research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206

Synthesis and structural characterization of a new dinuclear platinum(III) complex, [Pt2Cl4(NH3)2{μ-HN=C(O)But}2]

crossmark logo

aEuropean XFEL, Holzkoppel 4, Schenefeld, 22869, Germany, and bCRISMAT, CNRS, ENSICAEN, Université de Caen Normandie, Caen, France
*Correspondence e-mail: doriana.vinci@xfel.eu

Edited by K. E. Knope, Georgetown University, USA (Received 1 July 2022; accepted 30 September 2022; online 4 November 2022)

Cisplatin plays an important role in treating many tumors. However, its pharmacological potential is limited by several side effects (nephrotoxicity, neurotoxicity, hair loss) and the resistance of many cancer types. In the effort of overcome resistance to cisplatin, more than 3000 platinum complexes have been synthesized in the past 45 years, but only about 30 have demonstrated adequate pharmacological platinum(III) complexes have been synthesized and their structure determined, because some of these exhibit antitumor and catalytic activity for the oxidation of olefins. Herein the synthesis of a new dinuclear platinum(III) complex with two butyl­(amidato) bridging ligands, namely [Pt2Cl4(NH3)2{μ-HN=C(O)But}2], is reported. The complex was characterized by means of nuclear magnetic resonance (1H and 195Pt NMR) and infrared spectroscopy, and the crystal structure determined using X-ray single-crystal diffraction. This complex, to the best of our knowledge, is the first dinuclear complex of PtIII with two butyl­(amidato) bridging ligands with a head-to-tail configuration.

1. Introduction

The interest in platinum complexes arises from their anti­cancer and catalytic activities for the oxidation of olefins. The discovery of the anticancer properties of cisplatin and its clinical introduction in the 1970s represent an important milestone in the history of effective anticancer drugs. Cisplatin, cis-diamminedi­chloro­platinum(II), was discovered by Rosenberg et al. (1965[Rosenberg, B., Van Camp, L. & Krigas, T. (1965). Nature, 205, 698-699.]) during their experiments on the effect of electric fields on the cell division of bacteria. Cisplatin plays a significant role in the treatment of epithelial malignancies and in the cure of testicular cancer. It is used also as a principal component for the treatment of ovarian, head-and-neck, esophagus, stomach, colon, bladder, cervix and uterus cancers and as second-line treatment against most other advanced cancers including breast, pancreas, liver, kidney and prostate cancers. The primary cellular target of cisplatin is DNA, and the antitumor effects of platinum complexes are expected to arise from their ability to form several adducts with DNA, blocking its replication and transcription, and inducing cell death. The antitumor properties of cisplatin are attributed to the kinetics of the aqua­tion reactions of its chloride ligands leading to DNA crosslinking activities. This process is responsible of DNA bending, interferes with DNA replication, transcription and other nuclear functions, and blocks cancer cell proliferation and tumor growth (Boulikas et al., 2007[Boulikas, T., Pantos, A., Bellis, E. & Christofis, P. (2007). Cancer Ther. 5, 537-583.]). However, its pharmacological potential is limited by its side effects (nephrotoxicity, neurotoxicity, hair loss), and the resistance of many cancer types (O'Dwyer et al., 1999[O'Dwyer, P., Stevenson, J. & Johnson, S. (1999). In Cisplatin: Chemistry and Biochemistry of a Leading Anticancer Drug, edited by B. Lippert. Zurich: Verlag Helvetica Chimica Acta.]; Von Hoff et al., 1979[Von Hoff, D. D., Schilsky, R., Reichert, C. M., Reddick, R. L., Rozencweig, M., Young, R. C. & Muggia, F. M. (1979). Cancer Treat. Rep. 63, 1527-1531.]). In the past 45 years over 3000 platinum complexes have been synthesized in the effort of overcoming cisplatin limitations, but not more than 30 of them exhibited adequate pharmacological advantages over cisplatin. Only five of these latter complexes have been used in clinical applications, three FDA-approved (cisplatin, carboplatin and oxaliplatin), one used in Japan (nedaplatin) and one in China (lobaplatin). For a long time, the cisplatin geometry has been retained as a necessary requisite for a platinum complex to exhibit antitumoral activity (Kelland, 2007[Kelland, L. R. (2007). Nat. Rev. Cancer, 7, 573-584.]). Indeed, the trans-geometry of the platin complex and other transplatin analogs do not exhibit antitumor activity because of their kinetic instability which renders the complex unable to bind the target in the active form (Cornacchia et al., 2009[Cornacchia, D., Pellicani, R. Z., Intini, F. P., Pacifico, C. & Natile, G. (2009). Inorg. Chem. 48, 10800-10810.]). More recently, however, several researchers have reported that the trans-isomer can gain cytotoxicity comparable to that of the cis-isomer and even of cisplatin (Farrell et al., 1992[Farrell, N., Kelland, L. R., Roberts, J. D. & Van Beusichem, M. (1992). Cancer Res. 52, 5065-5072.]; Natile & Coluccia, 2001[Natile, G. & Coluccia, M. (2001). Coord. Chem. Rev. 216-217, 383-410.]; Montero et al., 1999[Montero, E. I., Díaz, S., González-Vadillo, A. M., Pérez, J. M., Alonso, C. & Navarro-Ranninger, C. (1999). J. Med. Chem. 42, 4264-4268.]; Kasparkova et al., 2003a[Kasparkova, J., Marini, V., Najajreh, Y., Gibson, D. & Brabec, V. (2003a). Biochemistry, 42, 6321-6332.],b[Kasparkova, J., Novakova, O., Farrell, N. & Brabec, V. (2003b). Biochemistry, 42, 792-800.]).

