Next Article in Journal
Synthesis of Medium-Sized Heterocycles by Transition-Metal-Catalyzed Intramolecular Cyclization
Next Article in Special Issue
Syntheses and Reactivity of New Zwitterionic Imidazolium Trihydridoborate and Triphenylborate Species
Previous Article in Journal
Potential Therapeutic Targets of Epigallocatechin Gallate (EGCG), the Most Abundant Catechin in Green Tea, and Its Role in the Therapy of Various Types of Cancer
Previous Article in Special Issue
Hybrid Gold(I) NHC-Artemether Complexes to Target Falciparum Malaria Parasites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-Heterocyclic Carbene Platinum(IV) as Metallodrug Candidates: Synthesis and 195Pt NMR Chemical Shift Trend

by
Mathilde Bouché
1,
Bruno Vincent
2,
Thierry Achard
1 and
Stéphane Bellemin-Laponnaz
1,*
1
Institut de Physique et Chimie des Matériaux de Strasbourg, Université de Strasbourg-CNRS UMR7504, 23 rue du Loess, BP 43 CEDEX 2, 67034 Strasbourg, France
2
Service de RMN, Fédération de Chimie Le Bel, Université de Strasbourg, CNRS FR2010, 1 rue Blaise Pascal, BP 296R8, CEDEX, 67008 Strasbourg, France
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(14), 3148; https://doi.org/10.3390/molecules25143148
Submission received: 19 June 2020 / Revised: 6 July 2020 / Accepted: 8 July 2020 / Published: 9 July 2020
(This article belongs to the Special Issue Carbon Ligands: From Fundamental Aspects to Applications)

Abstract

:
A series of octahedral platinum(IV) complexes functionalized with both N-heterocyclic carbene (NHC) ligands were synthesized according to a straightforward procedure and characterized. The coordination sphere around the metal was varied, investigating the influence of the substituted NHC and the amine ligand in trans position to the NHC. The influence of those structural variations on the chemical shift of the platinum center were evaluated by 195Pt NMR. This spectroscopy provided more insights on the impact of the structural changes on the electronic density at the platinum center. Investigation of the in vitro cytotoxicities of representative complexes were carried on three cancer cell lines and showed IC50 values down to the low micromolar range that compare favorably with the benchmark cisplatin or their platinum(II) counterparts bearing NHC ligands.

1. Introduction

In the treatment of solid tumors, platinum-based chemotherapy remains a top-notch drug thanks to its high anticancer efficiency and high remission rates in selected cancers, up to 90% in patients suffering testicular cancer [1]. However, the severe systemic toxicity and poor selectivity for tumors, with only 1% of the injected dose of cisplatin actually reaching their target, stress the importance to explore new strategies to stabilize the platinum center and prevent off-target interactions [2]. Lot of efforts have focused on adjusting the coordination sphere of the platinum or oxidizing the well established Pt(II) center into redox-activable Pt(IV) pro-drugs [3,4,5]. In particular, platinum complexes functionalized with N-heterocyclic carbenes have appeared as promising alternatives to cisplatin, as possible chemotherapeutics targeting the mitochondria [6,7,8,9]. In addition to investigating the redox potential of Pt(IV) complexes, 195Pt NMR is a valuable technique for the fast and standardized characterization of platinum complexes to complement routine characterization [10,11,12,13,14,15]. Moreover, investigating the chemical shift in 195Pt NMR is a prerequisite to further enable mechanistic and pharmacokinetic investigations of the platinum complex and its metabolites using 195Pt spectroscopy [16], as an alternative to LA-ICP-MS [17] or X-ray-based techniques [18,19,20,21]. Of note, Huynh et al. recently suggested the direct correlation of the platinum chemical shift in 195Pt NMR to both the electronic density at the platinum center and the electronic donation of the coordination sphere, in particular in the case of N-heterocyclic carbene (NHC)-platinum complexes [22,23,24]. Moreover, a linear correlation of the 195Pt NMR chemical shift with the in vitro anticancer activity (IC50) has been noted in azido-Pt(IV) complexes [25]. The chemistry in the solution of cisplatin and its derivatives have been studied by 195Pt NMR spectroscopy. In particular, they have been used to characterize related complexes with aqua, chloro, nitrato, sulfato, acetate, and phosphate ligands [26,27]. Therefore, we report herein a series of NHC-Pt(IV) complexes and a few examples of their Pt(II) metabolites possibly formed in vitro, that were synthesized and characterized using routine techniques. In vitro activities against three cancer cell lines of representative NHC-Pt(IV) complexes are also presented. Moreover, the shift in the platinum resonance signal in 195Pt NMR is investigated and discussed as a function of tuning their oxidation degree and coordination sphere.

