Oxidation of a rollover cycloplatinated(II) dimer by MeI: a kinetic study

Reza Babadi Aghakhanpoura, Mehdi Rashidi*a, Fatemeh Niroomand Hosseinib, Fatemeh Raoofa and S. Masoud Nabavizadeh*a
aDepartment of Chemistry, College of Sciences, Shiraz University, Shiraz, 71454, Iran. E-mail: rashidi@chem.susc.ac.ir; nabavi@chem.susc.ac.ir
bDepartment of Chemistry, Shiraz Branch, Islamic Azad University, Shiraz, Iran

Received 24th June 2015 , Accepted 30th July 2015

First published on 30th July 2015


Abstract

The known rollover cycloplatinated(II) complex [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, in which bpy-2H acts as a bridging rollover ligand, was reacted with MeI to give a new binuclear rollover cycloplatinated(IV) complex [Pt2Me4I2(PPh3)2(μ-bpy-2H)], 2. The stereochemistry of 2 was fully identified by NMR spectroscopy (1H and 31P) and confirmed by DFT calculations. To the best of our knowledge, 2 is the first example of a diplatinum(IV) complex having a bridging rollover bipyridine ligand. 1 has a 5dπ(Pt) → π*(bpy) MLCT band in the visible region which was used to easily follow the kinetics of its reaction with MeI; a double MeI oxidative addition was observed and classical SN2 mechanism was suggested for both steps of the reaction. The large negative entropy of activation (ΔS), found in each step, complies with an associative process. The rates are almost 3–5 times slower in the second step as compared to the first step, due the electronic effects transmitted through the rollover bpy ligand. The rates were also compared with that reported for the corresponding monomeric cyclometalated complex [PtMe(bpy-H)(PPh3)] and found to be higher (in step 1) and usually lower (in step 2). Theoretical computations of the geometry of the possible reaction transition states and intermediates revealed that each step of the reaction takes place via a transition state with a nearly linear arrangement of the I–CH3–Pt moiety. The computational results are in good agreement with the experimental findings, confirming the proposed mechanism.


1. Introduction

The chemistry of cyclometalated compounds is currently one of the most active topics in organic synthesis and homogeneous catalysis. The C–H bond activation using a directing group such as 2-phenylpyridine or 2,2′-bipyridine is a suitable method for achieving easy C–H bond functionalization.1–3 2-Phenylpyridine is a typical compound, and its reaction with a transition metal complex forms an orthometalated intermediate with the help of nitrogen coordination to the metal. A new class of cyclometalation is C–H bond metalation to form rollover complexes and the most studied ligand in this area is 2,2′-bipyridine (bpy).4 In this type of cyclometalation, the two nitrogen atoms of bpy at first coordinate to metal and then the coordination of one nitrogen to the metal is switched to carbon–metal bond by rotation of the pyridine ring (Scheme 1).5,6 Although rollover cyclometalated complexes are currently reported to be very easy to prepare, their reactivity have not been studied in-depth.
image file: c5ra12201e-s1.tif
Scheme 1 Classical and rollover cyclometalation.

Oxidative addition of MeI to platinum(II) centers is a fundamental process in organometallic chemistry with significant implications in catalysis. A classic example of such a system is the oxidative addition of MeI to [PtX4]2− complex in the Shilov process of activation of C–H bonds.7 It is well-known that electron-rich Pt(II) complexes (e.g. [PtMe2(NN)], in which NN are various diimine ligands such as 2,2′-bipyridine and 1,10-phenanthroline) undergo faster oxidative addition than electron-poor Pt(II) complexes.8–10 Electron density on the Pt(II) center of a [PtR2(LL)] complex can be manipulated by selecting R ligand, which has either electron-withdrawing or electron-donating ability.11–14

