Reactivity of a Tin(II) 1,3-Benzodi(thiophosphinoyl)methanediide

Treatment of 1 with 1.7 equiv of ZnEt2 in refluxing toluene afforded [{1,3-C6H4(PhPS)2CSnEt2}(μ-ZnS)]2 (3). Compound 1 also reacted with 1.7 equiv of...
0 downloads 0 Views 1MB Size
Article pubs.acs.org/Organometallics

Reactivity of a Tin(II) 1,3-Benzodi(thiophosphinoyl)methanediide Complex toward Gallium, Germanium, and Zinc Compounds Yi-Fan Yang, Rakesh Ganguly, Yongxin Li, and Cheuk-Wai So* Division of Chemistry and Biological Chemistry, School of Physical and Mathematical Sciences, Nanyang Technological University, 637371 Singapore S Supporting Information *

ABSTRACT: The reactivity of the tin(II) 1,3-benzodi(thiophosphinoyl)methanediide complex [{μ-1,3-C6H4(PhPS)2C}Sn]2 (1) toward GaCl3, GeCl4, and ZnEt2 is described. The reaction of 1 with 1.7 equiv of GeCl4 in CH2Cl2 at room temperature afforded [1,3-C6H4(PhPS)2C(GeCl3)SnCl]2 (2). Treatment of 1 with 1.7 equiv of ZnEt2 in refluxing toluene afforded [{1,3C6H4(PhPS)2CSnEt2}(μ-ZnS)]2 (3). Compound 1 also reacted with 1.7 equiv of GaCl3 in CH2Cl2 to afford [1,3C6H4(PhPS)2C(Sn)(GaCl3)] (4). Compounds 2−4 have been characterized by NMR spectroscopy and X-ray crystallography.



INTRODUCTION

Chart 1. Low-Valent Metallavinylidene Derivatives (A and D) and 1,3-Dimetallacyclobutane Derivatives (B and C)

The utilization of geminal dianions stabilized by phosphorus(V) substituents for the formation of a MC double bond has attracted much attention in the past few decades.1 Two of the well-studied examples are [(PPh2S)C2−] and [(PPh2NSiMe3)C2−].2 These geminal dianions can coordinate with a variety of main-group elements,3 transition metals,4 and lanthanides5 to form bis-phosphorus-stabilized carbene complexes of composition [(PPh2E)2CM] and [(PPh2E)2CMC(PPh2E)2]. In these complexes, the four electrons of the formal MC double bond are solely provided by the ligand, which are in contrast to the Fischer- and Schrock-type carbenes. Recently, low-valent group 14 metal derivatives [:MC(PPh2E)2]2 (M = Ge, Sn, Pb, E = NSiMe3, S) were synthesized.6 They have either a metallavinylidene structure (A, Chart 1) or a 1,3-dimetallacyclobutane structure (B and C). Moreover, the tin(II) derivative [(PPh2NSiMe3)(PPh2S)CSn:]2 (D),7 which comprises both iminophosphinoyl and thiophosphinoyl substituents, was synthesized. It has a metallavinylidene skeleton similar to that of the bisgermavinylidene [(Me3SiNPPh2)2CGe:]2 (A).6a The reactivities of low-valent group 14 metallavinylidene derivatives A8 and D9 have been well studied. They can act as a Lewis base or undergo an 1,2addition or cycloaddition toward various transition-metal complexes and organic and inorganic substrates.8,9 In contrast, the reactivity of low-valent group 14 1,3-dimetallacyclobutanes has been less studied. Only a handful of examples have been reported. For example, Leung et al. showed that the reaction of [Pb{μ-C(PPh2S)2}]2 (CII) with sulfur formed [PbS{C© 2013 American Chemical Society

(PPh2S)2}], in which the sulfur atom is inserted into the Cmethanediide−Pb bonds.6b In 2012, our group reported the synthesis of the tin(II) 1,3benzodi(thiophosphinoyl)methanediide complex [{μ-1,3C6H4(PhPS)2C}Sn]2 (1), which contains a 1,3-dimetallacyclobutane skeleton.10 Since the tin atoms in 1 contain a lone pair Received: February 19, 2013 Published: April 22, 2013 2643

dx.doi.org/10.1021/om400141j | Organometallics 2013, 32, 2643−2648

Organometallics

Article

Scheme 1. Synthesis of 2

Figure 1. Molecular structure of 2 with thermal ellipsoids at the 50% probability level. The solvent molecule and hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg): Sn(1)−S(1) = 2.6289(17), Sn(1)−S(2A) = 2.7274(19), Sn(1)−Cl(4) = 2.5192(19), C(1)− P(1) = 1.727(8), C(1)−P(2) = 1.731(8), C(1)−Ge(1) = 1.866(8), P(1)−S(1) = 2.036(3), P(2)−S(2) = 2.011(2); S(1)−Sn(1)−S(2A) = 87.12(6), S(1)−Sn(1)−Cl(4) = 83.55(6), S(2A)−Sn(1)−Cl(4) = 86.38(6), P(1)−C(1)−P(2) = 111.8(4), P(1)−C(1)−Ge(1) = 123.5(4), P(2)−C(1)− Ge(1) = 124.7(4).



