
MTT assay were used for both compounds.
As shown in Fig. 7 (upper panel), both JS-
56 and JS-92 significantly enhanced the ac-
tivity of caspase-9 compared to the control
in a dose-dependent manner. Sphingosine
(Sph), which promotes an increase in [Ca2+]
i in MCF-7 cells 33, and 2,5-di-t-butyl-1,4-
benzohydroquinone (BHQ), a known SERCA
reversible-inhibitor 34, also elevated the ac-
tivity of this initiating caspase. The positive
control Stau, which activates apoptosis in
MCF-7 cells 35, effectively activated caspase
9 at a level similar to the IC50 of the tetrahy-
droquinolines.
Fig. 5. Effect of tetrahydroquinolines JS-56 and JS-92 on SERCA activity. Each point represents the mean ±
SD of three independent experiments.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 253
Vol. 63(3): 243 - 261, 2022
Likewise, treatment of MCF-7 cells with
both tetrahydroquinolines raised the activity
of caspase 7 (Fig. 7, lower panel). Similar re-
sults were detected with the positive control
Stau and the SERCA inhibitor BHQ. The ac-
tivation-induced by JS-92 was slightly higher
compared to the rest of the compounds. This
effector caspase will be responsible for cleav-
ing specific cell targets to promote apopto-
sis in these cells.
Effect of tetrahydroquinolines JS-56 and
JS-92 on DNA fragmentation
An apoptotic cell is typically character-
ized by: changes in cell morphology, phos-
phatidylserine externalization, cytoplasm
contraction, shrinkage of the nucleus, and
DNA fragmentation 15. As a complement to
the results obtained in caspases 7 and 9 ac-
tivity assays, DNA fragmentation was ana-
lyzed through a TUNEL assay using Stau as a
positive control. In Fig. 8, cells were divided
into four populations displaying the percent-
age of cells in necrosis (propidium iodide,
IP+), apoptosis (fluorescein isothiocyanate,
FITC+), and alive (negative for both mark-
ers); apoptotic cells can either be IP- or IP+.
Apoptotic cells exhibit DNA fragmentation,
which allows FITC incorporation. However,
a subset of these apoptotic cells is also IP+,
denoting that these also have a loss of plas-
ma membrane integrity, corresponding to a
Fig. 6. Effect of tetrahydroquinolines JS-56 and JS-92 on the mitochondrial electrochemical potential in
MCF-7 cells. The fluorescent marker rhodamine 123 was used for these experiments, expressing re-
sults in arbitrary units (AU) of fluorescence. The arrows indicate the moment of the addition of the
different effectors: JS-56 or JS-92 (20 μM), thapsigargin (Tg, 2 μM), and FCCP (10 μM). The curves
are representative of at least three independent experiments.
254 Maksoud et al.
Investigación Clínica 63(3): 2022
later stage of apoptosis in which cells can
no longer control the passage of elements
across the membrane, with eventual cell ly-
sis in the final stage. The quantitative data is
represented in Table 2.
JS-56 treatment (10 μM) for 24 h in-
duced apoptosis, with 65.9% of FITC+ IP-
cells in this stage and only 4.3% of FITC+ IP+
cells. Treatment with Stau (1 μM) and JS-92
(5 μM) produced a significant percentage of
FITC+ IP- (31.5 and 46.5% for Stau and JS-
92, respectively) and FITC+ IP+ cells (60.6
and 35.5% for Stau and JS-92, respectively).
Hence, JS-92 has a more potent effect than
JS-56, which measurements of caspase ac-
tivity have also evidenced. These results are
statistically significant compared to the neg-
ative control (DMSO-treated MCF-7 cells).
Effect of tetrahydroquinolines JS-56
and JS-92 on NF-κB-dependent gene
expression
The nuclear transcription factor NF-κB
regulates numerous genes involved in multi-
ple cellular processes such as apoptosis, cell
proliferation, and differentiation. To evalu-
ate NF-κB-dependent gene expression, HeLa
tumor cells transfected with a luciferase re-
porter gene under the control of IL-6 pro-
moter, which contains critical binding sites
for NF-κB, were used.
