ArticlePDF Available

Antitumor Activity and Reductive Stress by Platinum(II) N‐Heterocyclic Carbenes based on Guanosine

Wiley
Chemistry - A European Journal
Authors:

Abstract and Figures

Platinum(II) complexes bearing N‐heterocyclic carbenes based guanosine and caffeine have been synthesized by unassisted C−H oxidative addition, leading to the corresponding trans‐hydride complexes. Platinum guanosine derivatives bearing triflate as counterion or bromide instead of hydride as co‐ligand were also synthesized to facilitate correlation between structure and activity. The hydride compounds show high antiproliferative activity against all cell lines (TC‐71, MV‐4‐11, U‐937 and A‐172). Methyl Guanosine complex 3, bearing a hydride ligand, is up to 30 times more active than compound 4, with a bromide in the same position. Changing the counterion has no significant effect in antiproliferative activity. Increasing bulkiness at N7, with an isopropyl group (compound 6), allows to maintain the antiproliferative activity while decreasing toxicity for non‐cancer cells. Compound 6 leads to an increase in endoplasmic reticulum and autophagy markers on TC71 and MV‐4‐11 cancer cells, induces reductive stress and increases glutathione levels in cancer cells but not in non‐cancer cell line HEK‐293.
This content is subject to copyright. Terms and conditions apply.
Antitumor Activity and Reductive Stress by Platinum(II)
N-Heterocyclic Carbenes based on Guanosine**
Maria Inês P. S. Leitão,[a] Maria Turos-Cabal,[b] Ana Maria Sanchez-Sanchez,[b]
Clara S. B. Gomes,[c, d, e] Federico Herrera,*[f] Vanesa Martin,*[b] and Ana Petronilho*[a]
Abstract: Platinum(II) complexes bearing N-heterocyclic car-
benes based guanosine and caffeine have been synthesized
by unassisted CH oxidative addition, leading to the corre-
sponding trans-hydride complexes. Platinum guanosine de-
rivatives bearing triflate as counterion or bromide instead of
hydride as co-ligand were also synthesized to facilitate
correlation between structure and activity. The hydride
compounds show high antiproliferative activity against all cell
lines (TC-71, MV-4-11, U-937 and A-172). Methyl Guanosine
complex 3, bearing a hydride ligand, is up to 30 times more
active than compound 4, with a bromide in the same
position. Changing the counterion has no significant effect in
antiproliferative activity. Increasing bulkiness at N7, with an
isopropyl group (compound 6), allows to maintain the
antiproliferative activity while decreasing toxicity for non-
cancer cells. Compound 6leads to an increase in endoplasmic
reticulum and autophagy markers on TC71 and MV-4-11
cancer cells, induces reductive stress and increases gluta-
thione levels in cancer cells but not in non-cancer cell line
HEK-293.
Introduction
Platinum-based drugs are responsible for circa 50 % of all
anticancer therapies worldwide, either in combination with
other therapies or as standalone treatment.[1,2] The first platinum
anticancer agent, cisplatin, began its clinical use more than
40 years ago and marked a milestone in the discovery and use
of metallodrugs.[3] Alongside with carboplatin and oxaliplatin,
cisplatin continues to play a major role in contemporary
oncology.[2,3] Cisplatin arrests the cell cycle at G2/M transition,
the magnitude of this cytostatic effect being dependent on the
duration and concentration of the treatment.[4] Despite their
effectiveness, the use of cisplatin and other metallodrugs has
several limitations. Their coordination is not restricted to DNA
nucleobases, and their binding to other biological substrates is
at the core of severe undesired side effects.[3] Cancer cells often
adapt to these drugs and become resistant, and tumour
recurrence is therefore common.[2,3] Subsequently, intense
research has been dedicated to the development of novel
platinum drugs.[2,5] A promising strategy is the coordination of
the metal moiety to naturally occurring and/or bioactive
ligands, which can provide a higher degree of selectivity.[6] In
this sense, modified nucleosides, such as organometallic
nucleosides,[7–9] show a great potential to fill this role. Modified
nucleosides are widely used in chemotherapy, acting as
antimetabolites that disrupt the synthesis of nucleic acids.[10,11]
The accessible incorporation of nucleoside analogues into
nucleic acids by the DNA repair machinery makes them
interesting candidates for combination with DNA-damaging
agents, such as cisplatin.[12] Indeed, combination therapies of
nucleosides and metallodrugs have proven effective for a
variety of cancer treatments.[2,13]
Previous work from our lab has established new method-
ologies for the synthesis of guanosine complexes based on
palladium and platinum.[14] In these complexes, the guanosine
ligand is bound to the metal centre as an N-heterocyclic
[a] Dr. M. I. P. S. Leitão, Dr. A. Petronilho
Instituto de Tecnologia Química e Biológica António Xavier
Avd República, 2780-157 Oeiras (Portugal)
E-mail: ana.petronilho@itqb.unl.pt
[b] M. Turos-Cabal, Dr. A. M. Sanchez-Sanchez, Prof. V. Martin
Facultad de Medicina, Dpto. Morfología y Biología Celular
Universidad de Oviedo- Instituto Universitario de Oncología del Principado
de Asturias (IUOPA)
Universidad de Oviedo and Instituto de Investigación Sanitaria del
Principado de Asturias (ISPA)
C/ Julián Clavería 6, 33006, Oviedo, Asturias (España)
E-mail: martinvanesa@uniovi.es
[c] Dr. C. S. B. Gomes
LAQV-REQUIMTE, Department of Chemistry. Associate Laboratory i4 HB
NOVA Faculdade de Ciências e Tecnologias, Universidade Nova de Lisboa
2829-516 Caparica (Portugal)
[d] Dr. C. S. B. Gomes
Institute for Health and Bioeconomy. UCIBIO
NOVA Faculdade de Ciências e Tecnologias, Universidade Nova de Lisboa
2829-516 Caparica (Portugal)
[e] Dr. C. S. B. Gomes
Applied Molecular Biosciences Unit, Department of Chemistry
NOVA Faculdade de Ciências e Tecnologias, Universidade Nova de Lisboa
2829-516 Caparica (Portugal)
[f] Prof. F. Herrera
BioISI Instituto de Biosistemas e Ciências integrativas, Dep. Química e
Bioquímica Faculdade de Ciências da Universidade de Lisboa Campo
Grande, 1749-016 Lisboa (Portugal)
E-mail: fherrera@fc.ul.pt
[**] A previous version of this manuscript has been deposited on a preprint
server (https://doi.org/10.26434/chemrxiv-2022-k8c22).