RESULTS AND DISCUSSION Reactivity of 1 toward GeCl4. The reaction of 1 with 1.7 equiv of GeCl4 in CH2Cl2 at room temperature afforded [1,3C6H4(PhPS)2C(GeCl3)SnCl]2 (2; Scheme 1) in 69% yield. It is proposed that the Ge−Cl bond of GeCl4 inserts into the C−Sn bonds in 1 to form intermediate E. Subsequently, the C−Sn bond in the intermediate is cleaved and the Sn atom coordinates with the thiophosphinoyl substituent of another molecule of E to form 2. The results also illustrate the nucleophilic character of the Cmethanediide atom in 1. The similar compound [(PPh2NSiMe3)(PPh2S)C{Rh(cod)}(SnCl)], in which the Cmethanediide atom bonds to two different metals, was prepared by a 1,2-addition of [RhCl(cod)]2 with the

of electrons, we explored its reactivity toward Lewis acidic aluminum trichloride. The result is astonishing, in that the lowvalent Sn atoms do not form any adduct with AlCl3. Instead, AlCl3 inserts into the Sn−C bonds and coordinates with the Cmethanediide atoms to form [1,3-C6H4(PhPS)2C(Sn)(AlCl3)]. This illustrates that the Cmethanediide−Sn bonds in 1 are polar and the Cmethanediide atom is nucleophilic. Hence, the lone pair of electrons on the tin atoms is less favorable for bond formation. In this study, we continue our effort to investigate the reactivity of 1 with other Lewis acids. Herein, the reactivity of 1 with germanium tetrachloride, diethylzinc, and gallium trichloride is reported. 2644

dx.doi.org/10.1021/om400141j | Organometallics 2013, 32, 2643−2648

Organometallics

Article

Scheme 2. Synthesis of 3

Figure 2. Molecular structure of 3 with thermal ellipsoids at the 50% probability level. Disordered ethyl and hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg): Zn(1)−S(2) = 2.3445(5), Zn(1)−S(3) = 2.3394(5), Zn(1)−S(1) = 2.3785(5), Zn(1)−S(1A) = 2.3741(5), Sn(1)−C(1A) = 2.1070(17), Sn(1)−C(2) = 2.162(7), Sn(1)−C(4) = 2.165(2), Sn(1)−S(1) = 2.4277(4), P(1)−C(1) = 1.7091(18), P(2A)−C(1) = 1.7117(18), P(1)−S(2) = 2.0324(6), P(2A)−S(3A) = 2.0279(6); Zn(1)−S(1)−Zn(1A) = 75.185(15), S(1)−Zn(1)−S(1A) = 104.816(15), S(2)−Zn(1)−S(3) = 109.848(18), P(2)−C(1A)−P(1A) = 111.62(10), P(2)−C(1A)−Sn(1) = 123.45(9), P(1A)−C(1A)−Sn(1) = 119.29(9), C(1A)−Sn(1)−S(1) = 103.20(5).

Å).11 The Sn(1)−S(2A) bond (2.7274(19) Å) is longer than the Sn(1)−S(1) bond but is comparable to that in 1 (2.7174(6) Å).10 The C−P bonds (1.727(8) and 1.731(8) Å) are shortened and P−S bonds (2.036(3) and 2.011(2) Å) are l e n g t h e n e d in c o m p a r i s o n w i t h t h o s e in [ 1 , 3 C6H4(PhPS)2CH2] (C−P = 1.8229(14), 1.8373(14) Å; P−S = 1.9463(5), 1.9498(5) Å).10 The results indicate that there is a considerable electron delocalization throughout the backbone of the ligand. The C(1)−Ge(1) bond (1.866(8) Å) is comparable to the reported C−Ge single bonds (1.90−2.05 Å).12 Reactivity of 1 toward ZnEt2. The reaction of 1 with 1.7 equiv of ZnEt2 in refluxing toluene afforded a mixture of [{1,3C6H4(PhPS)2CSnEt2}(μ-ZnS)]2 (3; Scheme 2), SnEt4, [1,3C6H4(PhPS)(PhP)CH2], and unidentified products, which was confirmed by NMR spectroscopy and mass spectrometry. The reaction mixture was filtered, and compound 3 was isolated as a highly air- and moisture-sensitive colorless crystalline solid in 16% yield after concentration of the filtrate. Although the mechanism of the reaction is unknown as yet, on the basis of the experimental results, we propose that the reaction proceeds