Tetrahydroquinolines JS-56 and JS-92,
when added alone, were not able to increase
the NF-κB-dependent gene expression, with
values similar to those observed in the nega-
tive control (Fig. 9). The luminescence mea-
sured in the negative control corresponds
to the basal NF-κB-dependent gene expres-
sion on MCF-7 cells. The NF-κB-dependent
gene expression is elevated when cells were
treated with 100 nM of PMA; it has been
Table 2
Percentage of living, apoptotic, and necrotic cells after treatment with JS-56 and JS-92.
Treatment group
Percentage
Living cells Apoptosis
(IP-)
Apoptosis
(IP+)
Necrosis
Control DMSO 93.5 ± 4.6 0.9 ± 0.6 0.1 ± 0.1 4.1 ± 1.2
Staurosporine 3.9 ± 1.3 31.5 ± 7.5 60.6 ± 6.8 1.2 ± 0.5
JS-56 25.5 ± 9.3 65.9 ± 4.7 4.3 ± 2.8 1.9 ± 1.6
JS-92 14.6 ± 6.3 46.5 ± 5.5 35.5 ± 4.8 3.4 ± 1.8
Fig. 7. Effect of tetrahydroquinolines JS-56 and JS-
92 on the activity of caspases 9 (initiator)
and 7 (effector). Stau (staurosporine), BHQ
(2,5-di-t-butyl-1,4-benzohydroquinone) and
Sph (sphingosine). The bars represent the
mean ± SD derived from three independent
experiments. * and ** p < 0.05 compared
to the control of untreated cells or treated
with JS-56, respectively.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 255
Vol. 63(3): 243 - 261, 2022
reported that PKC can activate IκBs kinase
complexes (IKKs), which eventually leads
to NF-κB activation 36. Interestingly, the re-
sults demonstrate that the NF-κB-dependent
gene expression increased significantly when
cells were treated with PMA plus JS-56 or
JS-92, indicating an apparent effect of both
compounds on the PKC-stimulated NF-κB-
dependent gene expression.
Molecular docking
The experimental results on SERCA
activity obtained for the tetrahydroquino-
lines JS-56 and JS-92 were confirmed using
molecular docking for SERCA protein (PDB
code 2AGV). Interestingly, both Tg and BHQ
SERCA inhibitors showed practically the
same disposition as crystallized inhibitors
at the respective protein binding sites, re-
flecting the computational package’s poten-
tial for this molecular docking. In general,
the tetrahydroquinolines showed a higher
preference for Tg binding sites than for the
BHQ active site. This effect generates a li-
gand-protein complex with binding energies
between -12 and -13 kcal/mol at the active
sites of Tg; this is in contrast to weak dock-
ing energies found at the BHQ active site for
both tetrahydroquinolines. Particularly, ex-
perimental results have demonstrated that
Tg exerted a more potent inhibitory activity
(IC50 of 0.2 nM) than BHQ (IC50 of 20 nM)
26. The preference of these large tetrahydro-
quinolines for Tg binding sites may be as-
sociated with the larger box size than BHQ
active sites, which may facilitate the stabi-
lization of the inhibitor-protein complex for
these bulky and inflexible molecules. It is
Fig. 8. Effect of tetrahydroquinolines JS-56 and JS-92 on DNA fragmentation in MCF-7 cells. The MCF-7 cells
were incubated with Fluorescein-12-dUTP (fragmented DNA marker) and Propidium Iodide (IP, core
marker). Cells were treated with staurosporine (1 μM), JS-56 (10 μM), and JS-92 (5 μM). The bars
represent the mean ± SD derived from three independent experiments. * p < 0.0001 compared to the
control of cells treated with DMSO.
Fig. 9. Effect of tetrahydroquinolines JS-56 and JS-
92 on NF-κB-dependent gene expression.