On the other hand, the number of dinuclear PtIII complexes containing a metal–metal single bond is steadily increasing as well (O'Halloran & Lippard, 1985[O'Halloran, T. V. & Lippard, S. J. (1985). Isr. J. Chem. 25, 130-137.]; Matsumoto & Sakai, 1999[Matsumoto, K. & Sakai, K. (1999). Adv. Inorg. Chem. 49, 375-427.]; González et al., 2000[González, V. M., Fuertes, M. A., Pérez-Alvarez, M. J., Cervantes, G., Moreno, V., Alonso, C. & Pérez, J. M. (2000). Biochem. Pharmacol. 60, 371-379.]; Saeki et al., 2003[Saeki, N., Nakamura, N., Ishibashi, T., Arime, M., Sekiya, H., Ishihara, K. & Matsumoto, K. (2003). J. Am. Chem. Soc. 125, 3605-3616.]). Dinuclear PtIII complexes are interesting for their potential use in catalysis and biomedicine, and their chemistry and reactivity has not yet been explored to the same extent as for PtII and PtIV species. The majority of PtIII complexes include a metal–metal single bond supported by two (Matsumoto & Sakai, 1999[Matsumoto, K. & Sakai, K. (1999). Adv. Inorg. Chem. 49, 375-427.]; Chen & Matsumoto, 2003[Chen, W. & Matsumoto, K. (2003). Inorg. Chim. Acta, 342, 88-96.]; Lippert, 1999[Lippert, B. (1999). Coord. Chem. Rev. 182, 263-295.]) or four (Fedotova et al., 1997[Fedotova, T. N., Minacheva, L. K., Kuznetsova, G. N., Sakharova, V. G., Gelfman, M. I. & Baranovskii, I. B. (1997). Russ. J. Inorg. Chem. 42, 1838-1846.]; Dolmella et al., 2002[Dolmella, A., Intini, F. P., Pacifico, C., Padovano, G. & Natile, G. (2002). Polyhedron, 21, 275-280.]; Bandoli et al., 2003[Bandoli, G., Dolmella, A., Intini, F. P., Pacifico, C. & Intini, G. (2003). Inorg. Chim. Acta, 346, 143-150.]) bridging ligands (Fig. 1[link]), namely `platinum blues'-derived and `lantern-type', respectively. Dinuclear platinum complexes with three bridging ligands (Fig. 1[link]) (Abe et al., 1991[Abe, T., Moriyama, H. & Matsumoto, K. (1991). Inorg. Chem. 30, 4198-4204.]) or unsupported by covalent bridges (Matsumoto et al., 1996[Matsumoto, K., Matsunami, J., Mizuno, K. & Uemura, H. (1996). J. Am. Chem. Soc. 118, 8959-8960.]; Lippert et al., 1983[Lippert, B., Schollhorn, H. & Thewalt, U. (1983). Z. Naturforsch. 38, 1441-1445.]) are rare. The chelating chain usually consists of three atoms of the type OXO (X = C,S,P), NCO, NCS, SCS or PXP (X = C,O) resulting in an overall five-membered ring including the platinum–platinum interaction (Cornacchia et al., 2009[Cornacchia, D., Pellicani, R. Z., Intini, F. P., Pacifico, C. & Natile, G. (2009). Inorg. Chem. 48, 10800-10810.]; Goodgame et al., 1986[Goodgame, D. M. L., Rollins, R. W., Slawin, A. M. Z., Williams, D. J. & Zard, P. W. (1986). Inorg. Chim. Acta, 120, 91-101.]; Umakoshi et al., 1987[Umakoshi, K., Kinoshita, I., Ichimura, A. & Ooi, S. (1987). Inorg. Chem. 26, 3551-3556.]; Cotton et al., 1997[Cotton, F. A., Matonic, J. H. & Murillo, C. A. (1997). Inorg. Chim. Acta, 264, 61-65.]; Bellitto et al., 1983[Bellitto, C., Flamini, A., Gastaldi, L. & Scaramuzza, L. (1983). Inorg. Chem. 22, 444-449.]; Usón et al., 1994[Usón, R., Forniés, J., Falvello, L. R., Tomas, M., Casas, J. M., Martin, A. & Cotton, F. A. (1994). J. Am. Chem. Soc. 116, 7160-7165.]). In addition to the bridging ligands, dinuclear PtIII complexes also have equatorial and axial ligands (Matsumoto, 2003[Matsumoto, K. (2003). Russ. Chem. Bull. 52, 2577-2587.]). A characteristic feature of dinuclear PtIII complexes is an unusually long bond distance between the platinum and the axial ligands, which is 10% longer than the corresponding distances in square-planar PtII and octahedral PtIV coordinations (O'Dwyer et al., 1999[O'Dwyer, P., Stevenson, J. & Johnson, S. (1999). In Cisplatin: Chemistry and Biochemistry of a Leading Anticancer Drug, edited by B. Lippert. Zurich: Verlag Helvetica Chimica Acta.]). Therefore, axial ligands are weaker bound, due to the strong trans labialization influence exerted by the intermetallic bond (Kuo et al., 2007[Kuo, M. T., Chen, H. H., Song, I. S., Savaraj, N. & Ishikawa, T. (2007). Cancer Metastasis Rev. 26, 71-83.]; Hartmann & Lipp, 2003[Hartmann, J. T. & Lipp, H. P. (2003). Expert Opin. Pharmacother. 4, 889-901.]; Dolmella et al., 2002[Dolmella, A., Intini, F. P., Pacifico, C., Padovano, G. & Natile, G. (2002). Polyhedron, 21, 275-280.]), and this characteristic feature may increase the reactivity toward the guanine bases of the DNA. The two-bridge PtIII complexes are characterized by a tilting of the two platinum coordination planes of 25° and a twist about the platinum–platinum vector averaging a torsion angle of 25° (O'Halloran & Lippard, 1985[O'Halloran, T. V. & Lippard, S. J. (1985). Isr. J. Chem. 25, 130-137.]). These distortions are due to the steric interactions between the non-bridging equatorial ligands.

[Figure 1]
Figure 1
(a) `Lantern-type' diplatinum(III) complex, (b) dinuclear PtIII with three bridging ligands and (c) a `platinum blues' derivate (this work). In this study, YZX = OCN, axial L = Cl, HHT is head-head-to-tail and HT is head-to-tail.