stannavinylidene derivative [(PPh2NSiMe3)(PPh2S)CSn:]2 (D).9b Compound 2 was isolated as a highly air- and moisturesensitive colorless crystalline solid, which is soluble only in CH2Cl2. The 1H and 13C NMR spectra display resonances for the phenyl protons. The 13C NMR spectrum shows a triplet at δ 54.10 ppm (JP−C = 15.3 Hz) for the Cmethanediide atom in 2. The 31P NMR spectrum displays one singlet at δ 46.44 ppm for the PCP nuclei. The 119Sn NMR resonance (δ 75.30 ppm, triplet, 2JSn−P = 11.9 Hz) lies between those of [(PPh2NSiMe3)(PPh2S)C{Rh(cod)}(SnCl)] (δ 110.7 ppm)9b and [HC(PPh2S)2SnCl] (δ −129.6 ppm).1b The molecular structure of 2 is shown in Figure 1. The C(8)C(13)P(2)C(1)P(1) least-squares plane is nearly orthogonal to the P(1)S(1)S(2)P(2) least-squares plane (dihedral angle 82.0°). The tin atoms are bridged by two thiophosphinoyl substituents and adopt a distorted-trigonal pyramidal-geometry (sum of bond angles 257.1°). These geometries indicate that the tin atoms possess a high-s-character lone pair. The Sn(1)− S(1) bond (2.6289(17) Å) is comparable to that of the aryl tin(II) dithiocarboxylate [:Sn(Ar){S2CAr}] (average 2.659 2645

dx.doi.org/10.1021/om400141j | Organometallics 2013, 32, 2643−2648

Organometallics

Article

Scheme 3. Canonical Forms of 3

Scheme 4. Synthesis of 4

of bridging Zn−S bond lengths (2.281−2.425 Å).15 The C(1A)−Sn(1) bond length (2.1070(17) Å) lies between those of 1 (2.334(2) and 2.301(2) Å)10 and the stannaethene [{(Me3Si)2CH}2SnC{(BtBu)2C(SiMe3)2}] (2.025(4) Å).16 The C(1A)−Sn(1) bond is also shorter than the C(2/4)− Sn(1) bonds (2.162(7) and 2.165(2) Å). The results indicate that the C(1A)−Sn(1) bond is intermediate between a single and a double bond. Moreover, on comparison of the bond lengths of the C(PS)2 skeleton (P−C = 1.7091(18), 1.7117(18) Å; P−S = 2.0324(6), 2.0279(6) Å) with those of [1,3-C6H4(PhPS)2CH2]10 and [{HC(PPh2S)2}2Zn] (P−C = 1.715(2), 1.722(2) Å; P−S = 2.0340(8), 2.0428(8) Å),14 there is an electron delocalization along the C(PS)2 skeleton. Thus, it is suggested that compound 3 should be a resonance hybrid of the canonical formulas I and II (Scheme 3). Reactivity of 1 toward GaCl3. The reaction of 1 with 1.7 equiv of GaCl3 in CH2Cl2 afforded [1,3-C6H4(PhPS)2C(Sn)(GaCl3)] (4; Scheme 4) in 71% yield, which was confirmed by NMR spectroscopy. Compound 4 was cocrystallized with its dimer 4d (Chart 2) in a ratio of 2:1 in CH2Cl2, which was