2. Results and Discussion

2.1. Synthesis of the Platinum(II) and Platinum(IV) Complexes

All NHC-Pt complexes were prepared using standard synthetic procedures as previously reported. The general scheme for the synthesis of the Pt(II) and Pt(VI) complexes is described in Scheme 1. First, platinum(II) NHC pyridine complexes were synthesized involving the in situ deprotonation of the imidazolium salt with K2CO3 and the coordination of the carbene to the PtCl2 precursor in dry amine with excess NaI overnight ((1), first step, Scheme 1) [28]. Chemical variation was then possible by the ligand substitution of the pyridine with various nitrogen-based ligands as shown in (1), second step, Scheme 1. The obtained (NHC)PtI2(pyridine) and (NHC)PtI2(amine) complexes could further be oxidized according to a procedure previously reported by us [9]. The aforementioned Pt(II) complexes were reacted with a 10-fold excess of bromine at 0 °C to obtain the corresponding (NHC)PtBr4(L) complexes ((2), Scheme 1). The reaction proceeded very quickly and cleanly to give the expected corresponding Pt(IV) species after only 5 min of reaction. The chlorinated complexes ((NHC)PtCl4L) were obtained by direct oxidation using a 2-fold excess of freshly prepared hypervalent iodine reagent PhICl2 ((3), Scheme 1). The reaction was complete after 1 h at 0 °C. All the platinum(IV) complexes were easily isolated by precipitation with pentane. They were usually obtained in high chemical yield and were stable under air in the solid state or in chlorinated solvents and showed increasing solubility in organic solvents in respect to the length of alkyl chains on the NHC or amine ligand.
Scheme 2 displays the molecular structures of the five platinum(II) NHC complexes used either as precursors for the Pt(IV) syntheses, or as a reference for the studies discussed here. The (NHC)Pt(II)(DMSO) complexes 3 and 4 were obtained by a transmetallation route from the bis(benzyl)imidazol-2-ylidene silver(I) bromide precursor reacted with platinum salt as previously published by us [7]. The NHC Pt(II) complexes 2 and 5 were obtained using the procedure described in (1), Scheme 1 using the corresponding salt NaBr or NaCl respectively.
Scheme 3 displays all the platinum(IV) that were synthesized and characterized. A series of (NHC)PtBr4(amine) complexes bearing a (methyl-, benzyl-)NHC, were obtained in a 99% yield with various trans amine ligands, i.e., dodecylamine, cyclohexylamine, morpholine and pyridine, corresponding to complexes 6, 8, 12 and 18, respectively. Identically, the (NHC)PtCl4(amine) complexes with varying amine ligands were obtained in good yields, the corresponding amine ligand being a cyclohexylamine for 22, a morpholine for 23 and a pyridine for 26. The versatile synthesis tolerated the NHC structural variations among the (NHC)PtBr4(amine) family, with N-substituents being a CH2-tert-butylacetate for 14, p-nitro-benzyl for 15, p-benzaldehyde for 16, a pentyl for 19, a cyclopentyl for 20 and a phenyl for 21, all obtained in 99% yield. The functionalization of the positions 4 and 5 of the NHC ligand did not hamper the oxidation reaction, and the (NHC)PtBr4(amine) complexes 9, 11, 13 and 17 were isolated in high yield, corresponding respectively to a benzimidazole, 4-metyl- and 5-aldehyde, 4-methylester, and 4,5-dichloro-NHC. Similarly, the (NHC)PtCl4(amine) complexes 24 bearing a 4,5-dichloro-NHC and 25 functionalized with a pentyl-N-substituted NHC were obtained also in a yield up to 99%. The characterization by the 1H NMR showed that all the proton signals displayed a shift to a lower field compared to their imidazolium precursors which proved typical for such complexes. Overall, the NHC-Pt(IV) complexes showed a signal duplication typically observed for all the protons in up to the 5J position to the platinum center, suggesting an enhanced coupling with the 195Pt isotope compared to their NHC-Pt(II) precursors. Of note, the very low solubility of (NHC)PtBr4(pyridine) complexes prevented the successful acquisition of the 13C NMR of complexes 4, 5, 11, 20 and 26, or rendered the carbenic carbon signal not visible. However, in the case of more lipophilic complexes, coupling between the carbenic carbon and the platinum center was observed in 13C NMR. Such a trend was found typical throughout all the NHC-Pt(IV) complexes, the carbenic carbon signal appearing as a singlet and doublet system, possibly due to the heavy atom effect of platinum [29]. Moreover, chemical shifts to a higher field of the carbenic carbon were also observed by 13C NMR spectroscopy, ca. δ 109–120 ppm in the case of NHC-Pt(IV) complexes, while (NHC)PtI2(amine) complexes previously reported by us [30,31] and others [32,33] typically show a signal shift at least 30 ppm greater.
Among these Pt(IV) complexes, the molecular structure of the (NHC)PtBr4(amine) complex 15 was determined by X-ray diffraction and is presented in Figure 1. The platinum center shows an octahedral geometry with bromine ligands forming a distorted square planar shape in equatorial position, comparable to other (NHC)PtBr4(amine) complexes previously reported by us [6,7]. The pyridine ligand is located in trans position to the NHC with a platinum-pyridine length of 2.128(6) Å while the NHC-platinum bond is found to be 2.057(8) Å. The molecular structure of the (NHC)PtCl4(amine) complex 23 revealed a comparable geometry with overall shorter bonds between the platinum center and the ligands, reflective of the influence of the coordination sphere on platinum’s electronic density, exemplified by the NHC-platinum length of 2.034(3) Å, and the chloride-platinum bonds in the range of 2.327(3)–2.336(3) Å [7].