In recent decades, the chemistry of dinuclear platinum complexes has attracted an increasing interest (O'Halloran & Lippard, 1985[O'Halloran, T. V. & Lippard, S. J. (1985). Isr. J. Chem. 25, 130-137.]; Matsumoto & Sakai, 1999[Matsumoto, K. & Sakai, K. (1999). Adv. Inorg. Chem. 49, 375-427.]), because it has been demonstrated that some of these dinuclear PtIII complexes exhibit antitumor (Cervantes et al., 1997[Cervantes, G., Prieto, M. J. & Moreno, V. (1997). Met.-Based Drugs, 4, 9-18.]) and catalytic activities for the oxidation of olefins (Matsumoto & Sakai, 1999[Matsumoto, K. & Sakai, K. (1999). Adv. Inorg. Chem. 49, 375-427.]; Saeki et al., 2003[Saeki, N., Nakamura, N., Ishibashi, T., Arime, M., Sekiya, H., Ishihara, K. & Matsumoto, K. (2003). J. Am. Chem. Soc. 125, 3605-3616.]; Ochiai et al., 2004[Ochiai, M., Lin, Y.-S., Yamada, J., Misawa, H., Arai, S. & Matsumoto, K. (2004). J. Am. Chem. Soc. 126, 2536-2545.]). From this perspective, our research efforts have focused on the synthesis and the structural determination of a new head-to-tail (HT) dinuclear PtIII complex (Fig. 1[link]), with chemical formula [Pt2Cl4(NH3)2{μ-HN=C(O)But}2]. The complex has been characterized by means of NMR and IR spectroscopies, while the structure was determined using X-ray single-crystal diffraction.

2. Experimental

2.1. Synthesis

(i) K2PtCl4 (2.1641 g, 5.21 mmol, Mr = 415.06 g mol−1) and KI (4.3874 g, 26.4 mmol, Mr = 165.99 g mol−1) were dissolved in water (15 ml). The solution was stirred at 55°C until a brown suspension consistent with the formation of the tetra­iodo salt, having formula K2PtI4, was observed. To this aqueous dispersion of potassium salt, an aqueous solution of NH4OH was added dropwise and the pH was controlled to not exceed 7.5 during the addition. The obtained solution was warmed at 55°C under stirring for 20 min to allow the formation of a yellow precipitate of cis-[PtI2(NH3)2].

[Scheme 1]

(ii) cis-[PtI2(NH3)2] (1.5229 g, 3.15 mmol, Mr = 482.92 g mol−1) was dissolved in water (20 ml), stirred with AgNO3 (1.0713 g, 6.31 mmol, Mr = 169.88 g mol−1) at 70°C for 20 min in the dark, and AgI was then removed by filtration from the yellowish suspension. The mother liquor was treated with NCBut (2 ml) and stirred at 70°C for 1 h, affording a blue solution of cis-[Pt(NH3)2(NCBut)2]. The resulting nitrile complex was stirred at room temperature for 1 h with KI (2.6192 g, 17.78 mmol) to give a green precipitate of trans-[PtI2(NH3)(NCBut)].

[Scheme 2]

(iii) trans-[PtI2(NH3)(NCBut)] (0.9457 g, 1.72 mmol, Mr = 549 g mol−1), dissolved in acetone (20 ml), was first stirred at 60°C to give a yellow solution, and then stirred with AgNO3 (0.5859 g, 3.45 mmol) at 70°C for 20 min in the dark, affording a green solution that was filtered. The mother solution was treated with KCl (1.2892 g, 17.3 mmol, Mr = 74.56 g mol−1) and then taken to dryness by evaporation of the solvent under reduced pressure. The solid residue was dissolved in water (20 ml), stirred at 50°C for 1 h, and dried in vacuum to obtain the formation of a green precipitate of trans-[PtCl2(NH3)(NCBut)] .

[Scheme 3]

(iv) trans-[PtCl2(NH3)(NCBut)] (0.1780 g, 0.49 mmol, Mr = 366 g mol−1) dissolved in chloro­form (30 ml) and Cl2 (2 ml) was stirred at room temperature for 30 min. The solvent was then evaporated under vacuum, forming a yellow precipitate of trans-[PtCl4(NH3)(NCBut)].

[Scheme 4]

(v) The novel binuclear PtIII complex can be prepared by the reaction of trans-[PtCl4(NH3)(NCBut)] and trans-[PtI2(NH3)(NCBut)] in water solution: trans-[PtCl4(NH3)(NCBut)] (0.0335 g, 76.8 µmol, Mr = 436 g mol−1) and trans-[PtI2(NH3)(NCBut)] (0.0415 g, 75.6 µmol) were first stirred at 60°C for 6 h in a water solution (10 ml), giving a brown suspension, and then heated for 20 h at 60°C. The solid residue was separated from the solution and crystallized in water by slow evaporation at room temperature. After three days, red crystals were formed (Fig. 2[link]). It is worth noting that in the final complex only chlorine ligands are in the equatorial and axial positions. This can be explained by assuming that the iodide anion (from the PtII complex) is a stronger base with respect to chlorine, and as a consequence exhibits a larger reactivity with H3O+ cations (water after reacting with the platinum precursors releases hydrogen ions).

[Scheme 5]
[Figure 2]
Figure 2
Single crystal selected for the X-ray diffraction characterization.

2.2. X-ray structure determination

The selected crystal (Fig. 1[link]) was mounted on a Bruker AXS X8 APEX CCD diffractometer equipped with a four-circle Kappa goniometer and a 4 K CCD detector (radiation Mo Kα). Data reduction and unit-cell refinement were carried out with the SAINT package (Bruker, 2003[Bruker (2003). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]). A total of 29306 reflections (θmax = 20.28°) was collected. The reflections were indexed, integrated and corrected for Lorentz, polarization and absorption effects with the program SADABS (Sheldrick, 2010[Sheldrick, G. M. (2010). SADABS. University of Gottingen, Gottingen, Germany.]). The unit-cell parameters were calculated from all reflections. The structure was solved using direct methods in space group C2/c and the model refined using full-matrix least-squares. The ADPs of non-hydrogen atoms were refined anisotropically, while hydrogen atoms were located by Fourier difference, except for those of the tert-butyl group which have been placed at calculated positions, and ADPs refined isotropically. All calculations and molecular graphics were carried out with SIR92 (Altomare et al., 1993[Altomare, A., Cascarano, G., Giacovazzo, C., Guagliardi, A., Camalli, M., Burla, M. C. & Polidori, G. (1993). Acta Cryst. A49, c55.]), PARST97 (Nardelli, 1995[Nardelli, M. (1995). J. Appl. Cryst. 28, 659-659.]), WinGX (Ferrugia, 1999[Ferrugia, L. J. (1999). J. Appl. Cryst. 32, 837-838.]), CRYSTALS (Betteridge et al., 2003[Betteridge, P. W., Carruthers, J. R., Cooper, R. I., Prout, K. & Watkin, D. J. (2003). J. Appl. Cryst. 36, 1487-1487.]), MERCURY (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]) and ORTEP-3 for Windows packages (Ferrugia, 2012[Ferrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]). Details of the experiment and crystal data are listed in Table 1[link]. Atomic positions are listed in Table 2[link]. Selected bond lengths, bond angles and atomic displacement parameters are given in the CIF (see supporting information). The CIF file has been deposited in the Crystallography Open Database (Gražulis et al., 2009[Gražulis, S., Chateigner, D., Downs, R. T., Yokochi, A. F. T., Quirós, M., Lutterotti, L., Manakova, E., Butkus, J., Moeck, P. & Le Bail, A. (2009). J. Appl. Cryst. 42, 726-729.]; ref. 3000395).