through an oxidative addition of 1 with ZnEt2 to form the stannaethene intermediate “1,3-C6H4(PhPS)2CSnEt2” (Scheme S1; see the Supporting Information). Subsequently, ZnEt2 r e a c t s w i t h 2 e q u i v o f t h e i n t e r m e d ia t e “1 , 3 C6H4(PhPS)2CSnEt2” to form 3, SnEt4, and transient “1,3C6H4(PhPS)(PhP)C:”. The latter abstracts two protons from the solvent to form [1,3-C6H4(PhPS)(PhP)CH2]. The results are similar to the oxidation of the base-stabilized stannavinylidene derivative [(PPh2NSiMe3)(PPh2S)CSn:]2 (D) with elemental sulfur to afford the unstable 2-stannathiomethanediide intermediate “(PPh2NSiMe3)(PPh2S)CSnS”, which then decomposes to form [(PPh2NSiMe3)(PPh2S)CH2] and [{(μS)SnC(PPh2NSiMe3)(PPh2S)}3Sn(μ3-S)].9a In addition, Chivers et al. reported that the intermediate [:C(PPh2S)2] dimerizes to form [(SPh2P)2C2(PPh2)2S2], which contains a sixmembered C2P2S2 ring.13 However, such a dimerization of “1,3-C6H4(PhPS)(PhP)C:” cannot be observed in the reaction; instead, [1,3-C6H4(PhPS)(PhP)CH2] is formed. Compound 3 is sparingly soluble in CH2Cl2. The 1H NMR spectrum displays resonances for the ethyl and phenyl protons. The 13C NMR signal for the PCP atom cannot be observed. The 31P spectrum shows a singlet with satellites due to coupling to the Sn nuclei at δ 51.65 ppm (2JSn−P = 43.3 Hz). The 119Sn NMR resonance (δ 53.11 ppm, triplet, 2JSn−P = 41.1 Hz) lies between those of the base-stabilized stannavinylidene derivative [(PPh2NSiMe3)(PPh2S)CSn:]2 (D, δ 132.1 ppm)9b and 1 (δ 26.3 ppm).10 The molecular structure of 3 is shown in Figure 2. With a C6H4P2C ring as a base, compound 3 has a triple-decker structure linked by the Sn and S atoms. The Zn atoms are bridged between the thiophosphinoyl substituents and coordinated with the S(1/1A) atoms, which adopt a tetrahedral geometry. The P(1)S(2)Zn(1)S(3)P(2)P(1A)S(2A)Zn(1A)S(3A) least-squares plane is nearly orthogonal to the Zn(1)S(1)Zn(1A)S(1A) least-squares plane (dihedral angle 82.6°). The Zn(1)−S(2) (2.3445(5) Å) and Zn(1)−S(3) bonds (2.3394(5) Å) are comparable to those in [{HC(PPh2S)2}2Zn] (2.3490(6) and 2.3500(6) Å).14 The Zn(1)−S(1) and Zn(1)− S(1A) bonds (2.3785(5) and 2.3741(5) Å) are within the range

Chart 2. Dimeric Form of 4

confirmed by X-ray crystallography (see below). The results show that the low-valent Sn atoms in 1 do not form any adduct with GaCl3. Instead, the reaction appears to proceed via an insertion of GaCl3 into the Sn−C bonds of 1 to form 4. The lighter congener [1,3-C6H4(PhPS)2C(Sn)(AlCl3)] was synthesized by the reaction of 1 with 2 equiv of AlCl3.10 2646

dx.doi.org/10.1021/om400141j | Organometallics 2013, 32, 2643−2648

Organometallics

Article

spectroscopy. The 31P NMR of cocrystals in CDCl3 shows a singlet with satellites due to coupling to the Sn nuclei at δ 58.27 ppm (2JSn−P = 164.7 Hz). The 31P NMR at −95 °C still shows a singlet with satellites at δ 57.53 ppm (2JSn−P = 147.4 Hz). The results indicate that compound 4d dissociates in solution to form compound 4. Moreover, the 119Sn NMR spectrum displays a triplet at δ −157.83 ppm (2JSn−P = 176.5 Hz), which is comparable to that of [1,3-C6H4(PhPS)2C(Sn)(AlCl3)] (−140.25 ppm).10 In conclusion, compound 1 shows different reactivity toward Lewis acids. The Ge−Cl bond of GeCl4 undergoes an insertion with 1 to form [1,3-C6H4(PhPS)2C(GeCl3)SnCl]2 (2). In contrast, GaCl3 undergoes an insertion with 1 to form [1,3C6H4(PhPS)2C(Sn)(GaCl3)] (4), in which the Cmethanediide atom coordinates to GaCl3. The reaction of 1 with ZnEt2 appears to proceed through the oxidative addition of the Sn atom in 1, followed by further rearrangement with ZnEt2 to form [{1,3-C6H4(PhPS)2CSnEt2}(μ-ZnS)]2 (3).