Our previous study has described in detail the effect of usual cyclometalated groups on rate of oxidative addition of MeI to the complexes of type [PtMe(C^N)L], in which C^N = 2-phenylpyridinate, 2-(p-tolyl)pyridinate or benzo[h]quinolate and L = PPh3 or PPh2Me.15 However, report on oxidative addition of alkyl halides on rollover cyclometalated platinum(II) complexes are very rare,9,16 with only one report existing on kinetic study of this type of complexes with MeI in which effect of rollover cyclometalated group on the rate of oxidative addition is investigated.9 There, a study of reaction between MeI and the monomeric Pt(II) complexes [PtMe(bpy-H)(L)] (bpy-H = κ2N,C-2,2′-bipyridine; L = PPh3 or PPh2Me)] led to formation of the Pt(IV) rollover complexes [PtMe2I(bpy-H)(L)]. Despite of this, any such reactions involving the related binuclear complexes are not reported. We report here, for the first time, on the reactivity toward oxidative addition of MeI with dimeric rollover cyclometalated Pt(II) complex, [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, to investigate the cooperative steric or electronic effects between the two adjacent metal centers. Finally, using density functional theory for structure optimizations on the related complexes, geometries of transition states and intermediates in the complete reaction sequence are proposed.

2. Experimental section

The 1H and 31P NMR spectra were recorded on a Bruker Avance DRX 400 MHz spectrometer in CDCl3 as solvent. The operating frequencies and references, respectively, are shown in parentheses as follows: 1H (400 MHz, TMS) and 31P (162 MHz, 85% H3PO4). The chemical shifts and coupling constants are in ppm and Hz, respectively. Kinetic studies were carried out by using a Perkin-Elmer Lambda 25 spectrophotometer with temperature control using an EYELA NCB-3100 constant temperature bath. The precursor [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, was prepared according to reported procedure.17

2.1. Synthesis of [Pt2Me4I2(PPh3)2(μ-bpy-2H)], 2

To a solution of [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, (50 mg in 25 mL toluene) was added an excess of MeI (550 μL) at room temperature and the mixture was stirred for 5 days. The solvent was removed under reduced pressure, and the residue was washed with diethyl ether and acetone. The product was dried under vacuum. Yield 40 mg; 65%, mp 203 °C (decomp.). Anal. calcd for C50H48I2N2P2Pt2: C, 43.4; H, 3.5; N, 2.0; found: C, 43.0; H, 3.2; N, 1.8. NMR data: δ(1H) 1.15 (d, 3JPH = 12.0 Hz, 2JPtH = 60.0 Hz, 6H, Me group trans to P), 1.64 (d, 3JPH = 12.0 Hz, 2JPtH = 75.0 Hz, 3H, Me groups trans to N), (aromatic protons): 7.0–7.9 (overlapping multiplets of phenyl groups of PPh3), 8.80 (d, 3JHH = 7.9 Hz, 2H, CH groups adjacent to coordinated N atoms), 6.84 (d, 3JHH = 12.0 Hz, 3JPtH = 64 Hz, 2H, CH groups adjacent to coordinated C atoms), 6.52 (t, 3JHH = 20.0 Hz, 2H); δ(31P) 10.4 (s, 1JPtP = 1004 Hz, 2P).

2.2. Kinetic study

The oxidative addition reaction was monitored using UV-vis (by monitoring the change in absorbance at λ = 375 nm) spectrophotometer. All kinetic measurements were monitored under pseudo-first-order conditions in CHCl3 solution. Concentration of [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, was 0.0003 M for UV-vis measurements. Concentration range used for MeI was 0.1–1.5 M. Kinetic measurements, under pseudo-first-order conditions for different concentrations of 1 (ranges from 2 × 10−4 to 4 × 10−4 M) at a constant [MeI], confirmed that the Pt complex concentration had no influence on value of the observed rate constant. The observed first-order rate constants were obtained from least-square fits of absorbance versus time data according to eqn (1).

2.3. Computational details

Gaussian 03 was used18 to fully optimize all the structures at the B3LYP level of density functional theory. The solvation energies were calculated by CPCM model in CHCl3. The effective core potential of Hay and Wadt with a double-ξ valence basis set (LANL2DZ) was chosen to describe Pt and I.19,20 The 6-31G(d) basis set was used for other atoms. Frequency calculations were carried out at the same level of theory to identify whether the calculated stationary point is a minimum (zero imaginary frequency) or a transition structure (one imaginary frequency).

3. Results and discussion

3.1. Synthesis and characterization of the rollover dimeric Pt(IV)–Pt(IV) complex 2

The studied reaction is depicted in Scheme 2.
image file: c5ra12201e-s2.tif
Scheme 2 Reaction studied in the present work.