Table 1
Experimental details

Crystal data
Chemical formula [Pt2Cl4(NH3)2{μ-HN=C(O)But}2]
Mr 766.33
Crystal system, space group Monoclinic, C2/c
Temperature (K) 293
a, b, c (Å) 19.3247 (3), 8.8795 (1), 12.7062 (2)
β (°) 108.332 (1)
V3) 2069.65 (5)
Z 4
Radiation type Mo Kα
μ (mm−1) 14.03
Crystal size (mm) 0.77 × 0.24 × 0.21
 
Data collection
Diffractometer Nonius KappaCCD
Absorption correction Multi-scan (SADABS)
Tmin, Tmax 0.03, 0.05
No. of measured, independent and observed [I > 2.0σ(I)] reflections 29306, 6496, 4944
Rint 0.027
(sin θ/λ)max−1) 0.910
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.022, 0.017, 1.12
No. of reflections 4707
No. of parameters 100
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 1.30, −1.30
Computer programs: COLLECT (Nonius, 2001[Nonius (2001). COLLECT. Nonius BV, Delft, The Netherlands.]), CrysAlis (Oxford Diffraction, 2002[Oxford Diffraction (2002). CrysAlis. Oxford Diffraction Ltd, Abingdon, Oxfordshire, England.]), SIR92 (Altomare et al., 1993[Altomare, A., Cascarano, G., Giacovazzo, C., Guagliardi, A., Camalli, M., Burla, M. C. & Polidori, G. (1993). Acta Cryst. A49, c55.]), CRYSTALS (Betteridge et al., 2003[Betteridge, P. W., Carruthers, J. R., Cooper, R. I., Prout, K. & Watkin, D. J. (2003). J. Appl. Cryst. 36, 1487-1487.]), CAMERON (Watkin et al., 1996[Watkin, D. J., Prout, C. K. & Pearce, L. J. (1996). CAMERON. Chemical Crystallography Laboratory, Oxford, England.]).

Table 2
Selected geometric parameters (Å, °)

Pt1—Pt1i 2.5661 (2) Pt1—N2i 1.9989 (18)
Pt1—Cl1 2.3214 (6) Pt1—O1 2.0243 (16)
Pt1—Cl2 2.4282 (6) C1—O1 1.293 (3)
Pt1—N1 2.0576 (19) C1—N2 1.294 (3)
Pt1i—Pt1—Cl2 170.852 (16) Cl1—Pt1—N1 90.92 (6)
Pt1i—Pt1—O1 85.94 (5) N2i—Pt1—O1 93.18 (8)
Pt1i—Pt1—N1 99.37 (6) N2i—Pt1—Cl2 90.82 (6)
O1—C1—N2 121.65 (19) N2i—Pt1—Cl1 87.80 (6)
Cl2—Pt1—O1 88.46 (5) N2i—Pt1—N1 177.99 (9)
Cl2—Pt1—N1 87.66 (7) N2i—Pt1—Pt1i 82.28 (6)
Cl1—Pt1—O1 179.02 (5)    
Symmetry code: (i) 1 − x, y, −z + ½.

3. Results and discussion

NMR and IR characterization. The 1H-NMR spectrum at 295 K in DMSO-d6 (Fig. S1) exhibits signals at frequencies ∼1.20, ∼5.00 and ∼6.70 ppm assigned to the tert-butyl [—C(CH3)3], ammine (—NH3) and amidate [—N(H)CO] protons, respectively. The [1H-195Pt] HSQC–NMR heterocorrelate spectrum (Fig. S2) recorded in DMSO-d6, exhibits two NH signals at 5.01 and 6.74 ppm correlated with the platinum signal at −347 ppm, indicative of a PtIII cation in a N2Cl2OPt coordination environment. The occurrence of the tert-butyl and ammine ligands in the complex is coherent with the IR spectrum with bands at 2964–2918 and 3283–3405 cm−1, respectively (Fig. S2).