Compounds 4 and 4d were isolated as a highly air- and moisture-sensitive colorless cocrystalline solid. In compound 4 (Figure 3), the Sn(1) atom is coordinated with the



EXPERIMENTAL SECTION

All manipulations were carried out under an inert atmosphere of dinitrogen gas by standard Schlenk techniques. CH2Cl2 was dried over and distilled over CaH2 prior to use. Toluene was dried and distilled over Na/K alloy prior to use. 1 was prepared as described in the literature.10 The 1H, 13C, 31P, 31P CPMAS, and 119Sn NMR spectra were recorded on a JEOL ECA 400 spectrometer. The chemical shifts δ are relative to external references SiMe4 for 1H and 13C, 85% H3PO4 for 31P, Me4Sn for 119Sn, and ammonium dihydrogen phosphate for 31 P CPMAS. Elemental analyses were performed by the Division of Chemistry and Biological Chemistry, Nanyang Technological University. Melting points were measured in sealed glass tubes and were not corrected. [1,3-C6H4(PhPS)2C(SnCl)(GeCl3)]2 (2). GeCl4 (1.0 mL, 1.0 mmol, 1.0 M in toluene) was added to a solution of 1 (0.57 g, 0.6 mmol) in CH2Cl2 (30 mL) at ambient temperature. The resulting white suspension was stirred overnight. After filtration and concentration of the filtrate, compound 2 was afforded as colorless crystals. The white residue is also compound 2, which was confirmed by NMR spectroscopy. Yield: 0.54 g (69%). Mp: 190 °C dec. Anal. Calcd for C38H28Cl8Ge2P4S4Sn2: C, 32.53; H, 2.01. Found: C, 32.17; H, 2.09. 1H NMR (395.9 MHz, THF-d8, 22.7 °C): δ 7.41−7.47 (m, 4H, Ph and C6H4), 7.51−7.62 (m, 16H, Ph and C6H4), 8.01−8.07 ppm (m, 8H, Ph and C6H4). 13C{1H} NMR (99.5 MHz, THF-d8, 21.8 °C): δ 54.10 (JP−C = 15.3 Hz, PCP), 128.38−129.52 (m, Ph and C6H4), 132.39− 132.30 (m, Ph and C6H4), 142.13 ppm (t, JP−C = 54.6 Hz, Cipso of Ph or C6H4). 31P{1H} NMR (160.3 MHz, THF-d8, 24.6 °C): δ 46.44 ppm. 119Sn{1H} NMR (147.6 MHz, THF-d8, 23.6 °C): δ 75.30 ppm (t, 2JSn−P = 11.9 Hz). [{1,3-C6H4(PhPS)2CSnEt2}(μ-ZnS)]2 (3). ZnEt2 (1.0 mL, 1.0 mmol, 1.0 M in hexane) was added to a suspension of 1 (0.57 g, 0.6 mmol) in toluene (20 mL). The resulting mixture was refluxed at 140 °C overnight to give an orange solution. After filtration and concentration of the filtrate, 3 was afforded as colorless crystals. Yield: 0.10 g (16%). Mp: 284 °C dec. Anal. Calcd for C46H48P4S6Sn2Zn2: C, 42.98; H, 3.76. Found: C, 42.88; H, 3.73. 1H NMR (399.5 MHz, CDCl3, 23.4 °C): δ 1.05 (t, 12H, 3JH−H = 7.8 Hz, CH2CH3), 1.17− 1.32 (m, 8H, CH2CH3), 7.30−7.40 (m, 12H, Ph and C6H4), 7.48− 7.52 (m, 4H, Ph and C6H4), 7.77−7.84 ppm (m, 12H, Ph and C6H4). 13 C{1H} NMR (99.5 MHz, CDCl3, 23.8 °C): δ 9.31 (t, 3JP−C = 3.7 Hz, SnCH2CH3), 10.60 (s, SnCH2CH3), 128.10 (s, Cpara of Ph), 128.20 (d, JP−C = 5.0 Hz, Ph or C6H4), 128.63−128.97 (m, Ph and C6H4), 130.28 (t, JP−C = 5.0 Hz, Ph or C6H4), 131.03 (d, JC−P = 69.6 Hz, Cipso of Ph or C6H4). 31P{1H} NMR (160.3 MHz, CDCl3, 24.6 °C): δ 51.65 ppm (2JSn−P = 43.3 Hz). 119Sn{1H} NMR (147.6 MHz, CDCl3, 24.5 °C): δ 53.11 ppm (t, 2JSn−P = 41.1 Hz). [1,3-C6H4(PhPS)2C(Sn)(GaCl3)] (4). A solution of 1 (0.57 g, 0.6 mmol) in CH2Cl2 (30 mL) was added to a solution of GaCl3 (0.18 g,