Supporting information for this article is available on the WWW under
https://doi.org/10.1002/chem.202301078
Chemistry—A European Journal
www.chemeurj.org
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (1 of 7) © 2023 Wiley-VCH GmbH
carbene, making the complex more stable.[15] These platinated
nucleosides were examined for their antiproliferative activity[14]
towards several human cell lines. Complex 1(Figure 1), bearing
an anionic guanosine derivative, has no relevant antiprolifer-
ative activity. By contrast, protic NHC complex 2(Figure 1) is
active for glioblastoma cell line U251, having a significant
activity when compared to cisplatin, while showing no cytotoxic
for non-cancer cells (HEK-293 cell line).[14]
We have recently reported the synthesis of a platinum
complex based on 7-methylguanosine by CH oxidative
addition leading to the formation of the hydride complex 3[15]
(Figure 1). Following our results previously described for
complex 2,[14] herein we report the antiproliferative activity of
compound 3. Additionally, and motivated by the results
obtained for 3, we examine the effect of synthetic variations at
the purine core on antiproliferative activity.
Results and Discussion
The anticancer activity of compound 3was evaluated via MTT
assay in four different human cancer cell lines (Figure 2), namely
TC-71 (Ewing’s sarcoma), MV-4-11 (myelomonocytic leukaemia),
U-937 (histiocytic lymphoma) and A-172 (glioblastoma). Com-
plexes 1and 2were active against the MV-4-11 cell line at
10 μM (2) and 100 μM (1), showing no activity against the three
other cell lines, even at the highest concentration tested. In
contrast, complex 3was active against the four cancer cell lines.
Differences in the antiproliferative activity between 1, bearing a
guanosynil ligand, and 2, with a guanosylidene ligand, were
previously reported by our group,[14] with complex 2with the
NHC ligand providing a higher antiproliferative activity. Since
the antiproliferative activity of 3is considerably higher than
that of 2, we hypothesized that the presence of the methyl
group and/or the presence of the hydride as co-ligands could
be responsible for the increased antiproliferative activity, in
addition to the beneficial effect of the NHC ligand.
To determine the main cause of the higher cytotoxicity, we
prepared complex 4. This complex bears a methyl group at N7
and a bromide instead of a hydride trans to the NHC. Complex
4is easily obtained by post-functionalization of compound 1
with methyl triflate (Scheme 1).
Characterization of 4by NMR spectroscopy shows similar-
ities with NHC 3, as expected. The most diagnostic feature is
found in the 13C{1H}, with the upfield of ca. 30 ppm of carbenic
carbon (C8) with respect to 3. In NHC 4, the C8 resonates as a
triplet at 155.3 ppm, while in NHC 3it resonates at 181.9 ppm.
This shift evidences the large differences in sigma donation of
the hydride versus bromide on C8 and the resulting trans
influence.[15] The antiproliferative activity was measured for
compound 4and the corresponding IC50 calculated and
compared with those of 3for the difference cell lines (Table 1).
Figure 1. Chemical structure of complexes 13.
Figure 2. Viability of four human cancer cell lines, namely MV-4-11, A-
172 TC-71 and U-937, after incubation with compounds A) 1, B) 2and C) 3,
for 48 h, assessed by evaluating its ability to metabolize MTT.
Chemistry—A European Journal
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (2 of 7) © 2023 Wiley-VCH GmbH
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Both compounds are active against all cell lines, compound
3being significantly more active than cisplatin and compound
4for all cell lines tested. For cell line U-937, from a non-solid
tumour (histiocytic lymphoma), complex 3is 30 times more
active than 4. These results indicate that the higher activity
found for 3could stem from a combination of a bulkier
substituent at N7 (when compared to 2) and the presence of a
hydride as co-ligand. Complexes 3and 4have different
counterions that could also contribute to the observed activity.
Thus, we synthesized complex 5via anion exchange of 3in
dichloromethane, using silver triflate (Scheme 2, see Supporting
Information for characterization), and evaluated the corre-
sponding antiproliferative activity. The IC50 values found for 5
are similar to those of 3, without significant differences (S.I.
Table S1). These results suggest that the counterion has no
significant influence on the antiproliferative activity.
Complex 3showed high cytotoxicity against four human
cancer cell lines, with IC50 values between 1.92 and 2.39 μM.
Motivated by these results, we performed further modifications
at the purine core to ascertain the overall effect on the activity
of these compounds. Variations of the steric environment at N7
at the purine backbone and at N9 were introduced to produce
complexes 6,7,[15] and 8(Figure 3). Compounds 6and 8were
synthesized by CH oxidative addition of the corresponding
ligand precursors to Pt(0), following the synthetic strategy
employed for compound 3. Complex 7was previously reported
by our group using a similar methodology.
For the synthesis of compound 6, acetate-protected
guanosine was quaternized with an isopropyl group to increase
steric crowding specifically at N7. 7-iPrGAc was reacted with
Pt(PPh3)4in dimethylformamide at 100°C for 7 h, yielding
complex 6in 68 % isolated yield (Scheme 3). Metallation at
position 8 results in a significant upfield shift of circa 1.1 ppm
for both methyl groups from the isopropyl moiety. The hydride
signal for complex 6resonates at 6.90 ppm, a shift of
0.70 ppm with respect to 4, which is indicative of a higher
electronic donation of the isopropyl group.[16] In the 13C
{1H} NMR, the carbenic C8 undergoes resonates at 181.4 ppm, in
agreement with that of previous complexes.[14]
For the synthesis of compound 8we employed 9-meth-
ylcaffeinium iodide (9-MeCaff), a more oxidized purine moiety.
9-MeCaff was then reacted with Pt(PPh3)4in dimeth-
ylformamide at 60°C for 24 h, affording complex 8in 42 %
isolated yield. Complex 8was characterized by NMR spectro-
scopy in DMSO-d6. In the 1H NMR analysis, the platinum-bonded
hydride resonates at 6.20 ppm.[15,17] In the 13C{1H} NMR analy-
sis, the metallated C8 resonates at 185.0 ppm, in line with the
values described for 3and 6. Of note, all compounds are stable
in DMSO-d6solutions for at least one week, as determined by
NMR.