X-ray diffraction analysis. The asymmetric unit comprises half a molecule of the [Pt2Cl4(NH3)2{μ-HN=C(O)But}2] complex and the structure is generated by the twofold axis at the midpoint of the Pt–Pt bond (Fig. 3[link]). The coordination geometry of each PtIII atom (Fig. 3[link]) can be considered a distorted octahedron, with one chlorine, one oxygen and two nitro­gen atoms in equatorial positions, and one chlorine and one platinum of the second subunit in axial positions. The Pt–Pt bond distance [2.5661 (2) Å] is larger than that observed for four-bridge complexes with the same amidate bridge {e.g. [Pt2{HN=C(But)O}4(9-EtG)2](NO3)2, 2.4512 (5) Å (Pacifico et al., 2010[Pacifico, C., Intini, F. P., Nushi, F. & Natile, G. (2010). Bioinorg. Chem. Appl. 2010, 102863.])}, and in the range of `platinum blue'-derivates with the head-to-tail configuration [2.582 (1)–2.547 (1) Å range given by O'Halloran & Lippard (1985[O'Halloran, T. V. & Lippard, S. J. (1985). Isr. J. Chem. 25, 130-137.])]. This behavior indicates that the Pt–Pt distance is influenced by the number of bridging ligands. The equatorial Pt1—Cl1 [2.3214 (6) Å], Pt1—N1 [2.0576 (19) Å] and Pt—O1 [2.0243 (16) Å] bond distances are within the range of those reported for doubly and quadruply bridged dinuclear PtIII (Fedotova et al., 1997[Fedotova, T. N., Minacheva, L. K., Kuznetsova, G. N., Sakharova, V. G., Gelfman, M. I. & Baranovskii, I. B. (1997). Russ. J. Inorg. Chem. 42, 1838-1846.]; Dolmella et al., 2002[Dolmella, A., Intini, F. P., Pacifico, C., Padovano, G. & Natile, G. (2002). Polyhedron, 21, 275-280.]; Hollis et al., 1983[Hollis, L. S., Roberts, M. M. & Lippard, S. J. (1983). Inorg. Chem. 22, 3637-3644.]), as well as for PtII and PtIV complexes (Erxleben et al., 2002[Erxleben, A., Metzger, S., Britten, J. F., Lock, C. J. L., Albinati, A. & Lippert, B. (2002). Inorg. Chim. Acta, 339, 461-469.]; Lippert et al., 1984[Lippert, B., Raudaschl, G., Lock, C. J. L. & Pilon, P. (1984). Inorg. Chim. Acta, 93, 43-50.]; Ali et al., 2005[Ali, M. S., Ali Khan, S. R., Ojima, H., Guzman, I. Y., Whitmire, K. H., Siddik, Z. H. & Khokhar, A. R. (2005). J. Inorg. Biochem. 99, 795-804.]; Sigel et al., 1999[Sigel, R. K. O., Sabat, M., Freisinger, E., Mower, A. & Lippert, B. (1999). Inorg. Chem. 38, 1481-1490.]; Shamsuddin et al., 2007[Shamsuddin, S., Ali, M. S., Whitmire, K. H. & Khokhar, A. R. (2007). Polyhedron, 26, 637-644.]). Instead, the axial Pt1−Cl2 bond distance [2.4282 (6) Å] is longer than those reported in the literature (Fedotova et al., 1997[Fedotova, T. N., Minacheva, L. K., Kuznetsova, G. N., Sakharova, V. G., Gelfman, M. I. & Baranovskii, I. B. (1997). Russ. J. Inorg. Chem. 42, 1838-1846.]; Dolmella et al., 2002[Dolmella, A., Intini, F. P., Pacifico, C., Padovano, G. & Natile, G. (2002). Polyhedron, 21, 275-280.]; Hollis et al., 1983[Hollis, L. S., Roberts, M. M. & Lippard, S. J. (1983). Inorg. Chem. 22, 3637-3644.]; Erxleben et al., 2002[Erxleben, A., Metzger, S., Britten, J. F., Lock, C. J. L., Albinati, A. & Lippert, B. (2002). Inorg. Chim. Acta, 339, 461-469.]; Lippert et al., 1984[Lippert, B., Raudaschl, G., Lock, C. J. L. & Pilon, P. (1984). Inorg. Chim. Acta, 93, 43-50.], 1986[Lippert, B., Schoellhorn, H. & Thewalt, U. (1986). Inorg. Chem. 25, 407-408.]; Ali et al., 2005[Ali, M. S., Ali Khan, S. R., Ojima, H., Guzman, I. Y., Whitmire, K. H., Siddik, Z. H. & Khokhar, A. R. (2005). J. Inorg. Biochem. 99, 795-804.]; Sigel et al., 1999[Sigel, R. K. O., Sabat, M., Freisinger, E., Mower, A. & Lippert, B. (1999). Inorg. Chem. 38, 1481-1490.]; Shamsuddin et al., 2007[Shamsuddin, S., Ali, M. S., Whitmire, K. H. & Khokhar, A. R. (2007). Polyhedron, 26, 637-644.]), possibly because of the strong trans influence exerted by the Pt—Pt bond. The doubly-bridged PtIII molecule is characterized by a twist around the platinum–platinum vector [N2—Pt1—Pt1—O2] with an average torsion angle of 19.58°. The steric interactions between the ammine (bound to the first platinum) and the chloride (bound to the platinum in the second subunit) ligands are responsible for these distortions. In the amidate moiety, the C—X distances are 1.294 (3) Å and 1.293 (3) Å for the C1—N2 and C1—O1 bonds, respectively, in accordance with an extensive π-bond delocalization over the O—C—N moiety. The complex exhibits an intramolecular hydrogen bond involving the ammine hydrogen and the chlorine ligand (Fig. 4[link])

[Figure 3]
Figure 3
ORTEP drawing of the final structure model of the [Pt2Cl4(NH3)2{μ-HN=C(O)But}] complex. Symmetry code: (i) 1 − x, y, ½ − z. Ellipsoids drawn at the 30% probability level. Atom color coding: white for hydrogen, green for chlorine, blue for nitrogen, gray for carbon and red for oxygen.
[Figure 4]
Figure 4
MERCURY drawing of [Pt2Cl4(NH3)2{μ-HN=C(O)But}2] showing the intra­molecular hydrogen bond interaction [N1⋯Cl1i 3.176 (3) Å, (N1)H12⋯Cl1i 2.46 Å, N1—H12⋯Cl1i 140°; symmetry code: (i) 1 − x, y, −z + ½]. Ellipsoids drawn at the 30% probability level. Atom color coding: white for hydrogen, green for chlorine, blue for nitrogen, gray for carbon and red for oxygen.

The molecular packing (Fig. 5[link]) is governed by intermolecular hydrogen bonds (Table 3[link]) and van der Waals interactions. (i) Intermolecular hydrogen bonds occur between the N(H) hydrogen of the ammine group and the chlorine ligand (see Fig. 6[link]); and between the axial chlorine and the N(H) hydrogen of the amidate chelating group (see Fig. 7[link]). (ii) van der Waals intermolecular forces involve chlorine and ammine ligands in equatorial position (see Fig. 8[link]).