HeLa tumor cells were transfected with a
luciferase reporter gene under the control
of the IL-6 promoter, which contains criti-
cal binding sites for NF-κB. Control, untrea-
ted MCF-7 cells; PMA, phorbol 12-myristate
13-acetate. * and ** p<0.05 compared to
the control of untreated cells or treated
with PMA, respectively.
256 Maksoud et al.
Investigación Clínica 63(3): 2022
essential to mention that our analysis was
focused on data obtained for the Tg active
site from chain-α, which exhibited the lower
binding energy complex.
DISCUSSION
Tetrahydroquinolines derivatives are an
essential group of natural or synthetic com-
pounds showing a wide range of biological
activities. Numerous compounds capable of
inducing a Ca2+ signal followed by a cellular
response have been reported. Ca2+ is one of
the essential ionic constituents in the body.
It has a central regulatory role in differentia-
tion, secretion, contraction, transcription,
phosphorylation, and apoptosis processes
10. Ca2+ is the component required by milk
to calcify newborn bones and teeth and
modulates the proliferation, differentiation,
and apoptosis of mammary gland epithelial
cells 37. The modulation of Ca2+ homeosta-
sis and its signaling could represent a viable
therapeutic approach in the treatment and/
or prevention of breast cancer. Therefore,
regulators of Ca2+ homeostasis or its signal-
ing in mammary gland tumors are possible
targets for drugs. In this investigation, we
used the tetrahydroquinolines JS-56 and JS-
92, modified derivatives from clove and cin-
namon plants, to study the effect of these
drugs on Ca2+ homeostasis and its possible
relationship with apoptosis in breast cancer
cells MCF-7.
In the interest of evaluating the cyto-
toxicity of tetrahydroquinolines JS-56 and
JS-92 on MCF-7 breast cancer cells, the MTT
assay was developed, allowing the determi-
nation of mitochondrial viability on treated
cells. Both compounds are cytotoxic to MCF-
7 cells in a dose-dependent manner, with an
IC50 of 9.74 μM and 5.031 μM for JS-56 and
JS-92, respectively. This difference may be
related to the bromine atom that JS-92, in-
stead of the cyanide atom present in JS-56.
IC50 values for these compounds in MCF-7
cells are significantly lower than those re-
ported previously in human dermis fibro-
blast, 92.87 μM and 42.48 μM for JS-56 and
JS-92, respectively 17.
Through fluorimetric techniques, we
were able to perform measurements of intra-
cellular Ca2+ in MCF-7 cells, aiming to find
possible mechanisms of action for these com-
pounds. The results revealed that JS-56 and
JS-92 promote an increase in [Ca2+]i. Like-
wise, the Tg effect on [Ca2+]i is diminished
when added after the impact generated by
the tetrahydroquinolines, compared to when
it is initially added, allowing us to conclude
that these compounds promote the release
of Ca2+ from a compartment sensitive to Tg,
namely the ER. Similar experiments were
done but using a solution without CaCl2. Un-
der these experimental conditions, the maxi-
mum Ca2+ peaks detected with JS-56, JS-92
or Tg are considerably reduced compared to
those achieved in the presence of extracellu-
lar Ca2+, and instead, returning to the basal
level. These results show an influx of Ca2+
from the outside. This effect may be caused
by a store-operated Ca2+ Entry (SOCE) or a
capacitive Ca2+ entry induced by emptying
Ca2+ from the ER 19,33.