Figure 3. Molecular structures of (top) 4 and (bottom) 4d (50% thermal ellipsoids) in an asymmetric unit. Solvent molecules and hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg): 4, C(1)−Sn(1) = 2.419(3), C(1)−Ga(1) = 1.973(3), Sn(1)−S(1) = 2.6970(8), Sn(1)−S(2) = 2.6662(7), P(1)−C(1) = 1.761(3), P(2)−C(1) = 1.760(3), P(1)−S(1) = 2.0134(10), P(2)− S(2) = 2.0219(10), Sn(1)−C(1)−Ga(1) = 98.80(11), S(1)−Sn(1)− S(2) = 107.73(2), S(1)−Sn(1)−C(1) = 74.01(6), S(2)−Sn(1)−C(1) = 73.60(6); 4d, C(20)−Sn(2) = 2.397(3), C(20)−Ga(2) = 1.996(3), Sn(2)−S(3) = 2.6434(7), Sn(2)−S(4) = 2.6667(7), P(3)−C(20) = 1.764(3), P(4A)−C(20) = 1.773(3), P(3)−S(3) = 2.0099(10), P(4)− S(4) = 2.0282(9), Sn(2)−C(20)−Ga(2) = 100.48(11), S(3)−Sn(2)− S(4) = 86.54(2), S(3)−Sn(2)−C(20) = 74.62(6), S(4)−Sn(2)− C(20) = 96.37(6).

thiophosphinoyl substituents and the Cmethanediide atom. The Sn(1)S(1)P(1)C(1)P(2)S(2) metallacycle adopts a pseudoboat conformation in which the C(1) and Sn(1) atoms are displaced from the P(1)S(1)S(2)P(2) least-squares plane by 1.033 and 1.512 Å, respectively. The geometry around the Sn(1) atom is distorted trigonal pyramidal (sum of bond angles 255.34°). This indicates that there is a stereoactive lone pair of electrons at the Sn(1) atom. Compound 4d is a dimer of compound 4, in which the tin atoms are bridged between the thiophosphinoyl substituents of two ligands. The C−Sn bonds in 4 (2.419(3) Å) and 4d (2.397(3) Å) are slightly longer than those in 1 (2.334(2) and 2.301(2) Å). The C−Ga bonds in 4 (1.973(3) Å) and 4d (1.996(3) Å) are comparable to that in [(IMes)GaCl 3 ] (1.954(4) Å; IMes = N,N′-bis(2,4,6trimethylphenyl)imidazol-2-ylidene).17 The coexistence of compounds 4 and 4d in the solid state can be confirmed by 31P CPMAS NMR spectroscopy, in which two signals at δ 55.97 and 59.82 ppm for 4d and one singlet at δ 57.25 ppm for 4 are observed. However, in solution, only compound 4 is observed, which is confirmed by NMR 2647