Table 3
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
Intramolecular
N1—H11⋯Cl2i 0.89 2.464 3.176 (3) 140
Intermolecular
N1—H11⋯Cl2ii 0.89 2.541 3.406 (2) 166
N2—H21⋯Cl2iii 0.87 2.669 3.361 (2) 138
Symmetry codes: (i) 1 + x, y, −z + ½; (ii) 1 − x, −y, 1 − z; (iii) x, 1 − y, z + ½.
[Figure 5]
Figure 5
Crystal packing of [Pt2Cl4(NH3)2{μ-HN=C(O)But}2], viewed along [010]. Ellipsoids drawn at the 30% probability level. Atom color coding: white for hydrogen, green for chlorine, blue for nitrogen, gray for carbon and red for oxygen.
[Figure 6]
Figure 6
MERCURY drawing of two molecules of [Pt2Cl4(NH3)2{μ-HN=C(O)But}2] linked by an intermolecular hydrogen bond [N1⋯Cl2i 3.406 (2) Å, (N1)H11⋯Cl2i 2.54 Å, N1—H1⋯Cl2i 166°; symmetry code: (i) 1 − x, −y, 1 − z]. Ellipsoids drawn at the 30% probability level. Atom color coding: white for hydrogen, green for chlorine, blue for nitrogen, gray for carbon and red for oxygen.
[Figure 7]
Figure 7
MERCURY drawing of two molecules of [Pt2Cl4(NH3)2{μ-HN=C(O)But}2] linked by an intermolecular hydrogen bond [N2⋯Cl2i 3.361 (2) Å, (N2)H21⋯Cl2i 2.67 Å, N2—H21⋯Cl2i 138°; symmetry code: (i) x, 1 − y, z + ½]. Ellipsoids drawn at the 30% probability level. Atom color coding: white for hydrogen, green for chlorine, blue for nitrogen, gray for carbon and red for oxygen.
[Figure 8]
Figure 8
MERCURY drawing of two molecules of [Pt2Cl4(NH3)2{μ-HN=C(O)But}2] linked by van der Waals intermolecular interactions [N1⋯Cl1i 3.496 (2) Å, (N1)H13⋯Cl1i 2.86 Å, N1—H13⋯Cl1i 130°; symmetry code: (i) 1 − x, −y, 1 − z]. Ellipsoids drawn at the 30% probability level. Atom color coding: white for hydrogen, green for chlorine, blue for nitrogen, gray for carbon and red for oxygen.

The structural results from the spectroscopic and diffraction analyses confirm that we have successfully designed the synthesis of the expected complex, and the absence of iodine ligands as hypothesized.

4. Conclusion

The necessity to overcome the limitations associated with the use of cisplatin, such as the occurrence of toxic side effects and drug-resistance behaviors, has led to the synthesis of new complexes with higher pharmacological properties. In this study, we have identified a strategy for synthesizing a novel dinuclear PtIII complex from the pivalo­nitrile derivatives of PtII and PtIV as precursors. The complex has been characterized by NMR (1H and 195Pt) and IR spectroscopies. The crystal structure was determined by X-ray crystallography. To our knowledge this complex is the first example of dinuclear PtIII species with two bridging ligands in HT configuration. Since it has been demonstrated that `lantern-shaped' PtIII complexes exhibit antitumor activity, it will be interesting to investigate the activity of our complex with biological assays.

Supporting information


Computing details top

Data collection: COLLECT (Nonius, 2001).; cell refinement: DENZO/SCALEPACK (Otwinowski & Minor, 1997); data reduction: CrysAlis, (Oxford Diffraction, 2002); program(s) used to solve structure: SIR92 (Altomare et al., 1994); program(s) used to refine structure: CRYSTALS (Betteridge et al., 2003); molecular graphics: CAMERON (Watkin et al., 1996); software used to prepare material for publication: CRYSTALS (Betteridge et al., 2003).