Crystals of 8were obtained from a saturated solution of 8in
DMSO-d6, allowing for their characterization by single-crystal X-
ray analysis (Figure 4). The crystal structure confirms the trans
orientation for the two phosphines. The N7C8N9 angle is
106.7(5)°, identical to other metal NHCs based on caffeine.[17–19]
The PtC8 bond length is 2.064(5) Å, longer than in the related
Scheme 1. Methylation of complex 1with methyl triflate, affording the NHC
4.
Table 1. IC50 sd values (μM) for compounds 3 and 4 towards four human
cancer cell lines after incubation for 48 h, expressed as mean of at least
three separate determinations. sd standard deviation
Cell line 3 4 Cisplatin
TC-71 2.100.26 59.266.90 4.30 1.10
MV-4-11 1.920.19 32.266.75 6.39 0.80
U-937 1.970.24 58.109.20 8.23 1.01
A-172 2.390.17 23.472.70 17.38 1.60
Scheme 2. Ion-exchange reaction of complex 3(I) with AgOTf, affording
complex 5(OTf).
Figure 3. Chemical structure of complexes 68.
Scheme 3. Synthesis of complex 6by CH oxidative addition of 7-iPrGAc to
Pt0.
Chemistry—A European Journal
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (3 of 7) © 2023 Wiley-VCH GmbH
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
systems having chlorine and bromine as co-ligands,[20,21] prob-
ably due to the trans influence of the hydride.[15]
Anticancer activity
The anticancer activity of these compounds was then evaluated
in the same four cancer cell lines and in the non-cancer cell line
HEK-293. The IC50 values obtained for complexes 3and 68are
indicated in Table 2.
All the variations introduced for complexes 68improved
the activity in relation to complex 3. Specifically, the introduc-
tion of a bulkier group at N7 (the isopropyl in 6, instead of a
methyl in 3) provides a beneficial effect in the anticancer
activity. Compound 6is more active against all cancer cell lines:
although it is still cytotoxic for HEK-293 cells, the safety index
values are more favourable for complex 6than for complex 3.
Replacing the sugar with a methyl group at N9 (complex 7)
leads to an antiproliferative activity of circa two to three times
higher in the four cell lines. However, the effect is even more
prominent in non-cancer HEK-293 cells, since complex 7is
slightly more toxic for this cell line than for cancer lines TC-71
and A-172. Finally, a more oxidized purine ligand (complex 8)
had no relevant benefit. Compound 8is more active than
complex 3, but it is also more toxic for the non-cancer cell line
HEK-293, in a similar degree to that of complex 7.
All IC50 values of guanine- and caffeine-derived NHCs are
very similar for each cell line. Compound 6, bearing an
isopropyl group at N7, is slightly less active than compounds 7
and 8for all cancer cell lines. However, complex 6is
approximately three times less toxic for the non-tumour cell
line HEK-293 with an IC50 of 2.98 μM (while 7and 8showed an
IC50 of 1.10 μM and 0.99 μM against HEK-293, respectively).
Although all compounds are toxic for non-tumour cells, the
safety index is above 1 in most cases (Table 3), indicating that
they are more toxic for cancer cells than normal proliferating
cells. In particular, the presence of a bulkier group at N7 in
compound 6resulted in a more favourable ratio between
anticancer activity and a lower toxicity for non-cancer cells, a
characteristic that warrants further development.
MTT assays detect a decrease in the mitochondrial activity
indicative of a decrease in the number of viable cells. However,
it does not allow to determine whether this decrease is due to
antiproliferative or cytotoxic effects. Trypan blue exclusion assay
allows to determine the number of both dead and viable cells:
viable cells possess intact cell membranes and exclude the
trypan blue dye, whereas dead cells do not. Thus, we used
trypan blue exclusion assay to confirm that there was a
reduction in the total number of cells (Dead +viable cells,
Figure 5A) and that this reduction was due, at least in part, to
induction of cell death (% dead cells vs total number of cells,
Figure 5B). Consistently, compound 6was neither cytotoxic nor
antiproliferative for HEK-293 cells at 1 μM (Figure 5A and B). Cell
cycle distribution analysis upon treatment with compound 6or
cisplatin for 24 h indicated that, besides cytotoxicity, there is an
early antiproliferative component in the action of these drugs
(Figure 5C and D). While compound 6induced an increase in
the proportion of cells in the G0/G1 phase, cisplatin caused cell
cycle arrest at G2/M, consistent with previous reports.[4]
We next explored the antitumoral mechanism of compound
6. We evaluated oxidative stress, apoptosis, endoplasmic
reticulum stress and autophagy, four common mechanisms of
action of antitumoral drugs. We evaluated these parameters
after 24 h with compound 6, to determine the events before
the loss of viability and proliferation observed at 48 h.
Surprisingly, compound 6did not induce oxidative stress.
Instead, 6reduced significantly the intracellular levels of
reactive oxygen species in cancer cell lines, but not in normal
HEK-293 cells (Figure 5E). This decrease was accompanied by an
increase in the mitochondrial membrane potential (Figure 5F).
This difference between non-cancer and cancer cells is not
Figure 4. Molecular structure of complex 8obtained by X-Ray crystallog-
raphy (H atoms omitted for clarity).
Table 2. IC50 sd values (μM) for compounds 3and 68towards four
human cancer cell lines after incubation for 48 h, expressed as mean of at
least three separate determinations. sd standard deviation
Cell line 3678
TC-71 2.100.26 1.120.13 1.23 0.15 0.83 0.11
MV-4-11 1.920.19 1.240.19 0.79 0.10 1.21 0.23
U-937 1.970.24 0.610.09 0.61 0.08 0.73 0.08
A-172 2.390.17 1.660.30 1.49 0.25 1.48 0.16
HEK-293 4.100.15 2.980.32 1.10 0.22 0.99 0.10
Table 3. Safety index values of PtH compounds 3and 68for four cancer
cell lines, calculated from MTT assays after 48 h of incubation with the
different compounds.[a]
Cell line 3 6 7 8
TC-71 1.95 2.66 0.89 1.19
MV-4-11 2.14 2.40 1.39 0.82
U-937 2.08 4.89 1.80 1.36
A-172 1.73 1.80 0.74 0.67
[a] Range of the ratios between the IC50 for the non-cancer cell line (HEK-
293) and the IC50 value for each cancer cell line presented in Table 2.