The known starting rollover methylplatinum(II) [Pt2Me2(PPh3)2(μ-bpy-2H)], 1,17 was reacted cleanly with an excess of MeI in toluene at room temperature to give the air-stable solid product [Pt2Me4I2(PPh3)2(μ-bpy-2H)], 2. The synthesized Pt(IV) complex was fully characterized using 1H and 31P NMR spectroscopy and elemental analysis (full data are collected in the Experimental section). Thus, structure of product of the oxidative addition, 2, is proposed to adopt an octahedral geometry in which the rollover bpy ligand, iodide and Me groups are located in an equatorial plane with the methyl and PPh3 ligands in the axial positions. In 1H NMR spectrum of 2, the two equivalent Me groups trans to P were observed at δ 1.15 as a doublet with 3JPH = 12.0 and 2JPtH = 60.0 Hz, while the two equivalent Me groups locating trans to N ligating atoms were appeared at δ 1.64, with a considerably higher 2JPtH value of 75.0 Hz, due to lower trans influence of N atom as compared with that of P atom; these data are in agreement with those reported for the related monomeric complexes, such as [PtMe2I(bpy-H)(PPh3)] with chemical shift values of 1.12 (with 2JPtH = 61 Hz) and 1.64 ppm (with 2JPtH = 71 Hz) for Me groups trans to P and N atoms, respectively.21 Hydrogens related to CH groups adjacent to ligating N and C atoms of the rollover bpy-2H ligand each appeared as a doublet at δ = 8.80 (with 3JHH = 7.9 Hz) and 6.84 (with 3JHH = 12.0 Hz and 3JPtH = 64.0 Hz), respectively. In the 31P NMR spectrum of [Pt2Me4I2(PPh3)2(μ-bpy-2H)], 2, a singlet at δ 10.4 (with 1JPtP = 1004 Hz), assigned to the two equivalent P ligating atoms, was observed with its 1JPtP value, as expected, being much lower than the corresponding value of 1JPtP = 2347 Hz found for the starting Pt(II) complex 1.17 Although the data comply well with relative disposition of different ligands on each Pt center of the complex 2 (with chirality at each Pt center), it is not possible to use the present data to actually propose any “frozen” conformer(s) for complex 2 resulting from rotation around one or two of the Pt–P bonds. As twice the “expected” numbers of signals were observed in NMR spectra of 2, the formation of a statistical 1[thin space (1/6-em)]:[thin space (1/6-em)]1 mixture of two stereoisomers may be a reasonable explanation.

3.2. Kinetic studies

Kinetics of oxidative addition of MeI in CHCl3 to the dimeric rollover cycloplatinated(II) complex 1 was studied using UV-vis spectroscopy. An excess of MeI was used and disappearance of the MLCT band was followed to monitor the reaction. Changes in the spectrum during a typical run are shown in Fig. 1. Absorbance–time curves monitored at 375 nm for reaction of 1 with MeI in CHCl3 solution were obtained in the presence of a large excess of MeI, such that its concentration remained effectively constant in each experiment.
image file: c5ra12201e-f1.tif
Fig. 1 Changes in the UV-vis spectrum during reaction of 1 with MeI in CHCl3; successive spectra were recorded at intervals of 150 s.

Time-dependence curves of spectra of the reaction at this condition are shown in Fig. 2. The data failed to fit properly in equation Abst = Abs + (Abs0 − Abs)[exp(−kobst)], which corresponds to a monophasic kinetic behavior. However, the data were successfully fitted in eqn (1) with two exponentials, showing the occurrence of two sequential first-order reactions. Thus, pseudo-first-order rate constants, kobs(1) and kobs(2), for the two steps of the reaction were calculated by nonlinear least-square fitting kinetic data to the biphasic first order equation (eqn (1) and Fig. 2).

 
Abst = α[exp(−kobs(1)t)] + β[exp(−kobs(2)t)] + Abs (1)


image file: c5ra12201e-f2.tif
Fig. 2 Absorbance–time curves for reaction of 1 with MeI (0.27–1.33 M with [MeI] increases reading downward) in CHCl3 at 15 °C. The inset shows the biphasic fit to the experimental curve when [MeI] = 0.27 M.