The lack of an additive effect of Tg with
tetrahydroquinolines on the [Ca2+]i suggests
a possible similar mechanism of action. The
possible inhibitory potential of these com-
pounds on SERCA activity was examined ac-
cordingly. Both tetrahydroquinolines inhibit
the enzyme activity with similar maximum
inhibitory effects of 56 and 51% for JS-56
and JS-92, respectively. This effect also in-
dicates that different from Tg, these com-
pounds could not completely inhibit the
ATPase activity. However, part of the overall
effect on the onset of apoptosis is the partial
inhibition of SERCA activity. As it has been
shown that Tg can induce ER stress via SER-
CA inhibition, JS-56 and JS-92 are likely po-
tential candidates for generating ER stress
through this identical mechanism. In this
context, it is worth mentioning that SERCA
targeting could represent an effective thera-
peutic strategy in treating pathogens and
cancer cells 38.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 257
Vol. 63(3): 243 - 261, 2022
The effect of both compounds on the
mitochondrial membrane Δψm was deter-
mined since the accumulation of mitochon-
drial Ca2+ is a necessary event in the initia-
tion of apoptosis via the intrinsic pathway.
The primary source of the cation derives
from its transfer from the ER 39. The results
revealed that both tetrahydroquinolines
generate an increase in rhodamine 123 fluo-
rescence. Although the increment in rhoda-
mine fluorescence caused by Tg was higher
than the compounds, its effect after the ad-
dition of the tetrahydroquinolines was less
pronounced, suggesting that they share part
of the mechanism of action on mitochondria.
The fluorescence increase after treating the
cells with both Tg and tetrahydroquinolines
was expected due to the partial dissipation
of mitochondrial Δψm. This change is gen-
erated by the entry of Ca2+ released from
the ER 6. When both tetrahydroquinolines
were added after Tg, an additional increase
in fluorescence was triggered, indicating a
superior dissipation of the Δψ. These unex-
pected results suggested that JS-56 and JS-
92 induced a cell response independent of
[Ca2+]i. The effects are also independent of
the Ca2+ electrophoretic uniporter. These
effects have been reported previously with
concentrations higher than 1 μM 40. Accord-
ingly, both tetrahydroquinolines, similar to
the SERCA inhibitor, could interact with the
mitochondria, generating a widespread ef-
fect.
Given that the Δψm loss of the mito-
chondrial membrane is a hallmark event of
apoptosis, the presumed activation of char-
acteristic processes associated with pro-
grammed cell death after applying the tet-
rahydroquinolines was investigated. In this
study, it was demonstrated that tetrahydro-
quinolines JS-56 and JS-92 are capable of
activating caspases 7 and 9 and promoting
DNA fragmentation. Activation of caspases
and the percentage of cells with DNA frag-
mentation was higher for JS-92, which coin-
cides with its superior cytotoxic effect. DNA
fragmentation in conjunction with caspase
activation and the results demonstrating
that these tetrahydroquinolines cause the
dissipation of mitochondria Δψm to support
the conclusion that these compounds acti-
vate apoptosis in MCF-7 cells. Additionally,
after the inhibition of SERCA and the conse-
quent reduction of the [Ca2+] at the ER, mis-
folded proteins could accumulate. The un-
folded protein response could be triggered,
leading the cells to death 12,13. By suppress-
ing this enzyme’s activity and promoting the
cation’s entry into the mitochondria, the
release of pro-apoptotic factors from this or-
ganelle would be facilitated 15. Thus, part of
the potential anticancer effect of these com-
pounds would be mediated through SERCA
inhibition. Based on the accumulated evi-
dence, we suggest that disturbances of the
Ca2+ signal induced by these tetrahydroquin-
olines would trigger apoptosis by promoting
ER stress and the release of proapoptotic
factors from the mitochondria after the en-
try of Ca2+ into this organelle.
According to our results, JS-56 and JS-
92 failed to elevate the NF-κB-dependent
gene expression. However, NF-κB-dependent
gene expression was augmented when cells
were treated with PMA, an activator of PKC
which upregulates NF-κB 35. Interestingly, the
NF-κB-dependent gene expression is further
elevated when cells were exposed to the com-
bination of PMA and tetrahydroquinolines,
possibly because these compounds enhance
the affinity of PKC for PMA or modulate the
PMA activity making it more potent and/or
stable. Since NF-κB is a regulator of the ex-
pression of multiple genes, additional experi-
ments are needed to determine its putative
participation in pro-apoptotic events. We do
not rule out non-specific binding of other
transcription factors to this NF-κB binding
site on the IL-6 promoter. Future research
will confirm the participation of this tran-
scription factor by determination of cytoplas-
mic/nuclear levels of p65 as well as confirma-
tion of these results in breast cancer cells.