2.2. In Vitro Activities against Cancer Cell Lines

Among the series of NHC-platinum complexes herein, a series of the most soluble complexes were selected for the evaluation of their in vitro anticancer activities. Overall, most NHC-Pt(IV) complexes were found to display comparable IC50 values to cisplatin in the range of 0.5–23 µM. Contrastingly, the complex 16 showed disparate anticancer activities depending on the cancer cell line with the IC50 values of 5.42 µM and 81.09 µM against the PC3 or HCT116 respectively (Table 1). Of note, the low solubility of this complex in aqueous media might explain the low IC50 values observed in this study. The series of the (NHC)PtBr4(amine) complexes 6, 8, 12 and 19 show potencies that compare favorably with the NHC-Pt(II) complexes which are expected to be the species released upon their redox activation. Such a result is in line with our previous findings suggesting their rapid reduction and release of the active species [6,7]. Remarkably, the (NHC)PtCl4(amine) complexes 22 and 25 show the most promising in vitro potencies with IC50 values in the low micromolar range against the three tested cancer cell lines.

2.3. 195Pt NMR Spectroscopy

The NHC-platinum complexes were further characterized using a 1H detection inverse NMR spectroscopy sequence which was preferred to direct the 195Pt measurement in regard of shorter acquisition time and enhanced sensibility. This was supported by a test experiment using complex 8 as a reference, comparing spectra obtained in direct 195Pt NMR or indirect HMQC 1H-195Pt NMR, and both showed a signal peak at δPt = −2168 ppm irrespective of the sequence used. Table 2 displays the 195Pt chemical shift NMR of all the complexes and carbenic carbon signal in the 13C NMR, when observed. The most significant variation in the platinum chemical shift was found as a function of the oxidation state of the platinum center. All the (NHC)PtBr4(amine) complexes 621 displayed a platinum chemical shift in the range of δPt −1901 to −2196 ppm while the (NHC)PtCl4(amine) complexes 2226 were observed at δPt −883 to −795 ppm and all other NHC-Pt(II) complexes 15 displayed a chemical shift below −3304 ppm.
Of note, the use of 1H detection inverse spectroscopy proved of high interest for most complexes to observe the 4JH-Pt long-range couplings between the platinum center and C3, C4 protons on the NHC backbone as well as the protons on the N-substituents of the NHC. This strong chemical coupling suggests a high electronic delocalization from the platinum center to the substituents of the NHC ligand which yet does not seem to significantly affect the chemical shift in 195Pt NMR. Thus, the series of NHC-Pt(IV) complexes 1416, 19 and 20 show a platinum chemical shift decrease from δPt = −2032 to −2070 ppm with the N-substituents following the trend Cy > C5H11 > Bn > CH2CO2tBu. Similarly, the functionalization of C3 and C4 positions on the NHC backbone of the NHC-Pt(IV) complexes is shown to have a negligible effect on the platinum shift with ΔδPt = 2 ppm between complexes 13 and 11. Moreover, a large platinum chemical shift variation ΔδPt = 64 ppm was observed between the imidazolin-2-ylidene ligand in 16Pt −2063 ppm) and the benzimidazolin-2-ylidene ligand in 9Pt −2127 ppm), which was found to correlate with the ΔδC = 23.1 ppm of their carbenic carbon observed by 13C NMR. Among the series of the NHC-Pt(IV) complexes, the variation of the trans amine ligand shows a trend in the platinum chemical shift that follows the amine’s basicity from δPt −2040 ppm for complex 18 to δPt −2196 ppm for complex 6. Thus, the trend in the platinum chemical shift is found to be 18 > 12 > 8 > 6, corresponding to a trans ligand being pyridine > morpholine > cyclohexylamine > dodecylamine. Of note, the same trend is visible while comparing their carbenic carbon shift as complex 18 bearing a pyridine shows a δC of 109.3 ppm while its cyclohexylamine counterpart 8 shows a shift up to 115.2 ppm. Moreover, the (NHC)PtCl4(amine) complexes follow the same trend with platinum chemical shifts being 26 > 23 > 22, corresponding to the trans amine ligand pyridine > morpholine > cyclohexylamine.