Chemistry—A European Journal
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (4 of 7) © 2023 Wiley-VCH GmbH
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 5. Compound 6 induces reductive stress and cell death in cancer cells. Human cancer cell lines TC-71, MV-4-11, U937 and A172, and normal
HEK293 cells were incubated with compound 6 (1 μM, white bars) or the vehicle (DMSO 0.1 % v/v, black bars). Incubation of cancer cells with compound 6 for
48 h induced a decrease in total cell number (A, dead cells +viable cells) partially explained by cell death (B, % of dead cells vs total number of cells), as
determined by trypan blue assays. After 24 h of incubation, C) compound 6 induced cell cycle arrest at G0/G1 phase, while D) Cisplatin (Cis-Pt) induced cell
cycle arrest at G2/M, suggesting different mechanisms of action. At the same time point, compound 6 induced a decrease in the levels of intracellular reactive
oxygen species (ROS), as E) determined by DCFHDA staining; an increase in mitochondrial membrane potential, as F) determined by 123-rhodamine staining;
and G) an increase in intracellular glutathione (GSH), H) CHOP and LC3-II levels; but not caspase activation. HEK293 cells did not respond to compound 6 as
cancer cells, with the exception of a slight increase in CHOP levels. I, Transient transfection of cells with LC3-EGFP (green) showed a dose-dependent
enlargement of autophagosomes induced by compound 6. Nuclei were counterstained with Hoechst (blue). Scale bar, 20 μm. Results are the mean SEM of 3
independent experiments. *, significant vs. CON, p<0.05 (t-student test).
Chemistry—A European Journal
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (5 of 7) © 2023 Wiley-VCH GmbH
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
unusual. Cancer cells often show higher basal levels of ROS, due
to an altered mitochondrial function.[22,23] What is more
uncommon is the co-existence of a reductive intracellular
environment and a higher mitochondrial potential, which we
found only once in the literature in conditions of reductive
stress.[24,25] Reductive stress is defined as a condition character-
ized by excessive and/or sustained accumulation of reducing
equivalents that jeopardizes cellular survival. Recent reports
have highlighted the potential of reductive stress as a mode of
action for the development of anticancer drugs, making use of
the so-called catalytic anticancer compounds. Reductive stress
can be promoted in cancer cells by selenium-containing
metabolites[26–28] or ruthenium complexes[29] and examples with
other metals have also been reported.[30–33]
Glutathione (GSH) is the most abundant intracellular
antioxidant, and its levels are often increased in reductive
stress.[34] Consistently, we found increased intracellular GSH
levels in two selected cancer cell lines upon incubation with
compound 6, but not in HEK293 cells (Figure 5G). Moreover,
Reductive stress is often associated with other forms of stress.
In TC71 and MV-4-11 cancer cells, compound 6increased the
levels of CHOP and LC3-II, two endoplasmic reticulum stress
and autophagy markers, respectively (Figure 5H). In contrast,
HEK293 cells showed a substantially lower increase in CHOP
expression and no change in LC3-II levels. Morphological
analysis of cells transfected with LC3-EGFP and treated with
compound 6showed a dose-dependent appearance of auto-
phagosomes and other vesicular structures in the cytoplasm,
consistent with autophagy (Figure 5I). The type of cancer cell
death induced by compound 6was not apoptotic, as indicated
by the absence of caspase-3 cleavage, necessary to execute
apoptosis.
Conclusions
In summary, the combination of steric hindrance at N7 and a
hydride trans to the NHC improves cytotoxicity against four
different cancer cell lines, with the best outcome found for
compound 6, bearing an isopropyl group at N7. The trypan
blue exclusion assay shows that for that the reduction in the
number of cells is due to cell death, but cell cycle arrest at G0/
G1 phase is also observed. Compound 6induces reductive
stress on cancer cells, with an increase in GSH levels and
endoplasmic reticulum and autophagy markers on TC71 and
MV-4-11 cancer cells (Ewing’s sarcoma and myelomonocytic
leukaemia, respectively). In contrast, non-tumour cells showed
neither signs of reductive stress, nor of autophagy, and the
putative increase in endoplasmic reticulum stress was lower
than in cancer cells. These platinated nucleosides thus show
anticancer activity with a different mode of action than that of
cisplatin, hence showing great potential to develop novel
metallodrugs able to circumvent the well know resistance
associated with cisplatin.
Experimental Section
Deposition Number: 2217136 (for 8) contain(s) the supplementary
crystallographic data for this paper. These data are provided free of
charge by the joint Cambridge Crystallographic Data Centre and
Fachinformationszentrum Karlsruhe Access Structures service.
Supporting Information
All experimental procedures, NMR and mass spectra, as well as
the crystallographic data, are included in the Supporting
Information. Additional references cited within the Supporting
Information.[35–39]
Acknowledgements
We thank Fernanda Murtinheira for the support on the
autophagy measurements. This work was supported by FCT
Fundação para a Ciência e a Tecnologia, I.P., through MOSTMI-
CRO-ITQB R&D Unit (UIDB/04612/2020, UIDP/04612/2020) and
LS4FUTURE Associated Laboratory (LA/P/0087/2020). M.I.P.S.L.
was supported by fellowships PD/BD/135483/2018 and COVID/
BD/152502/2022 from FCT. A.P. acknowledges the contract
CEECINST/00102/2018. The NMR spectra were acquired at
CERMAX–ITQB, supported by Infrastructure Project No. 022161
(co-financed by FEDER through COMPETE 2020, POCI, PORL and
FCT through PIDDAC). C.S.B. Gomes acknowledges the XTAL
Macromolecular Crystallography group for granting access to
the X-ray diffractometer. X-ray infrastructure financed by FCT-
MCTES through project RECI/BBBBEP/0124/2012. FH was sup-
ported by Centre grants from FCT to the BioISI Research Unit
(Refs. UIDB/04046/2020 and UIDP/04046/2020) and the Micro-
scopy facility at FCUL (as a node of the Portuguese Platform of
BioImaging, reference PPBI-POCI-01-0145-FEDER-022122), and
by individual grants through FCT (Ref. PTDC/FIS-MAC/2741/
2021) and the ARSACS Foundation (Canada). Maria Turos-Cabal
was supported by fellowship from the Spanish Association
Against Cancer (AECC, SV-19-AECC-FPI) and the Consejería de
Economía y Empleo del Principado de Asturias (FICYT, Severo-
Ochoa BP20-073).