Based on the results that both tetrahy-
droquinolines may interact with the Tg bind-
258 Maksoud et al.
Investigación Clínica 63(3): 2022
ing site of SERCA, we used a docking mod-
el to test the possibility of interaction. It
should be noted that both tetrahydroquino-
lines exhibited docking energies higher than
Tg, indicating that the Tg-Ca2+-ATPase com-
plex is more stable than the quinoline-Ca2+-
ATPase complex (Table 1). These theoretical
findings are in good agreement with experi-
mental data, where Tg exhibited the higher
experimental inhibitory activity (IC50 of 0.2
nM) compared to the tested tetrahydroquin-
olines (IC50 of 60 and 100 µM for JS-56 and
JS-92, respectively). The high stabilization
of the Tg-Ca2+-ATPase complex can be asso-
ciated with diverse intermolecular interac-
tions with residues located in the Tg1 active
site, including (i) a large number of hydro-
phobic interactions with at least 20 residues
in chain A and 13 residues in chain B; (ii)
from one to two strong H-interactions with
Ile-829 and Phe-834 for JS-56 and JS-92, re-
spectively; and (iii) H-π interactions between
phenolic, quinolinic, or indolic rings with
specific phenyl rings in Phe-834, Phe-256
or Tyr-837 (Table 1 and Fig. 10). However,
it is essential to mention that these diverse
interactions found for these tetrahydroquin-
olines are located in turn to a specific re-
Fig. 10. Molecular docking of the tetrahydroquinolines JS-56 and JS-92 on the TG1 active site. (A) TG1 (gray
ribbons) and TG2 (blue ribbons) active sites into SERCA protein; (B) Location of JS-56 (cyan color)
and JS-92 (orange color) into the TG1 active site pocket of SERCA protein; (C) Superimposition
between JS-56 (cyan) and JS-92 (orange) with TG1 substrate (yellow color); (D) Disposition and in-
teractions of JS-56 (cyan) (D) and JS-92 (orange) (E) into TG1 active site. Tentative H-π interactions
between phenolic, indolic, or quinolinic rings of JS-56 and JS-92 with Phe-236 (gray color), Phe-834
(blue color), and Phe-256 (green color) residues.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 259
Vol. 63(3): 243 - 261, 2022
gion of the active site. At the same time, Tg,
bearing a highly flexible octanoic acid resi-
due at C2, makes hydrophobic contacts with
non-polar patches on the outer surface of
SERCA’s trans-membrane domain and pos-
sibly with surrounding lipids, facilitating the
stabilization of the ligand-protein complex.
The molecular docking analysis shows that
hydrophobic ligands containing long and
flexible side chains are needed to design ef-
fective inhibitors of the Ca2+-ATPase enzyme
due to the high hydrophobic and extended
nature of the Ca2+-ATPase active site. Future
studies will continue to evaluate these types
of compounds and their effect on Ca2+ and
apoptosis, developing new analogs with dif-
ferent types of chemical modifications on a
more diverse panel of breast cancer cells and
other types of malignancies.
Both tetrahydroquinolines derivatives,
JS-56 and JS-92, induced cell death in MCF-
7 cells through a combination of events.
These events include the elevation of the
[Ca2+]i, inhibiting SERCA via direct interac-
tion, partially damaging the mitochondria,
which induced activation of apoptotic path-
ways and consequently MCF-7 cell death.
Therefore, these results show the potential
that these types of compounds would have
in anti-neoplastic treatment in humans. It
is hoped that more powerful analogs can be
developed in the future.