3. Materials and Methods

All the manipulations of the air- and moisture-sensitive compounds were carried out using standard Schlenk techniques under an argon atmosphere and the solvents were purified and degassed following standard procedures. All the reagents were purchased from commercial chemical suppliers (Acros (Illkirch, France), Alfa Aesar (Lancashire, UK), and TCI Europe (Paris, France)) and used without further purification. 1H and 13C nuclear magnetic resonance (NMR) spectra were recorded on a Brucker AVANCE 300 or Bruker AVANCE 500 spectrometer (Bruker, Wissembourg, France) using the residual solvent peak as a reference (CDCl3: δH = 7.26 ppm; δC = 77.16 ppm) at 295 K. The HMQC 1H-195Pt spectra were recorded on a Bruker AVANCE 600 spectrometer using the residual solvent peak as reference for the 1H calibration and an external reference for the 195Pt (H2PtCl6 in D2O: δPt = 0 ppm) at the Institut de Chimie NMR Facility of the University of Strasbourg. Positive mode electrospray ionization mass spectra (ESI-HRMS) analyses were carried out on microTOF, Bruker Daltonics (Bruker, Wissembourg, France).
All the syntheses and characterizations are available in the Supplementary Materials.

4. Conclusions

In the present work, a series of N-heterocyclic carbene-coordinated platinum(IV) complexes were synthesized in high yield according to a versatile procedure. All the complexes were found stable in the air and in chlorinated solvents for months. Some representative examples of these NHC-Pt(IV) complexes were selected for the in vitro evaluation of their cancer inhibitory properties and compared to their possible Pt(II) metabolites formed in the biological environment. Overall, the lipophilic (NHC)PtCl4(amine) complex 22 was found to induce the greater in vitro potencies toward selected cancer cell lines with IC50 values in the low micromolar range.
In the development of platinum-based metallodrugs, numerous parameters have to be considered in addition to the apparent electronic density at the platinum center that may be reflected by the 195Pt NMR chemical shift, namely lipophilicity and pharmacological properties and so forth. Moreover, the balance between the stability of the platinum drugs in the blood stream and their ability to form metabolites and interact with DNA is difficult to anticipate by finetuning the coordination sphere of the platinum. However, the 195Pt NMR has proved to be a helpful probe in investigating the biological activity of platinum-based drugs. For example, a recent study involving the monitoring of carboplatin after subcutaneous injection in rats was studied using 195Pt NMR [34]. Thus, all the complexes presented here were characterized with standard techniques and the influence of structural variations, i.e., on one hand the coordination sphere and on the other hand the NHC ligand’s functionalization, were correlated to their chemical shift in 195Pt NMR. All the (NHC)PtBr4(amine) complexes displayed platinum chemical shifts in the range of δPt −1900 to −2200 ppm while the (NHC)PtCl4(amine) complexes were observed at δPt −900 to −800 ppm. All other NHC–Pt(II) complexes displayed a chemical shift below −3304 ppm. The 195Pt NMR spectroscopy could then be used to monitor the kinetics and the mechanism of such platinum complexes with biological substances.

Supplementary Materials

The following are available online. 195Pt NMR spectra and characterization for all compounds.