The experimentally determined pseudo-first-order rate constants, kobs(1) and kobs(2), were converted to second order rate constants (k2 for the first step and k2 for the second step) by determining slope of the linear plots of kobs(1) and kobs(2) against the concentration of MeI reagent (Fig. 3). Non-zero intercepts implied that kobs(1) = k2[MeI] + k1 for the first step and kobs(2) = k2[MeI] + k1 for the second step and that the k1 or k1 step in the proposed reaction mechanism should exist. Thus, in addition to second-order terms k2 and k2, which are suggested for the simple associative SN2-type mechanism, a much smaller terms, k1 and k1, are also observed for associative substitutions by solvent. The results are given in Table 1. The findings are consistent with the two-step sequence shown in Scheme 3 for reaction of 1 with MeI. It should be noted that k2 > k2. The same method was used at other temperatures and activation parameters were obtained from the Eyring equation (eqn (2)):

 
image file: c5ra12201e-t1.tif(2)


image file: c5ra12201e-f3.tif
Fig. 3 Plots of first-order rate constants for the reaction of 1 with MeI at 25 °C versus [MeI] in CHCl3.
Table 1 Rate constantsa and activation parametersb for reaction of 1 with MeI in CHCl3
  Rate constants at different temperatures
15 °C 20 °C 25 °C 30 °C 40 °C ΔH/kJ mol−1 ΔS/J K−1 mol−1
a Estimated error in k values are ±5%.b Obtained from the Eyring equation.c From ref. 9.
103k2/(L mol−1 s−1) 2.83 3.72 4.83 7.02 11.74 40.9 ± 1.5 −152 ± 5
103k1/s−1 0.74 0.81 1.12 26.6 ± 9.9 −210 ± 33
103k2/(L mol−1 s−1) 0.59 0.78 0.99 2.22 3.16 51.7 ± 7.5 −127 ± 25
103k1/s−1 0.20 0.25 0.39 44.9 ± 9.5 −160 ± 33
103k2/(L mol−1 s−1) for [PtMe(bpy-H)(PPh3)]c 0.70 0.90 1.50 2.00 4.60 55.1 ± 3.2 −114 ± 10



image file: c5ra12201e-s3.tif
Scheme 3 Suggested mechanism for the two-step oxidative addition reaction of 1 with MeI (S = solvent).

In this equation ΔH = activation enthalpy, ΔS = activation entropy, k = rate constant, kB = Boltzmann's constant, T = temperature, h = Planck's constant, R = universal gas constant (see Fig. 4). The data are shown in Table 1.


image file: c5ra12201e-f4.tif
Fig. 4 Eyring plots for the reaction of 1 with MeI in CHCl3.

ΔS values are large and negative for the both steps, complying with the classical SN2 type mechanism. Oxidative addition of organoplatinum(II) complexes with alkyl halides have been extensively investigated8,22 and we have also studied a secondary α-deuterium KIE study involving the reaction of MeI/CD3I to confirm operation of SN2 mechanism in the oxidative addition of MeI to some organoplatinum(II) complexes.23–25 So it is well established that these reactions proceed by an SN2 mechanism. Although a redox pathway is also a possibility, it has never been demonstrated by experimental evidence. On the basis of these kinetic data and the related DFT calculations (described in the next section) we propose that (see Scheme 3) in the first step, one of the electron rich Pt(II) centers of [Pt2Me2(PPh3)2(μ-bpy-2H)], 1 (as a nucleophile), attacks on carbon atom of MeI through an SN2 type mechanism (with rate constant k2) and the mix valence Pt(II)–Pt(IV) kinetic binuclear intermediate IM-A is formed, via the transition state TS1, which is then equilibrated with the analogous intermediate IM-B with Me and I being in cis disposition at Pt(IV) center. The latter intermediate, IM-B, is then converted to the Pt(II)–Pt(IV) compound [PtMe(PPh3)(μ-bpy-2H)PtMe2I(PPh3)], A. The complex A is then reacted with MeI in the second step, again by an SN2 type mechanism, with the rate constant k2 being nearly 3–5 times smaller than value for rate constant of the first step (i.e. k2, see Table 1) and the Pt(IV)–Pt(IV) binuclear intermediate IM-C in equilibrium with the analogous intermediate IM-D is formed via the transition state TS2. The latter intermediate, IM-D, gives the final product [Pt2Me4I2(PPh3)2(μ-bpy-2H)], 2.