Funding
This work was supported by [Fondo Na-
cional de Ciencia, Tecnología e Investigación,
Venezuela (FONACIT)] (Grants 2017000274
and 2018000010), and the [Consejo de De-
sarrollo Científico y Humanístico-Univer-
sidad Central de Venezuela (CDCH-UCV)]
(Grant PG-03-8728-2013/2) to G.B.
Conflict of interest
The authors declare no conflicts of in-
terest.
Authors contributions
SM: Investigation, Methodology, Data
curation, Writing – original draft, Writing –
review & editing. AM: Methodology, Writing
– review & editing. LC: Methodology, Writing
– review & editing. FS: In silico analysis, Writ-
ing – review & editing. AP: NF-κB gene expres-
sion analysis. VK: In silico analysis, isolation
and purification of natural compounds. DMA:
Isolation and purification of natural com-
pounds. AR: Methodology, Writing – review &
editing. FA: In silico analysis, Writing – review
& editing. JS: Flow cytometry analysis, Writ-
ing – review & editing. GB: Conceptualiza-
tion, Formal analysis, Supervision, Funding
acquisition, Writing – review & editing.
Author’s ORCID numbers
Semer Maksoud:
0000-0001-9773-9415
Adriana Mayora:
0000-0002-6817-2066
Laura Colman:
0000-0002-5188-880X
Felipe Sojo:
0000-0002-6559-4845
Adriana A. Pimentel:
0000-0002-8862-9318
Vladimir V. Kouznetsov:
0000-0003-1417-8355
Diego R. Merchán-Arenas:
0000-0001-9243-5914
Ángel H. Romero:
0000-0001-8747-5153V
Francisco Arvelo:
0000-0003-1590-358x
Juan Bautista De Sanctis:
0000-0002-5480-4608
Gustavo Benaim:
0000-0002-9359-5546
260 Maksoud et al.
Investigación Clínica 63(3): 2022
REFERENCES
1. World Health Organization (WHO). Breast
Cancer (December 2021). https://www.
who.int/news-room/fact-sheets/detail/
breast-cancer.
2. Peart O. Breast intervention and breast
cancer treatment options. Radiol Technol
2015;86:535M-558M; quiz 559–62.
3. Sergeev IN, Li S, Colby J, Ho C-T, Dus-
henkov S. Polymethoxylated flavones indu-
ce Ca2+-mediated apoptosis in breast can-
cer cells. Life Sci 2006;80:245–253.
4. Sareen D, Darjatmoko SR, Albert DM,
Polans AS. Mitochondria, calcium, and cal-
pain are key mediators of Resveratrol-indu-
ced apoptosis in breast cancer. Mol Phar-
macol 2007;72:1466–1475.
5. Lee J-H, Li Y-C, Ip S-W, Hsu S-C, Chang
N-W, Tang N-Y, Yu C-S, Chou S-T, Lin
S-S, Lino C-C, Yang J-S, Chung J-G. The
role of Ca2+ in baicalein-induced apopto-
sis in human breast MDA-MB-231 cancer
cells through mitochondria- and caspa-
se-3-dependent pathway. Anticancer Res
2008;28:1701–1711.
6. Pimentel AA, Felibertt P, Sojo F, Colman
L, Mayora A, Silva ML, Rojas H, Dipolo
R, Suarez AI, Compagnone RS, Arvelo F,
Galindo-Castro I, De Sanctis JB, Chiri-
no P, Benaim G. The marine sponge toxin
agelasine B increases the intracellular Ca2+
concentration and induces apoptosis in hu-
man breast cancer cells (MCF-7). Cancer
Chemother Pharmacol 2012;69:71–83.
7. Zhang Z, Teruya K, Eto H, Shirahata S.
Induction of apoptosis by low-molecular-
weight fucoidan through calcium- and
caspase-dependent mitochondrial pathways
in MDA-MB-231 breast cancer cells. Biosci
Biotechnol Biochem 2013;77:235–242.