Author Contributions

S.B.-L. designed the research. M.B., T.A. and S.B.-L. conceived, designed and performed the chemical experiments. B.V. performed the NMR experiments. S.B.-L. and M.B. wrote the paper and T.A. and B.V. participated in manuscript writing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Strasbourg/CNRS-Program IDEX Interdisciplinaire. M.B. was granted by the French “Ministère de la Recherche”.

Acknowledgments

The authors gratefully acknowledge the Ministère de l’Enseignement Supérieur et de la Recherche for Ph.D. grants to M.B. Biological evaluations of cell proliferation inhibition have been performed at the Ciblothèque Cellulaire ICSN (Gif sur Yvette, France). The authors also thank Michel Sigrist for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gietema, J.A.; Meinardi, M.T.; Messerschmidt, J.; Gelevert, T.; Alt, F.; Uges, D.; Seijfer, D.T. Circulating plasma platinum more than 10 years after cisplatin treatment for testicular cancer. Lancet 2000, 355, 1075–1076. [Google Scholar] [CrossRef]
  2. Cheff, D.M.; Hall, M.D. A Drug of Such Damned Nature. 1 Challenges and Opportunities in Translational Platinum Drug Research: Miniperspective. J. Med. Chem. 2017, 60, 4517–4532. [Google Scholar] [CrossRef] [PubMed]
  3. Um, I.S.; Armstrong-Gordon, E.; Moussa, Y.E.; Gnjidic, D.; Wheate, N.J. Platinum drugs in the Australian cancer chemotherapy healthcare setting: Is it worthwhile for chemists to continue to develop platinums? Inorg. Chim. Acta 2019, 492, 177–181. [Google Scholar] [CrossRef]
  4. Gibson, D. Multi-action Pt (IV) anticancer agents; do we understand how they work? J. Inorg. Biochem. 2019, 191, 77–84. [Google Scholar] [CrossRef]
  5. Hall, M.D.; Hambley, T.W. Platinum (IV) antitumour compounds: Their bioinorganic chemistry. Coord. Chem. Rev. 2002, 232, 49–67. [Google Scholar] [CrossRef]
  6. Bouché, M.; Bonnefont, A.; Achard, T.; Bellemin-Laponnaz, S. Exploring diversity in platinum (IV) N-heterocyclic carbene complexes: Synthesis, characterization, reactivity and biological evaluation. Dalton Trans. 2018, 33, 11491–11502. [Google Scholar] [CrossRef]
  7. Bouché, M.; Dahm, G.; Wantz, M.; Fournel, S.; Achard, T.; Bellemin-Laponnaz, S. Platinum (IV) N-heterocyclic carbene complexes: Their synthesis, characterisation and cytotoxic activity. Dalton Trans. 2016, 45, 11362–11368. [Google Scholar] [CrossRef]
  8. Chardon, E.; Dahm, G.; Guichard, G.; Bellemin-Laponnaz, S. Derivatization of preformed platinum N-heterocyclic carbene complexes with amino acid and peptide ligands and cytotoxic activities toward human cancer cells. Organometallics 2012, 31, 7618–7621. [Google Scholar] [CrossRef]
  9. Bellemin-Laponnaz, S. N-Heterocyclic Carbene Platinum Complexes: A Big Step Forward for Effective Antitumor Compounds. Eur. J. Inorg. Chem. 2020, 2020, 10–20. [Google Scholar] [CrossRef] [Green Version]
  10. Priqueler, J.R.L.; Butler, I.S.; Rochon, D.D. High selectivity of colorimetric detection of p-nitrophenol based on Ag nanoclusters. Appl. Spectrosc. Rev. 2006, 41, 185–226. [Google Scholar] [CrossRef]
  11. Höfer, D.; Varbaniv, H.P.; Hejl, M.; Jakupec, M.A.; Roller, A.; Galanski, M.; Keppler, B.K. Impact of the equatorial coordination sphere on the rate of reduction, lipophilicity and cytotoxic activity of platinum (IV) complexes. J. Inorg. Biochem. 2017, 174, 119–129. [Google Scholar] [CrossRef] [PubMed]