On the basis of the UV-vis data for the two-step oxidative addition reaction studied in this work (see Table 1) and using the Espenson's approach, in which the biphasic equation was used in evaluating the molar absorptivities of intermediates and the two rate constants,26–28 the second step seems to be slower than the first step by a factor of nearly 3 to 5. This is suggested to be due to the fact that in the first step one of the Pt(II) centers of Pt(II)–Pt(II) complex 1 attacks on methyl group of MeI, while in the second step, MeI reacts with the Pt(II) center of the Pt(II)–Pt(IV) complex A, in which the adjacent Pt(IV) reduces nucleophilicity of the Pt(II) center (through the bridging rollover bpy-2H ligand) when compared with situation in the first step during which the attack is performed on a Pt(II)–Pt(II) species.

For the purpose of comparison, the kinetic data for reaction of the related monomeric complex [PtMe(bpy-H)(PPh3)] with MeI9 are also included in Table 1. Thus, on reaction with MeI, 1 in the first step reacts nearly 3 time faster than the monomer [PtMe(bpy-H)(PPh3)]; we attribute this to less electron withdrawing ability of the neighboring Pt(II) center in 1 to increase the electron density of the attacking Pt(II) center in Pt(II)–Pt(II) complex 1 when compared with the case of monomer. As such, 1 in the second step is reacted with MeI with usually lower rate constant.

3.3. DFT calculations

In order to shed some light on the suggested mechanism of double oxidative addition reaction involving [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, with MeI (depicted on Scheme 3), DFT calculations were carried out on the precursor complex 1, final product 2, potential transition states and intermediates. The Conductor-like Polarizable Continuum Model (CPCM) was used to model general solvation effects by chloroform. The reaction is initiated by nucleophilic attack by 5dz2 HOMO−1 of one of the Pt(II) centers of 1 on the σ* LUMO of MeI (Fig. 5). Substitution of iodide by the Pt center, followed by simultaneous partial removal of an iodide ion, results in formation of the transition state TS1 (see Fig. 6). In TS1, the bond angles I–CMe–Pt (177.6°) and Pt–CMe–H (89.7°) are close to 180° and 90°, respectively, showing a linear I⋯CMe⋯Pt arrangement with the hydrogen atoms of CH3 group forming a trigonal bipyramidal arrangement around the C center of methyl group. During the formation of TS1, the most significant changes in bond distances are observed for I⋯Me and Pt⋯Me (Me is from MeI) distances increasing from 2.227 Å in MeI to 2.698 Å in TS1, whereas the Pt⋯Me distance decreases from far apart (randomly chosen) in the reactant to 2.517 Å in the TS1. This confirms that oxidative addition reaction of MeI to the rollover dimeric complex 1 involves simultaneous cleavage of Me–I bond and formation of the Pt–Me bond, which is then followed by formation of the cationic intermediate IM-A by completely breaking and forming of Me–I and Pt–Me bonds, respectively. The intermediate IM-A has a square pyramidal geometry, with the incoming Me group locating in the apical position and the iodide ion in the outer sphere of dimer complex. As was mentioned before, oxidative addition of MeI to 1 yielded almost quantitatively the cis addition isomer 2, which is the stable thermodynamic product. Therefore the intermediate IM-A is quickly performing a facile trans to cis isomerization of Pt(IV) center to form the intermediate IM-B having Me and I in cis disposition. The free iodide ion coordinates to the platinum(IV) center of IM-B forming the mixed Pt(II)–Pt(IV) complex A having an octahedral geometry around the Pt(IV) center (see Scheme 3 and Fig. 6) with Pt–I and Pt–CMe bond lengths being 2.907 and 2.087 Å, respectively; notice that although the intermediate IM-A is theoretically shown to be slightly more stable (see Fig. 7) than the intermediate IM-B, from sterical point of view the latter is more favorable in forming A. As expected (see Fig. 6), in A, bond lengths of the Pt(II) center are shorter than those of the Pt(IV) center; for example Pt–CMe and Pt–P bonds (2.065 and 2.409 Å, respectively) in the Pt(II) center of A are shorter than those in the Pt(IV) center (2.087 and 2.608 Å, respectively).
image file: c5ra12201e-f5.tif
Fig. 5 (a) HOMO−1 of rollover dimer 1, (b) LUMO of MeI, (c) interactions of dz2 orbital of electron-rich Pt(II) center of dimer 1 with σ* LUMO of MeI and (d) HOMO of TS1.

image file: c5ra12201e-f6.tif
Fig. 6 The B3LYP/(LANL2DZ, 6-31G(d))/chloroform optimized structures involved during oxidative addition reaction of [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, with CH3I. The H atoms are removed for clarity (except for the methyl group of the CH3I). Selected bond angles (°) and bond lengths (Å) are also shown.

image file: c5ra12201e-f7.tif
Fig. 7 Energy profile for oxidative addition of MeI to [Pt2Me2(PPh3)2(μ-bpy-2H)], 1, in CHCl3.

A is then reacted with MeI in the second step, again by an SN2 type mechanism, upon which iodide separates from carbon of the MeI reagent and the transition state TS2 is formed (see Scheme 3 and Fig. 6 and 7), having an approximately collinear arrangement of the I–CH3–Pt moiety (the angle being 176.8°) with the Pt–CMe–H angle (91.8°) close to 90°, corresponding to an SN2 type of reaction. A similar trend is observed in the second step giving the transition state TS2 following by formation of the intermediate IM-C which is then equilibrate with the intermediate isomer IM-D,15 to finally give the thermodynamic isomer 2. Formation of the thermodynamic isomer 2 over the kinetic isomer (in which iodide groups are located trans to the incoming Me groups) is favored as in 2 the more sterically demanding PPh3 ligand is situated in “axial” position when compared with that in the kinetic isomer in which it is located in the “equatorial” position.

Although formation of 2 through a concerted pathway is a possibility, the theoretical suggested SN2 mechanism is consistent with the experimental finding.

Calculated structures and energies for the compounds, shown in Scheme 3, in chloroform solution (Fig. 6 and 7) indicate that, in consistent with the experimental observations, 1 is easily reacted with MeI to give A which is followed by a slower reaction to give 2. Calculated energies of the transition states TS1 (+45.5 kJ mol−1) and TS2 (+48.2 kJ mol−1) are in excellent agreement with the observed values of ΔH = 40.9 and 51.7 kJ mol−1 for the first and second steps, respectively.

A summary of calculated atomic charges (based on Atomic Polar Tensors population (APT) analysis) for the selected compounds of Scheme 3 is given in Table 2. Charges on Pt(II) centers of 1 are significantly lower than those on the other compounds, supporting that 1 has the most potential to act as nucleophile in the reaction with MeI (Fig. 5). Atomic charges on Pt(IV) and Pt(II) centers of the mixed valence Pt(IV)–Pt(II) complex A are +0.553 and −0.201, respectively. As one moves from 1 to 2, the electron population on the iodide atom increases. These results confirm the fact that the oxidative addition reaction in such systems happens through transfer of electron density from Pt center to Me group and then to the iodide atom.

Table 2 APT (Atomic Polar Tensors) atomic charges of the selected compounds
Compounds qPt(1)/qPt(2) qI(1)a/qI(2)b qC(1)a/qC(2)b
a First incoming MeI.b Second incoming MeI.
1 −0.204/−0.204 −0.236/−0.236 +0.274/+0.274
IM-B +0.440/−0.189 −1/−0.236 +0.203/+0.274
A +0.553/−0.201 −0.833/−0.236 +0.306/+0.274
IM-D +0.540/+0.469 −0.788/−1 +0.290/+0.211
2 +0.544/+0.554 −0.809/−0.809 +0.298/+0.298


A further rationale of oxidative addition process in terms of nature of frontier molecular orbitals of the Pt compounds (Fig. 8) may be helpful. Composition and energy of HOMO and LUMO of 1 and A are also reported in Fig. 8. The highest occupied molecular orbitals 1 consist mainly of the Pt orbitals (42% and 84% for HOMO and HOMO−1, respectively) with some contribution from orbitals of the bridged bpy-2H ligand (53 and 8% for HOMO and HOMO−1, respectively). The LUMO in 1 is predominately localized on the bpy-2H unit orbitals. The HOMO and HOMO−1 of A consist of the iodide ligand and the LUMO of this complex is mainly localized on the bpy-2H ligand (see Fig. 8). Value of energy separations between the HOMO and LUMO of 1 is greater than that of A and are equal to 4.157 and 3.588 eV, respectively, meaning that the kinetic stability of A is lower than 1. Main contribution to Pt⋯CMeI in the transition state TS1 comes from overlap of dz2 HOMO−1 of the platinum center with the LUMO of MeI (which is the C–I σ* antibonding molecular orbital, see Fig. 5). So it is reasonable to view the oxidation process formally as removal of electrons from HOMO−1 of 1 into the LUMO of MeI. The second step including the reaction of A with a second incoming MeI shows similar frontier orbitals (see Fig. 8).


image file: c5ra12201e-f8.tif
Fig. 8 Qualitative frontier molecular orbital scheme for 1 and A.

4. Conclusions

The rollover cyclometalated Pt(II)–Pt(II) dimeric complex [Pt2Me2(PPh3)2(μ-bpy-2H)], 1,17 is an interesting case of the less common complexes of this kind containing a bridging double deprotonated bipyridine ligand, in which the two Pt fragments are identical. As supported by experimental and theoretical investigations, a two-step oxidative addition reaction of MeI with 1 is suggested to occur via SN2 mechanism, during a number of stages (see Scheme 3), to finally give the thermodynamic Pt(IV)–Pt(IV) product [Pt2Me4I2(PPh3)2(μ-bpy-2H)], 2 (in which the added Me and I ligands on each of the two Pt centers are situated in cis position to each other); high steric demand of PPh3 ligands are probably responsible for isomerization of the kinetic product in each of the two steps (having the added Me and I ligands in trans position) to the thermodynamic analogous. Kinetics of the reaction, as investigated by UV-vis spectrophotometry and confirming by DFT studies, shows the following results:

(1) In the first step, one of the electron rich Pt centers of Pt(II)–Pt(II) starting complex 1 attacks the carbon of MeI to form A (suggested to be a Pt(II)–Pt(IV) complex) via the transition state TS1. The Pt(II) center in A is then reacted with MeI with a rate considerably slower (by a factor of nearly 3–5) than that in the first step involving the starting complex 1, which is in consistent with a significant electronic effect transmitted through the rollover bpy bridging ligand. Less importantly, the fact that coordination around A seems to be more sterically hindered than that of 1 is probably responsible for the rate differences.

(2) MeI in the first step reacts with 1 in CHCl3 at 25 °C with a rate constant k2 = 4.83 × 10−3 L mol−1 s−1, which is some 3 times faster than that with the related mononuclear cycloplatinated(II) complex [PtMe(bpy-H)(PPh3)] at the same condition (with k2 = 1.50 × 10−3 L mol−1 s−1). This is attributed to less electron withdrawing ability of neighboring Pt(II) center in 1 to increase the electron density of the under attacking Pt(II) center in the Pt(II)–Pt(II) complex 1 as compared to that in monomer [PtMe(bpy-H)(PPh3)].

(3) Rate of MeI oxidative addition in the second step involving 1 (see Table 1) is found to be usually lower than that involving the mononuclear complex [PtMe(bpy-H)(PPh3)].

Acknowledgements

This work has been supported by Shiraz University and the Iran National Science Foundation (Grant No. 93035108).