Conflict of Interests
The authors declare no conflict of interest.
Data Availability Statement
X-ray crystallographic data in CIF format are available from the
Cambridge Crystallographic Data Centre (Deposition Number
2217136).
Keywords: anticancer agents ·CH activation ·organometallic
nucleosides ·platinum complexes ·reductive stress
Chemistry—A European Journal
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (6 of 7) © 2023 Wiley-VCH GmbH
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[1] T. C. Johnstone, G. A. Y. Park, S. J. Lippard, Anticancer Res. 2014,34, 471–
476.
[2] S. Rottenberg, C. Disler, P. Perego, Nat. Rev. Cancer 2021,21, 37–50.
[3] S. Dilruba, G. V. Kalayda, Cancer Chemother. Pharmacol. 2016,77, 1103–
1124.
[4] C. M. Sorenson, A. Eastman, Cancer Res. 1988,48, 4484–4488.
[5] L. Kelland, Nat. Rev. Cancer 2007,7, 573–584.
[6] M. Hanif, J. Arshad, J. W. Astin, Z. Rana, A. Zafar, S. Movassaghi, E. Leung,
K. Patel, T. Söhnel, J. Reynisson, V. Sarojini, R. J. Rosengren, S. M. F.
Jamieson, C. G. Hartinger, Angew. Chem. Int. Ed. 2020,132, 14717–
14722.
[7] A. Collado, M. Gómez-Gallego, M. A. Sierra, Eur. J. Org. Chem. 2018,
2018, 1617–1623.
[8] H. Valdés, D. Canseco-González, J. M. Germán-Acacio, D. Morales-
Morales, J. Organomet. Chem. 2018,867, 51–54.
[9] K. Kowalski, Coord. Chem. Rev. 2021,432, 213705.
[10] C. M. Galmarini, J. R. Mackey, C. Dumontet, Lancet Oncol. 2002,3, 415–
424.
[11] L. P. Jordheim, D. Durantel, F. Zoulim, C. Dumontet, Nat. Rev. Drug
Discovery 2013,12, 447–464.
[12] T. Scattolin, I. Caligiuri, L. Canovese, N. Demitri, R. Gambari, I. Lampronti,
F. Rizzolio, C. Santo, F. Visentin, Dalton Trans. 2018,47, 13616–13630.
[13] Q. Peña, A. Wang, O. Zaremba, Y. Shi, H. W. Scheeren, J. M. Metselaar, F.
Kiessling, R. M. Pallares, S. Wuttke, T. Lammers, Chem. Soc. Rev. 2022,51,
2544–2582.
[14] M. I. P. S. Leitão, F. Herrera, A. Petronilho, ACS Omega 2018,3, 15653–
15656.
[15] M. I. P. S. Leitão, C. Gonzalez, G. Francescato, Z. Filipiak, A. Petronilho,
Chem. Commun. 2020,56, 13365–13368.
[16] L. Rocchigiani, J. Fernandez-Cestau, I. Chambrier, P. Hrobárik, M.
Bochmann, J. Am. Chem. Soc. 2018,140, 8287–8302.
[17] D. Chan, S. B. Duckett, S. L. Heath, I. G. Khazal, R. N. Perutz, S. Sabo-
Etienne, P. L. Timmins, Organometallics 2004,23, 5744–5756.
[18] G. Francescato, S. M. d. Silva, M. I. P. S. Leitão, A. Gaspar-Cordeiro, N.
Giannopoulos, C. S. B. Gomes, C. Pimentel, A. Petronilho, Appl. Organo-
met. Chem. 2022, e6687.
[19] V. Stoppa, T. Scattolin, M. Bevilacqua, M. Baron, C. Graiff, L. Orian, A.
Biffis, I. Menegazzo, M. Roverso, S. Bogialli, F. Visentin, C. Tubaro, New J.
Chem. 2021,45, 961–971.
[20] D. Brackemeyer, A. Hervé, C. S. to Brinke, M. C. Jahnke, F. E. Hahn, J. Am.
Chem. Soc. 2014,136, 7841–7844.
[21] F. Kampert, D. Brackemeyer, T. T. Y. Tan, F. E. Hahn, Organometallics
2018,37, 4181–4185.
[22] M. Azmanova, A. Pitto-Barry, ChemBioChem 2022,23, e202100641.
[23] J. D. Hayes, A. T. Dinkova-Kostova, K. D. Tew, Cancer Cell 2020,38, 167–
197.
[24] P. Fuchs, F. Bohle, S. Lichtenauer, J. M. Ugalde, E. Feitosa Araujo, B.
Mansuroglu, C. Ruberti, S. Wagner, S. J. Müller-Schüssele, A. J. Meyer, M.
Schwarzländer, Plant Cell 2022,34, 1375–1395.
[25] L. Zhang, K. D. Tew, Adv. Cancer Res. 2021,152, 383–413.
[26] X. Pan, X. Song, C. Wang, T. Cheng, D. Luan, K. Xu, B. Tang, Theranostics
2019,9, 1794–1808.
[27] V. V. Gandhi, P. P. Phadnis, A. Kunwar, Metallomics 2020,12, 1253–1266.
[28] V. V. Gandhi, K. A. Gandhi, L. B. Kumbhare, J. S. Goda, V. Gota, K. I.
Priyadarsini, A. Kunwar, Free Radical Biol. Med. 2021,175, 1–17.
[29] J. J. Soldevila-Barreda, I. Romero-Canelón, A. Habtemariam, P. J. Sadler,
Nat. Commun. 2015,6, 1–9.
[30] Z. Fan, J. Huang, H. Huang, S. Banerjee, ChemMedChem 2021,16, 2480–
2486.
[31] S. Banerjee, P. J. Sadler, RSC Chem. Biol. 2021,2, 12–29.