  12. Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. The next generation of platinum drugs: Targeted Pt (II) agents, nanoparticle delivery, and Pt (IV) prodrugs. Chem. Rev. 2016, 116, 3436–3486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Johnstone, T.C.; Alexander, S.M.; Wilson, J.J.; Lippard, S.J. Oxidative halogenation of cisplatin and carboplatin: Synthesis, spectroscopy, and crystal and molecular structures of Pt (IV) prodrugs. Dalton Trans. 2015, 44, 119–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bokach, N.A.; Kukushkin, V.Y.; Kuznetsov, M.L.; Garnovskii, D.A.; Natile, G.; Pombeiro, A.J.L. Direct addition of alcohols to organonitriles activated by ligation to a platinum (IV) center. Inorg. Chem. 2002, 41, 2041–2053. [Google Scholar] [CrossRef]
  15. Still, B.M.; Anil Kumar, P.G.; Aldrich-Wright, J.R.; Price, W.S. 195Pt NMR—Theory and application. Chem. Soc. Rev. 2007, 36, 665–686. [Google Scholar] [CrossRef]
  16. Hu, D.; Yang, C.; Lok, C.-N.; Xing, F.; Lee, P.-Y.; Fung, Y.M.E.; Jiang, H.; Che, C.-M. An Antitumor Bis (N-Heterocyclic Carbene) Platinum (II) Complex That Engages Asparagine Synthetase as an Anticancer Target. Angew. Chem. Int. Ed. 2019, 58, 10914–10918. [Google Scholar] [CrossRef]
  17. Matczuk, M.; Ruzik, L.; Alekssanko, S.S.; Keppler, B.K.; Jarosz, M.; Timerbaev, A.R. Analytical methodology for studying cellular uptake, processing and localization of gold nanoparticles. Anal. Chim. Acta 2019, 1052, 1–9. [Google Scholar] [CrossRef]
  18. Galvez, L.; Theiner, S.; Grabarics, M.; Kowol, C.R.; Keppler, B.K.; Hann, S.; Koellensperger, G. Critical assessment of different methods for quantitative measurement of metallodrug-protein associations. Anal. Bioanal. Chem. 2018, 410, 7211–7220. [Google Scholar] [CrossRef] [Green Version]
  19. Ahmad, S. Kinetic aspects of platinum anticancer agents. Polyhedron 2017, 138, 109–124. [Google Scholar] [CrossRef]
  20. Hall, M.D.; Daly, H.L.; Zhang, J.Z.; Zhang, M.; Alderden, R.A.; Pursche, D.; Foran, G.J.; Hambley, T.W. Quantitative measurement of the reduction of platinum (IV) complexes using X-ray absorption near-edge spectroscopy (XANES). Metallomics 2012, 4, 568–575. [Google Scholar] [CrossRef]
  21. Czapla-Masztafiak, J.; Kubas, A.; Kayser, Y.; Fernandes, D.L.A.; Kwiatek, W.M.; Lipiec, E.; Deacon, G.B.; Al-Jorani, K.; Wood, B.R.; Szlachetko, J.; et al. Mechanism of hydrolysis of a platinum (IV) complex discovered by atomic telemetry. J. Inorg. Biochem. 2018, 187, 56–61. [Google Scholar] [CrossRef] [PubMed]
  22. Huynh, H.V. The Organometallic Chemistry of N-heterocyclic Carbenes; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar]
  23. Teng, Q.Q.; Huynh, H.V. A Unified Ligand Electronic Parameter Based on 13 C NMR Spectroscopy of N-heterocyclic Carbene Complexes. Dalton Trans. 2017, 46, 614–627. [Google Scholar] [CrossRef]
  24. Teng, Q.Q.; Ng, P.S.; Leung, J.N.; Huynh, H.V. Donor strengths determination of pnictogen and chalcogen ligands by the Huynh electronic parameter and its correlation to sigma Hammett constants. Chem. Eur. J. 2019, 25, 13956–13963. [Google Scholar] [CrossRef] [Green Version]
  25. Tsipis, A.C.; Karapetsas, I.N. Prediction of 195Pt NMR of photoactivable diazido- and azine-Pt(IV) anticancer agents by DFT computational protocols. Magn. Reson. Chem. 2017, 55, 145–153. [Google Scholar] [CrossRef]