dx.doi.org/10.1021/om400141j | Organometallics 2013, 32, 2643−2648

Organometallics

Article

(4) (a) Fustier, M.; Le Goff, X. F.; Le Floch, P.; Mézailles, N. J. Am. Chem. Soc. 2010, 132, 13108. (b) Cavell, R. G.; Kamalesh Babu, R. P.; Kasani, A.; McDonald, R. J. Am. Chem. Soc. 1999, 121, 5805. (c) Cantat, T.; Ricard, L.; Mézailles, N.; Le Floch, P. Organometallics 2006, 25, 6030. (d) Jones, N. D.; Lin, G.; Gossage, R. A.; McDonald, R.; Cavell, R. G. Organometallics 2003, 22, 2832. (e) Cantat, T.; Mézailles, N.; Ricard, L.; Jean, Y.; Le Floch, P. Angew. Chem., Int. Ed. 2004, 43, 6382. (f) Cantat, T.; Demange, M.; Mézailles, N.; Ricard, L.; Jean, Y.; Le Floch, P. Organometallics 2005, 24, 4838. (g) FustierBoutignon, M.; Heuclin, H.; Le Goff, X. F.; Mézailles, N. Chem. Commun. 2012, 48, 3306. (5) (a) Cantat, T.; Jaroschik, F.; Ricard, L.; Le Floch, P.; Nief, F.; Mézailles, N. Organometallics 2006, 25, 1329. (b) Cantat, T.; Jaroschik, F.; Nief, F.; Ricard, L.; Mézailles, N.; Le Floch, P. Chem. Commun. 2005, 5178. (c) Cantat, T.; Arliguie, T.; Noël, A.; Thuéry, P.; Ephritikhine, M.; Le Floch, P.; Mézailles, N. J. Am. Chem. Soc. 2009, 131, 963. (d) Tourneux, J.-C.; Berthet, J.-C.; Thuéry, P.; Mézailles, N.; Le Floch, P.; Ephritikhine, M. Dalton Trans. 2010, 39, 2494. (e) Tourneux, J.-C.; Berthet, J.-C.; Cantat, T.; Thuéry, P.; Mézailles, N.; Le Floch, P.; Ephritikhine, M. Organometallics 2011, 30, 2957. (f) Tourneux, J.-C.; Berthet, J.-C.; Cantat, T.; Thuéry, P.; Mézailles, N.; Ephritikhine, M. J. Am. Chem. Soc. 2011, 133, 6162. (g) Ma, G.; Ferguson, M. J.; McDonald, R.; Cavell, R. G. Inorg. Chem. 2011, 50, 6500. (h) Mills, D. P.; Moro, F.; McMaster, J.; van Slageren, J.; Lewis, W.; Blake, A. J.; Liddle, S. T. Nat. Chem. 2011, 3, 454. (i) Cooper, O. J.; Mills, D. P.; McMaster, J.; Moro, F.; Davies, E. S.; Lewis, W.; Blake, A. J.; Liddle, S. T. Angew. Chem., Int. Ed. 2011, 50, 2383. (j) Ren, W.; Deng, X.; Zi, G.; Fang, D.-C. Dalton Trans. 2011, 40, 9662. (k) Mills, D. P.; Cooper, O. J.; Tuna, F.; McInnes, E. J. L.; Davies, E. S.; McMaster, J.; Moro, F.; Lewis, W.; Blake, A. J.; Liddle, S. T. J. Am. Chem. Soc. 2012, 134, 10047. (6) (a) Leung, W.-P.; Wang, Z.-X.; Li, H.-W.; Mak, T. C. W. Angew. Chem., Int. Ed. 2001, 40, 2501. (b) Leung, W.-P.; Wan, C.-L.; Kan, K.W.; Mak, T. C. W. Organometallics 2010, 29, 814. (7) Guo, J.; Lau, K.-C.; Xi, H.-W.; Lim, K. H.; So, C.-W. Chem. Commun. 2010, 46, 1929. (8) (a) Leung, W.-P.; So, C.-W.; Wang, J.-Z.; Mak, T. C. W. Chem. Commun. 2003, 248. (b) Leung, W.-P.; So, C.-W.; Kan, K.-W.; Chan, H.-S.; Mak, T. C. W. Inorg. Chem. 2005, 44, 7286. (c) Leung, W.-P.; So, C.-W.; Kan, K.-W.; Chan, H.-S.; Mak, T. C. W. Organometallics 2005, 24, 5033. (d) Leung, W.-P.; Kan, K.-W.; So, C.-W.; Mak, T. C. W. Appl. Organomet. Chem. 2007, 21, 814. (e) Leung, W.-P.; So, C.W.; Wang, Z.-X.; Wang, J.-Z.; Mak, T. C. W. Organometallics 2003, 22, 4305. (9) (a) Guo, J.-Y.; Xi, H.-W.; Nowik, I.; Herber, R. H.; Li, Y.; Lim, K. H.; So, C.-W. Inorg. Chem. 2012, 51, 3996. (b) Guo, J.-Y.; Li, Y.; Ganguly, R.; So, C.-W. Organometallics 2012, 31, 3888. (10) Yang, Y.-F.; Foo, C.; Ganguly, R.; Li, Y.; So, C.-W. Organometallics 2012, 31, 6538. (11) Weidenbruch, M.; Grobecker, U.; Saak, W.; Peters, E.-M.; Peters, K. Organometallics 1998, 17, 5206. (12) Baines, K. M.; Stibbs, W. G. Coord. Chem. Rev. 1995, 145, 157. (13) Konu, J.; Chivers, T. Chem. Commun. 2008, 4995. (14) Thirumoorthi, R.; Chivers, T. Eur. J. Inorg. Chem. 2012, 3061. (15) Bochmann, M.; Bwembya, G.; Grinter, R.; Lu, J.; Webb, K. J.; Williamson, D. J.; Hursthouse, M. B.; Mazid, M. Inorg. Chem. 1993, 32, 532. (16) Gamer, M. T.; Roesky, P. W. Z. Anorg. Allg. Chem. 2001, 627, 877. (17) Marion, N.; Escudero-Adán, E. C.; Benet-Buchholz, J.; Stevens, E. D.; Fensterbank, L.; Malacria, M.; Nolan, S. P. Organometallics 2007, 26, 3256. (18) Sheldrick, G. M. SHELXL-97; Universität Göttingen, Göttingen, Germany, 1997.