[32] K. Tyagi, T. Dixit, V. Venkatesh, Inorg. Chim. Acta 2022,533, 120754.
[33] S. Infante-Tadeo, V. Rodríguez-Fanjul, A. Habtemariam, A. M. Pizarro,
Chem. Sci. 2021,12, 9287–9297.
[34] H. Zhang, P. Limphong, J. Pieper, Q. Liu, C. K. Rodesch, E. Christians, I. J.
Benjamin, FASEB J. 2012,26, 1442–1451.
[35] F.-T. Luo, H.-K. Lo, J. Organomet. Chem. 2011,696, 1262–1265.
[36] T. M. Barbosa, R. Rittner, K. Alexander, R. Cosstick, R. J. Abraham,
Nucleosides Nucleotides Nucleic Acids 2016,35, 53–63.
[37] L. Krause, R. Herbst-Irmer, G. M. Sheldrick, D. Stalke, J. Appl. Crystallogr.
2015,48, 3–10.
[38] G. M. Sheldrick, Acta Crystallogr. Sect. C Struct. Chem. 2015,71, 3–8.
[39] L. J. Farrugia, J. Appl. Crystallogr. 2012,45, 849–854.
Manuscript received: April 4, 2023
Accepted manuscript online: April 27, 2023
Version of record online: ■■,
Chemistry—A European Journal
Research Article
doi.org/10.1002/chem.202301078
Chem. Eur. J. 2023, e202301078 (7 of 7) © 2023 Wiley-VCH GmbH
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
RESEARCH ARTICLE
Platinum(II) complexes bearing NHCs
based on guanosine undergo a sub-
stantial increase in antiproliferative
activity when changing the ligand
trans to the NHC from bromide to
hydride. Compound 6leads to an
increase in reductive stress and
increase in glutathione levels in
cancer cells but not in non-cancer
cells.
Dr. M. I. P. S. Leitão, M. Turos-Cabal,
Dr. A. M. Sanchez-Sanchez, Dr. C. S. B.
Gomes, Prof. F. Herrera*, Prof. V.
Martin*, Dr. A. Petronilho*
1 8
Antitumor Activity and Reductive
Stress by Platinum(II) N-Heterocyclic
Carbenes based on Guanosine
15213765, 0, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202301078 by Cochrane Portugal, Wiley Online Library on [16/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
... Similarly, organorhodium(III) complexes facilitated hydride transfer under physiological conditions, effectively inducing RS in ovarian cancer cells [165]. More recently, platinum(II) N-heterocyclic carbene-based organometallic nucleosides have been found to selectively increase GSH levels and induce RS across multiple cancer cell lines [166]. ...
Article
Full-text available
Reductive stress (RS), characterized by excessive accumulation of reducing equivalents such as NADH and NADPH, is emerging as a key factor in metabolic disorders and cancer. While oxidative stress (OS) has been widely studied, RS and its complex interplay with endocrine regulation remain less understood. This review explores molecular circuits of bidirectional crosstalk between metabolic hormones and RS, focusing on their role in diabetes, obesity, cardiovascular diseases, and cancer. RS disrupts insulin secretion and signaling, exacerbates metabolic inflammation, and contributes to adipose tissue dysfunction, ultimately promoting insulin resistance. In cardiovascular diseases, RS alters vascular smooth muscle cell function and myocardial metabolism, influencing ischemia-reperfusion injury outcomes. In cancer, RS plays a dual role: it enhances tumor survival by buffering OS and promoting metabolic reprogramming, yet excessive RS can trigger proteotoxicity and mitochondrial dysfunction, leading to apoptosis. Recent studies have identified RS-targeting strategies, including redox-modulating therapies, nanomedicine, and drug repurposing, offering potential for novel treatments. However, challenges remain, particularly in distinguishing physiological RS from pathological conditions and in overcoming therapy-induced resistance. Future research should focus on developing selective RS biomarkers, optimizing therapeutic interventions, and exploring the role of RS in immune and endocrine regulation.
... Currently known forms of RCD, including apoptosis, autophagy, ferroptosis, and necroptosis, each possesses specific molecular mechanisms and commonly exists in many human diseases, especially in cancer advancement. [28][29][30][31][32] The biological mechanisms of metal-NHC complexes are highly dependent on the type of metal ion, although other RCD subroutines, such as autophagy and necroptosis, [33][34][35][36][37][38][39] can also be observed, apoptosis induced by metal-NHCs is most common. Once transported into cells via organic cation transporters, metal-NHCs can cause a variety of changes, such as cell cycle arrest and mitochondrial dysfunction, leading to apoptosis by altering mitochondrial membrane potential (MMP), generating reactive oxygen species (ROS), releasing cytochrome c (cyt c), and activating caspases. ...
Article
Full-text available
Metal complexes based on N‐heterocyclic carbene (NHC) ligands have emerged as promising broad‐spectrum antitumor agents in bioorganometallic medicinal chemistry. In recent decades, studies on cytotoxic metal–NHC complexes have yielded numerous compounds exhibiting superior cytotoxicity compared to cisplatin. Although the molecular mechanisms of these anticancer complexes are not fully understood, some potential targets and modes of action have been identified. However, a comprehensive review of their biological mechanisms is currently absent. In general, apoptosis caused by metal–NHCs is common in tumor cells. They can cause a series of changes after entering cells, such as mitochondrial membrane potential (MMP) variation, reactive oxygen species (ROS) generation, cytochrome c (cyt c) release, endoplasmic reticulum (ER) stress, lysosome damage, and caspase activation, ultimately leading to apoptosis. Therefore, a detailed understanding of the influence of metal–NHCs on cancer cell apoptosis is crucial. In this review, we provide a comprehensive summary of recent advances in metal–NHC complexes that trigger apoptotic cell death via different apoptosis‐related targets or signaling pathways, including B‐cell lymphoma 2 (Bcl‐2 family), p53, cyt c, ER stress, lysosome damage, thioredoxin reductase (TrxR) inhibition, and so forth. We also discuss the challenges, limitations, and future directions of metal–NHC complexes to elucidate their emerging application in medicinal chemistry.