8. Al-Taweel N, Varghese E, Florea A-M, Büs-
selberg D. Cisplatin (CDDP) triggers cell
death of MCF-7 cells following disruption of
intracellular calcium ([Ca2+]i) homeosta-
sis. J Toxicol Sci 2014;39:765–774.
9. Sehgal P, Szalai P, Olesen C, Praetorius
HA, Nissen P, Christensen SB, Engedal
N, Møller J V. Inhibition of the sarco/en-
doplasmic reticulum (ER) Ca2+-ATPase by
thapsigargin analogs induces cell death via
ER Ca2+ depletion and the unfolded protein
response. J Biol Chem 2017;292:19656–
19673.
10. Carafoli E. Calcium signaling: A tale for all
seasons. Proc Natl Acad Sci 2002;99:1115–
1122.
11. Carafoli E. Intracellular calcium homeos-
tasis. Annu Rev Biochem 1987;56:395–433.
12. Breckenridge DG, Germain M, Mathai JP,
Nguyen M, Shore GC. Regulation of apop-
tosis by endoplasmic reticulum pathways.
Oncogene 2003;22:8608–8618.
13. Szegezdi E, Logue SE, Gorman AM, Sa-
mali A. Mediators of endoplasmic reticu-
lum stressinduced apoptosis. EMBO Rep
2006;7:880–885.
14. Rizzuto R, Bernardi P, Pozzan T. Mito-
chondria as allround players of the calcium
game. J Physiol 2000;529:37–47.
15. Hengartner MO. The biochemistry of apop-
tosis. Nature 2000;407:770–776.
16. Kouznetsov V V, Bello Forero JS, Amado
Torres DF. A simple entry to novel spiro
dihydroquinoline-oxindoles using Povarov
reaction between 3-N-aryliminoisatins and
isoeugenol. Tetrahedron Lett 2008; doi:
10.1016/j.tetlet.2008.07.096.
17. Kouznetsov V, R. Merchan Arenas D, Ar-
velo F, S. Bello Forero J, Sojo F, Munoz
A. 4-Hydroxy-3-methoxyphenyl substituted
3-methyl-tetrahydroquinoline derivatives
obtained through Imino Diels-Alder reac-
tions as potential antitumoral agents. Lett
Drug Des Discov 2010;7:632–639.
18. Mosmann T. Rapid colorimetric assay for
cellular growth and survival: Application to
proliferation and cytotoxicity assays. J Im-
munol Methods 1983;65:55–63.
19. Colina C, Flores A, Castillo C, Rosario Ga-
rrido M del, Israel A, DiPolo R, Benaim G.
Ceramide-1-P induces Ca2+ mobilization in
Jurkat T-cells by elevation of Ins(1,4,5)-P3
and activation of a store-operated calcium
channel. Biochem Biophys Res Commun
2005;336:54–60.
20. Grynkiewicz G, Poenie M, Tsien RY. A new
generation of Ca2+ indicators with greatly
improved fluorescence properties. J Biol
Chem 1985;260:3440–3450.
21. Eletr S, Inesi G. Phospholipid orientation
in sacroplasmic membranes: Spin-label ESR
and proton NMR studies. Biochim Biophys
Acta - Biomembr 1972;282:174–179.
Apoptotic effect of tetrahydroquinoline derivatives in breast cancer cells 261
Vol. 63(3): 243 - 261, 2022
22. Fiske CH, Subbarow Y. The colorimetric
determination of phosphorus. J Biol Chem
1925;66:375–400.
23. Benaim G, Cervino V, Lopez-Estraño C,
Weitzman C. Ethanol stimulates the plas-
ma membrane calcium pump from human
erythrocytes. Biochim Biophys Acta - Bio-
membr 1994;1195:141–148.
24. Benaim G, Sanders JM, Garcia-Marchán
Y, Colina C, Lira R, Caldera AR, Payares
G, Sanoja C, Burgos JM, Leon-Rossell A,
Concepcion JL, Schijman AG, Levin M,
Oldfield E, Urbina JA. Amiodarone has in-
trinsic anti-Trypanosoma cruzi activity and
acts synergistically with posaconazole. J
Med Chem 2006;49:892–899.