{2-amino-2-[mu-butyl(amidate)]-4-clhoride}diplatinum(III) top
Crystal data top
C10H26Cl4N4O2Pt2F(000) = 1416
Mr = 766.33Dx = 2.459 Mg m3
Monoclinic, C2/cMo Kα radiation, λ = 0.71073 Å
a = 19.3247 (3) ÅCell parameters from 9905 reflections
b = 8.8795 (1) Åθ = 3–40°
c = 12.7062 (2) ŵ = 14.03 mm1
β = 108.332 (1)°T = 293 K
V = 2069.65 (5) Å3Tabular, orange
Z = 40.77 × 0.24 × 0.21 mm
Data collection top
Nonius KappaCCD
diffractometer
4944 reflections with I > 2.0σ(I)
Graphite monochromatorRint = 0.027
φ & ω scansθmax = 40.3°, θmin = 3.3°
Absorption correction: multi-scan
SADABS
h = 3434
Tmin = 0.03, Tmax = 0.05k = 1616
29306 measured reflectionsl = 2223
6496 independent reflections
Refinement top
Refinement on FPrimary atom site location: structure-invariant direct methods
Least-squares matrix: fullHydrogen site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.022H-atom parameters constrained
wR(F2) = 0.017
S = 1.12(Δ/σ)max = 0.003
4707 reflectionsΔρmax = 1.30 e Å3
100 parametersΔρmin = 1.30 e Å3
0 restraints
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
C10.39504 (12)0.4284 (2)0.70417 (18)0.0252
C20.33011 (13)0.5356 (3)0.6762 (2)0.0328
C30.3542 (2)0.6860 (4)0.6415 (5)0.0736
C40.26775 (19)0.4685 (5)0.5828 (3)0.0604
C50.3040 (2)0.5602 (5)0.7774 (3)0.0615
Pt10.496514 (4)0.231576 (8)0.647481 (7)0.0213
Cl20.48843 (4)0.27493 (7)0.45529 (5)0.0362
Cl10.60019 (3)0.08476 (7)0.67596 (5)0.0328
N10.43076 (12)0.0464 (2)0.59437 (19)0.0321
O10.40504 (9)0.35648 (18)0.62161 (14)0.0292
N20.43773 (11)0.4095 (2)0.80464 (16)0.0281
H330.31590.75960.62980.1098*
H320.39650.72320.69810.1102*
H310.36620.67400.57390.1097*
H430.22600.53500.56670.0900*
H420.25460.37160.60550.0900*
H410.28210.45810.51690.0899*
H530.26170.62410.75770.0917*
H520.34160.60950.83480.0921*
H510.29230.46290.80410.0924*
H210.43220.46800.85600.0346*
H120.40380.02930.63600.0492*
H110.45710.03530.59450.0494*
H130.40170.06190.52530.0497*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
C10.0222 (8)0.0235 (8)0.0276 (9)0.0025 (6)0.0046 (7)0.0002 (7)
C20.0248 (10)0.0331 (11)0.0390 (12)0.0070 (8)0.0078 (9)0.0013 (9)
C30.056 (2)0.0418 (17)0.126 (4)0.0180 (15)0.032 (2)0.033 (2)
C40.0328 (14)0.080 (2)0.055 (2)0.0133 (15)0.0056 (13)0.0156 (17)
C50.0392 (16)0.095 (3)0.0502 (19)0.0221 (17)0.0139 (14)0.0106 (18)
Pt10.02004 (3)0.02133 (3)0.02118 (3)0.00012 (3)0.004499 (19)0.00125 (3)
Cl20.0471 (3)0.0360 (3)0.0259 (2)0.0030 (3)0.0119 (2)0.0024 (2)
Cl10.0290 (2)0.0361 (3)0.0340 (3)0.0074 (2)0.0107 (2)0.0035 (2)
N10.0307 (9)0.0298 (9)0.0342 (10)0.0058 (7)0.0081 (8)0.0074 (7)
O10.0271 (7)0.0315 (8)0.0254 (7)0.0071 (6)0.0029 (6)0.0019 (6)
N20.0286 (9)0.0270 (8)0.0264 (8)0.0072 (7)0.0053 (7)0.0015 (6)
Geometric parameters (Å, º) top
C1—C21.526 (3)C5—H520.959
C1—O11.293 (3)C5—H510.980
C1—N21.294 (3)Pt1—N2i1.9989 (18)
C2—C31.524 (4)Pt1—Pt1i2.5661 (2)
C2—C41.522 (4)Pt1—Cl22.4282 (6)
C2—C51.538 (5)Pt1—Cl12.3214 (6)
C3—H330.963Pt1—N12.0576 (19)
C3—H320.961Pt1—O12.0243 (16)
C3—H310.964N1—H120.864
C4—H430.969N1—H110.886
C4—H420.966N1—H130.891
C4—H410.965N2—H210.867
C5—H530.962
C2—C1—O1116.14 (19)H52—C5—H51110.0
C2—C1—N2122.2 (2)N2i—Pt1—Pt1i82.28 (6)
O1—C1—N2121.65 (19)N2i—Pt1—Cl290.82 (6)
C1—C2—C3108.1 (2)Pt1i—Pt1—Cl2170.852 (16)
C1—C2—C4109.2 (2)N2i—Pt1—Cl187.80 (6)
C3—C2—C4110.9 (3)Pt1i—Pt1—Cl194.280 (16)
C1—C2—C5110.8 (2)Cl2—Pt1—Cl191.44 (2)
C3—C2—C5109.1 (3)N2i—Pt1—N1177.99 (9)
C4—C2—C5108.7 (3)Pt1i—Pt1—N199.37 (6)
C2—C3—H33110.5Cl2—Pt1—N187.66 (7)
C2—C3—H32110.5Cl1—Pt1—N190.92 (6)
H33—C3—H32108.3N2i—Pt1—O193.18 (8)
C2—C3—H31110.0Pt1i—Pt1—O185.94 (5)
H33—C3—H31109.1Cl2—Pt1—O188.46 (5)
H32—C3—H31108.4Cl1—Pt1—O1179.02 (5)
C2—C4—H43109.2N1—Pt1—O188.10 (8)
C2—C4—H42109.5Pt1—N1—H12111.6
H43—C4—H42109.0Pt1—N1—H11111.0
C2—C4—H41109.9H12—N1—H11108.1
H43—C4—H41108.8Pt1—N1—H13109.5
H42—C4—H41110.4H12—N1—H13108.4
C2—C5—H53110.0H11—N1—H13108.2
C2—C5—H52109.4C1—O1—Pt1118.96 (13)
H53—C5—H52107.9C1—N2—Pt1i123.31 (15)
C2—C5—H51109.7C1—N2—H21118.4
H53—C5—H51109.9Pt1i—N2—H21117.4
Symmetry code: (i) x+1, y, z+3/2.
 

Acknowledgements

The authors gratefully acknowledge Professor F. P. Intini at the University of Bari for the NMR data collection. Open access funding enabled and organized by Projekt DEAL.