Article
Full-text available
Xanthines are purine derivatives predominantly found in plants. These include compounds such as caffeine, theophylline, and theobromine and exhibit a variety of pharmacological properties, demonstrating efficacy in treating neurodegenerative disorders, respiratory dysfunctions, and also cancer. The versatile attributes of these materials render them privileged scaffolds for the development of compounds for various biological applications. Xanthines are N‐heterocyclic carbene precursors that combine a pyrimidine and an imidazole ring. Owing to their biological relevance, xanthines have been employed as N‐heterocyclic carbenes in the development of metallodrugs for anticancer and antimicrobial purposes. In this conceptual review, we examine key examples of N‐heterocyclic carbene complexes derived from caffeine and other xanthines, elucidating their synthetic methods and describing their pertinent medicinal applications.
Article
Full-text available
Invasive fungal diseases affect more than two million people worldwide. The increasing incidence of invasive fungal infections is the result of many factors, including an increase in the resistance to current drugs. As such, there is an urgent need to obtain new drugs that are efficient, selective, and able to overcome existing resistance mechanisms. Candida yeasts are responsible for more than 70% of all nosocomial invasive fungal diseases. In this work, we describe the synthesis of nickel (II)(NHC) complexes based on xanthines, by direct metalation of xanthinium salts with nickelocene NiCp2to yield [NiCpI(NHC)] complexes. For methyl caffeine, a biscarbene complex [NiCp(NHC)2]⁺is also formed, resulting from carbene dissociation from the corresponding monocarbene. [NiCpI(NHC)] complexes are active as antifungals for Candida yeasts and show toxicity for human cells (HeLa) that is dependent on the substitution of N7 of the xanthine moiety. The biscarbene complex 5[NiCp(NHC)2]⁺is highly selective for Candida glabrata and shows very low toxicity for human cells, being a promising candidate for selective treatment of C. glabrata infections.
Article
Full-text available
Redox processes are at the heart of universal life processes, such as metabolism, signaling or folding of secreted proteins. Redox landscapes differ between cell compartments and are strictly controlled to tolerate changing conditions and to avoid cell dysfunction. While a sophisticated antioxidant network counteracts oxidative stress, our understanding of reductive stress responses remains fragmentary. Here, we observed root growth impairment in Arabidopsis thaliana mutants of mitochondrial alternative oxidase 1a (aox1a) in response to the model thiol reductant dithiothreitol (DTT). Mutants of mitochondrial uncoupling protein 1 (ucp1) displayed a similar phenotype indicating that impaired respiratory flexibility led to hypersensitivity. Endoplasmic reticulum (ER) stress was enhanced in the mitochondrial mutants and limiting endoplasmic reticulum oxidoreductin (ERO) capacity in the aox1a background led to synergistic root growth impairment by DTT, indicating that mitochondrial respiration alleviates reductive ER stress. The observations that DTT triggered NAD reduction in vivo and that the presence of thiols led to electron transport chain activity in isolated mitochondria offer a biochemical framework of mitochondrion-mediated alleviation of thiol-mediated reductive stress. Ablation of transcription factor ANAC017 impaired the induction of AOX1a expression by DTT and led to DTT hypersensitivity, revealing that reductive stress tolerance is achieved by adjusting mitochondrial respiratory capacity via retrograde signaling. Our data reveal an unexpected role for mitochondrial respiratory flexibility and retrograde signaling in reductive stress tolerance involving inter-organelle redox crosstalk.
Article
Full-text available
Excessive cellular oxidative stress is widely perceived as a key factor in pathophysiological conditions and cancer development. Healthy cells use several mechanisms to maintain intracellular levels of reactive oxygen species (ROS) and overall redox homeostasis to avoid damage to DNA, proteins, and lipids. Cancer cells, in contrast, exhibit elevated ROS levels and upregulated protective antioxidant pathways. Counterintuitively, such elevated oxidative stress and enhanced antioxidant defence mechanisms in cancer cells provide a therapeutic opportunity for the development of drugs with different anticancer mechanisms of action (MoA). In this review, oxidative stress and the role of ROS in cells are described. The tumour‐suppressive and tumour‐promotive functions of ROS are discussed, and these two different therapeutic strategies (increasing or decreasing ROS to fight cancer) are compared. Clinically approved drugs with demonstrated oxidative stress anticancer MoAs are highlighted followed by description of examples of metal‐based anticancer drug candidates causing oxidative stress in cancer cells via novel MoAs.
Article
Full-text available
Aquation is often acknowledged as a necessary step for metallodrug activity inside the cell. Hemilabile ligands can be used for reversible metallodrug activation. We report a new family of osmium(ii) arene complexes of formula [Os(η6-C6H5(CH2)3OH)(XY)Cl]+/0 (1-13) bearing the hemilabile η6-bound arene 3-phenylpropanol, where XY is a neutral N,N or an anionic N,O- bidentate chelating ligand. Os-Cl bond cleavage in water leads to the formation of the hydroxido/aqua adduct, Os-OH(H). In spite of being considered inert, the hydroxido adduct unexpectedly triggers rapid tether ring formation by attachment of the pendant alcohol-oxygen to the osmium centre, resulting in the alkoxy tethered complex [Os(η6-arene-O-κ1)(XY)] n+. Complexes 1C-13C of formula [Os(η6:κ1-C6H5(CH2)3OH/O)(XY)]+ are fully characterised, including the X-ray structure of cation 3C. Tether-ring formation is reversible and pH dependent. Osmium complexes bearing picolinate N,O-chelates (9-12) catalyse the hydrogenation of pyruvate to lactate. Intracellular lactate production upon co-incubation of complex 11 (XY = 4-Me-picolinate) with formate has been quantified inside MDA-MB-231 and MCF7 breast cancer cells. The tether Os-arene complexes presented here can be exploited for the intracellular conversion of metabolites that are essential in the intricate metabolism of the cancer cell.
Article
Full-text available
Considering the high increase in mortality caused by cancer in recent years, cancer drugs with novel mechanisms of anticancer action are urgently needed to overcome the drawbacks of platinum‐based chemotherapeutics. Recently, in the area of metal‐based cancer drug development research, the concept of catalytic cancer drugs has been introduced with organometallic RuII, OsII, RhIII and IrIII complexes. These complexes are reported as catalysts for many important biological transformations in cancer cells such as nicotinamide adenine dinucleotide (NAD(P)H) oxidation to NAD⁺, reduction of NAD⁺ to NADH, and reduction of pyruvate to lactate. These unnatural intracellular transformations with catalytic and nontoxic doses of metal complexes are known to severely perturb several important biochemical pathways and could be the antecedent of next‐generation catalytic cancer drug development. In this concept, we delineate the prospects of such recently reported organometallic RuII, OsII, RhIII and IrIII complexes as future catalytic cancer drugs. This new approach has the potential to deliver new cancer drug candidates.
Article
Metal complexes are extensively used for cancer therapy. The multiple variables available for tuning (metal, ligand, and metal-ligand interaction) offer unique opportunities for drug design, and have led to a vast portfolio of metallodrugs that can display a higher diversity of functions and mechanisms of action with respect to pure organic structures. Clinically approved metallodrugs, such as cisplatin, carboplatin and oxaliplatin, are used to treat many types of cancer and play prominent roles in combination regimens, including with immunotherapy. However, metallodrugs generally suffer from poor pharmacokinetics, low levels of target site accumulation, metal-mediated off-target reactivity and development of drug resistance, which can all limit their efficacy and clinical translation. Nanomedicine has arisen as a powerful tool to help overcome these shortcomings. Several nanoformulations have already significantly improved the efficacy and reduced the toxicity of (chemo-)therapeutic drugs, including some promising metallodrug-containing nanomedicines currently in clinical trials. In this critical review, we analyse the opportunities and clinical challenges of metallodrugs, and we assess the advantages and limitations of metallodrug delivery, both from a nanocarrier and from a metal-nano interaction perspective. We describe the latest and most relevant nanomedicine formulations developed for metal complexes, and we discuss how the rational combination of coordination chemistry with nanomedicine technology can assist in promoting the clinical translation of metallodrugs.
Article
Catalysis, a rapidly growing field of science and technology, has brought about a paradigm shift in the direction to achieve desirable synthetic transformations utilizing numerous organometallic compounds. These organometallic complexes impart great activity not only by metal and ligand framework, however, the subtle alterations take a dramatic change in its activity, which knocks the door of their applications in the arena of drugs discovery. In this review, we discussed about the recent trends (2015 onwards) in catalytic anticancer drugs and their mechanism of action against different potential targets (Fig. 1). We have comprehensively discussed about the transfer hydrogenation catalysis including, NADH/NAD⁺ (Nicotinamide adenine dinucleotide) imbalance and pyruvate reduction. We briefly explained the important role of different catalysts (such as photocatalysts, and bioorthogonal catalysts) in the context of cancer therapeutics. Finally, we shed light on the advantages, limitations, and future scope for designing effective catalytic anticancer agents.
Article
The aim of present study was to investigate the anticancer mechanisms of 3, 3′- diselenodipropionic acid (DSePA), a redox-active organodiselenide in human lung cancer cells. DSePA elicited a significant concentration and time-dependent cytotoxicity in human lung cancer cell line A549 than in normal WI98 cells. The cytotoxic effect of DSePA was preceded by an acute decrease in the level of basal reactive oxygen species (ROS) and a concurrent increase in levels of reducing equivalents (like GSH/GSSG and NADH/NAD) within cells. Further, a series of experiments were performed to measure the markers of intrinsic (Bax, cytochrome c and caspase-9), extrinsic (TNFR, FADR and caspase-8) and endoplasmic reticulum (ER) stress (protein ubiquitylation, calcium flux, Bip-2, CHOP and caspase-12) pathways in DSePA treated cells. DSePA treatment significantly increased the levels of all the above markers. Moreover, DSePA did not alter the expression and phosphorylation (Ser15) of p53 but caused a significant damage to mitochondria. Pharmacological modulation of GSH level by BSO and NAC in DSePA treated cells led to abrogation and augmentation of cell kill respectively. This established the role of reductive stress as a trigger for the apoptosis induced by DSePA treatment. Finally, in vitro anticancer activity of DSePA was also corroborated by its in vivo efficacy of suppressing the growth of A549 derived xenograft tumor in SCID mice. In conclusion, above results suggest that DSePA induces apoptosis in a p53 independent manner by involving extrinsic and intrinsic pathways together with ER stress which can an interesting strategy for lung cancer therapy.
Chapter
Reductive stress is defined as a condition characterized by excess accumulation of reducing equivalents (e.g., NADH, NADPH, GSH), surpassing the activity of endogenous oxidoreductases. Excessive reducing equivalents can perturb cell signaling pathways, change the formation of disulfide bonding in proteins, disturb mitochondrial homeostasis or decrease metabolism. Reductive stress is influenced by cellular antioxidant load, its flux and a subverted homeostasis that paradoxically can result in excess ROS induction. Balanced reducing equivalents and antioxidant enzymes that contribute to reductive stress can be regulated by Nrf2, typically considered as an oxidative stress induced transcription factor. Cancer cells may coordinate distinct pools of redox couples under reductive stress and these may link to biological consequences from both molecular and translational standpoints. In cancer, there is recent interest in understanding how selective induction of reductive stress may influence therapeutic management and disease progression.
Article
Two new gold(I) complexes with a di(N-heterocyclic carbene) ligand (diNHC) derived from caffeine have been synthesised by a base-assisted metalation of the appropriate di(azolium) salt in the presence of the gold precursor AuCl(SMe2). Under kinetically controlled conditions, the reaction affords a mononuclear cationic complex with the diNHC ligand chelating the gold centre, while the thermodynamically more stable complex is a dinuclear species with two bridging diNHC ligands. The two complexes have been characterised in solution by mass spectrometry, and 1H, 13C and DOSY NMR spectroscopies; the mononuclear–dinuclear transformation has been also followed with kinetic experiments giving a second-order rate constant k = (1.46 ± 0.01) × 10−2 dm3 mol−1 s−1 in CD3CN, at 313 K. Density functional theory (DFT) calculations have been performed to support these findings. The reactivity of the gold(I) complexes has been evaluated in the oxidative addition of halogens. The reaction allows accessing the corresponding mono- and dinuclear gold(III) complexes, which are stable and do not interconvert. With the mononuclear complex, species of general formula [AuX2(diNHC)](BF4) (X = Cl and I) have been isolated and a systematic structural investigation of the possible isomers has been performed through DFT calculations.