25. Kagawa S, Gu J, Honda T, McDonnell TJ,
Swisher SG, Roth JA, Fang B. Deficiency
of caspase-3 in MCF7 cells blocks Bax-me-
diated nuclear fragmentation but not cell
death. Clin Cancer Res 2001;7:1474–1480.
26. Obara K, Miyashita N, Xu C, Toyoshima I,
Sugita Y, Inesi G, Toyoshima C. Structural
role of countertransport revealed in Ca2+
pump crystal structure in the absence of
Ca2+. Proc Natl Acad Sci 2005;102:14489–
14496.
27. Guex N, Peitsch MC. SWISS-MODEL and
the Swiss-Pdb Viewer: An environment for
comparative protein modeling. Electropho-
resis 1997;18:2714–2723.
28. Fusi F, Saponara S, Gagov H, Sgaragli
G. 2,5-Di-t-butyl-1,4-benzohydroquinone
(BHQ) inhibits vascular L-type Ca2+ chan-
nel via superoxide anion generation. Br J
Pharmacol 2001;133:988–996.
29. Lytton J, Westlin M, Hanley MR. Thapsigar-
gin inhibits the sarcoplasmic or endoplas-
mic reticulum Ca-ATPase family of calcium
pumps. J Biol Chem 1991;266:17067–
17071.
30. Thompson M. ArgusLab 4.0 Planaria Soft-
ware. LLC: Seatle, WA. 2004.
31. Becke AD. A new mixing of Hartree–Fock
and local densityfunctional theories. J
Chem Phys 1993;98:1372–1377.
32. Young DC. Computational Drug Design.
Hoboken, NJ, USA: John Wiley & Sons, Inc.,
2009.
33. Colina C, Flores A, Rojas H, Acosta A,
Castillo C, Rosario Garrido M del, Israel
A, DiPolo R, Benaim G. Ceramide increase
cytoplasmic Ca2+ concentration in Jurkat
T cells by liberation of calcium from intra-
cellular stores and activation of a store-
operated calcium channel. Arch Biochem
Biophys 2005;436:333–345.
34. Moore GA, McConkey DJ, Kass GEN,
O’Brien PJ, Orrenius S. 2,5-Di(tert-
butyl)-1,4-benzohydroquinone - a novel in-
hibitor of liver microsomal Ca2+ sequestra-
tion. FEBS Lett 1987;224:331–336.
35. Tang D, Lahti JM, Kidd VJ. Caspase-8 acti-
vation and bid cleavage contribute to MCF7
cellular execution in a caspase-3-dependent
manner during staurosporine-mediated
apoptosis. J Biol Chem 2000;275:9303–
9307.
36. Schmitz ML, Bacher S, Dienz O. NFĸB
activation pathways induced by T cell costi-
mulation. FASEB J 2003;17:2187–2193.
37. Lee WJ, Monteith GR, Roberts-Thomson
SJ. Calcium transport and signaling in the
mammary gland: Targets for breast can-
cer. Biochim Biophys Acta - Rev Cancer
2006;1765:235–255.
38. Tadini-Buoninsegni F, Smeazzetto S, Gual-
dani R, Moncelli MR. Drug interactions
with the Ca2+-ATPase from sarco(endo)plas-
mic reticulum (SERCA). Front Mol Biosci
2018;5:36.
39. Demaurex N. Cell biology: Apoptosis--the
calcium connection. Science 2003;300:65–
67.
40. Vercesi AE, Moreno SN, Bernardes CF,
Meinicke AR, Fernandes EC, Docampo R.
Thapsigargin causes Ca2+ release and co-
llapse of the membrane potential of Trypa-
nosoma brucei mitochondria in situ and of
isolated rat liver mitochondria. J Biol Chem
1993;268:8564–8568.