  26. Appleton, T.G.; Berry, R.D.; Davis, C.A.; Hall, J.R.; Kimlin, H.A. Reactions of platinum(II) aqua complexes. I: Multinuclear (195Pt, 15N, and 31P) NMR study of reactions between the cis-diamminediaquaplatinum(II) cation and the oxygen-donor ligands hydroxide, perchlorate, nitrate, sulfate, phosphate, and acetate. Inorg. Chem. 1984, 23, 3514–3531. [Google Scholar] [CrossRef]
  27. Appleton, T.G.; Hall, J.R.; Ralph, S.F.; Thompson, C.S.M. Reactions of platinum(II) aqua complexes. 2. Platinum-195 NMR study of reactions between the tetraaquaplatinum(II) cation and chloride, hydroxide, perchlorate, nitrate, sulfate, phosphate, and acetate. Inorg. Chem. 1984, 23, 3521–3525. [Google Scholar] [CrossRef]
  28. Benhamou, L.; Chardon, E.; Lavigne, G.; Bellemin-Laponnaz, S.; César, V. Synthetic routes to N-heterocyclic carbene precursors. Chem. Rev. 2009, 111, 2705–2733. [Google Scholar] [CrossRef] [PubMed]
  29. Sutter, K.; Autschbach, J. Computational study and molecular orbital analysis of NMR shielding, spin–spin coupling, and electric field gradients of azido platinum complexes. J. Am. Chem. Soc. 2012, 134, 13374–13385. [Google Scholar] [CrossRef]
  30. Dahm, D.; Bailly, C.; Karmazin, L.; Bellemin-Laponnaz, S. Synthesis, structural characterization and in vitro anti-cancer activity of functionalized N-heterocyclic carbene platinum and palladium complexes. J. Organomet. Chem. 2015, 794, 115–124. [Google Scholar] [CrossRef]
  31. Chardon, E.; Puleo, G.-L.; Dahm, G.; Guichard, G.; Bellemin-Laponnaz, S. Direct functionalisation of group 10 N-heterocyclic carbene complexes for diversity enhancement. Chem. Commun. 2011, 47, 5864–5866. [Google Scholar] [CrossRef]
  32. Chtchigrovsky, M.; Eloy, L.; Jullien, H.; Saker, L.; Ségal-Bendirdjian, E.; Poupon, J.; Bombard, S.; Cresteil, T.; Retailleau, P.; Marinetti, A. Antitumor trans-N-Heterocyclic Carbene–Amine–Pt(II) Complexes: Synthesis of Dinuclear Species and Exploratory Investigations of DNA Binding and Cytotoxicity Mechanisms. J. Med. Chem. 2013, 56, 2074–2086. [Google Scholar] [CrossRef] [PubMed]
  33. Skander, M.; Retailleau, P.; Bourri, B.; Schio, L.; Mailliet, P.; Marinetti, A. N-heterocyclic carbene-amine Pt (II) complexes, a new chemical space for the development of platinum-based anticancer drugs. J. Med. Chem. 2010, 53, 2146–2154. [Google Scholar] [CrossRef] [PubMed]
  34. Becker, M.; Port, R.E.; Zabel, H.-J.; Zeller, W.J.; Bachert, P. Monitoring Local Disposition Kinetics of Carboplatinin Vivoafter Subcutaneous Injection in Rats by Means of 195Pt NMR. J. Magn. Reson. 1998, 133, 115–122. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Not available.
Scheme 1. General synthesis of the platinum (II) and platinum (IV) complexes.
Scheme 1. General synthesis of the platinum (II) and platinum (IV) complexes.
Molecules 25 03148 sch001
Scheme 2. Molecular structure of the N-heterocyclic carbene (NHC)-Pt(II) references.
Scheme 2. Molecular structure of the N-heterocyclic carbene (NHC)-Pt(II) references.
Molecules 25 03148 sch002
Scheme 3. Molecular structure of the NHC-Pt(IV) complexes.
Scheme 3. Molecular structure of the NHC-Pt(IV) complexes.
Molecules 25 03148 sch003