References

First citationAbe, T., Moriyama, H. & Matsumoto, K. (1991). Inorg. Chem. 30, 4198–4204.  CrossRef CAS Web of Science Google Scholar
First citationAli, M. S., Ali Khan, S. R., Ojima, H., Guzman, I. Y., Whitmire, K. H., Siddik, Z. H. & Khokhar, A. R. (2005). J. Inorg. Biochem. 99, 795–804.  Google Scholar
First citationAltomare, A., Cascarano, G., Giacovazzo, C., Guagliardi, A., Camalli, M., Burla, M. C. & Polidori, G. (1993). Acta Cryst. A49, c55.  Google Scholar
First citationBandoli, G., Dolmella, A., Intini, F. P., Pacifico, C. & Intini, G. (2003). Inorg. Chim. Acta, 346, 143–150.  Google Scholar
First citationBellitto, C., Flamini, A., Gastaldi, L. & Scaramuzza, L. (1983). Inorg. Chem. 22, 444–449.  CrossRef CAS Web of Science Google Scholar
First citationBetteridge, P. W., Carruthers, J. R., Cooper, R. I., Prout, K. & Watkin, D. J. (2003). J. Appl. Cryst. 36, 1487–1487.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBoulikas, T., Pantos, A., Bellis, E. & Christofis, P. (2007). Cancer Ther. 5, 537–583.  Google Scholar
First citationBruker (2003). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationCervantes, G., Prieto, M. J. & Moreno, V. (1997). Met.-Based Drugs, 4, 9–18.  Google Scholar
First citationChen, W. & Matsumoto, K. (2003). Inorg. Chim. Acta, 342, 88–96.  Google Scholar
First citationCornacchia, D., Pellicani, R. Z., Intini, F. P., Pacifico, C. & Natile, G. (2009). Inorg. Chem. 48, 10800–10810.  Google Scholar
First citationCotton, F. A., Matonic, J. H. & Murillo, C. A. (1997). Inorg. Chim. Acta, 264, 61–65.  Google Scholar
First citationDolmella, A., Intini, F. P., Pacifico, C., Padovano, G. & Natile, G. (2002). Polyhedron, 21, 275–280.  Google Scholar
First citationErxleben, A., Metzger, S., Britten, J. F., Lock, C. J. L., Albinati, A. & Lippert, B. (2002). Inorg. Chim. Acta, 339, 461–469.  Google Scholar
First citationFarrell, N., Kelland, L. R., Roberts, J. D. & Van Beusichem, M. (1992). Cancer Res. 52, 5065–5072.  Google Scholar
First citationFedotova, T. N., Minacheva, L. K., Kuznetsova, G. N., Sakharova, V. G., Gelfman, M. I. & Baranovskii, I. B. (1997). Russ. J. Inorg. Chem. 42, 1838–1846.  Google Scholar
First citationFerrugia, L. J. (1999). J. Appl. Cryst. 32, 837–838.  Google Scholar
First citationFerrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Google Scholar
First citationGonzález, V. M., Fuertes, M. A., Pérez-Alvarez, M. J., Cervantes, G., Moreno, V., Alonso, C. & Pérez, J. M. (2000). Biochem. Pharmacol. 60, 371–379.  Google Scholar
First citationGoodgame, D. M. L., Rollins, R. W., Slawin, A. M. Z., Williams, D. J. & Zard, P. W. (1986). Inorg. Chim. Acta, 120, 91–101.  Google Scholar
First citationGražulis, S., Chateigner, D., Downs, R. T., Yokochi, A. F. T., Quirós, M., Lutterotti, L., Manakova, E., Butkus, J., Moeck, P. & Le Bail, A. (2009). J. Appl. Cryst. 42, 726–729.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHartmann, J. T. & Lipp, H. P. (2003). Expert Opin. Pharmacother. 4, 889–901.  Google Scholar
First citationHollis, L. S., Roberts, M. M. & Lippard, S. J. (1983). Inorg. Chem. 22, 3637–3644.  CrossRef CAS Web of Science Google Scholar
First citationKasparkova, J., Marini, V., Najajreh, Y., Gibson, D. & Brabec, V. (2003a). Biochemistry, 42, 6321–6332.  Google Scholar
First citationKasparkova, J., Novakova, O., Farrell, N. & Brabec, V. (2003b). Biochemistry, 42, 792–800.  Google Scholar
First citationKelland, L. R. (2007). Nat. Rev. Cancer, 7, 573–584.  Google Scholar
First citationKuo, M. T., Chen, H. H., Song, I. S., Savaraj, N. & Ishikawa, T. (2007). Cancer Metastasis Rev. 26, 71–83.  Google Scholar
First citationLippert, B. (1999). Coord. Chem. Rev. 182, 263–295.  Google Scholar
First citationLippert, B., Raudaschl, G., Lock, C. J. L. & Pilon, P. (1984). Inorg. Chim. Acta, 93, 43–50.  Google Scholar
First citationLippert, B., Schoellhorn, H. & Thewalt, U. (1986). Inorg. Chem. 25, 407–408.  Google Scholar
First citationLippert, B., Schollhorn, H. & Thewalt, U. (1983). Z. Naturforsch. 38, 1441–1445.  Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMatsumoto, K. (2003). Russ. Chem. Bull. 52, 2577–2587.  Google Scholar
First citationMatsumoto, K., Matsunami, J., Mizuno, K. & Uemura, H. (1996). J. Am. Chem. Soc. 118, 8959–8960.  Google Scholar
First citationMatsumoto, K. & Sakai, K. (1999). Adv. Inorg. Chem. 49, 375–427.  CrossRef Google Scholar
First citationMontero, E. I., Díaz, S., González-Vadillo, A. M., Pérez, J. M., Alonso, C. & Navarro-Ranninger, C. (1999). J. Med. Chem. 42, 4264–4268.  Google Scholar
First citationNardelli, M. (1995). J. Appl. Cryst. 28, 659–659.  CAS Google Scholar
First citationNatile, G. & Coluccia, M. (2001). Coord. Chem. Rev. 216217, 383–410.  CrossRef CAS Google Scholar
First citationNonius (2001). COLLECT. Nonius BV, Delft, The Netherlands.  Google Scholar
First citationOchiai, M., Lin, Y.-S., Yamada, J., Misawa, H., Arai, S. & Matsumoto, K. (2004). J. Am. Chem. Soc. 126, 2536–2545.  Google Scholar
First citationO'Dwyer, P., Stevenson, J. & Johnson, S. (1999). In Cisplatin: Chemistry and Biochemistry of a Leading Anticancer Drug, edited by B. Lippert. Zurich: Verlag Helvetica Chimica Acta.  Google Scholar
First citationO'Halloran, T. V. & Lippard, S. J. (1985). Isr. J. Chem. 25, 130–137.  Google Scholar
First citationOxford Diffraction (2002). CrysAlis. Oxford Diffraction Ltd, Abingdon, Oxfordshire, England.  Google Scholar
First citationPacifico, C., Intini, F. P., Nushi, F. & Natile, G. (2010). Bioinorg. Chem. Appl. 2010, 102863.  Google Scholar
First citationRosenberg, B., Van Camp, L. & Krigas, T. (1965). Nature, 205, 698–699.  CrossRef PubMed CAS Web of Science Google Scholar
First citationSaeki, N., Nakamura, N., Ishibashi, T., Arime, M., Sekiya, H., Ishihara, K. & Matsumoto, K. (2003). J. Am. Chem. Soc. 125, 3605–3616.  Google Scholar
First citationShamsuddin, S., Ali, M. S., Whitmire, K. H. & Khokhar, A. R. (2007). Polyhedron, 26, 637–644.  Google Scholar
First citationSheldrick, G. M. (2010). SADABS. University of Gottingen, Gottingen, Germany.  Google Scholar
First citationSigel, R. K. O., Sabat, M., Freisinger, E., Mower, A. & Lippert, B. (1999). Inorg. Chem. 38, 1481–1490.  Google Scholar
First citationUmakoshi, K., Kinoshita, I., Ichimura, A. & Ooi, S. (1987). Inorg. Chem. 26, 3551–3556.  Google Scholar
First citationUsón, R., Forniés, J., Falvello, L. R., Tomas, M., Casas, J. M., Martin, A. & Cotton, F. A. (1994). J. Am. Chem. Soc. 116, 7160–7165.  Google Scholar
First citationVon Hoff, D. D., Schilsky, R., Reichert, C. M., Reddick, R. L., Rozencweig, M., Young, R. C. & Muggia, F. M. (1979). Cancer Treat. Rep. 63, 1527–1531.  CAS PubMed Web of Science Google Scholar
First citationWatkin, D. J., Prout, C. K. & Pearce, L. J. (1996). CAMERON. Chemical Crystallography Laboratory, Oxford, England.  Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Follow Acta Cryst. B
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds