DOI:
10.1039/C6RA21756G
(Paper)
RSC Adv., 2016,
6, 95073-95084
Carbon–sulfur bond reductive coupling from a platinum(II) thiolate complex†
Received
30th August 2016
, Accepted 29th September 2016
First published on 30th September 2016
Abstract
The room temperature addition of electrophilic alkyl halide reagents (RX = MeI, EtI and PhCH2Br) to complex [Pt(ppy)(η1-S-Spy)(PPh3)], 1a, in which ppyH = 2-phenylpyridine and pySH = pyridine-2-thiol, resulted in a rapid carbon–sulfur (C–S) bond reductive coupling to produce alkyl sulfides and corresponding halide complexes [Pt(ppy)(PPh3)X], X = I (2a) and Br (2b). A mechanism for this C–S bond formation reaction was suggested based on 1H and 31P {1H} NMR spectroscopic analyses. In the suggested mechanism, the reaction proceeded through a binuclear intermediate complex [{Pt(ppy)(PPh3)}2(μ2-Spy)]I, 3-I, which was separately synthesized by another counter anion (PF6) and it was fully characterized by multinuclear NMR spectroscopy and single X-ray crystallography. Also, density functional theory (DFT) calculations were used to theoretically assess the structures of intermediates and transition states in this bond formation reaction.
Introduction
Several prominent catalytic or stoichiometric systems were developed for C–H,1,2 C–C2–8 and C–X8–17 (X = different heteroatoms) bond formation reactions. Despite these extensive research efforts, C–S bond formation still has significant limitations identical to poisoning and/or deactivation effect of sulfur on reactive species or harsh reaction conditions.14–16 Such difficulties were addressed in some cases and numerous outstanding systems were established for C–S bond coupling16 using transition metals such as Ni,18 Cu,17 Pd8,16 and Pt.19 In addition, the growing interest in C–S bond construction chemistry is mostly due to its potential applications in organic synthesis,20 materials science,21 agricultural,15 and pharmaceutical industry.22
There are only a few reports on the formation of C–S bond using platinum transition metal complexes.19 This fact has led us to apply complex [Pt(ppy)(η1-S-Spy)(PPh3)], 1a, ppyH = 2-phenylpyridine and pySH = pyridine-2-thiol, in C–S bond reductive coupling with electrophilic reagents such as alkyl halides at room temperature. In addition, mechanistic aspect of this bond formation reaction was studied with multinuclear NMR spectroscopy by observing an interesting binuclear complex. Density functional theory (DFT) was employed to optimize structures of the related complexes, which were observed experimentally or proposed theoretically.
Result and discussion
Complex [Pt(ppy)(η1-S-Spy)(PPh3)], 1a, was prepared according to our recently published procedure (Scheme 1 (path I)),23 involving treatment of complex [Pt(ppy)(PPh3)Cl], A,24 with an ethanolic solution of sodium pyridine-2-thiolate (NaC5H4NS) under Ar atmosphere.
 |
| Scheme 1 (I) The synthetic route for the formation of complexes 1; (II) identified products in the reaction of different alkyl halides with complexes 1; (III) direct route for the formation of complexes 2 by halide substitution. All reactions were carried out under inert atmosphere at room temperature. | |
The reaction of complex 1a with stoichiometric ratio or an excess of MeI at room temperature was investigated. This reaction was extremely rapid and yielded unexpected products. As represented in Scheme 1 (path II), the reaction underwent fast C–S bond reductive coupling to cleanly produce an organic compound 2-(methylthio)pyridine (pySMe), O1,25 and complex [Pt(ppy)(PPh3)I], 2a.
Complex 2a was identified from the analysis of NMR spectroscopy data and X-ray single crystal diffraction data (see structural determination). As expected, the 31P {1H} NMR spectrum of complex 2a (in CDCl3) displayed only a singlet along with platinum satellites at δ = 22.7 with a significant value of 1JPtP = 4290 Hz. This large coupling constant confirmed that the phosphine ligand is positioned trans to nitrogen atom, same as complex A.24 In the 1H NMR spectrum of complex 2a, the existence of a doublet of doublet of doublets for H2 of ppy ligand at δ = 10.55 ppm with platinum satellites (3JPtH = 30.9 Hz) and also another doublet of doublet of doublets at δ = 6.67 ppm accompanied by platinum satellites (3JPtH = 52.1 Hz) for H9 of ppy ligand, confirmed that cyclometalating ligand was chelated in this complex (refer to the Experimental section for signal assignments). The later resonance exhibited a C–H⋯π interaction, as detected in the single X-ray crystal structure of complex 2a. In addition, complex 2a could be achieved directly by a nucleophilic substitution reaction of chloride ligand in complex A by iodide (I−) ion (Scheme 1 (path III)).
Other electrophilic substrates have been used to investigate the effect of these kind of reagents on the above reaction (Scheme 1 (path II)). The obtained products were similar to the resultant products from methyl iodide species. For example, addition of ethyl iodide (EtI) or benzyl bromide (PhCH2Br) to complex 1a resulted in products 2-(ethylthio)pyridine (pySEt), O2,25 and complex 2a or 2-(benzylthio)pyridine O3,25 and complex [Pt(ppy)(PPh3)Br], 2b, respectively (Scheme 1 (path II)). Complex 2b was fully characterized by NMR spectroscopy and single X-ray crystallography studies (see structural determination, Experimental section). Furthermore, complex 2b was obtained through salt metathesis reaction of complex A with KBr (Scheme 1 (path III)).
Mechanism of C–S coupling from complexes 1
In an effort to achieve more insight into the mechanism of reaction (carbon–sulfur bond formation) that was described in Scheme 1 (path II), it was monitored at different temperatures with NMR spectroscopy. The reactants were mixed and firstly NMR spectra were recorded at 25 °C. In 31P {1H} NMR spectra (Fig. 1), immediately after addition of MeI, the signals of the starting complex 1a (δ = 22.7 ppm) quickly disappeared (Fig. 1c) and a signal corresponding to complex 2a (δ = 22.4 ppm) grew rapidly as time elapsed. Apart from resonances of complex 1a (precursor) and complex 2a (product), an interesting Pt(II)–Pt(II) binuclear intermediate species was detected, suggesting the presence of two different phosphorus ligands with high platinum coupling constants (Fig. 1). Its structure was determined as a binuclear complex [{Pt(ppy)(PPh3)}2(μ2-Spy)]I, 3-I. Therefore, two singlet signals, one at δ = 21.5 ppm (1JPtP = 4407 Hz, Pa) and another one at δ = 21.0 ppm (1JPtP = 4245 Hz, Pb) were assigned to complex 3-I and phosphine ligands in this complex perhaps were located in cis position to the nitrogen and sulfur atoms of bridging thiolate ligand, respectively. This assignment is related to high cis influence of S atom over N atom.26,27 As time passed, the signal intensities of complex 3-I decreased while the signal intensities of complex 2a increased until full conversion (ca. 6 h) was obtained as compared to the pure form of complex 2a.
 |
| Fig. 1 Reaction of complex 1a with 3 eq. of MeI as monitored by 31P {1H} NMR spectroscopy in CD2Cl2 at room temperature. (a) Pure complex 1a, (b) immediately after addition of MeI, (c) 5 min (d) 15 min, (e) 30 min, (f) 60 min, (g) 120 min, (h) 180 min, (i) 240 min and (j) 480 min. The signal assignments are shown. | |
The related 1H NMR spectra (Fig. S1†) for this reaction were also obtained simultaneously at room temperature. This observation confirmed the formation of complex 2a, and especially compound O1, as products and complex 3-I as an intermediate by observing their expected signals. As shown in Fig. S1,† new resonances with increased signals at δ = 2.58 ppm and 8.46 ppm were assigned to methyl and H6′ (CH group adjacent to nitrogen atom) of compound O1,25 respectively. When complex 1 was treated with CD3I under the same conditions (CD2Cl2, room temperature), the methyl signal disappeared in the 1H NMR spectrum (Fig. S2†).
To detect other plausible intermediates, the reaction was monitored at low temperature (223 K). In this temperature the rate of reaction became very slow and an excess of MeI was used. However, no further intermediates were observed (Fig. S3†) and similar room temperature NMR spectra were obtained. This spectroscopic observation suggested that other possible intermediates were very short lived and they were quickly transformed to the products.
Attempts to prepare resemble of complex 3-I were successful. The synthetic strategy involved treatment of complex A with AgPF6 to abstract the chloride (Cl−) ligand from the platinum center as AgCl (separated from solution by filtration). Subsequently, complex 1a was added to the resulting solution and the binuclear complex [{Pt(ppy)(PPh3)}2(μ2-Spy)]PF6, 3-PF6, was formed as an orange solid. The multinuclear NMR spectroscopy and crystal structure determination confirmed the formation of this complex. The 31P {1H} NMR spectrum (Fig. 2) of complex 3-PF6 exhibited two singlet resonance flanked by platinum satellites at similar chemical shift (with equal coupling 1JPtP constant) of complex 3-I. It is notable that the counteranion (I− or PF6−) did not have impact effect on the cationic part of complex 3 in solution and only PF6− facilitated crystallization process. As expected, the 195Pt {1H} spectrum (Fig. 2) of complex displayed two doublets at −4081 (Pta) and −4296 (Ptb) ppm with 1JPtP values of 4413 and 4256 Hz, respectively, close to the values obtained from the 31P {1H} NMR spectrum.
 |
| Fig. 2 (a) 31P and (b) 31Pt NMR spectra of complex 3-PF6 in CD2Cl2 at room temperature. | |
The synthesized complex 3-I was stable in solution for several days and it did not display any decomposition products or rearrangements to complexes 1a and 2a. Also, complex 3-I was treated with MeI and it easily was reacted and lead to complex 2a and compound O1. When complex 3-PF6 was treated with MeI it did not show any reaction after three hours. According to this data, the binuclear complex 3, perhaps, after the reaction with MeI and subsequent C–S bond coupling needs two coordinating anion like complex 3-I (one from counterion and another from MeI). But complex 3-PF6 has a non-coordination counterion (PF6−), probably, this point stops desire bond formation.
Initial coordination site of alkyl halide species in complex 1a is important, because complex 1a contains three potential sites (nitrogen, sulfur of Spy and platinum center) for initial coordination of alkyl halide reagents. Therefore, some tests were conducted to clarify the nucleophilicity of each coordination site in complex 1a against these substrates (see below). The results from these experiments were helpful in proposing a mechanistic pathway for the observed C–S bond formation reaction in this study.
The free nitrogen of thiolate ligand in complex 1a with alkyl halide reagents could potentially undergo Menschutkin reaction28 to produce N-alkylated complexes (Scheme S1,† nitrogen path), although those were not observed in NMR spectroscopy studies (Fig. S1†). The possible N-methylated complex 1-N from the reaction of complex 1a with MeI could display a low field signal owing to the coordinated methyl on nitrogen.26 Supposedly, cationic complex 1-N was not stable and could not be detected with spectroscopic methods and it was converted to complex 2a and compound O1-t (N-methyl-2-thiopyridone).29 But, 1H NMR spectra29 did not show any signal for the formation of compound O1-t and only compound O1 was observed. It is notable that the conversion of compound O1-t to compound O1 (referenced to the thione-thiol tautomer)29,30 needs harsher reaction conditions.29 To solve this challenge, complex [Pt(ppy)(SPh)(PPh3)], 1b,23 in which PhSH = thiophenol, with no pendant nitrogen atom was selected for reaction with MeI.31 This reaction produced complex 2a and its corresponding C–S coupling product i.e. thioanisole O4
32 (Scheme 1 (path II)). Therefore, this test suggested that the uncoordinated nitrogen atom does not have an important role in the formation of yielded compounds and perhaps it has some effects on the formation of complex 3-I from complex 1a. Consequently, when the test reaction was followed with NMR spectroscopy (Fig. S4 and S5†) it merely displayed transformation of complex 1b to complex 2a without any detectable intermediates in comparison with complex 1a (Fig. 1). All above observations suggested that the nitrogen site has weak nucleophilicity.31 Comparable type of systems, each having at least one free nitrogen atom on the coordinated ligand to platinum center remain intact in reaction with alkyl halide because other sites of these complexes are stronger nucleophiles than N atom.31,33–35 Consequently, a different site in complexes 1a and 1b is involved in the initial coordination of alkyl halide reagents.
An alternative possibility for preliminary coordination of alkyl halides to complex 1a is sulfur atom which is known as S-alkylation.36–40 It is related to the metal complexes containing thiolate ligands especially dithiolate group,38–41 while these are treated with alkyl halides species (in thermal condition for most cases)38,39,42 S-alkylation is a plausible route.36,37,43 As shown in Scheme S1† (sulfur path), in the case of MeI, the presumed S-methylated complex 1-S ought to exhibit a signal with platinum satellite in relation to coordinated methyl to sulfur atom but this was not observed (Fig. S1†). Thus, another experiment was designed for unravelling this inconsistency. While complex 1a reacted with iodine or bromine reagents, it produced complexes 2a or 2b and its corresponding S–I or S–Br coupling products44 (Scheme 1 (path II)). This reaction was followed by NMR spectroscopy and an intermediate (complex 3-I) similar to the reaction of methyl iodide was detected (Fig. 1). Therefore, sulfur atom is not incipient coordination site and S-alkylation can be ruled out.43 Also, there are numerous reports on the oxidation of metal centers by alkyl substrates45–48 than sulfur oxidation (thiolate alkylation). For example, sicilia research group reported that the reaction of half-lantern compound (containing thiolate ligand) with MeI leaded to the oxidized complex and S-alkylation did not occur.48 Therefore, the proposed routes in Scheme S1† are not appropriate and platinum center is initial coordination site for alkyl halide substrates and oxidative addition path is adequate for suggested C–S coupling mechanism (Scheme 2).
 |
| Scheme 2 Proposed mechanism for the reaction of complex 1a with methyl iodide. | |
We also investigated the effect of phosphine ligand on the C–S bond formation by changing PPh3 with PPhMe2, and complex [Pt(ppy)(η1-S-Spy)(PPhMe2)], 1c, was prepared (Scheme 1 (path I)). When alkyl halide regent (MeI) was added to complex 1c in acetone-d6 at room temperature and the reaction was followed by NMR spectroscopy, rapid and complete conversion to complex [Pt(ppy)(PPhMe2)I], 2c, and O1 was observed (Scheme 1 (path II)). This reaction pathway is analogous to the reaction with complex 1a which contains PPh3, but C–S bond reductive elimination occurs at a faster rate. The increased rate of reaction is consistent with the greater nucleophilicity of complex 1c than complex 1a.33,49
Structural determination
Single-crystal X-ray diffraction investigation was carried out on complexes A, 2a, 2b and 3-PF6 to approve their molecular structures. Crystallographic data are collected in Table S1,† and selected bond distances and angles are quoted in Table S2.† The perspective drawing of these complexes are illustrated in Fig. 3 and S6.†
 |
| Fig. 3 Representations of the X-ray crystal structure of complex [{Pt(ppy)(PPh3)}2(μ2-Spy)]PF6, 3-PF6, showing all non-hydrogen atoms as 40% thermal ellipsoids. Hydrogen atoms, PF6 and a CH2Cl2 solvent molecule have been omitted for clarity. Pt1–C11 1.982(14), Pt2–C45 2.000(17), Pt1–N1 2.070(12), Pt2–N2 2.066(12), Pt1–N3 2.117(11), Pt1–P1 2.235(3), Pt2–P2 2.226(3), Pt2–S1 2.418(4), C11–Pt1–N1 81.0(5), C45–Pt2–N2 80.4(6), C11–Pt1–P1 95.0(4), C45–Pt2–P2 97.9(5), C11–Pt1–N3 172.2(4), C45–Pt2–S1 160.4(5), N1–Pt1–P1 173.5(3), N2–Pt2–P2 165.7(4), N1–Pt1–N3 91.2(4), N2–Pt2–S1 94.6(4), N3–Pt1–P1 92.8(3), P2–Pt2–S1 91.41(13). | |
All complexes A, 2a, 2b and 3-PF6 reveal a distorted square planar coordination environment around Pt(II) center. This distortion is due to small bite angle [81.0(5)–78.7(13)°] of the ppy cyclometalated ligand. This narrow bite angle with bond distances for cyclometalated chelate (i.e. Pt–Cppy and Pt–Nppy) are comparable to those observed in other five membered rings of 2-phenylpyridinate platinum(II) complexes.50–53 In each structure, the metalated carbon atom of the ppy ligand accommodates the coordinated phosphine ligand in cis arrangement of Pt–Cppy bond, as a result of high trans influence of carbon atom.54 Therefore, PPh3 ligands are located trans to nitrogen atom of ppy ligand and bond distances of the Pt–P are notably shorter than platinum–phosphorous bond lengths, positioned trans to carbon atom of cyclometalated ligand.50 Furthermore, halogen ligand completes the coordination sphere of platinum atom in complexes A, 2a and 2b or nitrogen and sulfur atoms of bridging Spy ligand in complex 3-PF6. Also, in the binuclear complex 3-PF6 the pyridine-2-thiolate group acts as a bridging ligand and binds two Pt centers through S^N coordination. The nitrogen and sulfur atoms of Spy ligand in this complex occupies a position trans to the Pt–Cσ bond of cyclometalated ligands.
Additional examination on the molecular packing of the crystal structures displayed the presence of inter- and intramolecular interactions (see below) such as C–H⋯π interaction, π⋯π stacking, without any notable Pt⋯Pt contacts. All molecular structures reveal a moderate short C–H⋯π intramolecular interaction (Fig. S7†) between hydrogen atom adjacent to coordinated carbon atom of cyclometalated ligand and the phenyl ring of the triphenylphosphine ligand (C–H⋯π; C2⋯Cph (PPh3) = 3.320–3.416 Å (A), 3.237–3.401 Å (2a), 3.291–3.365 Å (2b), for 3-PF6, C–H⋯π; C11 or C45⋯Cph (PPh3) = 3.244–3.493 Å, C–H⋯π; C1⋯CSpy = 3.462 Å).
The crystal structure of complexes A, 2a and 2b show that the halogen ligand is involved in a substantial intramolecular interaction with C–H group adjacent to coordinated nitrogen atom of ppy ligand (Fig. S8†). This C–H⋯X interaction causes deshielding of the proton involved in hydrogen bonding55,56 (∼1.3 to 1.9 ppm) in 1H NMR (Fig. S9†) relative to the hydrogen in free ppy ligand.57 The shift values has also increased according to atomic radii of halogen ligands in the order of Cl < Br < I.58
In complex 3-PF6 the ring of the pyridine-2-thiolate ligand exhibits weak intramolecular π⋯π interaction with one of the phenyl groups of the PPh3 ligand bound to the Pt1 with distance between the centroids of 3.888 Å (Fig. S10†). Complexes A and 2b are stacked through short intermolecular π⋯π interactions with an inter-planar distance of 3.335 Å and 3.329 Å, respectively. These interactions lead to formation of a dimer in a head to tail style, which are analogous to other platinum complexes having ppy fragments51 (Fig. S11†). Furthermore, in crystal network of complex 2b, each Br atom is bound to one hydrogen atom of the phosphine ligand from neighboring unit. In this molecular structure, each Pt(ppy) moiety is arranged through π⋯π interactions with another Pt(ppy) fragment in nearby chain and thus giving rise to a 2D network (Fig. S12†). Therefore, this complex indicates that the C–H groups (see above discussion) are involved in the formation of intra- and intermolecular hydrogen bonds59 through the coordinated Br ligand with bond distance of 2.583 Å and 3.045 Å, respectively (Fig. S8c and S12†). Complex 2a displays the comparatively long interlayer distances between nearby ppy ligands that are longer than 3.8 Å,60 demonstrating the lack of π⋯π interactions in this complex.
DFT investigation
In order to verify this approach (according to the experimental data), we calculated the energy and reaction profiles for the oxidative addition of Me–I to complex 1a and carbon–sulfur bond formation from the suggested intermediate structures (Scheme 2 and Fig. 4). It is also worth mentioning that the pyridyl group and its peripheral nitrogen donor atom could be orientated in four different fashions (Fig. S13†). In these modes, the py ring and N atom are located at a plane above or below of the molecule and near or far from the platinum center, respectively. However, theoretical calculations suggest all possible orientations have small variation in energy (∼2 kJ mol−1), perhaps due to free rotation around S–Cpy bond.61 But, the form a (Fig. S13†) is preferred over other structures since it is in good agreement with further theoretical calculations (see below).
 |
| Fig. 4 Calculated structures and relative energies for probable intermediates and transition states arising from the reaction of complex 1 and MeI in acetone solution. | |
The DFT calculations suggest that the electron rich platinum center of complex 1a (as a nucleophile) attacks on carbon atom of MeI (as an electrophile) to form transition state TS1 and energy barrier for this state is +52.9 KJ mol−1. This transition state illustrates a linear arrangement for I⋯CMeI⋯Pt with the bond angles close to 90° and 180° between H–CMeI⋯Pt (90–93°) and I⋯CMeI⋯Pt (178.5°), respectively. The most important modifications in bond distances are detected for I⋯CMeI and Pt⋯CMeI in transition state TS1. These calculations suggested that the bond distance increases for I⋯CMeI (2.73 Å) in this transition state as compared to MeI (2.20 Å), while Pt⋯CMeI (2.49 Å) bond length is reduced.62,63 This data supports the hypothesis that the formation of transition state TS1 is followed by concurrent bond breaking of methyl-iodide and a new bond formation between platinum and methyl, then the cationic 5-coordinate intermediate IM1 is obtained. This intermediate is equilibrated with the comparable intermediate IM2 with phosphine ligand and Me group being in trans disposition at platinum center (Fig. 4). Both intermediates IM1 and IM2 could potentially convert to the 6-coordinate compounds P1 and P2, respectively (Scheme 2). The calculation indicates that the energy barriers for these compounds are high (+42.9 and +50.3 kJ mol−1 for compounds P1 and P2, respectively) and subsequent C–S bond coupling is not favorable. It is recommended that this concerted elimination from the five-coordinate d6 metal complexes is more straightforward than direct elimination from the saturated six-coordinate metal complexes.11,64,65 Therefore, the carbon–sulfur bond formation from intermediate IM1 (selected based on its good agreement with theoretical calculations) proceeds via transition state TS2 as a concerted reductive elimination66 (Scheme 2). In TS2, the bond distances between platinum with methyl (2.42 Å) or thiolate (2.47 Å) groups are lengthened and the bond angle of CMe–Pt–SSpy (61.3°) is decreased and it suggested that this elimination is appropriate. Another possibility for the reductive elimination of C–S bond, probably, is the microscopic reverse of the popular SN2 mechanism.66,67 In this mechanism, pervious mechanistic evidences provided, the nucleophilic groups like NR−, OR−, I− dissociated from central metal, and then perform SN2 attack upon an alkyl ligand (Me group) of the cationic intermediates.12,65,68–70 Therefore, by addition of excess of nucleophilic groups to the reaction mixture the rate of reaction should be increased. By this strategy, when Spy− was added to the reaction mixture, the desire C–S bond coupling reaction was stopped. Due to Spy− was reacted with MeI and compound O1 and KI salt were formed without accelerating rate of reductive elimination of C–S from intermediate IM1. Consequently, the possibility of SN2 mechanism is not proper for this C–S bond formation. As compound O1 dissociates from this transition state (TS2), the new cationic T-shaped intermediate IM3
11,65,69,71 is produced (Scheme 2). Intermediate IM3 can then react by two routes: (i) migration of iodide to the platinum center and formation of complex 2a as a very stable complex (−71.3 KJ), or (ii) reaction with complex 1a to yield a cationic complex 3-I. The latter pathway is slower (deduced from NMR, Fig. 1 and S1†) and more favorable (−51.8 kJ, Fig. 4). Subsequently, complex 3-I in the presence of MeI acts as a nucleophile (perhaps through a classical SN2 mechanism)48 and leads to the formation of an intermediate IM4
48 (Pt(II)–Pt(IV)–Me). Finally, intermediate IM4 through a reductive elimination process (Me–Spy, compound O1) generates the stable final products (Fig. 4 and Scheme 2).
Conclusion
The platinum(II) complex 1a reacted readily with electrophilic alkyl sources at ambient temperature and through carbon–sulfur reductive coupling reaction produced corresponding halide complexes 2a or 2b and alkyl sulfides. This reaction was monitored by NMR spectroscopy at different temperatures and the results revealed (Scheme 2) an interesting binuclear complex 3-I was formed during this process. Also, we proposed that the platinum center is a stronger nucleophile than peripheral nitrogen or sulfur atom. This claim was approved by selection of complex 1c, which lacks a free nitrogen atom, and it underwent a similar C–S bond formation reaction. Moreover, in this experiment complex 1a did not display any Pt(IV) compound(s) and perhaps such intermediate(s) is(are) energetically unfavorable as supported by DFT calculations. Therefore, platinum–thiolate complexes of these nature served as an appropriate platform for oxidative addition reaction that impressively enhanced the formation of alkyl sulfides.
Experimental section
General procedures and materials
All NMR spectra (1H, 31P {1H} and 195P {1H}) were recorded on a Brucker Avance DPX 400 MHz instrument. References were TMS or the residual peak of the solvent, i.e. CD2Cl2, CDCl3 and acetone-d6 (1H), 85% H3PO4 (31P), and aqueous Na2PtCl6 (195Pt). The chemical shifts (δ) being reported as ppm and coupling constants (J) expressed in Hz. The microanalyses were performed using a vario EL CHNS elemental analyzer. Electrospray ion mass spectrum (ESI-MS) was recorded by a HP-5989B spectrometer using methanol–water as the mobile phase. All solvents were purified and dried according to standard procedures.72 2-phenylpyridine (ppyH), pyridine-2-thiol (pySH), thiophenol (PhSH), silver hexafluorophosphate (AgPF6), triphenylphosphine (PPh3) and dimethylphenylphosphine (PPhMe2) were purchased from Aldrich or Acros. Complexes [Pt(ppy)(DMSO)(Cl)],53 [Pt(ppy)(PPh3)Cl], A,24 [Pt(ppy)(η1-S-Spy)(PPh3)], 1a,23 [Pt(ppy)(SPh)(PPh3)], 1b,23 were prepared as reported in literature. The NMR labeling for all ligands are shown in Scheme 3 for clarifying the chemical shift assignments.
 |
| Scheme 3 Representative ligands with position labeling. | |
[Pt(ppy)(PPhMe2)Cl], B
To a solution of [Pt(ppy)(DMSO)Cl] (100 mg, 0.22 mmol) in acetone (10 mL) was added PPhMe2 (30.8 μL, 0.22 mmol). The mixture was stirred at room temperature for 1 h. Then, the solvent was removed under reduced pressure and the residue was triturated with n-hexane (3 × 2 mL) and the resultant yellow solid was dried under vacuum. Yield: 94 mg, 83%. Elem. anal. calcd for C19H19ClNPPt (522.87): C, 43.64; H, 3.66; N, 2.68; found: C, 43.91; H, 3.74; N, 2.57. 1H NMR (400 MHz, acetone-d6, 20 °C, δ): 9.80 (d, 3JPtH = 24.1, 3JHH = 5.8, 1H, H2), 8.11–8.04 (m, 4H, H4, H5 and H0 of PPhMe2), 7.69 (dd, 3JHH = 7.8, 4JHH = 1.3, 1H, H6), 7.53–7.48 (m, 4H, H3, Hp and Hm of PPhMe2), 7.00 (td, 3JHH = 7.8, 4JHH = 1.2, 1H, H7), 6.87 (d, 3JPtH = 56.1, 3JHH = 7.8, 1H, H9), 6.75 (td, 3JHH = 7.8, 4JHH = 1.2, 1H, H8), 2.00 (d, 2JPH = 10.4, 3JPtH = 39.7 Hz, 6H, Me group of PPhMe2). 31P {1H} NMR (162 MHz, acetone-d6, 20 °C, δ): −8.1 (s, 1JPtP = 4131, 1P).
[Pt(ppy)(η1-S-Spy)(PPhMe2)], 1c
Complex B (100 mg, 0.19 mmol) was added to an ethanolic solution of sodium pyridine-2-thiolate ligand (NaC5H4NS) under inert atmospheric condition [note: NaC5H4NS was prepared by dissolving sodium (5.8 mg, 0.25 mmol) in 10 mL of absolute ethanol and subsequent treatment with pyridine-2-thiol (21.3 mg, 0.19 mmol)]. The resulting orange colored mixture was allowed to react while stirring for 12 h. Then, solvent was removed under reduced pressure the residue was extracted with CH2Cl2 (20 mL). The obtained orange solution was filtered through celite and the filtrate was concentrated to a small volume (1 mL). Finally, n-hexane (5 mL) was added to precipitate complex 1c as an orange solid. Yield: 71 mg, 62%. Elem. anal. calcd for C24H23N2PPtS (597.57): C, 48.24; H, 3.88; N, 4.69; found: C, 47.96; H, 3.74; N, 4.76. 1H NMR (400 MHz, acetone-d6, 20 °C, δ): 10.05 (m, 3JPtH = 26.2, 1H, H2), 8.15–8.03 (m, 5H, H4, H5, H6′ and H0 of PPhMe2), 7.90 (dd, 3JHH = 7.8, 4JHH = 1.0, 1H, H6), 7.57 (d, 3JHH = 8.0, 1H, H3′), 7.48–7.36 (m, 4H, H3, Hp and Hm of PPhMe2), 7.19 (td, 3JHH = 7.8, 4JHH = 1.9, 1H, H7), 7.02 (t, 3JHH = 7.1, 1H, H4′), 6.98 (d, 3JPtH = 46.1, 3JHH = 7.8, 1H, H9), 6.80 (td, 3JHH = 7.8, 4JHH = 1.7, 1H, H8), 6.71 (dd, 3JHH = 5.1, 4JHH = 0.9, 1H, H5′), 2.01 (d, 2JPH = 10.9, 3JPtH = 40.1 Hz, 6H, Me group of PPhMe2). 31P {1H} NMR (162 MHz, acetone-d6, 20 °C, δ): −8.5 (s, 1JPtP = 4178, 1P).
[Pt(ppy)(PPh3)I], 2a
Method A. To a yellow suspension of complex 1a (100 mg, 0.14 mmol) in acetone (15 mL) was added an excess of MeI (174.3 μL, 2.8 mmol, 20 equiv.) at room temperature. The color of the mixture turned red instantly and then the mixture was stirred for 2 h. Then, the solvent was removed under reduced pressure and the residue was triturated with n-hexane (3 × 2 mL). The resulting green solid was dried under vacuum. Yield: 82 mg, 78%; mp 225 °C. Elem. anal. calcd for C29H23INPPt (738.03): C, 47.15; H, 3.14; N, 1.90; found: C, 47.81; H, 3.37; N, 1.97. 1H NMR (400 MHz, CDCl3, 20 °C, δ): 10.55 (ddd, 3JPtH = 30.9, 3JHH = 5.8, 4JHH = 0.9, 4JPH = 4.3, 1H, H2), 7.87–7.77 (m, 8H, H4, H5, H0 of PPh3), 7.50 (dd, 3JHH = 7.9, 4JHH = 1.1, 1H, H6), 7.43–7.34 (m, 9H, Hp and Hm of PPh3), 7.20 (tdd, 3JHH = 7.3, 4JHH = 1.3, 4JPH = 1.4, 1H, H3), 6.99 (td, 3JHH = 7.9, 4JHH = 0.9, 1H, H7), 6.67 (ddd, 3JPtH = 52.1, 3JHH = 7.3, 4JHH = 0.8, 4JPH = 4.0, 1H, H9), 6.52 (td, 3JHH = 8.0, 4JHH = 1.3, 1H, H8). 31P {1H} NMR (162 MHz, CDCl3, 20 °C, δ): 22.7 (s, 1JPtP = 4290, 1P). 1H NMR (400 MHz, acetone-d6, 20 °C, δ): 10.54 (ddd, 3JPtH = 31.1, 3JHH = 5.7, 4JHH = 1.0, 4JPH = 4.3, 1H, H2), 8.13–8.11 (m, 2H, H4, H5), 7.86–7.81 (m, 6H, H0 of PPh3), 7.70 (dd, 3JHH = 7.9, 4JHH = 1.4, 1H, H6), 7.50–7.40 (m, 10H, H3, Hp and Hm of PPh3), 6.99 (td, 3JHH = 7.9, 4JHH = 1.1, 1H, H7), 6.69 (ddd, 3JPtH = 52.7, 3JHH = 7.4, 4JHH = 0.8, 4JPH = 4.0, 1H, H9), 6.49 (td, 3JHH = 79, 4JHH = 1.5, 1H, H8). 31P {1H} NMR (162 MHz, acetone-d6, 20 °C, δ): 22.8 (s, 1JPtP = 4296, 1P). 1H NMR (400 MHz, CD2Cl2, 20 °C, δ): 10.49 (m, 3JPtH = not resolved, 1H, H2), 7.92–7.78 (m, 8H, H4, H5, H0 of PPh3), 7.53 (dd, 3JHH = 7.8, 4JHH = 1.1, 1H, H6), 7.44–7.38 (m, 9H, Hp and Hm of PPh3), 7.25 (t, 3JHH = 7.4, 1H, H3), 7.00 (t, 3JHH = 7.9, 1H, H7), 6.65 (dd, 3JPtH = 52.4, 3JHH = 7.5, 4JPH = 4.1, 1H, H9), 6.52 (t, 3JHH = 7.9, 1H, H8). 31P {1H} NMR (162 MHz, CD2Cl2, 20 °C, δ): 22.4 (s, 1JPtP = 4288, 1P).
Method B. To a yellow suspension of complex 1a (100 mg, 0.14 mmol) in acetone (15 mL) was added a solution of I2 (175.8 g, 0.70 mmol, 5 equiv.) in acetone (5 mL) at room temperature. A fast color change to red was observed. The resulting solution was stirred for 2 h. Then, the solvent was reduced under vacuum to a small volume (1 mL) and n-hexane (5 mL) was added to precipitate complex 2a as a green powder (73 mg, 71%).
Method C. To a green suspension of complex A (100 mg, 0.15 mmol) in acetone/CH2Cl2 (10
:
5 mL) was added an excess of KI (74.7 g, 0.45 mmol, 3 equiv.) in acetone (2 mL) at room temperature. The resulting mixture was allowed to react while stirring for 24 h. Then, the solvent was removed under reduced pressure and the residue was extracted with CH2Cl2 (20 mL). The obtained green solution was filtered through Celite and the filtrate was concentrated to a small volume (1 mL). Finally, n-hexane (5 mL) was added to yield complex 2a as a green solid which was dried under vacuum (48 mg, 43%).
[Pt(ppy)(PPh3)Br], 2b
Method A. To a yellow suspension of complex 1a (100 mg, 0.14 mmol) in acetone (15 mL) was added benzyl bromide (166.3 μL, 1.4 mmol, 10 equiv.) at room temperature and an orange solution was obtained quickly. The reaction mixture was stirred for 2 h. Then, the solvent was evaporated under vacuum and the residue was washed with n-hexane (3 × 2 mL), resulting in a green solid which was dried under vacuum. Yield: 86 mg, 89%; mp 221 °C. Elem. anal. calcd for C29H23BrNPPt (690.04): C, 50.43; H, 3.36; N, 2.03; found: C, 51.01; H, 3.59; N, 2.26. 1H NMR (400 MHz, CDCl3, 20 °C, δ): 10.18 (ddd, 3JPtH = 30.8, 3JHH = 5.7, 4JHH = 1.0, 4JPH = 4.1, 1H, H2), 7.89–7.77 (m, 8H, H4, H5, Ho of PPh3), 7.51 (dd, 3JHH = 7.8, 4JHH = 1.2, 1H, H6), 7.44–7.35 (m, 9H, Hp and Hm of PPh3), 7.27 (m, 1H, H3, this signal has overlapping with CHCl3 NMR solvent), 6.97 (td, 3JHH = 8.0, 4JHH = 0.9, 1H, H7), 6.67 (ddd, 3JPtH = 52.4, 3JHH = 7.4, 4JHH = 0.8, 4JPH = 3.9, 1H, H9), 6.52 (td, 3JHH = 8.0, 4JHH = 1.4, 1H, H8). 31P {1H} NMR (162 MHz, CDCl3, 20 °C, δ): 22.7 (s, 1JPtP = 4328, 1P).
Method B. To a yellow suspension of complex 1a (100 mg, 0.14 mmol) in acetone (15 mL) was added a solution of Br2 (36 μlit, 0.70 mmol, 5 equiv.) in acetone (5 mL) at room temperature. A fast color change to red was observed. The resulting solution was stirred for 2 h. Then, the solvent was reduced under vacuum to a small volume (1 mL) and n-hexane (5 mL) was added to precipitate complex 2b as a green powder (61 mg, 64%).
Method C. To a green suspension of complex A (100 mg, 0.15 mmol) in acetone/CH2Cl2 (10
:
5 mL) was added an excess of KBr (53.6 g, 0.45 mmol, 3 equiv.) in acetone (2 mL) at room temperature. The resulting mixture was allowed to react while stirring for 30 h. Then, the solvent was removed under reduced pressure and the solid residue was extracted with CH2Cl2 (20 mL). The obtained green solution was filtered through celite and the filtrate was concentrated to a small volume (1 mL). Finally, n-hexane (5 mL) was added to yield complex 2b as a green solid (39 mg, 38%).
[Pt(ppy)(PPhMe2)I], 2c
To an orange solution of complex 1c (100 mg, 0.17 mmol) in acetone (15 mL) was added an excess of MeI (208.3 μL, 3.4 mmol, 20 equiv.) at room temperature. The solution color turned red instantly and the mixture was stirred for 1 h. Then, the solvent was removed under reduced pressure and the resulting residue was triturated with n-hexane (3 × 2 mL). Finally, the resulting green solid was dried under vacuum. Yield: 64 mg, 62%. Elem. anal. calcd for C19H19INPPt (614.32): C, 37.15; H, 3.12; N, 2.28; found: C, 37.31; H, 3.24; N, 2.21. 1H NMR (400 MHz, acetone-d6, 20 °C, δ): 10.41 (m, 3JPtH = 32.3, 1H, H2), 8.10–8.05 (m, 4H, H4, H5 and H0 of PPhMe2), 7.70 (dd, 3JHH = 7.7, 4JHH = 1.4, 1H, H6), 7.51–7.43 (m, 4H, H3, Hp and Hm of PPhMe2), 7.05 (td, 3JHH = 7.7, 4JHH = 1.1, 1H, H7), 6.84 (ddd, 3JPtH = 55.3, 3JHH = 7.6, 4JHH = 1.0, 4JPH = 3.8, 1H, H9), 6.77 (td, 3JHH = 7.7, 4JHH = 1.3, 1H, H8), 2.21 (d, 2JPH = 10.8, 3JPtH = 39.9 Hz, 6H, Me group of PPhMe2). 31P {1H} NMR (162 MHz, acetone-d6, 20 °C, δ): −9.0 (s, 1JPtP = 4108, 1P).
[{Pt(ppy)(PPh3)}2(μ2-Spy)]PF6, 3-PF6
To a solution of complex A (100 mg, 0.15 mmol) in CH2Cl2 (15 mL) was added AgPF6 (51 mg, 0.20 mmol). The reaction mixture was stirred for 5 h in dark at room temperature, and then filtered through celite to remove AgCl. The green filtrate was treated with complex 1a (108.3 mg, 0.15 mmol) in CH2Cl2 (5 mL). The resulting orange solution was stirred for 2 h and then the solvent was removed under reduced pressure and concentrated to a small volume (∼1 mL) and n-hexane (3 mL) was added to obtain complex 3-PF6. Yield: 157 mg, 71%; mp 217 °C. MS ESI(+): m/z 1332.27 [M − PF6]+. Elem. anal. calcd for C63H50F6N3P3Pt2S (1477.22): C, 51.18; H, 3.41; N, 2.84; found: C, 51.47; H, 3.52; N, 2.91. 1H NMR (400 MHz, CDCl3, 20 °C, δ): 8.97 (m, 3JPtH = 32.7, 2H, H2), 7.98–6.31 (overlapping multiplets, 48H). 31P {1H} NMR (162 MHz, CD2Cl2, 20 °C, δ): 21.5 (s, 1JPtP = 4407, 1P, Pa cis with nitrogen atom of Spy ligand), 21.0 (s, 1JPtP = 4245, 1P, Pb cis with sulfur atom of Spy ligand), −144.4 (septet, 1JPF = 712, 1P, P of PF6). 195Pt {1H} NMR (85.7 MHz, CD2Cl2, 20 °C, δ): −4081 (d, 1JPtP = 4413, 1 Pt, Pta bound to the nitrogen atom of Spy ligand), −4296 (d, 1JPtP = 4256, 1 Pt, Ptb bound to the sulfur atom of Spy ligand).
Monitoring the reaction of complexes 1a–c with electrophilic reagents (MeI, CD3I, Br2 and I2) by 1H and 31P {1H} NMR spectroscopy
To a solution of complexes 1a–c (0.014 mmol) in CD2Cl2 (0.75 mL) in an NMR tube was added the appropriate electrophilic reagent at 298 K (0.020 mmol) or at 223 K (0.42 mmol). The tube was then placed in the probe of the NMR spectrometer, and NMR spectra were obtained at appropriate time intervals.
Crystal structure determination and refinement
The X-ray diffraction measurements were carried out on STOE IPDS-2/2T diffractometer with graphite-monochromated Mo Kα radiation. All single crystals were mounted on a glass fiber and used for data collection. Cell constants and an orientation matrix for data collection were obtained by least-square refinement of the diffraction data from 4977, 4763, 4350 and 10
828 for A, 2a, 2b and 3-PF6, respectively. Diffraction data were collected in a series of ω scans in 1° oscillations and integrated using the Stoe X-AREA73 software package. A numerical absorption correction was applied using X-RED74 and X-SHAAPE75 software. The data were corrected for Lorentz and polarizing effects. The structures were solved by direct methods76 and subsequent difference Fourier maps and then refined on F2 by a full-matrix least-squares procedure using anisotropic displacement parameters.77 Atomic factors are from the International Tables for X-ray Crystallography.78 All non-hydrogen atoms were refined with anisotropic displacement parameters. Hydrogen atoms were placed in ideal positions and refined as riding atoms with relative isotropic displacement parameters. A view of the structure is depicted in Fig. 3 (top) and S6.† All refinements were performed using the X-STEP32, SHELXL-2014 and WinGX-2013.3 programs.79–86 Also, the suitable crystals were obtained from CHCl3/n-hexane solution (complexes A, 2a and 2b) or CH2Cl2/n-hexane solution (complex 3-PF6) at room temperature.
Theoretical methods
Gaussian 09 was used87 to fully optimize all the structures at the DFT/B3LYP level of density functional theory. The solvation energies were calculated by CPCM model in acetone. The effective core potential of Hay and Wadt with a double-x valence basis set (LANL2DZ) was chosen to describe Pt and I.88 The 6-31G(d) basis set was used for other atoms.
Acknowledgements
This work was supported by the Institute for Advanced Studies in Basic Sciences (IASBS) Research Council. Technical support of the Chemistry Computational Center at Shahid Beheshti University is gratefully acknowledged. Thanks are also due to Dr F. Niroomand Hosseini, Islamic Azad University (Shiraz Branch), for helpful assistance and discussions on DFT calculations, Dr A. Neshat, IASBS, for valuable suggestions, Mr A. Biglari, the operator of Bruker NMR instrument at IASBS, for recording the NMR spectra and Prof. Elena Lalinde, Universidad de La Rioja, for ESI-MS measurements. This paper is dedicated to Professor Mehdi Rashidi on the occasion of his 65th birthday.
References
- M. G. Haghighi, S. M. Nabavizadeh, M. Rashidi and M. Kubicki, Dalton Trans., 2013, 42, 13369–13380 RSC.
- D. M. Crumpton-Bregel and K. I. Goldberg, J. Am. Chem. Soc., 2003, 125, 9442–9456 CrossRef CAS PubMed.
- H. R. Shahsavari, M. Rashidi, S. M. Nabavizadeh, S. Habibzadeh and F. W. Heinemann, Eur. J. Inorg. Chem., 2009, 3814–3820 CrossRef CAS.
- Á. Molnár, Chem. Rev., 2011, 111, 2251–2320 CrossRef PubMed.
- Á. Molnár, in Palladium-Catalyzed Coupling Reactions: Practical Aspects and Future Developments, Wiley-VCH Verlag GmbH & Co. KGaA, 2013 Search PubMed.
- J. M. Brown and N. A. Cooley, Chem. Rev., 1988, 88, 1031–1046 CrossRef CAS.
- P. G. Gildner and T. J. Colacot, Organometallics, 2015, 34, 5497–5508 CrossRef CAS.
- J. F. Hartwig, Inorg. Chem., 2007, 46, 1936–1947 CrossRef CAS PubMed.
- J. F. Hartwig, Nature, 2008, 455, 314–322 CrossRef CAS PubMed.
- J. Racowski and M. Sanford, in Higher Oxidation State Organopalladium and Platinum Chemistry, ed. A. J. Canty, Springer, Berlin, Heidelberg, 2011, vol. 35, pp. 61–84 Search PubMed.
- K. A. Grice, M. L. Scheuermann and K. I. Goldberg, in Higher Oxidation State Organopalladium and Platinum Chemistry, ed. A. J. Canty, Springer, 2011, vol. 35, pp. 1–38 Search PubMed.
- A. V. Pawlikowski, A. D. Getty and K. I. Goldberg, J. Am. Chem. Soc., 2007, 129, 10382–10393 CrossRef CAS PubMed.
- A. R. Dick, J. W. Kampf and M. S. Sanford, J. Am. Chem. Soc., 2005, 127, 12790–12791 CrossRef CAS PubMed.
- T. Kondo and T.-a. Mitsudo, Chem. Rev., 2000, 100, 3205–3220 CrossRef CAS PubMed.
- C. Shen, P. Zhang, Q. Sun, S. Bai, T. S. A. Hor and X. Liu, Chem. Soc. Rev., 2015, 44, 291–314 RSC.
- I. P. Beletskaya and V. P. Ananikov, Chem. Rev., 2011, 111, 1596–1636 CrossRef CAS PubMed.
- S. V. Ley and A. W. Thomas, Angew. Chem., Int. Ed., 2003, 42, 5400–5449 CrossRef CAS PubMed.
- J. Zhang, C. M. Medley, J. A. Krause and H. Guan, Organometallics, 2010, 29, 6393–6401 CrossRef CAS.
- E. Traversa, J. L. Templeton, H. Y. Cheng, M. M. Beromi, P. S. White and N. M. West, Organometallics, 2013, 32, 1938–1950 CrossRef CAS.
- M. Mellah, A. Voituriez and E. Schulz, Chem. Rev., 2007, 107, 5133–5209 CrossRef CAS PubMed.
- A. R. Murphy and J. M. J. Fréchet, Chem. Rev., 2007, 107, 1066–1096 CrossRef CAS PubMed.
- G. Liu, J. R. Huth, E. T. Olejniczak, R. Mendoza, P. DeVries, S. Leitza, E. B. Reilly, G. F. Okasinski, S. W. Fesik and T. W. von Geldern, J. Med. Chem., 2001, 44, 1202–1210 CrossRef CAS PubMed.
- M. Niazi, H. R. Shahsavari, M. G. Haghighi, M. R. Halvagar, S. Hatami and B. Notash, RSC Adv., 2016, 6, 76463–76472 RSC.
- H. Samouei, M. Rashidi and F. W. Heinemann, J. Iran. Chem. Soc., 2014, 11, 1207–1216 CrossRef CAS.
- N. Furukawa, F. Takahashi, T. Kawai, K. Kishimoto, S. Ogawa and S. Oae, Phosphorus, Sulfur Silicon Relat. Elem., 1983, 16, 167–180 CrossRef CAS.
- S. Fuertes, A. J. Chueca and V. Sicilia, Inorg. Chem., 2015, 54, 9885–9895 CrossRef CAS PubMed.
- L. Rigamonti, C. Manassero, M. Rusconi, M. Manassero and A. Pasini, Dalton Trans., 2009, 1206–1213 RSC.
- J. C. Schug, J. W. Viers and J. I. Seeman, J. Org. Chem., 1983, 48, 4892–4899 CrossRef CAS.
- P. Beak and J. T. Lee, J. Org. Chem., 1969, 34, 2125–2128 CrossRef CAS.
- M. D. Aseman, S. M. Nabavizadeh, H. R. Shahsavari and M. Rashidi, RSC Adv., 2015, 5, 22692–22702 RSC.
- M. S. McCready and R. J. Puddephatt, Organometallics, 2015, 34, 2261–2270 CrossRef CAS.
- T. Schaefer and J. D. Baleja, J. Magn. Reson., 1984, 60, 131–135 CAS.
- R. B. Aghakhanpour, S. M. Nabavizadeh, L. Mohammadi, S. Amini Jahromi and M. Rashidi, J. Organomet. Chem., 2015, 781, 47–52 CrossRef CAS.
- L. Maidich, A. Zucca, G. J. Clarkson and J. P. Rourke, Organometallics, 2013, 32, 3371–3375 CrossRef CAS.
- 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.
- L. F. Lindoy, S. E. Livingstone and T. N. Lockyer, Inorg. Chem., 1967, 6, 652–656 CrossRef CAS.
- L. F. Lindoy and S. E. Livingstone, Inorg. Chem., 1968, 7, 1149–1154 CrossRef CAS.
- J. J. Stace, P. J. Ball, V. Shingade, S. Chatterjee, A. Shiveley, W. L. Fleeman, A. J. Staniszewski, J. A. Krause and W. B. Connick, Inorg. Chim. Acta, 2016, 447, 98–104 CrossRef CAS.
- A. K. Fazlur-Rahman and J. G. Verkade, Inorg. Chem., 1992, 31, 5331–5335 CrossRef CAS.
- D. G. VanDerveer and R. Eisenberg, J. Am. Chem. Soc., 1974, 96, 4994–4996 CrossRef CAS.
- A. Vlčk Jr, Inorg. Chim. Acta, 1980, 43, 35–42 CrossRef.
- W. B. Connick and H. B. Gray, J. Am. Chem. Soc., 1997, 119, 11620–11627 CrossRef CAS.
- R. Zanella, R. Ros and M. Grazian, Inorg. Chem., 1973, 12, 2736–2738 CrossRef CAS.
- M. S. Chernov'yants, I. V. Burykin, Z. A. Starikova and N. E. Erofeev, J. Mol. Struct., 2011, 1006, 379–382 CrossRef.
- C.-H. Cheng, B. D. Spivack and R. Eisenberg, J. Am. Chem. Soc., 1977, 99, 3003–3011 CrossRef CAS.
- C.-H. Cheng and R. Eisenberg, Inorg. Chem., 1979, 18, 2438–2445 CrossRef CAS.
- J. K. Jawad, F. N. K. Al-Obaidy, J. A. Hammud and F. Al-Azab, J. Organomet. Chem., 2000, 599, 166–169 CrossRef CAS.
- V. Sicilia, M. Baya, P. Borja and A. Martín, Inorg. Chem., 2015, 54, 7316–7324 CrossRef CAS PubMed.
- 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.
- 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.
- J. R. Berenguer, E. Lalinde, A. Martín, M. T. Moreno, S. Ruiz, S. Sánchez and H. R. Shahsavari, Inorg. Chem., 2014, 53, 8770–8785 CrossRef CAS PubMed.
- Á. Díez, E. Lalinde and M. T. Moreno, Coord. Chem. Rev., 2011, 255, 2426–2447 CrossRef.
- A. Esmaeilbeig, H. Samouei, S. Abedanzadeh and Z. Amirghofran, J. Organomet. Chem., 2011, 696, 3135–3142 CrossRef CAS.
- A. Díez, J. Forniés, A. García, E. Lalinde and M. T. Moreno, Inorg. Chem., 2005, 44, 2443–2453 CrossRef PubMed.
- A. Zucca, G. L. Petretto, S. Stoccoro, M. A. Cinellu, M. Manassero, C. Manassero and G. Minghetti, Organometallics, 2009, 28, 2150–2159 CrossRef CAS.
- P. K. Byers and A. J. Canty, Organometallics, 1990, 9, 210–220 CrossRef CAS.
- M. Gholinejad and H. R. Shahsavari, Inorg. Chim. Acta, 2014, 421, 433–438 CrossRef CAS.
- D. Black, G. Deacon and G. Edwards, Aust. J. Chem., 1994, 47, 217–227 CrossRef CAS.
- E. S. Tabei, H. Samouei and M. Rashidi, Dalton Trans., 2011, 40, 11385–11388 RSC.
- C. Janiak, J. Chem. Soc., Dalton Trans., 2000, 3885–3896 RSC.
- M. Ferrer, D. Gómez-Bautista, A. Gutiérrez, G. Orduña-Marco, L. A. Oro, J. J. Pérez-Torrente, O. Rossell and E. Ruiz, Organometallics, 2016, 35, 336–345 CrossRef CAS.
- R. B. Aghakhanpour, M. Rashidi, F. N. Hosseini, F. Raoof and S. M. Nabavizadeh, RSC Adv., 2015, 5, 66534–66542 RSC.
- F. Raoof, M. Boostanizadeh, A. R. Esmaeilbeig, S. M. Nabavizadeh, R. B. Aghakhanpour, K. B. Ghiassi, M. M. Olmstead and A. L. Balch, RSC Adv., 2015, 5, 85111–85121 RSC.
- S. M. Nabavizadeh, F. N. Hosseini, N. Nejabat and Z. Parsa, Inorg. Chem., 2013, 52, 13480–13489 CrossRef CAS PubMed.
- K. I. Goldberg, J. Yan and E. M. Breitung, J. Am. Chem. Soc., 1995, 117, 6889–6896 CrossRef CAS.
- R. H. Crabtree, in The Organometallic Chemistry of the Transition Metals, John Wiley & Sons, Inc., 6th edn, 2014 Search PubMed.
- U. Fekl and K. I. Goldberg, Adv. Inorg. Chem., 2003, 54, 259 CrossRef CAS.
- K. I. Goldberg, J. Y. Yan and E. L. Winter, J. Am. Chem. Soc., 1994, 116, 1573–1574 CrossRef CAS.
- B. S. Williams, A. W. Holland and K. I. Goldberg, J. Am. Chem. Soc., 1999, 121, 252–253 CrossRef CAS.
- B. S. Williams and K. I. Goldberg, J. Am. Chem. Soc., 2001, 123, 2576–2587 CrossRef CAS PubMed.
- J. Procelewska, A. Zahl, G. Liehr, R. van Eldik, N. A. Smythe, B. S. Williams and K. I. Goldberg, Inorg. Chem., 2005, 44, 7732–7742 CrossRef CAS PubMed.
- B. S. Furniss, A. J. Hannaford, P. W. G. Smith and A. R. Tatchell, Vogel's
Textbook of Practical Organic Chemistry, Longman Scientific & Technical, 5th edn, 1989 Search PubMed.
- Stoe & Cie, X–AREA: Program for the Acquisition and Analysis of Data, Version 1.30, Stoe & Cie GmbH, Darmatadt, Germany, 2005 Search PubMed.
- Stoe & Cie, X–RED: Program for Data Reduction and Absorption Correction, Version 1.28b, Stoe & Cie GmbH, Darmatadt, Germany, 2005 Search PubMed.
- Stoe & Cie, X–SHAPE: Program for Crystal Optimization for Numerical Absorption Correction, Version 2.05, Stoe & Cie GmbH, Darmatadt, Germany, 2004 Search PubMed.
- G. M. Sheldrick, SHELX97. Program for Crystal Structure Solution, University of Göttingen, Germany, 1997 Search PubMed.
- G. M. Sheldrick, SHELX97. Program for Crystal Structure Refinement. University of Göttingen, Germany, 1997 Search PubMed.
- E. Prince, International Tables for X-ray Crystallography, Vol C, Kluwer Academic Publisher, Doordrecht, The Netherlands, 1995 Search PubMed.
- F. H. Allen, O. Johnson, G. P. Shields, B. R. Smith and M. Towler, J. Appl. Crystallogr., 2004, 37, 335–338 CrossRef CAS.
- L. J. Farrugia, J. Appl. Crystallogr., 1999, 32, 837–838 CrossRef CAS.
- C. F. Macrae, P. R. Edgington, P. McCabe, E. Pidcock, G. P. Shields, R. Taylor, M. Towler and J. van der Streek, J. Appl. Crystallogr., 2006, 39, 453–457 CrossRef CAS.
- M. N. Burnett and C. K. Johnson, ORTEP-III Report ORNL-6895, Oak Ridge National Laboratory, Tennessee, USA, 1996 Search PubMed.
- A. L. Spek, J. Appl. Crystallogr., 2003, 36, 7–13 CrossRef CAS.
- G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
- P. Coppens, L. Leiserowitz and D. Rabinovich, Acta Crystallogr., Sect. E: Struct. Rep. Online, 1965, 18, 1035–1038 CAS.
- Stoe & Cie, X-STEP32: Crystallographic Package, Version 1.07b, Stoe & Cie GmbH, Darmstadt, Germany, 2000 Search PubMed.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision B.01, 2010 Search PubMed.
- P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270–283 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: NMR spectra, crystallographic and computational details (PDF). CCDC 1457809, 1457810, 1457812 and 1496283. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra21756g |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.