References

  1. T. Shibata, S. Takayasu, S. Yuzawa and T. Otani, Org. Lett., 2012, 14, 5106–5109 CrossRef CAS PubMed.
  2. C. S. Yeung and V. M. Dong, Chem. Rev., 2011, 111, 1215–1292 CrossRef CAS PubMed.
  3. S. H. Cho, J. Y. Kim, J. Kwak and S. Chang, Chem. Soc. Rev., 2011, 40, 5068–5083 RSC.
  4. A. Zucca, D. Cordeschi, L. Maidich, M. I. Pilo, E. Masolo, S. Stoccoro, M. A. Cinellu and S. Galli, Inorg. Chem., 2013, 52, 7717–7731 CrossRef CAS PubMed.
  5. G. L. Petretto, J. P. Rourke, L. Maidich, S. Stoccoro, M. A. Cinellu, G. Minghetti, G. J. Clarkson and A. Zucca, Organometallics, 2012, 31, 2971–2977 CrossRef CAS.
  6. B. Butschke and H. Schwarz, Organometallics, 2010, 29, 6002–6011 CrossRef CAS.
  7. A. E. Shilov and G. B. Shul'pin, Chem. Rev., 1997, 97, 2879–2932 CrossRef CAS PubMed.
  8. M. Crespo, M. Martínez, S. M. Nabavizadeh and M. Rashidi, Coord. Chem. Rev., 2014, 279, 115–140 CrossRef CAS PubMed.
  9. R. B. Aghakhanpour, S. M. Nabavizadeh, L. Mohammadi, S. A. Jahromi and M. Rashidi, J. Organomet. Chem., 2015, 781, 47–52 CrossRef CAS PubMed.
  10. S. M. Nabavizadeh, H. Sepehrpour, R. Kia and A. L. Rheingold, J. Organomet. Chem., 2013, 745, 148–157 CrossRef PubMed.
  11. M. Rashidi, M. Nabavizadeh, R. Hakimelahi and S. Jamali, J. Chem. Soc. Dalton Trans., 2001, 3430–3434 RSC.
  12. S. M. Nabavizadeh, H. Sepehrpour, H. R. Shahsavari and M. Rashidi, New J. Chem., 2012, 36, 1739–1743 RSC.
  13. S. M. Nabavizadeh, E. S. Tabei, F. N. Hosseini, N. Keshavarz, S. Jamali and M. Rashidi, New J. Chem., 2010, 34, 495–499 RSC.
  14. S. M. Nabavizadeh, H. R. Shahsavari, H. Sepehrpour, F. N. Hosseini, S. Jamali and M. Rashidi, Dalton Trans., 2010, 39, 7800–7805 RSC.
  15. S. M. Nabavizadeh, H. Amini, F. Jame, S. Khosraviolya, H. R. Shahsavari, F. N. Hosseini and M. Rashidi, J. Organomet. Chem., 2012, 698, 53–61 CrossRef CAS PubMed.
  16. L. Maidich, A. Zucca, G. J. Clarkson and J. P. Rourke, Organometallics, 2013, 32, 3371–3375 CrossRef CAS.
  17. A. Zucca, G. L. Petretto, S. Stoccoro, M. A. Cinellu, M. Manassero, C. Manassero and G. Minghetti, Organometallics, 2009, 28, 2150–2159 CrossRef CAS.
  18. M. J. Frisch, N. Rega, G. A. Petersson, G. W. Trucks, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, J. C. Burant, T. Nakajima, Y. Honda, O. Kitao, H. B. Schlegel, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, J. M. Millam, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, G. E. Scuseria, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, S. S. Iyengar, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, M. A. Robb, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, J. Tomasi, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. R. Cheeseman, J. V. Ortiz, Q. Cui, A. G. Baboul, V. Barone, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, J. A. Montgomery Jr, R. L. Martin, D. J. Fox, B. Mennucci, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, T. Vreven, M. W. Wong, M. Cossi, C. Gonzalez, J. A. Pople, K. N. Kudin and G. Scalmani, Gaussian 03, Revision C.02, 2004 Search PubMed.
  19. P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270–283 CrossRef CAS PubMed.
  20. E. S. Raper, Coord. Chem. Rev., 1985, 61, 115–184 CrossRef CAS.
  21. A. Zucca, L. Maidich, L. Canu, G. L. Petretto, S. Stoccoro, M. A. Cinellu, G. J. Clarkson and J. P. Rourke, Chem. Eur. J., 2014, 20, 5501–5510 CrossRef CAS PubMed.
  22. L. M. Rendina and R. J. Puddephatt, Chem. Rev., 1997, 97, 1735–1754 CrossRef CAS PubMed.
  23. P. Hamidizadeh, M. Rashidi, S. M. Nabavizadeh, M. Samaniyan, M. D. Aseman, A. M. Owczarzak and M. Kubicki, J. Organomet. Chem., 2015, 791, 258–265 CrossRef CAS PubMed.
  24. M. D. Aseman, M. Rashidi, S. M. Nabavizadeh and R. J. Puddephatt, Organometallics, 2013, 32, 2593–2598 CrossRef CAS.
  25. S. Habibzadeh, M. Rashidi, S. M. Nabavizadeh, L. Mahmoodi, F. N. Hosseini and R. J. Puddephatt, Organometallics, 2010, 29, 82–88 CrossRef CAS.
  26. M. Shimura and J. H. Espenson, Inorg. Chem., 1984, 23, 4069–4071 CrossRef CAS.
  27. S. Jamali, S. M. Nabavizadeh and M. Rashidi, Inorg. Chem., 2005, 44, 8594–8601 CrossRef CAS PubMed.
  28. S. M. Nabavizadeh, M. D. Aseman, B. Ghaffari, M. Rashidi, F. N. Hosseini and G. Azimi, J. Organomet. Chem., 2012, 715, 73–81 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.