Figure 1. Molecular structure of complex 15. Selected bond distances (Å) and angles (deg): C(1)-Pt(1), 2.057(8); Br(1)-Pt(1), 2.4882(8); Br(2)-Pt(1), 2.4657(8); Br(3)-Pt(1), 2.4615(8); Br(4)-Pt, 2.4839(8); N(3)-Pt(1), 2.128(6); C(1)-Pt(1)-N(3), 179.2(3); C(1)-Pt(1)-Br(3), 92.9(2); N(3)-Pt(1)-Br(3), 87.10(16); Br(2)-Pt(1)-Br(3), 86.10(3); Br(1)-Pt(1)-Br(4), 177.05(3).
Figure 1. Molecular structure of complex 15. Selected bond distances (Å) and angles (deg): C(1)-Pt(1), 2.057(8); Br(1)-Pt(1), 2.4882(8); Br(2)-Pt(1), 2.4657(8); Br(3)-Pt(1), 2.4615(8); Br(4)-Pt, 2.4839(8); N(3)-Pt(1), 2.128(6); C(1)-Pt(1)-N(3), 179.2(3); C(1)-Pt(1)-Br(3), 92.9(2); N(3)-Pt(1)-Br(3), 87.10(16); Br(2)-Pt(1)-Br(3), 86.10(3); Br(1)-Pt(1)-Br(4), 177.05(3).
Molecules 25 03148 g001
Table 1. Half-inhibitory concentrations IC50 (µM) of the selected complexes toward the HCT116, MCF7 and PC3 cancer cells.
Table 1. Half-inhibitory concentrations IC50 (µM) of the selected complexes toward the HCT116, MCF7 and PC3 cancer cells.
Complex NumberStructureIC50 (µM)
HCT116 1
IC50 (µM)
MCF7 1
IC50 (µM)
PC3 1
Cisplatin(NH3)2PtCl23.57 ± 0.14.15 ± 0.73.10 ± 0.2
2(NHC)PtBr2(pyr)5.44 ± 17.73 ± 15.35 ± 1.6
3(NHC)PtBr2(DMSO)>100>100>100
4(NHC)PtCl2(DMSO)63 ± 580 ± 1365 ± 6
5(NHC)PtCl2(pyr)3.78 ± 0.13.48 ± 14.40 ± 0.9
6 (NHC)PtBr4(amine)7.5 ± 0.323 ± 510 ± 1
814 ± 25 ± 15 ± 1
1211 ± 0.33 ± 0.72 ± 0.5
1681.09 ± 217.22 ± 1.85.42 ± 0.5
195 ± 14 ± 0.25 ± 1
22(NHC)PtCl4(amine)0.5 ± 0.030.5 ± 0.091 ± 0.1
251.48 ± 0.21.78 ± 0.61.31 ± 0.2
1 HCT116, colon cancer cells; MCF7, breast carcinoma; PC3, prostate adenocarcinoma. (After 72 h of incubation; stock solutions in DMSO for all complexes; stock solution in H2O for cisplatin).
Table 2. Chemical shift evolution of the Pt signal as a function of the metal oxidation state, the coordination sphere of the metal and the NHC substituents (external reference for 195Pt: H2PtCl6 in D2O: δPt = 0 ppm).
Table 2. Chemical shift evolution of the Pt signal as a function of the metal oxidation state, the coordination sphere of the metal and the NHC substituents (external reference for 195Pt: H2PtCl6 in D2O: δPt = 0 ppm).
Complex Ox. StateδPt (ppm)
195Pt NMR
δC (ppm)
13C NMR
1+II−4313125.1
2+II−3814138.2
3+II−3356154.7
4+II−3351n.o. 1
5+II−3304n.o.
6+IV−2196n.o.
7+IV−2168113.4
8+IV−2168115.2
9+IV−2167133.9
10+IV−2083124.6
11+IV−2081n.o.
12+IV−2080112.7
13+IV−2079115.4
14+IV−2070n.o.
15+IV−2067n.o.
16+IV−2063110.8
17+IV−2058110.7
18+IV−2048109.3
19+IV−2040109.2
20+IV−2032n.o.
21+IV−1901n.o.
22+IV−883n.o.
23+IV−853n.o.
24+IV−825112.9
25+IV−810111.5
26+IV−795n.o.
1 n.o.: not observed.

Share and Cite

MDPI and ACS Style

Bouché, M.; Vincent, B.; Achard, T.; Bellemin-Laponnaz, S. N-Heterocyclic Carbene Platinum(IV) as Metallodrug Candidates: Synthesis and 195Pt NMR Chemical Shift Trend. Molecules 2020, 25, 3148. https://doi.org/10.3390/molecules25143148

AMA Style

Bouché M, Vincent B, Achard T, Bellemin-Laponnaz S. N-Heterocyclic Carbene Platinum(IV) as Metallodrug Candidates: Synthesis and 195Pt NMR Chemical Shift Trend. Molecules. 2020; 25(14):3148. https://doi.org/10.3390/molecules25143148

Chicago/Turabian Style

Bouché, Mathilde, Bruno Vincent, Thierry Achard, and Stéphane Bellemin-Laponnaz. 2020. "N-Heterocyclic Carbene Platinum(IV) as Metallodrug Candidates: Synthesis and 195Pt NMR Chemical Shift Trend" Molecules 25, no. 14: 3148. https://doi.org/10.3390/molecules25143148

Article Metrics

Back to TopTop