1.0 mmol) in CH2Cl2 (10 mL) at ambient temperature. The resulting yellow suspension was stirred overnight to give a white suspension. After filtration and concentration of the filtrate, 4 and its dimer 4d (2:1) were afforded as colorless cocrystals. Yield: 0.53 g (71%). Mp: 256 °C dec. Anal. Calcd for C19H14Cl3GaP2S2Sn: C, 34.41; H, 2.13. Found: C, 34.24; H, 2.10. 31P CPMAS (161.73 MHz, spinning speed 10 kHz): δ 55.97 (s, 4d), 57.25 (s, 4), 59.82 (s, 4d). In solution, only 4 is observed. 1H NMR (399.5 MHz, CDCl3, 23.0 °C): δ 7.42−7.54 (m, 6H, Ph and C6H4), 7.60−7.69 (m, 4H, Ph and C6H4), 7.78−7.84 ppm (m, 4H, Ph and C6H4). 13C{1H} NMR (99.5 MHz, CDCl3, 23.4 °C): δ 128.90−129.42 (m, Ph and C6H4), 132.35−132.57 (m, Ph and C6H4), 134.14−134.21 (m, Ph and C6H4), 134.40 ppm (s, Cpara of Ph). 31P{1H} NMR (160.3 MHz, CDCl3, 23.6 °C): δ 58.27 ppm (2JP−Sn = 164.7 Hz). 119Sn{1H} NMR (147.6 MHz, CDCl3, 23.5 °C): δ −157.83 ppm (t, 2JSn−P = 176.5 Hz). X-ray Data Collection and Structural Refinement. Intensity data for compounds 2−4 were collected using a Bruker APEX II diffractometer. The crystals of 2−4 were measured at 103(2) K. The structures were solved by direct phase determination (SHELXS-97) and refined for all data by full-matrix least-squares methods on F2.18 All non-hydrogen atoms were subjected to anisotropic refinement. The hydrogen atoms were generated geometrically and allowed to ride on their respective parent atoms; they were assigned appropriate isotopic thermal parameters and included in the structure factor calculations. The disordered ethyl substituents in compound 3 were treated with the appropriate restraints (see the Supporting Information). The X-ray crystallographic data of 2−4 are summarized in Table S1 (see the Supporting Information).



ASSOCIATED CONTENT



AUTHOR INFORMATION

* Supporting Information S

CIF files, tables, and a figure giving X-ray data for 2−4, crystallographic data of 2−4, a proposed mechanism for the formation of 3, and the restraints for solving the disorder of the ethyl groups in 3. This material is available free of charge via the Internet at http://pubs.acs.org. Corresponding Author

*E-mail for C.-W.S.: [email protected]. Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS This work was supported by the Academic Research Fund Tier 1 (RG 57/11). REFERENCES

(1) (a) Liddle, S. T.; Mills, D. P.; Wooles, A. J. Chem. Soc. Rev. 2011, 40, 2164. (b) Liddle, S. T.; Mills, D. P.; Wooles, A. J. Organomet. Chem. 2010, 36, 29. (c) Izod, K. Coord. Chem. Rev. 2002, 227, 153. (d) Cantat, T.; Mézailles, N.; Auffrant, A.; Le Floch, P. Dalton Trans. 2008, 1957. (e) Jones, N. D.; Cavell, R. G. J. Organomet. Chem. 2005, 690, 5485. (f) Panda, T. K.; Roesky, P. W. Chem. Soc. Rev. 2009, 38, 2782. (2) (a) Ong, C. M.; Stephan, D. W. J. Am. Chem. Soc. 1999, 121, 2939. (b) Cantat, T.; Ricard, L.; Le Floch, P.; Mézailles, N. Organometallics 2006, 25, 4965. (c) Hull, K. L.; Noll, B. C.; Henderson, K. W. Organometallics 2006, 25, 4072. (d) Hull, K. L.; Carmichael, I.; Noll, B. C.; Henderson, K. W. Chem. Eur. J. 2008, 14, 3939. (e) Orzechowski, L.; Jansen, G.; Harder, S. Angew. Chem., Int. Ed. 2009, 48, 3825. (f) Orzechowski, L.; Jansen, G.; Harder, S. J. Am. Chem. Soc. 2006, 128, 14676. (3) (a) Foo, C.; Lau, K.-C.; Yang, Y.-F.; So, C.-W. Chem. Commun. 2009, 6816. (b) Leung, W.-P.; Chan, Y.-C.; Mak, T. C. W. Inorg. Chem. 2011, 50, 10517. (c) Thirumoorthi, R.; Chivers, T.; VargasBaca, I. Dalton Trans. 2011, 40, 8086. 2648

dx.doi.org/10.1021/om400141j | Organometallics 2013, 32, 2643−2648